Skip to main content

Full text of "A course of pure mathematics"

See other formats


A     COURSE 

OF 

PURE    MATHEMATICS 


CAMBRIDGE  UNIVERSITY  PRESS 

C.  F.  CLAY,  Manager 

LONDON    :    FETTER  LANE,   E.C.  4 


NEW  YORK  :  THE  MACMILLAN  CO. 

BOMBAY       \ 

CALCUTTA  I  MACMILLAN  AND  CO.,  Ltd. 

MADRAS       ) 

TORONTO    :    THE   MACMILLAN  CO.   OF 

CANADA,  Ltd. 
TOKYO :  MARUZEN-KABUSHIKI-KAISHA 


ALL  RIGHTS  RESERVED 


A    COURSE 


OF 


PURE  MATHEMATICS 


BY 


G.    H.    HARDY,    M.A.,    F.R.S. 

FELLOW    OF    NEW    COLLEGE 

SAVILIAN    PROFESSOR    OF    GEOMETRY    IN    THE    UNIVERSITY 

OF    OXFORD 

LATE    FELLOW    OF    TRINITY    COLLEGE,    CAMBRIDGE 


THIRD   EDITION 


Cambridge 

at  the  University  Press 

1921 


First  Edition  1908 
Second  Edition  1914 
Third  Edition   1921 


PREFACE  TO  THE  THIRD  EDITION 

NO  extensive  changes  have  been  made  in  this  edition.    The  most 
important  are  in  §§  80-82,  which  I  have  rewritten  in  accord- 
ance with  suggestions  made  by  Mr  S.  Pollard. 

The  earlier  editions  contained  no  satisfactory  account  of  the 
genesis  of  the  circular  functions.  I  have  made  some  attempt  to 
meet  this  objection  in  §  158  and  Appendix  III.  Appendix  IV  is  also 
an  addition. 

It  is  curious  to  note  how  the  character  of  the  criticisms  I  have 
had  to  meet  has  changed.  I  was  too  meticulous  and  pedantic  for 
my  pupils  of  fifteen  years  ago:  I  am  altogether  too  popular  for  the 
Trinity  scholar  of  to-day.  I  need  hardly  say  that  I  find  such 
criticisms  very  gratifying,  as  the  best  evidence  that  the  book  has 
to  some  extent  fulfilled  the  purpose  with  which  it  was  written. 

G.  H.  H. 

August  1921 


EXTRACT  FROM  THE  PREFACE  TO 
THE  SECOND  EDITION 

THE  principal  changes  made  in  this  edition  are  as  follows. 
I  have  inserted  in  Chapter  I  a  sketch  of  Dedekind's  theory 
of  real  numbers,  and  a  proof  of  Weierstrass's  theorem  concerning 
points  of  condensation;  in  Chapter  IV  an  account  of  'limits  of 
indetermination '  and  the  'general  principle  of  convergence';  in 
Chapter  V  a  proof  of  the  '  Heine-Borel  Theorem ',  Heine's  theorem 
concerning  uniform  continuity,  and  the  fundamental  theorem 
concerning  implicit  functions ;  in  Chapter  VI  some  additional 
matter  concerning  the  integration  of  algebraical  functions ;  and 
in  Chapter  VII  a  section  on  differentials.  I  have  also  rewritten 
in  a  more  general  form  the  sections  which  deal  with  the  defini- 
tion of  the  definite  integral.  In  order  to  find  space  for  these 
insertions  I  have  deleted  a  good  deal  of  the  analytical  geometry 
and  formal  trigonometry  contained  in  Chapters  II  and  III  of 
the  first  edition.  These  changes  have  naturally  involved  a 
large  number  of  minor  alterations. 

G.  H.  H. 

October  1914 


EXTEACT  FROM   THE   PREFACE   TO   THE 
FIRST   EDITION 

THIS  book  has  been  designed  primarily  for  the  use  of  first 
year  students  at  the  Universities  whose  abilities  reach  or 
approach  something  like  what  is  usually  described  as  '  scholarship 
standard'.  I  hope  that  it  may  be  useful  to  other  classes  of 
readers,  but  it  is  this  class  whose  wants  I  have  considered  first. 
It  is  in  any  case  a  book  for  mathematicians :  I  have  nowhere 
made  any  attempt  to  meet  the  needs  of  students  of  engineering 
or  indeed  any  class  of  students  whose  interests  are  not  primarily 
mathematical. 

I  regard  the  book  as  being  really  elementary.  There  are 
plenty  of  hard  examples  (mainly  at  the  ends  of  the  chapters) :  to 
these  I  have  added,  wherever  space  permitted,  an  outline  of  the 
solution.  But  I  have  done  my  best  to  avoid  the  inclusion  of 
anything  that  involves  really  difficult  ideas.  For  instance,  I  make 
no  use  of  the  '  principle  of  convergence ' :  uniform  convergence, 
double  series,  infinite  products,  are  never  alluded  to :  and  I  prove 
no  general  theorems  whatever  concerning  the  inversion  of  limit- 

operations — I  never  even  define  ~L-  and  tt^-.     In  the  last  two 
r  oxoy  oyox 

chapters  I  have  occasion  once  or  twice  to  integrate  a  power-series, 

but  I  have  confined  myself  to  the  very  simplest  cases  and  given 

a  special  discussion  in  each  instance.     Anyone  who  has  read  this 

book  will  be  in  a  position   to  read  with  profit  Dr  Bromwich's 

Infinite  Series,  where  a  full  and  adequate  discussion  of  all  these 

points  will  be  found. 


September  1908 


CONTENTS 

CHAPTER   I 

REAL    VARIABLES 

SECT.  PAGE 

1-2.  Rational  numbers .                 ••.....  1 

3-7.             Irrational  numbers ■  3 

8.  Real  numbers 13 

9.  Relations  of  magnitude  between  real  numbers    .  .15 
10-11.  Algebraical  operations  with  real  numbers    ....  17 

12.              The  number  ,/2 19 

13-14.        Quadratic  surds .         .  19 

15.  The  continuum 23 

16.  The  continuous  real  variable 26 

17.  Sections  of  the  real  numbers.     Dedekind's  Theorem  .         .  27 

18.  Points  of  condensation .  29 

19.  Weierstrass's  Theorem  .        .         .         .         .         .         .         .  30 

Miscellaneous  Examples 31 

Decimals,  1.  Gauss's  Theorem,  6.  Graphical  solution  of  quadratic 
equations,  20.  Important  inequalities,  32.  Arithmetical  and  geometrical 
meaus,  32.  Schwarz's  Inequality,  33.  Cubic  and  other  surds,  34. 
Algebraical  numbers,  36. 


CHAPTER   II 


FUNCTIONS    OF    REAL   VARIABLES 


20. 

21. 

22. 

23. 

24-25. 

26-27. 

2S-29. 

30. 

31. 


The  idea  of  a  function 

The  graphical  representation  of  functions.     Coordinates 
Polar  coordinates  . 
Polynomials    . 


Rational  functions 
Algebraical  functions     . 
Transcendental  functions 
Graphical  solution  of  equations 
Functions    of    two    variables    and    their  graphical   repre 
sentation 


38 
41 
43 
44 
47 
49 
52 
58 

59 


Vlll 


CONTENTS 


SECT. 

32. 
33. 


Curves  in  a  plane 
Loci  in  space 

Miscellaneous  Examples 


PAGE 

60 
61 


65 


Trigonometrical  functions,  53.  Arithmetical  functions,  55.  Cylinders, 
62.  Contour  maps,  62.  Cones,  63.  Surfaces  of  revolution,  63.  Ruled 
surfaces,  64.  Geometrical  constructions  for  irrational  numbers,  66. 
Quadrature  of  the  circle,  68. 


CHAPTER   III 


COMPLEX    NUMBERS 


34-38. 

39-42. 

43. 

44. 

45. 

46. 

47-49. 


Displacements 

Complex  numbers 

The  quadratic  equation  with  real  coefficients 

Argand's  diagram 

de  Moivre's  Theorem 

Rational  functions  of  a  complex  variable     . 
Roots  of  complex  numbers  .... 

Miscellaneous  Examples 


69 

78 
81 
84 
86 


101 


Properties  of  a  triangle,  90,  101.  Equations  with  complex  coefficients, 
91.  Coaxal  circles,  93.  Bilinear  and  other  transformations,  94,  97,  104. 
Cross  ratios,  96.  Condition  that  four  points  should  be  concyclic,  97. 
Complex  functions  of  a  real  variable,  97.  Construction  of  regular  polygons 
by  Euclidean  methods,  100.     Imaginary  points  and  lines,  103. 


50. 

51. 

52. 

53-57. 

58-61. 

62. 

63-68. 

69-70. 

71. 

72 

73. 

74. 
75. 

76-77. 
78. 


allies  of  n 


CHAPTER   IV 

LIMITS    OF    FUNCTIONS    OF    A    POSITIVE    INTEGRAL    VARIABLE 

Functions  of  a  positive  integral  variable     . 

Interpolation  

Finite  and  infinite  classes     .... 
Properties  possessed  by  a  function  of  n  for  large 
Definition  of  a  limit  and  other  definitions  . 

Oscillating  functions 

General  theorems  concerning  limits 
Steadily  increasing  or  decreasing  functions 
Alternative  proof  of  Weierstrass's  Theorem 
The  limit  of  .vn 

1 


The  limit  of  ( 1  + 


Some  algebraical  lemmas 
The  limit  of  n(#a?-l). 
Infinite  series 
The  infinite  sreometrical  series 


106 
107 
108 
109 
116 
121 
125 
131 
134 
134 

137 

138 
139 
140 
143 


CONTENTS 


IX 


SECT 

79. 


PAGE 


The  representation  of  functions  of  a  continuous  real  variable 

by  means  of  limits 147 

80.  The  bounds  of  a  bounded  aggregate 149 

81.  The  bounds  of  a  bounded  function 149 

82.  The  limits  of  indetermination  of  a  bounded  function    .        .  150 

83-84.        The  general  principle  of  convergence 151 

85-86.        Limits  of  complex  functions  and  series  of  complex  terms  .  153 

87-88.        Applications  to  zn  and  the  geometrical  series     .         .        .  156, 

Miscellaneous  Examples 157 

Oscillation  of  sin  n6ir,  121,  123,  151.     Limits  of  nkxn,  Z]x,  ^n,  #(n!), 

Xn      /  771  \ 

— ,  I      )  xn,  136, 139.   Decimals,  143.   Arithmetical  series,  146.   Harmonical 
n\     \n  J 

series,  147.     Equation  xn+1=f(xn),  158.    Expansions  of  rational  functions, 
159.    Limit  of  a  mean  value,  160. 


CHAPTER   V 

LIMITS    OF    FUNCTIONS    OF    A    CONTINUOUS    VARIABLE. 
AND    DISCONTINUOUS    FUNCTIONS 


Continuous 


89-92.        Limits  as  x-*-cc  or  #-*-  —  oo 162 

93-97.        Limits  as  x-*-a 165- 

98-99.  Continuous  functions  of  a  real  variable       ....  174 

100-104.  Properties   of    continuous    functions.     Bounded    functions. 

The  oscillation  of  a  function  in  an  interval          .         .  179 

105-106.  Sets  of  intervals  on  a  line.     The  Heine-Borel  Theorem     .  185 

107.  Continuous  functions  of  several  variables    .         .        .         .190 

108-109.     Implicit  and  inverse  functions 191 

Miscellaneous  Examples 194 

Limits  and  continuity  of  polynomials  and  rational  functions,  169,  176. 


Limit  of 


,  171.     Orders  of  smallness  and  greatness,  172.     Limit  of 


■ ,  173.     Infinity  of  a  function,  177.     Continuity  of  cos  x  and  sin  a:,  177. 

x 

Classification  of  discontinuities,   178. 


CHAPTER   VI 


DERIVATIVES    AND    INTEGRALS 


110-112.    Derivatives 

113.  General  rules  for  differentiation  . 

114.  Derivatives  of  complex  functions 

115.  The  notation  of  the  differential  calculus 

116.  Differentiation  of  polynomials 

117.  Differentiation  of  rational  functions     . 

118.  Differentiation  of  algebraical  functions 


197 
203 

205 
205 
207 
209 
210 


CONTENTS 


SECT. 

119.  Differentiation  of  transcendental  functions 

120.  Eepeated  differentiation 

121.  General  theorems  concerning  derivatives. 
122-124.  Maxima  and  minima     .... 
125-126.  The  Mean  Value  Theorem   . 

127-128.  Integration.     The  logarithmic  function 

129.  Integration  of  polynomials  . 

130-131.  Integration  of  rational  functions 

132-139.  Integration     of    algebraical     functions. 

rationalisation.     Integration  by  parts 

140-144.  Integration  of  transcendental  functions 

145.  Areas  of  plane  curves  .... 

146.  Lengths  of  plane  curves 

Miscellaneous  Examples 


Rolle's  Theorem 


Integration     by 


PAGE 

212 
214 
217 
219 
226 
228 
232 
233 

236 
245 
249 
251 

253 


Derivative  of  xm,  201.  Derivatives  of  cos.r  and  sin.r,  201.  Tangent 
and  normal  to  a  curve,  201,  214.  Multiple  roots  of  equations,  208,  255. 
Rolle's  Theorem  for  polynomials,  209.  Leibniz'  Theorem,  215.  Maxima 
and  minima  of  the  quotient  of  two  quadratics,  223,  256.  Axes  of  a  conic, 
226.  Lengths  and  areas  in  polar  coordinates,  253.  Differentiation  of  a 
determinant,  254.  Extensions  of  the  Mean  Value  Theorem,  258.  Formulae 
of  reduction,  259. 


CHAPTER   VII 

ADDITIONAL    THEOREMS    IN    THE    DIFFERENTIAL    AND    INTEGRAL    CALCULUS 


147.  Taylor's  Theorem 

148.  Taylor's  Series 

149.  Applications  of  Taylor's  Theorem  to  maxima  and  minima 

150.  Applications  of  Taylor's  Theorem  to  the  calculation  of  limits 

151.  The  contact  of  plane  curves 
152-154.  Differentiation  of  functions  of  several  variables 

155.  Differentials 

156-161.  Definite  Integrals.     Areas  of  curves      . 

162.  Alternative  proof  of  Taylor's  Theorem 

163.  Application  to  the  binomial  series 

164.  Integrals  of  complex  functions     . 

Miscellaneous  Examples 


262 
266 

268 
268 
270 
274 
280 
283 
298 
299 
299 
300 


Newton's  method  of  approximation  to  the  roots  of  equations,  265. 
Series  for  cos  x  and  sin  x,  267.  Binomial  series,  267.  Tangent  to  a  curve, 
272,  283,  303.  Points  of  inflexion,  272.  Curvature,  273,  302.  Osculating 
conies,  274,  302.  Differentiation  of  implicit  functions,  283.  Fourier's 
integrals,  290,  294.  The  second  mean  value  theorem,  296.  Homogeneous 
functions,  302.  Euler's  Theorem,  302.  Jacobiaus,  303.  Schwarz's  in- 
equality for  integrals,  306.  Approximate  values  of  definite  integrals,  307. 
Simpson's  Rule,  307. 


CONTENTS 


XI 


CHAPTER     VIII 

THE    CONVERGENCE    OP    INFINITE    SERIES    AND    INFINITE    INTEGRALS 


SECT. 

165-168.    Series  of  positive  terms.     Cauchy's  and  d'Alembert's  tests 
of  convergence 

169.  Dirichlet's  Theorem 

170.  Multiplication  of  series  of  positive  terms     . 
171-174.    Further  tests  of  convergence.    Abel's  Theorem.     Maclaurin's 

integral  test    .         .         .         ... 

175.  The  series  2n~8 

176.  Cauchy's  condensation  test   .... 

177-182.  Infinite  integrals 

183.  Series  of  positive  and  negative  terms  . 

184-185.  Absolutely  convergent  series 

186-187.  Conditionally  convergent  series     . 

188.  Alternating  series 

189.  Abel's  and  Dirichlet's  tests  of  convergence 

190.  Series  of  complex  terms        .... 

191-194.  Power  series 

195.  Multiplication  of  series  in  general 

Miscellaneous  Examples 


PAGE 


sos- 
sis 

313 


315 

319 

320 

321 

335 

336 

338 

340 

342 

344. 

345. 

349 

350 


The  series  2nlr"  and  allied  series,  311.  Transformation  of  infinite 
integrals  by  substitution  and  integration  by  parts,  327,  328,  333.  The 
series  2ancos?i0,  2a„sin?i0,  338,  343,  344.  Alteration  of  the  sum  of  a 
series  by  rearrangement,  341.  Logarithmic  series,  348.  Binomial  series, 
348,  349.  Multiplication  of  conditionally  convergent  series,  350,  354. 
Recurring  series,  352.  Difference  equations,  353.  Definite  integrals,  355. 
Schwarz's   inequality  for  infinite  integrals,  356. 


CHAPTER   IX 

THE    LOGARITHMIC    AND    EXPONENTIAL    FUNCTIONS    OF    A    REAL    VARIABLE 


196-197.  The  logarithmic  function       .... 

198.  The  functional  equation  satisfied  by  log.r  . 

199-201.  The  behaviour  of  log  x  as  x  tends  to  infinity  or 

202.  The  logarithmic  scale  of  infinity 

203.  The  number  e 

204-206.  The  exponential  function       .... 

207.  The  general  power  ax 

208.  The  exponential  limit 

209.  The  logarithmic  limit 

210.  Common  logarithms 

211.  Losrarithmic  tests  of  convergence . 


to  zero 


357 
360 
300 
362 
363 
364 
366 
368 
369 
369 
374 


Xll 


CONTENTS 


SECT. 

212. 
213. 

214. 
215. 
216. 


The  exponential  series  .         .         .         . 
The  logarithmic  series  .         .         .        . 
The  series  for  arc  tan. v 
The  binomial  series       . 
Alternative  development  of  the  theory 

Miscellaneous  Examples 


IMGE 

378 
381 
382 
384 
386 

387 


Integrals  containing  the  exponential  function,  370.  The  hyperbolic 
functions,  372.  Integrals  of  certain  algebraical  functions,  373.  Euler's 
constant,  377,  389.  Irrationality  of  e,  380.  Approximation  to  surds  by  the 
binomial  theorem,  385.    Irrationality  of  logjon,  387.    Definite  integrals,  393. 


CHAPTER   X 

THE   GENERAL   THEORY   OF   THE   LOGARITHMIC,    EXPONENTIAL, 
AND    CIRCULAR    FUNCTIONS 


217-218. 
219. 

220. 

221. 

222-224. 

225-226. 

227-230. 

231. 

232. 

233. 

234-235. 

236. 

237. 


Functions  of  a  complex  variable  . 

Curvilinear  integrals      .... 

Definition  of  the  logarithmic  function 

The  values  of  the  logarithmic  function 

The  exponential  function 

The  general  power  a2    .... 

The  trigonometrical  and  hyperbolic  functions 

The    connection    between     the    logarithmic    and 

trigonometrical  functions 
The  exponential  series 
The  series  for  cos  2  and  sin  z 
The  logarithmic  series  . 
The  exponential  limit    . 
The  binomial  series 


Miscellaneous  Examples 


395 
396 
397 
399 
403 
404 
409 

413 
414 
416 
417 
421 
422 

425 


The  functional  equation  satisfied  by  Log  z,  402.  The  function  e2,  407. 
Logarithms  to  any  base,  408.  The  inverse  cosine,  sine,  and  tangent  of  a 
complex  number,  412.  Trigonometrical  series,  417,  420,  431.  Boots  of 
transcendental  equations,  425.  Transformations,  426,  428.  Stereographic 
projection,  427.  Mercator's  projection,  428.  Level  curves,  429.  Definite 
integrals,  432. 


Appendix  I.  The  proof  that  every  equation  has  a  root 

Appendix  II.  A  note  on  double  limit  problems  . 
Appendix  III.     The  circular  functions 

Appendix  IV.  The  infinite  in  analysis  and  geometry 


433 
439 

443 
445 


CHAPTER  I 

REAL    VARIABLES 

1.  Rational  numbers.  A  fraction  r  =p/q,  where  p  and  q 
are  positive  or  negative  integers,  is  called  a  rational  number.  We 
can  suppose  (i)  that  p  and  q  have  no  common  factor,  as  if  they 
have  a  common  factor  we  can  divide  each  of  them  by  it,  and 
(ii)  that  q  is  positive,  since 

pl(r  ?)  =  (-p)/q>   (-p)K-  q)=p/q- 

To  the  rational  numbers  thus  denned  we  may  add  the  '  rational 
number  0 '  obtained  by  taking  p  =  0. 

We  assume  that  the  reader  is  familiar  with  the  ordinary 
arithmetical  rules  for  the  manipulation  of  rational  numbers.  The 
examples  which  follow  demand  no  knowledge  beyond  this. 

Examples  I.  1.  If  r  and  s  are  rational  numbers,  then  r  +  s,  r  -  s,  rs,  and 
rjs  are  rational  numbers,  unless  in  the  last  case  s=0  (when  rjs  is  of  course 
meaningless). 

2.  If  X,  m,  and  n  are  positive  rational  numbers,  and  m  >  n,  then 
X(m2  —  ?i2),  2X«m,  and  X(m2  +  ;i2)  are  positive  rational  numbers.  Hence  show 
how  to  determine  any  number  of  right-angled  triangles  the  lengths  of  all  of 
whose  sides  are  rational. 

3.  Any  terminated  decimal  represents  a  rational  number  whose  denomi- 
nator contains  no  factors  other  than  2  or  5.  Conversely,  any  such  rational 
number  can  be  expressed,  and  in  one  way  only,  as  a  terminated  decimal. 

[The  general  theory  of  decimals  will  be  considered  in  Ch.  IV.] 

4.  The  positive  rational  numbers  may  be  arranged  in  the  form  of  a  simple 
scries  as  follows : 

l    i    l    J     2     14    3     2     1 

1)   1>   2  5    D    2»   3'    1)   2>   3  5   it"" 

Show  that  p/q  is  the  [^  (p  +  q  - 1)  (p  +  q  -  2)  +  q]th  term  of  the  series. 

[In  this  series  every  rational  number  is  repeated  indefinitely.  Thus  1 
occurs  as  {,  f ,  % , ....     We  can  of  course  avoid  this  by  omitting  every  number 

H.  1 


2  REAL   VARIABLES  [i 

which  has  already  occurred  in  a  simpler  form,  but  then  the  problem  of  deter- 
mining the  precise  position  of  pjq  becomes  more  complicated.] 

2.  The  representation  of  rational  numbers  by  points 
on  a  line.  It  is  convenient,  in  many  branches  of  mathematical 
analysis,  to  make  a  good  deal  of  use  of  geometrical  illustrations. 

The  use  of  geometrical  illustrations  in  this  way  does  not,  of 
course,  imply  that  analysis  has  any  sort  of  dependence  upon 
geometry :  they  are  illustrations  and  nothing  more,  and  are  em- 
ployed merely  for  the  sake  of  clearness  of  exposition.  This  being 
so,  it  is  not  necessary  that  we  should  attempt  any  logical  analysis 
of  the  ordinary  notions  of  elementary  geometry;  we  may  be  content 
to  suppose,  however  far  it  may  be  from  the  truth,  that  we  know 
what  they  mean. 

Assuming,  then,  that  we  know  what  is  meant  by  a  straight 
line,  a  segment  of  a  line,  and  the  length  of  a  segment,  let  us  take 
a  straight  line  A,  produced  indefinitely  in  both  directions,  and  a 
segment  A(>A1  of  any  length.  We  call  A0  the  origin,  or  the  point 
0,  and  J.!  the  point  1,  and  we  regard  these  points  as  representing 
the  numbers  0  and  1. 

In  order  to  obtain  a  point  which  shall  represent  a  positive 
rational  number  r=p/q,  we  choose  the  point  Ar  such  that 

A0ArjA0A1  =  r, 

A0Ar  being  a  stretch  of  the  line  extending  in  the  same  direction 
along  the  line  as  A0AU  a  direction  which  we  shall  suppose  to  be 
from  left  to  right  when,  as  in  Fig.  1,  the  line  is  drawn  horizontally 
across  the   paper.      In  order  to  obtain  a  point   to  represent  a 

1 j 1 1 ! 

A-s  A_i  A0  Ax  A, 

Fig.  1. 

negative  rational  number  r  =  —  s,  it  is  natural  to  regard  length  as 
a  magnitude  capable  of  sign,  positive  if  the  length  is  measured  in 
one  direction  (that  of  ^o^),  and  negative  if  measured  in  the 
other,  so  that  AB  =  —  BA  ;  and  to  take  as  the  point  representing 
r  the  point  A_s  such  that 

A0A_S  =  —  A_SA0  =  —  A0A8, 


1-3]  REAL   VARIABLES  3 

We  thus  obtain  a  point  Ar  on  the  line  corresponding  to  every 
rational  value  of  r,  positive  or  negative,  and  such  that 

A0Ar  =  r .  AqAx) 

and  if,  as  is  natural,  we  take  A0At  as  our  unit  of  length,  and  write 
A0A1=  1,  then  we  have 

A0Ar  =  r. 

We  shall  call  the  points  Ar  the  rational  points  of  the  line. 

3.  Irrational  numbers.  If  the  reader  will  mark  off  on  the 
line  all  the  points  corresponding  to  the  rational  numbers  whose 
denominators  are  1,  2,  3, ...  in  succession,  he  will  readily  convince 
himself  that  he  can  cover  the  line  with  rational  points  as  closely 
as  he  likes.  We  can  state  this  more  precisely  as  follows :  if  we 
take  any  segment  BG  on  A,  we  can  find  as  many  rational  points  as 
we  please  on  BG. 

Suppose,  for  example,  that  BG  falls  within  the  segment  AXA,. 
It  is  evident  that  if  we  choose  a  positive  integer  k  so  that 

k.BC>l    (1),* 

and  divide  A1A2  into  k  equal  parts,  then  at  least  one  of  the  points 
of  division  (say  P)  must  fall  inside  BG,  without  coinciding  with 
either  B  or  C.  For  if  this  were  not  so,  BG  would  be  entirely 
included  in  one  of  the  k  parts  into  which  AXA2  has  been  divided, 
which  contradicts  the  supposition  (1).  But  P  obviously  corre- 
sponds to  a  rational  number  whose  denominator  is  k.  Thus  at 
least  one  rational  point  P  lies  between  B  and  G.  But  then  we 
can  find  another  such  point  Q  between  B  and  P,  another  between 
B  and  Q,  and  so  on  indefinitely ;  i.e.,  as  we  asserted  above,  we  can 
find  as  many  as  we  please.  We  may  express  this  by  saying  that 
BG  includes  infinitely  many  rational  points. 

The  meaning  of  such  phrases  as  '■infinitely  many'  or  '•an  infinity  of,  in 
such  sentences  as  '  BG  includes  infinitely  many  rational  points '  or  '  there  are 
an  infinity  of  rational  points  on  BG'  or  'there  are  an  infinity  of  positive 
integers',  will  be  considered  more  closely  in  Ch.  IV.  The  assertion  'there  are 
an  infinity  of  positive  integers '  means  '  given  any  positive  integer  n,  however 
large,   we  can  find   more  than  n  positive  integers'.     This  is    plainly  true 

*  The  assumption  that  this  is  possible  is  equivalent  to  the  assumption  of  what 
is  known  as  the  Axiom  of  Archimedes. 

1—2 


4  KEAL  VAKIABLES  [i 

whatever  n  may  be,  e.g.  for  n  =  100,000  or  100,000,000.     The  assertion  means 
exactly  the  same  as  '  we  can  find  as  many  positive  integers  as  we  please '. 

The  reader  will  easily  convince  himself  of  the  truth  of  the  following 
assertion,  which  is  substantially  equivalent  to  what  was  proved  in  the  second 
paragraph  of  this  section  :  given  any  rational  number  r,  and  any  positive 
integer  n,  we  can  find  another  rational  number  lying  on  either  side  of  r  and 
differing  from  r  by  less  than  1/n.  It  is  merely  to  express  this  differently  to 
say  that  we  can  find  a  rational  number  lying  on  either  side  of  r  and  differing 
from  r  by  as  little  as  we  please.  Again,  given  any  two  rational  numbers 
}•  and  s,  we  can  interpolate  between  them  a  chain  of  rational  numbers  in 
which  any  two  consecutive  terms  differ  by  as  little  as  we  please,  that  is  to 
say  by  less  than  1/n,  where  n  is  any  positive  integer  assigned  beforehand. 

From  these  considerations  the  reader  might  be  tempted  to 
infer  that  an  adequate  view  of  the  nature  of  the  line  could  be 
obtained  by  imagining  it  to  be  formed  simply  by  the  rational 
points  which  lie  on  it.  And  it  is  certainly  the  case  that  if  we 
imagine  the  line  to  be  made  up  solely  of  the  rational  points, 
and  all  other  points  (if  there  are  any  such)  to  be  eliminated, 
the  figure  which  remained  would  possess  most  of  the  properties 
which  common  sense  attributes  to  the  straight  line,  and  would, 
to  put  the  matter  roughly,  look  and  behave  very  much  like 
a  line. 

A  little  further  consideration,  however,  shows  that  this  view 
would  involve  us  in  serious  difficulties. 

Let  us  look  at  the  matter  for  a  moment  with  the  eye  of 
common  sense,  and  consider  some  of  the  properties  which  we  may 
reasonably  expect  a  straight  line  to  possess  if  it  is  to  satisfy  the 
idea  which  we  have  formed  of  it  in  elementary  geometry. 

The  straight  line  must  be  composed  of  points,  and  any  segment 
of  it  by  all  the  points  which  lie  between  its  end  points.  With 
any  such  segment  must  be  associated  a  certain  entity  called  its 
length,  which  must  be  a  quantity  capable  of  numerical  measure- 
ment in  terms  of  any  standard  or  unit  length,  and  these  lengths 
must  be  capable  of  combination  with  one  another,  according  to 
the  ordinary  rules  of  algebra,  by  means  of  addition  or  multipli- 
cation. Again,  it  must  be  possible  to  construct  a  line  whose 
length  is  the  sum  or  product  of  any  two  given  lengths.  If  the 
length  PQ,  along  a  given  line,  is  a,  and  the  length  QR,  along 
the   same   straight   line,   is   b,   the   length   PR   must  be   a  +  b. 


3] 


REAL   VARIABLES 


Moreover,  if  the  lengths  OP,  OQ,  along  one  straight  line,  are 
1  and  a,  and  the  length  OR  along  another  straight  line  is  b, 
and  if  we  determine  the  length  OS  by  Euclid's  construction  (Euc. 
VI.  12)  for  a  fourth  proportional  to  the  lines  OP,  OQ,  OR,  this 
length  must  be  ab,  the  algebraical  fourth  proportional  to  1,  a,  b. 
And  it  is  hardly  necessary  to  remark  that  the  sums  and  products 
thus  defined  must  obey  the  ordinary  '  laws  of  algebra ' ;  viz. 

a  +  b  =  b  +  a,  a  +  (b  4-  c)==(a+b)  +  c, 
ab  =  ba,  a  (be)  =  (ab)  c,  a(b  +  c)  =  ab  +  ac. 
The  lengths  of  our  lines  must  also  obey  a  number  of  obvious 
laws  concerning  inequalities  as  well  as  equalities :  thus  if 
A,  B,  C  are  three  points  lying  along  A  from  left  to  right,  we  must 
have  AB<  AC,  and  so  on.  Moreover  it  must  be  possible,  on  our 
fundamental  line  A,  to  find  a  point  P  such  that  A0P  is  equal  to 
any  segment  whatever  taken  along  A  or  along  any  other  straight 
line.  All  these  properties  of  a  line,  and  more,  are  involved  in  the 
presuppositions  of  our  elementary  geometry. 

Now  it  is  very  easy  to  see  that  the  idea  of  a  straight  line  as 
composed  of  a  series  of  points,  each  corresponding  to  a  rational 
number,  cannot  possibly  satisfy  all  these  requirements.  There  are 
various  elementary  geometrical  constructions,  for  example,  which 
purport  to  construct  a  length  x  such  that  x2  =  2.    For  instance,  we 


Fig.  2. 

may  construct  an  isosceles  right-angled  triangle  ABC  such  that 
AB  =  AC=1.  Then  if  BC=  x,  x~  =  2.  Or  we  may  determine 
the  length  x  by  means  of  Euclid's  construction  (Euc.  vi.  13)  for 
a  mean  proportional  to  1  and  2,  as  indicated  in  the  figure.  Our 
requirements  therefore  involve  the  existence  of  a  length  measured 
by  a  number  x,  and  a  point  P  on  A  such  that 

AQP  =  x,     x2  =  2. 


6  REAL  VARIABLES  [i 

But  it  is  easy  to  see  that  there  is  no  rational  number  such  that 
its  square  is  2.  In  fact  we  may  go  further  and  say  that  there 
is  no  rational  number  whose  square  is  m/n,  where  mjn  is  any 
positive  fraction  in  its  lowest  terms,  unless  m  and  n  are  both 
perfect  squares. 

For  suppose,  if  possible,  that 

p"jq-  =  m/n. 

p  having  no  factor  in  common  with  q,  and  m  no  factor  m  common 
with  n.  Then  np2  =  mq2.  Every  factor  of  q2  must  divide  np2,  and 
as  p  and  q  have  no  common  factor,  every  factor  of  q2  must  divide 
n.  Hence  n=\q2,  where  A.  is  an  integer.  But  this  involves 
m  =  Xp2 :  and  as  m  and  n  have  no  common  factor,  A  must  be  unity. 
Thus  m  =p2,  n  =  q2,  as  was  to  be  proved.  In  particular  it  follows, 
by  taking  n ■—  1,  that  an  integer  cannot  be  the  square  of  a  rational 
number,  unless  that  rational  number  is  itself  integral. 

It  appears  then  that  our  requirements  involve  the  existence  of 
a  number  x  and  a  point  P,  not  one  of  the  rational  points  already 
constructed,  such  that  A0P  =  cc,  x2  =  2;  and  (as  the  reader  will 
remember  from  elementary  algebra)  we  write  x  =  *J2. 

The  following  alternative  proof  that  no  rational  number  can  have  its 
square  equal  to  2  is  interesting. 

Suppose,  if  possible,  that  p/q  is  a  positive  fraction,  in  its  lowest  terms, 
such  that  (p/q)2  =  2  or  p2  =  2q2.  It  is  easy  to  see  that  this  involves 
(2q—p)2  =  2(p-q)2;  and  so  (Zq-p)/(p  —  q)  is  another  fraction  having  the 
same  property.  But  clearly  q<p<2q,  and  so  p  —  q<q.  Hence  there  is 
another  fraction  equal  to  pjq  and  having  a  smaller  denominator,  which 
contradicts  the  assumption  that  p/q  is  in  its  lowest  terms. 

Examples  II.  1.  Show  that  no  rational  number  can  have  its  cube  equal! 
to  2. 

2.  Prove  generally  that  a  rational  fraction  p/q  in  its  lowest  terms  cannot 
be  the  cube  of  a  rational  number  unless  p  and  q  are  both  perfect  cubes. 

3.  A  more  general  proposition,  which  is  due  to  Gauss  and  includes  those 
which  precede  as  particular  cases,  is  the  following:  an  algebraical  equation 

xn+p1xn~1  +  p2xn-2  +  ...+pn=Q, 

with  integral  coefficients^  cannot  have  a  rational  but  non-integral  root. 

[For  suppose  that  the  equation  has  a  root  a/b,  where  a  and  6  are  integers 


3,  4]  REAL    VARIABLES  7 

without  a  common  factor,  and  b  is  positive.     Writing  a/6  for  x,  and  multiply- 
ing by  b11'1,  we  obtain 

an 
--j~=p1an-1+p2an-2b  +  ...+pnbn~\ 

a  fraction  in  its  lowest  terms  equal  to  an  integer,  which  is  absurd.    Thus  b=  1 
and  the  root  is  a.     It  is  evident  that  a  must  be  a  divisor  of  pn.\ 

4.  Show  that  if^>rt=l  and  neither  of 

1 +^1+^2+^3  +  ."  ,     *-pi+Ps-p3+:> 
is  zero,  then  the  equation  cannot  have  a  rational  root. 

5.  Find  the  rational  roots  (if  any)  of 

xa  _  4^3  _  8x2  + 1 3  v  +-i  ()  _  0. 

[The  roots  can  only  be  integral,  and  so  +1,  +2,  +5,  +10  are  the  only 
possibilities  :  whether  these  are  roots  can  be  determined  by  trial.  It  is  clear 
that  we  can  in  this  way  determine  the  rational  roots  of  any  such  equation.] 

4.  Irrational  numbers  (continued).  The  result  of  our 
geometrical  representation  of  the  rational  numbers  is  therefore  to 
suggest  the  desirability  of  enlarging  our  conception  of  '  number ' 
by  the  introduction  of  further  numbers  of  a  new  kind. 

The  same  conclusion  might  have  been  reached  without  the  use 
of  geometrical  language.  One  of  the  central  problems  of  algebra 
is  that  of  the  solution  of  equations,  such  as 

x~  =  1,     x'2  =  2. 

The  first  equation  has  the  two  rational  roots  1  and  —  1.  But, 
if  our  conception  of  number  is  to  be  limited  to  the  rational 
numbers,  we  can  only  say  that  the  second  equation  has  no  roots ; 
and  the  same  is  the  case  with  such  equations  as  x3  =  2,  x4  =  7. 
These  facts  are  plainly  sufficient  to  make  some  generalisation  of 
our  idea  of  number  desirable,  if  it  should  prove  to  be  possible. 

Let  us  consider  more  closely  the  equation  x1  —  2. 

We  have  already  seen  that  there  is  no  rational  number  x  which 
satisfies  this  equation.  The  square  of  any  rational  number  is 
either  less  than  or  greater  than  2.  We  can  therefore  divide  the 
rational  numbers  into  two  classes,  one  containing  the  numbers 
whose  squares  are  less  than  2,  and  the  other  those  whose  squares 
are  greater  than  2.  We  shall  confine  our  attention  to  the  positive 
rational  numbers,  and  we  shall  call  these  two  classes  the  class  L,  or 
the  lower  class,  or  the  left-hand  class,  and  the  class  R,  or  the  upper 


8  REAL   VARIABLES  [i 

class,  or  the  right-hand  class.  It  is  obvious  that  every  member  of 
R  is  greater  than  all  the  members  of  L.  Moreover  it  is  easy  to 
convince  ourselves  that  we  can  find  a  member  of  the  class  L  whose 
square,  though  less  than  2,  differs  from  2  by  as  little  as  we  please, 
and  a  member  of  R  whose  square,  though  greater  than  2,  also 
differs  from  2  by  as  little  as  we  please.  In  fact,  if  we  carry  out 
the  ordinary  arithmetical  process  for  the  extraction  of  the  square 
root  of  2,  we  obtain  a  series  of  rational  numbers,  viz. 

1,  14,    141.    1414,    14142,... 

whose  squares 

1,   1-96,    1-9881,   1-999396,   1-99996164,... 

are  all  less  than  2,  but  approach  nearer  and  nearer  to  it ;  and  by 
taking  a  sufficient  number  of  the  figures  given  by  the  process  we 
can  obtain  as  close  an  approximation  as  we  want.  And  if  we 
increase  the  last  figure,  in  each  of  the  approximations  given  above, 
by  unity,  we  obtain  a  series  of  rational  numbers 

2,  1-5,    1-42,    1-415,    1-4143,... 

whose  squares 

4,   2-25,   2-0164,    2-002225,    2-00024449,... 
are  all  greater  than  2  but  approximate  to  2  as  closely  as  we  please. 

The  reasoning  which  precedes,  although  it  will  probably  convince  the 
reader,  is  hardly  of  the  precise  character  required  by  modern  mathematics. 
We  can  supply  a  formal  proof  as  follows.  In  the  first  place,  we  can  find 
a  member  of  L  and  a  member  of  R,  differing  by  as  little  as  we  please.  For 
we  saw  in  §  3  that,  given  any  two  rational  numbers  a  and  b,  we  can  construct 
a  chain  of  rational  numbers,  of  which  a  and  b  are  the  first  and  last,  and  in 
which  any  two  consecutive  numbers  differ  by  as  little  as  we  please.  Let  us 
then  take  a  member  x  of  L  and  a  member  y  of  B,  and  interpolate  between 
them  a  chain  of  rational  numbers  of  which  oo  is  the  first  and  y  the  last,  and 
in  which  any  two  consecutive  numbers  differ  by  less  than  8,  8  being  any 
positive  rational  number  as  small  as  we  please,  such  as  -01  or  -0001  or  -000001. 
In  this  chain  there  must  be  a  last  which  belongs  to  L  and  a  first  which  belongs 
to  R,  and  these  two  numbers  differ  by  less  than  8. 

"We  can  now  prove  that  an  x  can  be  found  in  L  and  a  y  in  R  such  that 
2  —  x*  and  y2-2  are  as  small  as  we  please,  say  less  than  8.  Substituting  %8 
for  8  in  the  argument  which  precedes,  we  see  that  we  can  choose  x  and  y  so 
that  y  —  x  <  J  8  ;  and  we  may  plainly  suppose  that  both  x  and  y  are  less 
than  2.     Thus 

y  +  x<4,      y2-x-  =  (y  -x)(y+x)  <i{y  -x)<8  ; 


4,5] 


REAL    VARIABLES 


and  since  x"-<1  and  ?/2>2  it  follows  a  fortiori  that  2-x2  and  y- 
less  than  8. 


2  arc  each 


It  follows  also  that  there  can  be  no  largest  member  of  L  or 
smallest  member  of  R.  For  if  x  is  any  member  of  L,  then  x~  <  2. 
Suppose  that  x2  =  2  —  8.  Then  we  can  find  a  member  xt  of  L 
such  that  Xj2  differs  from  2  by  less  than  8,  and  so  oc?  >  x-  or  xx  >  x. 
Thus  there  are  larger  members  of  L  than  x;  and  as  x  is  any 
member  of  L,  it  follows  that  no  member  of  L  can  be  larger  than 
all  the  rest.  Hence  L  has  no  largest  member,  and  similarly  R  has 
no  smallest. 

5.  Irrational  numbers  (continued).  We  have  thus  divided 
the  positive  rational  numbers  into  two  classes,  L  and  R,  such  that 
(i)  every  member  of  R  is  greater  than  every  member  of  L,  (ii)  we 
can  find  a  member  of  L  and  a  member  of  R  whose  difference  is  as 
small  as  we  please,  (iii)  L  has  no  greatest  and  R  no  least  member. 
Our  common-sense  notion  of  the  attributes  of  a  straight  line,  the 
requirements  of  our  elementary  geometry  and  our  elementary 
algebra,  alike  demand  the  existence  of  a  number  x  greater  than  all 
the  member's  of  L  and  less  than  all  the  members  of  R,  and  of 
a  corresponding  point  P  on  A  such  that  P  divides  the  points  which 
correspond  to  members  of  L  from  those  which  correspond  to  members 
ofR. 


L  LL 


-i r-H- 


R   R 


Fist.  3. 


Let  us  suppose  for  a  moment  that  there  is  such  a  number  x, 
and  that  it  may  be  operated  upon  in  accordance  with  the  laws  of 
algebra,  so  that,  for  example,  x2  has  a  definite  meaning.  Then  a? 
cannot  be  either  less  than  or  greater  than  2.  For  suppose,  for 
example,  that  x2  is  less  than  2.  Then  it  follows  from  what  pre- 
cedes that  we  can  find  a  positive  rational  number  £  such  that  £2  lies 


10  REAL  VARIABLES  [i 

between  x2  and  2.  That  is  to  say,  we  can  find  a  member  of  L 
greater  than  x;  and  this  contradicts  the  supposition  that  a?  divides 
the  members  of  L  from  those  of  R.  Thus  x"  cannot  be  less  than 
2,  and  similarly  it  cannot  be  greater  than  2.  We  are  therefore 
driven  to  the  conclusion  that  x2  =  2,  and  that  x  is  the  number 
which  in  algebra  we  denote  by  t/2.  And  of  course  this  number 
\J2  is  not  rational,  for  no  rational  number  has  its  square  equal  to 
2.  It  is  the  simplest  example  of  what  is  called  an  irrational 
number. 

But  the  preceding  argument  may  be  applied  to  equations 
other  than  x2  =  2,  almost  word  for  word ;  for  example  to  x'2  =  N, 
where  N  is  any  integer  which  is  not  a  perfect  square,  or  to 

x3  =  S,     x3=7,     #4  =  23, 

or,  as  we  shall  see  later  on,  to  x3  =  Sx  +  8.  We  are  thus  led  to 
believe  in  the  existence  of  irrational  numbers  x  and  points  P  on 
A  such  that  x  satisfies  equations  such  as  these,  even  when  these 
lengths  cannot  (as  *J2  can)  be  constructed  by  means  of  elementary 
geometrical  methods. 

The  reader  will  no  doubt  remember  that  in  treatises  on  elementary  algebra 
the  root  of  such  an  equation  as  afl=n  is  denoted  by  v/n  or  n1^,  and  that  a 
meaning  is  attached  to  such  symbols  as 

by  means  of  the  equations 

np«i=(n1li)p,    »»"«»-J»/«=b1. 

And  he  will  remember  how,  in  virtue  of  these  definitions,  the  'laws  of  indices' 
such  as 

nr  x  n*  =  wr + ',    (nr)3  =  n™ 

are  extended  so  as  to  cover  the  case  in  which  r  and  s  are  any  rational  numbers 
whatever. 

The  reader  may  now  follow  one  or  other  of  two  alternative 
courses.  He  may,  if  he  pleases,  be  content  to  assume  that 
'irrational  numbers'  such  as  *J2,  \/3,  ...  exist  and  are  amenable  to 
the  algebraical  laws  with  which  he  is  familiar*.  If  he  does  this 
he  will  be  able  to  avoid  the  more  abstract  discussions  of  the  next 
few  sections,  and  may  pass  on  at  once  to  §§  13  et  seq. 

If,  on  the  other  hand,  he  is  not  disposed  to  adopt  so  naive  an 

*  This  is  the  point  of  view  which  was  adopted  in  the  first  edition  of  this  book. 


5,  6]  REAL   VARIABLES  11 

attitude,  he  will  be  well  advised  to  pay  careful  attention  to  the 
sections  which  follow,  in  which  these  questions  receive  fuller 
consideration  *. 

Examples  III.     1.    Find  the  difference  between  2  and  the  squares  of  the 
decimals  given  in  §  4  as  approximations  to  N/2. 

2.     Find  the  differences  between  2  and  the  squares  of 


3.  Show  that  if  mjn  is  a  good  approximation  to  s]%  then  (m  +  2n)l(m  +  n) 
is  a  better  one,  and  that  the  errors  in  the  two  cases  are  in  opposite  directions. 
Apply  this  result  to  continue  the  series  of  approximations  in  the  last 
example. 

4.  If  x  and  y  are  approximations  to  ,/2,  by  defect  and  by  excess  respec- 
tively, and  2  —  x2 < S,  y"  —  2 < 8,  then  y-x< 8. 

5.  The  equation  x2  =  4  is  satisfied  by  x=2.  Examine  how  far  the  argu- 
ment of  the  pi-eceding  sections  applies  to  this  equation  (writing  4  for  2 
throughout).  [If  we  define  the  classes  L,  R  as  before,  they  do  not  include  all 
rational  numbers.  The  rational  number  2  is  an  exception,  since  22  is  neither 
less  than  or  greater  than  4.] 

6.  Irrational  numbers  (continued).  In  Jjijwe  discussed 
a  special  mode  of  division  of  the  positive  rational  numbers  x  into 
two  classes,  such  that  x2  <  2  for  the  members  of  one  class  and 
x2  >  2  for  those  of  the  others.  Such  a  mode  of  division  is  called  a 
section  of  the  numbers  in  question.  It  is  plain  that  we  could 
equally  well  construct  a  section  in  which  the  numbers  of  the  twro 
classes  were  characterised  by  the  inequalities  x3  <  2  and  a?  >  2,  or 
x*<7  and  x*  >  7.  Let  us  now  attempt  to  state  the  principles 
of  the  construction  of  such  '  sections '  of  the  positive  rational 
numbers  in  quite  general  terms. 

Suppose  that  P  and  Q  stand  for  two  properties  which  are 
mutually  exclusive  and  one  of  which  must  be  possessed  by  every 
positive  rational  number.  Further,  suppose  that  every  such 
number  which  possesses  P  is  less  than  any  such  number  which 
possesses  Q.  Thus  P  might  be  the  property  '  x2  <  2 '  and  Q  the 
property  '  x2  >  2/  Then  we  call  the  numbers  which  possess  P  the 
lower  or  left-hand  class  L  and  those  which  possess  Q  the  upper  or 

*  In  these  sections  I  have  borrowed  freely  from  Appendix  I  of  Bromwich's 
Infinite  Series. 


12  REAL  VARIABLES  [i 

right-hand  class  R.  In  general  both  classes  will  exist ;  but  it  may- 
happen  in  special  cases  that  one  is  non-existent  and  that  every 
number  belongs  to  the  other.  This  would  obviously  happen,  for 
example,  if  P  (or  Q)  were  the  property  of  being  rational,  or  of 
being  jjositive.  For  the  present,  however,  we  shall  confine 
ourselves  to  cases  in  which  both  classes  do  exist ;  and  then  it 
follows,  as  in  §  4,  that  we  can  find  a  member  of  L  and  a  member 
of  R  whose  difference  is  as  small  as  we  please. 

In  the  particular  case  which  we  considered  in  §  4,  L  had  no 
greatest  member  and  R  no  least.  This  question  of  the  existence 
of  greatest  or  least  members  of  the  classes  is  of  the  utmost  im- 
portance. We  observe  first  that  it  is  impossible  in  any  case  that 
L  should  have  a  greatest  member  and  R  a  least.  For  if  I  were 
the  greatest  member  of  L,  and  r  the  least  of  R,  so  that  I  <  f,  then 
\  (I  +  r)  would  be  a  positive  rational  number  lying  between  I  and 
r,  and  so  could  belong  neither  to  L  nor  to  R ;  and  this  contradicts 
our  assumption  that  every  such  number  belongs  to  one  class  or  to 
the  other.  This  being  so,  there  are  but  three  possibilities,  which 
are  mutually  exclusive.  Either  (i)  L  has  a  greatest  member  I,  or 
(ii)  R  has  a  least  member  r,  or  (iii)  L  has  no  greatest  member  and 
R  no  least. 

The  section  of  §  4  gives  an  example  of  the  last  possibility.  An  example 
of  the  first  is  obtained  by  taking  P  to  be  '  x2  <  1 '  and  Q  to  be  '  x2  >  1 ' ; 
here  l—\.  If  P  is  (x2  <  1 '  and  Q  is  l  x2  >  1,'  we  have  an  example  of  the 
second  possibility,  with  r=l.  It  should  be  observed  that  we  do  not  obtain 
a  section  at  all  by  taking  P  to  be  '  x2  <  1 '  and  Q  to  be  '  x2>  1 ' ;  for  the  special 
number  1  escapes  classification  (cf.  Ex.  in.  5). 

7.  Irrational  numbers  (continued).  In  the  first  two  cases 
we  say  that  the  section  corresponds  to  a  positive  rational  number 
a,  which  is  I  in  the  one  case  and  r  in  the  other.  Conversely,  it  is 
clear  that  to  any  such  number  a  corresponds  a  section  which 
we  shall  denote  by  a*.  For  we  might  take  P  and  Q  to  be  the 
properties  expressed  by 

x  ^  a,     x  >  a 

respectively,  or  by  x  <  a  and  x  ^  a.  In  the  first  case  a  would  be 
the  greatest  member  of  L,  and  in  the  second  case  the  least  member 

*  It  will  be  convenient  to  denote  a  section,  corresponding  to  a  rational  number 
denoted  by  an  English  letter,  by  the  corresponding  Greek  letter. 


6-8]  REAL   VARIABLES  13 

of  R.  There  are  in  fact  just  two  sections  corresponding  to  any 
positive  rational  number.  In  order  to  avoid  ambiguity  we  select 
one  of  them  ;  let  us  select  that  in  which  the  number  itself  belongs 
to  the  upper  class.  In  other  words,  let  us  agree  that  we  will  consider 
only  sections  in  which  the  lower  class  L  has  no  greatest  number. 

There  being  this  correspondence  between  the  positive  rational 
numbers  and  the  sections  denned  by  means  of  them,  it  would  be 
perfectly  legitimate,  for  mathematical  purposes,  to  replace  the 
numbers  by  the  sections,  and  to  regard  the  symbols  which  occur 
in  our  formulae  as  standing  for  the  sections  instead  of  for  the 
numbers.  Thus,  for  example,  a  >  a'  would  mean  the  same  as 
a  >  a,  if  a  and  a  are  the  sections  which  correspond  to  a  and  a. 

But  when  we  have  in  this  way  substituted  sections  of  rational 
numbers  for  the  rational  numbers  themselves,  we  are  almost  forced 
to  a  generalisation  of  our  number  system.  For  there  are  sections 
(such  as  that  of  §  4)  which  do  not  correspond  to  any  rational 
number.  The  aggregate  of  sections  is  a  larger  aggregate  than  that 
of  the  positive  rational  numbers ;  it  includes  sections  corresponding 
to  all  these  numbers,  and  more  besides.  It  is  this  fact  which  we 
make  the  basis  of  our  generalisation  of  the  idea  of  number.  We 
accordingly  frame  the  following  definitions,  which  will  however  be 
modified  in  the  next  section,  and  must  therefore  be  regarded  as 
temporary  and  provisional. 

A  section  of  the  positive  rational  numbers,  in  which  both  classes 
exist  and  the  lower  class  has  no  greatest  member,  is  called  a 
positive  real  number. 

A  positive  real  number  which  does  not  correspond  to  a  positive 
rational  number  is  called  a  positive  irrational  number. 

8.  Real  numbers.  We  have  confined  ourselves  so  far  to 
certain  sections  of  the  positive  rational  numbers,  which  we  have 
agreed  provisionally  to  call  'positive  real  numbers.'  Before  we 
frame  our  final  definitions,  we  must  alter  our  point  of  view  a 
little.  We  shall  consider  sections,  or  divisions  into  two  classes, 
not  merely  of  the  positive  rational  numbers,  but  of  all  rational 
numbers,  including  zero.  We  may  then  repeat  all  that  we  have 
said  about  sections  of  the  positive  rational  numbers  in  §§  6,  7, 
merely  omitting  the  word  positive  occasionally. 


14  REAL   VARIABLES  [i 

Definitions.  A  section  of  the  rational  numbers,  in  which  both 
classes  exist  and  the  lower  class  has  no  greatest  member,  is  called 
a  real  number,  or  simply  a  number. 

A  real  number  ivhich  does  not  correspond  to  a  rational  number 
is  called  an  irrational  number. 

If  the  real  number  does  correspond  to  a  rational  number,  we 
shall  use  the  term  '  rational '  as  applying  to  the  real  number  also. 

The  term  'rational  number'  will,  as  a  result  of  our  definitions,  be 
ambiguous;  it  may  mean  the  rational  number  of  §  1,  or  the  corresponding 
real  number.  If  we  say  that  h  >  i -,  we  may  be  asserting  either  of  two  different 
propositions,  one  a  proposition  of  elementary  arithmetic,  the  other  a  proposition 
concerning  sections  of  the  rational  numbers.  Ambiguities  of  this  kind  are 
common  in  mathematics,  and  are  perfectly  harmless,  since  the  relations 
between  different  propositions  are  exactly  the  same  whichever  interpretation 
is  attached  to  the  propositions  themselves.  From  ?>>A  and  -jj>j  we  can 
infer  5  >  j  ;  the  inference  is  in  no  way  affected  by  any  doubt  as  to  whether 
\,  \,  and  j  are  arithmetical  fractions  or  real  numbers.  Sometimes,  of  course, 
the  context  in  which  {e.g.)  '  \ '  occurs  is  sufficient  to  fix  its  interpretation. 
When  we  say  (see  §  9)  that  ^<\Z(sX  we  niust  mean  by  '£'  the  real  number  \. 

The  reader  should  observe,  moreover,  that  no  particular  logical  importance 
is  to  be  attached  to  the  precise  form  of  definition  of  a  '  real  number '  that  we 
have  adopted.  We  defined  a  '  real  number '  as  being  a  section,  i.e.  a  pair  of 
classes.  We  might  equally  well  have  defined  it  as  being  the  lower,  or  the 
upper,  class  ;  indeed  it  would  be  easy  to  define  an  infinity  of  classes  of 
entities  each  of  which  would  possess  the  properties  of  the  class  of  real 
numbers.  What  is  essential  in  mathematics  is  that  its  symbols  should  be 
capable  of  some  interpretation ;  generally  they  are  capable  of  many,  and 
then,  so  far  as  mathematics  is  concerned,  it  does  not  matter  which  we  adopt. 
Mr  Beitrand  Kussell  has  said  that  'mathematics  is  the  science  in  which 
we  do  not  know  what  we  are  talking  about,  and  do  not  care  whether  what 
we  say  about  it  is  true',  a  remark  which  is  expressed  in  the  form  of  a 
paradox  but  which  in  reality  embodies  a  number  of  important  truths.  It 
would  take  too  long  to  analyse  the  meaning  of  Mr  Russell's  epigram  in  detail, 
but  one  at  any  rate  of  its  implications  is  this,  that  the  symbols  of  mathe- 
matics are  capable  of  varying  interpretations,  and  that  we  are  in  general  at 
liberty  to  adopt  whichever  we  prefer. 

There  are  now  three  cases  to  distinguish.  It  may  happen  that 
all  negative  rational  numbers  belong  to  the  lower  class  and  zero 
and  all  positive  rational  numbers  to  the  upper.  We  describe 
this  section  as  the  real  number  zero.  Or  again  it  may  happen 
that  the  lower  class  includes  some  positive  numbers.   Such  a  section 


8,  9]  REAL   VARIABLES  15 

we  describe  as  a  positive  real  number.  Finally  it  may  happen 
that  some  negative  numbers  belong  to  the  upper  class.  Such 
a  section  we  describe  as  a  negative  real  number*. 

The  difference  between  our  present  definition  of  a  positive  real  number  a 
and  that  of  §  7  amounts  to  the  addition  to  the  lower  class  of  zero  and  all  the 
negative  rational  numbers.  An  example  of  a  negative  real  number  is  given 
by  taking  the  property  P  of  §  6  to  be  .r  +  l<0  and  Q  to  be  .r+1^0. 
This  section  plainly  corresponds  io  the  negative  rational  number  - 1.  If  we 
took  P  to  be  x3<—2  and  Q  to  be  a^>  -  2,  we  should  obtain  a  negative  real 
number  which  is  not  rational. 

9.     Relations  of  magnitude  between  real  numbers.     It 

is  plain  that,  now  that  we  have  extended  our  conception  of 
number,  we  are  bound  to  make  corresponding  extensions  of  our 
conceptions  of  equality,  inequality,  addition,  multiplication,  and  so 
on.  We  have  to  show  that  these  ideas  can  be  applied  to  the  new 
numbers,  and  that,  when  this  extension  of  them  is  made,  all  the 
ordinary  laws  of  algebra  retain  their  validity,  so  that  we  can 
operate  with  real  numbers  in  general  in  exactly  the  same  way 
as  with  the  rational  numbers  of  §  1.  To  do  all  this  systematically 
would  occupy  a  considerable  space,  and  we  shall  be  content  to 
indicate  summarily  how  a  more  systematic  discussion  would 
proceed. 

We  denote  a  real  number  by  a  Greek  letter  such  as  a,  /3,  y,  . . . ; 
the  rational  numbers  of  its  lower  and  upper  classes  by  the  corre- 
sponding English  letters  a,  A;  b,  B;  c,  C;  ....  The  classes  them- 
•  selves  we  denote  by  (a),  (A),  ..... 

If  a  and  ft  are  two  real  numbers,  there  are  three  possibilities  : 

(i)  every  a  is  a  6  and  every  A&B\  in  this  case  (a)  is  identical 
with  (b)  and  (A)  with  (B); 

*  There  are  also  sections  in  which  every  number  belongs  to  the  lower  or  to 
the  upper  class.  The  reader  may  be  tempted  to  ask  why  we  do  not  regard  these 
sections  also  as  defining  numbers,  which  we  might  call  the  real  numbers  positive 
and  negative  infinity. 

There  is  no  logical  objection  to  such  a  procedure,  but  it  proves  to  be  incon- 
venient in  practice.  The  most  natural  definitions  of  addition  and  multiplication  do 
not  work  in  a  satisfactory  way.  Moreover,  for  a  beginner,  the  chief  difficulty  in  the 
elements  of  analysis  is  that  of  learning  to  attach  precise  senses  to  phrases  containing 
the  word  '  infinity ';  and  experience  seems  to  show  that  he  is  likely  to  be  confused  by 
any  addition  to  their  number. 


1G  KEAL   VARIABLES  [i 

(ii)  every  a  is  a  b,  but  not  all  A's  are  B's ;  in  this  case  (a)  is 
a  proper  part  of  (&)*,  and  (B)  a  proper  part  of  (A) ; 

(iii)    every  A  is  a  B,  but  not  all  a's  are  b's. 

These  three  cases  may  be  indicated  graphically  as  in  Fig.  4. 

In  case  (i)  we  write  a  =  6,  in  case  (ii)  a  <  6,  and  in  case 
(iii)  a>  8.     It  is  clear  that,  when 

a  and  8  are  both  rational,  these  ? (i) 

definitions  agree  with  the  ideas  of 

equality  and   inequality  between  + ' ("J 

rational  numbers  which  we  began 

by  taking  for  granted;   and  that  i + (iii) 

any  positive  number    is    greater  Fig.  4. 

than  any  negative  number. 

It  will  be  convenient  to  define  at  this  stage  the  negative  —  a 
ot  a  positive  number  a.  If  (a),  (A)  are  the  classes  which  consti- 
tute a,  we  can  define  another  section  of  the  rational  numbers  by 
putting  all  numbers  —  A  in  the  lower  class  and  all  numbers  —  a 
in  the  upper.  The  real  number  thus  defined,  which  is  clearly 
negative,  we  denote  by  —  a.  Similarly  we  can  define  —  a  when  a 
is  negative  or  zero ;  if  a  is  negative,  —  a  is  positive.  It  is  plain 
also  that  —  (—  0)  =  a.  Of  the  two  numbers  a  and  —  a  one  is  always 
positive  (unless  o  =  0).  The  one  which  is  positive  we  denote  by 
I  a  I  and  call  the  modulus  of  or. 

Examples  IV.    1.     Prove  that  0  =-  0. 

2.  Prove  that  /3  =  a,  /3<a,  or  /3>a  according  as  a=/3,  a>/3,  or  a</3. 

3.  If  a  =  j3  and  /3  =  y,  then  a  =  y. 

4.  If  a  ^  ft  /3<y,  or  a</3,  /3  ^  y,  then  a<y. 

5.  Prove  that  — /3=  —a   -/3<-a,  or  —  j3>  —a,  according  as  a  =  /3,  a</3, 
or  a>/3. 

6.  Prove  that  a>0  if  a  is  positive,  and  a<0  if  a  is  negative. 

7.  Prove  that  a  ^  |  o  | . 

8.  Prove  that  1<  <J2  <  ^3  <  2. 

9.  Prove  that,  if  a  and  £  are  two  different  real  numbers,  we  can  always 
find  an  infinity  of  rational  numbers  lying  between  a  and  /3. 

[All  these  results  are  immediate  consequences  of  our  definitions.] 

*  I.e.  is  included  in  but  not  identical  with  (&). 


9,  10]  REAL   VARIABLES  17 

10.     Algebraical  operations  with  real  numbers.    We  now 

proceed  to  define  the  meaning  of  the  elementary  algebraical  opera- 
tions such  as  addition,  as  applied  to  real  numbers  in  general. 

(i)  Addition.  In  order  to  define  the  sum  of  two  numbers 
a  and  ft,  we  consider  the  following  two  classes :  (i)  the  class  (c) 
formed  by  all  sums  c  =  a  +  b,  (ii)  the  class  (G)  formed  by  all  sums 
G  =  A  +  B.     Plainly  c  <  G  in  all  cases. 

Again,  there  cannot  be  more  than  one  rational  number  which 
does  not  belong  either  to  (c)  or  to  (0).  For  suppose  there  were 
two,  say  r  and  s,  and  let  s  be  the  greater.  Then  both  r  and  s 
must  be  greater  than  every  c  and  less  than  every  G;  and  so  G  —  c 
cannot  be  less  than  s  —  r.     But 

C-c  =  (A-a)  +  (B-b)] 

and  we  can  choose  a,  b,  A,  B  so  that  both  A  —  a  and  B—b 
are  as  small  as  we  like ;  and  this  plainly  contradicts  our 
hypothesis. 

If  every  rational  number  belongs  to  (c)  or  to  (0),  the  classes  (c), 
(G)  form  a  section  of  the  rational  numbers,  that  is  to  say,  a  number 
7.  If  there  is  one  which  does  not,  we  add  it  to  (G).  We  have 
now  a  section  or  real  number  7,  which  must  clearly  be  rational, 
since  it  corresponds  to  the  least  member  of  (G).  In  any  case 
■we  call  7  the  sum  of  a  and  ft,  and  write 

7  =  a  +  ft. 

If  both  a  and  /3  are  rational,  they  are  the  least  members  of  the  upper 
classes  (A)  and  (B).  In  this  case  it  is  clear  that  a+0  is  the  least  member 
of  (C%  so  that  our  definition  agrees  with  our  previous  ideas  of  addition. 

(ii)    Subtraction.     We  define  a  —  ft  by  the  equation 

«-/3  =  «+(-/3). 

The  idea  of  subtraction  accordingly  presents  no  fresh  difficulties. 

Examples  V.    1.    Prove  that  a  +  ( -  a)  =  0. 

2.  Prove  that  a  +  0=0  +  a  =  a. 

3.  Prove  that  a  +  /3  =  /3  +  a.  [This  follows  at  once  from  the  fact  that  the 
classes  (a+b)  and  (&  +  a),  or  (A+B)  and  (B+A),  are  the  same,  since,  e.g., 
a-\-b  =  b  +  a  when  a  and  b  are  rational.] 

4.  Prove  that  a  +  0  +  y )  =  (a + /3)  +  y. 

n.  2 


18  REAL   VARIABLES  [i 

5.  Prove  that  a  -  a  =  0. 

6.  Prove  that  a  -  /3=  -  (/3  -  a). 

7.  From  the  definition  of  subtraction,  and  Exs.  4,  1,  and  2  above,  it 
follows  that 

(a-/3)+0  =  {a  +  (-/3)]+/3  =  a  +  {(-/3)  +  /3}  =  a  +  O  =  a. 
We  might  therefore  define  the  difference  a-j3  =  y  by  the  equation  y+/3  =  a. 

8.  Prove  that  a-0-y)  =  a-|3  +  y. 

9.  Give  a  definition  of  subtraction  which  does  not  depend  upon  a  previous 
definition  of  addition.  [To  define  y=a  —  /3,  form  the  classes  (c),  (C)  for  which 
c  =  a  —  B,  C=A-b.  It  is  easy  to  show  that  this  definition  is  equivalent  to 
that  which  we  adopted  in  the  text.] 

10.  Prove  that 

||a|-|/9||Sja±|8|S|a|+|0|. 

11.  Algebraical  operations  with  real  numbers  {con- 
tinued), (iii)  Multiplication.  When  we  come  to  multiplication, 
it  is  most  convenient  to  confine  ourselves  to  'positive  numbers 
(among  which  we  may  include  0)  in  the  first  instance,  and  to  go 
back  for  a  moment  to  the  sections  of  positive  rational  numbers 
only  which  we  considered  in  §§  4 — 7.  We  may  then  follow  practi- 
cally the  same  road  as  in  the  case  of  addition,  taking  (c)  to  be  (ab) 
and  (C)  to  be  (AB).  The  argument  is  the  same,  except  when  we 
are  proving  that  all  rational  numbers  with  at  most  one  exception 
must  belong  to  (c)  or  (G).  This  depends,  as  in  the  case  of  addi- 
tion, on  showing  that  we  can  choose  a,  A,  b,  and  B  so  that  C—  c  is 
as  small  as  we  please.     Here  we  use  the  identity 

C-c  =  AB-ab  =  (A  -a)  B  +a(B- b). 

Finally  we  include  negative  numbers  within  the  scope  of  our 
definition  by  agreeing  that,  if  a  and  /3  are  positive,  then 

(-a)/3  =  -aft     a(-/3)  =  -«&     (-a)(-0)  =  a0. 

(iv)  Division.  In  order  to  define  division,  we  begin  by  de- 
fining the  reciprocal  1/a  of  a  number  a  (other  than  zero).  Con- 
fining ourselves  in  the  first  instance  to  positive  numbers  and 
sections  of  positive  rational  numbers,  we  define  the  reciprocal  of  a 
positive  number  a  by  means  of  the  lower  class  (1/A)  and  the  upper 
class  (1/a).  We  then  define  the  reciprocal  of  a  negative  number 
—  a  by  the  equation  l/(— a)  =  —  (1/a).  Finally  we  define  a//3  by 
the  equation 

«//3  =  ax(l//3). 


10-13]  REAL  VARIABLES 


19 


We  are  then  in  a  position  to  apply  to  all  real  numbers,  rational 
or  irrational,  the  whole  of  the  ideas  and  methods  of  elementary 
algebra.  Naturally  we  do  not  propose  to  carry  out  this  task  in 
detail.  It  will  be  more  profitable  and  more  interesting  to  turn 
our  attention  to  some  special,  but  particularly  important,  classes 
of  irrational  numbers. 

Examples  VI.  Prove  the  theorems  expressed  by  the  followin°- 
formulae  : 

1.     ax0  =  0xa  =  0.  2.     axl  =  lxa  =  a.  3.     ax(l/a)  =  l. 

4.     a/3  = /3a.  o.     a  (/3y)  =  (a/3)  y.  6.     a  (j8  +  y)  =  a/3  +  ay. 

7.     (a  +  (S)y  =  ay  +  j3y.  8.     |  a/3  |  =  |  a  |  |/3  |. 

12.  The  number  ^/2.  Let  us  now  return  for  a  moment  to 
the  particular  irrational  number  which  we  discussed  in  §§  4 — 5. 
We  there  constructed  a  section  by  means  of  the  inequalities 
x2  <  2,  x2  >  2.  This  was  a  section  of  the  positive  rational  numbers 
only ;  but  we  replace  it  (as  was  explained  in  §  8)  by  a  section  of 
all  the  rational  numbers.  We  denote  the  section  or  number  thus 
denned  by  the  symbol  *J2. 

The  classes  by  means  of  which  the  product  of  V2  by  itself  is 
defined  are  (i)  (aa),  where  a  and  a'  are  positive  rational  numbers 
whose  squares  are  less  than  2,  (ii)  {AA'),  where  A  and  A'  are 
positive  rational  numbers  whose  squares  are  greater  than  2.  These 
classes  exhaust  all  positive  rational  numbers  save  one,  which  can 
only  be  2  itself.     Thus 

(V2)2=V2\/2  =  2. 
Again 

(-  V2)2  =  (-  V2)  (-  V2)  =  V2  V2  =  (V2)2  =  2. 

Thus  the  equation  x-=2  has  the  two  roots  \]2  and  —  \]2.  Similarly 
we  could  discuss  the  equations  x*  =  3,  x3  =  7,  ...  and  the  corre- 
sponding irrational  numbers  \/3,  —  V3,  ^7,.... 

13.  Quadratic  surds.  A  number  of  the  form  +  \/a,  where 
a  is  a  positive  rational  number  which  is  not  the  square  of  another 
rational  number,  is  called  a  pure  quadratic  surd.  A  number  of 
the  form  a  ±  \/b,  where  a  is  rational,  and  *fb  is  a  pure  quadratic 
surd,  is  sometimes  called  a  mixed  quadratic  surd. 

2—2 


20  REAL   VARIABLES  [i 

The  two  numbers  a±Jb  are  the  roots  of  the  quadratic  equation 

.<v2-2cLv  +  a2-b  =  0. 

Conversely,  the  equation  x2  +  2px  +  q=0,  where  p  and  q  are  rational,  and 
p2-q>0,  has  as  its  roots  the  two  quadratic  surds  -p±^(p2  —  q)- 

The  only  kind  of  irrational  numbers  whose  existence  was 
suggested  by  the  geometrical  considerations  of  §  3  are  these 
quadratic  surds,  pure  and  mixed,  and  the  more  complicated 
irrationals  which  may  be  expressed  in  a  form  involving  the 
repeated  extraction  of  square  roots,  such  as 

V2  +  V(2  +  V2)  +  V{2  +  \/(2  +  \/2)}. 

It  is  easy  to  construct  geometrically  a  line  whose  length  is 
equal  to  any  number  of  this  form,  as  the  reader  will  easily  see  for 
himself.  That  irrational  numbers  of  these  kinds  only  can  be  con- 
structed by  Euclidean  methods  (i.e.  by  geometrical  constructions 
with  ruler  and  compasses)  is  a  point  the  proof  of  which  must 
be  deferred  for  the  present*  This  property  of  quadratic  surds 
makes  them  especially  interesting. 

Examples  VII.     1.     Give  geometrical  constructions  for 

V2     s/(2  +  V2),    V(2+V(2  +  V2)}. 

2.  The  quadratic  equation  ax2  +  2bx+c  —  0  has  two  real  roots +  if 
b2-ac>0.  Suppose  a,  b  c  rational.  Nothing  is  lost  by  taking  all  three 
to  be  integers,  for  we  can  multiply  the  equation  by  the  least  common 
multiple  of  their  denominators. 

The  reader  will  remember  that  the  roots  are  {-b±s/(b2-ac)}/a.  It  is 
easy  to  construct  these  lengths  geometrically,  first  constructing  sf(b'2-ac). 
A  much  more  elegant,  though  less  straightforward,  construction  is  the 
following. 


*  See  Ch.  II,  Misc.  Exs.  22. 

t  I.e.  there  are  two  values  of  X  for  which  ax2  +  2bx  +  c  =  0.  If  b2-ac<0  there 
are  no  such  values  of  x.  The  reader  will  remember  that  in  books  on  elementary 
algebra  the  equation  is  said  to  have  two  '  complex '  roots.  The  meaning  to  be 
attached  to  this  statement  will  be  explained  in  Ch.  III. 

When  b2  =  ac  the  equation  has  only  one  root.  For  the  sake  of  uniformity 
it  is  generally  said  in  this  case  to  have  '  two  equal '  roots,  but  this  is  a  mere 
convention. 


13,  14] 


REAL- VARIABLES 


21 


Draw  a  circle  of  unit  radius,  a  diameter  PQ,  and  the  tangents  at  the  ends 
of  the  diameters. 


Fig.  5- 

Take  PP'  =  -2a/b  and  QQ'  =  —c/2b,  having  regard  to  sign*.  Join  P'Q', 
cutting  the  circle  in  M  and  N.  Draw  PM  and  PX,  cutting  QQ'  in  X  and  Y. 
Then  QX  and  QY  are  the  roots  of  the  equation  with  their  proper  signs  t. 

The  proof  is  simple  and  we  leave  it  as  an  exercise  to  the  reader. 
Another,  perhaps  even  simpler,  construction  is  the  following.  Take  a  line 
AB  of  unit  length.  Draw  BC=  -2b/a  perpendicular  to  AB,  and  CD=c/a 
peipendicular  to  BC  and  in  the  same  direction  as  BA.  On  AD  as  diameter 
describe  a  circle  cutting  BC  in  X  and  Y.     Then  BX  and  BY  are  the  roots. 

3.  If  ac  is  positive  PP'  and  QQ'  will  be  drawn  in  the  same  direction. 
Verify  that  P'Q'  will  not  meet  the  circle  if  b'1<ac,  while  if  b2  =  ac  it  will  be 
a  tangent.  Verify  also  that  if  b2  =  ac  the  circle  in  the  second  construction 
will  touch  BC. 


4.      Prove  that 


>J(pq)  =  s/pxJq,   Kf(p2q)  =p  \fq- 


14.     Some  theorems  concerning  quadratic  surds.     Two 

pure  quadratic  surds  are  said  to  be  similar  if  they  can  be  ex- 
pressed as  rational  multiples  of  the  same  surd,  and  otherwise  to  be 
dissimilar.     Thus 

V8=2V2,     V¥=fV2, 

and  so  *JS,  y^  are  similar  surds.  On  the  other  hand,  if  M  and  N 
are  integers  which  have  no  common  factor,  and  neither  of  which 
is  a  perfect  square,  *JM  and  \fN  are  dissimilar  surds.  For  suppose, 
if  possible, 

q  V  u  s  V  u 

where  all  the  letters  denote  integers. 

*  The  figure  is  drawn  to  suit  the  case  in  which  b  and  c  have  the  same  and  a 
the  opposite  sign.     The  reader  should  draw  figures  for  other  cases. 

f  I  have  taken  this  construction  from  Klein's  Lecons  sur  certaines  questions  dc 
geometrie  elementaire  (French  translation  by  J.  Griess,  Paris,  1896). 


22  REAL   VARIABLES  [l 

Then  *JMN  is  evidently  rational,  and  therefore  (Ex.  II.  3) 
integral.  Thus  MN  =  P2,  where  P  is  an  integer.  Let  a,  b,  c,  ... 
be  the  prime  factors  of  P,  so  that 

MN=a2ab^c-y  ..., 

where  a,  ft,  y, ...  are  positive  integers.  Then  MN  is  divisible  by 
a?a,  and  therefore  either  (1)  M  is  divisible  by  a2a,  or  (2)  N  is 
divisible  by  a2a,  or  (3)  M  and  N  are  both  divisible  by  a.  The  last 
case  may  be  ruled  out,  since  M  and  N  have  no  common  factor. 
This  argument  may  be  applied  to  each  of  the  factors  a2a,  623,  c2y, . . . , 
so  that  M  must  be  divisible  by  some  of  these  factors  and  N  by 
the  remainder.    Thus 

M  =  P2,    N=P£, 

where  Pa2  denotes  the  product  of  some  of  the  factors  a2a,  b2fi,  c2V,  ... 
and  P22  the  product  of  the  rest.  Hence  M  and  N  are  both  perfect 
squares,  which  is  contrary  to  our  hypothesis. 

Theorem.    If  A,  B,  G,  D  are  rational  and 

A+*/B=C  +  s/D, 

then  either  (i)  A  =  C,  B  =  D  or  (ii)  B  and  D  are  loth  squares  of 
rational  numbers. 

For  B  —  D  is  rational,  and  so  is 

>JB-jD  =  G-A. 

If  B  is  not  equal  to  D  (in  which  case  it  is  obvious  that  A  is  also 
equal  to  C),  it  follows  that 

VP  +  *JD  =  (B  -  D)/WB  -  VP) 

is  also  rational.     Hence  *JB  and  *JD  are  rational. 

Corollary.    If  A  +  */B  =  C  +  */D,   then    A-*/B  =  C-^D 
(unless  \JB  and  \JD  are  both  rational). 

Examples  VIII.     1.     Trove  ab  initio  that  J2  and  */3  are  not  similar 
surds. 

2.  Prove  that  *Ja  and  J(lja),  where  a  is  rational,  are   similar    surds 
(unless  both  are  rational). 

3.  If  a  and  b  are  rational,  then  Ja  +  Jb  cannot  be  rational  unless  sfa  and 
*Jb  are  rational.     The  same  is  true  of  J  a  —  *]b,  unless  a  =  b. 


14,  15]  REAL   VARIABLES  23 

4.  If  JA+JB^JC+JD, 

then  either  (a)  ^  =  Cand  B=D,  or  (b)  A=D  and  B  =  C,  or  (c)  J  A,  JB,  <JC, 
*JD  are  all  rational  or  all  similar  surds.  [Square  the  given  equation  and 
apply  the  theorem  above.] 

5.  Neither  (a  +  s/b)3  nor  (a  -  *fb)3  can  be  rational  unless  >Jb  is  rational. 

6.  Prove  that  if  x=p  +  y/q,  where  p  and  q  are  rational,  then  xm,  where 
m  is  any  integer,  can  be  expressed  in  the  form  P+  QJq,  where  P  and  Q 
are  rational.     For  example, 

(P  +  Jq?  =p2  +  q  +  2p  Jq,     (p  +  V?)3  =p3  +  Spq  +  (3p*  +  q)  Jq. 
Deduce  that  any  polynomial  in  x  with  rational  coefficients  (i.e.  any  expression 
of  the  form 

a0tfl+a1xn-1  +  ...+an, 

where  a0,  ...  an  are  rational  numbers)  can  be  expressed  in  the  forni  P  +  QJq. 

7.  If  a  +  sjb,  where  b  is  not  a  perfect  square,  is  the  root  of  an  algebraical 
equation  with  rational  coefficients,  then  a  —  ^Jb  is  another  root  of  the  same 
equation. 

8.  Express  lj(p+Kfq)  in  the  form  prescribed  in  Ex.  6.  [Multiply 
numerator  and  denominator  by  p  -  sfq.] 

9.  Deduce  from  Exs.  6  and  8  that  any  expression  of  the  form  G  (x)/H  (x), 
where  G(x)  and  H(x)  are  polynomials  in  x  with  rational  coefficients,  can  be 
expressed  in  the  form  P  +  Q-Jq,  where  P  and  Q  are  rational. 

10.  If  p,  q,  and  p2-q  are  positive,  we  can  express  s/(p+\fq)  in  the  form 
•J®  +  <Jyi  where 

»-i  {p+s/(p2-q)},  y=h{p-s?(p2-<i)}' 

1 1 .  Determine  the  conditions  that  it  may  be  possible  to  express  J(p + Jq), 
where  p  and  q  are  rational,  in  the  form  slx  +  \ly,  where  x  and  y  are  rational. 

12.  If  a2  -  b  is  positive,  the  necessary  and  sufficient  conditions  that 

sj{a-hslb)  +  yl{a-s]b) 
should  be  rational  are  that  cp-b  and  \ {a  +  s/(a2 - b)}  should  both  be  squares 
of  rational  numbers. 

15.  The  continuum.  The  aggregate  of  all  real  numbers, 
rational  and  irrational,  is  called  the  arithmetical  continuum. 

It  is  convenient  to  suppose  that  the  straight  line  A  of  §  2 
is  composed  of  points  corresponding  to  all  the  numbers  of  the 
arithmetical  continuum,  and  of  no  others*.     The  points  of  the 

*  This  supposition  is  merely  a  hypothesis  adopted  (i)  because  it  suffices  for  the 
purposes  of  our  geometry  and  (ii)  because  it  provides  us  with  convenient  geometrical 
illustrations  of  analytical  processes.  As  we  use  geometrical  lauguage  only  for 
purposes  of  illustration,  it  is  not  part  of  our  business  to  study  the  foundations 
of  geometry. 


24  REAL   VARIABLES  [i 

line,  the  aggregate  of  which  may  be  said  to  constitute  the  linear 
continuum,  then  supply  us  with  a  convenient  image  of  the 
arithmetical  continuum. 

We  have  considered  in  some  detail  the  chief  properties  of  a 
few  classes  of  real  numbers,  such,  for  example,  as  rational  numbers 
or  quadratic  surds.  We  add  a  few  further  examples  to  show  how 
very  special  these  particular  classes  of  numbers  are,  and  how,  to 
put  it  roughly,  they  comprise  only  a  minute  fraction  of  the  infinite 
variety  of  numbers  which  constitute  the  continuum. 

(i)     Let  us  consider  a  more  complicated  surd  expression  such  as 
s  =  v/(4  +  v/15)  +  4/(4-v/15). 
Our  argument  for  supposing  that  the  expression  for  z  has  a  meaning  might  be 
as  follows.    We  first  show,  as  in  §12,  that  there  is  a  number  ?/  =  v/15  such  that 
y2  =  15,  and  we  can  then,  as  in  §  10,  define  the  numbers  4  +  ^/15,  4  — v/15. 
Now  consider  the  equation  in  zx, 

0!3=4  +  v/15. 
The  right-hand  side  of  this  equation  is  not  rational :  but  exactly  the  same 
reasoning  which  leads  us  to  suppose  that  there  is  a  real  number  x  such  that 
x3  =  2  (or  any  other  rational  number)  also  leads  us  to  the  conclusion  that  there 
is  a  number  zx  such  that  213=4  +  v/15.  We  thus  define  z1  =  £/(4+*J15),  and 
similarly  we  can  define  22=^(4—^15) ;  and  then,  as  in  §  10,  we  define  z=zx+z2. 

Now  it  is  easy  to  verify  that 

z3=3z  +  8. 

And  we  might  have  given  a  direct  proof  of  the  existence  of  a  unique  number 
z  such  that  z3=3z+8.  It  is  easy  to  see  that  there  cannot  be  two  such 
numbers.  For  if  zl3  =  3zl  +  8  and  z23  =  3z2  +  8,  we  find  on  subtracting  and 
dividing  by  Z\-z2  that  zl2+zlz2+z22=3.  But  if  zx  and  z2  are  positive  zl3>8, 
z23>8  and  therefore  zx>2,  z2>2,  z12+z1z2  +  z22>  12,  and  so  the  equation 
just  found  is  impossible.  And  it  is  easy  to  see  that  neither  zx  nor  z2  can 
be  negative.  For  if  zx  is  negative  and  equal  to  —  f,  f  is  positive  and 
£3-3f+8  =  0,  or  3-£2  =  8/(.  Hence  3-£2>0,  and  so  f<2.  But  then 
8/£>4,  and  so  8/f  cannot  be  equal  to  3-  f2,  which  is  less  than  3. 

Hence  there  is  at  most  one  z  such  that  z3  =  3z  +  8.  And  it  cannot  be 
rational.  For  any  rational  root  of  this  equation  must  be  integral  and  a 
factor  of  8  (Ex.  n.  3),  and  it  is  easy  to  verify  that  no  one  of  1,  2,  4,  8  is  a  root. 

Thus  z3  =  3z  +  8  has  at  most  one  root  and  that  root,  if  it  exists,  is  positive 
and  not  rational.  We  can  now  divide  the  positive  rational  numbers  x  into 
two  classes  L,  R  according  as  x3  <  3x  +  8  or  x3  >  3x  +  8.  It  is  easy  to  see  that 
if  ^3>3o;  +  8  and  y  is  any  number  greater  than  x,  then  also  y3  >  3_y +  8.  For 
suppose  if  possible  y3^3y  +  8.  Then  since  #3>3.r+8  we  obtain  on  sub- 
tracting y3  -  x3  <  3  (y  -  x),  or  y1  +  xy  +  x2  <  3,  which  is  impossible;  for  y  is 


15]  REAL   VARIABLES  25 

positive  and  x>2  (since  x3>8).  Similarly  we  can  show  that  if  x3 < 3x  +  8 
and  y  <  x  then  also  y3  <  3y  +  8. 

Finally,  it  is  evident  that  the  classes  L  and  R  both  exist ;  and  they  form 
a  section  of  the  positive  rational  numbers  or  positive  real  number  z  which 
satisfies  the  equation  z3  =  3z  +  8.  The  reader  who  knows  how  to  solve  cubic 
equations  by  Cardan's  method  will  be  able  to  obtain  the  explicit  expression  of 
z  directly  from  the  equation. 

(ii)  The  direct  argument  applied  above  to  the  equation 
a?  =  3x  +  8  could  be  applied  (though  the  application  would  be 
a  little  more  difficult)  to  the  equation 

x5  =  x  +  16. 

and  would  lead  us  to  the  conclusion  that  a  unique  positive  real 
number  exists  which  satisfies  this  equation.  In  this  case,  how- 
ever, it  is  not  possible  to  obtain  a  simple  explicit  expression 
for  x  composed  of  any  combination  of  surds.  It  can  in  fact 
be  proved  (though  the  proof  is  difficult)  that  it  is  generally 
impossible  to  .find  such  an  expression  for  the  root  of  an  equation 
of  higher  degree  than  4.  Thus,  besides  irrational  numbers  which 
can  be  expressed  as  pure  or  mixed  quadratic  or  other  surds,  or 
combinations  of  such  surds,  there  are  others  which  are  roots  of 
algebraical  equations  but  cannot  be  so  expressed.  It  is  only  in 
very  special  cases  that  such  expressions  can  be  found. 

(iii)  But  even  when  we  have  added  to  our  list  of  irrational 
numbers  roots  of  equations  (such  as  x5  =  x+  16)  which  cannot  be 
explicitly  expressed  as  surds,  we  have  not  exhausted  the  different 
kinds  of  irrational  numbers  contained  in  the  continuum.  Let  us 
draw  a  circle  whose  diameter  is  equal  to  A0A1}  i.e.  to  unity.  It  is 
natural  to  suppose  *  that  the  circumference  of  such  a  circle  has  a 
length  capable  of  numerical  measurement.  This  length  is  usually 
denoted  by  ir.  And  it  has  been  shown f  (though  the  proof  is  un- 
fortunately long  and  difficult)  that  this  number  ir  is  not  the 
root  of  any  algebraical  equation  with  integral  coefficients,  such, 
for  example,  as 

7T2  =  n,       7T3  =  71,       7T5  =  7T  +  ??, 

*  A  proof  will  be  found  in  Ch.  VII. 

t  See  Hobson's  Trigonometry  (3rd  edition),  pp.  305  et  seq.,  or  the  same  writer's 
Squaring  the  Circle  (Cambridge,  1913). 


26  REAL  VARIABLES  [i 

where  n  is  an  integer.  In  this  way  it  is  possible  to  define  a 
number  which  is  not  rational  nor  yet  belongs  to  any  of  the  classes 
of  irrational  numbers  which  we  have  so  far  considered.  And  this 
number  it  is  no  isolated  or  exceptional  case.  Any  number  of  other 
examples  can  be  constructed.  In  fact  it  is  only  special  classes  of 
irrational  numbers  which  are  roots  of  equations  of  this  kind,  just 
as  it  is  only  a  still  smaller  class  which  can  be  expressed  by  means 
of  surds. 

16.  The  continuous  real  variable.  The  'real  numbers' 
may  be  regarded  from  two  points  of  view.  We  may  think  of 
them  as  an  aggregate,  the  'arithmetical  continuum'  defined  in 
the  preceding  section,  or  individually.  And  when  we  think  of 
them  individually,  we  may  think  either  of  a  particular  specified 
number  (such  as  1,  —  £,  \/2,  or  rr)  or  we  may  think  of  any  number, 
an  unspecified  number,  the  number  x.  This  last  is  our  point  of 
view  when  we  make  such  assertions  as  'x  is  a  number',  'x  is  the 
measure  of  a  length',  'x  may  be  rational  or  irrational',  The  x 
which  occurs  in  propositions  such  as  these  is  called  the  continuous 
real  variable :  and  the  individual  numbers  are  called  the  values  of 
the  variable. 

A  'variable',  however,  need  not  necessarily  be  continuous. 
Instead  of  considering  the  aggregate  of  all  real  numbers,  we 
might  consider  some  partial  aggregate  contained  in  the  former 
aggregate,  such  as  the  aggregate  of  rational  numbers,  or  the 
aggregate  of  positive  integers.  Let  us  take  the  last  case.  Then 
in  statements  about  any  positive  integer,  or  an  unspecified  positive 
integer,  such  as  'n  is  either  odd  or  even',  n  is  called  the  variable, 
a  positive  integral  variable,  and  the  individual  positive  integers 
are  its  valuer. 

Naturally  ' x'  and  cn'  are  only  examples  of  variables,  the 
variable  whose  '  field  of  variation '  is  formed  by  all  the  real 
numbers,  and  that  whose  field  is  formed  by  the  positive  integers. 
These  are  the  most  important  examples,  but  we  have  often  to 
consider  other  cases.  In  the  theory  of  decimals,  for  instance,  we 
may  denote  by  x  any  figure  in  the  expression  of  any  number  as  a 
decimal.  Then  a;  is  a  variable,  but  a  variable  which  has  only  ten 
different  values,  viz.  0,  1,  2,  3,  4,  5,  6,  7,  8,  9.     The  reader  should 


15-17]  REAL   VARIABLES  27 

think  of  other  examples  of  variables  with  different  fields  of  varia- 
tion. He  will  find  interesting  examples  in  ordinary  life :  policeman 
x,  the  driver  of  cab  x,  the  year  x,  the  #th  day  of  the  week.  The 
values  of  these  variables  are  naturally  not  numbers. 

17.  Sections  of  the  real  numbers.  In  §§  4 — 7  we  con- 
sidered '  sections '  of  the  rational  numbers,  i.e.  modes  of  division  of 
the  rational  numbers  (or  of  the  positive  rational  numbers  only) 
into  two  classes  L  and  R  possessing  the  following  characteristic 
properties: 

(i)      that  every  number  of  the  type  considered  belongs  to  one 
and  only  one  of  the  two  classes ; 
(ii)     that  both  classes  exist ; 
(iii)     that  any  member  of  L  is  less  than  any  member  of  R. 

It  is  plainly  possible  to  apply  the  same  idea  to  the  aggregate 
of  all  real  numbers,  and  the  process  is,  as  the  reader  will  find  in 
later  chapters,  of  very  great  importance. 

Let  us  then  suppose*  that  P  and  Q  are  two  properties  which 
are  mutually  exclusive,  and  one  of  which  is  possessed  by  every 
real  number.  Further  let  us  suppose  that  any  number  which 
possesses  P  is  less  than  any  which  possesses  Q.  We  call  the 
numbers  which  possess  P  the  lower  or  left-hand  class  L,  and 
those  which  possess  Q  the  upper  or  right-hand  class  R. 

Thus  P  might  be  x  ^  N/2  and  Q  he  x>  J2.  It  is  important  to  observe 
that  a  pair  of  properties  which  suffice  to  define  a  section  of  the  rational 
numbers  may  not  suffice  to  define  one  of  the  real  numbers.  This  is  so,  for 
example,  with  the  pair  '  x  <  v/2 '  and  *  x  >  J2 '  or  (if  we  confine  ourselves 
to  positive  numbers)  with  '  x2  <  2 '  and  '  x2  >  2 '.  Every  rational  number 
possesses  one  or  other  of  the  properties,  but  not  every  real  number,  since  in 
either  case  >J2  escapes  classification. 

There  are  now  two  possibilities f.  Either  L  has  a  greatest 
member   I,  or  R  has  a  least  member  r,     Both  of  these  events 

*  The  discussion  which  follows  is  in  many  ways  similar  to  that  of  §  6.  We 
have  not  attempted  to  avoid  a  certain  amount  of  repetition.  The  idea  of  a  'section,' 
first  brought  into  prominence  in  Dedekind's  famous  pamphlet  Stetigkeii  and 
irrationale  Zahlen,  is  one  which  can,  and  indeed  must,  be  grasped  by  every  reader 
of  this  book,  even  if  he  be  one  of  those  who  prefer  to  omit  the  discussion  of  the 
notion  of  an  irrational  number  contained  in  §§  6 — 12. 

+  There  were  three  in  §  6. 


28  REAL   VARIABLES  [i 

cannot  occur.  For  if  L  had  a  greatest  member  I,  and  R  a  least 
member  r,  the  number  \{l  +  r)  would  be  greater  than  all  members 
of  L  and  less  than  all  members  of  R,  and  so  could  not  belong  to 
either  class.     On  the  other  hand  one  event  must  occur*. 

For  let  Xj  and  Rx  denote  the  classes  formed  from  L  and  R  by 
taking  only  the  rational  members  of  L  and  R.  Then  the  classes 
Zj  and  Rx  form  a  section  of  the  rational  numbers.  There  are  now 
two  cases  to  distinguish. 

It  may  happen  that  Lx  has  a  greatest  member  a.  In  this  case 
a  must  be  also  the  greatest  member  of  L.  For  if  not,  we  could  find 
a  greater,  say  /3.  There  are  rational  numbers  lying  between  a  and 
/3,  and  these,  being  less  than  j3,  belong  to  L,  and  therefore  to  Lx\ 
and  this  is  plainly  a  contradiction.  Hence  a  is  the  greatest 
member  of  L. 

On  the  other  hand  it  may  happen  that  Lx  has  no  greatest 
member.  In  this  case  the  section  of  the  rational  numbers  formed 
by  Li  and  Rx  is  a  real  number  a.  This  number  a  must  belong- 
to  L  or  to  R.  If  it  belongs  to  L  we  can  shew,  precisely  as  before, 
that  it  is  the  greatest  member  of  L,  and  similarly,  if  it  belongs 
to  R,  it  is  the  least  member  of  R. 

Thus  in  any  case  either  L  has  a  greatest  member  or  R  a 
least.  Any  section  of  the  real  numbers  therefore  'corresponds'  to 
a  real  number  in  the  sense  in  which  a  section  of  the  rational 
numbers  sometimes,  but  not  always,  corresponds  to  a  rational 
number.  This  conclusion  is  of  very  great  importance;  for  it  shows 
that  the  consideration  of  sections  of  all  the  real  numbers  does  not 
lead  to  any  further  genei-alisation  of  our  idea  of  number.  Starting 
from  the  rational  numbers,  we  found  that  the  idea  of  a  section  of 
the  rational  numbers  led  us  to  a  new  conception  of  a  number,  that 
of  a  real  number,  more  general  than  that  of  a  rational  number; 
and  it  might  have  been  expected  that  the  idea  of  a  section  of  the 
real  numbers  would  have  led  us  to  a  conception  more  general  still. 
The  discussion  which  precedes  shows  that  this  is  not  the  c,ase,  and 
that  the  aggregate  of  real  numbers,  or  the  continuum,  has  a  kind 
of  completeness  which  the  aggregate  of  the  rational  numbers 
lacked,  a  completeness  which  is  expressed  in  technical  language 
by  saying  that  the  continuum  is  closed. 

*  This  was  not  the  case  in  §  6. 


17,  18]  REAL   VARIABLES  29 

The  result  which  we  have  just  proved  may  be  stated  as  follows: 

Dedekind's  Theorem.  If  the  real  numbers  are  divided  into 
tivo  classes  L  and  R  in  such  a  way  that 

(i)      every  number  belongs  to  one  or  other  of  the  two  classes, 

(ii)     each  class  contains  at  least  one  number, 

(iii)    any  member  of  L  is  less  than  any  member  of  R, 
then  there  is  a  number  a,  which  has  the  property  that  all  the  numbers 
less  than  it  belong  to  L  and  all  the  numbers  greater  than  it  to  R. 
The  number  a  itself  may  belong  to  either  class. 

In  applications  we  have  often  to  consider  sections  not  of  all  numbers  but 
of  all  those  contained  in  an  interval  (8,  y),  that  is  to  say  of  all  numbers 
x  such  that  j3  g  x  £L  y.  A  '  section '  of  such  numbers  is  of  course  a  division  of 
them  into  two  classes  possessing  the  properties  (i),  (ii),  and  (iii).  Such 
a  section  may  be  converted  into  a  section  of  all  numbers  by  adding  to  L  all 
numbers  less  than  /3  and  to  R  all  numbers  greater  than  y.  It  is  clear  that 
the  conclusion  stated  in  Dedekind's  Theorem  still  holds  if  we  substitute  '  the 
real  numbers  of  the  interval  (3,  y) '  for  '  the  real  numbers ',  and  that  the 
number  a  in  this  case  satisfies  the  inequalities  05agy. 

18.  Points  of  accumulation.  A  system  of  real  numbers,  or 
of  the  points  on  a  straight  line  corresponding  to  them,  defined  in 
any  way  whatever,  is  called  an  aggregate  or  set  of  numbers  or 
points.  The  set  might  consist,  for  example,  of  all  the  positive 
integers,  or  of  all  the  rational  points. 

It  is  most  convenient  here  to  use  the  language  0l>  geometry*. 
Suppose  then  that  we  are  given  a  set  of  points,  which  we  will 
denote  by  S.  Take  any  point  £,  which  may  or  may  not  belong  to  8. 
Then  there  are  two  possibilities.  Either  (i)  it  is  possible  to  choose 
a  positive  number  &  so  that  the  interval  (£—  S,  £+  8)  does  not  con- 
tain any  point  of  S,  other  than  £  itself  f,  or  (ii)  this  is  not  possible. 

Suppose,  for  example,  that  S  consists  of  the  points  corresponding  to  all 
the  positive  integers.  If  £  is  itself  a  positive  integer,  we  can  take  8  to  be  any 
number  less  than  1,  and  (i)  will  be  true;  or,  if  £  is  halfway  between  two 
positive  integers,  we  can  take  §  to  be  any  number  less  than  ^.  On  the  other 
hand,  if  £  consists  of  all  the  rational  points,  then,  whatever  the  value  of  £, 
(ii)  is  true ;  for  any  interval  whatever  contains  an  infinity  of  rational  points. 

*  The  reader  will  hardly  require  to  be  reminded  that  this  course  is  adopted 
solely  for  reasons  of  linguistic  convenience. 

t  This  clause  is  of  course  unnecessary  if  £  does  not  itself  belong  to  S. 


30  REAL  VARIABLES  [i 

Let  us  suppose  that  (ii)  is  true.  Then  any  interval  (f  —  8,  f  +  8), 
however  small  its  length,  contains  at  least  one  point  £x  which 
belongs  to  S  and  does  not  coincide  with  f ;  and  this  whether  f 
itself  be  a  member  of  8  or  not.  In  this  case  we  shall  say  that  £  is 
a  point  of  accumulation  of  S.  It  is  easy  to  see  that  the  interval 
(£  —  $>  £+£)  must  contain,  not  merely  one,  but  infinitely  many 
points  of  8.  For,  when  we  have  determined  £ls  we  can  take  an 
interval  (£  —  B1}  £+  Sj)  surrounding  £  but  not  reaching  as  far  as  £x. 
But  this  interval  also  must  contain  a  point,  say  £2,  which  is  a 
member  of  S  and  does  not  coincide  with  £.  Obviously  we  may 
repeat  this  argument,  with  £2  in  the  place  of  £ ;  and  so  on 
indefinitely.     In  this  way  we  can  determine  as  many  points 

51  >     52)     ?3)    ••• 

as  we  please,  all  belonging  to  S,  and  all  lying  inside  the  interval 

(f -$.£  +  *)■ 

A  point  of  accumulation  of  $  may  or  may  not  be  itself  a  point 
of  S.    The  examples  which  follow  illustrate  the  various  possibilities. 

Examples  IX.  1.  If  S  consists  of  the  points  corresponding  to  the 
positive  integers,  or  all  the  integers,  there  are  no  points  of  accumulation. 

2.  If  S  consists  of  all  the  rational  points,  every  point  of  the  line  is  a 
point  of  accumulation. 

3.  If  S  consists  of  the  points  1,  \,  ^-,  ...,  there  is  one  point  of  accumula- 
tion, viz.  the  origin. 

4.  If  S  consists  of  all  the  positive  rational  points,  the  points  of  accumula- 
tion are  the  origin  and  all  positive  points  of  the  line. 

19.  Weierstrass's  Theorem.  The  general  theory  of  sets 
of  points  is  of  the  utmost  interest  and  importance  in  the  higher 
branches  of  analysis ;  but  it  is  for  the  most  part  too  difficult  to  be 
included  in  a  book  such  as  this.  There  is  however  one  funda- 
mental theorem  which  is  easily  deduced  from  Dedekind's  Theorem 
and  which  we  shall  require  later. 

Theorem.  If  a  set  8  contains  infinitely  many  points,  and  is 
entirely  situated  in  an  interval  (a,  /3),  then  at  least  one  point  of  the 
interval  is  a  point  of  accumulation  of  8. 

We  divide  the  points  of  the  line  A  into  two  classes  in  the 
following  manner.     The  point  P  belongs  to   L  if  there  are  an 


18,  19]  REAL   VARIABLES  31 

infinity  of  points  of  S  to  the  right  of  P,  and  to  R  in  the  contrary 
case.  Then  it  is  evident  that  conditions  (i)  and  (iii)  of  Dedekind's 
Theorem  are  satisfied;  and  since  a  belongs  to  L  and  /3  to  R, 
condition  (ii)  is  satisfied  also. 

Hence  there  is  a  point  £  such  that,  however  small  be  8,  £  —  8 
belongs  to  L  and  £  +  8  to  R,  so  that  the  interval  (£  —  8,  £  +  S) 
contains  an  infinity  of  points  of  S.  Hence  £  is  a  point  of  accumu- 
lation of  S. 

This  point  may  of  course  coincide  with  a  or  /3,  as  for  instance  when  a  =  0, 
/3  =  1,  and  £  consists  of  the  points  1,  ■§,  J,  ....  In  this  case  0  is  the  sole 
point  of  accumulation. 


MISCELLANEOUS   EXAMPLES   ON  CHAPTER  I. 

1.  What  are  the  conditions  that  ax  +  by  +  cz  =  0,  (1)  for  all  values  of 
x,  y,  z;  (2)  for  all  values  of  x,  y,  z  subject  to  ax+fty  +  yz=0;  (3)  for  all 
values  of  x,  y,  z  subject  to  both  ax+j3y  +  yz  =  0  and  Ax+By  +  Cz  =  01 

2.  Any  positive  rational  number  can  be  expressed  in  one  and  only  one 
way  in  the  form 

,       a2  a3  Civ 


1.371  1.2.3  '  *'•      1.8.3...*' 

where  aj,  «2) ...,  #&  are  integers,  and 

OgOu     0<a2<2,     0^  a3<3,  ...0<«t</-. 

3.  Any  positive  rational  number  can  be  expressed  in  one  and  one  way 
only  as  a  simple  continued  fraction 

a   I     l         l  l 

1     a2  +  a3  +  ...  +a„' 

where  al5  a2, ...  are  positive  integers,  of  which  the  first  only  may  be  zero. 

[Accounts  of  the  theory  of  such  continued  fractions  will  be  found  in  text- 
books of  algebra.  For  further  information  as  to  modes  of  representation  of 
rational  and  irrational  numbers,  see  Hobson,  Theory  of  Functions  of  a  Real 
Variable,  pp.  45 — 49.] 

4.  Find  the  rational  roots  (if  any)  of  9x%  -  6.v2  + 15^  -  10  =  0. 

5.  A  line  AB  is  divided  at  C  in  aurea  sectione  (Euc.  n.  11) — i.e.  so  that 
AB .  AC=BC2.     Show  that  the  ratio  ACjAB  is  irrational. 

[A  direct  geometrical  proof  will  be  found  in  Bromwich's  Infinite  Series, 
%  143,  p.  363.] 

6.  A  is  irrational.     In  what  circumstances  can  — -7 ,  where  a,  b,  c,  d 

cA  +  d 

are  rational,  be  rational  ? 


32  REAL   VARIABLES  [i 

7.  Some  elementary  inequalities.  In  what  follows  au  a2,  ...  de- 
note positive  numbers  (including  zero)  and  p,  q,  ...  positive  integers.  Since 
af  —  a2p  and  axi  -  a2q  have  the  same  sign,  we  have  (af  -  a2p)  («i*  -  a2q)  =0>  or 

«1P  +  9  +  a2P  +  9>a1Pa2«  +  a1«a2»' (1), 

an  inequality  which  may  also  be  written  in  the  form 

2~ 


aiP  +  g  +  r  +  ...  +  CT2P  +  g  +  r+-.  ^  /aiP  +  g2T 

2    '  =1       2~ 


&(*%*)  (*%*) w. 

By  repeated  application  of  this  formula  we  obtain 

and  in  particular S  ( — ^ — j     (4). 

"When  p  =  q=l  in  (1),  or  p  =  2  in  (4),  the  inequalities  are  merely  different 
forms  of  the  inequality  a12+«22S2a1a2,  which  expresses  the  fact  that  the 
arithmetic  mean  of  two  positive  numbers  is  not  less  than  their  geometric 
mean. 

8.  Generalisations  for  n  numbers.  If  we  write  down  the  hn(n-l) 
inequalities  of  the  type  (1)  which  can  be  formed  with  n  numbers  alt  a2,...,  an> 
and  add  the  results,  we  obtain  the  inequality 

%2aP  +  «>2aP2a« (5), 

or  (2aP  +  i)ln>{(2aP)fn}  {(^aq)/n} (6). 

Hence  we  can  deduce  an  obvious  extension  of  (3)  which  the  reader  may 
formulate  for  himself,  and  in  particular  the  inequality 

(2a*>)/»  >{(2a)/»}P (7). 

9.  The  general  form  of  the  theorem  concerning  the  arithmetic  and 
geometric  means.  An  inequality  of  a  slightly  different  character  is 
that  which  asserts  that  the  arithmetic  mean  of  alt  a2,  ...,  an  is  not  less 
than  their  geometric  mean.  Suppose  that  ar  and  a„  are  the  greatest  and 
least  of  the  u's  (if  there  are  several  greatest  or  least  a's  we  may  choose  any 
of  them  indifferently),  and  let  G  be  their  geometric  mean.  We  may  suppose 
0  >  0,  as  the  truth  of  the  proposition  is  obvious  when  G = 0.   If  now  we  replace 

ar  and  as  by 

ar'  =  G,    ctg  =  aratjG} 

we  do  not  alter  the  value  of  the  geometric  mean ;  and,  since 

ar'  +  at' - ar-as=(ar-G)  (as-  G)/G<0, 

we  certainly  do  not  increase  the  arithmetic  mean. 

It  is  clear  that  we  may  repeat  this  argument  until  we  have  replaced  each 
of  ax,  a2,  ...,  an  by  G;  at  most  n  repetitions  will  be  necessary.  As  the  final 
value  of  the  arithmetic  mean  is  G,  the  initial  value  cannot  have  been  less. 


REAL   VARIABLES  33 

10.  Schwarz's  inequality.  Suppose  that  aly  a2,  ...,  an  and  6n  b2,  ...,  6„ 
are  any  two  sets  of  numbers  positive  or  negative.  It  is  easy  to  verify  the 
identity 

{2arbrf = 2ar2  2as2  -  2  {a,.bs  -  asbr)2, 

where  r  and  s  assume  the  values  1,  2,  ...,  n.     It  follows  that 

(2arbry<2ar22br2, 
an  inequality  usually  known  as  Schwarz's  (though  due  originally  to  Cauchy). 

11.  If  a1}  a2,  ...,  an  are  all  positive,  and  sn  =  al  +  a2  +  ...  +  an,  then 

(l+a1)(l+a3)...(l+an)^l+*J4+jL+...+?iL. 

z  !  n  ! 

{Math.  Trip.  1909.) 

12.  If  ffj,  a2,  ...,  «„  and  61}  62,  •••5  K  are  two  sets  of  positive  numbers, 
arranged  in  descending  order  of  magnitude,  then 

{(ti+a2  +  ..,  +  an){bl  +  b2  +  ...+bn)£:n{a1b1+a2b2  +  ...  +  anbn), 

13.  If  a,  b,  c,...k  and  A,  B,  C,  ...  K  are  two  sets  of  numbers,  and  all  of 
the  first  set  are  positive,  then 

aA+bB+...+kK 
a  +  b  +  ...  +  k 

lies  between  the  algebraically  least  and  greatest  of  A,  B,  ...,  K. 

14.  If  sjp,  s/q  are  dissimilar  surds,  and  a  +  b  s?p  +  c  Jq+d  s/{pq)=0, 
where  a,  b,  c,  d  are  rational,  then  a  =  0,  6  =  0,  c  =  0,  d=0. 

[Express  *Jp  in  the  form  M+N \fq,  where  M  and  N  are  rational,  and  apply 
the  theorem  of  §  14.] 

15.  Show  that  if  a  J2  +  b  v/3  +  c  J 5  =  0,  where  a,  b,  c  are  rational  numbers, 
then  a  =  0,  6=0,  e=0. 

16.  Any  polynomial  in  *Jp  and  Jq,  with  rational  coefficients  (i.e.  any 
sum  of  a  finite  number  of  terms  of  the  form  A  {sjp)m  {\fq)n,  where  m  and  n 
are  integers,  and  A  rational),  can  be  expressed  in  the  form 

a  +  b  sip  +  c  sjq  +  d  s!pq, 

where  a,  6,  c,  d  are  rational. 

17.  Express  -= -~- — j-y-,  where  a,  6,  etc.  are  rational,  in  the  form 

a  +  e  y/p  \-j  slq 

A..+B  Jp+C  Jq+D  Jpg, 

where  A,  B,  C,  D  are  rational. 

[Evidently 
a  +  bsfp  +  c*Jq  _  (a  +  b  Kfp  +  c  slq)  (d+e  Jp-fs/q)       a+ff  sjp  +  y  Jq  +  8  Jpq 
d  +  esfp  +fs/q  ~  {d+e  sjpf  -f2q  e  +  {sfp 

where  a,  /3,  etc.  are  rational  numbers  which  can  easily  be  found.     The  required 
u.  3 


34  REAL   VARIABLES  [l 

reduction  may  now  be  easily  completed  by  multiplication  of  numerator  and 
denominator  by  e  —  £>Jp.     For  example,  prove  that 

1  111 

-5  +  W2-7V6.] 


\  +  s'2  +  JZ 

18.  If  a,  b,  x,  y  are  rational  numbers  such  that 

(ay  -  bx)2  +  4(a-x)(b-y)  =  0, 

then  either  (i)  x  =  a,  y  =  b  or  (ii)  1  —  ab  and  1  —  xy  are  squares  of  rational 
numbers.  (Math.  Trip.  1903.) 

19.  If  all  the  values  of  x  and  y  given  by 

ax2  +  2hxy  +  by2  =  1 ,     a'x2  +  2h'xy  +  b'y2  =  1 
(where  a,  h,  b,  a',  h',  V  are  rational)  are  rational,  then 

(h _  Kf  -(a-  a')  (b  -  b'),    (ab'  -  a'bf  +  4  (ah'  -  a'h)  (bh!  -  b'h) 
are  both  squares  of  rational  numbers.  (Math.  Trip.  1899.) 

20.  Show  that  $2  and  ^3  are  cubic  functions  of  ^2  +  ^/3,  with  rational 
coefficients,  and  that  $2  —  ^6  +  3  is  the  ratio  of  two  linear  functions  of 
^2  +  ^3.  (Math.  Trip.  1905.) 

21.  The  expression 

sj{a  +  2m  J  (a  -  m2)}  +  J  {a  -  2m  J  (a  -  on2)} 
is  equal  to  2m  if  2m2  >a> m2,  and  to  2  J  (a- m2)  if  a  >  2m2. 

22.  Show  that  any  polynomial  in  4^2,  with  rational  coefficients,  can  be 
expressed  in  the  form 

a  +  b*/2  +  c*/<l, 
where  a,  b,  e  are  rational. 

More  generally,  if  p  is  any  rational  number,  any  polynomial  in  typ  with 
rational  coefficients  can  be  expressed  in  the  form 

a0  +  ala  +  a2a2+  ...  +am-iam-1, 

where  a0,  au  ...  are  rational  and  a  =  ^/p.  For  any  such  polynomial  is  of  the 
form 

b0  +  b1a  +  b2a2-\- ...  +bkak, 

where  the  £'s  are  rational.  If  £<m  —  1,  this  is  already  of  the  form  required.  If 
h>m—  1,  let  ar  be  any  power  of  «  higher  than  the  (w  — l)th.  Then  r  =  Xm  +  s, 
where  X  is  an  integer  and  O^sg.m- 1 ;  and  a'=aKm+s=pKa.  Hence  we  can 
get  rid  of  all  powers  of  a  higher  than  the  (m  —  l)th. 

23.  Express  ($2 -If  and  (^2-l)/(#2+l)  in  the  form  a+b  1/2  +  0^4, 
where  a,  b,  c  are  rational.  [Multiply  numerator  and  denominator  of  the 
second  expression  by  $4-  $2  +  1.] 

24.  If  a  +  64/2  +  c^/4  =  0, 
where  a,  b,  c  are  rational,  then  a=0,  6  =  0,  c=0. 


REAL   VARIABLES  35 

[Let  y=%2.     Then  f=2  and 

cy2  +  by  +  a  =  0. 
Hence  2cy2  +  2by+ays=0  or 

ay2  +  2ey  +  2b  =  0. 

Multiplying  these  two  quadratic  equations  by  a  and  c  and  subtracting, 
we  obtain  (ab-2c2) y+a2-2bc  =  0,  or  y  =  -(a2-2bc)/(ab-  2c2),  a  rational 
number,  which  is  impossible.  The  only  alternative  is  that  ab-2c2=0, 
a2  -  26c  =  0. 

Hence  aft  =  2c2,  ai=4b2c2.  If  neither  a  nor  b  is  zero,  we  can  divide  the 
second  equation  by  the  first,  which  gives  a3=2b3 :  and  this  is  impossible, 
since  $2  cannot  be  equal  to  the  rational  number  ajb.  Hence  ab  =  0,  e  =  0, 
and  it  follows  from  the  original  equation  that  a,  6,  and  c  are  all  zero. 

As  a  corollary,  if  a  +  b^/2  +  c^=d+e^2+f^/4,  then  a  =  d,b  =  e,  c=f. 

It  may  be  proved,  more  generally,  that  if 

■  l/wi  •         i  {m-Vilm     n 

a0  +  »lP       +...+Om-lP  =0, 

p  not  being  a  perfect  with  power,  then  a0=a1=...  =  am_1  =  0;  but  the  proof  is 
less  simple.] 

25.  If  A  +  Z/B=C+VD,  then  either  A  =  C,  B=D,  or  B  and  D  are  both 
cubes  of  rational  numbers. 

26.  If  %A  +$B  +  %C=0,  then  either  one  of  A,  B,  C  is  zero,  and  the  other 
two  equal  and  opposite,  or  $Ay  £/B,  $C  are  rational  multiples  of  the  same 
surd  SfX. 

27.  Find  rational  numbers  a,  j8  such  that 

4/(7  +  5x/2)  =  a+/3v/2. 

28.  If  (a-b3)b>0,  then 

3/f    ,968  +  a       /{a-W\\   ,      »/(       9&3+^       /A*-&3\1 

is  rational.     [Each  of  the  numbers  under  a  cube  root  is  of  the  form 


Hx/(' 


3b   )) 


where  a  and  /3  are  rational.] 


29.     If  a  =  Z/p,  any  polynomial  in  a  is  the  root  of  an  equation  of  degree  n, 
with  rational  coefficients. 

[We  can  express  the  polynomial  (x  say)  in  the  form 
x  =  li  +  m1a+  ...+ria{n~1}, 
where  lu  mu  ...  are  rational,  as  in  Ex.  22. 


36 


Similarly 


REAL  VARIABLES 

xl  =  l2  +  m2a  +  ...+r2a(n~l\ 


[I 


xn=ln  +  mna  +  ...+rna{n~1). 
Hence  Lxx + L2x-  + . . .  +  LHxn  =  A, 

where  A  is  the  determinant 

^    m1  ...  9\ 
?2    m2  ...  r2 


ln    inn... 
and  Zi,  Z2,  ...  the  minors  of  llt  U,  ....] 

30.  Apply  this  process  to  x=p+*/q,  and  deduce  the  theorem  of  §  14. 

31.  Show  that  y  =  a  +  bp    +cp213  satisfies  the  equation 

yz  -  3ay2+ 7>y  (a2  —  bcp)  —  a3  —  b3p  —  c3/?2  +  3abcp  =  0. 

32.  Algebraical  numbers.     We  have  seen  that  some  irrational  numbers 
(such  as  s/2)  are  roots  of  equations  of  the  type 

a0xn  +  a1xn-1  +  ...  +  an=0, 

where  a0,  au  ...,  an  are  integers.  Such  irrational  numbers  are  called  alge- 
braical numbers :  all  other  irrational  numbers,  such  as  ir  (§  15),  are  called 
transcendental  numbers.    Show  that  if  x  is  an  algebraical  number,  then  so  are 

kx,  where  &  is  any  rational  number,  and  xm'n ,  where  m  and  n  are  any  integers. 

33.  If  x  and  y  are  algebraical  numbers,  then  so  are  x+v,x-y,  xy  and  x/y. 
[We  have  equations     a^x™  +  a^™  ~ l  + . . .  +  am = 0, 

&flr  +  61y»-i  +  ...+6B=0, 

where  the  a's  and  b's  are  integers.  Write  x+y=z,  y=z  —  x  in  the  second, 
and  eliminate  x.     We  thus  get  an  equation  of  similar  form 

c0zP  +  c1zi'-1  +  ...+cp  =  0, 

satisfied  by  z.     Similarly  for  the  other  cases.] 

34.  If  a0xn  +  a1xn-1  +  ...+all  =  O, 

where  a0,  ax,  ...,  an  are  any  algebraical  numbers,  then  x  is  an  algebraical 
number.     [We  have  n  + 1  equations  of  the  type 


a0,rar    r  +  al,rar    r      +  •••+«;, 


=  0    (r  =  0,  1,  ...,  n), 


in  which  the  coefficients  a0j  r,  alt  r,  ...  are  integers     Eliminate  a0.  ax,  ...,  an 
between  these  and  the  original  equation  for  x.] 

35.     Apply  this  process  to  the  equation  x2—  2.^^/2+^3=0. 

[The  result  is  Xs  -  16a;6 + 58a-4  -  48a;2  +  9  =  0.] 


REAL   VARIABLES  37 

36.  Find  equations,  with  rational  coefficients,  satisfied  by 

l  +  v/2  +  v/3,     ^|^|,     W3+V2RVV3-V2},     W+j/S. 

37.  If  x3=x  + 1,  then  x3n=allx  +  bH  +  c,Jx,  where 

38.  If  #° +^5  -  2a;4  -  x3  +  x2  + 1  =  0  and  y  =  x*  -  x2  4-  x  -  1 ,  then  y  satisfies 
a  quadratic  equation  with  rational  coefficients.  {Math.  Trip.  1903.) 

[It  will  be  found  that  f+y  +  1=0.] 


CHAPTER  II 

FUNCTIONS   OF   REAL   VARIABLES 

20.  The  idea  of  a  function.  Suppose  that  x  and  y  are 
two  continuous  real  variables,  which  we  may  suppose  to  be  repre- 
sented geometrically  by  distances  A0P  =  x,  B0Q  =  y  measured 
from  fixed  points  A0,  B0  along  two  straight  lines  A,  M.  And 
let  us  suppose  that  the  positions  of  the  points  P  and  Q  are  not 
independent,  but  connected  by  a  relation  which  we  can  imagine 
to  be  expressed  as  a  relation  between  x  and  y:  so  that,  when 
P  and  x  are  known,  Q  and  y  are  also  known.  We  might, 
for  example,  suppose  that  y  =  x,  or  y  —  2x,  or  \x,  or  x2  +  1.  In 
all  of  these  cases  the  value  of  x  determines  that  of  y.  Or 
again,  we  might  suppose  that  the  relation  between  x  and  y  is 
given,  not  by  means  of  an  explicit  formula  for  y  in  terms  of  x, 
but  by  means  of  a  geometrical  construction  which  enables  us  to 
determine  Q  when  P  is  known. 

In  these  circumstances  y  is  said  to  be  a  function  of  x.  This 
notion  of  functional  dependence  of  one  variable  upon  another  is 
perhaps  the  most  important  in  the  whole  range  of  higher  mathe- 
matics. In  order  to  enable  the  reader  to  be  certain  that  he 
understands  it  clearly,  we  shall,  in  this  chapter,  illustrate  it  by 
means  of  a  large  number  of  examples. 

But  before  we  proceed  to  do  this,  we  must  point  out  that 
the  simple  examples  of  functions  mentioned  above  possess  three 
characteristics  which  are  by  no  means  involved  in  the  general 
idea  of  a  function,  viz.: 

(1)  y  is  determined  for  every  value  of  x; 

(2)  to  each  value  of  x  for  which  y  is  given  corresponds  one 
and  only  one  value  of  y; 

(3)  the  relation  between  x  and  y  is  expressed  by  means  of 
an  analytical  formula,  from  which  the  value  of  y  corresponding  to 
a  given  value  of  x  can  be  calculated  by  direct  substitution  of  the 
latter. 


20]  FUNCTIONS   OF   REAL   VARIABLES  39 

It  is  indeed  the  case  that  these  particular  characteristics  are 
possessed  by  many  of  the  most  important  functions.  But  the  con- 
sideration of  the  following  examples  will  make  it  clear  that  they 
are  by  no  means  essential  to  a  function.  All  that  is  essential  is 
that  there  should  be  some  relation  between  x  and  y  such  that  to 
some  values  of  x  at  any  rate  correspond  values  of  y. 

Examples  X.  1.  Let  y=x  or  2x  or  lx  or  x^  +  l  Nothing  further  need 
be  said  at  present  about  cases  such  as  these. 

2.  Let  y  =  0  whatever  be  the  value  of  x.  Then  y  is  a  function  of  x,  for  we 
can  give  x  any  value,  and  the  corresponding  value  of  y  (viz.  0)  is  known.  In 
this  case  the  functional  relation  makes  the  same  value  of  y  correspond  to  all 
values  of  x.  The  same  would  be  true  were  y  equal  to  1  or  -  \  or  J2  instead 
of  0.     Such  a  function  of  x  is  called  a  constant. 

3.  Let  y2  =  x.  Then  if  x  is  positive  this  equation  defines  two  values  of  y 
corresponding  to  each  value  of  x,  viz.  ±Jx.  U  x=0,  y=0.  Hence  to  the 
particular  value  0  of  x  corresponds  one  and  only  one  value  of  y.  But  if  x  is 
negative  there  is  no  value  of  y  which  satisfies  the  equation.  That  is  to  say, 
the  function  y  is  not  defined  for  negative  values  of  x.  This  function  therefore 
possesses  the  characteristic  (3),  but  neither  (1)  nor  (2). 

4.  Consider  a  volume  of  gas  maintained  at  a  constant  temperature  and 
contained  in  a  cylinder  closed  by  a  sliding  piston*. 

Let  A  be  the  area  of  the  cross  section  of  the  piston  and  W  its  weight. 
The  gas,  held  in  a  state  of  compression  by  the  piston,  exerts  a  certain  pressure 
p0  per  unit  of  area  on  the  piston,  which  balances  the  weight  W,  so  that 

W=APo 

Let  vQ  be  the  volume  of  the  gas  when  the  system  is  thus  in  equilibrium. 
If  additional  weight  is  placed  upon  the  piston  the  latter  is  forced  downwards. 
The  volume  (v)  of  the  gas  diminishes ;  the  pressure  (p)  which  it  exerts 
upon  unit  area  of  the  piston  increases.  Boyle's  experimental  law  asserts  that 
the  product  of  p  and  v  is  very  nearly  constant,  a  correspondence  which,  if 
exact,  would  be  represented  by  an  equation  of  the  type 

pv=a  (i), 

where  a  is  a  number  which  can  be  determined  approximately  by  experiment. 

Boyle's  law,  however,  only  gives  a  reasonable  approximation  to  the  facts 
provided  the  gas  is  not  compressed  too  much.  When  v  is  decreased  and  p 
increased  beyond  a  certain  point,  the  relation  between  them  is  no  longer 
expressed  with  tolerable  exactness  by  the  equation  (i).     It  is  known  that  a 

*  I  borrow  this  instructive  example  from  Prof.  H.  S.  Carslaw's  Introduction  to 
the  Calculus. 


40  FUNCTIONS   OF   REAL  VARIABLES  [il 

much  better  approximation  to  the  true  relation  can  then  be  found  by  means 
of  what  is  known  as  '  van  der  Waals'  law:,  expressed  by  the  equation 


(P+S)(«-0H7    (ii), 


where  a,  /3,  y  are  numbers  which  can  also  be  determined  approximately  by 
experiment. 

Of  course  the  two  equations,  even  taken  together,  do  not  give  anything 
like  a  complete  account  of  the  relation  between  p  and  v.  This  relation  is  no 
doubt  in  reality  much  more  complicated,  and  its  form  changes,  as  v  varies, 
from  a  form  nearly  equivalent  to  (i)  to  a  form  nearly  equivalent  to  (ii).  But, 
from  a  mathematical  point  of  view,  there  is  nothing  to  prevent  us  from  con- 
templating an  ideal  state  of  things  in  which,  for  all  values  of  v  not  less  than 
a  certain  value  V,  (i)  would  be  exactly  true,  and  (ii)  exactly  true  for  all 
values  of  v  less  than  V.  And  then  we  might  regard  the  two  equations  as 
together  defining  p  as  a  function  of  v.  It  is  an  example  of  a  function  which 
for  some  values  of  v  is  defined  by  one  formula  and  for  other  values  of  v  is 
defined  by  another. 

This  function  possesses  the  characteristic  (2) .  to  any  value  of  v  only  one 
value  of  p  corresponds :  but  it  does  not  possess  (1).  For  p  is  not  defined  as 
a  function  of  v  for  negative  values  of  v ;  a  '  negative  volume '  means 
nothing,  and  so  negative  values  of  v  do  not  present  themselves  for  considera- 
tion at  all. 

5.  Suppose  that  a  perfectly  elastic  ball  is  dropped  (without  rotation) 
from  a  height  \grl  on  to  a  fixed  horizontal  plane,  and  rebounds  continually. 

The  ordinary  formulae  of  elementary  dynamics,  with  which  the  reader  is 
probably  familiar,  show  that  h  =  \gp  if  O^f^r,  h=\g  {2r-tf  if  tS^3t,  and 
generally 

h  =  hg{2nT-t)'i 

if  {2n-  l)r£t^(2»+l)r,  h  being  the  depth  of  the  ball,  at  time  t,  below  its 
original  position.  Obviously  h  is  a  function  of  t  which  is  only  defined  for 
positive  values  of  t. 

6.  Suppose  that  y  is  defined  as  being  the  largest  prime  factor  of  x.  This 
is  an  instance  of  a  definition  which  only  applies  to  a  particular  class  of  values 
of  x,  viz.  integral  values.  '  The  largest  prime  factor  of  -^  or  of  v/2  or  of  n ' 
means  nothing,  and  so  our  defining  relation  fails  to  define  for  such  values  of  x 
as  these.  Thus  this  function  does  not  possess  the  characteristic  (1).  It  does 
possess  (2),  but  not  (3),  as  there  is  no  simple  formula  which  expresses  y  in 
terms  of  x, 

7.  Let  y  be  defined  as  the  denominator  of  x  when  x  is  expressed  in  its 
lowest  terms.  This  is  an  example  of  a  function  which  is  defined  if  and  only 
if  x  is  rational.  Thus  y  =  7  if  x=  —  11/7 :  but  y  is  not  defined  for  x=*J2,  'the 
denominator  of  s/2 '  being  a  meaningless  form  of  words. 


20,  21] 


FUNCTIONS   OF   REAL   VARIABLES 


41 


8.  Let  y  be  defined  as  the  height  in  inches  of  policeman  Cx,  in  the 
Metropolitan  Police,  at  5.30  p.m.  on  8  Aug.  1907.  Then  y  is  defined  for  a 
certain  number  of  integral  values  of  x,  viz.  1,  2,  ...,  N,  where  N  is  the  total 
number  of  policemen  in  division  C  at  that  particular  moment  of  time. 

21.  The  graphical  representation  of  functions.  Sup- 
pose that  the  variable  y  is  a  function  of  the  variable  x.  It  will 
generally  be  open  to  us  also  to  regard  #  as  a  function  of  y,  in  virtue 
of  the  functional  relation  between  x  and  y.  But  for  the  present  we 
shall  look  at  this  relation  from  the  first  point  of  view.  We  shall 
then  call  x  the  independent  variable  and  y  the  dependent  variable; 
and,  when  the  particular  form  of  the  functional  relation  is  not 
specified,  we  shall  express  it  by  writing 

V  =/0) 
(or  F  (x),  <f>  (x),  ty  (x),  . . . ,  as  the  case  may  be). 

The  nature  of  particular  functions  may,  in  very  many  cases,  be 
illustrated  and  made  easily  intelligible  as  follows.  Draw  two  lines 
OX,  0  Y  at  right  angles  to  one  another  AY 

and  produced  indefinitely  in  both  direc- 
tions. We  can  represent  values  of  x 
and  y  by  distances  measured  from  0 
along  the  lines  OX,  OY  respectively, 
regard  being  paid,  of  course,  to  sign, 
and  the  positive  directions  of  measure- 
ment being  those  indicated  by  arrows 
in  Fig.  6. 

Let  a  be  any  value  of  x  for  which 
y  is  defined  and  has  (let  us  suppose) 
the  single  value  b.  Take  OA  =  a, 
OB  =  b,  and  complete  the  rectangle 
OAPB.  Imagine  the  point  P  marked  on  the  diagram.  This 
marking  of  the  point  P  may  be  regarded  as  showing  that  the 
value  of  y  for  x  =  a  is  b. 

If  to  the  value  a  oi  x  correspond  several  values  of  y  (say 
b,  V,  b"),  we  have,  instead  of  the  single  point  P,  a  number  of 
points  P,  P',  P". 

We  shall  call  P  the  point  (a,  b);  a  and  b  the  coordinates  of  P 
referred  to  the  axes  OX,  OY ;  a  the  abscissa,  b  the  ordinate  of  P; 
OX  and  OY  the  axis  of  x  and  the  axis  of  y,  or  together   the 


B' 

?' 

B 

P 

b 

O 

a 

A 

'X 

R" 

P" 

Fig.  6. 


42  FUNCTIONS   OF  REAL  VARIABLES  [ll 

axes  of  coordinates,  and  0  the  origin  of  coordinates,  or  simply 
the  origin. 

Let  us  now  suppose  that  for  all  values  a  of  x  for  which  y  is 
defined,  the  value  b  (or  values  b,  b',  b",  ...)  of  y,  and  the  corre- 
sponding point  P  (or  points  P,  P',  P",  ...),  have  been  determined. 
We  call  the  aggregate  of  all  these  points  the  graph  of  the 
function  y. 

To  take  a  very  simple  example,  suppose  that  y  is  defined  as 
a  function  of  x  by  the  equation 

Ax  +  By  +  G  =  0 (1), 

where  A,  B,  C  are  any  fixed  numbers*.  Then  y  is  a  function  of  # 
which  possesses  all  the  characteristics  (1),  (2),  (3)  of  §  20.  It  is 
easy  to  show  that  the  graph  of  y  is  a  straight  line.  The  reader  is 
in  all  probability  familiar  with  one  or  other  of  the  various  proofs 
of  this  proposition  which  are  given  in  text-books  of  Analytical 
Geometry. 

We  shall  sometimes  use  another  mode  of  expression.  We 
shall  say  that  when  x  and  y  vary  in  such  a  way  that  equation  (1) 
is  always  true,  the  locus  of  the  point  (x,  y)  is  a  straight  line,  and 
we  shall  call  (1)  the  equation  of  the  locus,  and  say  that  the  equation 
represents  the  locus.  This  use  of  the  terms  'locus',  'equation  of 
the  locus'  is  quite  general,  and  may  be  applied  whenever  the 
relation  between  x  and  y  is  capable  of  being  represented  by  an 
analytical  formula. 

The  equation  Ax  +  By  +  G  =  0  is  the  general  equation  of  the  first 
degree,  for  Ax  +  By  +  C  is  the  most  general  polynomial  in  x  and  y 
which  does  not  involve  any  terms  of  degree  higher  than  the  first 
in  x  and  y.  Hence  the  general  equation  of  the  first  degree  repre- 
sents a  straight  line.  It  is  equally  easy  to  prove  the  converse 
proposition  that  the  equation  of  any  straight  line  is  of  the  first 
degree. 

We  may  mention  a  few  further  examples  of  interesting  geo- 
metrical loci  defined  by  equations.     An  equation  of  the  form 
{x-af  +  {y-(3y  =  p\ 

*  If  B  =  0,  y  does  not  occur  in  the  equation.  We  must  then  regard  y  as  a 
function  of  x  defined  for  one  value  only  of  x,  viz.  x=  -  C\A,  and  then  having  all 
values. 


X 


21,  22]  FUNCTIONS   OF  REAL   VARIABLES  43 

or  x-  +  y2  +  2Gx  +  2Fy  +O  =  0, 

where  G2  +  F2  —  G  >  0,  represents  a  circle.     The  equation 
Ax2  +  2Hxy  +  By2  +  2Gx  +  2Fy  +  C  =  0 

(the  general  equation  of  the  second  degree)  represents,  assuming 
that  the  coefficients  satisfy  certain  inequalities,  a  conic  section, 
i.e.  an  ellipse,  parabola,  or  hyperbola.  For  further  discussion  of 
these  loci  we  must  refer  to  books  on  Analytical  Geometry. 

22.  Polar  coordinates.  In  what  precedes  we  have  determined 
the  position  of  P  by  the  lengths  of  its  coordinates  0M=x,  MP  =  y. 
If  0P  =  r  and  MOP  =  6,  0  being  an 
angle  between  0  and  2tt  (measured  in 
the  positive  direction),  it  is  evident  that 

x  =  r  cos  0,  y  =  r  sin  0, 

r  =  V(#2  +  y2),     cos  0 :  sin  0 : 1 : :  x :  y :  r, 

and  that  the  position  of  P  is  equally  well 

determined  by  a  knowledge  of  r  and  0. 

We  call  r  and  0  the  polar  coordinates  Fig-  7- 

of  P.     The  former,  it  should  be  observed,  is  essentially  positive*. 

If  P  moves  on  a  locus  there  will  be  some  relation  between  r 
and  0,  say  r=f(0)  or  0  =  F '(r).  This  we  call  the  polar  equation 
of  the  locus.  The  polar  equation  may  be  deduced  from  the  (x,  y) 
equation  (or  vice  versa)  by  means  of  the  formulae  above. 

Thus  the  polar  equation  of  a  straight  line  is  of  the  form 
rcos(0  —  0L)=p, 
where  p  and  a  are  constants.     The  equation  r  —  2a  cos  0  represents 
a  circle  passing  through  the  origin ;  and  the  general  equation  of 
a  circle  is  of  the  form 

r2  +  c2  -  2rc  cos  (0  -  o)  =  A2, 
where  A,  c,  and  a  are  constants. 

*  Polar  coordinates  are  sometimes  denned  so  that  r  may  be  positive  or  negative. 
In  this  case  two  pairs  of  coordinates — e.g.  (1,0)  and  (-1,  it) — correspond  to  the 
same  point.  The  distinction  between  the  two  systems  may  be  illustrated  by  means 
of  the  equation  llr  =  l-ecosd,  where  Z>0,  e>l.  According  to  our  definitions  r 
must  be  positive  and  therefore  cos0<l/e:  the  equation  represents  one  branch  only 
of  a  hyperbola,  the  other  having  the  equation  -  //r  =  l  -  e  cos  6.  With  the  system 
of  coordinates  which  admits  negative  values  of  r,  the  equation  represents  the  whole 
hyperbola. 


44 


FUNCTIONS   OF   REAL  VARIABLES 


[II 


23.  Further  examples  of  functions  and  their  graphical 
representation.  The  examples  which  follow  will  give  the 
reader  a  better  notion  of  the  infinite  variety  of  possible  types  of 
functions. 

A.  Polynomials.  A  'polynomial  in  a;  is  a  function  of  the 
form 

a0xm  +  a^™-1  +  . . .  +  am, 

where  aQ,  a1}  ...,  am  are  constants.     The  simplest  polynomials  are 

the  simple  powers  y=x,  x2,  a3, ...,  xm, The  graph  of  the  function 

xm  is  of  two  distinct  types,  according  as  m  is  even  or  odd. 

First  let  m  =  2.  Then  three  points  on  the  graph  are  (0,  0), 
(1,  1),  (—  1, 1).  Any  number  of  additional  points  on  the  graph 
may  be  found  by  assigning  other  special  values  to  x:  thus  the 
values 


x  =  \,  2,  3,-1   -2,  3 


give 


y  =  i  4,  9, 


4,  9. 


y  =  x* 


(0,0) 
Fig.  8. 


If  the  reader  will  plot  off  a  fair  number  of  points  on  the  graph,  he 
will  be  led  to  conjecture  that  the 
form  of  the  graph  is  something 
like  that  shown  in  Fig.  8.  If 
he  draws  a  curve  through  the 
special  points  which  he  has  proved 
to  lie  on  the  graph  and  then  tests 
its  accuracy  by  giving  x  new 
values,  and  calculating  the  cor- 
responding values  of  y,  he  will 
find  that  they  lie  as  near  to  the  curve  as  it  is  reasonable  to  expect, 
when  the  inevitable  inaccuracies  of  drawing  are  considered.  The 
curve  is  of  course  a  parabola. 

There  is,  however,  one  fundamental  question  which  we  cannot 
answer  adequately  at  present.  The  reader  has  no  doubt  some 
notion  as  to  what  is  meant  by  a  continuous  curve,  a  curve  without 
breaks  or  jumps ;  such  a  curve,  in  fact,  as  is  roughly  represented 
in  Fig.  8.  The  question  is  whether  the  graph  of  the  function 
y  =  x2  is  in  fact  such  a  curve.     This  cannot  be  proved  by  merely 


23] 


FUNCTIONS   OF   REAL  VARIABLES 


45 


constructing  any  number  of  isolated  points  on  the  curve,  although 
I    the  more  such   points  we  construct  the  more  probable  it  will 
appear. 

This  question  cannot  be  discussed  properly  until  Ch.  V.  In 
that  chapter  we  shall  consider  in  detail  what  our  common  sense 
idea  of  continuity  really  means,  and  how  we  can  prove  that  such 
graphs  as  the  one  now  considered,  and  others  which  we  shall 
consider  later  on  in  this  chapter,  are  really  continuous  curves. 
For  the  present  the  reader  may  be  content  to  draw  his  curves  as 
common  sense  dictates. 

It  is  easy  to  see  that  the  curve  y  =  x1  is  everywhere  convex  to  the  axis  of  x. 
Let  P0,  Px  (Fig.  8)  be  the  points  (xQ,  x02),  {xu  xf).  Then  the  coordinates  of 
a  point  on  the  chord  P0P1  are  x=\x0+fixu  y  =  Xx02 + fiXj2,  where  X  and  n  are 
positive  numbers  whose  sum  is  1.     And 

y  -  x-  =  (X  +  fi)  (\Xq2  +  fiXj2)  -(\x0  +  pXxf  =  X/i  (xx  - x0)2  >  0, 

so  that  the  chord  lies  entirely  above  the  curve. 

The  curve  y  =  x*  is  similar  to  y  —  a?  in  general  appearance,  but 
flatter  near  0,  and  steeper  beyond  the  points  A,  A'  (Fig.  9), 
and  y  =  xm,  where  m  is  even  and  greater  than  4,  is  still  more  so. 
As  m  gets  larger  and  larger  the  flatness  and  steepness  grow 
more  and  more  pronounced,  until  the  curve  is  practically  indis- 
tinguishable from  the  thick  line  in  the  figure. 


Fig.  9. 


Fig.  10. 


The  reader  should  next  consider  the  curves  given  by  y=xm, 
when  m  is  odd.  The  fundamental  difference  between  the  two 
cases  is  that  whereas  when  m  is  even  (—  x)m  =  xm,  so  that  the 
curve  is  symmetrical  about  OY,  when  m  is  odd  (—  x)m  =  —  xm,  so 


46  FUNCTIONS   OF   REAL   VARIABLES  [il 

that  y  is  negative  -when  x  is  negative.  Fig.  10  shows  the  curves 
y  —  x,  y  —  x3,  and  the  form  to  which  y  =  xm  approximates  for 
larger  odd  values  of  m 

It  is  now  easy  to  see  how  (theoretically  at  any  rate)  the  graph 
of  any  polynomial  may  be  constructed.  In  the  first  place,  from 
the  graph  of  y  —  xm  we  can  at  once  derive  that  of  Cxm,  where  C  is 
a  constant,  by  multiplying  the  ordinate  of  every  point  of  the 
curve  by  C.  And  if  we  know  the  graphs  of  f(x)  and  F(x),  we 
can  find  that  of  f(x)  +  F(x)  by  taking  the  ordinate  of  every  point 
to  be  the  sum  of  the  ordinates  of  the  corresponding  points  on  the 
two  original  curves. 

The  drawing  of  graphs  of  polynomials  is  however  so  much 
facilitated  by  the  use  of  more  advanced  methods,  which  will  be 
explained  later  on,  that  we  shall  not  pursue  the  subject  further 
here. 

Examples  XI.     1.     Trace  the  curves  y  =  7x\  y—-3x5,  y  =  x10. 

[The  reader  should  draw  the  curves  carefully,  and  all  three  should  be 
drawn  in  one  figure*.  He  will  then  realise  how  rapidly  the  higher  powers 
of  x  increase,  as  x  gets  larger  and  larger,  and  will  see  that,  in  such  a 
polynomial  as 

xw  +  3x5  +  7xi 

(or  even  .r10  +  30.v5  +  700.t4),  it  is  the  Jirst  term  which  is  of  really  preponderant 
importance  when  x  is  fairly  large.  Thus  even  when  x=4,  x10  >  1,000,000, 
while  30.Z5  <  35,000  and  700a-4  <  180,000;  while  if  a  =  10  the  preponderance 
of  the  first  term  is  still  more  marked.] 

2.  Compare  the  relative  magnitudes  of  a;12,  1,000,000a'6,  l,000,000,000,000.v 
when  x=l,  10,  100,  etc. 

[The  reader  should  make  up  a  number  of  examples  of  this  type  for  himself. 
This  idea  of  the  relative  rate  of  growth  of  different  functions  of  x  is  one  with 
which  we  shall  often  be  concerned  in  the  following  chapters.] 

3.  Draw  the  graph  of  ax2  +  2bx  +  c 

[Here  y  —  {{ac-  b2)ja)  =a  {x  +  (b/a)}2.  If  we  take  new  axes  parallel  to  the 
old  and  passing  through  the  point  x=  —  b/a,  y  =  (ac  —  b2)/a,  the  new  equation 
isy'  =  ax'2.     The  curve  is  a  parabola.] 

4.  Trace  the  curves  y = xz  -  3x  + 1 ,  y = x2  (x  —  1 ),  y = x  (x  - 1  )2. 

*  It  will  be  found  convenient  to  take  the  scale  of  measurement  along  the  axis 
of  y  a  good  deal  smaller  than  that  along  the  axis  of  x,  in  order  to  prevent  the 
figure  becoming  of  an  awkward  size. 


23,  24]  FUNCTIONS   OF   REAL   VARIABLES  47 

24.  B.  Rational  Functions.  The  class  of  functions  which 
ranks  next  to  that  of  polynomials  in  simplicity  and  importance 
is  that  of  rational  functions.  A  rational  function  is  the  quotient 
of  one  polynomial  by  another :  thus  if  P  (x),  Q  (x)  are  polynomials, 
we  may  denote  the  general  rational  function  by 

In  the  particular  case  when  Q  (x)  reduces  to  unity  or  any  other 
constant  {i.e.  does  not  involve  x),  R  (x)  reduces  to  a  polynomial : 
thus  the  class  of  rational  functions  includes  that  of  polynomials 
as  a  sub-class.  The  following  points  concerning  the  definition 
should  be  noticed. 

(1)  "We  usually  suppose  that  P{x)  and  Q  (x)  have  no  common  factor  x  +  a 
or  xv  +  ax13'1  +  bxp~2  + ...  +  k,  all  such  factors  being  removed  by  division. 

(2)  It  should  however  be  observed  that  this  removal  of  common  factors 
does  as  a  rule  change  the  function.  Consider  for  example  the  function  x/x, 
which  is  a  rational  function.  On  removing  the  common  factor  x  we  obtain 
1/1  =  1.  But  the  original  function  is  not  always  equal  to  1 :  it  is  equal  to  1 
only  so  long  as  #4=0.  If  #  =  0  it  takes  the  form  0/0,  which  is  meaningless. 
Thus  the  function  x/x  is  equal  to  1  if  x  4=0  and  is  undefined  when  x  =  0.  It 
therefore  differs  from  the  function  1,  which  is  always  equal  to  1. 

(3)  Such  a  function  as 


\x+l  +  x-l)/  \x  +  x-2J 


may  be  reduced,  by  the  ordinary  rules  of  algebra,  to  the  form 

x*(x-2) 
(x-l)2(x+l)' 

which  is  a  rational  function  of  the  standard  form.  But  here  again  it  must  be 
noticed  that  the  reduction  is  not  always  legitimate.  In  order  to  calculate  the 
value  of  a  function  for  a  given  value  of  x  we  must  substitute  the  value  for  as 
in  the  function  in  the  form  in  which  it  is  given.  In  the  case  of  this  function 
the  values  x=  -1,  1,  0,  2  all  lead  to  a  meaningless  expression,  and  so  the 
function  is  not  defined  for  these  values.  The  same  is  true  of  the  reduced 
form,  so  far  as  the  values  —  1  and  1  are  concerned.  But  x  =  0  and  x  =  2  give 
the  value  0.     Thus  once  more  the  two  functions  are  not  the  same. 

(4)  But,  as  appeal's  from  the  particular  example  considered  under  (3), 
there  will  generally  be  a  certain  number  of  values  of  x  for  which  the  function 
is  not  defined  even  when  it  has  been  reduced  to  a  rational  function  of  the 
standard  form.  These  are  the  values  of  x  (if  any)  for  which  the  de- 
nominator vanishes.  Thus  (x2 - 7)/(#2 - 3#  +  2)  is  not  defined  when  x=l 
or  2. 


48  FUNCTIONS   OF   REAL  VARIABLES  [il 

(5)  Generally  we  agree,  in  dealing  with  expressions  such  as  those  con- 
sidered in  (2)  and  (3),  to  disregard  the  exceptional  values  of  x  for  which  such 
processes  of  simplification  as  were  used  there  are  illegitimate,  and  to  reduce 
our  function  to  the  standard  form  of  rational  function.  The  reader  will 
easily  verify  that  (on  this  understanding)  the  sum,  product,  or  quotient  of 
two  rational  functions  may  themselves  be  reduced  to  rational  functions  of 
the  standard  type.  And  generally  a  rational  function  of  a  rational  function 
is  itself  a  rational  function:  i.e.  if  in  z  =  P  (y)lQ(y),  where  P  and  Q  are 
polynomials,  we  substitute  y  —  Px  {x)jQl  (x),  we  obtain  on  simplification  an 
equation  of  the  form  z=P2(x)/Q2{x). 

(6)  It  is  in  no  way  presupposed  in  the  definition  of  a  rational  function 
that  the  constants  which  occur  as  coefficients  should  be  rational  numbers. 
The  word  rational  has  reference  solely  to  the  way  in  which  the  variable  x 
appears  in  the  function.     Thus 

x2  +  x  +  >/3 
x$2-ir 
is  a  rational  function 

The  use  of  the  word  rational  arises  as  follows.  The  rational  function 
P  (X)IQ  (x)  may  De  generated  from  x  by  a  finite  number  of  operations  upon 
x,  including  only  multiplication  of  x  by  itself  or  a  constant,  addition  of  terms 
thus  obtained,  and  division  of  one  function,  obtained  by  such  multiplications 
and  additions,  by  another.  In  so  far  as  the  variable  x  is  concerned,  this  pro- 
cedure is  very  much  like  that  by  which  all  rational  numbers  can  be  obtained 
from  unity,  a  procedure  exemplified  in  the  equation 
5  =  1  +  1  +  1  +  1  +  1 
3  1+1+1        ' 

Again,  any  function  which  can  be  deduced  from  x  by  the  elementary 
operations  mentioned  above,  using  at  each  stage  of  the  process  functions 
which  have  already  been  obtained  from  x  in  the  same  way,  can  be  reduced  to 
the  standard  type  of  rational  function.  The  most  general  kind  of  function 
which  can  be  obtained  in  this  way  is  sufficiently  illustrated  by  the  example 
x 


x*  +  l+' 


ltr-.3V2)/(17+W' 
"*"     &r  +  l     '/ 


which  can  obviously  be  reduced  to  the  standard  type  of  rational  function. 

25.  The  drawing  of  graphs  of  rational  functions,  even  more 
than  that  of  polynomials,  is  immensely  facilitated  by  the  use  of 
methods  depending  upon  the  differential  calculus.  We  shall 
therefore  content  ourselves  at  present  with  a  very  few  examples. 

Examples  XII.     1 .     Draw  the  graphs  ofy  =  l/x,y  =  l/x2,  y  =  l/xz, .... 

[The  figures  show  the  graphs  of  the  first  two  curves.  It  should  be 
observed  that,  since  1/0,  I/O2,  ...  are  meaningless  expressions,  these  functions 
are  not  defined  for  A'  =  0.] 


24-26] 


FUNCTIONS   OF   REAL   VARIABLES 


49 


2.     Trace  y=x+(ljx\   ar-(l/#),  ^  +  (l/x2),  o,-2-(l/.^)    and    a.v+(b/x) 
taking  various  values,  positive  and  negative,  for  a  and  b. 


3.     Trace 


_x±l        /a?+l\2 


^+1 


(.v-1)2'     a^-l" 

4.  Trace  y=l/(a?-a)(a?-6),  lj(x-a)(x-b)  (#-c),  where  a<6<c. 

5.  Sketch  the  general  form  assumed  by  the  curves  v=llxm  as  »i 
becomes  larger  and  larger,  considering  separately  the  cases  in  which  m  is 
odd  or  even. 


(-1,-1) 


y  =  llx 


y  =  ]/x2 


Fig.  11. 


Fig.  12. 


26.  C.  Explicit  Algebraical  Functions.  The  next  im- 
portant class  of  functions  is  that  of  explicit  algebraical  functions. 
These  are  functions  which  can  be  generated  from  a;  by  a  finite 
number  of  operations  such  as  those  used  in  generating  rational 
functions,  together  with  a  finite  number  of  operations  of  root 
extraction.     Thus 


vci  +  ao+vxi-*-)' 

fx2  +  x  +  \/3\! 


*/x  +  \J(x  +  \/x), 


fx"  +  x+^/S\i 

\X~$"2-7T   ) 


X$2-TT 

are  explicit  algebraical  functions,  and  so  is  xm/n  (i.e.  tyx™),  where  m 
and  n  are  any  integers. 

It  should  be  noticed  that  there  is  an  ambiguity  of  notation 

involved  in  such  an  equation  as  y  =  \/x.     We  have,  up  to  the 

present,  regarded  (e.g.)  V2  as  denoting  the  positive  square  root 

of  2,  and  it  would  be  natural  to  denote  by  \Jx,  where  x  is  any 

n.  4 


50  FUNCTIONS   OF   REAL   VARIABLES  [il 

positive  number,  the  positive  square  root  of  x,  in  which  case 
y  =  sjx  would  be  a  one-valued  function  of  x.  It  is  however 
often  more  convenient  to  regard  \jx  as  standing  for  the  two-valued 
function  whose  two  values  are  the  positive  and  negative  square 
roots  of  x. 

The  reader  will  observe  that,  when  this  course  is  adopted,  the 
function  \fx  differs  fundamentally  from  rational  functions  in  two 
respects.  In  the  first  place  a  rational  function  is  always  defined 
for  all  values  of  x  with  a  certain  number  of  isolated  exceptions. 
But  six  is  undefined  for  a  whole  range  of  values  of  x  (i.e.  .  all 
negative  values).  Secondly  the  function,  when  x  has  a  value 
for  which  it  is  defined,  has  generally  two  values  of  opposite  signs. 

The  function  tyx,  on  the  other  hand,  is  one-valued  and  defined 
for  all  values  of  x. 

Examples  XIII.  1.  »J{(x-a)(b-x)},  where  a<b,  is  defined  only  for 
«£,i'  <6.     If  a<x<b  it  has  two  values  :  if  x  =  a  or  b  only  one,  viz.  0. 

2,  Consider  similarly 

J{(x  -  a)  (x  -  b)  (x  -  c)}    (a  <  b  <  c), 

«/{*(*"-<£)},     y{(x-a)*(b-x)}     (a<b), 

J(1+*W(1-*)         .,        f) 

7(i+*)+n/(i-*)'   w+«a]' 

3,  Trace  the  curves      y2=x,    y*  =  x,    y2=xi. 

4,  Draw  the  graphs  of  the  functions  y  =  >J(a2  —  x-),    y=b<JQ  —  (a?9/a2)}. 

27.     D.     Implicit  Algebraical  Functions.      It  is  easy  to 

verify  that  if 

^  =V(l+s)-.y(l-tt) 

V     VCl  +  aO  +  vXl-ff)' 

/l  +  3/\6_  (1  +xf 


then 


or  if  y  =  \/x  +  \/(x  +  \fx), 

then  if  -  (4/  +  4y  + 1 )  x  =  0. 

Each  of  these  equations  may  be  expressed  in  the  form 

tfm  +  R1ym-1  +  ...+Bm=0 (1), 

where  Ru  R2,  ...,  Rm  are  rational  functions  of  x:  and  the  reader 
will  easily  verify  that,  if  y  is  any  one  of  the  functions  considered 
in  the  last  set  of  examples,  y  satisfies  an  equation  of  this  form. 


20,  27]  FUNCTIONS   OF  REAL   VARIABLES  51 

It  is  naturally  suggested  that  the  same  is  true  of  any  explicit 
algebraic  function.  And  this  is  in  fact  true,  and  indeed  not 
difficult  to  prove,  though  we  shall  not  delay  to  write  out  a  formal 
proof  here.  An  example  should  make  clear  to  the  reader  the  lin.cs 
on  which  such  a  proof  would  proceed.     Let 

_  x  +  \/x  +  \J{x  +  *Jx]  +  y/(l  +  Od) 

J  ~  x  -  sjx  +  V{«  +  V*}  -  ^(1  +  *')  ' 
Then  we  have  the  equations 

x  +  u  +  v  +  w 
y  = , 

X  —  U  +  V  —  w 

u*  =  x,      v-  =  X  +  U,      Wz  =  1  +  X, 
and  we  have  only  to  eliminate  u,  v,  w  between  these  equations  in 
order  to  obtain  an  equation  of  the  form  desired. 

We  are  therefore  led  to  give  the  following  definition  :  a  function 
y=f(x)  will  be  said  to  be  an  algebraical  function  of  x  if  it  is  the 
root  of  an  equation  such  as  (1),  i.e.  the  root  of  an  equation  of  the 
mth  degree  in  y,  ivhose  coefficients  are  rational  functions  of  x.  There 
is  plainly  no  loss  of  generality  in  supposing  the  first  coefficient  to 
be  unity. 

This  class  of  functions  includes  all  the  explicit  algebraical 
functions  considered  in  §  26.  But  it  also  includes  other  functions 
which  cannot  be  expressed  as  explicit  algebraical  functions.  For 
it  is  known  that  in  general  such  an  equation  as  (1)  cannot  be 
solved  explicitly  for  y  in  terms  of  x,  when  m  is  greater  than  4, 
though  such  a  solution  is  always  possible  if  m  =  1,  2,  3,  or  4  and 
in  special  cases  for  higher  values  of  m. 

The  definition  of  an  algebraical  function  should  be  compared 
with  that  of  an  algebraical  number  given  in  the  last  chapter 
(Misc.  Exs.  32). 

Examples  XIV.     1.     If  m  =  l,  y  is  a  rational  function. 

2.     If  m  =  2,  the  equation  is  y~  +  Rxy  +  -ff2=0,  s0  that 

This  function  is  defined  for  all  values  of  x  for  which  /?x224i?2.  It  has  two 
values  if  R^>ARo,  and  one  if  fil2  =  4R2. 

If  m=3  or  4,  we  can  use  the  methods  explained  in  treatises  on  Algebra  for 
the  solution  of  cubic  and  biquadratic  equations.  But  as  a  rule  the  process  is 
complicated  and  the  results  inconvenient  in  form,  and  we  can  generally  study 
the  properties  of  the  function  better  by  means  of  the  original  equation. 

4—2 


52  FUNCTIONS   OF   REAL   VARIABLES  [il 

3.  Consider  the  functions  denned  by  the  equations 

y2-2y-x2=0,    y2 -2y  +  x2  =  0,    #*-  2y2  +  x2=0, 
in  each  case  obtaining  y  as  an  explicit  function  of  x,  and  stating  for  what 
values  of  x  it  is  defined. 

4.  Find  algebraical  equations,  with  coefficients  rational  in  x,  satisfied  by 
each  of  the  functions 

n/*  +  V(1/*),     Vx  +  V(llx),    *J(x  +  s/x),    *J{x  +  »J(x  +  s/x)}. 

5.  Consider  the  equation  yi=x2. 

[Here  y2  =  ±  ff.  If  #  is  positive,  y=*Jx:  if  negative,  y = V  ( -  a?).  Thus  the 
function  has  two  values  for  all  values  of  x  save  #=0.] 

6.  An  algebraical  function  of  an  algebraical  function  of  x  is  itself  an 
algebraical  function  of  x. 

[For  we  have 

.  ym+i?i(z)ym-1+-+fim(z)=o, 

where  zn+Sl  (x)  zn~l  + ...  +  Sn  (x)  =0. 

Eliminating  z  we  find  an  equation  of  the  form 

yP  +  T1  (x)  yr>-i  +  ...  +  Tp  (x)  =0. 
Here  all  the  capital  letters  denote  rational  functions.] 

7.  An  example  should  perhaps  be  given  of  an  algebraical  function  which 
cannot  be  expressed  in  an  explicit  algebraical  form.  Such  an  example  is  the 
function  y  defined  by  the  equation 

y3-y  —  x=0- 
But  the  proof  that  we  cannot  find  an  explicit  algebraical  expression  for  y  in 
terms  of  x  is  difficult,  and  cannot  be  attempted  here. 

28.  Transcendental  functions.  All  functions  of  x  which 
are  not  rational  or  even  algebraical  are  called  transcendental 
functions.  This  class  of  functions,  being  denned  in  so  purely- 
negative  a  manner,  naturally  includes  an  infinite  variety  of  whole 
kinds  of  functions  of  varying  degrees  of  simplicity  and  importance. 
Among  these  we  can  at  present  distinguish  two  kinds  which  are 
particularly  interesting. 

E.  The  direct  and  inverse  trigonometrical  or  circular 
functions.  These  are  the  sine  and  cosine  functions  of  elementary 
trigonometry,  and  their  inverses,  and  the  functions  derived  from 
them.  We  may  assume  provisionally  that  the  reader  is  familiar 
with  their  most  important  properties  *. 

*  The  definitions  of  the  circular  functions  given  in  elementary  trigonometry  pre- 
suppose that  any  sector  of  a  circle  has  associated  with  it  a  definite  number  called  its 
area.    How  this  assumption  is  justified  will  appear  in  Ch.  VII. 


27,  28] 


FUNCTIONS   OF   REAL   VARIABLES 


53 


Examples  XV.    1.    Draw  the  graphs  of  cos  x,  sin  x,  and  a  cos  x  +  b  sin  >•. 

[Since  acosx  +  bsinx=@co$(x-a),  where  $=sJ(a2->rb2),  and  a  is  an  angle 
whose  cosine  and  sine  are  a/^/(a2  +  b2)  and  b/»J(a2  +  b2),  the  graphs  of  these 
three  functions  are  similar  in  character.] 

2.  Draw  the  graphs  of  cos2  x,  sin2^,  a  cos2  x  +  b  sin2  x. 

3.  Suppose  the  graphs  of  f(x)  and  F(x)  drawn.     Then  the  graph  of 

/  (x)  cos2  x + F  (x)  sin2  x 
is  a  wavy  curve  which  oscillates  between  the  curves  y=f{x),  y  =  F(x).    Draw 
the  graph  when  f(x)  =  x,  F(x)=x2. 

4.  Show  that  the  graph  of  cospx  +  cosqx  lies  between  those  of 
2  cos \(p -  q) x  and  —  2 cos  |(p  +  q)  x,  touching  each  in  turn.  Sketch  the 
graph  when  (p-q)l(p  +  q)  is  small.  (Math.  Trip.  1908.) 

5.  Draw  the  graphs  of  x  +  sin  x,  (l/#)  +  sin#,  #sin#,  (sinx)/x. 

6.  Draw  the  graph  of  sin  (Ifx). 

[If  y =sin  (1/x),  then  y  =  0  when  .r  =  1/mn-,  where  «&  is  any  integer.  Similarly 
?/  =  l  when  a*=1/(2»i+|)7t  and  y=—  1  when  a?=l/(2»i— ^) w.  The  curve  is 
entirely  comprised  between  the  lines  y=  —1  and  y=l  (Fig.  13).  It  oscillates 
up  and  down,  the  rapidity  of  the  oscillations  becoming  greater  and  greater  as 
x  approaches  0.  For  x=0  the  function  is  undefined.  When  x  is  large  y  is 
small  *.  The  negative  half  of  the  curve  is  similar  in  character  to  the  positive 
half.] 

7.  Draw  the  graph  of  x  sin  (1/x). 

[This  curve  is  comprised  between  the  lines  y=  -  x  and  y=x  just  as  the 
last  curve  is  comprised  between  the  lines  y=  —1  and  y=l  (Fig.  14).] 


Fig.  13.  Fig.  14. 

*  See  Chs.  IV  and  V  for  explanations  as  to  the  precise  meaning  of  this  phrase. 


54  FUNCTIONS   OF   REAL   VARIABLES  [il 

8.  Draw  the  graphs  of  #2sin  (l/#),  (l/o?)  sin  (ljx),  sin2  (l/#),  {.r  sin  (l/.r)}2, 
a  cos2  (I/a) +  6  sin2  (l/#),  sin  x  +  sin  (l/x),  Bin  x  sin  (l/#). 

9.  Draw  the  graphs  of  cos  a2,  sin  a;2,  a  cos  a2  +  b  sin  a2. 

10.  Draw  the  graphs  of  arc  cos  x  and  arc  sin  x. 

[If  y=. arc  cos  a,  a = cosy.  This  enables  us  to  draw  the  gi'aph  of  x,  con- 
sidered as  a  function  of  y,  and  the  same  curve  shows  y  as  a  function  of  x. 
It  is  clear  that  y  is  only  defined  for  -l£.r£l,  and  is  infinitely  many- 
valued  for  these  values  of  x.  As  the  reader  no  doubt  remembers,  there  is, 
when  -  k,f<l,  a  value  of  y  between  0  and  -it,  say  a,  and  the  other  values 
of  y  are  given  by  the  formula  2mv  ±  a,  where  n  is  any  integer,  positive  or 
negative.] 

1 1 .  Draw  the  graphs  of 

tan  x,     cot  x,     sec  x,     cosec  x,     tan2  x,     cot2  x,     sec2  x,     cosec2  x. 

12.  Draw  the  graphs  of  arc  tan  x,  arc  cot  x,  arc  sec  x,  arc  cosec  x.  Give 
fornmlae  (as  in  Ex.  10)  expressing  all  the  values  of  each  of  these  functions 
in  terms  of  any  particular  value. 

13.  Draw  the  graphs  of  tan  (l/.r),  cot(l/.v),  sec(l/.v),  cosec  (1  jx) 

14.  Show  that  cos  a  and  sin.r  are  not  rational  functions  of  x. 

[A  function  is  said  to  be  periodic,  with  period  a,  if  f(x)=f(x  +  a)  for  all 
values  of  x  for  which  f(x)  is  defined.  Thus  cos  a  and  sin  x  have  the  period 
27T..  It  is  easy  to  see  that  no  periodic  function  can  be  a  rational  function, 
unless  it  is  a  constant.     For  suppose  that 

f(x)  =  P(x)/Q(x), 

where  P  and  Q  are  polynomials,  and  that/(.r)  =/(.r  +  a),  each  of  these  equations 
holding  for  all  values  of  x.  Let  /(0)  =  /•.  Then  the  equation  P  (x)  —  kQ  (x)  =  0 
is  satisfied  by  an  infinite  number  of  values  of  x,  viz.  x=0,  a,  2a,  etc.,  and 
therefore  for  all  values  of  x.  Thus  f(x)  =  k  for  all  values  of  x,  i.e.  f(x)  is  a 
constant.] 

15.  Show,  more  generally,  that  no  function  with  a  period  can  be  an 
algebraical  function  of  x. 

[Let  the  equation  which  defines  the  algebraical  function  be 

y™+Ritf»-i  +  ...+Rm=0 (1) 

where  R\,  ...  are  rational  functions  of  x.     This  may  be  put  in  the  form 

p0ri+Piri'1+-+pm=o, 

where  P0,  Pt,  ...  are  polynomials  in  x.     Arguing  as  above,  we  see  that 


28,  29] 


FUNCTIONS   OF   HEAL   VARIABLES 


DO 


for  all  values  of  x.     Hence  y  =  k  satisfies  the  equation  (1)  for  all  values  of  x 
and  one  set  of  values  of  our  algebraical  function  reduces  to  a  constant. 

Now  divide  (1)  by  y  -  k  and  repeat  the  argument.  Our  final  conclusion  is 
that  our  algebraical  function  has,  for  any  value  of  x,  the  same  set  of  values 
k,  k',  ... ;  i.e.  it  is  composed  of  a  certain  number  of  constants.] 

16.  The  inverse  sine  and  inverse  cosine  are  not  rational  or  algebraical 
functions.  [This  follows  from  the  fact  that,  for  any  value  of  x  between  -  1 
and  +  1,  arc  sin#  and  arc  cos  a;  have  infinitely  many  values.] 


29.    F.    Other  classes  of  transcendental  functions.    Next 

in  importance  to  the  trigonometrical  functions  come  the  expo- 
nential and  logarithmic  functions,  which  will  be  discussed  in 
Chs.  IX  and  X.  But  these  functions  are  beyond  our  range  at 
present.  And  most  of  the  other  classes  of  transcendental  func- 
tions whose  properties  have  been  studied,  such  as  the  elliptic 
functions,  Bessel's  and  Legendre's  functions,  Gamma-functions, 
and  so  forth,  lie  altogether  beyond  the  scope  of  this  book. 
There  are  however  some  elementary  types  of  functions  which, 
though  of  much  less  importance  theoretically  than  the  rational, 
algebraical,  or  trigonometrical  functions,  are  particularly  instruc- 
tive as  illustrations  of  the  possible  varieties  of  the  functional 
relation. 

Examples  XVI.  1.  Let  y=\x\  where  [x]  denotes  the  greatest  integer 
not  greater  than  x.  The  graph  is  shown  in  Fig.  15  a.  The  left-hand  end 
points  of  the  thick  lines,  but  not  the  right-hand  ones,  belong  to  the  graph. 

2.     y=x-[x],     (Fig.  15  6.) 


Fig.  15  a. 


Fig.  156. 


56  FUNCTIONS   OF   REAL   VARIABLES  [il 

3.    y=V{*-[*]}.     (^g.  15  c.)  4    y  =  [x]  +  J{x-[x]}.     (Fig.  15o?.) 

5.    y=(ff-[>])2,     [*]  +  (*-[*])". 


Fig.  15  c. 


Fig.  15  d. 


7.  Let  y  be  defined  as  the  largest  prime  factor  of  x  (cf.  Exs.  x.  6). 
Then  y  is  defined  only  for  integral  values  of  x.     If 

a?=l,  2,  3,  4,  5,  6,  7,  8,  9,  10,  11,  12,  13,  ... , 

then  y  =  \,  2,  3,  2,  5,  3,  7,  2,  3,    5,11,    3,13,.... 

The  graph  consists  of  a  number  of  isolated  points. 

8.  Let  y  be  the  denominator  of  x  (Exs.  x.  7).  In  this  case  y  is  defined 
only  for  rational  values  of  x.  We  can  mark  off  as  many  points  on  the  graph 
as  we  please,  but  the  result  is  not  in  any  ordinary  sense  of  the  word  a  curve, 
and  there  are  no  points  corresponding  to  any  irrational  values  of  x. 

Draw  the  straight  line  joining  the  points  (N-  1,  JV),  (iV,  JV),  where  N  is  a 
positive  integer.  Show  that  the  number  of  points  of  the  locus  which  lie  on 
this  line  is  equal  to  the  number  of  positive  integers  less  than  and  prime  to  N. 

9.  Let  #  =  0  when  x  is  an  integer,  y=x  when  x  is  not  an  integer.  The 
graph  is  derived  from  the  straight  line  y  =  x  by  taking  out  the  points 

...(-1,  -1),     (0,0),     (1,1),     (2,2),... 
and  adding  the  points  (  — 1,  0),  (0,  0),  (1,  0),  ...  on  the  axis  of  x. 

The  reader  may  possibly  regard  this  as  an  unreasonable  function.  Why, 
he  may  ask,  if  y  is  equal  to  x  for  all  values  of  x  save  integral  values,  should  it 
not  be  equal  to  x  for  integral  values  too  1  The  answer  is  simply,  why  should 
itl  The  function  y  does  in  point  of  fact  answer  to  the  definition  of  a 
function :  there  is  a  relation  between  x  and  y  such  that  when  x  is  known  y  is 
known.  We  are  perfectly  at  liberty  to  take  this  relation  to  be  what  we  please, 
however  arbitrary  and  apparently  futile.  This  function  y  is,  of  course,  a  quite 
different  function  from  that  one  which  is  always  equal  to  x,  whatever  value, 
integral  or  otherwise,  x  may  have. 


29] 


FUNCTIONS    OF   REAL  VARIABLES 


57 


10.  Let  y  =  1  when  x  is  rational,  but  y  =  0  when  x  is  irrational.  The  graph 
consists  of  two  series  of  points  arranged  upon  the  lines  y  =  \  and  y— -0.  To 
the  eye  it  is  not  distinguishable  from  two  continuous  straight  lines,  but  in 
reality  an  infinite  number  of  points  are  missing  from  each  line. 

11.  Let  y=x  when  x  is  irrational  and  y  =  s'{{\  +p2)/(l+q2)}  when  x  is  a 
rational  fraction  pjq. 


Fig.  16. 


The  irrational  values  of  x  contribute  to  the  graph  a  curve  in  reality  dis- 
continuous, but  apparently  not  to  be  distinguished  from  the  straight  line  y=x. 

Now  consider  the  rational  values  of  x.  First  let  x  be  positive.  Then 
s/{(l+p2)/(l+q2)}  cannot  be  equal  to  pjq  unless  p  =  q,  i.e.  x=l.  Thus  all 
the  points  which  correspond  to  rational  values  of  x  lie  off  the  line,  except 
the  one  point  (1,  1).  Again,  if  p<q,  v^{(l+p2)/(l+?2)}  >plq  5  if  P  >  ?> 
V/{(1 4-jo2)/(l  +  q2)}  <plq.  Thus  the  points  lie  above  the  line  y=x  if  0 <x <  1, 
below  if  x  >  1.  If  p  and  q  are  large,  v/{(l  +p2)/(l  +  q2)}  is  nearly  equal  to  pjq. 
Near  any  value  of  x  we  can  find  any  number  of  rational  fractions  with  large 
numerators  and  denominators.  Hence  the  graph  contains  a  large  number  of 
points  which  crowd  round  the  line  y=x.  Its  general  appearance  (for  positive 
values  of  x)  is  that  of  a  line  surrounded  by  a  swarm  of  isolated  points  which 
gets  denser  and  denser  as  the  points  approach  the  line. 

The  part  of  the  graph  which  corresponds  to  negative  values  of  x  consists 
of  the  rest  of  the  discontinuous  line  together  with  the  reflections  of  all  these 
isolated  points  in  the  axis  of  y.  Thus  to  the  left  of  the  axis  of  y  the  swarm 
of  points  is  not  round  y  =  x  but  round  y  =  —  x,  which  is  not  itself  part  of  the 
graph.     See  Fig.  16. 


58  FUNCTIONS   OF   REAL    VARIABLES  [il 

30.  Graphical  solution  of  equations  containing  a  single 
unknown  number.  Many  equations  can  be  expressed  in  the 
form 

/(*)  =  *(*) (1), 

where  f(x)  and  </>  (%)  are  functions  whose  graphs  are  easy  to  draw. 
And  if  the  curves 

y-f(®)>   y  =  4>(x) 

intersect  in  a  point  P  whose  abscissa  is  £,  then  £  is  a  root  of  the 
equation  (1). 

Examples  XVII.  1.  The  quadratic  equation  ax2  +  2bx  +  c=0.  This 
may  be  solved  graphically  in  a  variety  of  ways.     For  instance  we  may  draw 

the  graphs  of 

y  =  ax  +  2b,    y  =  -  cjx, 

whose  intersections,  if  any,  give  the  roots.     Or  we  may  tako 

y  =  x2,    y=-(2bx  +  c)ja. 

But  the  most  elementary  method  is  probably  to  draw  the  circle 

a(x2+y2)  +  2bx  +  c=0, 

whose  centre  is  (  —  b/a,  0)  and  radius   {J(b2  —  ac)}/a.     The  abscissae  of  its 
intersections  with  the  axis  of  x  are  the  roots  of  the  equation, 

2.  Solve  by  any  of  these  methods 

x2  +  2x-3=0,    x2-7x+4  =  0,     3#2  +  2a' -  2  =  0. 

3.  The  equation  xm  +  ax  +  b  =  0.     This  may  be  solved  by  constructing 

the  curves  y  =  xm,  y=-  ax—b.     Verify  the  following  table  for  the  number  of 

roots  of 

xm  +  ax  +  b  =  0: 

b  positive,  two  or  none, 


[b  positive,  two  o 

(a)  m  even  <  ,  x 
v  '  [b  negative,  two ' 

,  ,    (a  positive,  one, 

(b)  in  odd   \  ,.        . 

{ a  negative,  tare 


ree  or  one. 
Construct  numerical  examples  to  illustrate  all  possible  cases. 

4.  Show  that  the  equation  tan^=a^  +  6has  always  an  infinite  number 
of  roots. 

5.  Determine  the  number  of  roots  of 

sin#  =  .r,     sin#  =  J#,     sinx  =  |x,    Bin  $  =  j^x , 

6.  Show  that  if  a  is  small  and  positive  (e.g.  a  =  '01),  the  equation 

x  —  a  =  \i7  sin2  x 
has  three  roots.     Consider  also  the  case  in  which  a  is  small  and  negative. 
Explain  how  the  number  of  roots  varies  as  a  varies. 


30,  31]  FUNCTIONS   OF   REAL   VARIABLES  59 

31.  Functions  of  two  variables  and  their  graphical 
representation.  In  §  20  we  considered  two  variables  connected 
by  a  relation.  We  may  similarly  consider  three  variables  (x,  y, 
and  z)  connected  by  a  relation  such  that  when  the  values  of  x  and 
y  are  both  given,  the  value  or  values  of  z  are  known.  In  this  case 
we  call  z  a  function  of  the  two  variables  x  and  y;  x  and  y  the 
independent  variables,  z  the  dependent  variable ;  and  we  express 
this  dependence  of  z  upon  x  and  y  by  writing 

The  remarks  of  §  20  may  all  be  applied,  mutatis  mutandis,  to  this 
more  complicated  case. 

The  method  of  representing  such  functions  of  two  variables 
graphically  is  exactly  the  same  in  principle  as  in  the  case  of 
functions  of  a  single  variable.  We  must  take  three  axes,  OX,  0  Y, 
OZ  in  space  of  three  dimensions,  each  axis  being  perpendicular 
to  the  other  two.  The  point  (a,  b,  c)  is  the  point  whose  distances 
from  the  planes  YOZ,  ZOX,  XOY,  measured  parallel  to  OX,  OY, 
OZ,  are  a,  b,  and  c.  Regard  must  of  course  be  paid  to  sign, 
lengths  measured  in  the  directions  OX,  OY,  OZ  being  regarded 
as  positive.  The  definitions  of  coordinates,  axes,  origin  are  the 
same  as  before. 

Now  let  z  =f(x,  y). 

As  x  and  y  vary,  the  point  (x,  y,  z)  will  move  in  space.  The 
aggregate  of  all  the  positions  it  assumes  is  called  the  locus  of  the 
point  (x,  y,  z)  or  the  graph  of  the  function  z  =f(x,  y).  When  the 
relation  between  x,  y,  and  z  which  defines  z  can  be  expressed  in  an 
analytical  formula,  this  formula  is  called  the  equation  of  the  locus. 
It  is  easy  to  show,  for  example,  that  the  equation 

Ax  +  By+Cz+D  =  0 

(the  general  equation  of  the  first  degree)  represents  a  plane,  and 
that  the  equation  of  any  plane  is  of  this  form.     The  equation 

(x-ay-  +  (y-{3y+(z-ryy  =  P\ 

or  x2+y2  +  z2  +  2Fx+2Gy  +  2Hz  +  C=0, 

where  F2  +  G2  +  H2  —  C  >  0,  represents  a  sphere ;  and  so  on.  For 
proofs  of  these  propositions  we  must  again  refer  to  text-books  of 
Analytical  Geometry. 


GO  FUNCTIONS  OF   REAL  VARIABLES  [il 

32.    Curves  in  a  plane.    We  have  hitherto  used  the  notation 

y-/(«) (!) 

to  express  functional  dependence  of  y  upon  x.  It  is  evident  that 
this  notation  is  most  appropriate  in  the  case  in  which  y  is  ex- 
pressed explicitly  in  terms  of  x  by  means  of  a  formula,  as  when 
for  example 

y  —  x-,     sin  x,     a  cos2  x  +  b  sin2  x. 

We  have  however  very  often  to  deal  with  functional  relations 
which  it  is  impossible  or  inconvenient  to  express  in  this  form. 
If,  for  example,  y5  —  y  —  x  =  0  or  x5  +  y5  —  ay  =  0,  it  is  known 
to  be  impossible  to  express  y  explicitly  as  an  algebraical  function 
of  x.     If 

X-  +  y*  +  2Gx  +  2Fy  +C  =  0, 
y  can  indeed  be  so  expressed,  viz.  by  the  formula 
y  =  -  F+  V(^2  -  «2  -  2Gx  -  C); 
but  the  functional  dependence  of  y  upon  x  is  better  and  more 
simply  expressed  by  the  original  equation. 

It  will  be  observed  that  in  these  two  cases  the  functional 
relation  is  fully  expressed  by  equating  a  function  of  the  two 
variables  x  and  y  to  zero,  i.e.  by  means  of  an  equation 

ffry)=o (2). 

We  shall  adopt  this  equation  as  the  standard  method  of 
expressing  the  functional  relation.  It  includes  the  equation  (1) 
as  a  special  case,  since  y—f(x)  is  a  special  form  of  a  function  of  x 
and  y.  We  can  then  speak  of  the  locus  of  the  point  (x,  y)  subject 
tof(x,  y)  =  0,  the  graph  of  the  function  y  defined  by  f(x,  y)  =  0, 
the  curve  or  locus  f(x,  y)  —  0,  and  the  equation  of  this  curve  or 
locus. 

There  is  another  method  of  representing  curves  which  is  often 
useful.  Suppose  that  x  and  y  are  both  functions  of  a  third 
variable  t,  which  is  to  be  regarded  as  essentially  auxiliary  and 
devoid  of  any  particular  geometrical  significance.     We  may  write 

x=f(t),    y  =  F{t)  (3). 

If  a  particular  value  is  assigned  to  t,  the  corresponding  values  of 
x  and  of  y  are  known.     Each  pair  of  such  values  defines  a  point 


32,  33]  FUNCTIONS   OF   REAL   VARIABLES  61 

(x,  y).  If  Ave  construct  all  the  points  which  correspond  in  this 
way  to  different  values  of  t,  we  obtain  the  graph  of  the  locus 
defined  by  the  equations  (3).     Suppose  for  example 

x  =  a  cos  t,  y  =  a  sin  t. 
Let  t  vary  from  0  to  2ir.  Then  it  is  easy  to  see  that  the  point 
(x,  y)  describes  the  circle  whose  centre  is  the  origin  and  whose 
radius  is  a.  If  t  varies  beyond  these  limits,  (x,  y)  describes  the 
circle  over  and  over  again.  We  can  in  this  case  at  once  obtain 
a  direct  relation  between  x  and  y  by  squaring  and  adding:  we 
find  that  x2  +  y2  =  a2,  t  being  now  eliminated. 

Examples  XVIII.  1.  The  points  of  intersection  of  the  two  curves  whose 
equations  are  f(x,  y)=0,  <£  (x,  y)  =  0,  where  /  and  cf>  are  polynomials,  can  be 
determined  if  these  equations  can  be  solved  as  a  pair  of  simultaneous  equations 
in  x  and  y.  The  solution  generally  consists  of  a  finite  number  of  pairs  of 
values  of  x  and  y.  The  two  equations  therefore  generally  represent  a  finite 
number  of  isolated  points. 

2.  Trace  the  curves  (x+y)2=l,  xy  —  l,  x2—y2=l. 

3.  The  curve  f(x,y)  +  'K<p(x,y)  =  0  represents  a  curve  passing  through 
the  points  of  intersection  of/=0  and  <£  =  0. 

4.  What  loci  are  represented  by 

(a)    x=at+b,    y  =  ct+cl,      (j8)    xla  =  2t/(l  +  t2\    yla=(l-t2)/(l  +  t2), 
when  t  varies  through  all  real  values  ? 

33.  Loci  in  space.  In  space  of  three  dimensions  there  are 
two  fundamentally  different  kinds  of  loci,  of  which  the  simplest 
examples  are  the  plane  and  the  straight  line. 

A  particle  which  moves  along  a  straight  line  has  only  one 
degree  of  freedom.  Its  direction  of  motion  is  fixed ;  its  position 
can  be  completely  fixed  by  one  measurement  of  position,  e.g.  by 
its  distance  from  a  fixed  point  on  the  line.  If  we  take  the  line  as 
our  fundamental  line  A  of  Chap.  I,  the  position  of  any  of  its  points 
is  determined  by  a  single  coordinate  x.  A  particle  which  moves 
in  a  plane,  on  the  other  hand,  has  two  degrees  of  freedom;  its 
position  can  only  be  fixed  by  the  determination  of  two  coordinates. 

A  locus  represented  by  a  single  equation 

z  =/0,  y) 
plainly  belongs  to  the  second  of  these  two  classes  of  loci,  and  is 
called  a  surface.     It  may  or  may  not  (in  the  obvious  simple  cases 


62  FUNCTIONS   OF   REAL   VARIABLES  [ll 

it  will)  satisfy  our  common-sense  notion  of  what  a  surface 
should  be. 

The  considerations  of  §  31  may  evidently  be  generalised  so 
as  to  give  definitions  of  a  function  j  (x,  y,  z)  of  three  variables  (or 
of  functions  of  any  number  of  variables).  And  as  in  §  32  we 
agreed  to  adopt  f(x,  y)  =  0  as  the  standard  form  of  the  equation 
of  a  plane  curve,  so  now  we  shall  agree  to  adopt 

f(x,y,z)  =  0 

as  the  standard  form  of  equation  of  a  surface. 

The  locus  represented  by  two  equations  of  the  form  z  =f(x,  y) 
or  fix,  y,  z)  =  0  belongs  to  the  first  class  of  loci,  and  is  called 
a  curve.  Thus  a  straight  line  may  be  represented  by  two  equations 
of  the  type  Ax  +  By+Cz  +  D  =  0.  A  circle  in  space  may  be 
regarded  as  the  intersection  of  a  sphere  and  a  plane ;  it  may 
therefore  be  represented  by  two  equations  of  the  forms 

(x  -  a)2  +  (y  -  /3)2  +  (z-  yf  =  p",     Ax  +  By+Cz  +  D  =  0. 

Examples  XIX.  1.  What  is  represented  by  three  equations  of  the  type 
f(x,y,z)  =  0l 

2.  Three  linear  equations  in  general  represent  a  single  point.  "What  are 
the  exceptional  cases  ? 

3.  What  are  the  equations  of  a  plane  curve /(.r,  y)=0  in  the  plane  XOY, 
when  regarded  as  a  curve  in  space  1     [/(#,  y)=0,  2  =  0.] 

4.  Cylinders.  What  is  the  meaning  of  a  single  equation  f(x,y)  =  0, 
considered  as  a  locus  in  space  of  three  dimensions  ? 

[All  points  on  the  surface  satisfy  /  (x,  y)  =  0,  whatever  be  the  value  of  z.  The 
curve  /(#,  y)=0,  2=0  is  the  curve  in  which  the  locus  cuts  the  plane  XOY. 
The  locus  is  the  surface  formed  by  drawing  lines  parallel  to  OZ  through  all 
points  of  this  curve.     Such  a  surface  is  called  a  cylinder.] 

5     Graphical  representation  of  a  surface  on  a  plane.  Contour  Maps. 

It  might  seem  to  be  impossible  to  represent  a  surface  adequately  by  a 
drawing  on  a  plane  ;  and  so  indeed  it  is :  but  a  very  fair  notion  of  the 
nature  of  the  surface  may  often  be  obtained  as  follows.  Let  the  equation  of 
the  surface  be  z=f(x,  y). 

If  we  give  z  a  particular  value  a,  we  have  an  equation  f(x,  y)  =  a,  which 
we  may  regard  as  determining  a  plane  curve  on  the  paper.  We  trace  this 
curve  and  mark  it  (a).     Actually  the  curve  (a)  is  the  projection  on  the  plane 


33] 


FUNCTIONS   OF   REAL   VARIABLES 


63 


XOY  of  the  section  of  the  surface  by  the  plane  z  —  a.  We  do  this  for  all 
values  of  a  (practically,  of  course,  for  a  selection  of  values  of  a).  We  obtain 
some  such  figure  as  is  shown  in  Fig.  17.  It  will  at  once  suggest  a  contoured 
Ordnance  Survey  map :  and  in  fact  this  is  the  principle  on  which  such  maps 
are  constructed.  The  contour  line  1000  is  the  projection,  on  the  plane  of  the 
sea  level,  of  the  section  of  the  surface  of  the  land  by  the  plane  parallel  to  the 
plane  of  the  sea  level  and  1000  ft.  above  it*. 


3000 


6.  Draw  a  series  of  contour  lines  to  illustrate  the  form  of  the  surface 
2z=3.vy. 

7.  Right  circular  cones.  Take  the  origin  of  coordinates  at  the 
vertex  of  the  cone  and  the  axis  of  z  along  the  axis  of  the  cone ;  and  let  a  be 
the  semi-vertical  angle  of  the  cone.  The  equation  of  the  cone  (which  must 
be  regarded  as  extending  both  ways  from  its  vertex)  is  .r2  +  ?/2- c2tan2  a  =  0. 

8.  Surfaces  of  revolution  in  general.  The  cone  of  Ex.  7  cuts  ZOX  in 
two  lines  whose  equations  may  be  combined  in  the  equation  x2  =  z2  tan2  a. 
That  is  to  say,  the  equation  of  the  surface  generated  by  the  revolution  of 
the  curve  y  =  0,  #2=22tan2  a  round  the  axis  of  z  is  derived  from  the  second  of 
these  equations  by  changing  x2  into  x2+y2.  Show  generally  that  the  equation 
of  the  surface  generated  by  the  revolution  of  the  curve  y  =  0,  x=f(z),  round 
the  axis  of  z,  is 

9.  Cones  in  general.  A  surface  formed  by  straight  lines  passing 
through  a  fixed  point  is  called  a  cone:  the  point  is  called  the  vertex.  A 
particular  case  is  given  by  the  right  circular  cone  of  Ex.  7.  Show  that  the 
equation  of  a  cone  whose  vertex  is  0  is  of  the  form  f(z/x,  z/y)  =  0,  and  that  any 
equation  of  this  form  represents  a  cone.  [If  (x,  y,  z)  lies  on  the  cone,  so  must 
(X.r,  Ay,  \z),  for  any  value  of  X.] 


We  assume  that  the  effects  of  the  earth's  curvature  may  be  neglected. 


64 


FUNCTIONS   OF   REAL   VARIABLES 


[II 


10.    Ruled  surfaces. 

composed  of  straight  lines. 


Cylinders  and  cones  are  special  cases  of  surfaces 
Such  surfaces  are  called  ruled  surfaces. 


The  two  equations 

x=az  +  b,    y  =  cz+d (1) 

represent  the  intersection  of  two  planes,  i.e.  a  straight  line.  Now  suppose 
that  a,  b,  c,  d  instead  of  being  fixed  are  functions  of  an  auxiliary  variable  t. 
For  any  particular  value  of  t  the  equations  (1)  give  a  line.  As  t  varies, 
this  line  moves  and  generates  a  surface,  whose  equation  may  be  found  by 
eliminating  t  between  the  two  equations  (1).  For  instance,  in  Ex.  7  the 
equations  of  the  line  which  generates  the  cone  are 

x=z  tan  a  cos  t,    y  =  z  tan  a  sin  t, 

where  t  is  the  angle  between  the  plane  XOZ  and  a  plane  through  the  line  and 
the  axis  of  z. 

Another  simple  example  of  a  ruled  surface  may  be  constructed  as  follows. 
Take  two  sections  of  a  right  circular  cylinder  perpendicular  to  the  axis  and 
at  a  distance  I  apart  (Fig.  18  a).  We  can  imagine  the  surface  of  the  cylinder 
to  be  made  up  of  a  number  of  thin  parallel  rigid  rods  of  length  I,  such  as  PQ, 
the  ends  of  the  rods  being  fastened  to  two  circular  rods  of  radius  a. 

Now  let  us  take  a  third  circular  rod  of  the  same  radius  and  place  it 
round  the  surface  of  the  cylinder  at  a  distance  h  from  one  of  the  first  two 
rods  (see  Fig.  18  a,  where  Pq  —  li).  Unfasten  the  end  Q  of  the  rod  PQ  and 
turn  PQ  about  P  until  Q  can  be  fastened  to  the  third  circular  rod  in  the 
position  Q'.     The  angle  qOQ'  =  a  in  the  figure  is  evidently  given  by 

P  -  h2  =  qQ'2  =  (2a  sin  \af. 

Let  all  the  other  rods  of  which  the  cylinder  was  composed  be  treated  in  the 
same  way.  We  obtain  a  ruled  surface  whose  form  is  indicated  in  Fig.  18  b. 
It  is  entirely  built  up  of  straight  lines;  but  the  surface  is  curved  everywhere, 
and  is  in  general  shape  not  unlike  certain  forms  of  table-napkin  rings  (Fig.  18c). 


fZ      ^ 

p 

~           9 

-^ 

Q 

Fie.  18  a. 


Fig.  186. 


Fig.  18  c. 


FUNCTIONS   OF   REAL   VARIABLES  65 


MISCELLANEOUS   EXAMPLES   ON   CHAPTEI1   IL 

1 .  Show  that  if  y  =/ {x)  =  {ax  +  b)/{cx  -  a)  then  x  =f  {y). 

2.  If  f{x)=f{  —  x)  for  all  values  of  x,f{x)  is  called  an  even  function. 
Iff(x)  =  —/(  —  #),  it  is  called  an  odd  function.  Show  that  any  function  of  x, 
denned  for  all  values  of  x,  is  the  sum  of  an  even  and  an  odd  function  of  x. 

[Use  the  identity /(*)  =  \  tf(«) +/(-*)} +*{/(*)  -/(-*)}•] 

3.  Draw  the  graphs  of  the  functions 

3  sin  x  +  4  cos  x,     sin  f  -t=  sin  a; ) .      {Math.  Trip.  1 896.) 

4.  Draw  the  graphs  of  the  functions 

sin  x {a cos2 x+b  sin2*'),    - — —  («  cos2 #  +  6  sin2 #),     ( — ' 
"a;  \  x 

5.  Draw  the  graphs  of  the  functions  #  [l/.r],  [.f]/^. 

6.  Draw  the  graphs  of  the  functions 

(i)     arc  cos  {2x2  — 1)-2  arc  cos  x, 

(ii)     arc  tan arc  tan  a  —  arc  tan  x, 

I  -ax 

where  the  symbols  arc  cos  a,  arc  tan  a  denote,  for  any  value  of  a,  the  least 
positive  (or  zero)  angle,  whose  cosine  or  tangent  is  a. 

7.  Verify  the  following  method  of  constructing  the  graph  of  /{<£  (x)}  by 
means  of  the  line  y  =  x  and  the  graphs  of  /  {x)  and  0  (x) :  take  OA  =  x  along 
OX,  draw  AB  parallel  to  OF  to  meet  y  =  (j>  (x)  in  B,  BC  parallel  to  OX  to 
meet  y—x  in  C,  CD  parallel  to  OY  to  meet  y=f{x)  in  Z),  and  DP  parallel  to 
OX  to  meet  AB  in  P;  then  P  is  a  point  on  the  graph  required. 

8.  Show  that  the  roots  of  x3+px  +  q  =  0  are  the  abscissae  of  the  points  of 
intersection  (other  than  the  origin)  of  the  parabola  y=x2  and  the  circle 

x2+y2  +  {p~  \)y  +  qx=0. 

9.  The  roots  of  xi  +  nx3  +px2  +  qx  +  r  =  0  are  the  abscissae  of  the  points  of 
intersection  of  the  parabola  x2=y  —  hix  and  the  circle 

x2+y2  +  {l?i2-jtpn  +  %n  +  q)x  +  {p-l  -  \n2)y-s!-r=0. 

10.  Discuss  the  graphical  solution  of  the  equation 

xm+ax2-\-bx+c  =  0 

by  means  of  the  curves  y  =  xm,  y  =  —  ax2  —  bx  —  c.     Draw  up  a  table  of  the 
various  possible  numbers  of  roots. 

11.  Solve  the  equation  sec  #  +  cosec  0  =  2  x/2;  and  show  that  the  equation 
sec  0  +  cosec  6  =  o  has  two  roots  between  0  and  2tt  if  c2<8  and  four  if  c2>8. 


66  FUNCTIONS   OF   KEAL   VARIABLES  [il 

12.  Show  that  the  equation 

2x  =  (1n  +  \)TT  (1-cos^), 
where  n  is  a  positive  integer,  has   2?i  +  3  roots  and   no   more,  indicating 
their  localities  roughly.  {Math.  Trip.  1896.) 

13.  Show  that  the  equation  §.«sin^  =  l  has  four  roots  between  -it 
and  7r. 

14.  Discuss  the  number  and  values  of  the  roots  of  the  equations 

(1)     cot.^+.r-|7T  =  0,         (2)     ^2  +  sin2.r=l,         (3)     t^nx=2x/(l-\-x°-), 
(4)     sin.r-A,  +  ^'3=0,        (5)     (1-cos.r)  tana-.r+sino7=0. 

15.  The  polynomial  of  the  second  degree  which  assumes,  when  x=a,  b,  c 
the  values  a,  0,  y  is 

(x-b)(x-c)        (x-c)(x-a)        (x  -  a)  (x  —  b) 
a(a-b)(a-c)+P  (b-c)(b-a)  +y  (c-a)(c-b)' 

Give  a  similar  formula  for  the  polynomial  of  the  (?i-l)th  degree  which 
assumes,  when  x=ax,a^  ...  an,  the  values  a1?  a2,  ...  an. 

16.  Find  a  polynomial  in  x  of  the  second  degree  which  for  the  values 
0,  1,  2  of  x  takes  the  values  1/c,  l/(c  +  l),  l/(c  +  2);  and  show  that  when 
x  =  c  +  2  its  value  is  l/(c  +  l).  (Math.  Trip.  1911.) 

17.  Show  that  if,.r  is  a  rational  function  of  y,  and  y  is  a  rational  function 
of  x,  then  Axy  +  Bx  +  Cy  +  D=0. 

18.  If  y  is  an  algebraical  function  of  x,  then  x  is  an  algebraical  function 
of  y- 

19.  Verify  that  the  equation 

.t*2 


COS  i  nX  =  1  -  • 


•+<■-» -s/C-r) 


is  approximately  true  for  all  values  of  x  between  0  and  1.     [Take  „r=0,  J,  J, 
i>  |>  f  >  1>  and  use  tables.     For  which  of  these  values  is  the  formula  exact1?] 

20.  What  is  the  form  of  the  graph  of  the  functions 

*  =  !>] +  l>]>     z  =  x+y-[x]-[y]l 

21.  What  is  the  form  of  the  graph  of  the  functions  z  =  smx  +  smy, 
z = sin  x  sin  y,  z  =  sin  xy,  z = sin  {x1  +y2)  ? 

22.  Geometrical  constructions  for  irrational  numbers.  In  Chapter  I 
we  indicated  one  or  two  simple  geometrical  constructions  for  a  length  equal  to 
N/2,  starting  from  a  given  unit  length.  We  also  showed  how  to  construct 
the  roots  of  any  quadratic  equation  axi  +  2bx  +  c  =  Q,  it  being  supposed  that 
we  can  construct  lines  whose  lengths  are  equal  to  any  of  the  ratios  of  the 
coefficients  a,  b,  c,  as  is  certainly  the  case  if  a,  b,  c  are  rational.  All  these  con- 
structions were  what  may  be  called  Euclidean  constructions  ;  they  depended 
on  the  ruler  and  compasses  only. 


FUNCTIONS   OF   REAL   VARIABLES  67 

It  is  fairly  obvious  that  we  can  construct  by  these  methods  the  length 
measured  by  any  irrational  number  which  is  defined  by  any  combination  of 
square  roots,  however  complicated.     Thus 

7/     //17  +  3V11N  //17-3V11M 

V1VV17-3V1V  VV17  +  3V1VJ 
is  a  case  in  point.  This  expression  contains  a  fourth  root,  but  this  is  of 
course  the  square  root  of  a  square  root.  We  should  begin  by  constructing 
x/11,  e.g.  as  the  mean  between  1  and  11 :  then  17  +  3^11  and  17-3>/ll,  and 
so  on.  Or  these  two  mixed  surds  might  be  constructed  directly  as  the  roots  of 
A-2_  34^+190  =  0. 

Conversely,  only  irrationals  of  this  kind  can  be  constructed  by  Euclidean 
methods.  Starting  from  a  unit  length  we  can  construct  any  rational  length. 
And  hence  we  can  construct  the  line  Ax  +  By  +  C=0,  provided  that  the  ratios 
of  A,  B,  C  are  rational,  and  the  circle 

(.r_a)2  +  (y_/3)2  =  p2 

(or  x2+y2  +  2gx  +  2fy+c  =  0),  provided  that  a,  /3,  p  are  rational,  a  condition 
which  implies  that  g,  f,  c  are  rational. 

Now  in  any  Euclidean  construction  each  new  point  introduced  into  the 
figure  is  determined  as  the  intersection  of  two  lines  or  circles,  or  a  line  and 
a  circle.     But  if  the  coefficients  are  rational,  such  a  pair  of  equations  as 

Ax  +  By  +  C=0,  x2+y2  +  2gx  +  2fy+c  =  0 
give,  on  solution,  values  of  x  and  y  of  the  form  m  +  n^p,  where  m,  n,  p  are 
rational :  for  if  we  substitute  for  x  in  terms  of  y  in  the  second  equation  we 
obtain  a  quadratic  in  y  with  rational  coefficients.  Hence  the  coordinates  of 
all  points  obtained  by  means  of  lines  and  circles  with  rational  coefficients 
are  expressible  by  rational  numbers  and  quadratic  surds.  And  so  the  same 
is  true  of  the  distance  sj  {{xi  —  x2)2-\-{yi-y-2)2}  between  any  two  points  so 
obtained. 

"With  the  irrational  distances  thus  constructed  we  may  proceed  to  construct 
a  number  of  lines  and  circles  whose  coefficients  may  now  themselves  involve 
quadratic  surds.  It  is  evident,  however,  that  all  the  lengths  which  we  can 
construct  by  the  use  of  such  lines  and  circles  are  still  expressible  by  square 
roots  only,  though  our  surd  expressions  may  now  be  of  a  more  complicated 
form.  And  this  remains  true  however  often  our  constructions  are  repeated. 
Hence  Euclidean  methods  will  construct  any  surd  expression  involving  square 
roots  only,  and  no  others. 

One  of  the  famous  problems  of  antiquity  was  that  of  the  duplication  of 
the  cube,  that  is  to  say  of  the  construction  by  Euclidean  methods  of  a 
length  measured  by  $2.  It  can  be  shown  that  $2  cannot  be  expressed  by 
means  of  any  finite  combination  of  rational  numbers  and  square  roots,  and  so 
that  the  problem  is  an  impossible  one.  See  Hobson,  Squaring  the  Circle, 
pp.  47  et  seq. ;  the  first  stage  of  the  proof,  viz.  the  proof  that  S]2  cannot  be  a 
root  of  a  quadratic  equation  ax2  +  2bx  +  c  =  0  with  rational  coefficients,  was 
given  in  Ch.  I  (Misc.  Exs.  24). 

5—2 


68  FUNCTIONS   OF   REAL   VARIABLES  [ll 

23.  Approximate  quadrature  of  the  circle.  Let  0  be  the  centre  of 
a  circle  of  radius  R.  On  the  tangent  at  A  take  AP=11R  &m\AQ  =  'i?LR, 
in  the  same  direction.  On  AO  take  AN=OP  and  draw  JVM  parallel  to 
OQ  and  cutting  AP  in  M.     Show  that 

/U//tf  =  ifv/146, 

and  that  to  take  AM  as  being  equal  to  the  circumference  of  the  circle  would 
lead  to  a  value  of  tt  correct  to  five  places  of  decimals.  If  R  is  the  earth's 
radius,  the  error  in  supposing  AM to  be  its  circumference  is  less  than  11  yards. 

24.  Show  that  the  only  lengths  which  can  be  constructed  with  the  ruler 
only,  starting  from  a  given  unit  length,  are  rational  lengths. 

25.  Constructions  for  Z]2.  0  is  the  vertex  and  S  the  focus  of  the 
parabola  y2  =  4x,  and  P  is  one  of  its  points  of  intersection  with  the  parabola 
x2=2y.  Show  that  OP  meets  the  latus  rectum  of  the  first  parabola  in  a  point 
Q  such  that  SQ=$2. 

26.  Take  a  circle  of  unit  diameter,  a  diameter  OA  and  the  tangent  at  A. 
Draw  a  chord  OBC  cutting  the  circle  at  B  and  the  tangent  at  C.  On  this 
line  take  OM=BC.  Taking  0  as  origin  and  OA  as  axis  of  x,  show  that  the 
locus  of  M  is  the  curve 

(x2 + y2)  x — y2  =  0 

(the  Oissoid  of  Diodes).  Sketch  the  curve.  Take  along  the  axis  of  y  a  length 
0D  =  2.  Let  AD  cut  the  curve  in  P  and  OP  cut  the  tangent  to  the  circle 
at  A  in  Q.    Show  that  A  Q=%2. 


CHAPTER   III 

COMPLEX    NUMBERS 

34.     Displacements  along  a  line  and  in  a  plane.     The 

*  real  number '  x,  with  which  we  have  been  concerned  in  the  two 
preceding  chapters,  may  be  regarded  from  many  different  points 
of  view.  It  may  be  regarded  as  a  pure  number,  destitute  of 
geometrical  significance,  or  a  geometrical  significance  may  be 
attached  to  it  in  at  least  three  different  ways.  It  may  be  re- 
garded as  the  measure  of  a  length,  viz.  the  length  A0P  along  the 
line  A  of  Chap.  I.  It  may  be  regarded  as  the  mark  of  a  point, 
viz.  the  point  P  whose  distance  from  A0  is  x.  Or  it  may  be 
regarded  as  the  measure  of  a  displacement  or  change  of  position 
on  the  line  A.  It  is  on  this  last  point  of  view  that  we  shall  now 
concentrate  our  attention. 

Imagine  a  small  particle  placed  at  P  on  the  line  A  and  then 
displaced  to  Q.  We  shall  call  the  displacement  or  change  of 
position  which  is  needed  to  transfer  the  particle  from  P  to  Q  the 
displacement  PQ.  To  specify  a  displacement  completely  three 
things  are  needed,  its  magnitude,  its  sense  forwards  or  backwards 
along  the  line,  and  what  may  be  called  its  point  of  application, 
i.e.  the  original  position  P  of  the  particle.  But,  when  we  are 
thinking  merely  of  the  change  of  position  produced  by  the  dis- 
placement, it  is  natural  to  disregard  the  point  of  application  and 
to  consider  all  displacements  as  equivalent  whose  lengths  and 
senses  are  the  same.  Then  the  displacement  is  completely  speci- 
fied by  the  length  PQ  =  x,  the  sense  of  the  displacement  being 
fixed  by  the  sign  of  x.  We  may  therefore,  without  ambiguity, 
speak  of  the  displacement  [#]*,  and  we  may  write  PQ  =  [x\. 

*  It  is  hardly  necessary  to  caution  the  reader  against  confusing  this  use  of  the 
symbol  [x]  and  that  of  Chap.  II  (Exs.  xvi.  and  Misc.  Exs.). 


70  COMPLEX    NUMBERS  [ill 

We  use  the  square  bracket  to  distinguish  the  displacement  [x] 
from  the  length  or  number  x*.  If  the  coordinate  of  P  is  a,  that 
of  Q  will  be  a  +  cc;  the  displacement  [x]  therefore  transfers  a 
particle  from  the  point  a  to  the  point  a  +  x. 

We  come  now  to  consider  displacements  in  a  plane.  We  may- 
define  the  displacement  PQ  as  before.  But  now  more  data  are 
required  in  order  to  specify  it  completely.  We  require  to  know : 
(i)  the  magnitude  of  the  displacement,  i.e.  the  length  of  the 
straight  line  PQ ;  (ii)  the  direction  of  the  displacement,  which  is 
determined  by  the  angle  which  PQ  makes  with  some  fixed  line  in 
the  plane;  (iii)  the  sense  of  the  displacement;  and  (iv)  its  point 
of  application.  Of  these  requirements  we  may  disregard  the 
fourth,  if  we  consider  two  displacements  as  equivalent  if  they  are 
the  same  in  magnitude,  direction,  and  sense.  In  other  words,  if 
PQ  and  RS  are  equal  and  parallel,  and  the  sense  of  motion  from 
P  to  Q  is  the  same  as  that  of 
motion  from  R  to  S,  we  regard 
the  displacements  PQ  and  RS  as 
equivalent,  and  write 

PQ=M 

Now   let   us  take  any  pair  of 
coordinate  axes  in  the  plane  (such      "  Fi    19 

as  OX,  0Y  in  Fig.  19).  Draw  a 
line  OA  equal  and  parallel  to  PQ,  the  sense  of  motion  from  0 
to  A  being  the  same  as  that  from  P  to  Q.  Then  PQ  and  OA 
are  equivalent  displacements.  Let  x  and  y  be  the  coordinates 
of  A.  Then  it  is  evident  that  OA  is  completely  specified 
if  x  and  y  are  given.  We  call  OA  the  displacement  [x,  y]  and 
write 

0A  =  PQ  =  RS=[x,y]. 

*  Strictly  speaking  we  ought,  by  some  similar  difference  of  notation,  to  dis- 
tinguish the  actual  length  x  from  the  number  x  which  measures  it.  The  reader 
will  perhaps  be  inclined  to  consider  such  distinctions  futile  and  pedantic.  But 
increasing  experience  of  mathematics  will  reveal  to  him  the  great  importance  of 
distinguishing  clearly  between  things  which,  however  intimately  connected,  are  not 
the  same.  If  cricket  were  a  mathematical  science,  it  would  be  very  important  to 
distinguish  between  the  motion  of  the  batsman  between  the  wickets,  the  run  which 
he  scores,  and  the  mark  which  is  put  down  in  the  score-book. 


34-3G]  COMPLEX   NUMBERS  71 

35.  Equivalence  of  displacements.  Multiplication  of 
displacements  by  numbers.  If  f  and  rj  are  the  coordinates 
of  P,  and  (■'  and  77'  those  of  Q,  it  is  evident  that 

sb  =  g  —  £     y  =  V  -  77. 

The  displacement  from  (£,  77)  to  (£',  77')  is  therefore 

[£'-£  */  —  «?]■ 

It  is  clear  that  two  displacements  \x,  y\  \x',  y"\  are  equivalent 
if,  and  only  if,  x  =  x',y  =  y.     Thus  [x,  y]  =  [x,  y']  if  and  only  if 

x  =  x',  y  =  y' (1). 

The  reverse  displacement  QP  would  be  [£  -  f ',  77  —  77'],  and  it 
is  natural  to  agree  that 

QP  =  -PQ, 

these  equations  being  really  definitions  of  the  meaning  of  the 
symbols  —  [£'  —  £  77'  —  77],  —  PQ.     Having  thus  agreed  that 

- I>>  y]  =  [-  ^  -  yl 

it  is  natural  to  agree  further  that 

a[x,  y]  =  [ax,  ay] (2), 

where  a  is  any  real  number,  positive  or  negative.  Thus  (Fig.  19) 
if  OB  =  -\OA  then 

dB  =  -lOA  =  -i[x,y-\  =  [-\x,-\y\ 

The  equations  (1)  and  (2)  define  the  first  two  important  ideas 
connected  with  displacements,  viz.  equivalence  of  displacements, 
and  multiplication  of  displacements  by  numbers. 

36.  Addition  of  displacements.  We  have  not  yet  given 
any  definition  which  enables  us  to  attach  any  meaning  to  the 
expressions 

PQ  +  P7®,   [x,y]  +  W,y'l 

Common  sense  at  once  suggests  that  we  should  define  the  sum 
of  two  displacements  as  the  displacement  which  is  the  result 
of  the  successive  application  of  the  two  given  displacements.     In 


72 


COMPLEX    NUMBERS 


[III 


other  words,  it  suggests  that  if  QQX  be  drawn  equal  and  parallel 
to  P'Q',  so  that  the  result  of  successive  displacements  PQ,  P'Q'  on 
a  particle  at  P  is  to  transfer  it  first  to  Q  and  then  to  Ql}  then  we 
should  define  the  sum  of  PQ  and  P'Q'  as  being  PQ^  If  then  we 
draw  OA  equal  and  parallel  to  PQ,  and  OB  equal  and  parallel  to 
PQ',  and  complete  the  parallelogram  OAGB,  we  have 

PQ  +  PQ '  =  PQl  =  OA  +  OB=OC. 

.P' 


Fig.  20. 

Let  us  consider  the  consequences  of  adopting  this  definition. 
If  the  coordinates  of  B  are  x ,  y',  then  those  of  the  middle  point  of 
AB  are  ^  (x  +  x),  ^(y+y'),  and  those  of  C  are  x+x',  y+y'.   Hence 
[x,y]  +  [x',y']  =  [x  +  x',y  +  y'] (3), 

which  may  be  regarded  as  the  symbolic  definition  of  addition  of 
displacements.     We  observe  that 

[ob,  y']  +  [x,  y]  =  [x  +  x,y'  +  y] 

=  \x  +  of,  y  +  y.]  =  [x,  y]  +  [x ,  y'] 

In  other  words,  addition  of  displacements  obeys  the  commutative 
laiv  expressed  in  ordinary  algebra  by  the  equation  a  +  b  =  b  +  a. 
This  law  expresses  the  obvious  geometrical  fact  that  if  we  move 
from  P  first  through  a  distance  PQ2  equal  and  parallel  to  P'Q', 
and  then  through  a  distance  equal  and  parallel  to  PQ,  we  shall 
arrive  at  the  same  point  Q1  as  before. 


36]  COMPLEX  NUMBERS  73 

In  particular 

[*,y]=[«,0]  +  [0,y]  (4). 

Here  [x,  0]  denotes  a  displacement  through  a  distance  x  in 
a  direction  parallel  to  OX.  It  is  in  fact  what  we  previously 
denoted  by  [a],  when  we  were  considering  only  displacements 
along  a  line.  We  call  [x,  0]  and  [0,  y]  the  components  of  [x,  y], 
and  [x,  y]  their  resultant. 

When  we  have  once  defined  addition  of  two  displacements, 
there  is  no  further  difficulty  in  the  way  of  •defining  addition  of 
any  number.     Thus,  by  definition, 

[x,  y]  +  [x\  y']  +  [x",  y"]  =  ([>,  y]  +  [x',  y'})  +  [x",  f] 

=  [x  +  x',y  +  y']  +  [x",  y"]  =[x  +  x'  +  x",  y  +  y'  +  y"]. 

We  define  subtraction  of  displacements  by  the  equation 

[x,  y]  -  [>',  y']  =  [x,  y]  +  (-  [>',  y'}) (5), 

which  is  the  same  thing  as  [x,  y]  +  [—  x',  —  y']  or  as  [x  —  x',y  —  y'\ 
In  particular 

[x,  y]-[x,y]  =  [0,  0]. 

The  displacement  [0,  0]  leaves  the  particle  where  it  was;  it  is 
the  zero  displacement,  and  we  agree  to  write  [0,  0J  =  0. 

Examples  XX.     1.    Prove  that 

(i)         a  [px,  &y] = /3  [ax,  ay]  =  [o/3.r,  a/3y], 

(ii)       (0  y]  +  [>',  y ])  +  [>",  y"]  =  O,  y ]  +  flV,  y']  +  [x",  y"]), 

(iii)      [x,  y]  +  [x1,  y']  =  [x',  y']  +  [x,  y], 

(iv)       (a  +  0)  [,v,  y]  =  a  [x,  y]  +/3  [x,  y], 

( v)      « {fa  y] + W,  y']\ = a  [v,  y]  +  a  [.>/,  /]. 

[We  have  already  proved  (iii).  The  remaining  equations  follow  with  equal 
ease  from  the  definitions.  The  reader  should  in  each  case  consider  the 
geometrical  significance  of  the  equation,  as  we  did  above  in  the  case  of  (hi).] 

2.  Ifl/is  the  middle  point  of  PQ,  then  0ll=\(JTP+0Q).  More  generally, 
if  M  divides  PQ  in  the  ratio  p  :  X,  then 

3.  If  G  is  the  centre  of  mass  of  equal  particles  at  Pu  P2,  ...,  Pn,  then 

OG  =  ( OPi  +  OP2  + . . .  +  0Pn)/n. 


74 


COMPLEX   NUMBERS 


[III 


4.  If  P,  Q,  R  are  collinear  points  in  the  plane,  then  it  is  possible  to  find 
real  numbers  a,  /3,  y,  not  all  zero,  and  such  that 

a.OP+^.OQ  +  y.OR=0; 
and  conversely.     [This  is  really  only  another  way  of  stating  Ex.  2.] 

5.  If  A~B  and  A~C  are  two  displacements  not  in  the  same  straight  line, 
and  

a  .  AB  +  p.  AC=y  .  AB  +  8  .  AC, 
then  a  =  y  and  0  =  8. 

[Take  ABl  =  a.  AB,  AC1  =  (B  .  AC.  Complete  the  parallelogram  AB^P^. 
Then  AP1  =  a  .  AB  +  0  .  AC.  It  is  evident  that  APX  can  only  be  expressed 
in  this  form  in  one  way,  whence  the  theorem  follows.] 

6.  A  BCD  is  a  parallelogram.      Through   Q,  a  point  inside  the  paral- 
lelogram, RQS  and  TQU  are  drawn 
parallel  to  the  sides.      Show  that 
RU,  TS  intersect  on  AC. 

[Let  the  ratios  AT:  AB,AR:AD 
be  denoted  by  a,  0.     Then 

AT=a.AB,    ~AR=p.AD, 

A~U=a.AB  +  AD,    AS=TB  +  p.~AD. 

Let  RU  meet  AC  in  P.     Then, 
since  R,  U,  P  are  collinear,  Fig-  21. 

AP  =  ^-AR+-^AU, 

A+fi.  A+/Z 

where  //./A  is  the  ratio  in  which  P  divides  RU.     That  is  to  say 


AP= 


a/i. 


AB+^±*AD. 


But  since  P  lies  on  AC,  AP  is  a  numerical  multiple  of  AC ;  say 
I7>=£ .  AC=  k  .  A~B+k .  AD. 
Hence  (Ex.  5)  ap  =  0\  +  /x  =  (A  +  /a)  k,  from  which  we  deduce 


a+0-1 

The  symmetry  of  this  result  shows  that  a  similar  argument  would  also  give 

n0 


AP  = 


AC, 


a+0-1 

if  P'  is  the  point  where  TS  meets  AC.     Hence  P  and  P'  are  the  same  point.] 

7.     A  BCD  is  a  parallelogram,  and  M  the  middle  point  of  AB.     Show  that 
D.M  trisects  and  is  trisected  by  A  C*. 

*  The  two  preceding  examples  are  taken  from  Willard  Gibbs'  Vector  Analysis. 


36,  37]  COMPLEX    NUMBERS  75 

37.  Multiplication  of  displacements.  So  far  we  have 
made  no  attempt  to  attach  any  meaning  whatever  to  the  notion 
of  the  product  of  two  displacements.  The  only  kind  of  multipli- 
cation which  we  have  considered  is  that  in  which  a  displacement 
is  multiplied  by  a  number.     The  expression 

[x,  y]  x  [x'}  2/'] 

so  far  means  nothing,  and  we  are  at  liberty  to  define  it  to  mean 
anything  we  like.  It  is,  however,  fairly  clear  that  if  any  definition 
of  such  a  product  is  to  be  of  any  use,  the  product  of  two  displace- 
ments must  itself  be  a  displacement. 

We  might,  for  example,  define  it  as  being  equal  to 

0  +  x,  y  +  y'] ; 

in  other  words,  we  might  agree  that  the  product  of  two  displace- 
ments was  to  be  always  equal  to  their  sum.  But  there  would  be 
two  serious  objections  to  such  a  definition.  In  the  first  place  our 
definition  would  be  futile.  We  should  only  be  introducing  a  new 
method  of  expressing  something  which  we  can  perfectly  well 
express  without  it.  In  the  second  place  our  definition  would  be 
inconvenient  and  misleading  for  the  following  reasons.  If  a  is 
a  real  number,  we  have  already  defined  a  [x,  y]  as  [ax,  ay].  Now, 
as  we  saw  in  §  34,  the  real  number  a  may  itself  from  one  point  of 
view  be  regarded  as  a  displacement,  viz.  the  displacement  [a] 
along  the  axis  OX,  or,  in  our  later  notation,  the  displacement 
[a,  0].  It  is  therefore,  if  not  absolutely  necessary,  at  any  rate 
most  desirable,  that  our  definition  should  be  such  that 

[a,  0]  [x,  y]  =  [ax,  ay], 

and  the  suggested  definition  does  not  give  this  result. 

A  more  reasonable  definition  might  appear  to  be 

[x,  y]  [x,  y']  =  [xx,  yy']. 
But  this  would  give 

[a,  0]  [x,  y]  =  [ax,  0] ; 

and  so  this  definition  also  would  be  open  to  the  second  objection. 

In  fact,  it  is  by  no  means  obvious  what  is  the  best  meaning 
to  attach  to  the  product  [x,  y~\  [x,  y'].  All  that  is  clear  is  (1)  that, 
if  our  definition  is  to  be  of  any  use,  this  product  must  itself  be 


76 


COMPLEX    NUMBERS 


[III 


a  displacement  whose  coordinates  depend  on  x  and  y,  or  in  other 
words  that  we  must  have 

[x,  y]  [x,  y']  =  [X,  Y], 

where  X  and   Y  are  functions  of  x,  y,  x,  and  y ';  (2)  that  the 
definition  must  be  such  as  to  agree  with  the  equation 

[x,  0]  [>',  y']  =  \xx,  xy'] ; 

and  (3)  that  the  definition  must  obey  the  ordinary  commutative, 
distributive,  and  associative  laws  of  multiplication,  so  that 

[x,  y]  [x\  y']  =  [>',  y']  [x,  y], 

([x,  y]  +  K  y'])  [x",  y"]  =  [x,  y]  [x'\  y"]  +  [x\  y]  [x",  f\ 

[x,  y]  ([x,  y']  +  [x",  y"])  =  [x,  y]  [x,  y']  +  [x,  y]  \x" ,  y"\ 

and        [x,  y]  {[x ,  y']  [x",  y"])  =  ([x,  y]  [x',  y'])  [x",  y"]. 

38.  The  right  definition  to  take  is  suggested  as  follows.  We 
know  that,  if  OAB,  OCD  are  two  similar  triangles,  the  angles 
corresponding  in  the  order  in  which  they  are  written,  then 

OB/OA  =  OD/OC, 

or  OB  .  OG  =  OA  .  OD.     This  suggests  that  we  should  try  to  define 
multiplication  and  division  of  displacements  in  such  a  way  that 

OBIOA  =  OB/W,    W.W=OA.OB. 

Now  let 

OB  =  [x,yl    OG  =  [x',y'],    OD  =  [X,  Y], 


37,  38]  COMPLEX    NUMBERS  77 

and  suppose  that  A  is  the  point  (1,  0),  so  that  OA  =  [1,  0].     Then 

OA.OD  =  [l,0][X,  F]  =  [Z,  F], 
and  so  [x,  y]  \x',  y']  =  [X,  F]. 

The  product  OB .  OG  is  therefore  to  be  defined  as  OB,  D  being 
obtained  by  constructing  on  OG  a  triangle  similar  to  OAB.  In 
order  to  free  this  definition  from  ambiguity,  it  should  be  observed 
that  on  OG  we  can  describe  two  such  triangles,  OGD  and  OGD', 
We  choose  that  for  which  the  angle  GOD  is  equal  to  A  OB  in  sign 
as  well  as  in  magnitude.  We  say  that  the  two  triangles  are  then 
similar  in  the  same  sense. 

If  the  polar  coordinates  of  B  and  G  are  (p,  6)  and  (cr,  <f>),  so 
that 

x  =  p  cos  0,     y  =  p  sin  0,    x'  =  a  cos  <p,     y'  =  a  sin  cf>, 

then  the  polar  coordinates  of  B  are  evidently  pa  and  0+cf>.   Hence 

X  =  pa  cos  (0  +  <f))  =  xx  —  yy ', 

Y=  pa  sin  (0  +  </>)  =  xy'  +  yx. 

The  required  definition  is  therefore 

\^     [x,  y]  [x,  y']  =  [ocx'  -  yy',  xy'  +  yx']  (6). 

We  observe  (1)  that  if  y  =  0,  then  X  =  xx,  Y  =  xy',  as  we 
desired ;  (2)  that  the  right-hand  side  is  not  altered  if  we  inter- 
change x  and  x,  and  y  and  y',  so  that 

[x,  y] [x,  y]  =  [x,  y] [x,  y] ; 
and  (3)  that 

{[x,  y]  +  [x'}  y']}  \x",  y"]  =  j>  +  x',y  +  y]  [x",  y"] 

=  [(x  +  x')  x"  ~(y  +  y')  y",  (x  +  x)  y"  +  (y  +  y')  x"] 

=  W  ~  yy",  ^j"  +  yx"]  +  OV'  -  y'y",  x'y"  +  y'x"] 

=  [x,y]W,y"]  +  [x',y'][x",y"l 

Similarly  we  can  verify  that  all  the  equations  at  the  end  of  §  37 
are  satisfied.  Thus  the  definition  (6)  fulfils  all  the  requirements 
which  we  made  of  it  in  §  37. 

Example.  Show  directly  from  the  geometrical  definition  given  above 
that  multiplication  of  displacements  obeys  the  commutative  and  distributive 
laws.  [Take  the  commutative  law  for  example.  The  product  OB  .  OC  is  OD 
(Fig.  22),  COD  being  similar  to  A  OD.     To  construct  the  product  ~0C .  OD  we 


78  COMPLEX   NUMBERS  [ill 

should  have  to  construct  on  OB  a  triangle  BOD{  similar  to  AOC ';  and  so  what 
we  want  to  prove  is  that  D  and  Dx  coincide,  or  that  BOD  is  similar  to  AOC. 
This  is  an  easy  piece  of  elementary  geometry.] 

39.  Complex  numbers.  Just  as  to  a  displacement  [x]  along 
OX  correspond  a  point  (x)  and  a  real  number  x,  so  to  a  displace- 
ment [x,  y]  in  the  plane  correspond  a  point  (x,  y)  and  a  pair 
of  real  numbers  x,  y. 

We  shall  find  it  convenient  to  denote  this  pair  of  real  numbers 

x,  y  by  the  symbol 

x  +  yi. 

The  reason  for  the  choice  of  this  notation  will  appear  later. 
For  the  present  the  reader  must  regard  x  +  yi  as  simply  another 
way  of  writing  [x,  y\.  The  expression  x  +  yi  is  called  a  complex 
number. 

We  proceed  next  to  define  equivalence,  addition,  and  multiplica- 
tion of  complex  numbers.  To  every  complex  number  corresponds 
a  displacement.  Two  complex  numbers  are  equivalent  if  the 
corresponding  displacements  are  equivalent.  The  sum  or  product 
of  two  complex  numbers  is  the  complex  number  which  corresponds 
to  the  sum  or  product  of  the  two  corresponding  displacements. 

Thus 

as  +  yi  =  x  +  yi (1 ), 

if  and  only  if  x  =  x',  y  =  y'', 

(x+yi)  +  (x+y'i)  =  (x  +  x')  +  (y  +  y')i (2); 

(x  +  yi)  (x  +  yi)  =  xx  —  yy'  +  (xyr  +  yx)  i (3). 

In  particular  we  have,  as  special  cases  of  (2)  and  (3), 

x  +  yi  =  (x  +  Oi)  +  (0  +  yi), 

(x  +  Oi)  {x  +  y'i)  =  xx  +  xy'i; 

and  these  equations  suggest  that  there  will  be  no  danger  of 
confusion  if,  when  dealing  with  complex  numbers,  we  write  x  for 
x  +  Oi  and  yi  for  0  +  yi,  as  we  shall  henceforth. 

Positive  integral  powers  and  polynomials  of  complex  numbers 
are  then  defined  as  in  ordinary  algebra.  Thus,  by  putting  x  =  x, 
y  —  y  in  (3),  we  obtain 

(x  +  yi)2  =  (x  +  yi)  (x  +  yi)  =  x2  —y2+  2xyi. 


38,  39]  COMPLEX    NUMBERS  79 

The  reader  will  easily  verify  for  himself  that  addition  and 
multiplication  of  complex  numbers  obey  the  laws  of  algebra 
expressed  by  the  equations 

m  +  yi  +  (x  +  y'i)  =  (x'  +  y'i)  +  (x  +  yi), 
{(x  +  yi)  +  (x  +  y'i)}  +  (x"  +  y'i)  =  (x  +  yi)  +  {(x  +  y'i)  +  (x"  +  y"i)}, 

O  +  yi)  (x'  +  y'i)  =  {x  +  y'i)  (x  +  yi), 
(x+yi)  {(x'+y'i)  +  (x"  +  y"i))  =  (x  +  yi)  (x'  +  y'i)  +  (x  +  yi)  (x"  +  y"i), 

{(oe+yi)+(tf+y'i)}  (*"+y"*)=(; + 2/0  O"  +  y"*)  +  0'  +  y'i)  O"  +  y'% 

(x  +  yi)  \(x'  +  y'i)  {x"  +  y"i))  =  {{x  +  yi)  (x'  +  y'i)}  (x"  +  y"i), 

the  proofs  of  these  equations  being  practically  the  same  as  those 
of  the  corresponding  equations  for  the  corresponding  displace- 
ments. 

Subtraction  and  division  of  complex  numbers  are  defined  as 
in  ordinary  algebra.     Thus  we  may  define  (x  +  yi)  —  (x'  +  y'i)  as 

(x  +  yi)  +  {- (x  +  y'i)}  =  x  +  yi  +  (- x  -  y'i)  =  (x-  x)  +  (y  -  y') i  ; 

or  again,  as  the  number  f  +  rji  such  that 

(x'  +  y'i)  +  (£  +  rji)  =  x  +  yi, 

which  leads  to  the  same  result.  And  (x  +  yi)l(x'j^j[i)  is  defined 
as  being  the  complex  number  £  +  rji  such  that 

(as'  +  y'i)  (f  +  rji)  =  x  +  yi, 

or  x'l;  —  y'rj  +  (x'rj  +  y'%)  i  =  x  +  yi, 

or  x'g  —  y'rj  =  x,    x'rj  +  y'i;  =  y    (4). 

Solving  these  equations  for  £  and  rj,  we  obtain 


xx'  +  yy'         _  yx'  —  xy 
x'2  +  y'2  '  x'2  +  y'2 


This  solution  fails  if  x  and  y  are  both  zero,  i.e.  if  x'  +  y'i  =  0. 
Thus  subtraction  is  always  possible;  division  is  always  possible 
unless  the  divisor  is  zero. 


80 


COMPLEX   NUMBERS 


[III 


Examples.     (1)    From  a  geometrical  point  of  view,  the  problem  of  the 
division  of  the  displacement  OB  by  OC  is  that  of  finding  D  so  that  the  triangles 
COB,  AOD  are  similar,  and  this  is 
evidently  possible  (and  the  solution 
unique)  unless  C  coincides  with  0,  or 
OC=0. 

(2)    The  numbers  %+yi,  x-yi  are 
said  to  be  conjugate.     Verify  that 

(x+yi)  (x  -  yi) = x2 +y2, 
so  that  the  product  of  two  conjugate 
numbers  is  real,  and  that 

x+yi  _  (x + yi)  (x'  —  y'i) 
x'  +y'i  ~  {x'  +y'i)  (x'  -  y'i) 

_  xx'  +yy'  +  (x'y  -  xy')  i 
~  x'2Ty''2      ''  ~~ ' 


Fig.  23. 


40.  One  most  important  property  of  real  numbers  is  that 
known  as  the  factor  theorem,  which  asserts  that  the  product  of  two 
numbers  cannot  be  zero  unless  one  of  the  two  is  itself  zero.  To 
prove  that  this  is  also  true  of  complex  numbers  we  put  x  =  0, 
y  =  0  in  the  equations  (4)  of  the  preceding  section.     Then 

x'%  —  y'n  =  0,     xt)  +  y'%  =  0. 

These  equations  give  £  =  0,  77  =  0,  i.e. 

%  +  vi  =  0, 

unless  x'  =  0  and  y  =  0,  or  x'  +  y'i  =  0.     Thus  x  +  yi  cannot  vanish 
unless  either  x  +  y'i  or  £  +  rji  vanishes. 

41.  The  equation  i2  =  —  1.  We  agreed  to  sinnilify  our 
notation  by  writing  x  instead  of  x  +  Oi  and  yi  instead  of  0  +  yi. 
The  particular  complex  number  \i  we  shall  denote  simply  by  i. 
It  is  the  number  which  corresponds  to  a  unit  displacement  along 
OY.    Also 

# = a = (o + ii)  (o  + 10 = (o .  0  - 1 . 1) + (o .  1  + 1 .  o)  %  =  - 1. 

Similarly  (—  if  =  —  1.     Thus    the    complex    numbers   i   and  —  i 
satisfy  the  equation  x2  =  —  1. 

The  reader  will  now  easily  satisfy  himself  that  the  upshot  of 
the  rules  for  addition  and  multiplication  of  complex  numbers  is 
this,  that  we  operate  ruith  complex  numbers  in  exactly  the  same 
way  as  with  real  numbers,  treating  the  symbol  i  as  itself  a  number, 


39-43]  COMPLEX   NUMBERS  81 

but  replacing  the  product  ii  =  i2  by  —  1  whenever  it  occurs.  Thus, 
for  example, 

(x  +  yi)  (x'  +  y'i)  =  aeaf  +  xy'i  +  yx'i  +  yy'i2 

=  (««'  -  yy)  +  (xy'  +  yd)  i 

42.  The  geometrical  interpretation  of  multiplication 
by  %.     Since 

(x  +  yi)  i  =  —  y  +  xi, 

it  follows  that  if  x  4-  yi  corresponds  to  OP,  and  OQ  is  drawn  equal 
to  OP  and  so  that  POQ  is  a  positive  right  angle,  then  (x  +  yi)  i 
corresponds  to  OQ.  In  other  words,  multiplication  of  a  complex 
number  by  i  turns  the  corresponding  displacement  through  a  right 
angle. 

We  might  have  developed  the  whole  theory  of  complex 
numbers  from  this  point  of  view.  Starting  with  the  ideas  of 
x  as  representing  a  displacement  along  OX,  and  of  i  as  a  symbol 
of  operation  equivalent  to  turning  x  through  a  right  angle,  we 
should  have  been  led  to  regard  yi  as  a  displacement  of  magnitude 
y  along  OY.  It  would  then  have  been  natural  to  define  x  +  yi  as 
m  §§  37  and  40,  and  (x  +  yi)  i  would  have  represented  the  dis- 
placement obtained  by  turning  x  +  yi  through  a  right  angle, 
i.e.  —y  +  xi.  Finally,  we  should  naturally  have  defined  (x  +  yi)  x 
as  xx'  +  yx'i,  (x  +  yi)  y'i  as  —  yy'  +  xy'i,  and  (x  +  yi)  (x'  +  y'i)  as  the 
sum  of  these  displacements,  i.e.  as 

xx'  -  yy'  +  (xy'  +  yx)  i. 

43.  The  equations  z2  + 1  =  0,  az*  +  2bz  +  c  =  0.  There  is  no 
real  number  z  such  that  z2  +  1  =  0 ;  this  is  expressed  by  saying 
that  the  equation  has  no  real  roots.  But,  as  we  have  just  seen, 
the  two  complex  numbers  i  and  —  i  satisfy  this  equation.  We 
express  this  by  saying  that  the  equation  has  the  two  complex  roots 
i  and  —  i.  Since  i  satisfies  z2  =  —  1,  it  is  sometimes  written  in  the 
form  «/(—  1). 

Complex  numbers  are  sometimes  called  imaginary*.  The 
expression  is  by  no  means  a  happily  chosen  one,  but  it  is  firmly 

*  The  phrase  'real  number'  was  introduced  as  au  antithesis  to  'imaginary 
number '. 

H.  G 


82  COMPLEX   NUMBERS  [ill 

established  and  has  to  be  accepted.  It  cannot,  however,  be  too 
strongly  impressed  upon  the  reader  that  an  'imaginary  number' 
is  no  more  '  imaginary ',  in  any  ordinary  sense  of  the  word,  than  a 
'  real '  number ;  and  that  it  is  not  a  number  at  all,  in  the  sense  in 
which  the  'real'  numbers  are  numbers,  but,  as  should  be  clear  from 
the  preceding  discussion,  a  pair  of  numbers  (x}y),  united  symbolically, 
for  purposes  of  technical  convenience,  in  the  form  x  +  yi,_  Such 
a  pair  of  numbers  is  no  less  'real'  than  any  ordinary  number 
such  as  {,  or  than  the  paper  on  which  this  is  printed,  or  than 
the  Solar  System.     Thus 

i  =  0  +  li 

stands  for  the  pair  of  numbers  (0,  1),  and  may  be  represented 
geometrically  by  a  point  or  by  the  displacement  [0,  1].  And 
when  we  say  that  i  is  a  root  of  the  equation  z2  +  1  =  0,  what  we 
mean  is  simply  that  we  have  defined  a  method  of  combining  such 
pairs  of  numbers  (or  displacements)  which  we  call  '  multiplica- 
tion', and  which,  when  we  so  combine  (0,  1)  with  itself,  gives  the 
result  (- 1,  0). 

Now  let  us  consider  the  more  general  equation 
az"  +  2bz  +  c  =  Q, 

where  a,  b,  c  are  real  numbers.  If  b2  >  ac,  the  ordinary  method  of 
solution  gives  two  real  roots 

{-  b  ±  V(69  -  ac)}fa. 
If  b2  <  ac,  the  equation  has  no  real  roots.     It  may  be  written  in 

the  form 

{z  +  (b/a)}2  =  -(ac-b2)la^, 

an  equation  which  is  evidently  satisfied  if  we  substitute  for 
z  +  (b/a)  either  of  the  complex  numbers  ±%  *J(ac  —  b2)/a*.  We 
express  this  by  saying  that  the  equation  has  the  two  complex  roots 
{-  b  ±  i  s/(ac  -  b2)}/a, 
If  we  agree  as  a  matter  of  convention  to  say  that  when  b2  =  ac 
(in  which  case  the  equation  is  satisfied  by  one  value  of  x  only, 
viz.  —  b/a),  the  equation  has  two  equal  roots,  we  can  say  that 
a  quadratic  equation  with  real  coefficients  has  two  roots  in  all 
cases,  either  two  distinct  real  roots,  or  two  equal  real  roots,  or  two 
distinct  complex  roots. 

*  We  shall  sometimes  write  x  +  iy  instead  of  x  +  yi  for  convenience  in  printing. 


43]  COMPLEX   NUMBERS  83 

The  question  is  naturally  suggested  whether  a  quadratic 
equation  may  not,  when  complex  roots  are  once  admitted,  have 
more  than  two  roots.  It  is  easy  to  see  that  this  is  not  possible. 
Its  impossibility  may  in  fact  be  proved  by  precisely  the  same 
chain  of  reasoning  as  is  used  in  elementary  algebra  to  prove  that 
an  equation  of  the  ?ith  degree  cannot  have  more  than  n  real 
roots.  Let  us  denote  the  complex  number  x  +  yi  by  the  single 
letter  z,  a  convention  which  we  may  express  by  writing 
z  =  x  +  yi.  Let  /  (z)  denote  any  polynomial  in  z,  with  real  or 
complex  coefficients.     Then  we  prove  in  succession : 

(1)  that  the  remainder,  when  f(z)  is  divided  by  z  —  a,  a  being 
any  real  or  complex  number,  is  /(a) ; 

(2)  that  if  a  is  a  root  of  the  equation  f(z)  =  0,  then  f{z)  is 
divisible  by  z  —  a ; 

(3)  that  if  f(z)  is  of  the  ?ith  degree,  and  f(z)  =  0  has  the 
n  roots  a1}  a2, ...,  an,  then 

f{z)  =  A  (z  -  rtj)  (z  -  o2) . . .  (z  -  an), 

where  A  is  a  constant,  real  or  complex,  in  fact  the  coefficient 
of  zn  in  f(z).  From  the  last  result,  and  the  theorem  of  §  40, 
it  follows  that  f(z)  cannot  have  more  than  n  roots. 

We  conclude  that  a  quadratic  equation  with  real  coefficients  has 
exactly  two  roots.  We  shall  see  later  on  that  a  similar  theorem  is 
true  for  an  equation  of  any  degree  and  with  either  real  or  complex 
coefficients :  an  equation  of  the  nth  degree  has  exactly  n  roots. 
The  only  point  in  the  proof  which  presents  any  difficulty  is  the 
first,  viz.  the  proof  that  any  equation  must  have  at  least  one 
root.  This  we  must  postpone  for  the  present*.  We  may,  however, 
at  once  call  attention  to  one  very  interesting  result  of  this  theorem. 
In  the  theory  of  number  we  start  from  the  positive  integers  and 
from  the  ideas  of  addition  and  multiplication  and  the  converse 
operations  of  subtraction  and  division.  We  find  that  these 
operations  are  not  always  possible  unless  we  admit  new  kinds  of 
numbers.  We  can  only  attach  a  meaning  to  3  —  7  if  we  admit 
negative  numbers,  or  to  f  if  we  admit  rational  fractions.  When 
we  extend  our  list  of  arithmetical  operations  so  as  to  include  root 
extraction  and  the  solution  of  equations,  we  find  that  some  of 

*  See  Appendix  I. 

6-2 


84  COMPLEX   NUMBERS  [ill 

them,  such  as  that  of  the  extraction  of  the  square  root  of  a  number 
which  (like  2)  is  not  a  perfect  square,  are  not  possible  unless  we 
widen  our  conception  of  a  number,  and  admit  the  irrational 
numbers  of  Chap.  I. 

Others,  such  as  the  extraction  of  the  square  root  of  —  1,  are 
not  possible  unless  we  go  still  further,  and  admit  the  complex 
numbers  of  this  chapter.  And  it  would  not  be  unnatural  to 
suppose  that,  when  we  come  to  consider  equations  of  higher 
degree,  some  might  prove  to  be  insoluble  even  by  the  aid  of 
complex  numbers,  and  that  thus  we  might  be  led  to  the  con- 
siderations of  higher  and  higher  types  of,  so  to  say,  hyper-complex 
numbers.  The  fact  that  the  roots  of  any  algebraical  equation 
whatever  are  ordinary  complex  numbers  shows  that  this  is  not  the 
case.  The  application  of  any  of  the  ordinary  algebraical  operations 
to  complex  numbers  will  yield  only  complex  numbers.  In  technical 
language  '  the  field  of  the  complex  numbers  is  closed  for  algebraical 
operations '. 

Before  we  pass  on  to  other  matters,  let  us  add  that  all 
theorems  of  elementary  algebra  which  are  proved  merely  by 
the  application  of  the  rules  of  addition  and  multiplication  are 
true  whether  the  numbers  which  occur  in  them  are  real  or  com- 
plex, since  the  rules  referred  to  apply  to  complex  as  well  as 
real  numbers.  For  example,  we  know  that,  if  a  and  j3  are  the 
roots  of 

az2  +  2bz  +  c  =  0, 

then  a  +  /3  =  -  (26/a),      a£  =  (c/a). 

Similarly,  if  a,  /3,  7  are  the  roots  of 

az3  +  Sbz2  +  3cz  +  d  =  0, 
then 

a  +  /3  +  7  =  -  (36/a),     £7  +  7a  +  a/3  =  (3c/a),     aj3y  =  -(d/a). 

All  such  theorems  as  these  are  true  whether  a,  b,  ...  a,  fi,  ...  are 
real  or  complex. 

44.  Argand's  diagram.  Let  P  (Fig.  24)  be  the  point  (x,  y\ 
r  the  length  OP,  and  6  the  angle  XOP,  so  that 

x  =  r  cos  6,  y  =  r  sin  6,  r  =  *J(x2  +  y2),  cos  6  :  sin  6  : 1 : :  x  :  y  :  r. 


43,  44] 


COMPLEX   NUMBERS 


85 


We  denote  the  complex  number  x  +  yi  by  z,  as  in  §  43,  and 
we  call  z  the  complex  variable. 
We  call  P  the  point  z,  or 
the  jpoint  corresponding  to  z\ 
z  the  argument  of  P,  x  the 
real  part,  y  the  imaginary 
part,  r  the  modulus,  and 
6  the  amplitude  of  z ;  and  we 
write 


x  =  B.{z),     y=I(z), 
6  =  am  z. 


r=  \z 


Fig.  24. 


When  y  =  0  we  say  that  2  is  ?'m£,  when  x  =  0  that  2  is  purely 
imaginary.  Two  numbers  #  +  yt,  #  —  yi  which  differ  only  in 
the  signs  of  their  imaginary  parts,  we  call  conjugate.  It  will  be 
observed  that  the  sum  2x  of  two  conjugate  numbers  and  their 
product  #2  +  2/2  are  both  real,  that  they  have  the  same  modulus 
V(#2  +  y2)  and  that  their  product  is  equal  to  the  square  of  the 
modulus  of  either.  The  roots  of  a  quadratic  with  real  coefficients, 
for  example,  are  conjugate,  when  not  real. 

It  must  be  observed  that  6  or  am  z  is  a  many- valued  function  of 
x  and  y,  having  an  infinity  of  values,  which  are  angles  differing  by 
multiples  of  2tt*.  A  line  originally  lying  along  OX  will,  if  turned 
through  any  of  these  angles,  come  to  lie  along  OP.  We  shall 
describe  that  one  of  these  angles  which  lies  between  —  it  and 
it  as  the  principal  value  of  the  amplitude  of  z.  This  de- 
finition is  unambiguous  except  when  one  of  the  values  is  it, 
in  which  case  —  it  is  also  a  value.  In  this  case  we  must  make 
some  special  provision  as  to  which  value  is  to  be  regarded  as 
the  principal  value.  In  general,  when  we  speak  of  the  amplitude 
of  z  we  shall,  unless  the  contrary  is  stated,  mean  the  principal 
value  of  the  amplitude. 

Fig.  24  is  usually  known  as  Argand's  diagram. 

*  It  is  evident  that  |  z  |  is  identical  with  the  polar  coordinate  r  of  P,  and  that 
the  other  polar  coordinate  0  is  one  value  of  am  z.  This  value  is  not  necessarily 
the  principal  value,  as  defined  below,  for  the  polar  coordinate  of  §  22  lies  between 
0  and  2tt,  and  the  principal  value  between  -  w  and  ir. 


86  COMPLEX    NUMBERS  [ill 

45.  De  Moivre's  Theorem.  The  following  statements 
follow  immediately  from  the  definitions  of  addition  and  multi- 
plication. 

(1)  The  real  (or  imaginary)  part  of  the  sum  of  two  conyplex 
numbers  is  equal  to  the  sum  of  their  real  (or  imaginary)  parts. 

(2)  The  modulus  of  the  product  of  two  complex  numbers  is 
equal  to  the  product  of  their  moduli. 

(3)  The  amplitude  of  the  product  of  two  complex  numbers  is 
either  equal  to  the  sum  of  their  amplitudes,  or  differs  from  it  by  2ir. 

It  should  be  observed  that  it  is  not  always  true  that  the  principal  value  of 
am  (zz1)  is  the  sum  of  the  principal  values  of  am  z  and  am  z'.  For  example,  if 
z=z'=  —  l+i,  then  the  principal  values  of  the  amplitudes  of  z  and  z'  are  each 
fn-.     But  zz'  =  —  2i,  and  the  principal  value  of  am  (zz')  is  —\n  and  not  fn-. 

The  two  last  theorems  may  be  expressed  in  the  equation 
r  (cos  0  +  i  sin  0)  x  p  (cos  cp  +  i  sin  cp) 

=  rp  {cos  (0  +  cp)  +  i  sin  (6  +  <f>)}, 
which  may  be  proved  at  once  by  multiplying  out  and  using  the 
ordinary  trigonometrical  formulae  for  cos  (0  +  cp)  and  sin  (0  +  cp). 
More  generally 

i\  (cos  dx  +  i  sin  0t)  x  r2  (cos  02  +  i  sin  0.2)  x  ...  x  rn  (cos  0n  +  i  sin  6n) 
=  nr2  . . .  rn  {cos  (d1  +  02  +  . . .  +  6n)  +  i  sin  (01  +  02+...+  0n)}. 
A  particularly  interesting  case  is  that  in  which 

n  =  r2  =  . . .  =  rn  =  1,     01  =  02=...  =  0n=  0 
We  then  obtain  the  equation 

(cos  0  +  i  sin  0)n  =  cos  n0  +  i  sin  n0, 
where  n  is  any  positive  integer:  a  result  known  as  De  Moivre's 
Theorem*. 

Again,  if  z  =  r  (cos  0  +  i  sin  0) 

then  \\z  =  (cos  0  -  i  sin  0)/r. 

Thus  the  modulus  of  the  reciprocal  of  z  is  the  reciprocal  of  the 
modulus  of  z,  and  the  amplitude  of  the  reciprocal  is  the  negative  of 
the  amplitude  of  z.  We  can  now  state  the  theorems  for  quotients 
which  correspond  to  (2)  and  (3). 

*  It  will  sometimes  be  convenient,  for  the  sake  of  brevity,  to  denote  cos  d  +  i  sin  d 
by  Cis0:  in  this  notation,  suggested  by  Profs.  Harkness  and  Morley,  De  Moivre's 
theorem  is  expressed  by  the  equation  (Cis  0)"  =  Cisn0. 


45] 


COMPLEX    NUMBERS 


87 


(4)  The  modulus  of  the  quotient  of  two  complex  numbers  is 
equal  to  the  quotient  of  their  moduli. 

(5)  The  amplitude  of  the  quotient  of  two  complex  numbers 
either  is  equal  to  the  difference  of  their  amplitudes,  or  differs  from 
it  by  27r. 

Again        (cos  9  +  i  sin  6)~n  =  (cos  6  —  i  sin  6)n 

-  {cos  (-  0)  +  i  sin  (-  6))n 

=  cos  (—  nO)  +  i  sin  (—  nd). 

Hence  Be  Moivres  Theorem  holds  for  all  integral  values  of  n, 

positive  or  negative. 

To  the  theorems  (1) — (5)  we  may  add  the  following  theorem, 
which  is  also  of  very  great  importance. 

(6)  The  modulus  of  the  sum  of  any  number  of  complex 
numbers  is  not  greater  than  the  sum  of  their  moduli. 


Let  OP,  OP',  ...  be  the  displacements  corresponding  to  the 
various  complex  numbers.  Draw  PQ  equal  and  parallel  to  OP', 
QR  equal  and  parallel  to  OP",  and  so  on.  Finally  we  reach  a 
point  U,  such  that 

0U=OP  +  0P'  +  0P7,+  .... 
The    length    OU  is   the    modulus   of  the    sum    of  the    complex 
numbers,  whereas  the  sum  of  their  moduli   is  the   total   length 
of  the  broken  line  OPQR...U,  which  is  not  less  than  OU. 

A  purely  arithmetical  proof  of  this  theorem  is  outlined  in 
Exs.  xxi.  1. 


88  COMPLEX   NUMBERS  [ill 

46.  We  add  some  theorems  concerning  rational  functions  of 
complex  numbers.  A  rational  function  of  the  complex  variable  z 
is  defined  exactly  as  is  a  rational  function  of  a  real  variable  x, 
viz.  as  the  quotient  of  two  polynomials  in  z. 

Theorem  1.  Any  rational  function  R  (z)  can  be  reduced  to 
the  form  X  +  Yi,  where  X  and  Y  are  rational  functions  of  x  and 
y  with  real  coefficients. 

In  the  first  place  it  is  evident  that  any  polynomial  P  (x  +  yi) 
can  be  reduced,  in  virtue  of  the  definitions  of  addition  and  multi- 
plication, to  the  form  A  +  Bi,  where  A  and  B  are  polynomials 
in  x  and  y  with  real  coefficients.  Similarly  Q  (x  +  yi)  can  be 
reduced  to  the  form  G  +  Di.     Hence 

R  (x  +  yi)  =  P(x  +  yi)/Q  (x  +  yi) 

can  be  expressed  in  the  form 

(A  +  Bi)J(G  +  Di)  =  (A  +  Bi)  (G  -  Di)/(C  +  Di)  (G  -  Di) 

AG  +  BD     BG-AD . 
~   6'2  +  X>2   +  C2  +  £2  *' 

which  proves  the  theorem. 

Theorem  2.  If  R(x  +  yi)  =  X+  Yi,  R  denoting  a  rational 
function  as  before,  but  with  real  coefficients,  then  R(x—yi)=X—Yi. 

In  the  first  place  this  is  easily  verified  for  a  power  (x  +  yi)n 
by  actual  expansion.  It  follows  by  addition  that  the  theorem  is 
true  for  any  polynomial  with  real  coefficients.  Hence,  in  the 
notation  used  above, 

A_-Bi_AG  +  BD     BG-AD  . 
M(x     y%)—(}-M~   C2  +  i)2"      C2  +  Z>2   *' 

the  reduction  being  the  same  as  before  except  that  the  sign  of  i 
is  changed  throughout.  It  is  evident  that  results  similar  to  those 
of  Theorems  1  and  2  hold  for  functions  of  any  number  of  complex 
variables. 

Theorem  3.     The  roots  of  an  equation 

a0zn+a1zn~1+  ...+an  =  0, 

whose  coefficients  are  real,  may,  in  so  far  as  they  are  not  themselves 
real,  be  arranged  in  conjugate  pairs. 


46]  COMPLEX   NUMBERS  89 

For  it  follows  from  Theorem  2  that  if  x  +  yi  is  a  root  then  so  is 
x  —  yi.  A  particular  case  of  this  theorem  is  the  result  (§  43)  that 
the  roots  of  a  quadratic  equation  with  real  coefficients  are  either 
real  or  conjugate. 

This  theorem  is  sometimes  stated  as  follows :  in  an  equation 
with  real  coefficients  complex  roots  occur  in  conjugate  pairs.  It 
should  be  compared  Avith  the  result  of  Exs.  vm.  7,  which  may  be 
stated  as  follows :  in  an  equation  with  rational  coefficients  irrational 
roots  occur  in  conjugate  pairs*. 

Examples  XXI.  1.  Prove  theorem  (6)  of  §  45  directly  from  the 
definitions  and  without  the  aid  of  geometrical  considerations. 

[First,  to  prove  that  |  z  +  z'  \  £  |  z  \  +  \  z'  |  is  to  prove  that 
(x + x'f  +  (y  +yj-  <  y (.r2  +  */2)  +  J(af*  +/2)}2. 
The  theorem  is  then  easily  extended  to  the  general  case.] 

2.  The  one  and  only  case  in  which 

\z\  +  \z'\+...  =  \z  +  z'+...\, 
is  that  in  which  the  numbers  z,  z',  ...  have  all  the  same  amplitude.     Prove 
this  both  geometrically  and  analytically. 

3.  The  modulus  of  the  sum  of  any  number  of  complex  numbers  is  not 
less  than  the  sum  of  their  real  (or  imaginary)  parts. 

4.  If  the  sum  and  product  of  two  complex  numbers  are  both  real,  then 
the  two  numbers  must  either  be  real  or  conjugate. 

5.  If  a  +  bj2  +  (c  +  dAj2)i=A  +  Bs/2  +  (C+Ds/2)i, 
where  a,  b,  c,  d,  A,  B,  C,  D  are  real  rational  numbers,  then 

a=A,    b  =  B,    c=C,    d=D. 

6.  Express  the  following  numbers  in  the  form  A  +  Bi,  where  A  and  B  are 
real  numbers : 

euw    (£)•    £*)■    &&    «-» 

where  X  and  p  are  real  numbers. 

7.  Express  the  following  functions  of  z=x+yi  in  the  form  X +  Yi,  where 
Xand  Fare  real  functions  of  x  and  y:  z2,  z3,  zn,  \}z}  a-f  (l/«),  (a  +  /32)/(y+8?), 
where  a,  /3,  y,  8  are  real  numbers. 

8.  Find  the  moduli  of  the  numbers  and  functions  in  the  two  preceding 
examples. 

*  The  numbers  a  +  ^Jb,  a  -  sjb,  where  a,  b  are  rational,  are  sometimes  said  to  be 
'  conjugate '. 


90  COMPLEX   NUMBERS  [ill 

9.  The  two  lines  joining  the  points  z  =  a,  z  =  b  and  z  =  c,  z=d  will  be 
perpendicular  if 

am(£r!),*±**i 

i.e.  if  (a  —  b)l(c-d)  is  purely  imaginary.  What  is  the  condition  that  the  lines 
should  be  parallel  ? 

10.  The  three  angular  points  of  a  triangle  are  given  by  z  =  a,  z=&,  z  =  y, 
where  a,  /3,  y  are  complex  numbers.     Establish  the  following  propositions : 

(i)      the  centre  of  gravity  is  given  by  z  =  \  (a  +  fi  +  y)  ; 
(ii)     the  circum-centre  is  given  by  \z—a\  =  \z-f&\  =  \z-  y\  ; 
(iii)    the  three  perpendiculars  from  the  angular  points  on  the  opposite 
sides  meet  in  a  point  given  by 

(iv)     there  is  a  point  P  inside  the  triangle  such  that 
CBP=ACP=BAP=a>, 
and  cot  &>  =  cot  A  +  cot  B  +  cot  G. 

[To  prove  (iii)  we  observe  that  if  A,  B,  G  are  the  vertices,  and  P  any 
point  z,  then  the  condition  that  AP  should  be  perpendicular  to  BG  is  (Ex.  9) 
that  (z  —  a)/(/3  -  y)  should  be  purely  imaginary,  or  that 

R(s-a)R(/3-y)+I(«-a)I(0-y)=O. 

This  equation,  and  the  two  similar  equations  obtained  by  permuting  a,  /3,  y 
cyclically,  are  satisfied  by  the  same  value  of  z,  as  appears  from  the  fact  that 
the  sum  of  the  three  left-hand  sides  is  zero. 

To  prove  (iv),  take  BG  parallel  to  the  positive  direction  of  the  axis  of  x. 
Then* 

y-P=a,     a-y=-bC\s(-G),     0-a= -cCisZ?. 

We  have  to  determine  z  and  w  from  the  equations 

(3-a)(fto-ao)^(z-ff)(yo-0o)^(z-y)(ao-yo)  =  Cia2tj 
(20-a0)(/3-a)      (so-/30)(y-i8)      (*0"yo)(a-y) 
where  z0,  a0,  /30,  y0  denote  the  conjugates  of  z,  a,  /3,  y. 

Adding  the  numerators  and  denominators  of  the  three  equal  fractions, 
and  using  the  equation 

i  cot  co  =  (1  +  Cis  2co)/(l  -  Cis  2w), 
we  find  that 

icot(0=  03-y)(/3o-yo)  +  (y-q)(yo-«o)  +  (a-/3)(qo-/3o\ 
/3y0  -  A>y + y«o  -  yo«  +  a£o  -  ao0 

From  this  it  is  easily  deduced  that  the  value  of  cot  co  is  (a2+62  +  c2)/4A, 
where  A  is  the  area  of  the  triangle ;  and  this  is  equivalent  to  the  result  given. 

*  We  suppose  that  as  we  go  round  the  triangle  in  the  direction  ABC  we  leave 
it  on  our  left. 


46]  COMPLEX   NUMBERS  91 

To  determine  z,  we  multiply  the  numerators  and  denominators  of  the 
equal  fractions  by  (y0-/30)/(/3-a),  (a0-y0)/(y-P),  (/30 -  a0)/(a  -  y),  and  add 
to  form  a  new  fraction.     It  will  be  found  that 

aaCis^4-6j3Cis5  +  cvCisC  , 

z = ' . 

aCis  A  +  b  Oisi>'  +  cCis  C 

11.  The  two  triangles  whose  vertices  are  the  points  a,  b,  c  and  x,  y,  z 
respectively  will  he  similar  if 

1     1     1    =0 

a     b     c 
x    y     z 

[The  condition  required  is  that  ABIAC=XY/XZ  (large  letters  denoting 
the  points  whose  arguments  are  the  corresponding  small  letters),  or 
(b-a)l(c-a)=-(y -x)l{z-x),  which  is  the  same  as  the  condition  given.] 

12.  Deduce  from  the  last  example  that  if  the  points  x,  y,  z  are  collinear 
then  we  can  find  real  numbers  a,  /3,  y  such  that  ct-f /3+y  =  0  and  ax+(3y+yz=0, 
and  conversely  (cf.  Exs.  XX.  4).  [Use  the  fact  that  in  this  case  the  triangle 
formed  by  x,  y,  z  is  similar  to  a  certain  line-triangle  on  the  axis  OX,  and 
apply  the  result  of  the  last  example.] 

13.  The  general  linear  equation  with  complex  coefficients.     The 

equation  az  +  j3  —  0  has  the  one  solution  z  =  -  (/3/a),  unless  a  =  0.     If  we  put 

a  =  a+Ai,     (3  =  b-\-Bi,     z=x+yi, 

and  equate  real  and  imaginary  parts,  we  obtain  two  equations  to  determine 
the  two  real  numbers  x  and  y.  The  equation  will  have  a  real  root  if  y  =  0, 
which  gives  ax  +  b  =  0,  Ax  +  B  =  0,  and  the  condition  that  these  equations 
should  be  consistent  is  aB  —  bA  =  0. 

14.  The  general  quadratic  equation  with  complex  coefficients.  This 
equation  is 

(a  +  Ai)z*+2(b  +  Bi)z+(c  +  Ci)  =  0. 

Unless  a  and  A  are  both  zero  we  can  divide  through  by  a  +  iA.  Hence 
we  may  consider 

z2  +  2(b  +  Bi)z  +  (c  +  Ci)  =  0    (1) 

as  the  standard  form  of  our  equation.  Putting  z  =  x+yi  and  equating  real 
and  imaginary  parts,  we  obtain  a  pair  of  simultaneous  equations  for  x  and  y, 
viz. 

xi  -y2  +  2(bx-By)  +  c  =  0      2xy  +  2  (by+Bx)  +  C=  0. 

If  we  put 

x  +  b  =  £,    y  +  B=rj,     b2-B2-c  =  h,     2bB-C=Jc, 

these  equations  become  £2  —  r]2=k,     2^rj=k. 


92  COMPLEX   NUMBERS  [ill 

Squaring  and  adding  we  obtain 

We  must  choose  the  signs  so  that  £77  has  the  sign  of  k  :  i.e.  if  k  is  positive 
we  must  take  like  signs,  if  k  is  negative  unlike  signs. 

Conditions  for  equal  roots.  The  two  roots  can  only  be  equal  if  both  the 
square  roots  above  vanish,  i.e.  if  A=0,  £=0,  or  if  c  =  b2-B2,  C=2bB.  These 
conditions  are  equivalent  to  the  single  condition  c  +  Ci=(b  +  Bi)2,  which 
obviously  expresses  the  fact  that  the  left-hand  side  of  (1)  is  a  perfect  square. 

Condition  for  a  real  root.  If  x2  +  2  (b  +  Bi)  x  +  (c  +  Ci)=0,  where  x  is 
real,   then   x2  +  2bx-\-c=0,    2Bx-\-C=0.      Eliminating  x  we  find   that  the 

required  condition  is 

C2-4bBC+lcB2  =  Q. 

Condition  for  a  purely  imaginary  root.     This  is  easily  found  to  be 
C2-4bBC-4b2c  =  0. 

Conditions  for  a  pair  of  conjugate  complex  roots.  Since  the  sum  and  the 
product  of  two  conjugate  complex  numbers  are  both  real,  b  +  Bi  and  c  +  Ci 
must  both  be  real,  i.e.  B  =  0,  C=0.  Thus  the  equation  (1)  can  have  a  pair  of 
conjugate  complex  roots  only  if  its  coefficients  are  real.  The  reader  should 
verify  this  conclusion  by  means  of  the  explicit  expressions  of  the  roots. 
Moreover,  if  62Sc,  the  roots  will  be  real  even  in  this  case.  Hence  for  a  pair 
of  conjugate  roots  we  must  have  B=0,  C=0,  b2<c. 

15.     The  Cubic  equation.     Consider  the  cubic  equation 
z3  +  ZHz  +  G=0, 
where  G  and  H  are  complex  numbers,  it  being  given  that  the  equation  has 
(a)  a  real  root,  (6)  a  purely  imaginary  root,  (c)  a  pair  of  conjugate  roots      If 
H='K+fxi,  C  =  p  +  a-i,  we  arrive  at  the  following  conclusions. 

(a)  Conditions  for  a  real  root.  If  p.  is  not  zero,  then  the  i*eal  root  is  -  <r/3p, 
and  o-3  +  27X/xV  —  27p3p  =  0.  On  the  other  hand,  if  p  =  0  then  we  must  also 
have  o-  =  0,  so  that  the  coefficients  of  the  equation  are  real.  In  this  case  there 
may  be  three  real  roots. 

(b)  Conditions  for  a  purely  imaginary  root.  If  p  is  not  zero  then  the  purely 
imaginary  root  is  (p/3p)  i,  and  p3-27Ap2p  —  27p3cr=0.  If  p  =  0  then  alsop  =  0, 
and  the  root  is  yi,  where  y  is  given  by  the  equation  y3  —  3Xy  -  o-  =  0,  which  has 
real  coefficients.     In  this  case  there  may  be  three  purely  imaginary  roots. 

(c)  Conditions  for  a  pair  of  conjugate  complex  roots.  Let  these  be  x+yi 
and  x—yi.  Then  since  the  sum  of  the  three  roots  is  zero  the  third  root 
must  be  —  2x.  From  the  relations  between  the  coefficients  and  the  roots  of 
an  equation  we  deduce 

y*  _  Zx2  =  3ff,     2x  (x2+y2)  =  G. 

Hence  G  and  H  must  both  be  real. 

In  each  case  we  can  either  find  a  root  (in  which  case  the  equation  can 
be  reduced  to  a  quadratic  by  dividing  by  a  known  factor)  or  we  can  reduce 
the  solution  of  the  equation  to  the  solution  of  a  cubic  equation  with  real 
coefficients. 


46] 


COMPLEX   NUMBERS 


93 


16.  The  cubic  equation  xz  +  ahv2  +  a2x  +  a3  =  0,  where  a1  =  A1  +  A  /?',  . . . ,  has 
a  pair  of  conjugate  complex  roots.  Prove  that  the  remaining  root  is 
—  Ai'a3jA3',  unless  A3'  =  0.     Examine  the  case  in  which  A3'  =  0. 

1 7.  Prove  that  if  z3  +  ZHz  +  G=0  has  two  complex  roots  then  the  equation 

8a3  +  6aH-G=0 
has  one  real  root  which  is  the  real  part  a  of  the  complex  roots  of  the 
original  equation  ;  and  show  that  a  has  the  same  sign  as  G. 

18.  An  equation  of  any  order  with  complex  coefficients  will  in  general 
have  no  real  roots  nor  pairs  of  conjugate  complex  roots.  How  many  con- 
ditions must  be  satisfied  by  the  coefficients  in  order  that  the  equation  should 
have  (a)  a  real  root,  (6)  a  pair  of  conjugate  roots  ? 

19.  Coaxal  circles.     In  Fig.  26,  let  a,  b,  z  be  the  arguments  of  A,  B,  P 

z-b 


Then 


—  =  APB, 

z-a 


if  the  principal  value  of  the  amplitude  is  chosen.  If  the  two  circles  shown 
in  the  figure  are  equal,  and  z',  zu  z{  are  the  arguments  of  P\  Pu  P±y 
and  A  PB  —  6,  it  is  easy  to  see  that 


z'-b 

am   ■; =TT-i 

z  —a 


h-b 


and 


z{-  b 
am  —, =  -  it  + 1 


z{  -  a 
The  locus  defined  by  the  equation 

am =  6, 

z-a 

where  &  is  constant,  is  the  arc  APB.  By 
writing  it  —  6,  -0,  —  n  +  6  for  $,  we  obtain 
the  other  three  arcs  shown. 

The  system  of  equations  obtained  by 
supposing  that  6  is  a  parameter,  varying 
from  —  it  to  +7r,  represents  the  system  of 
circles  which  can  be  drawn  through  the 
points  A,  B.  .  It  should  however  be  ob- 
served that  each  circle  has  to  be  divided 
into  two  parts  to  which  correspond  different 
values  of  8. 

20.     Now  let  us  consider  the  equation 

\z-a\~ 
where  \  is  a  constant. 

Let  A' be  the  point  in  which  the  tangent  to  the  circle  ABP  at  P  meets 
AB.     Then  the  triangles  KPA,  EBP  are  similar,  and  so 
AP/PB = PK\BK=  KA/KP  =  X. 


Fig.  26. 


•(1), 


94)  COMPLEX   NUMBERS  [ill 

Hence  KAjKB  =  \\  and  therefore  K  is  a  fixed  point  for  all  positions  of  P 
which  satisfy  the  equation  (1).  Also  KP2=KA.KB,  and  so  is  constant. 
Hence  the  locus  of  P  is  a  circle  -whose  centre  is  K. 

The  system  of  equations  obtained  by  varying  X  represents  a  system  of 
circles,  and  every  circle  of  this  system  cuts  at  right  angles  every  circle  of  the 
system  of  Ex.  19. 

The  system  of  Ex.  19  is  called  a  system  of  coaxal  circles  of  the  common 
point  kind.  The  system  of  Ex.  20  is  called  a  system  of  coaxal  circles  of  the 
limiting  point  kind,  A  and  B  being  the  limiting  points  of  the  system.  If  X 
is  very  large  or  very  small  then  the  circle  is  a  very  small  circle  containing  A 
or  B  in  its  interior. 

21.    Bilinear  Transformations.    Consider  the  equation 

z=Z+a (1), 

where  z  =  x+yi  and  Z=X+Yi  are  two  complex  variables  which  we  may 
suppose  to  be  represented  in  two  planes  xoy,  XOY.  To  every  value  of  z 
corresponds  one  of  Z,  and  conversely.     If  a  =  a+@i  then 

x=X+a,    y=Y+p, 

and  to  the  point  (x,  y)  corresponds  the  point  (X,  Y).  If  (x,  y)  describes  a 
curve  of  any  kind  in  its  plane,  (X,  Y)  describes  a  curve  in  its  plane.  Thus 
to  any  figure  in  one  plane  corresponds  a  figure  in  the  other.  A  passage  of 
this  kind  from  a  figure  in  the  plane  xoy  to  a  figure  in  the  plane  XOY  by 
means  of  a  relation  such  as  (1)  between  z  and  Z  is  called  a  transformation. 
In  this  particular  case  the  relation  between  corresponding  figures  is  very 
easily  defined.  The  (X,  Y)  figure  is  the  same  in  size,  shape,  and  orientation 
as  the  (x,  y)  figure,  but  is  shifted  a  distance  a  to  the  left,  and  a  distance  /3 
downwards.     Such  a  transformation  is  called  a  translation. 

Now  consider  the  equation 

*=pZ (2), 

where  p  is  real.  This  gives  x  —  pX,y=pY.  The  two  figures  are  similar  and 
similarly  situated  about  their  respective  origins,  but  the  scale  of  the  (x,  y) 
figure  is  p  times  that  of  the  (X,  Y)  figure.  Such  a  transformation  is  called 
a  magnification. 

Finally  consider  the  equation 

s  =  (cos  cp  +  ^sin  (p)  Z (3). 

It  is  clear  that  |  z  |  =  |  Z  |  and  that  one  value  of  am  z  is  am  Z+  <p,  and  that  the 
two  figures  differ  only  in  that  the  (x,  y)  figure  is  the  (X,  Y)  figure  turned 
about  the  origin  through  an  angle  <p  in  the  positive  direction.  Such  a  trans- 
formation is  called  a  rotation. 

The  general  linear  transformation 

z  =  aZ-vb (4) 


46]  COMPLEX   NUMBERS  95 

is  a  combination  of  the  three  transformations  (1),  (2),  (3).  For,  if  |  a\=p  and 
am  a  —  <p,  we  can  replace  (4)  by  the  three  equations 

z  =  z'  +  b,     z'=pZ',     Z'  =  (cos  (p  +  i  sin  <p)  Z. 

Thus  the  general  linear  transformation  is  equivalent  to  the  combination  of  a 
translation,  a  magnification,  and  a  rotation. 

Next  let  us  consider  the  transformation 

2=1/^ (5). 

If  \Z\  =  R  and  amZ=0,  then  \z\  =  l/R  and  am  z  =  -9,  and  to  pass  from 
the  (a?,  y)  figure  to  the  (X,  Y)  figure  we  invert  the  former  with  respect  to  o, 
with  unit  radius  of  inversion,  and  then  construct  the  image  of  the  new  figure 
in  the  axis  ox  {i.e.  the  symmetrical  figure  on  the  other  side  of  ox). 

Finally  consider  the  transformation 

aZ+b 

z=WVd (6)- 

This  is  equivalent  to  the  combination  of  the  transformations 

z  =  {ajc)  +  (bc-ad)(z'lc),     z'=l/Z\     Z'  =  cZ+d, 

i.e.  to  a  certain  combination  of  transformations  of  the  types  already  con- 
sidered. 

The  transformation  (6)  is  called  the  general  bilinear  transformation. 
Solving  for  Z  we  obtain 

7^}dz  —  b 
cz  —  a' 

The  general  bilinear  transformation  is  the  most  general  type  of  trans- 
formation for  which  one  and  only  one  value  of  z  corresponds  to  each  value  of 
Z,  and  conversely. 

22.  The  general  bilinear  transformation  transforms  circles  into  circles. 
This  may  be  proved  in  a  variety  of  ways.  We  may  assume  the  well-known 
theorem  in  pure  geometry,  that  inversion  transforms  circles  into  circles 
(which  may  of  course  in  particular  cases  be  straight  lines).  Or  we  may 
use  the  results  of  Exs.  19  and  20.     If,  e.g.,  the  (x,  y)  circle  is 

\(m- *)/(*- p)\=\ 
and  we  substitute  for  z  in  terms  of  Z,  we  obtain 

\(Z-<r')/(Z-p')\  =  X, 


a  —  pc 


,  ,         b  —  ad         ,        b  —  pd         •. 

where  o-  = ,     p= —  ,      X  = 

a—o-c  a  —  pc 

23.  Consider  the  transformations  z  =  \\Z,  z=(l+Z)/(l-Z),  and  draw 
the  (X,  Y)  curves  which  correspond  to  (1)  circles  whose  centre  is  the  origin, 
(2)  straight  lines  through  the  origin. 


96  COMPLEX   NUMBERS  [ill 

24.  The  condition  that  the  transformation  z  =  (aZ+b)j(cZ+d)  should 
make  the  circle  x2+y2  =  l  correspond  to  a  straight  line  in  the  (X,  Y)  plane 

is  \a\  =  \c\. 

25.  Cross  ratios.     The  cross  ratio  [z\Zi,  z$Z\)  is  defined  to  be 

(g1-g3)(z2-g4) 
fa -24)  02-%)' 

If  the  four  points  zu  22,  23,  zi  are  on  the  same  line,  this  definition  agrees 
with  that  adopted  in  elementary  geometry.  There  are  24  cross  ratios  which 
can  be  formed  from  zx,  z2,  z3,  z±  by  permuting  the  suffixes.  These  consist  of 
six  groups  of  four  equal  cross  ratios.  If  one  ratio  is  X,  then  the  six  distinct 
cross  ratios  are  X,  1  -  X,  1/X,  1/(1  -X),  (X-  1)/X,  X/(X-  1).  The  four  points  are 
said  to  be  harmonic  or  harmonically  related  if  any  one  of  these  is  equal  to 
—  1,     In  this  case  the  six  ratios  are  —  1,  2,  —  1,  J,  2,  \. 

If  any  cross  ratio  is  real  then  all  are  real  and  the  four  points  lie  on  a 
circle.     For  in  this  case 

am(;i-%)(^r^) 
(zi-Zi){z2-z3) 

must  have  one  of  the  three  values  —  n,  0,  -rr,  so  that  am  {{z1  —  s3)/(«i  -  24)}  and 
a,m{(z2-  z3)/(z2-  zi)}  must  either  be  equal  or  differ  by  ir  (cf.  Ex.  19). 

If  (2^2)  z3Zi)=  -  1,  we  have  the  two  equations 

Z-i  —  Zo  Zn  —  Z3  Z\  —  Zo\        \Zn  —  Zo\ 

am  -1— -=  ±7r  +  am- — -;        - — -  = — -  . 

zl  ~  Zi  zi~z\  I  zl  —  z4  I         I  z2  ~  -4  I 

The  four  points  AX,  A2,  A3,  A4  lie  on  a  circle,  Ax  and  A2  being  separated 
by  A3  and  Av  Also  AXA3\AX Ai  =  A2A3jA2Ai.  Let  0  be  the  middle  point  of 
A3Ai.      The  equation 

(zi-z3)(^-zi)^     1 

(Zl-Zi)(z2-Z3) 

may  be  put  in  the  form 

(h  +  22)  (*3  +  Zi)  =  2  (zl z2  +  hzl\ 

or,  what  is  the  same  thing, 

{h  -i(h+ 24)}  {22  -  i  (23 + 24)}  =  {i(z3  -  ?4)}2. 

But  this  is  equivalent  to  OAx .  OA2=~OA32=OAi2.  Hence  OAx  and  0A2 
make  equal  angles  with  A3At,  and  OA1.OA2  =  OA32=OAi2.  It  will  be  ob- 
served that  the  relation  between  the  pairs  Alf  A2  and  A3,  A4  is  symmetrical. 
Hence,  if  C  is  the  middle  point  of  AXA2,  0'A3  and  0'Ai  are  equally  inclined 
to  AXA2,  and  0'A3.  0'Ai=0'A12  =  0'A2*. 

26.  If  the  points  Alt  A2  are  given  by  az2  +  2bz  +  c  =  0,  and  the  points 
A3,  Ai  by  a'z2  +  2b'z+c'  =  0,  and  0  is  the  middle  point  of  A3A4,  and 
ac'  +  a'c -  2bb'  =  0,  then  O.^,  0A2  are  equally  inclined  to  A3A4  and 
OJi .  0J2  =  OA32  =  ai42.  (Jfatf.  Trip.  1901.) 


46]  COMPLEX   NUMBERS  97 

27.  AB,  CD  are  two  intersecting  lines  in  Argand's  diagram,  and  P, 
Q  their  middle  points.  Prove  that,  if  A  B  bisects  the  angle  CPD  and 
PA2  =  PBi=PC.PD,  then  CD  bisects  the  angle  AQB  and  QC2=QD2  =  QA .  QB. 

{Math.  Trip.  1909.) 

28.  The  condition  that  four  points  should  lie  on  a  circle.  A 
sufficient  condition  is  that  one  (and  therefore  all)  of  the  cross  ratios 
should  be  real  (Ex.  25)  ;  this  condition  is  also  necessary.  Another  form 
of  the  condition  is  that  it  should  be  possible  to  choose  real  numbers 
a,  |8,  y  such  that 

1  1  1  =0. 

a  /3  y 

Z\Zi  +  Z2Zz      Z23i  +  Z3zl       Z^  +  Z^ 

[To  prove  this  we  observe  that  the  transformation  Z=  1/(2  —  z4)  is  equivalent 
to  an  inversion  with  respect  to  the  point  24,  coupled  with  a  certain  reflexion 
(Ex.  21).  If  zt,  z2,  z3  lie  on  a  circle  through  24,  the  corresponding  points 
^  =  1/(24-24),  Z2  =  l/(z2  —  zt),  Z3=  l/(23  —  24)  lie  on  a  straight  line.  Hence 
(Ex.  12)  we  can  find  real  numbers  a,  /3',  y  such  that  a'  +  /3'  +  y'  =  0  and 
a'l(zl-zi)+fi'l(z2  —  zl)  +  y'j(z3-zi)  =  0,  and  it  is  easy  to  prove  that  this  is 
equivalent  to  the  given  condition.] 

29.  Prove  the  following  analogue  of  De  Moivre's  Theorem  for  real 
numbers:   if  (pi,  (p2,  $3,  -••  is  a  series  of  positive  acute  angles  such  that 

tan  <p  m  +  j = tan  <p  m  sec  <fi  j  4-  sec  <f)m  tan  <f>  t , 
then  tan  <£m  +  n=tan  <£„,  sec  0,,  +  sec  (pm  tan  c/>,v, 

sec  0m  +  n=sec  (pm  sec  <£„  +  tan  4>m  tan  0„, 
and  tan  (f>m  +  seo  <pm=(tan  (/>!  +  sec  <pi)m. 

[Use  the  method  of  mathematical  induction.] 

30.  The  transformation  z=Zn.  In  this  case  r=Rm,  and  6  and  me 
differ  by  a  multiple  of  2n\  If  Z  describes  a  circle  round  the  origin  then  z 
describes  a  circle  round  the  origin  m  times. 

The  whole  (a?,  y)  plane  corresponds  to  any  one  of  m  sectors  in  the  (A',  Y) 
plane,  each  of  angle  2jr/»i.  To  each  point  in  the  (a?,  y)  plane  correspond 
m  points  in  the  (X,  Y)  plane. 

31.  Complex  functions  of  a  real  variable.  If  fit),  <f>  (t)  are  two  real 
functions  of  a  real  variable   t  defined  for  a  certain  range  of  values  of  t, 

we  call 

z=f(t)  +  i<f>(t)  (1) 

a  complex  function  of  t.     We  can  represent  it  graphically  by  drawing  the 
curve 

*=/(*),      Jf=0  (0  i 

h.  7 


98  COMPLEX   NUMBERS  [ill 

the  equation  of  the  curve  may  be  obtained  by  eliminating  t  between  these 
equations.  If  z  is  a  polynomial  in  t,  or  rational  function  of  t,  with  complex 
coefficients,  we  can  express  it  in  the  form  (1)  and  so  determine  the  curve 
represented  by  the  function. 

(i)     Let  z  =  a  +  (b-a)t, 

where  a  and  b  are  complex  numbers.     If  a  =  a  +  a'i,  &  =  /3  +  j3'i,  then 
A?=a+(/3-aK     y  =  a'  +  (P'~a')t. 

The  curve  is  the  straight  line  joining  the  points  z  =  a  and  z=b.  The  seg- 
ment between  the  points  corresponds  to  the  range  of  values  of  t  from  0 
to  1.  Find  the  values  of  t  which  correspond  to  the  two  produced  segments 
of  the  line. 

where  p  is  positive,  then  the  curve  is  the  circle  of  centre  c  and  radius  p.  As 
t  varies  through  all  real  values  z  describes  the  circle  once. 

(iii)  In  general  the  equation  z  =  (a  +  bt)l(c+dt)  represents  a  circle. 
This  can  be  proved  by  calculating  x  and  y  and  eliminating :  but  this  process 
is  rather  cumbrous.  A  simpler  method  is  obtained  by  using  the  result  of 
Ex.  22.  Let  z=(a+bZ)/(c  +  dZ),  Z—t.  As  t  varies  Z  describes  a  straight 
line,  viz.  the  axis  of  X.     Hence  z  describes  a  circle. 

(iv)     The  equation  z=a  +  2bt  +  ct2 

represents  a  parabola  generally,  a  straight  line  if  b/c  is  real. 

(v)     The  equation  z  =  (a  +  2bt  +  ct2)/(a  +  2fit  +  yt2),  where  a,  /3,  y  are  real, 
represents  a  conic  section. 
[Eliminate  t  from 
Z=(A+2Bt+Cti)/(a  +  2pt  +  yf-),     y  =  (A'  +  2B't  +  C't2)/(a  +  2j3t+yf-), 
where  A+A'i=a,  B  +  B'i=b,  C+C'i=c] 

47.  Roots  of  complex  numbers.  We  have  not,  up  to  the 
present,  attributed  any  meaning  to  symbols  such  as  ya,  amln, 
when  a  is  a  complex  number,  and  m  and  n  integers.  It  is, 
however,  natural  to  adopt  the  definitions  which  are  given  in 
elementary  algebra  for  real  values  of  a.  Thus  we  define  v'a  or 
aVn,  where  n  is  a  positive  integer,  as  a  number  z  which  satisfies 
the  equation  z11  =  a ;  and  amln,  where  m  is  an  integer,  as  (a1"1)"1. 
These  definitions  do  not  prejudge  the  question  as  to  whether 
there  are  or  are  not  more  than  one  (or  any)  roots  of  the  equation. 

48.  Solution  of  the  equation  zn  =  a.     Let 

a  =  p  (cos  <f>  +i  sin  tp), 
where  p  is  positive  and  </>  is  an  angle  such  that  —  ir  <  (f>  g  tt.     If 


46-48]  COMPLEX   NUMBERS  99 

we  put  z  =  r  (cos  6  +  i  sin  6),  the  equation  takes  the  form 

rn  (cos  lid  +  i  sin  nO)  =  p  (cos  cp  +  i  sin  cp) ; 

so  that  rn  =  p,     cos  nO  =  cos  cp,     sin  nd  =  sine/)    (1). 

The  only  possible  value  of  r  is  typ,  the  ordinary  arithmetical 
?ith  root  of  p ;  and  in  order  that  the  last  two  equations  should  be 
satisfied  it  is  necessary  and  sufficient  that  nO  =  <p  +  2kir,  where  k 
is  an  integer,  or 

0  =  (tj>  +  2hir)\n. 

If  k=pn  +  q,  where  p  and  q  are  integers,  and  0^q<n,  the 
value  of  6  is  2p7r  +  (cp  4-  2qir)/ii,  and  in  this  the  value  of  p  is  a 
matter  of  indifference.     Hence  the  equation 

zn  =  a  =  p  (cos  cp  +  i  sin  0) 
Aas  ?i  roofe  ancZ  ?i  only,  given  by  z  =  r  (cos  0  +  i  sin  0),  where 
r  =  typ,     0  =  (<p  +  2qTr)/n,     (q=  0,  1,  2,  ...  n-1). 

That  these  ?i  roots  are  in  reality  all  distinct  is  easily  seen 
by  plotting  them  on  Argand's  diagram.     The  particular  root 

"HJp  {cos  (<p/ri)  +  i  sin  (cpjn)) 

is  called  the  principal  value  of  %/a. 

The  case  in  which  a=l,  p  =  l,  <f>  =  0  is  of  particular  interest. 
The  n  roots  of  the  equation  xn  =  1  are 

cos  (2q7r/n)  +  i  sin  (2q7r/n),     (q  =  0,  1,  . . .  n  —  1). 
These  numbers  are  called  the  nth  roots  of  unity;  the  principal 
value  is  unity  itself.     If  we  write  oon  for  cos  (2Tr/n)  +  i  sin  (2-irjn), 
we  see  that  the  nth.  roots  of  unity  are 

1,    con,    col,  •••  ^'T1- 

Examples  XXII.  1.  The  two  square  roots  of  1  are  1,  - 1  ;  the  three 
cube  roots  are  1,  £(  —  l+i^/3),  ^(  —  l—i^/3);  the  four  fourth  roots  are  1, 
i,  —  1,   —  i;  and  the  five  fifth  roots  are 

1,     i[     ^5-1+^(10  +  2^/5}],      ^[-^5- 1  +  ^/(10-2^/5}], 

i[_s/5-l-iv/(10-2V5}],     i[     J5-l-iJ{l0  +  2s/5}]. 

2.  Prove  that  l  +  con+G)^+...  +  o>"~  =0. 

3.  Prove  that  (x+ycos  +  zo>l)(x+ya>l+za>3)=x2+y2-\-z2  — yz  —  zx-xy. 

4.  The  wth  roots  of  a  are  the  products  of  the  ?ith  roots  of  unity  by  the 
principal  value  of  Hja. 

7—2 


100  COMPLEX   NUMBERS  [ill 

5.  It  follows  from  Exs.  xxi.  14  that  the  roots  of 

are  ±  v'[fc  W(a*+&)+a}]±ilM  W(«2+/32)  -  a}J 

like  or  unlike  signs  being  chosen  according  as  j3  is  positive  or  negative.     Show 
that  this  result  agrees  with  the  result  of  §  48. 

6.  Show  that  (x2m-  a2m)/(a;2  -  a2)  is  equal  to 

#-  — 2a#  cos  —  +  cr )  (  ^J-  zturcos \-w  ) ...    a--  2«.r  cos — Ya~ ) . 

\  m       J  \  in        J      \  m  J 

[The  factors  of  x2m  -  a2m  ai'e 

(x-a),     (x-aa2m),     (x-aJ2J,  ...  (x-aa>^~1). 

The  factor  x  -  aa>™m  is  x  +  a.  The  factors  (a?  -  aog,,,),  (a;  -  «w^"s)  taken  together 
give  a  factor  #2  —  2ax  cos  (sir/in)  +  a2.] 

7.  Kesolve  A,2",  +  1-a2"l  +  1,  #2m  +  a2m,  and  xim  +  1  +  a2m  +  1  into  factors  in  a 
similar  way. 

8.  Show  that  x2n  -  2xnan  cos  8  +  a2n  is  equal  to 


Lp- 


8       „\  /  „     _            0  +  2tt       , 
2.ra  cos  -  +  cr )    ar  -  2.ra  cos h  a' 


0  +  2(w-lW      2 

\xa  cos — Yar 

n 


...(x*-2xc 

[Use  the  formula 
x2n  -  2xnan  cos  8  +  a2n = {a,*1  -  an  (cos  8  + 1  sin  5)}  {#»  -  an  (cos  0  -  j  sin  8)}, 
and  split  up  each  of  the  last  two  expressions  into  n  factors.] 

9.  Find  all  the  roots  of  the  equation  xG  -  2x3  +  2  =  0.     (Math.  Trip.  1910.) 

10.  The  problem  of  finding  the  accurate  value  of  a>n  in  a  numerical  form 
involving  square  roots  only,  as  in  the  formula  co3  =  f  (-  l  +  z'^/3),  is  the 
algebraical  equivalent  of  the  geometrical  problem  of  inscribing  a  regular 
polygon  of  n  sides  in  a  circle  of  unit  radius  by  Euclidean  methods,  i.e.  by  ruler 
and  compasses.  For  this  construction  will  be  possible  if  and  only  if  we  can 
construct  lengths  measured  by  cos  (2ir/ri)  and  sin  (2ir/n)  ;  and  this  is  possible 
(Ch.  II,  Misc.  Exs.  22)  if  and  only  if  these  numbers  are  expressible  in  a  form 
involving  square  roots  only. 

Euclid  gives  constructions  for  n  =  o,  4,  5,  6,  8,  10,  12,  and  15.  It  is 
evident  that  the  construction  is  possible  for  any  value  of  n  which  can  be 
found  from  these  by  multiplication  by  any  power  of  2.  There  are  other 
special  values  of  n  for  which  such  constructions  are  possible,  the  most  inter- 
esting being  n  =  \7. 


48,  49]  COMPLEX   NUMBERS  101 

49.     The    general  form    of  De  Moivre's  Theorem.     It 

follows  from  the  results  of  the  last  section  that  if  q  is  a  positive 
integer  then  one  of  the  values  of  (cos  0  +  i  sin  6)1,q  is 

cos  (0  /  q)  +  i  sin  (0/q). 

Raising  each  of  these  expressions  to  the  power  p  (where  p  is  any 
integer  positive  or  negative),  we  obtain  the  theorem  that  one  of 
the  values  of  (cos  0  +  i  sin  0)pIv  is  cos  (p0/q)  +  i  sin  (p0/q),  or  that  if 
a  is  any  rational  number  then  one  of  the  values  of  (cos  0  +  i  sin  0)a  is 

cos  a0  +  %  sin  a.0. 

This  is  a  generalised  form  of  De  Moivre's  Theorem  (§  45). 


MISCELLANEOUS   EXAMPLES   ON  CHAPTER  III. 

1.  The  condition  that  a  triangle  {xyz)  should  be  equilateral  is  that 

x2+y2+z2  —  yz  — zx  —  xy  =  0. 

[Let  XYZ  be  the  triangle.     The  displacement  ZX  is  YZ  turned  through 
an  angle  fn-  in    the    positive    or    negative    direction.      Since   Cis  §7r  =  <o3, 

Cis(-§7r)  =  l/<»3=a>3,  we  have  x-z  =  (z-y)  w3  or  x-z  =  (z—  y)  <a\.      Hence 

x+ya>3+zodl=0  or  x+yul  +  za.^0.     The  result  follows  from  Exs.  xxn.  3.] 

2.  If  X  YZ,  X'  Y'Z  are  two  triangles,  and 


YZ.  Y'Z'  =  ZX .  Z'X'  =  XY.  XT, 

then  both  triangles  are  equilateral.     [From  the  equations 

(y  -  2)  (/  -  z')  =  (z-x)  (z'  -x')  =  (x  -  y)  (x'  -  y')  =  k% 
say,  we  deduce  2  l/(/  -  z)  =  0,  or  2x"2  -  2y'z'=0.     Now  apply  the  result  of  the 
last  example.] 

3.  Similar  triangles  BOX,  CAY,  ABZ  are  described  on  the  sides  of  a 
triangle  ABC.   Show  that  the  centres  of  gravity  of  ABC,  XYZ  are  coincident. 

[We  have   (x-c)/(b  —  c)  =  (y-a)/(c-a)  =  (z-b)/(a  —  b)  =  \,   say.      Express 
i|  (x+y+z)  in  terms  of  a,  b,  <?.] 

4.  If  X,  Y,  Z  are  points  on  the  sides  of  the  triangle  ABC,  such  that 

BX/XC=  CYI YA  =  AZ\ZB  =  r, 
and  if    ABC,   XYZ  are    similar,   then   either  r=l   or   both  triangles   are 
equilateral. 

5.  If  A ,  B,  C,  D  are  four  points  in  a  plane,  then 

AD.BC^BD.CA  +  CD.AB. 


102  COMPLEX    NUMBERS  [ill 

[Let  2X,  z2,  Z3,  24  be  the  complex  numbers  corresponding  to  A,  B,  C,  D. 
Then  we  have  identically 

(Xi  -  #4)  (a?2  -  Xz)  +  (*2  -  #4)  0*3  -  #l)  +  (#3  ~  #4)  (#1  -  #2)  =  0- 

Henco 

I  (#1  ~  #4)  (#2  ~  #s)  !  =  I  (#2  -  #4)  (#3  ~  X\)  +  (#3  -  %i)  (#1  -  #2)  I 

<  I  (d?2  -  *«)  (#3  -  #l)  I  + 1  (#3 -  #4)  (#1  -  ^2)  I-] 

6.  Deduce  Ptolemy's  Theorem  concerning  cyclic  quadrilaterals  from  the 
fact  that  the  cross  ratios  of  four  concyclic  points  are  real.  [Use  the  same 
identity  as  in  the  last  example.] 

7.  If  22  +  2'2=  1,  then  the  points  2,  z'  are  ends  of  conjugate  diameters  of  an 
ellipse  whose  foci  are  the  points  1,  - 1.  [If  CP,  CD  are  conjugate  semi- 
diameters  of  an  ellipse  and  S,  H  its  foci,  then  CD  is  parallel  to  the  external 
bisector  of  the  angle  SPH,  and  SP .  HP=  CD2.] 

8.  Prove  that  \a  +  b\2+\a-b\2=2{\a\2  +  \b\2}.  [This  is  the  analytical 
equivalent  of  the  geometrical  theorem  that,  if  M  is  the  middle  point  of  PQ, 
then  OP2  +  OQ2=20M2  +  2MP2.] 

9.  Deduce  from  Ex.  8  that 

\a  +  <J{a2-b2)\  +  \a->J(a2-b2)\  =  \a  +  b\  +  \a-b\. 

[If  a+V(«2 - 1'2)  =  zu  a~  <sA>2  -  b2)  =  z2,  we  have 

\h\2+\^\2^\zl  +  z2\2+h\z1-z2\2=2\a\2  +  2\a2-b2\, 

and  so     (\zi\  +  \z2\)2=2{\a\2+\a2-b2\  +  \b\2}  =  \a  +  b\2  +  \a-b\2  +  2\a2-b2\. 

Another  way  of  stating  the  result  is  :  if  zx  and  2,  are  the  roots  of 
az2+2pz  +  y  =  0,  then 

|*1M*2|  =  (l/|«|){(|-/3Way|)  +  (|-/Way|)}.] 

10.  Show  that  the*  necessary  and  sufficient  conditions  that  both  the  roots 
of  the  equation  z2  +  az  +  b=0  should  be  of  unit  modulus  are 

|a|  =  2,     I  6  |  =  1,     am  6  =  2  am  a. 
[The  amplitudes  have  not  necessarily  their  principal  values.] 

11.  If  %i+4a1x3  +  Qa2.v2  +  4a3x+ai=0  is  an  equation  with  real  coefficients 
and  has  two  real  and  two  complex  roots,  concyclic  in  the  Argand  diagram,  then 

«32  +  «i2«4  +  «23  -  «2«4  -  2a1a2«3  =  0. 

12.  The  four  roots  of  a^  +  4aix3  +  (>a2x2  +  4a3.v  +  a4  =  0  will  be  harmonic- 
ally related  if 

a0a32  +  afai  +  a23  -  a0a2a4  -  2axa2az  =■  0. 

[Express    ^23,14^31,24^12,34,    where    ^23.14= (21-22)  (23-24) +(%-%)  («2-«4) 

and  2j,  z2,  23,  24  are  the  roots  of  the  equation,  in  terms  of  the  coefficients.] 


COMPLEX    NUMBERS  103 

13.  Imaginary  points  and  straight  lines.  Let  ax+by  +  c  =  0  be 
an  equation  with  complex  coefficients  (which  of  course  may  be  real  in  special 

cases). 

If  we  give  x  any  particular  real  or  complex  value,  we  can  find  the  corre- 
sponding value  of  y.  The  aggregate  of  pairs  of  real  or  complex  values  of  x 
and  y  which  satisfy  the  equation  is  called  an  imaginary  straight  line  ;  the 
pairs  of  values  are  called  imaginary  points,  and  are  said  to  lie  on  the  line. 
The  values  of  x  and  y  are  called  the  coordinates  of  the  point  (x,  y).  When 
x  and  y  are  real,  the  point  is  called  a  real  point :  when  a,  b,  c  are  all  real  (or 
can  be  made  all  real  by  division  by  a  common  factor),  the  line  is  called  a  real 
line.  The  points  x=a+j3i,  y  =  y  +  8i  and  x=a  —  fti,  y=y  —  Si  are  said  to  be 
conjugate  ;  and  so  are  the  lines 

(A  +  A'i)  x  +  (B  +  B'i)y  +  G+  C'i=0,     (A  -  A'i) x+(B - B'i)y  +  C-  C'i=0. 

Verify  the  following  assertions  : — every  real  line  contains  infinitely  many 
pairs  of  conjugate  imaginary  points  ;  an  imaginary  line  in  general  contains 
one  and  only  one  real  point ;  an  imaginary  line  cannot  contain  a  pair  of 
conjugate  imaginary  points  : — and  find  the  conditions  (a)  that  the  line 
joining  two  given  imaginary  points  should  be  real,  and  (b)  that  the  point 
of  intersection  of  two  imaginary  lines  should  be  real. 

14.  Prove  the  identities 

(x+y  +  z)(x+y<03+za>l)  (x +ya>;+  zo>3) = Xs +y3  +  z3-  3xyz, 
(x+y+z)  (x+y<o5+zco45)  (x+y<o\+  za>35)  (x+ya>\+  za>l)  (x+ycol+z^) 
= x5  +y5  +  z6  —  bxhjz  +  5xy2z2. 

15.  Solve  the  equations 

x3-3ax+(a3  +  l)  =  0,     ^5-5a^+5a2.c  +  (a5  +  l)  =  0. 

1 6.  If/  (x)  =  «0  +  axx  + . . .  +  akxk,  then 

{/(•*')  +f(<°x)  +  •••+/  (*>"  ~  l*))l* = «o + <vn  +  v* + •  •  •  +  <vAw, 

<o  being  any  root  of  xn  =  l  (except  x=l),  and  \n  the  greatest  multiple  of  n 
contained  in  h.     Find  a  similar  formula  for  a  +a  +  „•*"  +  «„ +2>4'*;2'l+  •••• 

1 7.  If  (1  +x)n=p0  +Pix  +p,x2  +..., 
n  being  a  positive  integer,  then 

Po ~ Pz+P\ -•■■  =  23"  cos  frnr,    pi  -p3+Po -  •••  =22M sin \nir. 

18.  Sum  the  series 

X  X2  X3  xnl3 

2!%-2!+5!  «-5!+8!  n-8\  +  "'+n-ll' 

n  being  a  multiple  of  3.  (Math.  Trip.  1899.) 

19.  It  t  is  a  complex  number  such  that  |£|  =  1,  then  the  point 
x=(at  +  b)j(t  —  c)  describes  a  circle  as  t  varies,  unless  |c|  =  l,  when  it 
describes  a  straight  line. 


104  COMPLEX   NUMBERS  [ill 

20.  If  t  varies  as  in  the  last  example  then  the  point  x=^{at  +  (b/t)}  in 
general  describes  an  ellipse  whose  foci  are  given  by  x2—ab,  and  whose  axes 
are  |  a  |  + 1  b  |  and  |  a  \  -  \  b  |.  But  if  |  a  |  =  |  b  |  then  x  describes  the  finite  straight 
line  joining  the  points  -J(ab),  y/(ab). 

21.  Prove  that  if  t  is  real  and  z=t2-  l+J(t^  —  t2),  then,  when  t2<l,  z  is 
represented  by  a  point  which  lies  on  the  circle  x2+y2+x=0.  Assuming  that, 
when  t2>l,  sj(ft-t2)  denotes  the  positive  square  root  of  fi-t2,  discuss  the 
motion  of  the  point  which  represents  z,  as  t  diminishes  from  a  large  positive 
value  to  a  large  negative  value.  {Math.  Trip.  1912.) 

22.  The  coefficients  of  the  transformation  z=(aZ+b)/(cZ+d)  are  subject 
to  the  condition  ad— bc  =  l.  Show  that,  if  c4=0,  there  are  two  fixed  points 
a,  /3,  i.e.  points  unaltered  by  the  transformation,  except  when  (a  +  d)2=4,  when 
there  is  only  one  fixed  point  a  ;  and  that  in  these  two  cases  the  transforma- 
tion may  be  expressed  in  the  forms 

z  —  a       „  Z  —  a  1  1  „ 

Show  further  that,  if  c=0,  there  will  be  one  fixed  point  a  unless  a  =  d, 

and  that  in  these  two  cases  the  transformation   may  be  expressed   in  the 

forms 

Z-a=K{Z-a),     z=Z+K. 

Finally,  if  a,  b,  c,  d  are  further  restricted  to  positive  integral  values  (in- 
cluding zero),  show  that  the  only  transformations  with  less  than  two  fixed 
points  are  of  the  forms  (l/a)  =  (l/Z)  +  A',  z  =  Z+K.  (Math.  Trip.  1911.) 

23.  Prove  that  the  relation  z  =  (l+Zi)/(Z+i)  transforms  the  part  of  the 
axis  of  x  between  the  points  z=\  and  z=  —  1  into  a  semicircle  passing 
through  the  points  Z=  1  and  Z=  —  1.  Find  all  the  figures  that  can  be  obtained 
from  the  originally  selected  part  of  the  axis  of  x  by  successive  applications  of 
the  transformation.  (Math.  Trip.  1912.) 

24.  If  z  =  2Z+Z2  then  the  circle  \Z\  =  \  corresponds  to  a  cardioid  in  the 
plane  of  z. 

25.  Discuss  the  transformation  z=^{Z+(l/Z)},  showing  in  particular 
that  to  the  circles  X2+  Y2  =  a2  correspond  the  confocal  ellipses 

-»+  „     V   „„  =  !• 


SH)F  fiC-Dr 


26.  If  (2  +  l)2  =  4/Zthen  the  unit  circle  in  the  2-plane  corresponds  to  the 
parabola  Acos2|e  =  l  in  the  ^-plane,  and  the  inside  of  the  circle  to  the 
outside  of  the  parabola. 

27.  Show  that,  by  means  of  the  transformation  z  =  {(Z—  ci)l(Z+ci)}2, 
the  upper  half  of  the  2-plane  may  be  made  to  correspond  to  the  interior  of 
a  certain  semicircle  in  the  Z-plane. 


COMPLEX   NUMBERS  10o 

28.  If  z  =  Z2-l,  then  as  z  describes  the  circle  |*|  =  k,  the  two  corre- 
sponding positions  of  Z  each  describe  the  Cassinian  oval  pip2=K  where 
Pi,  p2  are  the  distances  of  Z  from  the  points  -1,  1.  Trace  the  ovals  for 
different  values  of  k. 

29.  Consider  the  relation  az2  +  2hzZ+  bZ2  +  2gz  +  2/Z+  c  =  0.  Show  that 
there  are  two  values  of  Z  for  which  the  corresponding  values  of  z  are  equal, 
and  vice  versa.  We  call  these  the  branch  points  in  the  Z  and  2-planes  re- 
spectively. Show  that,  if  z  describes  an  ellipse  whose  foci  are  the  branch 
points,  then  so  does  Z. 

[We  can,  without  loss  of  generality,  take  the  given  relation  in  the  form 

z2  +  2zZcos(o  +  Z2=l  : 

the  reader  should  satisfy  himself  that  this  is  the  case.  The  branch  points  in 
either  plane  are  cosec  a>  and  —  cosec  co.  An  ellipse  of  the  form  specified  is 
given  by 

|  z  -f  cosec  a>  |  + 1  z  —  cosec  a>\  =  C, 

where  C  is  a  constant.     This  is  equivalent  (Ex.  9)  to 

I  z+ */(z2  ~  cosec2  w)  |  + 1  z  -  s](z2  —  cosec2  a)  \  =  C. 

Express  this  in  terms  of  Z.] 

30      If  z=aZm  +  bZn,  where  m,  n  are  positive  integers  and  a,  b  real,  then 
as  Z  describes  the  unit  circle,  z  describes  a  hypo-  or  epi-cycloid. 

31.  Show  that  the  transformation 

Jp+di)Z0+b 

cZq  -  (a  -  di) ' 

where  a,  b,  c,  d  are  real  and  a2  +  d2  +  bc  >  0,  and  Z0  denotes  the  conjugate  of 
Z,  is  equivalent  to  an  inversion  with  respect  to  the  circle 

c  (x2  +y2)  -  2ax  -  2dy  -  b = 0. 

What  is  the  geometrical  interpretation  of  the  transformation  when 

a2  +  d2  +  bc<0l. 

32.  The  transformation 

i-g  _n-z^ 

1+Z~\l+Z; 

where  c  is  rational  and  0  <  c  <  1,  transforms  the  circle  |  z  \  =  1  into  the  boundary 
of  a  circular  lune  of  angle  njc. 


CHAPTER  IV 

LIMITS   OF   FUNCTIONS  OF   A   POSITIVE  INTEGRAL  VARIABLE 

50.     Functions    of   a    positive    integral    variable.      In 

Chapter  II  we  discussed  the  notion  of  a  function  of  a  real 
variable  x,  and  illustrated  the  discussion  by  a  large  number  of 
examples  of  such  functions.  And  the  reader  will  remember  that 
there  was  one  important  particular  with  regard  to  which  the 
functions  which  we  took  as  illustrations  differed  very  widely. 
Some  were  defined  for  all  values  of  x,  some  for  rational  values 
only,  some  for  integral  values  only,  and  so  on. 

Consider,  for  example,  the  following  functions  :  (i)  .r,  (ii)  njx,  (iii)  the 
denominator  of  x,  (iv)  the  square  root  of  the  product  of  the  numerator  and 
the  denominator  of  x,  (v)  the  largest  prime  factor  of  x,  (vi)  the  product  of 
*Jx  and  the  largest  prime  factor  of  x,  (vii)  the  #th  prime  number,  (viii)  the 
height  measured  in  inches  of  convict  x  in  Dartmoor  prison. 

Then  the  aggregates  of  values  of  x  for  which  these  functions  are  defined 
or,  as  we  may  say,  the  fields  of  definition  of  the  functions,  consist  of  (i)  all 
values  of  x,  (ii)  all  positive  values  of  x,  (iii)  all  rational  values  of  x,  (iv)  all 
positive  rational  values  of  x,  (v)  all  integral  values  of  at,  (vi),  (vii)  all  positive 
integral  values  of  x,  (viii)  a  certain  number  of  positive  integral  values  of  x, 
viz.,  1,  2,  ...,  N,  where  N  is  the  total  number  of  convicts  at  Dartmoor  at  a 
given  moment  of  time*. 

Now  let  us  consider  a  function,  such  as  (vii)  above,  which  is 
defined  for  all  positive  integral  values  of  x  and  no  others.     This 

*  In  the  last  case  N  depends  on  the  time,  and  convict  x,  where  x  has  a  definite 
value,  is  a  different  individual  at  different  moments  of  time.  Thus  if  we  take 
different  moments  of  time  into  consideration  we  have  a  simple  example  of  a 
function  y  =  F  (x,  t)  of  two  variables,  defined  for  a  certain  range  of  values  of  t,  viz. 
from  the  time  of  the  establishment  of  Dartmoor  prison  to  the  time  of  its  abandon- 
ment, and  for  a  certain  number  of  positive  integral  values  of  x,  this  number 
varying  with  t. 


50,  51]     FUNCTIONS    OF   A    POSITIVE    INTEGRAL   VARIABLE  107 

function  may  be  regarded  from  two  slightly  different  points  of 
view.  We  may  consider  it,  as  has  so  far  been  our  custom,  as  a 
function  of  the  real  variable  x  defined  for  some  only  of  the  values 
of  w,  viz.  positive  integral  values,  and  say  that  for  all  other  values 
of  x  the  definition  fails.  Or  we  may  leave  values  of  x  other 
than  positive  integral  values  entirely  out  of  account,  and  regard 
our  function  as  a  function  of  the  positive  integral  variable  n, 
whose  values  are  the  positive  integers 

1,2,3,4,.... 

In  this  case  we  may  write 

and  regard  y  now  as  a  function  of  n  defined  for  all  values  of  n. 

It  is  obvious  that  any  function  of  x  defined  for  all  values  of  x 
gives  rise  to  a  function  of  n  defined  for  all  values  of  n.  Thus  from 
the  function  y  =  x2  we  deduce  the  function  y  =  rC-  by  merely 
omitting  from  consideration  all  values  of  x  other  than  positive 
integers,  and  the  corresponding  values  of  y.  On  the  other  hand 
from  any  function  of  n  we  can  deduce  any  number  of  functions 
of  x  by  merely  assigning  values  to  y,  corresponding  to  values  of  x 
other  than  positive  integral  values,  in  any  way  we  please. 

51.  Interpolation.  The  problem  of  determining  a  function  of  x  which 
shall  assume,  for  all  positive  integral  values  of  x,  values  agreeing  with  those 
of  a  given  function  of  n,  is  of  extreme  importance  in  higher  mathematics. 
It  is  called  the  problem  of  functional  interpolation. 

Were  the  problem  however  merely  that  of  finding  some  function  of  x  to 
fulfil  the  condition  stated,  it  would  of  course  present  no  difficulty  whatever. 
We  could,  as  explained  above,  simply  fill  in  the  missing  values  as  we  pleased  : 
we  might  indeed  simply  regard  the  given  values  of  the  function  of  n  as  all 
the  values  of  the  function  of  x  and  say  that  the  definition  of  the  latter 
function  failed  for  all  other  values  of  x.  But  such  purely  theoretical  solutions 
are  obviously  not  what  is  usually  wanted.  What  is  usually  wanted  is  some 
formula  involving  x  (of  as  simple  a  kind  as  possible)  which  assumes  the  given 
values  for  #=1,  2,  .... 

In  some  cases,  especially  when  the  function  of  n  is  itself  defined  by  a 
formula,  there  is  an  obvious  solution.  If  for  example  y  =  <fi  («),  where  cf>  {n) 
is  a  function  of  n,  such  as  n2  or  cos  mr,  which  would  have  a  meaning  even 
were  n  not  a  positive  integer,  we  naturally  take  our  function  of  x  to  be 
y  =  (f)(x).  But  even  in  this  very  simple  case  it  is  easy  to  write  down  other 
almost  equally  obvious  solutions  of  the  problem.     For  example 

y  =  cf)  (x)  +  sin  xn 
assumes  the  value  (p  (n)  for  x  =  n,  since  sin  ?in-  =  0. 


108  LIMITS   OF    FUNCTIONS   OF   A  [iV 

In  other  cases  (/>  (n)  may  be  defined  by  a  formula,  such  as  (  —  l)n,  which 
ceases  to  define  for  some  values  of  x  (as  here  in  the  case  of  fractional  values 
of  x  with  even  denominators,  or  irrational  values).  But  it  may  be  possible 
to  transform  the  formula  in  such  a  way  that  it  does  define  for  all  values  of 
x.     In  this  case,  for  example, 

(  — l)n  =  COS«7T, 

if  n  is  an  integer,  and  the  problem  of  interpolation  is  solved  by  the  function 


In  other  cases  <p(x)  may  be  defined  for  some  values  of  x  other  than 
positive  integers,  but  not  for  all.  Thus  from  y  =  nn  we  are  led  to  y=xx. 
This  expression  has  a  meaning  for  some  only  of  the  remaining  values  of  x. 
If  for  simplicity  we  confine  ourselves  to  positive  values  of  x,  then  xx  has 
a  meaning  for  all  rational  values  of  .r,  in  virtue  of  the  definitions  of 
fractional  powers  adopted  in  elementary  algebra.  But  when  x  is  irrational 
xx  has  (so  far  as  we  are  in  a  position  to  say  at  the  present  moment)  no 
meaning  at  all.  Thus  in  this  case  the  problem  of  interpolation  at  once 
leads  us  to  consider  the  question  of  extending  our  definitions  in  such  a 
way  that  xx  shall  have  a  meaning  even  when  x  is  irrational.  We  shall  see 
later  on  how  the  desired  extension  may  be  effected. 

Again,  consider  the  case  in  which 

y  =  1 .  2  . . .  n  =  n  ! . 

In  this  case  there  is  no  obvious  formula  in  x  which  reduces  to  n  !  for  x=n, 
as  x\  means  nothing  for  values  of  x  other  than  the  positive  integers.  This 
is  a  case  in  which  attempts  to  solve  the  problem  of  interpolation  have  led  to 
important  advances  in  mathematics.  For  mathematicians  have  succeeded 
in  discovering  a  function  (the  Gamma-function)  which  possesses  the  desired 
property  and  many  other  interesting  and  important  properties  besides. 

52.  Finite  and  infinite  classes.  Before  we  proceed  further 
it  is  necessary  to  make  a  few  remarks  about  certain  ideas  of  an 
abstract  and  logical  nature  which  are  of  constant  occurrence  in 
Pure  Mathematics. 

In  the  first  place,  the  reader  is  probably  familiar  with  the 
notion  of  a  class.  It  is  unnecessary  to  discuss  here  any  logical 
difficulties  which  may  be  involved  in  the  notion  of  a  'class': 
roughly  speaking  we  may  say  that  a  class  is  the  aggregate  or 
collection  of  all  the  entities  or  objects  which  possess  a  certain 
property,  simple  or  complex.  Thus  we  have  the  class  of  British 
subjects,  or  members  of  Parliament,  or  positive  integers,  or  real 
numbers. 


51-53]  POSITIVE   INTEGRAL   VARIABLE  109 

Moreover,  the  reader  has  probably  an  idea  of  what  is  meant 
by  a  finite  or  infinite  class.  Thus  the  class  of  British  subjects 
is  a  finite  class:  the  aggregate  of  all  British  subjects,  past, 
present,  and  future,  has  a  finite  number  n,  though  of  course  we 
cannot  tell  at  present  the  actual  value  of  n.  The  class  of  present 
British  subjects,  on  the  other  hand,  has  a  number  n  which  could 
be  ascertained  by  counting,  were  the  methods  of  the  census 
effective  enough. 

On  the  other  hand  the  class  of  positive  integers  is  not  finite 
but  infinite.  This  may  be  expressed  more  precisely  as  follows. 
If  n  is  any  positive  integer,  such  as  1000, 1,000,000  or  any  number 
we  like  to  think  of,  then  there  are  more  than  n  positive  integers. 
Thus,  if  the  number  we  think  of  is  1,000,000,  there  are  obviously 
at  least  1,000,001  positive  integers.  Similarly  the  class  of  rational 
numbers,  or  of  real  numbers,  is  infinite.  It  is  convenient  to 
express  this  by  saying  that  there  are  an  infinite  number  of 
positive  integers,  or  rational  numbers,  or  real  numbers.  But  the 
reader  must  be  careful  always  to  remember  that  by  saying  this 
we  mean  simply  that  the  class  in  question  has  not  a  finite  number 
of  members  such  as  1000  or  1,000,000. 

53.  Properties  possessed  by  a  function  of  n  for  large 
values  of  n.  We  may  now  return  to  the  '  functions  of  n '  which  we 
were  discussing  in  §§  50 — 51.  They  have  many  points  of  difference 
from  the  functions  of  x  which  we  discussed  in  Chap.  II.  But  there 
is  one  fundamental  characteristic  which  the  two  classes  of  func- 
tions have  in  common :  the  values  of  the  variable  for  which  they 
are  defined  form  an  infinite  class.  It  is  this  fact  which  forms  the 
basis  of  all  the  considerations  which  follow  and  which,  as  we  shall 
see  in  the  next  chapter,  apply,  mutatis  mutandis,  to  functions  of  x 
as  well. 

Suppose  that  $(n)  is  any  function  of  n,  and  that  P  is  any 
property  which  <£  (n)  may  or  may  not  ha,ve,  such  as  that  of  being 
a  positive  integer  or  of  being  greater  than  1.  Consider,  for  each 
of  the  values  n—1,  2,  3,  ...,  whether  <f>(n)  has  the  property  P  or 
not.     Then  there  are  three  possibilities: — 

(a)  <f>  (n)  may  have  the  property  P  for  all  values  of  n,  or  for 
all  values  of  n  except  a  finite  number  N  of  such  values  : 


HO  LIMITS   OF   FUNCTIONS   OF   A  [IV 

(b)  <f)  (n)  may  have  the  property  for  no  values  of  n,  or  only  for 
a  finite  number  N  of  such  values : 

(c)  neither  (a)  nor  (b)  may  be  true. 

If  (6)  is  true,  the  values  of  n  for  which  $  (n)  has  the  property 
form  a  finite  class.  If  (a)  is  true,  the  values  of  n  for  which  $  (n) 
has  not  the  property  form  a  finite  class.  In  the  third  case  neither 
class  is  finite.     Let  us  consider  some  particular  cases. 

(1)  Let  0  (n)  =  n,  and  let  P  be  the  property  of  being  a  positive  integer. 
Then  0  (ft)  has  the  property  P  for  all  values  of  ft. 

If  on  the  other  hand  P  denotes  the  property  of  being  a  positive  integer 
greater  than  or  equal  to  1000,  then  <f>  (ft)  has  the  property  for  all  values  of  n 
except  a  finite  number  of  values  of  n,  viz.  1,  2,  3,  ...,  999.  In  either  of 
these  cases  (a)  is  true. 

(2)  If  cj)(n)  =  n,  and  P  is  the  property  of  being  less  than  1000,  then  (6)  is 
true. 

(3)  If  (f)  (ft)  =  n,  and  P  is  the  property  of  being  odd,  then  (c)  is  true.  For 
<f)  (n)  is  odd  if  n  is  odd  and  even  if  n  is  even,  and  both  the  odd  and  the  even 
values  of  ft  form  an  infinite  class. 

Example.  Consider,  in  each  of  the  following  cases,  whether  (a),  (b),  or 
(c)  is  true : 

(i)        <p  (n)  =  n,  P  being  the  property  of  being  a  perfect  square, 

(ii)       4>(n)=Pny  where  pn  denotes  the  nth  prime  number,  P  being  the 

property  of  being  odd, 
(iii)      <f)  (n)=pu,  P  being  the  property  of  being  even, 
(iv)      $  (n)=pa,  P being  the  property  cf>  (n)>n, 
(v)       <£  (n)  =  1  -  (-  l)n  (I In),  P  being  the  property  <£  («)<1 , 
(vi)      <f>(n)  =  l-(-  l)n  (Jin),  P  being  the  property  0  (n)<2, 
( vii)     0  (n)  =  1000  {1  +  ( -  l)»}/n,  P  being  the  property  <j>(n)<  1, 
(viii)    (f>  («)  =  1/ft,  P  being  the  property  <£  («)<  -001, 
(ix)      $  (»)  =  ( -  l)n/n,  P  being  the  property  |  (p  (ft)  |  <  -001, 
(x)       4>(n)  =  10000 In,  or  (-l)»10000/ft,  P  being  either  of  the  properties 

4>(n)< -001  or  |  0  (ft)  |  < -001, 
(xi)      (f>  (n)  =  (n-  l)/(ft  + 1),  P  being  the  property  1-0  (?«■)< -0001. 

54.  Let  us  now  suppose  that  <f>  (n)  and  P  are  such  that  the 
assertion  (a)  is  true,  i.e.  that  <f>  (?i)  has  the  property  P,  if  not  for 
all  values  of  n,  at  any  rate  for  all  values  of  n  except  a  finite 
number  N  of  such  values.  We  may  denote  these  exceptional 
values  by 

nu  n.2,  ...,  nN. 


53,  54]  POSITIVE   INTEGRAL   VARIABLE  111 

There  is  of  course  no  reason  why  these  JV  values  should  be  the 
first  N  values  1,  2,  ...,  N,  though,  as  the  preceding  examples 
show,  this  is  frequently  the  case  in  practice.  But  whether  this 
is  so  or  not  we  know  that  <f>  (n)  has  the  property  P  if  n  >  nN. 
Thus  the  ??th  prime  is  odd  if  n  >  2,  n  =  2  being  the  only  exception 
to  the  statement;  and  l/n<  '001  if  n  >  1000,  the  first  1000  values 
of  n  being  the  exceptions ;  and 

1000  {1  +  (-  l)»}/n  <  1 

if  ra >  2000,  the  exceptional  values  being  2,  4,  6,  ...,  2000.  That 
is  to  say,  in  each  of  these  cases  the  property  is  possessed  for  all 
values  of  n  from  a  definite  value  onwards. 

We  shall  frequently  express  this  by  saying  that  <£  (?z)  has  the 
property  for  large,  or  very  large,  or  all  sufficiently  large  values  of  n. 
Thus  when  we  say  that  <£  (n)  has  the  property  P  (which  will  as  a 
rule  be  a  property  expressed  by  some  relation  of  inequality)  for 
large  values  of  n,  what  we  mean  is  that  we  can  determine  some 
definite  number,  n0  say,  such  that  cf>  (n)  has  the  property  for  all 
values  of  n  greater  than  or  equal  to  n0.  This  number  n0,  in  the 
examples  considered  above,  may  be  taken  to  be  any  number 
greater  than  nN,  the  greatest  of  the  exceptional  numbers:  it  is 
most  natural  to  take  it  to  be  nN+l. 

Thus  we  may  say  that  'all  large  primes  are  odd',  or  that  '1/n  is 
less  than  "001  for  large  values  of  n  '.  And  the  reader  must  make 
himself  familiar  with  the  use  of  the  word  large  in  statements  of 
this  kind.  Large  is  in  fact  a  word  which,  standing  by  itself,  has 
no  more  absolute  meaning  in  mathematics  than  in  the  language 
of  common  life.  It  is  a  truism  that  in  common  life  a  number 
which  is  large  in  one  connection  is  small  in  another ;  6  goals  is  a 
large  score  in  a  football  match,  but  6  runs  is  not  a  large  score  in  a 
cricket  match;  and  400  runs  is  a  large  score,  but  £400  is  not 
a  large  income :  and  so  of  course  in  mathematics  large  generally 
means  large  enough,  and  what  is  large  enough  for  one  purpose 
may  not  be  large  enough  for  another. 

We  know  now  what  is  meant  by  the  assertion  '  <£  (n)  has  the 
property  P  for  large  values  of  n '.  It  is  with  assertions  of  this 
kind  that  we  shall  be  concerned  throughout  this  chapter. 


112  LIMITS   OF   FUNCTIONS   OF   A  [iV 

55.  The  phrase  'n  tends  to  infinity'.  There  is  a  some- 
what different  way  of  looking  at  the  matter  which  it  is  natural  to 
adopt.  Suppose  that  n  assumes  successively  the  values  1,  2,  3, .... 
The  word  'successively'  naturally  suggests  succession  in  time,  and 
we  may  suppose  n,  if  we  like,  to  assume  these  values  at  successive 
moments  of  time  (e.g.  at  the  beginnings  of  successive  seconds). 
Then  as  the  seconds  pass  n  gets  larger  and  larger  and  there  is 
no  limit  to  the  extent  of  its  increase.  However  large  a  number 
we  may  think  of  (e.g.  2147483647),  a  time  will  come  when  n  has 
become  larger  than  this  number. 

It  is  convenient  to  have  a  short  phrase  to  express  this  unending 
growth  of  11,  and  we  shall  say  that  n  tends  to  infinity,  or  n  **»oo  , 
this  last  symbol  being  usually  employed  as  an  abbreviation  for 
'infinity'.  The  phrase  'tends  to'  like  the  word  'successively' 
naturally  suggests  the  idea  of  change  in  time,  and  it  is  convenient 
to  think  of  the  variation  of  n  as  accomplished  in  time  in  the 
manner  described  above.  This  however  is  a  mere  matter  of  con- 
venience. The  variable  n  is  a  purely  logical  entity  which  has  in 
itself  nothing  to  do  with  time. 

The  reader  cannot  too  strongly  impress  upon  himself  that 
when  we  say  that  n  '  tends  to  oo '  we  mean  simply  that  n  is 
supposed  to  assume  a  series  of  values  which  increase  continually 
and  without  limit,  There  is  no  number  ' infinity':  such  an 
equation  as 

n=  oo 

is  as  it  stands  absolutely  meaningless :  n  cannot  be  equal  to  oo , 
because  "  equal  to  oo  '  means  nothing.  So  far  in  fact  the  symbol 
oo  means  nothing  at  all  except  in  the  one  phrase  '  tends  to  co ', 
the  meaning  of  which  we  have  explained  above.  Later  on  we 
shall  learn  how  to  attach  a  meaning  to  other  phrases  involving 
the  symbol  oo ,  but  the  reader  will  always  have  to  bear  in  mind 

(1)  that  oo  by  itself  means  nothing,  although  phrases  con- 
taining it  sometimes  mean  something, 

(2)  that  in  every  case  in  which  a  phrase  containing  the 
symbol  oo  means  something  it  will  do  so  simply  because  we  have 
previously  attached  a  meaning  to  this  particular  phrase  by  means 
of  a  special  definition. 


55,  56]  POSITIVE    INTEGRAL    VARIABLE  113 

Now  it  is  clear  that  if  <£  (n)  has  the  property  P  for  large  values 
of  n,  and  if  n  '  tends  to  oo ',  in  the  sense  which  we  have  just 
explained,  then  n  will  ultimately  assume  values  large  enough  to 
ensure  that  <f>(n)  has  the  property  P.  And  so  another  way  of 
putting  the  question  '  what  properties  has  <£  (n)  for  sufficiently 
large  values  of  n  ? '  is  '  how  does  <f>  (n)  behave  as  n  tends  to  oo  ? ' 

56.  The  behaviour  of  a  function  of  n  as  n  tends  to 
infinity.  We  shall  now  proceed,  in  the  light  of  the  remarks 
made  in  the  preceding  sections,  to  consider  the  meaning  of  some 
kinds  of  statements  which  are  perpetually  occurring  in  higher 
mathematics.  Let  us  consider,  for  example,  the  two  following 
statements  :  (a)  1/n  is  small  for  large  values  of  n,  (b)  1  —  (1/n)  is 
nearly  equal  to  1  for  large  values  of  n.  Obvious  as  they  may 
seem,  there  is  a  good  deal  in  them  which  will  repay  the  reader's 
attention.     Let  us  take  (a)  first,  as  being  slightly  the  simpler. 

We  have  already  considered  the  statement  '  1/n  is  less  than  "01 
for  large  values  of  n\  This,  we  saw,  means  that  the  inequality 
1/n  < '01  is  true  for  all  values  of  n  greater  than  some  definite 
value,  in  fact  greater  than  100.  Similarly  it  is  true  that  '  1/n  is 
less  than  '0001  for  large  values  of  n' :  in  fact  1/n <  "0001  if 
n  >  10000.  And  instead  of  '01  or  -0001  we  might  take  -000001  or 
•00000001,  or  indeed  any  positive  number  we  like. 

It  is  obviously  convenient  to  have  some  way  of  expressing  the 
fact  that  any  such  statement  as  '  1/n  is  less  than  "01  for  large 
values  of  n'  is  true,  when  we  substitute  for  "01  any  smaller 
number,  such  as  "0001  or  "000001  or  any  other  number  we  care 
to  choose.  And  clearly  we  can  do  this  by  saying  that  '  however 
small  8  may  be  (provided  of  course  it  is  positive),  then  l/n<8  for 
sufficiently  large  values  of  n '.  That  this  is  true  is  obvious.  For 
l/n<  S  if  n>  1/8,  so  that  our  'sufficiently  large'  values  of  n  need 
only  all  be  greater  than  1/8.  The  assertion  is  however  a  complex 
one,  in  that  it  really  stands  for  the  whole  class  of  assertions  which 
we  obtain  by  giving  to  8  special  values  such  as  '01.  And  of  course 
the  smaller  8  is,  and  the  larger  1/8,  the  larger  must  be  the  least  of 
the  '  sufficiently  large '  values  of  n :  values  which  are  sufficiently 
large  when  8  has  one  value  are  inadequate  when  it  has  a  smaller. 

The  last  statement  italicised  is  what  is  really  meant  by  the 
statement    (a),   that    1/n   is    small    when    n  is    large.     Similarly 
h.  '  8 


114  LIMITS   OF   FUNCTIONS   OF   A  [iV 

(b)  really  means  "if  <\>{n)  =  !  —  {!/ n),  then  the  statement  ll  —  <f>(n)<8 
for  sufficiently  large  values  of  n'  is  true  whatever  positive  value 
{such  as  '01  or  "0001)  we  attribute  to  8  ".  That  the  statement  (b) 
is  true  is  obvious  from  the  fact  that  1  —  <}>(n)=  1/n. 

There  is  another  way  in  which  it  is  common  to  state  the  facts 
expressed  by  the  assertions  (a)  and  (6).  This  is  suggested  at  once 
by  §  55.  Instead  of  saying  '  1/n  is  small  for  large  values  of  n '  we 
say  '  1/n  tends  to  0  as  n  tends  to  oo \  Similarly  we  say  that 
'  1  —  (1/n)  tends  to  1  as  n  tends  to  oo  ' :  and  these  statements  are 
to  be  regarded  as  precisely  equivalent  to  (a)  and  (b).     Thus  the 

statements 

'  1/n  is  small  when  n  is  large ', 

'  1/n  tends  to  0  as  n  tends  to  oo  ', 

are  equivalent  to  one  another  and  to  the  more  formal  statement 

'if  8  is  any  positive  number,  however  small,  then  1/n <  8 
for  sufficiently  large  values  of  n ', 

or  to  the  still  more  formal  statement 

'  if  8  is  any  positive  number,  however  small,  then  we  can 
find  a  number  n0  such  that  1/n  <  8  for  all  values  of  n  greater 
than  or  equal  to  n0\ 

The  number  n0  which  occurs  in  the  last  statement  is  of  course 
a  function  of  8.  We  shall  sometimes  emphasize  this  fact  by 
writing  n0  in  the  form  n0  (8). 

The  reader  should  imagine  himself  confronted  by  an  opponent  who 
questions  the  truth  of  the  statement.  He  would  name  a  series  of  numbers 
growing  smaller  and  smaller.  He  might  begin  with  -001.  The  reader  would 
reply  that  l/?i<'001  as  soon  as  %>1000.  The  opponent  would  be  bound  to 
admit  this,  but  would  try  again  with  some  smaller  number,  such  as  •0000001 . 
The  reader  would  reply  that  l/n< -0000001  as  soon  as  n>  10000000:  and  so 
on.  In  this  simple  case  it  is  evident  that  the  reader  would  always  have  the 
better  of  the  argument. 

We  shall  now  introduce  yet  another  way  of  expressing  this 
property  of  the  function  1/n.  We  shall  say  that  '  the  limit  of  1/n 
as  n  tends  to  oo  is  0 ',  a  statement  which  we  may  express  symboli- 
cally in  the  form 

lim    -  =  0, 


56,  57]  POSITIVE    INTEGRAL    VARIABLE  115 

or  simply  lim  (1/n)  =  0.    We  shall  also  sometimes  write 

as  n  -*  oo  ',  which  may  be  read  '  1/n  tends  to  0  as  n  tends  to  oo  ' ;  or 
simply  '  1/n  -*  0  '.     In  the  same  way  we  shall  write 

lim    fl--)  =  l,     limfl-.-W 
n^oc  \       nj  \       nj 

or  1  -  (1/n)  -^  1. 

57.  Now  let  us  consider  a  different  example :  let  <f>  (n)  =  n2. 
Then  '  n2  is  large  when  n  is  large '.  This  statement  is  equivalent 
to  the  more  formal  statements 

'  if  A  is  any  positive  number,  however  large,  then  n2  >  A 
for  sufficiently  large  values  of  n  \ 

*  we  can  find  a  number  n0  (A)  such  that  n2  >  A  for  all  values 
of  n  greater  than  or  equal  to  n0  (A) '. 

And  it  is  natural  in  this  case  to  say  that  '  n2  tends  to  oo  as  n 
tends  to  oo ',  or  '  n2  tends  to  oo  with  n ',  and  to  write 

n-  -»-  oo . 

Finally  consider  the  function  <£(n)  =  —  n2.  In  this  case  <f>(ri) 
is  large,  but  negative,  when  n  is  large,  and  we  naturally  say  that 
'  —  n2  tends  to  —  oo  as  n  tends  to  oo  '  and  write 

—  n2  -*■  —  oo  . 

And  the  use  of  the  symbol  —  oo  in  this  sense  suggests  that  it 
will  sometimes  be  convenient  to  write  n2  -*■  +  oo  for  n2  -*■  oo  and 
generally  to  use  +  oo  instead  of  oo ,  in  order  to  secure  greater 
uniformity  of  notation. 

But  we  must  once  more  repeat  that  in  all  these  statements 
the  symbols  oo  ,  +  oo  ,  —  oo  mean  nothing  whatever  by  themselves, 
and  only  acquire  a  meaning  when  they  occur  in  certain  special 
connections  in  virtue  of  the  explanations  which    we  have  just 

given. 

8—2 


116  LIMITS    OF  FUNCTIONS   OF   A  [iV 

58.  Definition  of  a  limit.  After  the  discussion  which 
precedes  the  reader  should  be  in  a  position  to  appreciate  the 
general  notion  of  a  limit.  Roughly  we  may  say  that  (f>  (n)  tends 
to  a  limit  I  as  n  tends  to  co  if  0  (n)  is  nearly  equal  to  I  when  n  is 
large.  But  although  the  meaning  of  this  statement  should  be 
clear  enough  after  the  preceding  explanations,  it  is  not,  as  it 
stands,  precise  enough  to  serve  as  a  strict  mathematical  definition. 
It  is,  in  fact,  equivalent  to  a  whole  class  of  statements  of  the 
type  'for  sufficiently  large  values  of  n,  <f>(n)  differs  from  I  by  less 
than  8 '.  This  statement  has  to  be  true  for  8  =  *01  or  "0001  or  any 
positive  number ;  and  for  any  such  value  of  8  it  has  to  be  true  for 
any  value  of  n  after  a  certain  definite  value  n0(8),  though  the 
smaller  8  is  the  larger,  as  a  rule,  will  be  this  value  n0  (8). 

We  accordingly  frame  the  following  formal  definition : 

Definition  I.  The  function  cf>  (n)  is  said  to  tend  to  the  limit 
I  as  n  tends  to  oo ,  if,  however  small  be  the  positive  number  8, 
0  (n)  differs  from  I  by  less  than  8  for  sufficiently  large  values  of  n; 
that  is  to  say  if,  however  small  be  the  positive  number  8,  ive  can 
determine  a  number  n0(8)  corresponding  to  8,  such  that  <f>(n)  differs 
from  I  by  less  than  8  for  all  values  of  n  greater  than  or  equal  to  n0  (8). 

It  is  usual  to  denote  the  difference  between  </>(«)  and  I,  taken' 
positively,  by  |  </>  (n)  —  I  |.  It  is  equal  to  <f> (n)  —  I  or  to  I  —  <f> (n), 
whichever  is  positive,  and  agrees  with  the  definition  of  the 
modulus  of  </>  (n)  —  I,  as  given  in  Chap.  Ill,  though  at  present 
we  are  only  considering  real  values,  positive  or  negative. 

With  this  notation  the  definition  may  be  stated  more  shortly 
as  follows:  'if,  given  any  positive  number,  8,  however  small,  we 
can  find  n0  (8)  so  that  |  <£  (n)  —  1 |  <  8  when  n~  n0  (8),  then  we  say 
that  <f>  (n)  tends  to  the  limit  I  as  n  tends  to  oo ,  and  write 

lim  </>  (n)  =  I '. 

Sometimes  we  may  omit  the  ln-*-ao ' ;  and  sometimes  it  is  convenient,  for 
brevity,  to  write  <fi  (n)-*-l. 

The  reader  will  find  it  instructive  to  work  out,  in  a  few  simple  cases,  the 
explicit  expression  of  n0  as  a  function  of  8.  Thus  if  (f>  (x)  =  ljn  then  £=0,  and 
the  condition  reduces  to  \jn<8  for  n>n0,  which  is  satisfied  if  %0=1  +  [1/5J*. 
There  is  one  and  only  one  case  in  which  the  same  7i0  will  do  for  all  values  of  8. 

*  Here  and  henceforward  we  shall  use  [x]  in  the  sense  of  Chap.  II,  i.e.  as  the 
greatest  integer  not  greater  than  x. 


58-60] 


POSITIVE   INTEGRAL   VARIABLE 


117 


If,  from  a  certain  value  N  of  n  onwards,  $  (n)  is  constant,  say  equal  to  C,  then 
it  is  evident  that  (f>  (n)  —  C=0  for  n  giV,  so  that  the  inequality  |  <£  (n)  —  C  \  <8 
is  satisfied  for  n'Z.N  and  all  positive  values  of  8.  And  if  |  <f>(n)  —  l\  <8  for 
n^N  and  all  positive  values  of  8,  then  it  is  evident  that  $  (n)=l  when  n  >X, 
so  that  (p  (n)  is  constant  for  all  such  values  of  n. 

59.  The  definition  of  a  limit  may  be  illustrated  geometrically 
as  follows.  The  graph  of  <f>  (n)  consists  of  a  number  of  points 
corresponding  to  the  values  n  =  l,  2,  3,  .... 

Draw  the  line  y  —  I,  and  the  parallel  lines  y  =  1  —  8,  y  =  1  +  8 
at  distance  8  from  it.     Then 

lim  (f)  (n)  =  I, 


Fig.  27 


if,  when  once  these  lines  have  been  drawn,  no  matter  how  close 
they  may  be  together,  we  can  always  draw  a  line  x  =  na,  as  in  the 
figure,  in  such  a  way  that  the  point  of  the  graph  on  this  line,  and 
all  points  to  the  right  of  it,  lie  between  them.  We  shall  find 
this  geometrical  way  of  looking  at  our  definition  particularly 
useful  when  we  come  to  deal  with  functions  defined  for  all  values 
of  a  real  variable  and  not  merely  for  positive  integral  values. 

60.  So  much  for  functions  of  n  which  tend  to  a  limit  as  n 
tends  to  oo  .  We  must  now  frame  corresponding  definitions  for 
functions  which,  like  the  functions  if-  or  —  n-,  tend  to  positive  or 
negative  infinity.  The  reader  should  by  now  find  no  difficulty  in 
appreciating  the  point  of 

Definition  II.  The  function  <f>(n)  is  said  to  tend  to  +cc 
{positive  infinity)  ivith  n,  if,  when  any  number  A,  however  large, 
is  assigned,  we  can  determine  n0  (A)  so  that  <f>  (n)  >  A  ivhen  n  =  n0  (A); 


118  LIMITS   OF   FUNCTIONS   OF   A  [iV 

that  is  to  say  if,  however  large  A  may  be,  </>  (n)  >  A  for  sufficiently 
large  values  of  n. 

Another,  less  precise,  form  of  statement  is  '  if  we  can  make 
<£  (n)  as  large  as  we  please  by  sufficiently  increasing  n '.  This  is 
open  to  the  objection  that  it  obscures  a  fundamental  point,  viz. 
that  $  (n)  must  be  greater  than  A  for  all  values  of  n  such  that 
n  =  n0  (A),  and  not  merely  for  some  such  values.  But  there  is  no 
harm  in  using  this  form  of  expression  if  we  are  clear  what  it 
means. 

When  $  (n)  tends  to  +  oo  we  write 

<£  (n)  -*■  +  oo  . 

We    may   leave   it   to   the   reader   to   frame   the   corresponding 
definition  for  functions  which  tend  to  negative  infinity. 

61.  Some  points  concerning  the  definitions.  The  reader 
should  be  careful  to  observe  the  following  points. 

(1)  We  may  obviously  alter  the  values  of  <f)(n)  for  any 
finite  number  of  values  of  n,  in  any  way  we  please,  without  in 
the  least  affecting  the  behaviour  of  <£  (n)  as  n  tends  to  oo .  For 
example  1/n  tends  to  0  as  n  tends  to  oo .  We  may  deduce  any 
number  of  new  functions  from  1/n  by  altering  a  finite  number  of 
its  values.  For  instance  we  may  consider  the  function  <f>  (n)  which 
is  equal  to  3  for  n  =  \,  2,  7,  11,  101,  107,  109,  237  and  equal  to 
1/n  for  all  other  values  of  n.  For  this  function,  just  as  for  the 
original  function  1/n,  lim  </>  (n)  =  0.  Similarly,  for  the  function 
cf>  (n)  which  is  equal  to  3  if  ft  =  1,  2,  7,  11,  101,  107,  109,  237,  and 
to  n2  otherwise,  it  is  true  that  <f>  (ft)  -*■  +  oo  . 

(2)  On  the  other  hand  we  cannot  as  a  rule  alter  an  infinite 
number  of  the  values  of  </>  (n)  without  affecting  fundamentally  its 
behaviour  as  n  tends  to  oo  .  If  for  example  we  altered  the  function 
1/n  by  changing  its  value  to  1  whenever  n  is  a  multiple  of  100, 
it  would  no  longer  be  true  that  lim  <£  (n)  =  0.  So  long  as  a  finite 
number  of  values  only  were  affected  we  could  always  choose  the 
number  n0  of  the  definition  so  as  to  be  greater  than  the  greatest 
of  the  values  of  n  for  which  <£  (ft)  was  altered.  In  the  examples 
above,  for  instance,  we  could  always  take  n0  >  237,  and  indeed  we 
should  be  compelled  to  do  so  as  soon  as  our  imaginary  opponent 


60,  61]  POSITIVE   INTEGRAL   VARIABLE  119 

of  §  56  had  assigned  a  value  of  8  as  small  as  3  (in  the  first 
example)  or  a  value  of  A  as  great  as  3  (in  the  second).  But 
now  however  large  n0  may  be  there  will  be  greater  values  of  n  for 
which  <f)  (n)  has  been  altered. 

(3)  In  applying  the  test  of  Definition  I  it  is  of  course 
absolutely  essential  that  we  should  have  |  <p(n)  —  1 1  <  8  not  merely 
when  n  =n0  but  when  n  ^  n0>  i.e.  for  n0  and  for  all  larger  values 
of  n.  It  is  obvious,  for  example,  that,  if  <f>  (n)  is  the  function  last 
considered,  then  given  8  we  can  choose  n0  so  that  |  (f>(n)  |  <  8  when 
n  =  n0 :  we  have  only  to  choose  a  sufficiently  large  value  of  n 
which  is  not  a  multiple  of  100.  But,  when  n0  is  thus  chosen,  it 
is  not  true  that  |  <f>  (n)  \  <  8  when  n^n0:  all  the  multiples  of  100 
which  are  greater  than  n0  are  exceptions  to  this  statement. 

(4)  If  (f>(n)  is  always  greater  than  I,  we  can  replace 
|  $  (n)  -l\  by  $(n)  —  l.  Thus  the  test  whether  1/n  tends  to  the 
limit  0  as  n  tends  to  oo  is  simply  whether  1/n  <  8  when  n  ^  n0. 
If  however  </>  («)  =  (—  \)njn,  then  I  is  again  0,  but  <£  (n)  —  I  is  some- 
times positive  and  sometimes  negative.  In  such  a  case  we  must 
state  the  condition  in  the  form  |  <f>  (n)  —  1 1  <  8,  for  example,  in 
this  particular  case,  in  the  form  |  <j>  (n)  |  <  8. 

(5)  The  limit  I  may  itself  be  one  of  the  actual  values  of 
(f>(n).  Thus  if  </>(w)  =  0  for  all  values  of  n,  it  is  obvious  that 
lim  </>  (n)  =  0.  Again,  if  we  had,  in  (2)  and  (3)  above,  altered 
the  value  of  the  function,  when  n  is  a  multiple  of  100,  to  0 
instead  of  to  1,  we  should  have  obtained  a  function  $  (n)  which 
is  equal  to  0  when  n  is  a  multiple  of  100  and  to  1/n  otherwise. 
The  limit  of  this  function  as  n  tends  to  cc  is  still  obviously  zero. 
This  limit  is  itself  the  value  of  the  function  for  an  infinite  number 
of  values  of  n,  viz.  all  multiples  of  100. 

On  the  other  hand  the  limit  itself  need  not  {and  in  general  will 
not)  be  the  value  of  the  function  for  any  value  of  n.  This  is 
sufficiently  obvious  in  the  case  of  <f>  (n)  =  1/n.  The  limit  is  zero ; 
but  the  function  is  never  equal  to  zero  for  any  value  of  n. 

The  reader  cannot  impress  these  facts  too  strongly  on  his 
mind.  A  limit  is  not  a  value  of  the  function  :  it  is  something 
quite  distinct  from  these  values,  though  it  is  defined  by  its  relations 


120  LIMITS   OF   FUNCTIONS   OF   A  [iV 

to  them  and  may  possibly  be  equal  to  some  of  them.     For  the 

functions 

<f>(n)  =  0,  1, 

the  limit  is  equal  to  all  the  values  of  </>  (n) :  for 

0(n)-l/n>    i-lTIn,    1  +  (1/n),    l  +  {(-l)"/ra) 

it  is  not  equal  to  any  value  of  <f>  (n)  :  for 

<£  (n)  =  (sin  hi7r)/n,     1  +  {(sin  £w7t)/m} 

(whose  limits  as  n  tends  to  oo  are  easily  seen  to  be  0  and  1,  since 
sin  \mr  is  never  numerically  greater  than  1)  the  limit  is  equal  to 
the  value  which  cf>  (n)  assumes  for  all  even  values  of  n,  but  the 
values  assumed  for  odd  values  of  n  are  all  different  from  the  limit 
and  from  one  another. 

(6)  A  function  may  be  alwaj'S  numerically  very  large  when 
n  is  very  large  without  tending  either  to  +  oo  or  to  —  oo  .  A 
sufficient  illustration  of  this  is  given  by  <£  (n)  =  (—  l)n  n.  A  function 
can  only  tend  to  +  go  or  to  —  oo  if,  after  a  certain  value  of  n, 
it  maintains  a  constant  sign. 

Examples  XXIII.  Consider  the  behaviour  of  the  following  functions 
of  x  as  n  tends  to  oo  : 

1.  <f)  (n)=nk,  where  k  is  a  positive  or  negative  integer  or  rational  fraction. 
If  k  is  positive,  then  nk  tends  to  +  oo  with  n.  If  k  is  negative,  then  lim  nk=0. 
If  £  =  0,  then  nk=l  for  all  values  of  n.     Hence  lim»*=l. 

The  reader  will  find  it  instructive,  even  in  so  simple  a  case  as  this,  to 
write  down  a  formal  proof  that  the  conditions  of  our  definitions  are  satisfied. 
Take  for  instance  the  case  of  k>0.  Let  A  be  any  assigned  number,  however 
large.  We  wish  to  choose  7i0  so  that  »fc>A  when  n2:n0.  We  have  in  fact  only 
to  take  for  n0  any  number  greater  than  ^/A.  If  e.g.  k  =  4,  then  ni>  10900  when 
W>11,  %4>100000000  when  »>101,  and  so  on. 

2.  <f)(n)=p„,  where  £>,t  is  the  nth.  prime  number.  If  there  were  only 
a  finite  number  of  primes  then  cj)  (n)  would  be  defined  only  for  a  finite  number 
of  values  of  n.  There  are  however,  as  was  first  shown  by  Euclid,  infinitely 
many  primes.  Euclid's  proof  is  as  follows.  If  there  are  only  a  finite 
n-umber  of  primes,  let  them  be  1,  2,  3,  5,  7,  11,  ...  N.  Consider  the  number 
1  + (1.2. 3. 5. 7. 11  ...  JY).  This  number  is  evidently  not  divisible  by 
any  of  2,  3,  5,  ...  JV,  since  the  remainder  when  it  is  divided  by  any  of 
these  numbers  is  1.  It  is  therefore  not  divisible  by  any  prime  save  1,  and 
is  therefore  itself  prime,  which  is  contrary  to  our  hypothesis. 

It  is  moreover  obvious  that  <p  {n)>n  for  all  values  of  n  (save  n  =  \,  2,  3). 
Hence  <\>  (n)  -*•  +  oo . 


61,  62]  POSITIVE    INTEGRAL   VARIABLE  121 

3.  Let  0  (ft)  be  the  number  of  primes  less  than?*.    Here  again  0  (n)-*  +  oo . 

4.  0  (n)  =  [an],  where  a  is  any  positive  number.     Here 

0(«)=O    (0<M<l/a),     0(%)  =  ]    (l/o<n<2/o), 
and  so  on ;  and  0  («)-»- +  oo  . 

5.  If  0  (n)  =  1000000/ft,  then  lim  0  (n)  =  0  :  and  if  ^  (n)  =  ft/1000000,  then 
^(»)-»4oo.  These  conclusions  are  in  no  way  affected  by  the  fact  that  at  first 
0  (ft)  is  much  larger  than  yjr  (ft),  being  in  fact  larger  until  n  =  1000000. 

6.  0  (ft)  =  l/{ft  -  ( - 1)»},  ft  -  ( - 1)«,  ft  {1  -  ( - 1)"}.  The  first  function  tends 
to  0,  the  second  to  +  °o  ,  the  third  does  not  tend  either  to  a  limit  or  to  -f  oo  . 

7.  0  (ft)  =  (sin  ndn)jn,  where  6  is  any  real  number.  Here  [0(n)|<l/?i, 
since  |  sin  ndn  [  ^  1,  and  lim  0  (ft)  =  0. 

8.  0  (ft)  =  (sin  ndn)jsln,  (a  cos2  nd  +  b  sin2  nd)/n,  where  a  and  b  are  any  real 
numbers. 

9.  0(ft)=sinft#7r.  If  6  is  integral  then  0(«)  =  O  for  all  values  of  ft,  and 
therefore  lim  0  (ft) =0. 

Next  let  6  be  rational,  e.g.  8=p/q,  where  p  and  q  are  positive  integers. 
Let  n  =  aq  +  b  where  a  is  the  quotient  and  b  the  remainder  when  n  is  divided 
by  q.  Then  sin  (npnjq)  =  ( -  l)ap  sin  (bpnlq).  Suppose,  for  example,  p  even  ; 
then,  as  n  increases  from  0  to  q—  1,  0  (ft)  takes  the  values 

0,  sin  (pirfq),  sin  (2pir/q),  ...  sin  {(q-1) pn/q}. 
When  n  increases  from  q  to  2q-l  these  values  are  repeated  ;  and  so  also 
as  n  goes  from  2q  to  3^-1,  3q  to  4q  —  1,  and  so  on.  Thus  the  values  of  0  (ft) 
form  a  perpetual  cyclic  repetition  of  a  finite  series  of  different  values.  It  is 
evident  that  when  this  is  the  case  0  (ft)  cannot  tend  to  a  limit,  nor  to  +  co , 
nor  to  —  oo ,  as  ft  tends  to  infinity. 

The  case  in  which  6  is  irrational  is  a  little  more  difficult.  It  is  discussed 
in  the  next  set  of  examples. 

62.  Oscillating  Functions.  Definition.  When  (f>  (n)  does 
not  tend  to  a  limit,  nor  to  +  oo ,  nor  to  —  oo ,  as  n  tends  to  oo ,  ive 
say  that  <j>(n)  oscillates  as  n  tends  to  cc  . 

A  function  <f)(n)  certainly  oscillates  if  its  values  form,  as 
in  the  case  considered  in  the  last  example  above,  a  continual 
repetition  of  a  cycle  of  values.  But  of  course  it  may  oscillate 
without  possessing  this  peculiarity.  Oscillation  is  defined  in  a 
purely  negative  manner :  a  function  oscillates  when  it  does  not  do 
certain  othei  things. 


122  LIMITS   OF   FUNCTIONS   OF   A  [IV 

The  simplest  example  of  an  oscillatory  function  is  given  by 

which  is  equal  to  + 1  when  n  is  even  and  to  —  1  when  n  is  odd. 
In  this  case  the  values  recur  cyclically.     But  consider 

4>00  =  (-i)n+(iM 

the  values  of  which  are 

-1  +  1,     l  +  (l/2),     -l+(l/3)5     l  +  (l/4),     -l  +  (l/5),  .... 

When  n  is  large  every  value  is  nearly  equal  to  +1  or  —  1,  and 
obviously  cf>  (n)  does  not  tend  to  a  limit  or  to  +  oo  or  to  —  oo  ,  and 
therefore  it  oscillates :  but  the  values  do  not  recur.  It  is  to  be 
observed  that  in  this  case  every  value  of  <f>  (n)  is  numerically  less 
than  or  equal  to  3/2.     Similarly 

cf)  (n)  =  (-iynoo  + (1000 /n) 

oscillates.  When  n  is  large,  every  value  is  nearly  equal  to  100 
or  to  —100.  The  numerically  greatest  value  is  900  (for  n  =  l). 
But  now  consider  <£(?i)  =  (—  l)nn,  the  values  of  which  are  —  1,  2, 
—  3,  4,  —5,  ....  This  function  oscillates,  for  it  does  not  tend  to  a 
limit,  nor  to  +  oo  ,  nor  to  —  oo  .  And  in  this  case  we  cannot  assign 
any  limit  beyond  Avhich  the  numerical  value  of  the  terms  does 
not  rise.  The  distinction  between  these  two  examples  suggests  a 
further  definition. 

Definition.  If  <f>  (n)  oscillates  as  n  tends  to  co ,  then  </>  (n)  will 
be  said  to  oscillate  finitely  or  infinitely  according  as  it  is  or  is  not 
possible  to  assign  a  number  K  such  that  all  the  values  of  <f>  (n)  are 
numerically  less  than  K,  i.e.  |  </>  (n)  \  <  K  for  all  values  of  n. 

These  definitions,  as  well  as  those  of  §§  58  and  GO,  are  further 
illustrated  in  the  following  examples. 

Examples  XXIV.  Consider  the  behaviour  as  n  tends  to  oo  of  the 
following  functions : 

1.  (-1)*,    5+3(-l)»    (1000000/?*)  + (-1)",    1000000  (-l)»  +  (l/»). 

2.  (-l)nn,    1000000  +  (-l)B». 

3.  1000000-%,    ( -1)»  (1000000 -n)- 

4.  n{l  +  (-  1)"}.     In  this  case  the  values  of  0 (n)  are 

0,    4,    0,    8,    0,    12,    0,    16,  .... 
The    odd   terras    are   all   zero    and    the  even   terms  tend   to    +  oo :    <fr  (n) 
oscillates  infinitely. 


62]  POSITIVE    INTEGRAL   VARIABLE  123 

5.  n2  +  (  —  l)ll2n.  The  second  term  oscillates  infinitely,  but  the  first  is 
very  much  larger  than  the  second  when  n  is  large.  In  fact  $  (n)  >n2  -  2n  and 
n2-2n  =  (n-l)2-  1  is  greater  than  any  assigned  value  A  if  n>l  +  J(A  +  l). 
Thus  cf)  (»)-*■  +  oo.  It  should  be  observed  that  in  this  case  <£(2£+l)  is 
always  less  than  0  (2k),  so  that  the  function  progresses  to  infinity  by  a  con- 
tinual series  of  steps  forwards  and  backwards.  It  does  not  however  'oscillate' 
according  to  our  definition  of  the  term. 

6.  n2{l  +  (-l)»},         (-l)nn2  +  n,         n3  +  (-l)nn2. 

7.  sin  ndn.  We  have  already  seen  (Exs.  xxm.  9)  that  <p(n)  oscillates 
finitely  when  d  is  rational,  unless  d  is  an  integer,  when  $  (n)  =  0,  0  (?i)-»-0. 

The  case  in  which  d  is  irrational  is  a  little  more  difficult.  But  it  is  not 
difficult  to  see  that  <f)(n)  still  oscillates  finitely.  We  can  without  loss  of 
generality  suppose  O<0<1.  In  the  first  place  |</>(w)|<l.  Hence  (\>(n) 
must  oscillate  finitely  or  tend  to  a  limit.  We  shall  consider  whether  the 
second  alternative  is  really  possible.     Let  us  suppose  that 

lim  sin  ndn  =  l. 
Then,  however  small  8  may  be,  we  can  choose  n0  so  that  sin  ndn  lies  between 
1-8  and   1+8  for  all  values   of   n  greater  than   or   equal   to  ?i0.     Hence 
sin  (n+l)  dn- sin  ndn  is  numerically  less  than  28  for  all  such  values  of  n, 
and  so  |  sin^n-  cos  (n  +  \)  dn  \<8. 

Hence        cos  (n+%)  dn  =  cosndn  cos  hdn  —  sin  ndn  sin|07r 
must  be  numerically  less  than  8/ 1  sin  h  dn  |.     Similarly 

cos  (n  -h)6n  =  cos  ndn  cos  \6n  +  sin  ndn  sin  \  6tt 
must  be  numerically  less  than  8j  \  sin  ^  6n  \ ;  and  so  each  of  cos  ndn  cos  1 6n, 
sin  n6n  sin  ^  6n  must  be  numerically  less  than  8j  |  sin  ^  On  |.  That  is  to  say, 
cos  ndn  cos  1 6n  is  very  small  if  n  is  large,  and  this  can  only  be  the  case 
if  cos  n6n  is  very  small.  Similarly  sin  ndn  must  be  very  small,  so  that  I 
must  be  zero.  But  it  is  impossible  that  cos  ndn  and  sin  ndn  can  both  be 
very  small,  as  the  sum  of  their  squares  is  unity.  Thus  the  hypothesis  that 
sin  ndn  tends  to  a  limit  I  is  impossible,  and  therefore  sin  ndn  oscillates 
as  n  tends  to  oo . 

The  reader  should  consider  with  particular  care  the  argument 
'  cos  ndn  cos  \  dn  is  very  small,  and  this  can  only  be  the  case  if  cos  ndn 
is  very  small'.  Why,  he  may  ask,  should  it  not  be  the  other  factor  cos^#7r 
which  is  'very  small'?  The  answer  is  to  be  found,  of  course,  in  the  meaning 
of  the  phrase  '  very  small '  as  used  in  this  connection.  When  we  say  '  $  (n) 
is  very  small'  for  large  values  of  n,  we  mean  that  we  can  choose  n0  so  that 
<f)  («.)  is  numerically  smaller  than  any  assigned  number,  if  n  is  sufficiently 
large.  Such  an  assertion  is  palpably  absurd  when  made  of  a  fixed  number 
such  as  cos  |  dn,  which  is  not  zero. 

Prove  similarly  that  cos  ndn  oscillates  finitely,  unless  d  is  an  even  integer. 

8.  sin  ndn  +  (I  I  n),     sinra&r+l,     siandn  +  n,     (-l)nsinndn. 

9.  a  cos  ndn +  b  sin  ndn,    sin2  ndn,    a  cos2  ndn +  b  sin2  ndn. 


124  LIMITS   OF   FUNCTIONS   OF   A  [iV 

10.  a  +  hi  +  ( -  l)n  (c  +  dn)  +e  cos  ndn  +/sin  ndir. 

11.  n  sin  ndir.  If  n  is  integral,  then  #  (w)=°>  $  ('0-*"0-  If  #  is  rational 
but  not  integral,  or  irrational,  then  cf>  (n)  oscillates  infinitely. 

12.  n  (a  cos2  n6 W  +  b  sin2  ndir).  In  this  case  (f>(n)  tends  to  +cc  if  a  and 
6  are  both  positive,  but  to  -  oo  if  both  are  negative.  Consider  the  special 
cases  in  which  a  =  0,  b>0,  or  a>0,  6=0,  or  a=0,  b  =  0.  If  a  and  6  have 
opposite  signs  cj>  (n)  generally  oscillates  infinitely.  Consider  any  excep- 
tional cases. 

13.  sin  (nWn).  If  6  is  integral,  then  0  (»)-*0.  Otherwise  0  (n)  oscillates 
finitely,  as  may  be  shown  by  arguments  similar  to  though  more  complex 
than  those  used  in  Exs.  xxm.  9  and  xxiv.  7*. 

14.  s\x\(n\6ir).  If  6  has  a  rational  value  pjq,  then  n\  6  is  certainly 
integral  for  all  values  of  n  greater  than  or  equal  to  q.  Hence  0  (n)  -»-0.  The 
case  in  which  6  is  irrational  cannot  be  dealt  with  without  the  aid  of  considera- 
tions of  a  much  more  difficult  character. 

15.  cos  (n  !  dn),  a  cos2  (n  !  6n)-\-b  sin2  («  !  6n),  where  6  is  rational. 

16.  an-[bn\  (-l)n  (an-[bn]).  17.     [sjn\  (-l)"[vH  >Jn-[shi]. 

18.  The  smallest  prune  factor  of  n.  When  n  is  a  prime,  0  (n)  =  n.  When 
n  is  even,  (f>  (n)  =  2.     Thus  0  (?&)  oscillates  infinitely. 

19.  The  largest  prime  factor  of  n. 

20.  The  number  of  days  in  the  year  n  a.d. 

Examples  XXV.  1.  If  </>  (»i) -*■  +  °°  and  >fr  (»)><£(«)  for  all  values  of 
n,  then  ^  (n)  -*-  +  oo  . 

2.  If  0  («)-*- 0,  and  |  ^  (?i)  |  ^  ]  0  («)  |  for  all  values  of  n,  then  yj/  («)-*■ 0. 

3.  If  lim  |  0  (?i)  |  =  0,  then  lim  <£  (n)  =  0. 

4.  If  ^  («)  tends  to  a  limit  or  oscillates  finitely,  and  |  \^  (n)  \^\<j>  (n)  |  when 
n£n0)  then  yj/(n)  tends  to  a  limit  or  oscillates  finitely. 

5.  If  (f>  (n)  tends  to  +00  ,  or  to  —  00 ,  or  oscillates  infinitely,  and 

|*(n)|>|*(»)| 
when  «Sn0,  then  ^  (n)  tends  to  +  00  or  to  -00  or  oscillates  infinitely. 

6.  'If  4>(n)  oscillates  and,  however  great  be  n0,  we  can  find  values  of  n 
greater  than  n0  for  which  >//■  (n)  >  <fi  (71),  and  values  of  n  greater  than  n0  for 
which  -ty (n)  <  0 (n),  then  ^(«)  oscillates'.  Is  this  true?  If  not  give  an 
example  to  the  contrary. 

7.  If  cf>  {n)-*-l  as  n-*-co ,  then  also  <£  (»+£>) -»■/,  £>  being  any  fixed  integer. 
[This  follows  at  once  from  the  definition.  Similarly  we  see  that  if  0  («)  tends 
to  +oo  or  —00  or  oscillates  so  also  does  <£(n +£>).] 

8.  The  same  conclusions  hold  (except  in  the  case  of  oscillation)  if  p  varies 
with  n  but  is  always  numerically  less  than  a  fixed  positive  integer  N  ;  or  if  p 
varies  with  n  in  any  way,  so  long  as  it  is  always  positive. 

*  See  Bromwich's  Infinite  Series,  p.  485. 


62,  63]  POSITIVE   INTEGRAL   VARIABLE  125 

9.  Determine  the  least  value  of  n0  for  which  it  is  true  that 

(a)    «2  +  2;j>999999    (n>n0),  (b)    n2  +  2n>  1000000    (»>»„). 

10.  Determine  the  least  value  of  n0  for  which  it  is  true  that 

(a)     «  +  (-l)'l>1000     («>%),         (b)     ?i  +  (-l)»>1000000     («>>i0). 

11.  Determine  the  least  value  of  ?i0  for  which  it  is  true  that 

(a)     ?i2+2>i>A     (n>n0),  (6)     ra+(-l)»>A     (»>w0), 

A  being  any  positive  number. 

[(«)  n0=[y/(A  +  l)]:  (6)  «0=1+[A]  or  2  +  [a],  according  as  [a]  is  odd  or 
even,  i.e.  rc0=l  +  [A]  +  i  {l  +  (-lp}.] 

12.  Determine  the  least  value  of  n0  such  that 

(a)    «/(^2  +  l)<-0001,        (b)    (l/%)  +  {(-l)»/%2}< -ooooi, 
when  «S;i0.     [Let  us  take  the  latter  case.     In  the  first  place 
(l/n)+{(-l)»M^(«+l)/«2 

and  it  is  easy  to  see  that  the  least  value  of  n0,  such  that  (n+l)/n2< -000001 
when  n  >«.0,  is  1000002.  But  the  inequality  given  is  satisfied  by  n=  1000001, 
and  this  is  the  value  of  n0  required.] 

63.  Some  general  theorems  with  regard  to  limits. 
A.  The  behaviour  of  the  sum  of  two  functions  whose 
behaviour   is   known. 

Theorem  I.  If  (f>(n)  and  yfr(n)  tend  to  limits  a,  b,  then 
<f>  (n)  +  y\r  (n)  tends  to  the  limit  a  +  b. 

This  is  almost  obvious*.     The  argument  which  the  reader  will 

*  There  is  a  certain  ambiguity  in  this  phrase  which  the  reader  will  do  well  to 
notice.  When  one  says  '  such  and  such  a  theorem  is  almost  obvious '  one  may 
mean  one  or  other  of  two  things.  One  may  mean  '  it  is  difficult  to  doubt  the  truth 
of  the  theorem',  '  the  theorem  is  such  as  common-sense  instinctively  accepts',  as 
it  accepts,  for  example,  the  truth  of  the  propositions  '2  +  2  =  4'  or  'the  base-angles 
of  an  isosceles  triangle  are  equal'.  That  a  theorem  is  'obvious'  in  this  sense  does 
not  prove  that  it  is  true,  since  the  most  confident  of  the  intuitive  judgments  of 
common  sense  are  often  found  to  be  mistaken;  and  even  if  the  theorem  is  true, 
the  fact  that  it  is  also  '  obvious '  is  no  reason  for  not  proving  it,  if  a  proof  can  be 
found.  The  object  of  mathematics  is  to  prove  that  certain  premises  imply  certain 
conclusions;  and  the  fact  that  the  conclusions  may  be  as  'obvious'  as  the  premises 
never  detracts  from  the  necessity,  and  often  not  even  from  the  interest  of  the  proof. 

But  sometimes  (as  for  example  here)  we  mean  by  '  this  is  almost  obvious  ' 
something  quite  different  from  this.  We  mean  '  a  moment's  reflection  should  not 
only  convince  the  reader  of  the  truth  of  what  is  stated,  but  should  also  suggest  to 
him  the  general  Hues  of  a  rigorous  proof.  And  often,  when  a  statement  is 
'  obvious  '  in  this  sense,  one  may  well  omit  the  proof,  not  because  the  proof  is  in 
any  sense  unnecessary,  but  because  it  is  a  waste  of  time  and  space  to  state  in  detail 
what  the  reader  can  easily  supply  for  himself. 


126  LIMITS   OF    FUNCTIONS   OF   A  [iV 

at  once  form  in  his  mind  is  roughly  this :  '  when  n  is  large,  <£  (w)  is 
nearly  equal  to  a  and  yjr  (n)  to  b,  and  therefore  their  sum  is  nearly 
equal  to  a  +  b '.  It  is  well  to  state  the  argument  quite  formally, 
however. 

Let  8  be  any  assigned  positive  number  (e.g.  "001,  0000001,  ...). 
We  require  to  show  that  a  number  n0  can  be  found  such  that 

\<f>(n)  +  ^(n)-a-b\<8 (1), 

when  )i  ^  «0,  Now  by  a  proposition  proved  in  Chap.  Ill  (more 
generally  indeed  than  we  need  here)  the  modulus  of  the  sum  of 
two  numbers  is  less  than  or  equal  to  the  sum  of  their  moduli. 
Thus 

I  <t>  (V)  +  yjr  (n)  -a-b\^\<f>  (n)  -a\  +  \yfr  (n)  -b\. 

It  follows  that  the  desired  condition  will  certainly  be  satisfied  if 
n0  can  be  so  chosen  that 

\(f)(n)-a\  +  \ylr(n)-b\  <  8  (2), 

when  n  =  n0.  But  this  is  certainly  the  case.  For  since  lim  <f>(n)  =  a 
we  can,  by  the  definition  of  a  limit,  find  nx  so  that  j  <£  (n)  —  a\  <8' 
when  n  =  nu  and  this  however  small  8'  may  be.  Nothing  prevents 
our  taking  8'  =  %8,  so  that  |  <f>  (n)  —  a  j  <  £S  when  n  S  nx.  Similarly 
we  can  find  n2  so  that  |  yfr  (n)  —  b\  <  ^8  when  n~n2.  Now  take  ??0 
to  be  the  greater  of  the  two  numbers  nu  n2.  Then  |  0  (n)  —  a\<^8 
and  \yfr(n)  —  b\  <^8  when  n  =n0,  and  therefore  (2)  is  satisfied  and 
the  theorem  is  proved. 

The   argument   may  be   concisely  stated   thus:   since  lim<f)(n)  =  a  and 
lim  ^r(n)  =  b,  we  can  choose  n^,  n2  so  that 

\4>(n)-a\<%8   (n>»h),    |^(»0-&|<i*  (n>n2); 
and  then,  if  n  is  not  less  than  either  iii  or  n2, 

I  <t>  (n)+^  («)  -  a  -  b  \  <\  $  (n)  -  a\  +  \(f>  (n)  -  b  \<8 ; 
and  therefore  lim  {$  (n)  +  y\r(n))  =  a  +  b. 

64.     Results  subsidiary  to  Theorem  I.     The  reader  should 
have  no  difficulty  in  verifying  the  following  subsidiary  results. 

1.  If  <f>(n)  tends  to  a  limit,  but  yfr  (n)  tends  to  +  <x>  or  to  —  oo 
or  oscillates  finitely   or   infinitely,  then  <j>  (n)  +  ^  (n)  behaves  like 
-f  O). 

2.  If  <j>  (n)  -^  +  oo  ,  and   ty  (>i)  -^  +  oo    or  oscillates  finitely, 
then  <f>  (n)  +  ty  (n)  -^  +  oo  . 


63,  64]  POSITIVE   INTEGRAL    VARIABLE  127 

In  this  statement  we  may  obviously  change  +  oo    into  —  oo 
throughout. 

3.  If  <f>  (n) -*  <x>  and  ^p  (?i) -*»  —  oo  ,  then  </>  (n)  +  yfr  (n)  may 
tend  either  to  a  limit  or  to  +  oo  or  to  —  oo  or  may  oscillate  either 
finitely  or  infinitely. 

These  five  possibilities  are  illustrated  in  order  by  (i)  (f)(n)  =  7i,  ^(n)  =  —11, 
(ii)  <j>(n)  =  n2,  y^(7i)=-7i,  (iii)  0  (n)=n,  ^(n)=-n2,  (iv)  tj)(n)=n  +  (-  1)» 
^  (n)  =  - 71,  (v)  (f)  (11)  =n?+ (  —  l)n»,  >//•  (?i)  =  -  »8.  The  reader  should  construct 
additional  examples  of  each  case. 

4.  If  </>  (n)  -*>  +  00  and  i|r  (?i)  oscillates  infinitely,  then 
<£  (n)  +  -»/r  (n)  way  tena*  to  +  00  or  oscillate  infinitely,  but  cannot 
tend  to  a  limit,  or  to  —  00 ,  or  oscillate  finitely. 

For  yjr(n)  =  {$  (71)  +  yp-  (»)}  -  <£  (»i) ;  and,  if  $  («)  +  >//•  («)  behaved  in  any  of  the 
three  last  ways,  it  would  follow,  from  the  previous  results,  that  \^  (n)  -*-  —  00 , 
which  is  not  the  ease.  As  examples  of  the  two  cases  which  are  possible, 
consider  (i)  cf>(7i)  =  7i2,  i//-(n)  =  (-l)"»,  (ii)  (f>(n)  =  n,  ^  (»)  =  (-  \)nn\  Here 
again  the  signs  of  +  oo  and  —  00  may  be  permuted  throughout. 

5.  If  <f>  (n)  and  ty  (n)  both  oscillate  finitely,  then  <f)  (n)  +  yjr  (n) 
must  tend  to  a  limit  or  oscillate  finitely. 

As  examples  take 

(i)    0(tO=(-1)»,   *(?0=(-l)»+1,        (ii)*(n)=^(»)-(-l)». 

6.  If  <f>  (n)  oscillates  finitely,  and  yjr  (n)  infinitely,  then 
<f>  (n)  +  yfr  (11)  oscillates  infinitely. 

For  <f>  (n)  is  in  absolute  value  always  less  than  a  certain  constant,  say  K. 
On  the  other  hand  -\|/-  (n),  since  it  oscillates  infinitely,  must  assume  values 
numerically  greater  than  any  assignable  number  (e.g.  10  A,  100 A', ...).  Hence 
<p  (n)  +  >//•  (n)  must  assume  values  numerically  greater  than  any  assignable 
number  (e.g.  9  A,  99 A",  ...).  Hence  <£(?£) +  ^(n)  must  either  tend  to  +00  or 
-  oo  or  oscillate  infinitely.     But  if  it  tended  to  +  oo  then 

would  also  tend  to  +cc ,  in  virtue  of  the  preceding  results.  Thus  (f>  (7i)  +  yjs  (n) 
cannot  tend  to  +00,  nor,  for  similar  reasons,  to  —00:  hence  it  oscillates 
infinitely. 

7.  If  both  <j>  (n)  and  yjr  (n)  oscillate  infinitely,  then  </>  (n)  +  i/r  (n) 
may  tend  to  a  limit,  or  to  4-  00  ,  or  to  —  00 ,  or  oscillate  either  finitely 
or  infinitely. 

Suppose,  for  instance,  that  <fi  (u)  =  (  —  l)n7i,  while  -^(n)  is  in  turn  each  of 
the  functions  (-l)»  +  1?j,  {l  +  (-l)»  +  1}  n,  -  {l  +  (-  l)n}n,  (- l)"  +  1(?i  +  l), 
(  —  l)n7i.     We  thus  obtain  examples  of  all  five  possibilities. 


128  LIMITS    OF    FUNCTIONS    OF    A  [iV 

The  results  1 — 7  cover  all  the  cases  which  are  really  distinct. 
Before  passing  on  to  consider  the  product  of  two  functions,  we 
may  point  out  that  the  result  of  Theorem  I  may  be  immediately 
extended  to  the  sum  of  three  or  more  functions  which  tend  to 
limits  as  n  -*■  oo  . 

65.  B.  The  behaviour  of  the  product  of  two  functions 
whose  behaviour  is  known.  We  can  now  prove  a  similar 
set  of  theorems  concerning  the  product  of  two  functions.  The 
principal  result  is  the  following. 

Theorem  II.     If  lim  (/>  (n)  =  a  and  Km  y}r  (n)  =  b,  then 

Km  cj>  (n)  yjr  (n)  =  ab. 
Let  <£  (n)  =  a  +  fa  (n),     yjr  (n)  =  b  +  ijr1  (n), 

so  that  lim  fa  (n)  =  0  and  lim  y\rl  (n)  =  0.     Then 

(f>(n)yfr  (n)  =  ab  +  a^  (n)  +  bfa  (n)  +  fa  (n)  -v^!  («). 

Hence  the  numerical  value  of  the  difference  <f>  (n)  yjr  (n)  —  ab  is 
certainly  not  greater  than  the  sum  of  the  numerical  values  of 
a^x  (n),  bfa  (ri),  fa  (n)  ^  (n).     From  this  it  follows  that 

lim  { cf>  (n)  yjr  (?i)  —  ab]  =  0, 
which  proves  the  theorem. 

The  following  is  a  strictly  formal  proof.     We  have 

|  (f)  (n)  \}s  (n)  -ab  \  <  |  atyx  (n)  \  +  \bfa(n)\  +  \  fa  (?i)  1 1  fa  (n)  |. 
Assuming  that  neither  a  nor  b  is  zero,  we  may  choose  n0  so  that 

\fa(n)\<l8l\b\,     \fa(n)\<iSI\a\, 
when  n  £n0.     Then 

|(/,(n)^(W)-a6l<iS  +  p  +  {i82/(|a||&|)}, 

which  is  certainly  less  than  8  if  8  <£  |  a  1 1  6 1.  That  is  to  say  we  can  choose 
h0  so  that  |  <b  (n)  ty  (n)  —  ab  \  <  8  when  »  >  n0 ,  and  so  the  theorem  follows.  The 
reader  should  supply  a  proof  for  the  case  in  which  at  least  one  of  a  and  b  is 
zero. 

We  need  hardly  point  out  that  this  theorem,  like  Theorem  I, 
may  be  immediately  extended  to  the  product  of  any  number  of 
functions  of  n.  There  is  also  a  series  of  subsidiary  theorems 
concerning  products  analogous  to  those  stated  in  §  64  for  sums. 
We  must  distinguish  now  six  different  ways  in  which  <f>  (n)  may 
behave  as  n  tends  to  co .     It  may  (1)  tend  to  a  limit  other  than 


C4-66]  POSITIVE    INTEGRAL    VARIABLE  129 

zero,  (2)  tend  to  zero,  (3a)  tend  to  +  oo ,  (36)  tend  to  -  co , 
(4)  oscillate  finitely,  (5)  oscillate  infinitely.  It  is  not  necessary,  as 
a  rule,  to  take  account  separately  of  (3a)  and  (36),  as  the  results 
for  one  case  may  be  deduced  from  those  for  the  other  by  a  change 
of  sign. 

To  state  these  subsidiary  theorems  at  length  would  occupy  more  space 
than  we  can  afford.  We  select  the  two  which  follow  as  examples,  leaving  the 
verification  of  them  to  the  reader.  He  will  find  it  an  instructive  exercise  to 
formulate  some  of  the  remaining  theorems  himself. 

(i)  If  <fi  (n)  -»-  +  oo  and  \^  (n)  oscillates  finitely,  then  cp  (n)  ^  (??)  must  tend 
to  +oo  or  to  -co  or  oscillate  infinitely. 

Examples  of  these  three  possibilities  may  be  obtained  by  taking  $  (»)  to 
be  n  and  ^  (n)  to  be  one  of  the  three  functions  2  +  ( -  1)",  —  2  —  ( -  1)»  ( -  l)n. 

(ii)  If  <fr  (n)  and  yjr  (n)  oscillate  finitely,  then  <£  (n)  \|/-  (n)  must  tend  to  a 
limit  {which  may  be  zero)  or  oscillate  finitely. 

For  examples,  take  (a)  <\>(n)  =  y\r  (»)=(-  1)",  (b)  cf>  (»)  =  l+(-l)n, 
\^(h)  =  1 -(-1)",     and   (c)  (f>  (n)  =  cos  %nn,  \|/-  (?i)  =  sin  Jhtt. 

A  particular  case  of  Theorem  II  -which  is  important  is  that 
in  which  yfr  (n)  is  constant.  The  theorem  then  asserts  simply 
that  lim  kef)  (n)  =  ka  if  lim  <f)  (n)  =  a.  To  this  we  may  join  the 
subsidiary  theorem  that  if  <f>(n)-^  +  cc  then  k(f)  (n)  -**  +  go  or 
k  <f>  (n)  -^-  —  oc  ,  according  as  k  is  positive  or  negative,  unless  k  =  0, 
when  of  course  k<f)  (n)  =  0  for  all  values  of  n  and  lim  kcf>  (n)  =  0. 
And  if  (f>  (n)  oscillates  finitely  or  infinitely,  then  so  does  k<j>  (n), 
unless  k  =  0. 

66.  C.  The  behaviour  of  the  difference  or  quotient  of 
two  functions  whose  behaviour  is  known.  There  is,  of 
course,  a  similar  set  of  theorems  for  the  difference  of  two  given 
functions,  which  are  obvious  corollaries  from  what  precedes.  In 
order  to  deal  with  the  quotient 

we  begin  with  the  following  theorem. 

Theorem  III.     If  lim  </>  (n)  =  a,  and  a  is  not  zero,  then 

1         1 
lim  rpr  =  - . 
9  (n)      a 

Let  (f>  (n)  =  a  +  ^  (n), 

u.  9 


130  LIMITS   OF   FUNCTIONS   OF   A  [lV 

so  that  lira  fa  (n)  =  0.     Then 


1         1 


<f>  (n)     a 


fain) 


a  1 1  a  +  fa  (n) 


and  it  is  plain,  since  lim  fa  (n)  =  0,  that  we  can  choose  n0  so  that 
this  is  smaller  than  any  assigned  number  8  when  n  =  n0 . 

From  Theorems  II  and  III  we  can  at  once  deduce  the  principal 
theorem  for  quotients,  viz. 

Theorem  IV.  If  lim  <j>  (n)  =  a  and  lim  yjr  (n)  =  b,  and  b  is  not 
zero,  then 

limiM     a 
i/r  (n)     b 

The  reader  will  again  find  it  instructive  to  formulate,  prove, 
and  illustrate  by  examples  some  of  the  '  subsidiary  theorems ' 
corresponding  to  Theorems  III  and  IV. 

67.  Theorem  V.  IfR{j>  (n),  -»/r  (n),  %  (n), ...}  is  any  rational 
function  of  $>(n),  yjr  (?i),  %(n),  ... ,  i.e.  any  function  of  the  form 

P{cf>(n),  ir(n),  X(n),  ...J/Qft(n),  ^(n),  X(n),  ...}, 
where  P  and  Q  denote 'polynomials  in  <j>(n),  ty(n),  %(n), ...:  and  if 

lim  $ (n)  =  a,     lim  \|r  (n)  =  b,     lim  %  (n)  =  c, ... , 
and  Q(a,  b,  c,  ...)=f  0  ; 

then  lim  R  {<f>  (n),  ty  (n),  % (n),  ...}  =  R  (a,  b,  c, . ..). 

For  P  is  a  sum  of  a  finite  number  of  terms  of  the  type 

where  A  is  a  constant  and  p,  q,  ...  positive  integers.  This  term, 
by  Theorem  II  (or  rather  by  its  obvious  extension  to  the  product 
of  any  number  of  functions)  tends  to  the  limit  Aapbq ...,  and  so  P 
tends  to  the  limit  P  (a,  b.  c, ...),  by  the  similar  extension  of 
Theorem  I.  Similarly  Q  tends  to  Q  (a,  b,  c,  ...);  and  the  result 
then  follows  from  Theorem  IV. 

68.  The  preceding  general  theorem  may  be  applied  to  the 
following  very  important  particular  problem  :  luhat  is  the  behaviour 
of  the  most  general  rational  function  of  n,  viz. 

o/     _  a0nv  +  a^iP-1  +  .  •  •  +  ap 
*{n)~b0n«+b1n«-i  +  ...+bq' 
as  n  tends  to  oo  *  ? 

*  We  naturally  suppose  that  neither  a0  nor  b0  is  zero. 


66-69]  POSITIVE    INTEGRAL   VARIABLE  131 

In  order  to  apply  the  theorem  we  transform  S  (n)  by  writing 
it  in  the  form 

The  function  in  curly  brackets  is  of  the  form  R{4>(n)},  where 
</>  (w)  =  1/n,  and  therefore  tends,  as  n  tends  to  oo  ,  to  the  limit 
R(0)=*aQ/b0.  Now  rc*-«-^0  if  p  <  q  ;  n*-4=l  and  n*"-?— 1  if 
p  =  gr  j  and  nP-Q-*-  +  go  if  p  >  5.    Hence,  by  Theorem  II, 

lim  8  (n)  =  0     (p  <q), 

lim  £  (n)  =  Oo/60     (i>  =  q), 

S  (n)  -*■  +  go      (p>  q,  a0/b0  positive), 

S  (n)  -*■  —  00      (p>  q,  aQ/b0  negative). 

Examples  XXVI.     1.     What  is  the  behaviour  of  the  functions 
'?i-l\2       ?i2  +  l       .     ,.    H2+l 


0£)'.  (-nS)-  =£• 1-* 


as  ?i  -s-  co 


2.  Which  (if  any)  of  the  functions 

l/(cos2 ^mr  +n  sin2  \nir),     \j{n  (cos2 \nn +n  sin2 1  wtt)} , 
(h  cos2 hnu  +  sin2 ^?itt)/{?i  (cos2 %nir+n  sm2^mr)} 
tend  to  a  limit  as  n  -*-  00  ? 

3.  Denoting  by  aS1  (n)  the  general  rational  function  of  n  considered  above, 
show  that  in  all  cases 

,im%ti>  um*{.+,(WU 

aS  (n)  o  (») 

69.     Functions  of  n  which  increase  steadily  with  n.    A 

special  but  particularly  important  class  of  functions  of  n  is  formed 
by  those  whose  variation  as  n  tends  to  00  is  always  in  the  same 
direction,  that  is  to  say  those  which  always  increase  (or  always 
decrease)  as  n  increases.  Since  —  <f>  (n)  always  increases  if  <£  (n) 
always  decreases,  it  is  not  necessary  to  consider  the  two  kinds  of 
functions  separately ;  for  theorems  proved  for  one  kind  can  at 
once  be  extended  to  the  other. 

Definition.    The  function  <f>  (n)  will  be  said  to  increase  steadily 
with  n  if  (j>  (n  +  1)  =  (f>  (n)  for  all  values  of  n. 

9—2 


132  LIMITS   OF   FUNCTIONS   OF   A  [lV 

It  is  to  be  observed  that  we  do  not  exclude  the  case  in  which 
$  (n)  has  the  same  value  for  several  values  of  n ;  all  we  exclude  is 
possible  decrease.     Thus  the  function 

<f>(n)  =  2n+(-iy\ 

whose  values  for  n  =  0,  1,  2,  3,  4,  . . .  are 

1,  1,5,5,9,  9,... 

is  said  to  increase  steadily  with  n,  Our  definition  would  indeed 
include  even  functions  which  remain  constant  from  some  value  of  n 
onwards;  thus  $(w)  =  l  steadily  increases  according  to  our  definition. 
However,  as  these  functions  are  extremely  special  ones,  and  as 
there  can  be  no  doubt  as  to  their  behaviour  as  n  tends  to  00 ,  this 
apparent  incongruity  in  the  definition  is  not  a  serious  defect. 

There  is  one  exceedingly  important  theorem  concerning 
functions  of  this  class. 

Theorem,  If  </>  (?i)  steadily  increases  with  n,  then  either 
(i)  </>(n)  tends  to  a  limit  as  n  tends  to  00,  or  (ii)  <£  (n)  -*  +  go  . 

That  is  to  say,  while  there  are  in  general  five  alternatives  as  to 
the  behaviour  of  a  function,  there  are  two  only  for  this  special 
kind  of  function. 

This  theorem  is  a  simple  corollary  of  Dedekind's  Theorem 
(§  17).  We  divide  the  real  numbers  £  into  two  classes  L  and  R, 
putting  f  in  L  or  R  according  as  <£  (n)  =  £  for  some  value  of  n 
(and  so  of  course  for  all  greater  values),  or  <j>  (n)  <  £  for  all 
values  of  n. 

The  class  L  certainly  exists ;  the  class  R  may  or  may  not. 
If  it  does  not,  then,  given  any  number  A,  however  large,  <f>  (n)  >  A 
for  all  sufficiently  large  values  of  n,  and  so 

0  (n)  -*■  +  go  . 

If  on  the  other  hand  R  exists,  the  classes  L  and  R  form  a 
section  of  the  real  numbers  in  the  sense  of  §  17.  Let  a  be  the 
number  corresponding  to  the  section,  and  let  8  be  any  positive 
number.  Then  <j>  (n)  <  a  +  8  for  all  values  of  n,  and  so,  since  8  is 
arbitrary,  </>  (71)  £  a.  On  the  other  hand  $  (n)  >  a  —  8  for  some 
value  of  n,  and  so  for  all  sufficiently  large  values.    Thus 

a—  8<<j>  (71)  <  a 


G9,  70]  POSITIVE   INTEGRAL   VARIABLE  133 

for  all  sufficiently  large  values  of  n;  i.e. 

It  should  be  observed  that  in  general  <£  (n)  <  a  for  all  values  of  n ;  for  if 
(j>  (n)  is  equal  to  a  for  any  value  of  n  it  must  be  equal  to  a  for  all  greater 
values  of  n.  Thus  <£  (n)  can  never  be  equal  to  a  except  in  the  case  in  which 
the  values  of  (f>  (n)  are  ultimately  all  the  same.  If  this  is  so,  a  is  the  largest 
member  of  L ;  otherwise  L  has  no  largest  member. 

Cor.  1.  If(f>  (n)  increases  steadily  with  n,  then  it  will  tend  to  a 
limit  or  to  +  cc  according  as  it  is  or  is  not  possible  to  find  a  number 
K  such  that  (j>  (n)  <  K  for  all  values  of  n. 

We  shall  find  this  corollary  exceedingly  useful  later  on. 

Cor.  2.  If  <£  (?i)  increases  steadily  with  n,  and  $  (n)  <  K  for 
all  values  ofn,  then  <f>  (n)  tends  to  a  limit  and  this  limit  is  less  than 
or  equal  to  K. 

It  should  be  noticed  that  the  limit  may  be  equal  to  K:  if  e.g. 
(f>  (n)  =  3  —  (1/w),  then  every  value  of  (f>  (n)  is  less  than  3,  but  the 
limit  is  equal  to  3. 

Cor.  3.     If  $  (n)  increases  steadily  luith  n,  and  tends  to  a  limit, 

then 

4>  (n)  =  lim  <f>  (n) 
for  all  values  ofn. 

The  reader  should  write  out  for  himself  the  corresponding 
theorems  and  corollaries  for  the  case  in  which  </>  (n)  decreases  as  n 
increases. 

70.  The  great  importance  of  these  theorems  lies  in  the  fact 
'  that  they  give  us  (what  we  have  so  far  been  without)  a  means  of 
deciding,  in  a  great  many  cases,  whether  a  given  function  of  n 
does  or  does  not  tend  to  a  limit  as  n-^cc ,  without  requiring  us  to 
be  able  to  guess  or  otherwise  infer  beforehand  what  the  limit  is.  If 
we  know  what  the  limit,  if  there  is  one,  must  be,  we  can  use  the 

test 

|  (f>  O)  -l\<8         (n^  n0) : 

as  for  example  in  the  case  of  <f>  (n)  =  1/n,  where  it  is  obvious  that 
the  limit  can  only  be  zero.  But  suppose  we  have  to  determine 
whether 


*w-(1+i 


134  LIMITS  OF   FUNCTIONS   OF   A  [IV 

tends  to  a  limit.  In  this  case  it  is  not  obvious  what  the  limit,  if 
there  is  one,  will  be :  and  it  is  evident  that  the  test  above,  which 
involves  I,  cannot  be  used,  at  any  rate  directly,  to  decide  whether 
I  exists  or  not. 

Of  course  the  test  can  sometimes  be  used  indirectly,  to  prove  by  means  of 
a  reductio  ad  absurdum  that  I  cannot  exist.  If  e.g.  $  (»)=(-  l)'v,  it  is  clear 
that  I  would  have  to  be  equal  to  1  and  also  equal  to  - 1,  which  is  obviously 
impossible. 

71.  Alternative  proof  of  Weierstrass's  Theorem  of  §  19.  The  results 
of  §  69  enable  us  to  give  an  alternative  proof  of  the  important  theorem 
pi-oved  in  §  19. 

If  we  divide  PQ  into  two  equal  parts,  one  at  least  of  them  must  contain 
infinitely  many  points  of  S.  We  select  the  one  which  does,  or,  if  both  do,  we 
select  the  left-hand  half ;  and  we  denote  the  selected  half  by  Pt  Qx  (Fig.  28). 
If  PiQi  is  the  left-hand  half,  Px  is  the  same  point  as  P. 

p3  Q3 

Pi I  T  Qi 


Q2  Q 


p4  Q4 


Fig.  28. 
Similarly,  if  we  divide  Pt  Qt  into  two  halves,  one  at  least  of  them  must 
contain  infinitely  many  points  of  S.     "We  select  the  half  P2Q2  which  does  so, 
or,  if  both  do  so,  we  select  the  left-hand  half.     Proceeding  in  this  way  we  can 
define  a  sequence  of  intervals 

PQi  PiQii  P2Q21  PzQzi  •••> 

each  of  which  is  a  half  of  its  predecessor,  and  each  of  which  contains  infinitely 
many  points  of  S. 

The  points  P,  Pu  P2,  ...  progress  steadily  from  left  to  right,  and  so  Pn 
tends  to  a  limiting  position  T.  Similarly  Qn  tends  to  a  limiting  position  T'. 
But  TT'  is  plainly  less  than  PnQH,  whatever  the  value  of  n;  and  PnQn,  being 
equal  to  PQ/2n,  tends  to  zero.  Hence  T'  coincides  with  T,  and  Pn  and  Qn 
both  tend  to  T. 

Then  T  is  a  point  of  accumulation  of  S.  For  suppose  that  £  is  its 
coordinate,  and  consider  any  interval  of  the  type  (£-8,  £+8).  If  n 
is  sufficiently  large,  PnQn  will  lie  entirely  inside  this  interval*.  Hence 
(£-£>  £+8)  contains  infinitely  many  points  of  S. 

72.     The  limit  of  x11  as  n  tends  to  00 .     Let  us  apply  the 

results    of  §    69    to    the   particularly   important   case   in   which 

</>  (n)  =  xn.   If  x  =  1  then  <f>  (n)  =  1,  lim  </>  (n)  =  1,  and  if  x  =  G  then 

<f>(n)  =  0,  lim  (f>(n)=0,  so  that  these  special  cases  need  not  detain  us. 

*  This  will  certainly  be  the  case  as  soon  as  PQj2n  <  8. 


70-72]  POSITIVE   INTEGRAL    VARIABLE  135 

First,  suppose  x  positive.  Then,  since  <f>  (n  +  1)  =  x<$>  (n),  </>  (n) 
increases  with  n  if  x  >  1,  decreases  as  n  increases  if  x  <  1. 

If  x>l,  then  xn  must  tend  either  to  a  limit  (which  must 
obviously  be  greater  than  1)  or  to  +  oo .  Suppose  it  tends  to  a 
limit  I.     Then  lim  <£  (n  +  1)  =  lim  cf>  (71)  =  I,  by  Exs.  xxv.  7  ;  but 

lim  <f>  (n  +  1)  =  lim  x<j>  (n)  =  x  lim  <f>  (n)  =  xl, 

and  therefore  l  —  xl:  and  as  x  and  I  are  both  greater  than  1,  this 
is  impossible.     Hence 

xn  -*+  CO  (x  >  1  ). 

Example.  The  reader  may  give  an  alternative  proof,  showing  by  the 
binomial  theorem  that  xn>l  +  n8  if  8  is  positive  and  x=  1  +  8,  and  so  that 

xn  -»-  +  oo . 

On  the  other  hand  xn  is  a  decreasing  function  if  x  <  1,  and 
must  therefore  tend  to  a  limit  or  to  —  00 .     Since  xn  is  positive 
the  second  alternative  may  be  ignored.     Thus  lim  xn  =  I,  say,  and 
as  above  I  =  xl,  so  that  I  must  be  zero.     Hence 
lima?"  =  0         (0<«<1). 

Example.  Prove  as  in  the  preceding  example  that  (l/x)n  tends  to  +00  if 
0<1r<l,  and  deduce  that  x11  tends  to  0. 

We  have  finally  to  consider  the  case  in  which  x  is  negative. 
If  —  1  <  x  <  0  and  x  =  —  y,  so  that  0  <  y  <  1,  then  it  follows  from 
what  precedes  that  lim  yn  =  0  and  therefore  lim  x11  =  0.  If  x  =  —  1 
it  is  obvious  that  xn  oscillates,  taking  the  values  —  1,  1  alterna- 
tively. Finally  if  x  <  —  1,  and  x  =  —  y,  so  that  y  >  1,  then  yn  tends 
to  +  00  ,  and  therefore  xn  takes  values,  both  positive  and  negative, 
numerically  greater  than  any  assigned  number.  Hence  xn  oscillates 
infinitely.     To  sum  up : 

<£  (n)  =  xn  -*»  +  00  0>1), 

lim  0  (n)  =  1  (x  =  1), 

lim<£(¥)  =  0  (—1<x<1), 

<f)  (n)  oscillates  finitely  (x  =  —  l), 

<f>  (n)  oscillates  infinitely       (x  <  —  1). 

Examples  XXVII*.  1.  If  0  (n)  is  positive  and  (f>(n  +  l)>K(f)(n),  where 
A">1,  for  all  values  of  n,  then  0  («)-*- 4-  30  . 

*  These  examples  are  particularly  important  and  several  of  them  will  be  made 
use  of  later  in  the  text.     They  should  therefore  be  studied  very  carefully. 


136  LIMITS   OF   FUNCTIONS   OF   A  [lV 

[For  <f>(n)>K<t>(n-l)>K*<j>(n-2)  ...  >A»-i0(l), 

from  which  the  conclusion  follows  at  once,  as  A"n-»-oo.] 

2.  The  same  result  is  true  if  the  conditions  above  stated  are  satisfied 
only  when  n  >  n0. 

3.  If  4>{n)  is  positive  and  <f)(n  +  l)<K(f>(ri),  where  0<A"<1,  then 
lim  (f>  (n)  =  0.  This  result  also  is  true  if  the  conditions  are  satisfied  only  when 
n  Sn0. 

4.  If  |<£(«  +  l)|<A~|0(n)|  when  n^.n0,  and  0<A"<1,  then  lim  <£(«)  =  0. 

5.  If  0(w)  is  positive  and  lim  {<p(n  +  l)}/{(f>(n)}  =  l>l,  then  0(?i)-»-+oo. 
[For  we  can  determine  n0  so  that  {<£  (n  + 1)} I {(f)  (n)}>K>l  when  n^n^ :  wc 

may,  e.g.,  take  A' half-way  between  1  and  I.     Now  apply  Ex.  1.] 

6.  If  lim  {0  (w  +  l)}/{0  (w)}  =  ?,  where  I  is  numerically  less  than  unity, 
then  Yim(f)(n)  —  0.     [This  follows  from  Ex.  4  as  Ex.  5  follows  from  Ex.  1.] 

7.  Determine  the  behaviour,  as  ?i-»-co,  of  (f>(n)=nrxn,  where  r  is  any 
positive  integer. 

[If  x=0  then  (f)(n)=0  for  all  values  of  n,  and  0  («)-»•  0.     In  all  other  cases 
(f)(n  +  l)  _  fn  +  V 


(f)  (n)         \   n 

First  suppose  x  positive.  Then  0(?i)-*  +  °°  if  x>\  (Ex.  5)  and  (f>(n)-*-0  if 
x<l  (Ex.  G).  If  x—1,  then  0  (»)=M.r-»-+ oo  .  Next  suppose  x  negative. 
Then  |  (f>(n)  \=nr\  x  |"  tends  to  +oo  if  |#||£l  and  to  0  if  |.r|<l.  Hence 
(f>  (n)  oscillates  infinitely  if  xS,-  1  and  0  (»)-»0  if  -  l<.v<0.] 

8.  Discuss  n~rxn  in  the  same  way.  [The  results  are  the  same,  except 
that  (f>(n)-*-0  when  x  =  l  or   —  1.] 

9.  Draw  up  a  table  to  show  how  nkxn  behaves  as  ?i-*-oo,  for  all  real 
values  of  x,  and  all  positive  and  negative  integral  values  of  k. 

[The  reader  will  observe  that  the  value  of  k  is  immaterial  except  in  the 
special  cases  when  x  =  l  or  —  1.  Since  \\m  {(n+l)/n}k=l,  whether  k  be 
positive  or  negative,  the  limit  of  the  ratio  (f>(n  +  l)/(f)(n)  depends  only  on 
x,  and  the  behaviour  of  0  (»)  is  in  general  dominated  by  the  factor  x'\  The 
factor  nk  only  asserts  itself  when  x  is  numerically  equal  to  1.] 

10.  Prove  that  if  x  is  positive  then  ?]x-*-\  as  n^*-  so  .  [Suppose,  e.g.,  x>\. 
Then  at,  -Jx,  ^'x,  ...  is  a  decreasing  sequence,  and  ^/.r>l  for  all  values  of  n. 
Thus  Hjx-*-l,  where  1^.1.  But  if  l>\  we  can  find  values  of  n,  as  large  as 
we  please,  for  which  f/x>l  or  x>ln ;  and,  since  £"-*-  +  oo  as  ?i-*oo,  this 
is  impossible.] 

11.  Z/n-*\.  [For  B  +  V(«  +  l)<\/»  if  (»  +  l)B<flB  +  1  or  {1+(1/«)}W<«, 
which  is  certainly  satisfied  if  «23  (see  §  73  for  a  proof).  Thus  yfn  decreases 
as  n  increases  from  3  onwards,  and,  as  it  is  always  greater  than  unity,  it  tends 
to  a  limit  which  is  greater  than  or  equal  to  unity.  But  if  tfn-*-l,  where  l>l, 
then  n>ln,  which  is  certainly  untrue  for  sufficiently  large  values  of  n, 
since  ?"/«-*■ +  oo  with  n  (Exs.  7,  8).] 


72,  73]  POSITIVE   INTEGRAL   VARIABLE  137 

12.  Z/(n  !)-*-  +  co  .  [However  large  A  may  be,  n !  >  An  if  n  is  large  enough. 
For  if  un  =  An/n  !  then  un  +  i/un=AJ(n+l),  which  tends  to  zero  as  » -»■<»,  so 
that  un  does  the  same  (Ex.  6).] 

13.  Show  that  if —1<^7<1  then 

m  (m  —  1) ...  (m  —  n+1) 
n ! 
tends  to  zero  as  n  -*■  co . 

[If  m  is  a  positive  integer,  «n=0  for  n  >  m.     Otherwise 


unless  x=0.] 

73.     The  limit  of  ( I  +-J  .     A  more  difficult  problem  which 
can  be  solved  by  the  help  of  §  69  arises  when  <f>(n)  =  {1  4-  l/w]n. 

It  follows  from  the  binomial  theorem*  that 

,       IV1     .,  1      n(n-l)l  n(n-l)...(n-n  +  l)  1 

1+-    =l+n.-+     \    a  '-  +  ...  + 


nj  n  1 .  2     n2  1.2...  n  n1 

The  (p  +  l)th  term  in  this  expression,  viz. 

1      ^i_lVi_?V..A_£=i 


1  .  2  ...£>  \       nj \       n; 
is  positive  and   an   increasing   function   of  n,   and   the   number 

of  terms  also  increases  with  n.     Hence  ( 1  +  - )    increases  with  n, 
and  so  tends  to  a  limit  or  to  +  oo  ,  as  n  -*■  x . 
But 

1  +  -     <l  +  l  +  =— a  +  ^     ^    o  +•••  + 


nj  1.2     1.2.3  1.2.3...» 


Thus  ( 1  +  - )    cannot  tend  to  +  oo ,  and  so 
lim  (  1  +  - )    =  e, 


where  e  is  a  number  such  that  2  <  e  ^  3. 

*  The  binomial  theorem  for  a  positive  integral  exponent,  which  is  what  is  used 
here,  is  a  theorem  of  elementary  algebra.  The  other  cases  of  the  theorem  belong 
to  the  theory  of  infinite  series,  and  will  be  considered  later. 


138  LIMITS   OF    FUNCTIONS   OF    A  [IV 

74.  Some  algebraical  lemmas.  It  will  be  convenient  to  prove  at 
this  stage  a  number  of  elementary  inequalities  which  will  be  useful  to  us 
later  on. 

(i)     It  is  evident  that  if  a>l  and  ;•  is  a  positive  integer  then 
rar>ar-'l  +  ar--+...  +  l. 
Multiplying  both  sides  of  this  inequality  by  o—  1,  we  obtain 

rar(a-l)>ar-l  ; 
and  adding  r  (ar-  1)  to  each  side,  and  dividing  by  r  (?'+l),  we  obtain 

~7+r>—    <a>]) (1)- 

Similarly  we  can  prove  that 

L-B—<l—E       (0</3<l) (2). 

It  follows  that  if  r  and  s  are  positive  integers,  and  r>s,  then 

ar_l       a._l        1-gr        1-ffl 

~^>^~>  ~F~<~7~    (u}- 

Hei-e  0</3<l<a.     In  particular,  when  5=1,  we  have 

ar-l>r(a-l),      l-^<r(l-$)   (4). 

(ii)  The  inequalities  (3)  and  (4)  have  been  proved  on  the  supposition 
that  r  and  s  are  positive  integers.  But  it  is  easy  to  see  that  they  hold  under 
the  more  general  hypothesis  that  r  and  s  are  any  positive  rational  numbers. 
Let  us  consider,  for  example,  the  first  of  the  inequalities  (3).  Let  r=a/b, 
s=c/d,  where  a,  b,  c,  d  are  positive  integers;  so  that  ad>bc.  If  we  put 
a =yhd,  the  inequality  takes  the  form 

(y'"l-l)lad>(y!«-l)lbc; 

and  this  we  have  proved  already.     The  same  argument  applies  to  the  re- 
maining inequalities ;  and  it  can  evidently  be  proved  in  a  similar  manner  that 

n8-l<s(a-l),      l-j3«>*(l-j8)  (5), 

if  s  is  a  positive  rational  number  less  than  1. 

(iii)  In  what  follows  it  is  to  be  understood  that  all  the  letters  denote 
positive  numbers,  that  r  and  s  are  rational,  and  that  a  and  r  are  greater 
than  1,  /3  and  s  less  than  1.  Writing  1//3  for  a,  and  1/a  for  /3,  in  (4),  we 
obtain 

a'-Kra'-^a-l),     1  -^rp-1  (1  -/9)    (6). 

Similarly,  from  (5),  we  deduce 

a»-l>sa*-l(a-l),      1  -  ^S/S8"1  (1  -/3)    (7). 

Combining  (4)  and  (6),  we  see  that 

rar-1(fl-l)>dr-l>r(a-l)  (8). 


74,  75]  POSITIVE   INTEGRAL   VARIABLE  139 

Writing  xjy  for  a,  we  obtain 

rxr~ a  (.v  -  y)>xr  -  yr>ryr~1  (x  -  y) (9) 

if  x>y>0.    And  the  same  argument,  applied  to  (5)  and  (7),  leads  to 

sx*-1  {x-y)<xs-y»<syi~1  (x-y)    (10). 

Examples  XXVIII     1.     Verify  (9)  for  r-=2,  3,  and  (10)  for  s=|,  A. 

2.  Show  that  (9)  and  (10)  are  also  true  if  y>x>0. 

3.  Show  that  (9)  also  holds  for  r<0.  [See  Chrystal's  Algebra,  vol.  ii, 
pp.  43—45.] 

4.  If  $(ri)-~l,  where  l>0,  as  n-*>oo,  then  cf)k-*lk,  k  being  any  rational 
number. 

[We  may  suppose  that  k  >  0,  in  virtue  of  Theorem  III  of  §  66 ;  and  that 
%l<(f><  21,  as  is  certainly  the  case  from  a  certain  value  of  n  onwards.  If 
k>  1 

£0* ~1((f)-l)>cf)k-  lk  >Hk~l  (<f>  - 1) 

or  H*"1  (l-<f))>lk-(j)k>k(f)k-i(l-(j)), 

according  as  $>l  or  <f)<l.  It  follows  that  the  ratio  of  \<f>k  —  lk\  and  \<j>  — 1\ 
lies  between  k  {\l)k~l  and  k  (2l)k-1.  The  proof  is  similar  when  0  <k  <1.  The 
result  is  still  true  when  £=0,  if  k  >  0.] 

5.  Extend  the  results  of  Exs.  xxvu.  7,  8,  9  to  the  case  in  which  r  or  k 
are  any  rational  numbers. 

75.  The  limit  of  n  (#x  -1).  If  in  the  first  inequality  (3)  of  §  74  we 
put  r  =  l/(n  —  1),  8  =  1 /n,  we  see  that 

(»-l)(n-#a-l)>»(#a-l) 

when  a>l.  Thus  if  <£  (n)  =  n  (tya  —  l)  then  0  («)  decreases  steadily  as  ?i  in- 
creases. Also  (j>  (n)  is  always  positive.  Hence  cf>  (n)  tends  to  a  limit  I  as 
n-*-aa  ,  and  l=0. 

Again  if,  in  the  first  inequality  (7)  of  §  74,  we  put  s  =  1  ju,  we  obtain 

7i(Va-l)>ya(l-l)>l~. 

Thus  I  >  1  -  ( 1  /a)  >  0.     Hence,  if  a  >  1 ,  we  have 

lim  n(Z/a-l)=f(a), 

M-*oo 

where  /(a)>0. 

Next  suppose  /3<l,andlet/3  =  l/a;  thenn  (%/$-!)  =  -n(^a-l)/^/a.  Now 
M(v^a-l)-*-/(a),  and  (Exs.  xxvu.  10) 

Hence,  if  /3  =  l/a<l,  we  have 

w«//9-l)— /(a). 
Finally,  if  #=1,  then  n  (Z/x-l)  =  0  for  all  values  of  n. 


140  LIMITS  OF   FUNCTIONS  OF  A  [IV 

Thus  we  arrive  at  the  result :  the  limit 

\\m.n{Zjx—  1) 

defines  a  function  of  x  for  all  positive  values  of  x.     This  function  f  (x)  possesses 

the  properties 

f(l/x)=-f(x),    /(1)=0, 

and  is  positive  or  negative  according  as  x>\  or  x<\.     Later  on  we  shall  he 

able  to  identify  this  function  with  the  Napierian  logarithm  of  x 

Example.     Prove  that/ (xy)=f{x)  +f{y).    [Use  the  equations 

f(xy)  =  \im  n  ($xy  -  l)  =  lim  {n  (#*-l)  ^+71(^  —  1)}.] 

76.     Infinite  Series.     Suppose  that  u  (»)  is  any  function  of 
n  denned  for  all  values  of  n,     If  we  add  up  the  values  of  u  (v) 
for  v  =  l,  2, ...  n,  we  obtain  another  function  of  n,  viz. 
s  (n)  =  u  (1)  +  u  (2)  + ...  +  u  (n), 

also  denned  for  all  values  of  n.     It  is  generally  most  convenient 
to  alter  our  notation  slightly  and  write  this  equation  in  the  form 
sn  =  u1  +  u2  +  ...+un, 

n 

or,  more  shortly,  sn  —  2  w„. 

v  =  l 

If  now  we  suppose  that  sn  tends  to  a  limit  s  when  n  tends 
to  oo  ,  we  have 

n 

lim    2  M»  =  s. 

n-*-oo  j/  =  l 

This  equation  is  usually  written  in  one  of  the  forms' 

00 

2  «„  =  s,    it!  +  u2  +  u3  +  . . .  =  s, 

the  dots  denoting  the  indefinite  continuance  of  the  series  of  m's. 

The  meaning  of  the  above  equations,  expressed  roughly,  is 
that  by  adding  more  and  more  of  the  m's  together  we  get  nearer 
and  nearer  to  the  limit  s.  More  precisely,  if  any  small  positive 
number  8  is  chosen,  we  can  choose  n0  (8)  so  that  the  sum  of  the  first 
n0  (S)  terms,  or  any  of  greater  number  of  terms,  lies  between  s  —  S 
and  s  +  8;  or  in  symbols 

s  —  B<sn<s  +  8, 
if  n  =  n0(8).     In  these  circumstances  we  shall  call  the  series   • 

Mi  +  u2+  ... 
a  convergent  infinite  series,  and  we  shall  call  s  the  sum  of  the 
series,  or  the  sum  of  all  the  terms  of  the  series. 


75-77]  POSITIVE   INTEGRAL  VARIABLE  141 

Thus  to  say  that  the  series  Wj  +  u2+  ...  converges  and  has  the 
sum  s,  or  converges  to  the  sum  s  or  simply  converges  to  s,  is  merely 
another  way  of  stating  that  the  sum  sn  =  Wj  +  u2+  ...  +  un  of  the 
first  n  terms  tends  to  the  limit  s  as  n  -»■  oo ,  and  the  consideration 
of  such  infinite  series  introduces  no  new  ideas  beyond  those  with 
which  the  early  part  of  this  chapter  should  already  have  made 
the  reader  familiar.  In  fact  the  sum  sn  is  merely  a  function  cf>  (n), 
such  as  we  have  been  considering,  expressed  in  a  particular  form. 
Any  function  <p  (n)  may  be  expressed  in  this  form,  by  writing 
4>  (»)  =  0  (1)  +  {</>  (2)  -0  (1)}  +  ...  +  {0  (n)  -<j>  (n-  1)}  ; 
and  it  is  sometimes  convenient  to  say  that  <f>(n)  converges  (instead 
of  '  tends ')  to  the  limit  I,  say,  as  n  -*-  oo  . 

If  sn  -*■  +  oo  or  sn  -*■  —  oo  ,  we  shall  say  that  the  series  u1  +  u2+  ... 
is  divergent  or  diverges  to  +  oo ,  or  —  oo ,  as  the  case  may  be. 
These  phrases  too  may  be  applied  to  any  function  </>  (n) :  thus  if 
0  («)  -*■  +  oo  we  may  say  that  </>  (n)  diverges  to  +  go  .  If  sn  does 
not  tend  to  a  limit  or  to  +  oo  or  to  —  oo ,  then  it  oscillates  finitely  or 
infinitely :  in  this  case  Ave  say  that  the  series  Wj  +u2  +  ...  oscillates 
finitely  or  infinitely*. 

77.  General  theorems  concerning  infinite  series.  When 
we  are  dealing  with  infinite  series  we  shall  constantly  have 
occasion  to  use  the  following  general  theorems. 

(1)  If  u1  +  ti,,  + ...  is  convergent,  and  has  the  sum  s,  then 
a  +  u1  -f-  u2  +  ...  is  convergent  and  has  the  sum  a  +  s.  Similarly 
a  +  b  +  c  +  . . .  +  k  +  ^l1  +  u2  +  . . .  is  convergent  and  has  the  sum 
a  +  b  +  c  +  ...  +  k  +  s. 

(2)  If  MJ  +  W2+...  is  convergent  and  has  the  sum  s,  then 
w»i+i  +  u-fit+a  +  •  •  •  is  convergent  and  has  the  sum 

(3)  If  any  series  considered  in  (1)  or  (2)  diverges  or  oscillates, 
then  so  do  the  others. 

(4)  If  Mj  +  Ma+...  is  convergent  and  has  the  sum  s,  then 

leu,!  +  ku2  +  ...  is  convergent  and  has  the  sum  Jcs. 

*  The  reader  should  be  warned  that  the  words  '  divergent '  and  *  oscillatory ' 
are  used  differently  by  different  writers.  The  use  of  the  words  here  agrees  with 
that  of  Bromwich's  Infinite  Series.  In  Hobson's  Theory  of  Functions  of  a  Real 
Variable  a  series  is  said  to  oscillate  only  if  it  oscillates  finitely,  series  which 
oscillate  infinitely  being  classed  as '  divergent'.  Many  foreign  writers  use  'divergent ' 
as  meaning  merely  '  not  convergent '. 


142  LIMITS   OF   FUNCTIONS    OF   A  [iV 

(5)  If  the  first  series  considered  in  (4)  diverges  or  oscillates, 
then  so  does  the  second,  unless  k  =  0. 

(6)  If  Wj  +  u2  +  ...  and  vx  +  v2  +  ...  are  both  convergent,  then 
the  series  (z<j  +  fli)  +  (u2  +  v2)  +  • . .  is  convergent  and  its  sum  is  the 
sum  of  the  first  two  series. 

All  these  theorems  are  almost  obvious  and  may  be  proved  at 
once  from  the  definitions  or  by  applying  the  results  of  §§  63 — 66  to 
the  sum  sn  =  u^  +  u2  +  ...  +  un.  Those  which  follow  are  of  a  some- 
what different  character. 

(7)  If  ux  +  u2  +  . . .  is  convergent,  then  lim  un  =  0. 

For  un  =  sn  —  5n_!,  and  sn  and  sn_x  have  the  same  limit  s. 
Hence  lim  un  =  s  —  s  =  0. 

The  reader  may  be  tempted  to  think  that  the  converse  of  the  theorem  is 
true  and  that  if  lim  un=0  then  the  series  2un  must  be  convergent.  That  this 
is  not  the  case  is  easily  seen  from  an  example.     Let  the  series  be 

1 +*+*+*+■•• 

so  that  un=\jn.     The  sum  of  the  first  four  terms  is 

i+l+|+i>i+Hj=1+Hi 

The  sum  of  the  next  four  terms  is  |+J  +  ^+|>f  =  |;  the  sum  of  the  next 
eight  terms  is  greater  than  fy  =  | ,  and  so  on.     The  sum  of  the  first 

4  +  4  +  8  +  16-f  ...+2"  =  2n  +  1 
terms  is  greater  than 

2+J+|+J+...+|=i(B+3)1 
and  this  increases  beyond  all  limit  with  n :  hence  the  series  diverges  to  +  co . 

(8)  If  ux  +u2  +  u3  +  ...  is  convergent,  then  so  is  any  series 
formed  by  grouping  the  terms  in  brackets  in  any  way  to  form  new 
single  terms,  and  the  sums  of  the  two  series  are  the  same. 

The  reader  will  be  able  to  supply  the  proof  of  this  theorem.  Here  again 
the  converse  is  not  true.    Thus  1-1  +  1-1  +  .. .  oscillates,  while 

(1-1)+(1-1)+... 

or  0  +  0  +  0  +  ...  converges  to  0. 

(9)  If  every  term  un  is  positive  (or  zero),  then  the  series  ~Xun 
must  either  converge  or  diverge  to  +  oo .  If  it  converges,  its  sum 
must  be  positive  (unless  all  the  terms  are  zero,  when  of  course  its 
sum  is  zero). 

For  sn  is  an  increasing  function  of  n,  according  to  the  definition 
of  §  69,  and  we  can  apply  the  results  of  that  section  to  sn. 


77,  78]  POSITIVE   INTEGRAL   VARIABLE  143 

(10)  If  every  term  un  is  'positive  (or  zero),  then  the  necessary 
and  sufficient  condition  that  the  series  Xun  shoidd  be  convergent  is 
that  it  should  be  possible  to  find  a  number  K  such  that  the  sum  of 
any  number  of  terms  is  less  than  K;  and,  if  K  can  be  so  found,  then 
the  sum  of  the  series  is  not  greater  than  K. 

This  also  follows  at  once  from  §  69.  It  is  perhaps  hardly 
necessary  to  point  out  that  the  theorem  is  not  true  if  the  condition 
that  every  un  is  positive  is  not  fulfilled.     For  example 

1-1+1-1+  ... 
obviously  oscillates,  sn  being  alternately  equal  to  1  and  to  0. 

(11)  If  «!  +  u.2  +  . . .,  vl  +  v2  +  . . .  are  two  series  of  positive  (or 
zero)  terms,  and  the  second  series  is  convergent,  and  if  un  ^  Kvn, 
where  K  is  a  constant,  for  all  values  of  n,  then  the  first  series  is  also 
convergent,  and  its  sum  is  less  than  or  equal  to  that  of  the  second. 

For  if  vx  +  vz  +  . . .  =  t  then  v1-\-v2  +  ...+vn£t  for  all  values  of 
n,  and  so  2i1  +  u.z  +  . . .  +  un  ^  Kt ;  which  proves  the  theorem. 

Conversely,  if  %un  is  divergent,  and  vn  =  Kun,  then  %vn  is 
divergent. 

78.  The  infinite  geometrical  series.  We  shall  now  con- 
sider the  '  geometrical '  series,  whose  general  term  is  un  =  rn~\  In 
this  case 

sn  =  1  +  r  +  r2  +  . . .  +  rn~l  =  (1  -  rn)/(l  -  r), 
except  in  the  special  case  in  which  r  —  1,  when 

sn  =  1  +  1  +  ...  +l=n. 
In  the  last  case  sn  -*■  +  oo  .     In  the  general  case  sn  will  tend  to  a 
limit  if  and  only  if  rn  does  so.     Referring  to  the  results  of  §  72 
we  see  that  the  series  1  +  r  +  r2  +  . . .  is  convergent  and  has  the  sum 
1/(1  —  r)  if  and  only  if  —  1  <  r  <  1. 

If  r  =  1,  then  sn  =  n,  and  so  sn  -^  +  oo  ;  i.e.  the  series  diverges  to 
+  oo.  If  r  =  —  1,  then  sn  =  1  or  sn  =  0  according  as  n  is  odd  or 
even:  i.e.  sn  oscillates  finitely.  If  r  <  —  1,  then  sn  oscillates  infinitely. 
Thus,  to  sum  up,  ilte  series  1  +  r  +  r2  +  . . .  diverges  to+ccifr^l, 
converges  to  1/(1  —  r)  if  —  1  <  r  <  1,  oscillates  finitely  if  r  =  —  1, 
and  oscillates  infinitely  if  r<  —  1. 

Examples  XXIX.      1.    Recurring  decimals.    The  commonest  example 
of  an  infinite  geometric  series  is  given  by  an  ordinary  recurring  decimal. 


144  LIMITS   OF   FUNCTIONS   OF  A  [iV 

Consider,  for  example,  the  decimal  '21713      This  stands,  according  to  the 
ordinary  rules  of  arithmetic,  for 
2        1         7        J_      _3_      J^       3_  217       13   //,       1\       2687 


i/C1-^)- 


10T108      103      104      105^  106^107  '     '       1000  '  105/  V       lO2/       12375' 
The  reader  should  consider  where  and  how  any  of  the  general  theorems  of 
§  77  have  been  used  in  this  reduction. 

2.  Show  that  in  general 

Cl\  <X2 . . .  &m  (X\...  (ln  —  Of  j  Ct2  •  •  •  Ctrl 

•a1a„...ama1a2...au=  yy\..y00...0  ' 

the  denominator  containing  %  9's  and  m  O's. 

3.  Show  that  a  pure  recurring  decimal  is  always  equal  to  a  proper 
fraction  whose  denominator  does  not  contain  2  or  5  as  a  factor. 

4.  A  decimal  with  m  non-recurring  and  n  recurring  decimal  figures  is 
equal  to  a  proper  fraction  whose  denominator  is  divisible  by  2m  or  5m  but  by 
no  higher  power  of  either. 

5.  The  converses  of  Exs.  3,  4  are  also  true.  Let  r—p/g,  and  suppose  first 
that  g  is  prime  to  10.  If  we  divide  all  powers  of  10  by  g  we  can  obtain  at  most 
g  different  remainders.  It  is  therefore  possible  to  find  two  numbers  n1  and 
n2,  where  n2>n1,  such  that  10n'  and  10,i2  give  the  same  remainder.  Hence 
10"'  -  10"2=  10"2  (10"1-"2- 1)  is  divisible  by  g,  and  so  10n-  1,  where  w,=n1-%> 
is  divisible  by  g.  Hence  r  may  be  expressed  in  the  form  Pj(lOn  —  1),  or  in  the 
form 

P_  J^_ 
10"  +  lO2™  +  ""' 
i.e.  as  a  pure  recurring  decimal  with  n  figures.  If  on  the  other  hand  g  =  2a5^  Q, 
where  Q  is  prime  to  10,  and  m  is  the  greater  of  a  and  /3,  then  10mr  has  a  de- 
nominator prime  to  10,  and  is  therefore  expressible  as  the  sum  of  an  integer 
and  a  pure  recurring  decimal.  But  this  is  not  true  of  lO'V,  for  any  value  of 
fjL  less  than  m  ;  hence  the  decimal  for  r  has  exactly  m  non-recurring  figures. 

6.  To  the  results  of  Exs.  2 — 5  we  must  add  that  of  Ex.  i.  3.  Finally,  if 
we  observe  that 

10^102^103^""       ' 
we  see  that  every  terminating  decimal  can  also  be  expressed  as  a  mixed 
recurring  decimal  whose  recurring  part   is  composed  entirely  of  9's.     For 
example,   -217  =  "2169.     Thus   every  proper  fraction  can   be  expressed  as  a 
recurring  decimal,  and  conversely. 

7.  Decimals  in  general.  The  expression  of  irrational  numbers  as 
non-recurring  decimals.  Any  decimal,  whether  recurring  or  not,  corresponds 
to  a  definite  number  between  0  and  1.  For  the  decimal  •a1a-2a2ai...  stands 
for  the  series 

«i_  ,   <H_  ,   «s   , 

l6i"l02i"l03*1"•,•, 


78]  POSITIVE   INTEGRAL   VARIABLE  145 

Since  all  the  digits  ar  are  positive,  the  sum  sn  of  the  first  n  terms  of  this 
series  increases  with  n,  and  it  is  certainly  not  greater  than  -9  or  1.  Hence 
sn  tends  to  a  limit  between  0  and  1. 

Moreover  no  two  decimals  can  correspond  to  the  same  number  (except  in 
the  special  case  noticed  in  Ex.  6).  For  suppose  that  ■a1a.2a3 ...,  -b^b.jb^...  are 
two  decimals  which  agree  as  far  as  the  figures  ar_i,  br_1,  while  ar>br. 
Then  ar>br  +  l>br  ■  br  +  ibr  +  2...  (unless  6r  +  1,  6r+2,  ...  are  all  9's),  and  so 

■a1a2...arar  +  1 ...  >  -bib2...  brbri.x  ... . 

It  follows  that  the  expression  of  a  rational  fraction  as  a  recurring  decimal 
(Exs.  2 — 6)  is  unique.  It  also  follows  that  every  decimal  which  does  not 
recur  represents  some  irrational  number  between  0  and  1.  Conversely,  any 
such  number  can  be  expressed  as  such  a  decimal.  For  it  must  lie  in  one  of 
the  intervals 

0,     1/10;     1/10,     2/10;     ...  ;     9/10,      1. 

If  it  lies  between  r/10  and  (r+ 1)/10,  then  the  first  figure  is  r.  By  subdividing 
this  interval  into  10  parts  we  can  determine  the  second  figure;  and  so  on. 
But  (Exs.  3,  4)  the  decimal  cannot  recur.  Thus,  for  example,  the  decimal 
1-414...,  obtained  by  the  ordinary  process  for  the  extraction  of  J 2,  cannot 
recur. 

8.  The  decimals  -1010010001000010...  and  -2020020002000020...,  in 
which  the  number  of  zeros  between  two  l's  or  2's  increases  by  one  at  each 
stage,  represent  irrational  numbers. 

9.  The  decimal  -11101010001010...,  in  which  the  nth  figure  is  1  if  n  is 
prime,  and  zero  otherwise,  represents  an  irrational  number.  [Since  the 
number  of  primes  is  infinite  the  decimal  does  not  terminate.  Nor  can  it 
recur :  for  if  it  did  we  could  determine  m  and  p  so  that  m,  m  +p,  m  +  2p, 
m  +  3p, ...  are  all  prime  numbers  ;  and  this  is  absurd,  since  the  series  includes 
m  +  mp.]* 

Examples  XXX.  1.  The  series  rm+rm+1  +  ...  is  convergent  if  -1  <r<  1, 
and  its  sum  is  1/(1  -r)-l  -r-  ...  -r"'-1(§  77,  (2)). 

2.  The  series  rm  +  rm  +  1+  ...  is  convergent  if  -  1<  r  <  1,  and  its  sum  is 
r'n/(l  -  r)  (§  77,  (4)).    Verify  that  the  results  of  Exs.  1  and  2  are  in  agreement. 

3.  Prove  that  the  series  l  +  2r+2;-2+  ...  is  convergent,  and  that  its  sum 
is  (l-H-)/(l-r),  (a)  by  writing  it  in  the  form  -  l  +  2(l+r  +  ?-2+ ...),  (/3)  by 
writing  it  in  the  form  1  +2  (r  +  r2+ ...),  (y)  by  adding  the  two  series 
l+r  +  r2+  ...,  r+r2+  ....  In  each  case  mention  which  of  the  theorems  of 
§  77  are  used  in  your  proof. 

*  All  the  results  of  Exs.  xxix  may  be  extended,  with  suitable  modifications,  to 
decimals  in  any  scale  of  notation.  For  a  fuller  discussion  see  Bromwich,  Infinite 
Series,  Appendix  I. 

H.  10 


146  LIMITS   OF   FUNCTIONS   OF  A  [IV 

4.  Prove  that  the  'arithmetic'  series 

a  +  (a  +  b)  +  (a  +  2b)+... 

is  always  divergent,  unless  both  a  and  b  are  zero.  Show  that,  if  b  is  not 
zero,  the  series  diverges  to  +00  or  to  —  00  according  to  the  sign  of  &,  while  if 
6  =  0  it  diverges  to  +00  or  —  00  according  to  the  sign  of  a. 

5.  What  is  the  sum  of  the  series 

(l-r)  +  (r-r")  +  (fa-r8)+„* 

when  the  series  is  convergent  ?  [The  series  converges  only  if  -l<r:gl.  Its 
sum  is  1,  except  when  ?•  =  ],  when  its  sum  is  0.] 


r2  ,.2 

6.  Sum  the  series  r2  + 9+ ,.  ,    ...„  +  ....     [The  series  is  always  con- 

1+H      (l  +  r-y  L 

vergent.     Its  sum  is  1  +r2,  except  when  r=0,  when  its  sum  is  0.] 

7.  If  we  assume  that  1  +  r +r2  +  . ..  is  convergent  then  we  can  prove  that  its 
sum  is  1/(1  —  r)  by  means  of  §  77,  (1)  and  (4).     For  if  l+r  +  r2+  ...  =s  then 

s  =  l+r  (l  +  r2  +  ...)  =  l+rs. 

v  v 

8.  Sum  the  series  r+  — h  — - — r;,  +  ... 

l+r     (l+r)2 

when  it  is  convergent.  [The  series  is  convergent  if  —  l<l/(l+r)<l,  i.e.  if 
r<  -  2  or  if  r>0,  and  its  sum  is  1  +r.  It  is  also  convergent  when  r=0,  when 
its  sum  is  0.] 

9.  Answer  the  same  question  for  the  series 


l+r  '  (l+r)2    '"'  l-r  +  (l-r)« 


»'-i— +  fnr^2-->       J,+  TZ_T.  +  n—^+-. 


i-r+7-Hrr^ 


l-r\l-r) 


10.  Consider  the  convergence  of  the  series 

(l+r)  +  (r2  +  ?-?)  +  ...,     (l+r+r*)+(r3+f*+r6)  +  ..., 

l-2r+r2  +  r3-2r4+?-5  +  ...,     (1  -2r  +  r2)  +  (r3-2ri  +  r5)  + ... , 
and  find  their  sums  when  they  are  convergent. 

11.  If  0^«,(<1  then  the  series  a0  +  °V+ «2?*2+  •••  is  convergent  for 
0£r<l,  and  its  sum  is  not  greater  than  1/(1 -r). 

12.  If  in  addition  the  series  ao  +  ai  +  a2  + ...  is  convergent,  then  the  series 
a0  +  a1r+a2r2  +  ...  is  convergent  for  O^r^l,  and  its  sum  is  not  greater  than 
the  lesser  of  a0  +  «i  +  a2  +  ...  and  1/(1—  r). 

13.  The  series  1+I+T~2+  1    2   3  +  '" 
is  convergent.     [For  1/(1 .  2...?i)  <  1/2""1.] 


78,  79]  POSITIVE   INTEGRAL   VARIABLE  147 

14.  The  series 

1+T72  +  1.2.3.4+-'      I  +  r70  +  1.2.3.4."5+- 

are  convergent. 

15.  The  general  harmonic  series 

a     a+6     a+26    -  "' 

where  a  and  b  are  positive,  diverges  to  +  oo . 

[For  «„=l/(a+n6)>l/{«(a  +  6)}.  Now  compare  with  1 +£+£+....] 

16.  Show  that  the  series 

is  convergent  if  and  only  if  un  tends  to  a  limit  as  «->•<». 

17.  If  Ui  +  u2  +  ii3  +  ...   is   divergent  then   so  is  any  series  formed  by 
grouping  the  terms  in  brackets  in  any  way  to  form  new  single  terms. 

18.  Any  series,  formed  by  taking  a  selection  of  the  terms  of  a  convergent 
series  of  positive  terms,  is  itself  convergent. 

79.  The  representation  of  functions  of  a  continuous 
real  variable  by  means  of  limits.  In  the  preceding  sections 
we  have  frequently  been  concerned  with  limits  such  as 

lim  <f>n  (x), 

and  series  such  as 

ax  (x)  +  u2(%)  + ...  =  lim  {ih (x)  +  u2  (x)  +  ...  +  w„  (so)}, 

in  which  the  function  of  n  whose  limit  we  are  seeking  involves, 
besides  n,  another  variable  x.  In  such  cases  the  limit  is  of  course 
a  function  of  x.     Thus  in  §  75  we  encountered  the  function 

fix)  =  lim  n  (\/x  —  1): 

and  the  sum  of  the  geometrical  series  1  +  x  +  x-  +  . . .  is  a  function 
of  x,  viz.  the  function  which  is  equal  to  1/(1  —  x)  if  —  1  <  x  <  1  and 
is  undefined  for  all  other  values  of  x. 

Many  of  the  apparently  '  arbitrary '  or  '  unnatural '  functions 
considered  in  Ch.  II  are  capable  of  a  simple  representation  of  this 
kind,  as  will  appear  from  the  following  examples. 

10—2 


148  LIMITS   OF   FUNCTIONS   OF   A  [iV 

Examples  XXXI.  1.  <])n(x)=x.  Here  n  does  not  appear  at  all  in  the 
expression  of  $„  (x),  and  $  (#)=lim  $n  (x)=x  for  all  values  of  x. 

2.  (f>n  (x) = xjn.     Here  <f>  (x)  =  lim  $n  (a?)  =  0  for  all  values  of  x. 

3.  $n(x)=nx.  If  x>0,  <pn(x)-^  +  <x>  ;  if  ,r<0,  (f>n(x)-*--cc  :  only  when 
.(,  =  0  has  (f)n(x)  a  limit  (viz.  0)  as  »-*-oo.  Thus  <£(a)  =  0  when  #=0  and  is 
not  defined  for  any  other  value  of  x. 

4.  cf)n(x)  =  lj?ix,  nx/{nx  +  l). 

5.  <pn(x)  =  xn.  Here  #(.r)  =  0,  (-K.r<l) ;  cj>(x)  =  l,  (x=l);  and  $  (#) 
is  not  defined  for  any  other  value  of  x. 

6.  <j)n(z)  =  xn(l-x).  Here  <£(**)  differs  from  the  <j>(x)  of  Ex.  5  in  that 
it  has  the  value  0  when  x—1. 

7.  4>n  (#)  =  #"/*&•  Here  <fi  (x)  differs  from  the  (f>  (#)  of  Ex.  6  in  that  it  has 
the  value  0  when  x=  —  1  as  well  as  when  x=l. 

8.  <f)n  (x)  =  «•/(*■  + 1 ).     [0  (a?) = 0,  ( -  Kx<l) ;  <f>  (a?) = £,  (a?=  1) ;  $  (x)  =  1 , 

(x<  —  1  or  a?>l) ;  and  $(.£)  is  not  defined  when  x=  —  1.] 

9.  0,  (#)=*"/(*»- 1),    1/(^+1),     1/(^-1),    l/(^n  +  x-n),     lj(x"-x-»). 

10.  0ft (#)  =  (#"-].)/(#» +1),  (H.r»-l)/(?u-n  +  l),  (#»-»)/(#»+»).  [In  the 
first  case  $  (#)=  1  when  |#|>1,  $  (#)  =  -  1  when  \x\  <1,  <j)(x)=0  when  .r=l 
and  0  (#)  is  not  defined  when  a  =  —  1.  The  second  and  third  functions  differ 
from  the  first  in  that  they  are  defined  both  when  x=l  and  when  x—  —1 :  the 
second  has  the  value  1  and  the  third  the  value  -  1  for  both  these  values  of  x.~] 

11.  Construct  an  example  in  which  cj)(x)  =  l,  (|#|>1);  <£(o;)=-l, 
(|#|<1);  and  0(.^)  =  O,  (#=1  and  x=  —  1). 

12.  <j)n  {x)  =  x  {(x2n-  l)l(x2n+ 1)} 2,     nj(xn  +  x-n  +  n). 

13.  0n  (x)  =  {*"/(*)  +  g  (*)}/(*"  +  1).  [Here  0  (x)  =/  (x),  (|  x  \  >  1) ; 
<$>(x)=g(x),  (|#|<1);  <p  (x)  — %{/(%)+ 9 (x)}>  (*=1)  >  and  $(#)  is  undefined 
when  x=  —  1.] 

14.  0n(A-)  =  (2/7r)  arc  tan  (?«?).  [0(a?)  =  l,  (#>0);  0(.r)  =  O,  (a?=0); 
(f)(x)=  —  1,  (a'<0).  This  function  is  important  in  the  Theory  of  Numbers, 
and  is  usually  denoted  by  s^?i  x.~\ 

15.  0n  (#)  =  sin  nxn.  [(f>(x)  =  Q  when  #  is  an  integer;  and  (f>(x)  is 
otherwise  undefined  (Ex.  xxiv.  7).] 

16.  If  $„  (#)  =  sin  («! ,T7r)  then  cf)(x)  =  0  for  all  rational  values  of  x  (Ex. 
xxiv.  14).    [The  consideration  of  irrational  values  presents  greater  difficulties.] 

17.  4>n  (x)  =  (cos2  xn)n.  [(f)(x)  =  0  except  when  x  is  integral,  when 
0(.r)  =  l.] 

18.  If  i\r  >  1752  then  the  number  of  days  in  the  year  iV  a.d.  is 

lim  {365  +  (cos2  J^tt)"  -  (cos2  TfoiV»r)»  +  (cos2  jfoxVir)"}. 


79-81]  POSITIVE  INTEGRAL  VARIABLE  14!) 

80.  The  bounds  of  a  bounded  aggregate.  Let  S  be  any  system  or 
aggregate  of  real  numbers  s.  If  there  is  a  number  K  such  that  s<  K  for 
every  s  of  S,  we  say  that  S  is  bounded  above.  If  there  is  a  number  k  such  that 
s  >  h  for  every  s,  we  say  that  S  is  bounded  below.  If  S  is  both  bounded  above 
and  bounded  below,  we  say  simply  that  S  is  bounded. 

Suppose  first  that  S  is  bounded  above  (but  not  necessarily  below).  There 
will  be  an  infinity  of  numbers  which  possess  the  property  possessed  by  A'  ; 
any  number  greater  than  A,  for  example,  possesses  it.  We  shall  prove  that 
among  these  numbers  there  is  a  least*,  which  we  shall  call  M.  This  number  J/ 
is  not  exceeded  by  any  member  of  S,  but  every  number  less  than  M  is  exceeded 
by  at  least  one  member  of  >S. 

We  divide  the  real  numbers  |  into  two  classes  L  and  R,  putting  £  into  L  or 
R  according  as  it  is  or  is  not  exceeded  by  members  of  S.  Then  every  £  belongs 
to  one  and  one  only  of  the  classes  L  and  R.  Each  class  exists  ;  for  any 
number  less  than  any  member  of  S  belongs  to  L,  while  K  belongs  to  R. 
Finally,  any  member  of  L  is  less  than  some  member  of  S,  and  therefore  less 
than  any  member  of  R.  Thus  the  three  conditions  of  Dedekind's  Theorem 
(§  17)  are  satisfied,  and  there  is  a  number  M  dividing  the  classes. 

The  number  M  is  the  number  whose  existence  we  had  to  prove.  In  the 
first  place,  M  cannot  be  exceeded  by  any  member  of  S.  For  if  there  were  such 
a  member  s  of  S,  we  could  write  s=  M+rj,  where  rj  is  positive.  The  number 
M-\-\rj  -would  then  belong  to  L,  because  it  is  less  than  s,  and  to  R,  because  it  is 
greater  than  M ;  and  this  is  impossible.  On  the  other  hand,  any  number  less 
than  M  belongs  to  Z,  and  is  therefore  exceeded  by  at  least  one  member  of  S. 
Thus  31  has  all  the  properties  required. 

This  number  M  we  call  the  upper  bound  of  »S",  and  we  may  enunciate  the 
following  theorem.  Any  aggregate  S  which  is  bounded  above  has  an  upper 
bound  M.  No  member  of  S  exceeds  M ;  but  any  number  less  than  M  is  exceeded 
by  at  least  one  member  of  S. 

In  exactly  the  same  way  we  can  prove  the  corresponding  theorem  for  an 
aggregate  bounded  below  (but  not  necessarily  above).  Any  aggregate  S  which 
is  bounded  below  has  a  lower  bound  m.  No  member  of  S  is  less  than  m;  but 
there  is  at  least  one  member  of  S  \ohich  is  less  than  any  number  greater  than  m. 

It  will  be  observed  that,  when  S  is  bounded  above,  M  <  A',  and  when  S  is 
bounded  below,  m  >  k.     When  S  is  bounded,  k  i  m  g  M  <  K. 

81.  The  bounds  of  a  bounded  function.  Suppose  that  0  (n)  is  a  func- 
tion of  the  positive  integral  variable  n.  The  aggregate  of  all  the  values  0  (n) 
defines  a  set  S,  to  which  we  may  apply  all  the  arguments  of  §  80.  If  S  is 
bounded  above,  or  bounded  below,  or  bounded,  we  say  that  $  (n)  is  bounded 

*  An  infinite  aggregate  of  numbers  does  not  necessarily  possess  a  least  member. 
The  set  consisting  of  the  numbers 

11  1 


*'    >'    f 


for  example,  has  no  least  member. 


150  LIMITS   OF    FUNCTIONS   OF   A  [iV 

above,  or  bounded  below,  or  bounded.  If  $  (»)  is  bounded  above,  that  is  to 
say  if  there  is  a  number  K  such  that  <£  (n)  <  K  for  all  values  of  n,  then  there 
is  a  number  M  such  that 

(i)  <f>  (n)  5|  M  for  all  values  of  n  ; 

(ii)  if  8  is  any  positive  member  then  §  (n)  >  M—  8  for  at  least  one  value  of  n. 
This  number  M  we  call  the  upper  bound  of  $  (n).  Similarly,  if  (f>  (n)  is 
bounded  below,  that  is  to  say  if  there  is  a  number  k  such  that  (f>(n)  ^k  for  all 
values  of  n,  then  there  is  a  number  m  such  that 

(i)  <p  (n)  ^  m  for  all  values  of  n  ; 

(ii)  if  8  is  any  positive  number  then  $  (n)  <  m  +  8  for  at  least  one  value  of  n. 
This  number  m  we  call  the  lower  bound  of  cf>  («). 

If  K  exists,  M^K ;  if  k  exists,  m  zik;  and  if  both  k  and  K  exist  then 
k<m<M<K. 

82.  The  limits  of  indetermination  of  a  bounded  function.  Suppose 
that  <p  (n)  is  a  bounded  function,  and  M  and  m  its  upper  and  lower  bounds. 
Let  us  take  any  real  number  £,  and  consider  now  the  relations  of  inequality 
which  may  hold  between  £  and  the  values  assumed  by  <£  (»)  for  large  values 
of  n.     There  are  three  mutually  exclusive  possibilities  : 

(1)  $  ^  (f)  (n)  for  all  sufficiently  large  values  of  n ; 

(2)  $  ^4>  (n)  for  all  sufficiently  large  values  of  n ; 

(3)  £  <  <$>  (n)  for  an  infinity  of  values  of  n,  and  also   £  >  0  (n)  for  an 

infinity  of  values  of  a. 

In  case  (1)  we  shall  say  that  £  is  a  superior  number,  in  case  (2)  that  it  is 
an  inferior  number,  and  in  case  (3)  that  it  is  an  intermediate  number.  It  is 
plain  that  no  superior  number  can  be  less  than  m,  and  no  inferior  number 
greater  than  M. 

Let  us  consider  the  aggregate  of  all  superior  numbers.  It  is  bounded 
below,  since  none  of  its  members  are  less  than  rn,  and  has  therefore  a  lower 
bound,  which  we  shall  denote  by  A.  Similarly  the  aggregate  of  inferior 
numbers  has  an  upper  bound,  which  we  denote  by  X. 

We  call  A  and  X  respectively  the  upper  and  lower  limits  of  indetermination 
of  <f>  (n)  as  n  tends  to  infinity  ;  and  write 

A  =  lim  $  («),     X  =  lim  <f>  (n). 
These  numbers  have  the  following  properties  : 

(1)  «<)igAiI; 

(2)  A  and  X  are  the  upper  and  lower  bounds  of  the  aggregate  of  intermediate 
numbers,  if  any  such  exist  ; 

(3)  if  8  is  any  positive  number,  then  <f>  (n)  <  A  +  S  for  all  sufficiently  large 
values  of  n,  and  cf)(n)>  A  —  8  for  an  infinity  of  values  of  n  ; 

(4)  similarly  <f>  (n)  >  X  -  8  for  all  sufficiently  large  values  of  n,  and 
(f>(n)<\  +  8  for  an  infinity  of  values  of  n\ 


81,  82]  POSITIVE    INTEGRAL   VARIABLE  151 

(5)  the  necessary  and  sufficient  condition  that  <£  (n)  should  tend  to  a  limit 
is  that  A  =  X,  and  in  this  case  the  limit  is  I,  the  common  value  of  X  and  A. 

Of  these  properties,  (1)  is  an  immediate  consequence  of  the  definitions  ; 
and  we  can  prove  (2)  as  follows.  If  A=X  =  ?,  there  can  be  at  most  one  inter- 
mediate number,  viz.  I,  and  there  is  nothing  to  prove.  Suppose  then  that 
A  >  X.  Any  intermediate  number  £  is  less  than  any  superior  and  greater  than 
any  inferior  number,  so  that  X  ^  £  <  A.  But  if  X  <  £  <  A  then  £  must  be 
intermediate,  since  it  is  plainly  neither  superior  nor  inferior.  Hence  there  are 
intermediate  numbers  as  near  as  we  please  to  either  X  or  A. 

To  prove  (3)  we  observe  that  \  +  8  is  superior  and  A- 8  intermediate  or 
inferior.  The  result  is  then  an  immediate  consequence  of  the  definitions ;  and 
the  proof  of  (4)  is  substantially  the  same. 

Finally  (5)  may  be  proved  as  follows.     If  A  =  \  =  l,  then 

l-8«t>(n)<l+b 

for  every  positive  value  of  8  and  all  sufficiently  large  values  of  n,  so  that 
0  (n)  ->- 1.  Conversely,  if  <£  (n)  -*■  I,  then  the  inequalities  above  written  hold 
for  all  sufficiently  large  values  of  n.  Hence  1-8  is  inferior  and  1  +  8  superior, 
so  that 

\>l-8,     \^l  +  8, 

and  therefore  A  —  X  ^  28.     As  A  —  X  21 0,  this  can  only  be  true  if  A  =  X. 

Examples  XXXII.  1.  Neither  A  nor  X  is  affected  by  any  alteration  in 
any  finite  number  of  values  of  <fi  (n). 

2.  If  <fi  (n)  —  a  for  all  values  of  n,  then  m  =  X  =  A  =  M=  a. 

3.  If  <f)  (n)  =  l/n,  then  m=X=A=0  and  M=l. 

4.  If  (j>  iii)  =  ( - 1  )n,  then  m  =  X  =  - 1  and  A  =  M =  1 . 

5.  If  <£(?*)  =  (- l)nM  then  m=-l,  X  =  A  =  0,  Jf=|. 

6.  If  0(?O  =  (- 1)"{1  +  (l/n)},  then  m= -2,  X=-l,  A=l,  if=f. 

7.  Let  </>  (n)  =  sin  n6ir,  where  6>  0.  If  6  is  an  integer  then  m  =  X  =  A  =  M=  0. 
If  6  is  rational  but  not  integral  a  variety  of  cases  arise.  Suppose,  e.g.,  that 
B=plq,  p  and  q  being  positive,  odd,  and  prime  to  one  another,  and  q>\. 
Then  $(?i)  assumes  the  cyclical  sequence  of  values 

smipn/q),     sin {2pwjq),  ,     sin{(2g--  l)pn/q},     sin  (2qpirjq\  

It  is  easily  verified  that  the  numerically  greatest  and  least  values  of  <p  (n)  are 
cos(7r/2^)  and  —cos{irj2q),  so  that 

m  =  \=  -cos(7r/2<7),     A=J/=cos(7r/2g-). 
The  reader  may  discuss  similarly  the  cases  which  arise  when  p  and  q  are 
not  both  odd. 

The  case  in  which  6  is  irrational  is  more  difficult :  it  may  be  shown  that 
in  this  case  m  =  \=  -  1  and  A  =  M=l.  It  may  also  be  shown  that  the  values 
of  </>  (n)  are  scattered  all  over  the  interval  ( - 1,  1 )  in  such  a  way  that,  if  £  is 


152                                   LIMITS   OF   FUNCTIONS   OF   A  [iV 

any  number  of  the  interval,  then  there  is  a  sequence  Wj,  n2,  such  that 

The  results  are  very  similar  when  <f)  (n)  is  the  fractional  part  of  nO. 

83.  The  general  principle  of  convergence  for  a  bounded  function. 

The  results  of  the  preceding  sections  enable  us  to  formulate  a  very  important 
necessary  and  sufficient  condition  that  a  bounded  function  cf>  (n)  should  tend 
to  a  limit,  a  condition  usually  referred  to  as  the  general  principle  of  convergence 
to  a  limit. 

Theorem  1.  The  necessary  and  sufficient  condition  that  a  bounded  function 
cf)  (n)  should  tend  to  a  limit  is  that,  when  any  positive  number  8  is  given,  it  should 
be  possible  to  find  a  number  n0  (8)  such  that 

l0(«»)-*(»h)l<8 
for  all  values  of  %  and  n2  such  that  n2  >  n\  ^  n0  (8). 

In  the  first  place,  the  condition  is  necessary.  For  if  ${n)-*-l  then  we 
can  find  n0  so  that 

l-±8<(p(n)<l  +  ho- 
when  n  >  n0 ,  and  so 

|0(»2)-0("l)l<S (1) 

when  nx  >  n0  and  n2  =  no- 

In  the  second  place,  the  condition  is  sufficient.  In  order  to  prove  this  we 
have  only  to  show  that  it  involves  X  =  A.  But  if  X  <  A  then  there  are,  however 
small  8  may  be,  infinitely  many  values  of  n  such  that  4>(?i)<\  +  8  and 
infinitely  many  such  that  cf>(n)>  A-8  ;  and  therefore  we  can  find  values  of 
ni  and  n2,  each  greater  than  any  assigned  number  «0,  and  such  that 

0(w2)-0(%1)>A-X-28, 
which  is  greater  than  i(A-X)  if  8  is  small  enough.     This  plainly  contradicts 
the  inequality  (1).     Hence  X  =  A,  and  so  (f>  (n)  tends  to  a  limit. 

84.  Unbounded  functions.  So  far  we  have  restricted  ourselves  to 
bounded  functions  ;  but  the  '  general  principle  of  convergence '  is  the  same 
for  unbounded  as  for  bounded  functions,  and  the  words  'a  bounded  functio7i' 
may  be  omitted  from  the  enunciation  of  Theorem  1. 

In  the  first  place,  if  <p  (n)  tends  to  a  limit  I  then  it  is  certainly  bounded  ;  for 
all  but  a  finite  number  of  its  values  are  less  than  1  +  8  and  greater  than  1-8. 
In  the  second  place,  if  the  condition  of  Theorem  1  is  satisfied,  we  have 

|0O2)-<M'"i)|<s 

whenever  /!,  >k0  and  «2="o-  Let  us  choose  some  particular  value  nx  greater 
than  n0.     Then 

0  («!)  -  8  <  (j)  (n2)  <  0  (%)  +  8 

when  n2  2in0.  Hence  (f>  (n)  is  bounded ;  and  so  the  second  part  of  the  proof  of 
the  last  section  applies  also. 

*  A  number  of  simple  proofs  of  this  result  are  given  by  Hardy  and  Littlewood, 
"  Some  Problems  of  Diophantine  Approximation",  Acta  Mathematica,  vol.  xxxvii. 


83-85]  POSITIVE  INTEGRAL   VARIABLE  153 

The  theoretical  importance  of  the  '  general  principle  of  convergence '  can 
hardly  be  overestimated.  Like  the  theorems  of  §  69,  it  gives  us  a  means  of 
deciding  whether  a  function  <f>(7i)  tends  to  a  limit  or  not,  without  requiring 
us  to  be  able  to  tell  beforehand  what  the  limit,  if  it  exists,  must  be  ;  and 
it  has  not  the  limitations  inevitable  in  theorems  of  such  a  special  character 
as  those  of  §  69.  But  in  elementary  work  it  is  generally  possible  to  dispense 
with  it,  and  to  obtain  all  we  want  from  these  special  theorems.  And  it  will 
be  found  that,  in  spite  of  the  importance  of  the  principle,  practically  no 
applications  are  made  of  it  in  the  chapters  which  follow.*  We  will  only 
remark  that,  if  we  suppose  that 

4>  in)  =  s)l  =  ul  +  u2  +  ...  +  un , 

we  obtain  at  once  a  necessary  and  sufficient  condition  for  the  convergence  of 
an  infinite  series,  viz  : 

Theorem  2.  The  necessary  and  sufficient  condition  for  the  convergence 
of  the  series  ■m1-H<2  +  ---  &  that,  given  any  positive  number  8,  it  should  be 
possible  to  find  n0  so  that 

for  all  values  of  nx  and  n2  such  that  ?i2>n1>7i0. 

85.  Limits  of  complex  functions  and  series  of  complex 
terms.  In  this  chapter  we  have,  up  to  the  present,  concerned 
ourselves  only  with  real  functions  of  n  and  series  all  of  whose 
terms  are  real.  There  is  however  no  difficulty  in  extending  our 
ideas  and  definitions  to  the  case  in  which  the  functions  or  the 
terms  of  the  series  are  complex. 

Suppose  that  <p(ri)  is  complex  and  equal  to 

p  (n)  +  i<r  (n), 

where  p  (n),  cr  (n)  are  real  functions  of  n.  Then  if  p  (n)  and  a  (n) 
converge  respectively  to  limits  r  and  s  as  n  -»-  go  ,  ive  shall  say  that 
<j>  (n)  converges  to  the  limit  I  =  r  +  is,  and  write 

lim  $  (n)  =  I. 

Similarly,  when  nn  is  complex  and  equal  to  vn  +  nvn,  we  shall  say 
that  the  series 

Ml  +  «2  +U3+  ... 

is  convergent  and  has  the  sum  I  =  r  +  is,  if  the  series 
v1  +  v,  +  v3+  ...,     wJ  +  iu,  +  ws+  ... 
are  convergent  and  have  the  sums  r,  s  respectively. 

*  A  few  proofs  given  in  Ch.  VIII  can  be  simplified  by  the  use  of  the  principle. 


154  LIMITS   OF   FUNCTIONS   OF   A  [iV 

To  say  that  ut  +  ?/2  +  u3  +  ...  is  convergent  and  has  the  sum 
I  is  of  course  the  same  as  to  say  that  the  sum 

Sn  =  U1  +  Ui+  ...  +lln  =  (v1+V2  +  ...+Vn)  +  i(lV1+W2  +  ...  +  Wn) 

converges  to  the  limit  I  as  n  -*-  go  . 

In  the  case  of  real  functions  and  series  we  also  gave  definitions 
of  divergence  and  oscillation,  finite  or  infinite.  But  in  the  case 
of  complex  functions  and  series,  where  we  have  to. consider  the 
behaviour  both  of  p  (n)  and  of  <r  (n),  there  are  so  many  possibilities 
that  this  is  hardly  worth  while.  When  it  is  necessary  to  make 
further  distinctions  of  this  kind,  we  shall  make  them  by  stating 
the  way  in  which  the  real  or  imaginary  parts  behave  when  taken 
separately. 

86.  The  reader  will  find  no  difficulty  in  proving  such 
theorems  as  the  following,  which  are  obvious  extensions  of 
theorems  already  proved  for  real  functions  and  series. 

(1)  If  lim  <p  (n)  =  I  then  lim  <£  (n  +  p)  =  £  for  any  fixed  value 
of  p. 

(2)  If  wa  +  u2  +  . . .  is  convergent  and  has  the  sum  I,  then 
a  -f  b  +  c  +  . . .  +  k  +  Uj  +  v2  +  . . .  is  convergent  and  has  the  sum 
a  +  b  +  c+  ...  +  k  +  1,  and  up+l  +  up+2  +  ...  is  convergent  and  has 
the  sum  I  —  1^  —  ti2  —  ...  —  up. 

(3)  If  lim  0  (n)  =  I  and  lim  yjr  (n)  =  m,  then 

lim  {(f>  (n)  +  ty  (n)}  =1  +  m. 

(4)  If  lim  (p  (n)  =  I,  then  lim  k<f>  (n)  =  hi. 

(5)  If  lim  <p  (11)  =  I  and  lim  \jr  (n)=m,  then  lim  <p(n)yjs(n)=lm. 

(6)  If  w1  +  wa+ ...  converges  to  the  sum  I,  and  v1  +  v2+...  to 
the  sum  m,  then  (ux  +  v^  +  (u2+ v2)  + . . .  converges  to  the  sum  l  +  m. 

(7)  If  Wj  +  m2  +  . ..  converges  to  the  sum  I  then  ki^  +  ku2  +  ... 
converges  to  the  sum  kl. 

(8)  If  Wj  +  u2+  u3  +  ...  is  convergent  then  lim  un  =  0. 

(9)  If  «2  +  Ms  +  us  +  ...  is  convergent,  then  so  is  any  series 
formed  by  grouping  the  terms  in  brackets,  and  the  sums  of  the  two 
series  are  the  same. 


85,  86]  POSITIVE   INTEGRAL   VARIABLE  155 

As  an  example,  let  us  prove  theorem  (5).     Let 

0(»)=p(»)+*V(»)j     M»)=p' (»)+&/(»),     l  =  r  +  is,     m=r'  +  is'. 

Then  p(ri)-*-r,     a(n)-**s,     p'(n)-^r',     o-'  (»)-*-*'. 

But  (p  (n)  \js  (n)  =  pp  -  (to-'  +  i  (pa'  +  p'v), 

and  pp'  -  cro-'  -*-  r/  -  ss',    p<r'  + p'<r  -*-  rs'  +  r's  ; 

so  that  <p  (n)  \|/-  (n)  ->-  rr'  —  ss'  +  i  (rs'  +  r's), 

i.e.  <p  (n)  yjr  (n) -*- (r + is)  (?•'  +  is')  =  Im. 

The  following  theorems  are  of  a  somewhat  different  character. 

(10)  In   order  that  <f>  (n)  =  p  (n)  +  i<r  (n)  should   converge   to 
zero  as  n  -*-  co ,  it  is  necessary  and  sufficient  that 

\4>  00 1 =s/Up  (*)}'+{*  OOP] 

should  converge  to  zero. 

If  p(n)  and  <r(n)  both  converge  to  zero  then  it  is  plain  that  sf(p2  +  cr'i) 
does  so.  The  converse  follows  from  the  fact  that  the  numerical  value  of  p  or 
cr  cannot  be  greater  than  ,/(p2  +  a-2). 

(11)  More  generally,  in  order  that  <p(n)  should  converge  to  a 
limit  I,  it  is  necessary  and  sufficient  that 

l*00-»l 

should  converge  to  zero. 

For  (p  (n)  —  I  converges  to  zero,  and  we  can  apply  (10). 

(12)  Theorems  1  and  2  of  §§  83 — 84  are  still  true  when 
<p  (n)  and  un  are  complex. 

We  have  to  show  that  the  necessary  and  sufficient  condition  that  cp(n) 
should  tend  to  I  is  that 

I  <P  (%)-<?(%)  I  <8  (l) 

when  n2>  nx  ^.nQ. 

If  <p  (n)-*-l  then  p  (%)-*»»■  and  a  (n)->-s,  and  so  we  can  find  numbers  n0'  and 
n0"  depending  on  8  and  such  that 

I P  W  -  p  Oi)  I  <  i  8,     I  <r  W  -  o-  (»i)  ]  <  i  5, 
the  first  inequality  holding  when  «2  >  ni  =  no'j  and  the  second  when  n2  >  ?ix  ^ «0". 
Hence 

I  <P  (»2)~  <P  (»l)  I  ^  I  P  (%)~P  («l)  I  +  I  *  ("2)  -  O-  («l)  |<8 

when  w2>^i  =  «o>  where  n0  is  the  greater  of  n0'  and  «0".     Thus  the  condition 
(1)  is  necessary.     To  prove  that  it  is  sufficient  we  have  only  to  observe  that 

I  p  (n2)  -  p  (nx)  I  ^  I  <p  (n2)  -  cp  (%)  |  <  3 
when  n2>n1^n0.     Thus  p (n)  tends  to  a  limit  r,  and  in  the  same  way  it  may 
be  shown  that  a  (n)  tends  to  a  limit  s. 


156  LIMITS   OF   FUNCTIONS   OF   A  [iV 

87.  The   limit   of  zn  as   n  -*■  oc ,  z  being   any  complex 

number.  Let  us  consider  the  important  case  in  which  <£  (n)  =  z11. 
This  problem  has  already  been  discussed  for  real  values  of  z  in 

§72. 

If  zn  —  I  then  zn+l  -*  I,  by  (1)  of  §  86.     But,  by  (4)  of  §  86, 

zn+i  =  zzn  _^  2l} 

and  therefore  I  =  zl,  which  is  only  possible  if  (a)  I  =  0  or  (b)  z  ==  1. 
If  z  =  1  then  lim2n=  1.  Apart  from  this  special  case  the  limit, 
if  it  exists,  can  only  be  zero. 

Now  if  z  =  r  (cos  6  +  i  sin  6),  where  r  is  positive,  then 

zn  =  r11  (cos  nd  +  i  sin  n6), 

so  that  |  zn  |  =  rn.  Thus  |  zn  \  tends  to  zero  if  and  only  if  r  <  1 ; 
and  it  follows  from  (10)  of  §  86  that 

lim  z11  =  0 

if  and  only  if  r  <  1.  In  no  other  case  does  zn  converge  to  a  limit, 
except  when  z  =  1  and  zn  -*■  1. 

88.  The    geometric     series     1  +  z  +  z-  +  . . .    when   z    is 
complex.     Since 

sn=l+z  +  z*-+...  +  zn-*  =  (l-zn)/(l-z), 

unless  z  =  1,  when  the  value  of  s}l  is  ??,  it  follows  that  the  series 
1  +  z  +  z2  +  ...  is  convergent  if  and  only  if  r  =  |  z  \  <  1.  And  its 
sum  when  convergent  is  1/(1  —  z). 

Thus  if  z  =  r  (cos  9  +  i  sin  9)  =  ?*  Cis  9,  and  r  <  1,  we  have 

l+*  +  *2+...  =  l/(l-rCis0), 

or      1  +  r  Cis  9  +  r"  Cis  29  +  ...  =  1/(1  -  r  Cis  9) 

=  (1  —  r  cos  0  +  ir  sin  6)1(1  —  2?-  cos  6  +  r2). 

Separating  the  real  and  imaginary  parts,  we  obtain 

1  +  r  cos  9  +  r2  cos  29  +  . . .  =  (1  —  r  cos  6)1(1  -  2r  cos  6  +  r2), 

r  sin  0  +  r2  sin  20  +  . . .  =  r  sin  9/(1  -  2r  cos  0  +  r2), 

provided  r<l.  If  we  change  6  into  0  +  7r,  we  see  that  these 
results  hold  also  for  negative  values  of  r  numerically  less  than  1. 
Thus  they  hold  when  —  1  <  r  <  1. 


87,  88]  POSITIVE    INTEGRAL   VARIABLE  157 

Examples  XXXIII.  1.  Prove  directly  that  cf>  (n)  =  rn  cos  n8  converges 
to  0  when  r<  1  and  to  1  when  r=  1  and  6  is  a  multiple  of  2w.  Prove  further 
that  if  r  —  \  and  6  is  not  a  multiple  of  2n,  then  <p(n)  oscillates  finitely;  if 
r>\  and  6  is  a  multiple  of  2ir,  then  <f>  (n)-*-  +  <x>;  and  if  r>l  and  6  is  not  a 
multiple  of  2n,  then  <£  (n)  oscillates  infinitely. 

2.  Establish  a  similar  series  of  results  for  $  (m)  =  ?•"  sin  nd. 

3.  Prove  that  zm + zm  +  1+  ...  =  zmj(l  - z), 

2»*  +  22'»  +  i  +  2sm  +  2+  ...  =zm(l+z)/(l-z), 
if  and  only  if  j  z  |  <  1.     "Which  of  the  theorems  of  §  86  do  you  use  ? 

4.  Prove  that  if  - 1  <  r  <  1  then 

l  +  2rcos0  +  2r2cos20+...=(l-r2)/(l-2rcos0  +  ?-2). 

5.  The  series  I  +  t^-  +  (V^Y  +••• 

converges  to  the  sum  1  /  ( 1  -  ■=— -  J  =  1  +  z  if  1 2/(1  +  z)  |  <  1.     Show  that  this 
condition  is  equivalent  to  the  condition  that  z  has  a  real  part  greater  than  —\. 

MISCELLANEOUS   EXAMPLES   ON  CHAPTER  IV. 

1.  The  function  $(«)  takes  the  values  1,  0,  0,  0,  1,  0,  0,  0,  1,  ...  when 
n  =  0,  1,  2, ....  Express  (f>(n)  in  terms  of  n  by  a  formula  which  does  not 
involve  trigonometrical  functions.     [4>(n)  =  ^{l  +  (  —  l)"  +  i'l  +  (-z)"}.] 

2.  If  <£  (n)  steadily  increases,  and  \|/-  (a)  steadily  decreases,  as  n  tends  to 
00,  and  if  \|/-  (n)  >  <f)  («)  for  all  values  of  n,  then  both  <j>(n)  and  \^(n)  tend  to 
limits,  and  lim0(?i)^lim\^(n).  [This  is  an  intermediate  corollary  from 
§69.] 

3.  Prove  that,  if 


*«-(l+^,     +{n)=(l-?)~n, 


then  (f)(n  +  l)>cf)  (n)  and  ^  (/i  + 1) < ^  (to).    [The  first  result  has  already  been 
proved  in  §  73.] 

4.  Prove  also  that  ^  (to)  >  (f)  (to)  for  all  values  of  to :  and  deduce  (by  means 
of  the  preceding  examples)  that  both  <£  (to)  and  \js  (to)  tend  to  limits  as  to 
tends  to  oo .  * 

5.  The  arithmetic  mean  of  the  products  of  all  distinct  pairs  of  positive 
integers  whose  sum  is  to  is  denoted  by  Sn.      Show  that  lim  (SJn2)  =  1/6. 

(Math.  Trip.  1903.) 

*  A  proof  that  lim  {\p  (n)  -  <p  (ft)}=0,  and  that  therefore  each  function  tends  to 
the  limit  e,  will  be  found  in  Chrystal's  Algebra,  vol.  ii,  p.  78.  We  shall  however 
prove  this  in  Ch.  IX  by  a  different  method. 


158  LIMITS   OF   FUNCTIONS   OF   A  [iV 

6.     Prove  that  if  Xi  =  ^{x  +  (Ajx)},  x2~\{xl-\-{Ajx\)}1  and  so  on,  x  and 
A  being  positive,  then  lim#n=,vM' 


[Prove  first  that  — - — rT=  I  — - — r.  )    .] 
1  xn+JA      \x+,JAJ 


7.  If  <£  (»)  is  a  positive  integer  for  all  values  of  n,  and  tends  to  oo  with  n, 
then  x  tends  to  0  if  0<x<\  and  to  +oo  if  x>l.  Discuss  the  behaviour 
of  x      ,  as  w-*-oo ,  for  other  values  of  x. 

8.*  If  an  increases  or  decreases  steadily  as  n  increases,  then  the  same  is 
true  of  (ax  +  «2  +  •  •  •  +  an)ln- 

9.  If  xn  +  i  =  s!(k+xn),  and  h  and  xx  are  positive,  then  the  sequence  x\,  x.2, 
x3,  ...  is  an  increasing  or  decreasing  sequence  according  as  xx  is  less  than  or 
greater  than  a,  the  positive  root  of  the  equation  x2  —  x  +  k;  and  in  either  case 
xn-»-a  as  n->~<x> . 

10.  If  snn+i=kj(l+x^),  ar»d  &  and  X\  are  positive,  then  the  sequences 
#!,  x3,  x$,  ...  and  #2>  xa  x§i  •••  are  one  au  increasing  and  the  other  a  decreasing 
sequence,  and  each  sequence  tends  to  the  limit  a,  the  positive  root  of  the 
equation  a?+x  =  k. 

11.  The  function  f(x)  is  increasing  and  continuous  (see  Ch.  V)  for  all 
values  of  x,  and  a  sequence  xlt  x2,  x3,  ...  is  denned  by  the  equation 
xn  +  i=f(xn).  Discuss  on  general  graphical  grounds  the  question  as  to 
whether  xn  tends  to  a  root  of  the  equation  x-f(x).  Consider  in  particular 
the  case  in  which  this  equation  has  only  one  root,  distinguishing  the  cases  in 
which  the  curve  y=f(%)  crosses  the  line  y  =  x  from  above  to  below  and  from 
below  to  above. 

12.  If  Xy ,  x2  are  positive  and  xn  +  j  =  \  (xn+xn  _  i),  then  the  sequences  xx ,  x3, 
.r5,  ...  and  x2,  x\,  x6,  ...  are  one  a  decreasing  and  the  other  an  increasing 
sequence,  and  they  have  the  common  limit  J  (xy  +  2x2). 

13.  Draw  a  graph  of  the  function  y  defined  by  the  equation 

y=  limx*-»Sm^rrx+x^  ^Math   THp  19Q1  } 

14.  The  function  y=  lim t~= — . 

w-*w  l+nsm2Trx 

is  equal  to  0  except  when  x  is  an  integer,  and  then  equal  to  1.     The  function 
lim  ^(x)  +  ncj)(x)sm2nx 
n-*~«>         1+71  sin2  nx 
is  equal  to  <f>  (x)  unless  x  is  au  integer,  and  then  equal  to  \|/-  (x). 

15.  Show  that  the  graph  of  the  function 

..      xn(b(x)  +  x-n\lf(x) 
y=  hm  — J-^-z !-i— 

*  Exs.  8 — 12  are  taken  from  Broniwich's  Infinite  Series. 


POSITIVE   INTEGRAL   VARIABLE  159 

is  composed  of  parts  of  the  graphs  of  cf>(x)  and  \^(x),  together  with  (as  a  rule) 
two  isolated  points.     Is  y  defined  when  (a)  x=l,  (b)  x=  - 1,  (c)  x  —  0  ] 

16.     Prove  that  the  function  y  which  is  equal  to  0  when  x  is  rational,  and 
to  1  when  x  is  irrational,  may  be  represented  in  the  form 

y  =  lim  sgn{sia2(m\  irx)}, 
where  sgn  x—  lim  (2/rr)  arc  tan  (nx)t 

as  in  Ex.  xxxi.  14.  [If  x  is  rational  then  sin2  (m !  nx),  and  therefore 
sgn  {sin2  (m !  nx)},  is  equal  to  zero  from  a  certain  value  of  m  onwards  :  if 
x  is  irrational  then  sin2  (m  !  nx)  is  always  positive,  and  so  sgn  {sin2  (m !  irx)} 
is  always  equal  to  1.] 

Prove  that  y  may  also  be  represented  in  the  form 


1  —  lim   [lim  {cos  (m  !  7i\i)}2"]. 


17.     Sum  the  series 


i  *(i/  +  l)'        iv{»  +  l)...{v+k)' 
[Since 

I =  1/ ! 1  ) 

,      ,  »  1  =  1  f 1 1 } 

We    aVC       iV{v  +  l)...(v+k)     k\l.2...&     (n  +  l)(n  +  2)...{n+k)) 

and  so  2  — ; rr — ; r,  =  ,  /f  ,, .  1 

18.  If  |2|<|aL  then     -A.  =  -^fi  +  £  +  f!  +  ..V 

3-a  a   \         a       a-  / 

andif  |s|>|a|,  then  —  =      -  (l  +  -  +-,  +  ... ). 

1   '      '    '  z  —  a  z  \         z      zl  j 

19.  Expansion  of  (Az  +  B)/(az2  +  2bz  +  c)  in  powers  of  z.  Let  a,  8 
be  the  roots  of  az2  +  2bz  +  c=0,  so  that  az2  +  2bz  +  c  =  a(z-  a){z~8).  We 
shall  suppose  that  A,  B,  a,  b,  c  are  all  real,  and  a  and  8  unequal.  It  is  then 
easy  to  verify  that 

Az+B  1        /Aa  +  B  _  A8  +  B\ 

az2  +  2bz  +  c~  a(a-8)\  z-a  z~8J' 

There  are  two  cases,  according  as  b2  >  ac  or  b2  <  ac. 

(1)  If  b2>ac  then  the  roots  a,  8  are  real  and  distinct.  If  \z\  is  less  than 
either  \a\  or  \8\  we  can  expand  l/(z  — a)  and  lftz—8)  in  ascending  powers  of  s 
(Ex.  18).  If  U  |  is  greater  than  either  |  a  |  or  |  /3 1  we  must  expand  in  descending 
powers  of  z;  while  if  \z\  lies  between  \a\  and  \8\  one  fraction  must  be  ex- 
panded in  ascending  and  one  in  descending  powers  of  z.  The  reader  should 
write  down  the  actual  results.  If  \z\  is  equal  to  \a\  or  \8\  then  no  such 
expansion  is. possible. 


160  LIMITS   OF   FUNCTIONS    OF   A  [iV 

(2)     If  b2<ac  then  the  roots  are  conjugate  complex  numbers  (Ch.  Ill 
$5  43),  and  we  can  write 

a  =  p  Cis<p,       )3=p  Cis  (~4>), 
where     p2  =  a/3  =  c/a,     p  cos  cp  =  £  (a  +  /3)  =  -  6/a,    so    that    cos  0  =  -  s](b2lac), 
sin  0  =  ^/(1  -  (b2Jac)}. 

If  \z\<p  then  each  fraction  may  be  expanded  in  ascending  powers  of  z. 
The  coefficient  of  zn  will  be  found  to  be 

Ap  sin  n<p  +  2?  sin  {(n  +  I)<p} 
apn  +  ]  sin  <p 

If  |«|>p  we  obtain  a  similar  expansion  in  descending  powers,  while  if  |s|  =  p 
no  such  expansion  is  possible. 

20.  Show  that  if  | z\ <  1  then 

1  +  22  +  32"+.  ..  +  («  +  l)s»+...  =  l/(l-2)2. 
J  2n  ?l2n 

[The  sum  to  n  terms  is  r, .„  -  = .1 

L  (1  —  z)-      1  —  z  J 

21.  Expand  Lj{z-a)2  in  powers  of  2,  ascending  or  descending  according 

as  \z\<  |  a  |  or  |  z\  >\a\. 

22.  Show  that  if  b2=ac  and  |  az  |<  |  b  |  then 

Jz  +  2?     _°g      n 

az2  +  2bz  +  c     /nZ> 

where  pll  =  {(-a)n/bn  +  2}  {{n  +  l)aB-?ibA};  and  find   the  corresponding  ex- 
pansion, in  descending  powers  of  z,  which  holds  when  |«2]>|6|. 

23.  Verify  the  result  of  Ex.  19  in  the  case  of  the  fraction  1/(1  +z2).     [We 
have  l/(l+sa)  =  S8nsin  {h(n  +  I)n}  =  l-z2+zi- ....] 


24.     Prove  that  if  U|<1  then 


2     oo 

2zn  sin  {g(n  +  l)  tt}. 


1  +  2  +  22      V3  0 

25.  Expand  (l+*)/(l  +22),  (1  +22)/(l  +  2^)  and  (1  +z  +  z2)/(l  +  zi)  in  ascend- 
ing powers  of  2.     For  what  values  of  2  do  your  results  hold  ? 

26.  If  a/(a  +  bz  +  cz2)  =  1  +fiz +p2z2  + . . .  then 

i  .      2    .      2  2  ,  «+C2  a2 


a  -  as  a2  -  (62  -  2ac)  z  +  c2z2 ' 

{Math.  Trip.  1900.) 
27.     If  lim  sn  =  Z  then 

lim  «i+*+-.+^ 

}i-»-00  " 

[Let  sn=£-K„.     Then  we  have  to  prove  that  (ti  +  t2  +  ...  +  tn)/n  tends  to 
zero  if  tn  does  so. 


POSITIVE    INTEGRAL   VARIABLE  ]61 

We  divide  the  numbers  t1}  t2,  ...  tn  into  two  sets  tu  t2,  ...,  tp  and  tp+1, 
tp  +  2,  ••;  tn-  Here  we  suppose  that  p  is  a  function  of  n  which  tends  to  oo 
as  »-*-co ,  but  more  slowly  than  n,  so  that  p-*-cc  and p/n-*-0  ;  e.g.  we  might 
suppose  p  to  be  the  integral  part  of  Jn. 

Let  f  be  any  positive  number.  However  small  8  may  be,  we  can  choose 
n0  so  that  ?p  +  1,  tp  +  2,...,tn  are  all  numerically  less  than  ^8  when  n>n0,  and  so 

\(tp+i+tp+2+... +tn)/n\<$8(n-p)ln<$8. 

But,  if  A  is  the  greatest  of  the  moduli  of  all  the  numbers  tlt  t2,  ...,  we 
have 

\(tl  +  t2  +  ...  +  tp)/n\<pA/n, 

and  this  also  will  be  less  than  |S  when  n>n0,  if  n0  is  large  enough,  since 
pjn-*-0  as  «-»-oo .     Thus 

l(^+?2+...+0/H|<|(^+^+...  +  g/»|  +  |(^+1+...  +  ^)/?i|<§ 

when  ?^tto  >  which  proves  the  theorem. 

The  reader,  if  he  desires  to  become  expert  in  dealing  with  questions  about 
limits,  should  study  the  argument  above  with  great  care.  It  is  very  often 
necessary,  in  proving  the  limit  of  some  given  expression  to  be  zero,  to  split  it 
into  two  parts  which  have  to  be  proved  to  have  the  limit  zero  in  slightly 
different  ways.     When  this  is  the  case  the  proof  is  never  very  easy. 

The  point  of  the  proof  is  this  :  we  have  to  prove  that  (t1  +  t2  +  ...  +  tn)jn  is 
small  when  n  is  large,  the  t's  being  small  when  their  suffixes  are  large.  We 
split  up  the  terms  in  the  bracket  into  two  groups.  The  terms  in  the  first 
group  are  not  all  small,  but  their  number  is  small  compared  with  n.  The 
number  in  the  second  group  is  not  small  compared  with  n,  but  the  terms  are 
all  small,  and  their  number  at  any  rate  less  than  n,  so  that  their  sum  is  small 
compared  with  n.  Hence  each  of  the  parts  into  which  (t1+t2+...  +  tn)ln 
has  been  divided  is  small  when  n  is  large.] 

28.  If  4>(,n)-  4>{n-  l)-*-l  as  n-s-co,  then  $  (n)ln-*-l. 

[If  <f)  (n)  =  st  +  S2+  ...+sn  then  <p  (n)  —  $  (n—  l)  =  sn,  and  the  theorem  re- 
duces to  that  proved  in  the  last  example.] 

29.  If  slt  =  i  {1  -  ( - 1)"},  so  that  sn  is  equal  to  1  or  0  according  as  n  is  odd 
or  even,  then  {s1  +  s2  +  ...+sn)l7i-^^  as  ?i-*-ao  . 

[This  example  proves  that  the  converse  of  Ex.  27  is  not  true  :  for  su 
oscillates  as  n-*-oo .] 

30.  If  cn ,  sn  denote  the  sums  of  the  first  n  terms  of  the  series 

i  +  cos#  +  cos2<9+...,     sin  6  +  sin  20+..., 
then 

lim  (c! + cz+ . ..  +  cn)jn  =  0,       lim  (st + s2  + . . .  +  sn)jn  =  \  cot  \6. 

11 


CHAPTER  V 

LIMITS    OF   FUNCTIONS    OF    A    CONTINUOUS    VARIABLE. 
CONTINUOUS   AND   DISCONTINUOUS    FUNCTIONS 

89.     Limits  as  x  tends  to   oo .     We   shall   now  return  to 

functions  of  a  continuous  real  variable.  We  "shall  confine  our- 
selves entirely  to  one-valued  functions*,  and  we  shall  denote  such 
a  function  by  <f>  (x).  We  suppose  x  to  assume  successively  all 
values  corresponding  to  points  on  our  fundamental  straight  line 
A,  starting  from  some  definite  point  on  the  line  and  progressing 
always  to  the  right.  In  these  circumstances  we  say  that  x 
tends  to  infinity,  or  to  oo  ,  and  write  x  ->■  oo  .  The  only  difference 
between  the  '  tending  of  n  to  oo  '  discussed  in  the  last  chapter,  and 
this  '  tending  of  x  to  oo  ',  is  that  x  assumes  all  values  as  it  tends 
to  oo ,  i.e.  that  the  point  P  which  corresponds  to  x  coincides  in 
turn  with  every  point  of  A  to  the  right  of  its  initial  position, 
whereas  n  tended  to  oo  by  a  series  of  jumps.  We  can  express  this 
distinction  by  saying  that  x  tends  continuously  to  oo  . 

As  we  explained  at  the  beginning  of  the  last  chapter,  there  is 
a  very  close  correspondence  between  functions  of  x  and  functions 
of  n.  Every  function  of  n  may  be  regarded  as  a  selection  from 
the  values  of  a  function  of  x.  In  the  last  chapter  Ave  discussed 
the  peculiarities  which  may  characterise  the  behaviour  of  a 
function  c£  (n)  as  n  tends  to  oo .  Now  we  are  concerned  with  the 
same  problem  for  a  function  </>  (x) ;  and  the  definitions  and 
theorems  to  which  we  are  led  are  practically  repetitions  of  those 
of  the  last  chapter.  Thus  corresponding  to  Def.  1  of  §  58  we 
have : 

*  Thus  Klx  stands  in  this  chapter  for  the  one-valued  function  +  K!x  and  not  (as 
in  §  26)  for  the  two-valued  function  whose  values  are  +  s!x  and  -  s!x. 


89]  LIMITS   OF   FUNCTIONS    OF   A   CONTINUOUS   VARIABLE      163 

Definition  1.  The  function  <f>(x)  is  said  to  tend  to  the  limit  I 
as  x  tends  to  oo  if,  when  any  positive  number  8,  however  small,  is 
assigned,  a  number  x0  (8)  can  be  chosen  such  that,  for  all  values  of 
x  equal  to  or  greater  than  x0  (8),  0  (x)  differs  from  I  by  less  than  8, 
i.e.  if 

\<j>(x)-l\<S 
when  x  =  x0  (S). 

When  this  is  the  case  we  may  write 
Km  <f>  (x)  =  I, 

x-*  oo 

or,  when  there  is  no  risk  of  ambiguity,  simply  Km  $  (x)  =  I,  or 
<b  (x)  ->■ 1.     Similarly  we  have  : 

Definition  2.     The  function  <^(x)  is  said  to  tend  to  oo  with 
i  x  if  when  any  number  A,  however  large,  is  assigned,  we  can  choose 
a  number  #'0(A)  such  that 

<f>  (x)  >  A 
wlien  x  =  x0(A). 

We  then  write 

<f)  (x)  -^  oo  . 

Similarly  we  define  <£  (x)  -*■  —  cc  *.     Finally  we  have : 

Definition  3.     If  the  conditions  of  neither  of  the  two  preceding 
definitions  are  satisfied,  then  <£  (x)  is  said  to  oscillate  as  x  tends 
l  to  oo  .     If  |  (f>  (x)  I  is  less  than  some  constant  K  when  x  =  x0f,  then 
<b(x)  is  said  to  oscillate  finitely,  and  otherwise  infinitely. 

The  reader  will  remember  that  in  the  last  chapter  we  con- 
sidered very  carefully  various  less  formal  ways  of  expressing  the 
facts  represented  by  the  formulae  (f)(n)~^l,  <j>(n)  -»»  oo  .  Similar 
modes  of  expression  may  of  course  be  used  in  the  present  case. 
Thus  we  may  say  that  <£  (x)  is  small  or  nearly  equal  to  I  or  large 
when  x  is  large,  using  the  words  '  small ',  '  nearly ',  '  large '  in 
a  sense  similar  to  that  in  which  they  were  used  in  Ch.  IV. 

*  We  shall  sometimes  find  it  convenient  to  write  +00,  x  -*■  +  00  ,  <j>  (x)  -*-  +  00 
instead  of  00  ,  x  -*-  00  ,  <p  (x)  -*■  oo . 

t  In  the  corresponding  definition  of  §  62,  we  postulated  that  |  <j>  (n)  \  <K  for  all 
values  of  n,  and  not  merely  when  n  >  ?i0 .  But  then  the  two  hypotheses  would  have 
beeu  equivalent ;  for  if  |  cp  (n)  |  <  K  when  11  2: 7i0,  then  |  <p  (n)  \  <  K'  for  all  values 
of  n,  where  K'  is  the  greatest  of  0(1),  0(2),  ...  ,  (p(n0-l)  and  K.  Here  the 
matter  is  not  quite  so  simple,  as  there  are  infinitely  many  values  of  x  less  than  x0. 

11—2 


1G4  LIMITS   OF   FUNCTIONS  [V 

Examples  XXXIV.  1.  Consider  the  behaviour  of  the  following  functions 
as  .1  '  -co  :  1/x,  l  +  {l/:c),  x\  x\  [x],  x-[x],  [x]  +  J{x-[x]}. 

The  first  four  functions  correspond  exactly  to  functions  of  n  fully  dis- 
cussed in  Ch.  IV.  The  graphs  of  the  last  three  were  constructed  in  Ch.  II 
(Exs.  xvi.  1,  2,  4),  and  the  reader  will  see  at  once  that  [#]-»■  <x>,  x-  [x]  oscillates 
finitely,  and  [x]  +  J{x  -  [x]}  -*-  co . 

One  simple  remark  may  be  inserted  here.  The  function  cj)(x)  =  x-[x] 
oscillates  between  0  and  1,  as  is  obvious  from  the  form  of  its  graph.  It  is 
equal  to  zero  whenever  x  is  an  integer,  so  that  the  function  cf>(n)  derived 
from  it  is  always  zero  and  so  tends  to  the  limit  zero.     The  same  is  true  if 

(f)  {%)  =  sin  xn,     (f>  (n)  =  sin  mr  =0. 

It  is  evident  that  <fr{x)-*-l  or  $(.v)-»-ao  or  <t>{x)-*-  —  cc  involves  the  corre- 
sponding property  for  0  (n),  but  that  the  converse  is  by  no  means  always 
true. 

2.  Consider  in  the  same  way  the  functions : 

(sin  xn)lx,    ^sin.r7r,     {x  sin  xir)2,     tan.^n-,     a  cos2  xn  +  b  sin2  xn, 
illustrating  your  remarks  by  means  of  the  graphs  of  the  functions. 

3.  Give  a  geometrical  explanation  of  Def.  1,  analogous  to  the  geometrical 
explanation  of  Ch.  IV,  §  59. 

4.  If  0  (x)  ->■  I,  and  I  is  not  zero,  then  <£  (x)  cos  xtv  and  0  (x)  sin  xn  oscillate 
finitely.  If  (f>(x)-*-<x>  or  (f>  (x) ->-  —  co  ,  then  they  oscillate  infinitely.  The 
graph  of  either  function  is  a  wavy  curve  oscillating  between  the  curves 
y=<p{x)  andy=-<£(.r). 

5.  Discuss  the  behaviour,  as  x->~cc ,  of  the  function 

y  —f  (x) cos2  xtc + f  (x) sm2  xirt 

where  f{x)  and  F{x)  are  some  pair  of  simple  functions  {e.g.  x  and  x2).  [The 
graph  of  y  is  a  curve  oscillating  between  the  curves  y =f{%),  y  =  F{x).] 

90.  Limits  as  x  tends  to  —  co .  The  reader  will  have  no 
difficulty  in  framing  for  himself  definitions  of  the  meaning  of  the 
assertions  '  x  tends  to  —  co  ',  or  '  x  -*■  —  co  '  and 

lim  <f)  (x)  =1,         <j)  (%)  -*■  co ,         0  (#)-*-  —  co  . 

a:-*  —  oo 

In  fact,  if  x  =  —  y  and  <f>  (x)  =  <f>  (—  y)  =  yjr  (y),  then  y  tends 
to  co  as  x  tends  to  —  co ,  and  the  question  of  the  behaviour  of 
</>  (x)  as  x  tends  to  —  oo  is  the  same  as  that  of  the  behaviour  of 
\}r  (y)  as  y  tends  to  co  . 


89-93]  OF   A    CONTINUOUS   VARIABLE  165 

91.  Theorems    corresponding    to    those    of    Ch.    IV,    §§  63—67. 

The  theorems  concerning  the  sums,  products,  and  quotients  of  functions 
proved  in  Ch.  IV  are  all  true  (with  obvious  verbal  alterations  which  the 
reader  will  have  no  difficulty  in  supplying)  for  functions  of  the  continuous 
variable  x.  Not  only  the  enunciations  but  the  proofs  remain  substantially 
the  same. 

92.  Steadily  increasing  or  decreasing  functions.  The  definition 
which  corresponds  to  that  of  §  69  is  as  follows :  the  function  <£  (x)  will 
be  said  to  increase  steadily  with  x  if  <f)(x2)5:<fi  (.i\)  ivhenever  x2>x1.  In 
many  cases,  of  course,  this  condition  is  only  satisfied  from  a  definite  value 
of  x  onwards,  i.e.  when  xi>xl  >x0.  The  theorem  which  follows  in  that  section 
requires  no  alteration  but  that  of  n  into  x :  and  the  proof  is  the  same,  except 
for  obvious  verbal  changes. 

If  </>(a'2)>0(#i),  tne  possibility  of  equality  being  excluded,  whenever 
x-2>xu  then  <f)(x)  will  be  said  to  be  steadily  increasing  in  the  stricter  sense. 
"We  shall  find  that  the  distinction  is  often  important  (cf.  §§  108 — 109). 

The  reader  should  consider  whether  or  no  the  following  functions 
increase  steadily  with  x  (or  at  any  rate  increase  steadily  from  a  certain 
value  of  x  onwards):  x~-x,  x-\-sinx,  #+2sin#,  .r2  +  2sina-,  [x],  [#]  +  sina;, 
[a?]+V{#~ [#]}■     AH  these  functions  tend  to  oo  as  x-*-  co  . 

93.  Limits  as  x  tends  to  0.  Let  <£  (x)  be  such  a  function 
of  x  that  lim  <f>  (x)  =  /,  and  let  y  =  1/x.     Then 

*(«)-*(l/y)-*(y), 

say.  As  x  tends  to  co  ,  y  tends  to  the  limit  0,  and  ^  (y)  tends  to 
the  limit  I. 

Let  us  now  dismiss  x  and  consider  ty  (y)  simply  as  a  function 
of  y.  We  are  for  the  moment  concerned  only  with  those  values 
of  y  which  correspond  to  large  positive  values  of  x,  that  is  to  say 
with  small  positive  values  of  y.  And  ty  (y)  has  the  property  that 
by  making  y  sufficiently  small  we  can  make  yfr  (y)  differ  by  as 
little  as  we  please  from  I.  To  put  the  matter  more  precisely, 
the  statement  expressed  by  lim  ${x)  =  l  means  that,  when  any 
positive  number  8,  however  small,  is  assigned,  we  can  choose 
ccQ  so  that  \<f>(cc)  —  l\<  8  for  all  values  of  x  greater  than  or  equal 
to  xQ.  But  this  is  the  same  thing  as  saying  that  we  can  choose 
y0  =  l/#„  so  that  |  yjr  (y)  —  1 1  <  8  for  all  positive  values  of  y  less  than 
or  equal  to  y0. 

We  are  thus  led  to  the  following  definitions : 


166  LIMITS   OF   FUNCTIONS  [v 

A.  If,  when  any  positive  number  8,  however  small,  is  assigned, 
we  can  choose  y0  (8)  so  that 

\${y)-l\<8 

when  0  <  y  ^  y0  (8),  then  we  say  that  <f>  (y)  tends  to  the  limit  I  as  y 
tends  to  0  by  positive  values,  and  we  write 

lim  <f>  (y)  =  I. 

B.  If,  when  any  number  A,  however  large,  is  assigned,  we  can 
choose  y0  (A)  so  that 

<f>  (y)  >  A 

ivhen  0  <y  ^y0 (A),  then  we  say  that  <f> (y)  tends  to  co  as  y  tends 
to  0  by  positive  values,  and  we  write 

$  (y)  -*-  oo  . 

We  define  in  a  similar  way  the  meaning  of  '  <f>  (y)  tends  to 
the  limit  I  as  y  tends  to  0  by  negative  values ',  or  '  lim  ty(y)  =  l 
when  y  ->■  —  0 '.  We  have  in  fact  only  to  alter  0  <  y  £  y0  (8)  to 
—  y0  (8)  ^  y  <  0  in  definition  A.  There  is  of  course  a  corresponding 
analogue  of  definition  B,  and  similar  definitions  in  which 

as  y  -*■  +  0  or  y  -*-  —  0. 

If  lim  (j>(y)  =  l  and     lim  §(y)  =  l,  we  write  simply 

lim  <£  (3/)  =  I. 
y-*-o 

This  case  is  so  important  that  it  is  worth  while  to  give  a  formal 
definition. 

If,  when  any  positive  number  8,  however  small,  is  assigned,  we 

can  choose  y0  (8)  so  that,  for  all  values  of  y  different  from  zero  but 

numerically  less  than  or  equal  to  y0  (8),  </>  (y)  differs  from  I  by  less 

than  8,  then  we  say  that  <f>  (y)  tends  to  the  limit  I  as  y  tends  to  0, 

and  write 

lim  <£  (y)  =  I 

So  also,  if  (f)  (y)  -*■  co  as  y  -*■  +  0  and  also  as  y  -*■  —  0,  we  say 
that  <£  (?/)  -*■  oo  as  y  -*•  0.  We  define  in  a  similar  manner  the 
statement  that  $  (y)  -*■  —  co  as  y  -*■  0. 


93,  94]  OF   A   CONTINUOUS   VARIABLE  167 

Finally,  if  <£  (?/)  does  not  tend  to  a  limit,  or  to  oo ,  or  to 
—  cc  ,  as  y  •-»■  +  0,  we  say  that  <f>  (y)  oscillates  as  y  -*»  +  0,  finitely 
or  infinitely  as  the  case  may  be;  and  we  define  oscillation  as 
y  -*■  —  0  in  a  similar  manner. 

The  preceding  definitions  have  been  stated  in  terms  of  a 
variable  denoted  by  y :  what  letter  is  used  is  of  course  immaterial, 
and  we  may  suppose  x  written  instead  of  y  throughout  them. 

94.  Limits  as  x  tends  to  a.  Suppose  that  $(y)-*-l  as 
y  -*■  0,  and  write 

y  =  x-a,     $  (y)  =  $  (x  -  a)  =  f  (x). 

If  y  -*-  0  then  x  -*•  a  and  -fr  (x)  -*•  I,  and  we  are  naturally  led  to 
write 

lim  \fr  (x)  =  I, 

or  simply  lim  yfr  (x)  =  I  or  yp-  (x)  -*■  I,  and  to  say  that  yjr  (x)  tends  to 
the  limit  I  as  x  tends  to  a.  The  meaning  of  this  equation  may 
be  formally  and  directly  defined  as  follows :  if,  given  8,  we  can 
always  determine  e(8)  so  that 

\4>(x)-l\<8 

when  0  <  |  x  —  a  \  ^  e  (8),  then 

lim  (f>  (x)  =  I. 

By  restricting  ourselves  to  values  of  x  greater  than  a,  i.e.  by 

replacing  0  <  |  x  —  a  \  ^  e  (8)  by  a  <  x  ^  a  +  e  (8),  we  define  '  cf>  (x) 

tends  to  I  when  x  approaches  a  from  the  right',  which  we  may 

write  as 

lim    <f>  (x)  —  I. 

In  the  same  way  we  can  define  the  meaning  of 

lim    <j)  (x)  =  I. 

x-*-a-0 

Thus  lim  4>(x)  =  1  is  equivalent  to  the  two  assertions 
lim   <f)  (x)  =  I,      lim  0  (x)  =  I. 

x^-a  +  0  .T-»-a-0 

We  can  give  similar  definitions  referring  to  the  cases  in  which 
cf)(x)^x>  or  </>(#)-* — ao  as  x-^a  through  values  greater  or  less 
than  a ;  but  it  is  probably  unnecessary  to  dwell  further  on  these 
definitions,  since  they  are  exactly  similar  to  those  stated  above  in 


1G8  LIMITS   OF   FUNCTIONS  [V 

the  special  case  when  a  =  0,  and  we  can  always  discuss  the 
behaviour  of  <£(#)  as  x~^a  by  putting  x  —  a  =  y  and  supposing 
that  2/^-0. 

95.  Steadily  increasing  or  decreasing  functions.  If  there  is  a  number 
c  such  that  0  (x') 2g$  (x")  whenever  a-e<x'<x"<a  +  e,  then  <f>(x)  will  be 
said  to  increase  steadily  in  the  neighbourhood  of  x=a. 

Suppose  first  that  x<a,  and  put  y=lj(a-x).  Then  y-z-ao  as  x-*-a— 0, 
and  0  (x)  =  \//-  (y)  is  a  steadily  increasing  function  of  ?/,  never  greater  than  $  (a). 
It  follows  from  §  92  that  0  (x)  tends  to  a  limit  not  greater  than  <£  (a).  We 
shall  write 

lim    <fi(x)  =  (}>(a+0). 

We  can  define  <j>  (a  -  0)  in  a  similar  manner ;  and  it  is  clear  that 

0  (a-0) <(f)  (a)^0  (a  +  0). 

It  is  obvious  that  similar  considerations  may  be  applied  to  decreasing 
functions. 

If  (f>(x')<<fi(x"),  the  possibility  of  equality  being  excluded,  whenever 
a  —  e<x'<x"<a  +  e,  then  (f)(x)  will  be  said  to  be  steadily  increasing  in  the 
stricter  sense. 

96.  Limits  of  indetermination  and  the  principle  of  convergence. 

All  of  the  argument  of  §§  80 — 84  may  be  applied  to  functions  of  a  con- 
tinuous variable  x  which  tends  to  a  limit  a.  In  particular,  if  (f>  (x)  is 
bounded  in  an  interval  including  a  (i.e.  if  we  can  find  e,  H,  and  K  so  that 
H<<f)  (x)<K  when  a  —  e<^'^a  +  e)*,  then  we  can  define  X  and  A,  the  lower  and 
upper  limits  of  indetermination  of  <£  (x)  as  x->-a,  and  prove  that  the  necessary 
and  sufficient  condition  that  0  (x)-*-l  as  x-*-a  is  that  X  =  A  =  l.  We  can  also 
establish  the  analogue  of  the  principle  of  convergence,  i.e.  prove  that  the 
necessary  and  sufficient  condition  that  <£  (x)  shoidd  tend  to  a  limit  as  x-*-a  is 
that,  when  8  is  given,  we  can  choose  e  (8)  so  that  |0(#2)-(M-'ri)|<S  when 
0<\x2-a\  <  |#i-a|<e  (S). 

Examples  XXXV.    1.    If 

as  x  ->- a,  then  <£  (x)  +  \//-  (#)  -a~  £  +  /',  $  (x)  ^  (x)  -»■  W,  and  0  (.r)/^  (a?)  -*■  1 1 1', 
unless  in  the  last  case  l'  =  0. 

[We  saw  in  §  91  that  the  theorems  of  Ch.  IV,  §§  63  et  seq.  hold  also  for 
functions  of  x  when  x ->-  oc  or#-*--oo.  By  putting  x=\jy  we  may  extend 
them  to  functions  of  y,  when  y-*-0,  and  by  putting  y  =  z  —  a  to  functions  of  z, 
when  z-*-a. 

*  For  some  further  discussion  of  the  notion  of  a  function  bounded  in  an  interval 
see  §  102. 


94-97]  OF   A   CONTINUOUS   VARIABLE  169 

The  reader  should  however  try  to  prove  them  directly  from  the  formal 
definition  given  above.  Thus,  in  order  to  obtain  a  strict  direct  proof  of  the 
first  result  he  need  only  take  the  proof  of  Theorem  I  of  §  63  and  write 
throughout  x  for  n,  a  for  oo  and  0<|#-a|<e  for  n>nQ.] 

2.  If  m  is  a  positive  integer  then  #m-»-0  as  x-*-0. 

3.  If  m  is  a  negative  integer  then  xm-*-  +  <x>  as  x-*-  +  0,  while  a?"-*-— oo  or 
xm -*-  +  <&  &sx-*--0,  according  as  m  is  odd  or  even.  If  »j  =  0  theu  xm=l 
and  xm-*-l. 

4.  li  m  (a  +  bx  +  ex2  +  . . .  +  kxm)  =  a. 

5.  lim  \(a  +  bx+  ...  +kxm)J(a  +  0x  +  ...  +KX^)\  =  a!a,  unless  a  =  0.    If  a=0 

and  «#=0,  (34=0,  then  the  function  tends  to  +oo  or  -  oo ,  as  x-*-  +  0,  according 
as  a  and  /3  have  like  or  unlike  signs;  the  case  is  reversed  if  a--»--0.  The 
case  in  which  both  a  and  a  vanish  is  considered  in  Ex.  xxxvi.  5.  Discuss  the 
cases  which  arise  when  a=t=0  and  more  than  one  of  the  first  coefficients  in  the 
denominator  vanish. 

6.  lim  xm  =  am,  if  m  is  any  positive  or  negative  integer,  except  when  a  =  Q 

x-*-a 

and  in  is  negative.  [If  m>0,  put  x=y-\-a  and  apply  Ex.  4.  When  to<0, 
the  result  follows  from  Ex.  1  above.  It  follows  at  once  that  lim  P  (x)  =  P  (a), 
if  P  {x)  is  any  polynomial.] 

7.  lim  R  (x)  =  R  (a),  if  R  denotes  any  rational  function  and  a  is  not  one 

x->-a 
of  the  roots  of  its  denominator. 

8.  Show  that  lim  xm=am  for  all  rational  values  of  m,  except  when  a=0 

x-s-a 
and  m  is  negative.     [This  follows  at  once,  when  a  is  positive,  from  the  in- 
equalities (9)  or  (10)  of  §  74.    For  \xm—am\<H\x-a],  where  H  is  the  greater 
of  the  absolute  values  of  mxm~1  and  mam~1  (cf.  Ex.  xxvin.  4).    If  a  is  negative 
we  write  x=  -y  and  a=  -  b.     Then 

lim  xm  =  lim  ( - 1  )mfn  =  ( -  1  )'n  bm  =  am.] 

97.  The  reader  will  probably  fail  to  see  at  first  that  any  proof 
of  such  results  as  those  of  Exs.  4,  5,  6,  7,  8  above  is  necessary. 
He  may  ask  '  why  not  simply  put  x  =  0,  or  x  =  a  ?  Of  course 
we  then  get  a,  a/a,  am,  P  (a),  R  (a) '  It  is  very  important  that  he 
should  see  exactly  where  he  is  wrong.  We  shall  therefore  consider 
this  point  carefully  before  passing  on  to  any  further  examples. 

The  statement  lim  cf)  (x)  =  I 

is  a  statement  about  the  values  of  (f>(x)  when  x  has  any  value 


170  LIMITS   OF   FUNCTIONS  [V 

distinct  from  but  differing  by  little  from  zero  *.  It  is  not  a  statement 
about  the  value  of  <f>  (x)  when  x  =  0.  When  we  make  the  state- 
ment we  assert  that,  when  x  is  nearly  equal  to  zero,  <f>(x)  is  nearly- 
equal  to  I.  We  assert  nothing  whatever  about  what  happens 
when  x  is  actually  equal  to  0.  So  far  as  we  know,  $  (x)  may 
not  be  defined  at  all  for  x  =  0 ;  or  it  may  have  some  value 
other  than  I.  For  example,  consider  the  function  defined  for  all 
values  of  x  by  the  equation  c/>  (x)  =  0.     It  is  obvious  that 

lim</><»  =  0    (1). 

Now  consider  the  function  i/r  (x)  which  differs  from  <f>  (x)  only  in 
that  -v/r  (x)  =  1  when  x  =  0.     Then 

\imyjr(x)  =  0    (2), 

for,  when  x  is  nearly  equal  to  zero,  i|r  (x)  is  not  only  nearly  but 
exactly  equal  to  zero.  But  ty  (0)  =  1.  The  graph  of  this  function 
consists  of  the  axis  of  x,  with  the  point  x  =  0  left  out,  and  one 
isolated  point,  viz.  the  point  (0,  1).  The  equation  (2)  expresses 
the  fact  that  if  we  move  along  the  graph  towards  the  axis  of  y, 
from  either  side,  then  the  ordinate  of  the  curve,  being  always  equal 
to  zero,  tends  to  the  limit  zero.  This  fact  is  in  no  way  affected 
by  the  position  of  the  isolated  point  (0,  1). 

The  reader  may  object  to  this  example  on  the  score  of 
artificiality :  but  it  is  easy  to  write  down  simple  formulae  repre- 
senting functions  which   behave   precisely  like  this   near  x  =  0. 

One  is 

^(x)  =  [l-x2l 

where  [1  —  x-"]  denotes  as  usual  the  greatest  integer  not  greater 
than  1  —  x2.  For  if  x  =  0  then  ty  (x)  =  [1]  =  1 ;  while  if  0  <  x  <  1, 
or  —  1  <  x  <  0,  then  0  <  1  —  x2  <  1  and  so  ■yjr  (x)  =  [1  —  x2]  =  0. 

Or  again,  let  us  consider  the  function 

y  =  x/x 

already  discussed  in  Ch.  II,  §  24,  (2).  This  function  is  equal 
to  1  for  all  values  of  x  save  x  =  0.  It  is  not  equal  to  1  when 
x  =  0 :  it  is  in  fact  not  defined  at  all  for  x  =  0.     For  when  we  say 

*  Thus  in  Def.  A  of  §  93  we  make  a  statement  about  values  of  y  such  that 
0<y£.y0,  the  first  of  these  inequalities  being  inserted  expressly  in  order  to 
exclude  the  value  y  =  0. 


97]  OF   A   CONTINUOUS   VARIABLE  171 

that  <fr(x)  is  defined  for  x  =  0  we  mean  (as  we  explained  in  Ch.  II, 
I.e.)  that  we  can  calculate  its  value  for  x  =  0  by  putting  x  =  0 
in  the  actual  expression  of  $  (cc).  In  this  case  we  cannot.  When 
we  put  x  =  0  in  <f>(x)  we  obtain  0/0,  which  is  a  meaningless 
expression.  The  reader  may  object  'divide  numerator  and  de- 
nominator by  x '.  But  he  must  admit  that  when  x  =  0  this  is 
impossible.  Thus  y  =  xjx  is  a  function  which  differs  from  y  =  1 
solely  in  that  it  is  not  defined  for  x  =  0.     None  the  less 

lim  {xjx)  =  1, 

for  xjx  is  equal  to  1  so  long  as  x  differs  from  zero,  however  small 
the  difference  may  be. 

Similarly  <f)  (x)  =  {(x  +  l)2  —  l}/x  =  x  +  2  so  long  as  x  is  not 
equal  to  zero,  but  is  undefined  when  x  =  0.  None  the  less 
lim  (j)  (x)  =  2. 

On  the  other  hand  there  is  of  course  nothing  to  prevent  the 
limit  of  <f>  (x)  as  x  tends  to  zero  from  being  equal  to  <f>  (0),  the  value 
of  cj)  (x)  for  x  =  0.  Thus  if  <j)(x)=x  then  <f>  (0)  =  0  and  lim  <f>  (x)  =  0. 
This  is  in  fact,  from  a  practical  point  of  view,  i.e.  from  the  point 
of  view  of  what  most  frequently  occurs  in  applications,  the 
ordinary  case. 

Examples  XXXVI.     1 .    lim  (x2  -  a?)/(x  -a)= 2a. 

2.  lim  {xm-am)[(x  —  a)  =  mam~1,  if  m  is  any  integer  (zero  included). 

3.  Show  that  the  result  of  Ex.  2  remains  true  for  all  rational  values 
of  m,  provided  a  is  positive.  [This  follows  at  once  from  the  inequalities 
(9)  and  (10)  of  §  74.] 

4.  lim  (x7 -  2.v5  +  l)l(x3 -3x2+2)  =  l.      [Observe  that  x-l  is  a  factor  of 

ZH.-1 

both  numerator  and  denominator.] 

5.  Discuss  the  behaviour  of 

<j)(x)  =  (a0xm  +  a1xm  +  1+  ...  +akxm  +  *)/(b0xn  +  biXn  +  1+  ...  +blx7t+l) 
as  x  tends  to  0  by  positive  or  negative  values. 

[If  m  >  n,  lim  0  (a?) = 0.  If  m  =  n,  lim  $  (x)  =  or0/60 .  If  m  <  n  and  n  -  m  is 
exer\,(j)(x)-*-  +  cc  or  0  (.r) -*- -  oo  according  as  a0/b0>0  or  a0/b0<0.  Ifm<?iand 
n  -  m  is  odd,  0  (x)  -^  +  »  as  x->-  +  0  and  (f>  (x) ^  -  co  as  a;-*-  -  0,  or  $ (x)  ^  -  co 
as  :i'-9-  +  0  and  <£(#)-*  +  »  as  a?-*- - 0,  according  as  ff0/60>°  or  «<A<0-] 


172  LIMITS   OF    FUNCTIONS  [V 

6.  Orders  of  smallness.  When  x  is  small  x2  is  very  much  smaller, 
x3  much  smaller  still,  and  so  on :  in  other  words 

lira  (.vVx)  =  0,     lim(A-3/*'2)  =  0,     .... 

Another  way  of  stating  the  matter  is  to  say  that,  when  x  tends  to  0, 
x2,  x3,  ...  all  also  tend  to  0,  but  x2  tends  to  0  more  rapidly  than  x,  x3  than 
x2,  and  so  on.  It  is  convenient  to  have  some  scale  by  which  to  measure 
the  rapidity  with  which  a  function,  whose  limit,  as  x  tends  to  0,  is  0, 
diminishes  with  x,  and  it  is  natural  to  take  the  simple  functions  x,  x2,  x3,  ... 
as  the  measures  of  our  scale. 

We  say,  therefore,  that  <p(x)  is  of  the  first  order  of  smallness  if  (f>(x)/x 
tends  to  a  limit  other  than  0  as  x  tends  to  0.  Thus  2x+3x2+xr  is  of  the 
first  order  of  smallness,  since  lim  (2x  +  3x2  +  x7)/x  =  2. 

Similarly  we  define  the  second,  third,  fourth,  ...  orders  of  smallness.  It 
must  not  be  imagined  that  this  scale  of  orders  of  smallness  is  in  any  way 
complete.  If  it  were  complete,  then  every  function  <fr  (x)  which  tends  to  zero 
with  x  would  be  of  either  the  first  or  second  or  some  higher  order  of  smallness. 
This  is  obviously  not  the  case.  For  example  <^{x)  =  x71'0  tends  to  zero  more 
rapidly  than  x  and  less  rapidly  than  x2. 

The  reader  may  not  unnaturally  think  that  our  scale  might  be  made 
complete  by  including  in  it  fractional  orders  of  smallness.  Thus  we  might 
say  that  x7/&  was  of  the  £th  order  of  smallness.  We  shall  however  see  later 
on  that  such  a  scale  of  orders  would  still  be  altogether  incomplete.  And 
as  a  matter  of  fact  the  integral  orders  of  smallness  defined  above  are  so 
much  more  important  in  applications  than  any  others  that  it  is  hardly 
necessary  to  attempt  to  make  our  definitions  more  precise. 

Orders  of  greatness.  Similar  definitions  are  at  once  suggested  to 
meet  the  case  in  which  <p  (x)  is  large  (positively  or  negatively)  when  x  is 
small.  We  shall  say  that  <fi  (x)  is  of  the  £th  order  of  greatness  when  x  is  small 
if  (p  (x)jx~k=xk(f)(x)  tends  to  a  limit  different  from  0  as  x  tends  to  0. 

These  definitions  have  reference  to  the  case  in  which  x-*-0.  There  are  of 
course  corresponding  definitions  relating  to  the  cases  in  which  x-*~  go  or  x  -*■  a. 
Thus  if  xkcj)(x)  tends  to  a  limit  other  than  zero,  as  o,-^-oo ,  then  we  say  that 
<j)(x)  is  of  the  kth  order  of  smallness  when  x  is  large:  while  if  (x-a)k<j>(x) 
tends  to  a  limit  other  than  zero,  as  x-*-a,  then  we  say  that  (f>  (x)  is  of  the  £th 
order  of  greatness  when  x  is  nearly  equal  to  a. 

*7.  limN/(l+.r)  =  hniv'(l-..i;)  =  b  [Put  l+x=y  or  \-x=y,  and  use 
Ex.  xxxv.  8.] 

8.  lim{N/(l+.r)-v/(l  —x)}jx=\.  [Multiply  numerator  and  denominator 
hy^(l+x)  +  s/(l-x).] 

*  In  the  examples  which  follow  it  is  to  be  assumed  that  limits  as  x-*0  are 
required,  unless  (as  in  Exs.  19,  22)  the  contrary  is  explicitly  stated. 


97]  OF   A   CONTINUOUS   VARIABLE  173 

9.  Consider  the  behaviour  of  {N/(l  +xm)  -  v/(l  -  xm)}/xn  as  #-*-<),  m  and  n 
being  positive  integers. 

10.  lina  {v/(l  +.V+A'2)  -  l}lx=$. 

nmv/(l-*2W(l-*) 

12.  Draw  a  graph  of  the  function 

r-  {j^t  +^i +^j + ^}  /{^r + irj + j^t  +  ,4j}  • 

Has  it  a  limit  as  x-*-01    [Here  y=l  except  for  x=l,  i,  A,  J,  when  y  is 
not  defined,  and  y-*~l  as  ,r-*-0.] 

13.  limS-^=l. 

x 

[It  may  be  deduced  from  the  definitions  of  the  trigonometrical  ratios*  that 
if  x  is  positive  and  less  than  \rt  then 

sin#<#<tan# 

sin  x     , 

or  cos.r< <1 

x 

ot  0<1 -<l-cos#  =  2sin2i.r. 

x 

But2sin2^<2(ia')2<^'2    Hence  lim  (l  -  —  ")=(),  and   lim   —  =1. 

x-*-+o\  x    J  z-*~+0    x 

As  — -  is  an  even  function,  the  result  follows.] 
x  J 

t.      ..     1-cosa'     .  ,_      ,.     sino»  T    ,,.    ,        .»       _. 

14.  lim — —  =\.  15.     lim  — —  =a.     Is  this  true  it  a  =  0  2 

16.  lim =  1.    [Put.r  =  siny.] 

. .     tan  ax         . .     arc  tan  ax 

17.  hm =  a,  lim =a. 

x  x 

..     cosec  x  -  cot  x      ,  ,«     t     l  +  cos7r.r     , 

18.  lim =  \.  19.     lim— — ., — =*. 

x  x-*-\  t*n  nX 


*  The  proofs  of  the  inequalities  which  are  used  here  depend  on  certain  pro- 
perties of  the  area  of  a  sector  of  a  circle  which  are  usually  taken  as  geometrically 
intuitive ;  for  example,  that  the  area  of  the  sector  is  greater  than  that  of  the 
triangle  inscribed  in  the  sector.  The  justification  of  these  assumptions  must  be 
postponed  to  Ch.  VII. 


174 


CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS 


[V 


20.  How  do  the  functions  sin(l/x),  (l/.r)sin  (l/x),  xsin(llx)  behave 
as  x-*-0?  [The  first  oscillates  finitely,  the  second  infinitely,  the  third 
tends  to  the  limit  0.     None  is  defined  when  #  =  0.     See  Exs.  xv.  6,  7,  8.] 


21.     Does  the  function 


y=(Binf)/(Si1^ 


tend  to  a  limit  as  x  tends  to  0  ?  [No.  The  function  is  equal  to  1  except  when 
sin(l/#)  =  0;  i.e.  when#=l/jr,  1 /2tt,  ...,-1/tt,  -1/2it,  ....  For  these  values  the 
formula  for  y  assumes  the  meaningless  form  0/0,  and  y  is  therefore  not  defined 
for  an  infinity  of  values  of  x  near  x  —  0.] 

22.     Prove  that  if  m  is  any  integer  then   [#]-»-«i   and  x  —  [x]-a-0  as 
x-*-m +0,  and  [#]-»- m -I,  x- [#]-*- 1  as  x-*-m - 0. 

98.     Continuous    functions    of    a    real   variable.     The 

reader  has  no  doubt  some  idea  as  to  what  is  meant  by  a  continuous 

curve.     Thus  he  would  call  the  curve  C  in  Fig.  29  continuous, 

the  curve  C  generally  continuous  but  discontinuous  for  x  =  g  and 

x  =  £  . 

Y 

C 


Fig.  29. 

Either  of  these  curves  may  be  regarded  as  the  graph  of  a 
function  </>  (x).  It  is  natural  to  call  a  function  continuous  if  its 
graph  is  a  continuous  curve,  and  otherwise  discontinuous.  Let  us 
take  this  as  a  provisional  definition  and  try  to  distinguish  more 
precisely  some  of  the  properties  which  are  involved  in  it. 

In  the  first  place  it  is  evident  that  the  property  oi  the 
function  y  =  <f)(x)  of  which  C  is  the  graph  may  be  analysed  into 
some  property  possessed  by  the  curve  at  each  of  its  points. 
To  be  able  to  define  continuity  for  all  values  of  x  we  must  first 
define  continuity  for  any  particular  value  of  x.  Let  us  there- 
fore  fix   on   some   particular   value    of    x,    say    the    value   x  =  £ 


97,  98]         CONTINUOUS   AND   DISCONTINUOUS    FUNCTIONS  175 

corresponding   to    the   point   P   of    the   graph.      What   are    the 
characteristic  properties  of  <£  (x)  associated  with  this  value  of  x  ? 

In  the  first  place  $  (x)  is  defined  for  x  =  f .  This  is  obviously 
essential.  If  <p  (£)  were  not  defined  there  would  be  a  point 
missing  from  the  curve. 

Secondly  <j>  (x)  is  defined  for  all  vahies  of  x  near  x  =  £;  i.e.  we 
can  find  an  interval,  including  x  =  f  in  its  interior,  for  all  points 
of  which  <f>  (x)  is  denned. 

Thirdly  if  x  approaches  the  value  %  from  either  side  then  <f>  (x) 
approaches  the  limit  </>  (£)• 

The  properties  thus  defined  are  far  from  exhausting  those 
which  are  possessed  by  the  curve  as  pictured  by  the  eye  of 
common  sense.  This  picture  of  a  curve  is  a  generalisation  from 
particular  curves  such  as  straight  lines  and  circles.  But  they  are 
the  simplest  and  most  fundamental  properties :  and  the  graph  of 
any  function  which  has  these  properties  would,  so  far  as  drawing 
it  is  practically  possible,  satisfy  our  geometrical  feeling  of  what  a 
continuous  curve  should  be.  We  therefore  select  these  properties 
as  embodying  the  mathematical  notion  of  continuity.  We  are  thus 
led  to  the  following 

Definition.     The  function  <f>  (x)  is  said  to  be  continuous  for . 
x=g  if  it  tends  to  a  limit  as  x  tends  to  £  from  either  side,  and 
each  of  these  limits  is  equal  to  <£  (£). 

We  can  now  define  continuity  throughout  an  interval.  The 
function  <f>  (x)  is  said  to  be  continuous  throughout  a  certain 
interval  of  values  of  x  if  it  is  continuous  for  all  values  of  x  in  that 
interval.  It  is  said  to  be  continuous  everywhere  if  it  is  continuous 
for  every  value  of  x.  Thus  [x]  is  continuous  in  the  interval 
(e,  1  —  e),  where  e  is  any  positive  number  less  than  ^;  and  1  and  x 
are  continuous  everywhere- 

If  we  recur  to  the  definitions  of  a  limit  we  see  that  our 
definition  is  equivalent  to  '  <f>(x)  is  continuous  for  x=  £  if,  given  8, 
we  can  choose  e  (8)  so  that  |  </>  (x)  —  <£  (£)  |  <  8  if  0  ^  |  x  —  f  |  ^  e  (8)'. 

We  have  often  to  consider  functions  defined  only  in  an  interval 
(a,  b).     In  this  case  it  is  convenient  to  make  a  slight  and  obvious 


/ 


176 


CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS 


[V 


change  in  our  definition  of  continuity  in  so  far  as  it  concerns  the 
particular  points  a  and  b.  We  shall  then  say  that  $  (x)  is  con- 
tinuous for  x  =  a  if  </>  (a  +  0)  exists  and  is  equal  to  $  (a),  and  for 
x  =  b  if  (/>  (b  —  0)  exists  and  is  equal  to  (f>  (b). 

99.  The  definition  of  continuity  given  in  the  last  section  may 
be  illustrated  geometrically  as  follows.  Draw  the  two  horizontal 
lines  y  =  <j>  (f )  -  8  and  y  =  0  (f )  +  8.  Then  |  0  (a?)  -  <f>  (f )  |  <  S  ex- 
presses the  fact  that  the  point  on  the  curve  corresponding  to  x  lies 


f+f 


Fig.  30. 


between  these  two  lines.  Similarly  |  x  —  £  |  ^  e  expresses  the  fact 
that  x  lies  in  the  interval  (£— e,  £+e).  Thus  our  definition  asserts 
that  if  we  draw  two  such  horizontal  lines,  no  matter  how  close 
together,  we  can  always  cut  off  a  vertical  strip  of  the  plane  by 
two  vertical  lines  in  such  a  way  that  all  that  part  of  the  curve 
which  is  contained  in  the  strip  lies  between  the  two  horizontal 
lines.  This  is  evidently  true  of  the  curve  G  (Fig.  29),  whatever 
value  £  may  have. 

We  shall  now  discuss  the  continuity  of  some  special  types  of 
functions.  Some  of  the  results  which  follow  were  (as  we  pointed 
out  at  the  time)  tacitly  assumed  in  Ch.  II. 

Examples  XXXVII.  1.  The  sum  or  product  of  two  functions  continuous 
at  a  point  is  continuous  at  that  point.  The  quotient  is  also  continuous 
unless  the  denominator  vanishes  at  the  point.  [This  follows  at  once  from 
Ex.  xxxv.  1.] 

2.  Any  polynomial  is  continuous  for  all  values  of  x.  Any  rational 
fraction  is  continuous  except  for  values  of  x  for  which  the  denominator 
vanishes.     [This  follows  from  Exs.  xxxv.  6,  7.] 


99]  CONTINUOUS    AND    DISCONTINUOUS    FUNCTION'S  177 

3.  v/.r  is  continuous  for  all  positive  values  of  x  (Ex.  xxxv.  8).  It  is  not 
defined  when  x  <  0,  but  is  continuous  for  x  =  0  in  virtue  of  the  remark  made  at 
the  end  of  §  98-  The  same  is  true  of  x'"'n,  where  m  and  n  are  any  positive 
integers  of  which  n  is  even. 

4.  The  function  xm'n,  where  n  is  odd,  is  continuous  for  all  values  of  x. 

5.  l/.r  is  not  continuous  for  .r  =  0.  It  has  no  value  for  x  =  0,  nor  does  it 
tend  to  a  limit  as  x-*-0-  In  fact  \jx-*-  +oo  or  I/a?-*-  —  oo  according  as  x-*-Q 
by  positive  or  negative  values. 

6.  Discuss  the  continuity  of  x~m>'1,  where  m  and  n  are  positive  integers, 
for  x=0. 

7.  The  standard  rational  function  R(x)  =  P  (x)jQ(x)  is  discontinuous  for 
x=a,  where  a  is  any  root  of  Q (x)  =  0.  Thus  (x2  +  l)j(x2-Sx  +  2)  is  discon- 
tinuous for  x=l.  It  will  be  noticed  that  in  the  case  of  rational  functions  a 
discontinuity  is  always  associated  with  (a)  a  failure  of  the  definition  for  a 
particular  value  of  x  and  (b)  a  tending  of  the  function  to  +  oo  or  —  x  as  x 
approaches  this  value  from  either  side.  Such  a  particular  kind  of  point  of 
discontinuity  is  usually  described  as  an  infinity  of  the  function.  An  'infinity' 
is  the  kind  of  discontinuity  of  most  common  occurrence  in  ordinary  work. 

8.  Discuss  the  continuity  of 

v'  {(x  -a)(b-  x%  #  {(x  -  a)  (b  -  x)},  v'{(.r  -  «)/(&  -  .r)},   #{(*  -  a)/(6  -  .r)} 

9.  sin  x  and  cos  x  are  continuous  for  all  values  of  x. 
[We  have  sin  (x  +  h)  —  sin  x=2  sin  \h  cos  (x+\h). 

which  is  numerically  less  than  the  numerical  value  of  A.J 

10.  For  what  values  of  x  are  tana;,  cot  x,  sec  x,  and  cosec  v  continuous 
or  discontinuous  ? 

11.  If  f  (y)  is  continuous  for  ?/  =  »?,  and  0  (x)  is  a  continuous  function  of 
x  which  is  equal  to  rj  when  $=£,  then /{<£  (.r)}  is  continuous  for  #=£. 

12.  If  (f)  (x)  is  continuous  for  any  particular  value  of  x,  then  any  poly- 
nomial in  (p  (x),  such  as  a  {<f>  (x)\m+ ...,  is  so  too. 

13.  Discuss  the  continuity  of 

l/(acos2.r-r-&sin2.r),    ^/C2,  +  cosx),    x'(l -f  sin.r),     l/s/(l  +  sin.  r). 

14.  sin  (l/.r),  .r  sin  (l/x),  and  a?2  sin  (l/.r)  are  continuous  except  for  x  =  0. 

15.  The  function  which  is  equal  to  x  sin  (l/x)  except  when  x  =  0,  and  to 
zero  when  x  =  0,  is  continuous  for  all  values  of  x. 

16.  [r]  and  x  —  [x\  are  discontinuous  for  all  integral  values  of  x. 

17.  For  what  (if  any)  values  of  x  are  the  following  functions  discon- 
tinuous :  [x*\  Wx],  <J(x -[*]),  [x]  +  J(x-[x]),  [2x],  [#]+[-#]? 

II.  12 


178  CONTINUOUS    AND   DISCONTINUOUS    FUNCTIONS  [v 

18.  Classification  of  discontinuities.  Some  of  the  preceding  examples 
suggest  a  classification  of  different  types  of  discontinuity. 

(1)  Suppose  that  0  (x)  tends  to  a  limit  as  x->-a  either  by  values  less 
than  or  by  values  greater  than  a.  Denote  these  limits,  as  in  §  95,  by  0  (a  -  0) 
and  0  (a  +  0)  respectively.  Then,  for  continuity,  it  is  necessary  and  sufficient 
that  0  (x)  should  be  denned  for  x= a,  and  that  0  (a  -  0)  =  0  (a)  -  0  (a +0).  Dis- 
continuity may  arise  in  a  variety  of  ways. 

(a)  0(a-O)  may  be  equal  to  0  (a+0),  but  0(a)  may  not  be  denned,  or 
may  differ  from  0 (a- 0)  and  0(a+O).  Thus  if  0  (#)  =  #sin  (l/x)  and  a  =  0, 
0  (0  -  0)  =  0  (0  +  0) =0,  but  0  (x)  is  not  defined  for  x  =  0.  Or  if  0  (x)  =  [1  -  x2] 
anda  =  0,  0  (0-0)  =  $  (0  +  0)  =  0,  but  0(0)  =  1. 

(/3)  0  (a  -  0)  and  0  (a +  0)  may  be  unequal.  In  this  case  0  (a)  may  be 
equal  to  one  or  to  neither,  or  be  undefined.  The  first  case  is  illustrated 
by  0  (#)  =  [>],  for  which  0  (0  -  0)=  -  1,  0  (0  +  0)=$  (0)  =  0  ;  the  second  by 
0  (a?)  =  [*•]  -  [  -  x],  for  which  0  (0  -  0)  =  - 1 ,  $  (0 + 0)  =  1 ,  $  (0)  =  0 ;  and  the  third 
by  0(.r)  =  [.»;]  +  x sin  (l/.f),  for  which  $  (0-0)=  -  1,  0  (0  +  0)  =  0,  and  $  (0)  is 
undefined. 

In  any  of  these  cases  we  say  that  0  (x)  has  a  simple  discontinuity  at 
x  =  a.  And  to  these  cases  we  may  add  those  in  which  (f>(x)  is  defined  only 
on  one  side  of  x— a,  and  0  (a  -0)  or  $  (a  +  0),  as  the  case  may  be,  exists,  but 
0  (x)  is  either  not  defined  when  x  =  a  or  has  when  x=a  a  value  different  from 
0(a-O)  or  cj){a  +  0). 

It  is  plain  from  §  95  that  a  function  which  increases  or  decreases  steadily 
in  the  neighbourhood  of  x  =  a  can  have  at  most  a  simple  discontinuity  for  x=a. 

(2)  It  may  be  the  case  that  only  one  (or  neither)  of  0  (a  -  0)  and  0  (a  +  0) 
exists,  but  that,  supposing  for  example  0  (a  +  0)  not  to  exist,  0  (#)-»- +  oo  or 
0  (x)-*~-  oo  as  x-^a  +  0,  so  that  0  (x)  tends  to  a  limit  or  to  +oo  or  to  -  qo  as 
x  approaches  a  from  either  side.  Such  is  the  case,  for  instance,  if  0  (x)  =  l/x  or 
0  (x)  =  Ijx2,  and  a  =  0.  In  such  cases  we  say  (cf.  Ex.  7)  that  x=a  is  an  infinity 
of  0  (x).  And  again  we  may  add  to  these  cases  those  in  which  0  (.r)-»-  +oo 
or  (p  (x)-*-  —  cc&sx-*-a  from  one  side,  but  0  (x)  is  not  defined  at  all  on  the 
other  side  of  x  —  a. 

(3)  Any  point  of  discontinuity  which  is  not  a  point  of  simple  discon- 
tinuity nor  an  infinity  is  called  a  point  of  oscillatory  discontinuity.  Such 
is  the  point  x  =  0  for  the  functions  sin  (l/x),  (ljx)  sin  (l/#). 

19.  What  is  the  nature  of  the  discontinuities  at  x  =  0  of  the  functions 
(smx)lx,  [x]  +  [-x],  cosec  x,  <J(l/x),  f/(llx),  cosec(l/.r),  sin  (1 J x)/ sin  (l/x)  ? 

20.  The  function  which  is  equal  to  1  when  x  is  rational  and  to  0  when 
x  is  irrational  (Ch.  II,  Ex.  xvi.  10)  is  discontinuous  for  all  values  of  x.  So  too 
is  any  function  which  is  defined  only  for  rational  or  for  irrational  values  of  x. 


99,  100]      CONTINUOUS    AND    DISCONTINUOUS   FUNCTIONS  179 

21.  The  function  which  is  equal  to  x  when  x  is  irrational  and  to 
\/{(1+P2)/(1  +  ?2)}  wnen  x  is  a  rational  fraction  p/q  (Ch.  II,  Ex.  xvi.  11)  is 
discontinuous  for  all  negative  and  for  positive  rational  values  of  .v,  but 
continuous  for  positive  irrational  values. 

22.  For  what  points  are  the  functions  considered  in  Ch.  IV,  Exs.  xxxi 
discontinuous,  and  what  is  the  nature  of  their  discontinuities  ?  [Consider, 
e.g.,  the  function  y  =  lini  xn  (Ex.  5).  Here  y  is  only  defined  when  -l<.r<l  : 
it  is  equal  to  0  when  -\<x<\  and  to  1  when  x=\.     The  points  x=l  and 

x=  —  1  are  points  of  simple  discontinuity.] 

100.  The  fundamental  property  of  a  continuous  function. 

It  may  perhaps  be  thought  that  the  analysis  of  the  idea  of  a  con- 
tinuous curve  given  in  §  98  is  not  the  simplest  or  most  natural 
possible.  Another  method  of  analysing  our  idea  of  continuity  is  the 
following.  Let  A  and  B  be  two  points  on  the  graph  of  0  (x)  whose 
coordinates  are  x0,  <f>(x0)  and  x1}  <}>(%i)  respectively.  Draw  any 
straight  line  X  which  passes  between  A  and  B.  Then  common 
sense  certainly  declares  that  if  the  graph  of  </>  (x)  is  continuous  it 
must  cut  X. 

If  we  consider  this  property  as  an  intrinsic  geometrical 
property  of  continuous  curves  it  is  clear  that  there  is  no  real 
loss  of  generality  in  supposing  X  to  be  parallel  to  the  axis  of  x. 
In  this  case  the  ordinates  of  A  and  B  cannot  be  equal :  let  us 
suppose,  for  defmiteness,  that  <f)  (^2)  >  <p  (x0).  And  let  X  be  the 
line  y  =  ?/,  where  $  (x0)  <  r/  <<f>  (x,).  Then  to  say  that  the  graph 
of  <f)(x)  must  cut  X  is  the  same  thing  as  to  say  that  there  is  a 
value  of  x  between  x0  and  a^  for  which  <£  (x)  =  rj. 

We  conclude  then  that  a  continuous  function  <f>  (x)  must 
possess  the  following  property :    if 

4>{oc0)  =  ya,     <f)(x1)  =  !/1, 

and  y()<ri<  y1,then  there  is  a  value  of  x  between  x0  and  xxfor  which 
cf)  (#)  =  rj.  In  other  words  as  x  varies  from  x0  to  x1}  y  must  assume 
at  least  once  every  value  between  y0  and  yx. 

We  shall  now  prove  that  if  </>  (x)  is  a  continuous  function  of  x  in 
the  sense  denned  in  §  98  then  it  does  in  fact  possess  this  property. 
There  is  a  certain  range  of  values  of  x,  to  the  right  of  x0,  for  which 
<b(.r)<v.     For  (p{xu)<r/,  and  so  <£  (x)  is  certainly  less  than  ?;  if 

12—2 


180  CONTINUOUS   AND    DISCONTINUOUS    FUNCTIONS  [V 

<f>  (x)  —  <f>  (x0)  is  numerically  less  than  77  —  <£  (x0).  But  since  <£  (x) 
is  continuous  for  x  =  x0,  this  condition  is  certainly  satisfied  if  x  is 
near  enough  to  x0.  Similarly  there  is  a  certain  range  of  values, 
to  the  left  of  x1}  for  which  cf>  (x)  >  77. 

Let  us  divide  the  values  of  x  between  x0  and  xx  into  two  classes 
L,  R  as  follows : 

(1)  in  the  class  L  we  put  all  values  f  of  x  such  that  <f>  (x)  <  77 
when  x  =  £  and  for  all  values  of  x  between  x0  and  £ ; 

(2)  in  the  class  R  we  put  all  the  other  values  of  x,  i.e.  all 
numbers  f  such  that  either  </>  (£)  =  7;  or  there  is  a  value  of  x  between 
x0  and  £  for  which  <f>  (as)  =  77. 

Then  it  is  evident  that  these  two  classes  satisfy  all  the 
conditions  imposed  upon  the  classes  L,  R  of  §  17,  and  so  constitute 
a  section  of  the  real  numbers.  Let  f0  be  the  number  corresponding 
to  the  section. 

First  suppose  <£  (£0)  >  77,  so  that  £0  belongs  to  the  upper  class : 
and  let  (f>  (£0)  =  V  +  k,  say.     Then  <£  (£')  <  77  and  so 

for  all  values  of  £'  less  than  £0,  which  contradicts  the  condition  of 
continuity  for  x  =  £0. 

Next  suppose  <£  (£„)  =  r/  —  k  <  rj.  Then,  if  £'  is  any  number 
greater  than  £„,  either  <£(£')  =  77  or  we  can  find  a  number  £" 
between  £0  and  £'  such  that  <£(£")  =  77.  In  either  case  we  can 
find  a  number  as  near  to  £0  as  we  please  and  such  that  the  corre- 
sponding values  of  (f>  (x)  differ  by  more  than  Jc.  And  this  again 
contradicts  the  hypothesis  that  <f>  (x)  is  continuous  for  x  =  £0. 

Hence  <f>  (£0)  =  77,  and  the  theorem  is  established.  It  should 
be  observed  that  we  have  proved  more  than  is  asserted  explicitly 
in  the  theorem ;  we  have  proved  in  fact  that  |0  is  the  least  value 
of  x  for  which  cf>  (x)  =  tj.  It  is  not  obvious,  or  indeed  generally 
true,  that  there  is  a  least  among  the  values  of  x  for  which  a 
function  assumes  a  given  value,  though  this  is  true  for  continuous 
functions. 

It  is  easy  to  see  that  the  converse  of  the  theorem  just  proved  is  not 
true.     Thus  such  a  function  as  the  function  $  (.r)  whose  graph  is  represented 


100-102]     CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  181 


by  Fig.  31  obviously  assumes  at  least  once  every  value  between  0  (.v0)  and 
0(-yi) :  ye^  $  (x)  *s  discontinuous.  Indeed  it  is  not  even  true  that  0  (x)  must 
be  continuous  when  it  assumes  each  value  once  and  once  only.  Thus  let  0  (x) 
be  denned  as  follows  from  .r=0  to  x  =  l.  If  ,x-  =  0  let  0  (x)=0;  if  0  <  x  <  1 
let  cf)(x)  =  l-x;  and  if  x=l  let  <f>(x)  =  l.  The  graph  of  the  function  is 
shown  in  Fig.  32;  it  includes  the  points  0,  C  but  not  the  points  A,  D.  It 
is  clear  that,  as  x  varies  from  0  to  1,  0  (x)  assumes  once  and  once  only  every 
value  between  0  (0)  =  0  and  0(1)  =  1 ;  but  0(.r)  is  discontinuous  for  x  =  0  and 
x=l. 


Fig.  31. 


As  a  matter  of  fact,  however,  the  curves  which  usually  occur  in  elementary 
mathematics  are  composed  of  a  finite  number  of  pieces  along  which  y  always 
varies  in  the  same  direction.  It  is  easy  to  show  that  if  ?/  =  0  (x)  always  varies 
in  the  same  direction,  i.e.  steadily  increases  or  decreases,  as  x  varies  from 
.rtl  to  Xi,  then  the  two  notions  of  continuity  are  really  equivalent,  i.e.  that  if 
0  (x)  takes  every  value  between  0  (x0)  and  0  fa)  then  it  must  be  a  continuous 
function  in  the  sense  of  §  98  For  let  £  be  any  value  of  x  between  x0  and 
xv  As. r-»-£  through  values  less  than  £,  0(#)  tends  to  the  limit  0(£  — 0) 
(§  95).  Similarly  as  x-^^  through  values  greater  than  £,  0  (x)  tends  to  the 
limit  0(|+O).     The  function  will  be  continuous  for  *'=£  if  and  only  if 

<M£-o)=0(£)=0(|+o) 

But  if  either  of  these  equations  is  untrue,  say  the  first,  then  it  is  evident  that 
<f>(x)  never  assumes  any  value  which  lies  between  0  (£  — 0)  and  0  (£),  which 
is  contrary  to  our  assumption.  Thus  0  (x)  must  be  continuous.  The  net 
result  of  this  and  the  last  section  is  consequently  to  show  that  our  common- 
sense  notion  of  what  we  mean  by  continuity  is  substantially  accurate,  and 
capable  of  precise  statement  in  mathematical  terms. 

101.    In  this  and  the  following  paragraphs  we  shall  state  and 
prove  some  general  theorems  concerning  continuous  functions. 


182  CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  [v 

Theorem  1.  Suppose  that  <f>(x)  is  continuous  for  x=%,  and 
that  <£(£)  is  positive.  Then  we  can  determine  a  positive  number  e 
such  that  <£  (£)  is  positive  throughout  the  interval  (£  —  e,  £  +  e). 

For,  taking  8  =  \$  (£)  in  the  fundamental  inequality  of  p.  175, 
we  can  choose  e  so  that 

|*(«)--*(0I<**(0 

throughout  (£  —  e,  £  +  e),  and  then 

*(*)s*(0-i£(*)-*<0i>to(e)>o, 

so  that  <£  {x)  is  positive.    There  is  plainly  a  corresponding  theorem 
referring  to  negative  values  of  <£  (x). 

Theorem  2.  If(f>  ix)  ^s  continuous  for  x  =  f,  and  </>  (a;)  vanishes 
for  values  of  x  as  near  to  £  as  we  please,  or  assumes,  for  values  of 
x  as  near  to  £  as  we  please,  both  positive  and  negative  values,  then 

*({)- a 

This  is  an  obvious  corollary  of  Theorem  1.  If  (/>(£)  is  not  zero, 
it  must  be  positive  or  negative  ;  and  if  it  were,  for  example,  positive, 
it  would  be  positive  for  all  values  of  x  sufficiently  near  to  £,  which 
contradicts  the  hypotheses  of  the  theorem. 

102.    The  range  of  values  of  a  continuous  function.    Let 

us  consider  a  function  (f>  (x)  about  which  we  shall  only  assume  at 
present  that  it  is  defined  for  every  value  of  x  in  an  interval  (a,  b). 

The  values  assumed  by  <£  (x)  for  values  of  x  in  (a,  b)  form  an 
aggregate  S  to  which  we  can  apply  the  arguments  of  §  80,  as  we 
applied  them  in  §  81  to  the  aggregate  of  values  of  a  function  of  n. 
If  there  is  a  number  K  such  that  <j>  (x)  %  K,  for  all  values  of  x  in 
question,  we  say  that  cf>  (x)  is  bounded  above.  In  this  case  cf)  (x) 
possesses  an  upper  bound  M :  no  value  of  <£  (x)  exceeds  M,  but  any 
number  less  than  M  is  exceeded  by  at  least  one  value  of  </>  (an). 
Similarly  we  define  'bounded  below',  'lower  bound',  'bounded',  as 
applied  to  functions  of  a  continuous  variable  x. 

Theorem  1.     If  <£  (x)  is  continuous  throughout  (a,  b),  then  it  is 

bounded  in  (a,  b). 


102]  CONTINUOUS   AND    DISCONTINUOUS   FUNCTIONS  183 

We  can  certainly  determine  an  interval  (a,  £),  extending  to 
the  right  from  a,  in  which  (f>  (x)  is  bounded.  For  since  (/>  (as)  is 
continuous  for  x  =  a,  we  can,  given  any  positive  number  8  however 
small,  determine  an  interval  (a,  £)  throughout  which  <£  (x)  lies 
between  <f>(a)—8  and  <£ (a)  +  8;  and  obviously  </> (x)  is  bounded  in 
:his  interval. 

Now  divide  the  points  £  of  the  interval  (a,  b)  into  two  classes 
L,  R,  putting  £  in  L  if  </>  (£)  is  bounded  in  (a,  £),  and  in  R  if  this 
is  not  the  case.  It  follows  from  what  precedes  that  L  certainly 
exists:  what  we  propose  to  prove  is  that  R  does  not.  Suppose 
that  R  does  exist,  and  let  ft  be  the  number  corresponding  to  the 
section  whose  lower  and  upper  classes  are  L  and  R.  Since  <$>  (x) 
is  continuous  for  x  =  ft,  we  can,  however  small  8  may  be,  determine 
an  interval  (ft  —  rj,  ft  +  rj)*  throughout  which 

<f>(ft)-8<c}>(x)<(f>(ft)  +  8. 
Thus  $  (x)  is  bounded  in  (ft  —  n,  ft  +  y).  Now  ft  —  n  belongs  to  L. 
Therefore  <f)(x)  is  bounded  in  (a,  ft  —  ri):  and  therefore  it  is 
bounded  in  the  whole  interval  (a,  ft  +  ?;).  But  ft  +  77  belongs  to  R 
and  so  <£  (x)  is  not  bounded  in  (a,  ft  +  rj).  This  contradiction 
shows  that  R  does  not  exist.  And  so  <f>  (x)  is  bounded  in  the 
whole  interval  (a,  b). 

Theorem  2.  If  <fi  (x)  is  continuous  throughout  (a,  b),  and  M 
and  m  are  its  upper  and  lower  bounds,  then  <£  (x)  assumes  the  values 
M  and  m  at  least  once  each  in  the  interval. 

For,  given  any  positive  number  8,  we  can  find  a  value  of  x  for 
which  M  -  <£>  O)  <  8  or  1/{M  -  </>  (x)}  >  1/8.  Hence  1/[M  -  0  (x)} 
is  not  bounded,  and  therefore,  by  Theorem  1,  is  not  continuous. 
But  M—cf)(x)  is  a  continuous  function,  and  so  l/[M—<f>(x)}  is 
continuous  at  any  point  at  which  its  denominator  does  not  vanish 
(Ex.  xxxvil.  1).  There  must  therefore  be  one  point  at  which 
the  denominator  vanishes:  at  this  point  tf>(x)  =  M.  Similarly  it 
may  be  shown  that  there  is  a  point  at  which  0  (x)  =  in. 

The  proof  just  given  is  somewhat  subtle  and  indirect,  and  it 
may  be  well,  in  view  of  the  great  importance  of  the  theorem, 
to  indicate  alternative  lines  of  proof.  It  will  however  be  con- 
venient to  postpone  these  for  a  moment f. 

*  If  j3  =  b  we  must  replace  this  interval  by  (8-ij,  /3),  and  /3  +  i?  by  /3,  throughout 
the  argument  which  follows, 
f  See  §  101. 


184  CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  [v 

Examples  XXXVIII.  1.  If  0(a?)  =  l/#  except  when  .r  =  0,  and  <p(x)  =  0 
when  x  —  0,  then  <f>(x)  has  neither  an  upper  nor  a  lower  bound  in  any 
interval  which  includes  x  =  0  in  its  interior,  as  e.g.  the  interval  (  — 1,  +1). 

2.  If  cf>(x)  =  l/x2  except  when  x  =  0,  and  <p(x)  =  0  when  x=0,  then  <fi(x) 
has  the  lower  bound  0,  but  no  upper  bound,  in  the  interval  (—  1,  +1). 

3.  Let  <£(#)  =  sin  (1/x)  except  when  x  =  0,  and  cf>(x)  =  0  when  x—0.  Then 
(f>  (x)  is  discontinuous  for  x  —  0.  In  any  interval  ( -  8,  +  8)  the  lower  bound  is 
—  1  and  the  upper  bound  4-1,  and  each  of  these  values  is  assumed  by  $  (x)  an 
infinity  of  times. 

4.  Let  <p  (x)  =  x  -  [x].  This  function  is  discontinuous  for  all  integral 
values  of  x.  In  the  interval  (0,  1)  its  lower  bound  is  0  and  its  upper  bound  1. 
It  is  equal  to  0  when  a?=0  or  x=\,  but  it  is  never  equal  to  1.  Thus  <j>(x) 
never  assumes  a  value  equal  to  its  upper  bound. 

5.  Let  <p(x)  =  0  when  x  is  irrational,  and  $  (•v)  =  q  when  x  is  a  rational 
fraction  pjq.  Then  <£  (x)  has  the  lower  bound  0,  but  no  upper  bound,  in  any 
interval  (a,  b).  But  if  <£  (x)  =  (  —  l)p  q  when  x=pjq,  then  <f>  (x)  has  neither  an 
upper  nor  a  lower  bound  in  any  interval. 

103.     The  oscillation  of  a  function  in  an  interval.     Let 

(f)  (%)  be  any  function  bounded  throughout  (a,  b),  and  M  and  m 
its  upper  and  lower  bounds.  We  shall  now  use  the  notation 
M  (a,  b),  m(a,  b)  for  M,  m,  in  order  to  exhibit  explicitly  the  de- 
pendence of  M  and  in  on  a  and  b,  and  we  shall  write 

0(a,  b)  =  M  (a,  b)  -  m  {a,  b). 

This  number  0  (a,  b),  the  difference  between  the  upper  and 
lower  bounds  of  <f>  (a)  in  (a,  b),  we  shall  call  the  oscillation  of  </>  (x) 
in  (a,  b).  The  simplest  of  the  properties  of  the  functions  M  (a,  b), 
m  (a,  b),   0  (a,  b)  are  as  follows. 

(1)  If  a  ^  c  £  b  then  M  (a,  b)  is  equal  to  the  greater  of  M  (a,  c) 
and  M(c,  b),  and  m  (a,  b)  to  the  lesser  of  m  (a,  c)  and  m  (c,  b). 

(2)  M  (a,  b)  is  an  increasing,  in  (a,  b)  a  decreasing,  and  0  (a,  b) 
an  increasing  function  of  b. 

(3)  0(a,b)^0(a,c)  +  0(c,b). 

The  first  two  theorems  are  almost  immediate  consequences  of 
our  definitions.  Let  yu,  be  the  greater  of  M  (a,  c)  and  M  (c,  b),  and 
let  B  be  any  positive  number.  Then  </>  (x)  ^  /u,  throughout  (a,  c) 
and  (c,  b),  and  therefore  throughout  (a,  b) ;  and  <p  (w)  >  /x  -  8 
somewhere  in  (a,  c)  or  in  (c,  6),  and  therefore  somewhere  in  (a,  b). 


102-105]     CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  185 

Hence  M  (a,  b)  =  /x.     The  proposition  concerning  m  may  be  proved 
similarly.     Thus  (1)  is  proved,  and  (2)  is  an  obvious  corollary. 

Suppose  now  that  Mx  is  the  greater  and  M2  the  less  of  if  (a,  c) 
and  M  (c,  b),  and  that  my  is  the  less  and  m,  the  greater  of  m  (a,  c) 
and  m  (c,  b).  Then,  since  c  belongs  to  both  intervals,  </>  (c)  is  not 
greater  than  M2  nor  less  than  m,.  Hence  M2  ^  m2,  whether  these 
numbers  correspond  to  the  same  one  of  the  intervals  (a,  c)  and 
(c,  b)  or  not,  and 

0(a,b)  =  My  -  wi,  ^  Mx  +  M2  -  w,  -  m2. 
But  0  (a,  c)  +  0  (c,  b)  =  Mx  +  M,  -  mx  -  m, ; 

and  (3)  follows. 

104.  Alternative  proofs  of  Theorem  2  of  §  102.  The  most  straight- 
forward proof  of  Theorem  2  of  §  102  is  as  follows.  Let  £  be  any  number  of 
the  interval  (a,  b).  The  function  J/  (a,  £)  increases  steadily  with  £  and  never 
exceeds  M.  We  can  therefore  construct  a  section  of  the  numbers  £  by 
putting  £  in  L  or  in  R  according  as  M  (a,  £)  <M  or  M(a,  £)  =  M.  Let  (i  be 
the  number  corresponding  to  the  section.     If  a<8  <b,  we  have 

M(a,8-r,)<M,     M (a,  8  + 1,)  =  J/ 
for  all  positive  values  of  77,  and  so 

#(/3-i7,  I3  +  t,)  =  M, 
by  (1)  of  §  103.     Hence  $  (.r)  assumes,  for  values  of  x  as  near  as  we  please  to 
8,  values  as  near  as  we  please  to  M,  and  so,  since  $  (x)  is  continuous,  (f)  (8) 
must  be  equal  to  M. 

If  8  =  a  then  J/ (a,  a  +17)=  J/.  And  if  /3  =  6  then  M  (a,  6-17)  <  J/,  and 
so  Jf  (6  —  77,  b)  =  M.  In  either  case  the  argument  may  be  completed  as 
before. 

The  theorem  may  also  be  proved  by  the  method  of  repeated  bisection 
used  in  §  71.  If  M  is  the  upper  bound  of  cf>  (x)  in  an  interval  PQ,  and  PQ 
is  divided  into  two  equal  parts,  then  it  is  possible  to  find  a  half  Py  Qy  in  which 
the  upper  bound  of  $  (x)  is  also  M.  Proceeding  as  in  §  71,  we  construct  a 
sequence  of  intervals  PQ,  PyQy,  P2Q2,  •  ••  in  each  of  which  the  upper  bound 
of  0  (x)  is  M.  These  intervals,  as  in  §  71,  converge  to  a  point  T,  and  it  is 
easily  proved  that  the  value  of  <fi  (x)  at  this  point  is  M. 

105.  Sets  of  intervals  on  a  line.  The  Heine-Borel 
Theorem.  We  shall  now  proceed  to  prove  some  theorems  con- 
cerning the  oscillation  of  a  function  which  are  of  a  somewhat 
abstract  character  but  of  very  great  importance,  particularly,  as 
we  shall  see  later,  in  the  theory  of  integration.  These  theorems 
depend  upon  a  general  theorem  concerning  intervals  on  a  line. 


186  CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  [V 

Suppose  that  we  are  given  a  set  of  intervals  in  a  straight 
line,  that  is  to  say  an  aggregate  each  of  whose  members  is  an 
interval  (a,  ft).  We  make  no  restriction  as  to  the  nature  of 
these  intervals ;  they  may  be  finite  or  infinite  in  number ;  they 
may  or  may  not  overlap*;  and  any  number  of  them  may  be 
included  in  others- 
It  is  worth  while  in  passing  to  give  a  few  examples  of  sets  of  intervals  to 
which  we  shall  have  occasion  to  return  later. 

(i)  If  the  interval  (0,  1)  is  divided  into  n  equal  parts  then  the  n  intervals 
thus  formed  define  a  finite  set  of  non-overlapping  intervals  which  just  cover 
up  the  line. 

(ii)  We  take  every  point  £  of  the  interval  (0,  1),  and  associate  with  £  the 
interval  (£  — e,  £  +  e),  where  e  is  a  positive  number  less  than  1,  except  that 
with  0  we  associate  (0,  e)  and  with  1  we  associate  (1  —  e,  1),  and  in  general  we 
reject  any  part  of  any  interval  which  projects  outside  the  interval  (0,  1).  We 
thus  define  an  infinite  set  of  intervals,  and  it  is  obvious  that  many  of  them 
overlap  with  one  another. 

(iii)  We  take  the  rational  points  pjq  of  the  interval  (0,  1),  and  associate 
\M\th.  plq  the  interval 

(P_e     P  +  l 

\q      if     q      q- 

where  e  is  positive  and  less  than  1.  We  regard  0  as  0/1  and  1  as  1/1  :  in 
these  two  cases  we  reject  the  part  of  the  interval  which  lies  outside  (0,  1).  We 
obtain  thus  an  infinite  set  of  intervals,  which  plainly  overlap  with  one  another, 
since  there  are  an  infinity  of  rational  points,  other  than  p/q,  in  the  interval 
associated  with  pjq. 

The  Heine-Borel  Theorem.  Suppose  that  we  are  given  an 
interval  (a,  b),  and  a  set  of  intervals  I  each  of  whose  members  is 
included  in  (a,  b).  Suppose  further  that  I  possesses  the  following 
properties : 

(i)  every  point  of  (a,  b),  other  than  a  and  b,  lies  inside~\  at 
least  one  interval  of  I ; 

(ii)  a  is  the  left-hand  end  point,  and  b  the  right-hand  end 
point,  of  at  least  one  interval  of  I. 

Then  it  is  possible  to  choose  a  finite  number  of  intervals  from 
the  set  I  which  form  a  set  of  intervals  possessing  the  properties  (i) 
and  (ii). 

*  The  word  overlap  is  used  in  its  obvious  sense :  two  intervals  overlap  if  they 
have  points  in  common  which  are  not  end  points  of  either.  Thus  (0,  f)  aud  (|,  1) 
overlap.     A  pair  of  intervals  such  as  (0,  ^)  and  (£,  1)  may  be  said  to  abut. 

t  That  is  to  say  '  in  and  not  at  an  end  of. 


105]  CONTINUOUS   AND   DISCONTINUOUS    FUNCTIONS  187 

We  know  that  a  is  the  left-hand  end  point  of  at  least  one 
interval  of  /,  say  (a,  a^.  We  know  also  that  at  lies  inside  at  least 
one  interval  of  /,  say  (a/,  a2).  Similarly  a2  lies  inside  an  interval 
(a.2,  a3)  of  /.  It  is  plain  that  this  argument  may  be  repeated  in- 
definitely, unless  after  a  finite  number  of  steps  an  coincides  with  b. 

If  an  does  coincide  with  b  after  a  finite  number  of  steps  then 
there  is  nothing  further  to  prove,  for  we  have  obtained  a  finite  set 
of  intervals,  selected  from  the  intervals  of  /,  and  possessing  the 
properties  required.  If  an  never  coincides  with  b,  then  the  points 
a1}  «2,  a3,  ...  must  (since  each  lies  to  the  right  of  its  predecessor) 
tend  to  a  limiting  position,  but  this  limiting  position  may,  so  far 
as  we  can  tell,  lie  anywhere  in  (a,  b). 

Let  us  suppose  now  that  the  process  just  indicated,  starting 
from  a,  is  performed  in  all  possible  ways,  so  that  we  obtain  all 
possible  sequences  of  the  type  a1}  a2,  a3,  ....  Then  we  can  prove 
that  there  must  be  at  least  one  suck  sequence  which  arrives  at  b 
after  a  finite  number  of  steps. 


S    «s         4'      Sn      4"  \ 


b0 

Fig.  33. 


There  are  two  possibilities  with  regard  to  any  point  £  between 
a  and  b.  Either  (i)  £  lies  to  the  left  of  some  point  an  of  some 
sequence  or  (ii)  it  does  not.  We  divide  the  points  £  into  two 
classes  L  and  R  according  as  to  whether  (i)  or  (ii)  is  true.  The 
class  L  certainly  exists,  since  all  points  of  the  interval  (a,  o2) 
belong  to  L.  We  shall  now  prove  that  R  does  not  exist,  so  that 
every  point  f  belongs  to  L. 

If  R  exists  then  L  lies  entirely  to  the  left  of  R,  and  the  classes 
L,  R  form  a  section  of  the  real  numbers  between  a  and  b,  to 
which  corresponds  a  number  |0.  The  point  £0  lies  inside  an  interval 
of  /,  say  (£',  f "),  and  £'  belongs  to  L,  and  so  lies  to  the  left  of 
some  term  an  of  some  sequence.  But  then  we  can  take  (£',  £") 
as  the  interval  (an',  an+1)  associated  with  an  in  our  construction 
of  the  sequence  a1}  a2,  a3,  ...;  and  all  points  to  the  left  of  £" 
lie  to  the  left  of  an+1.  There  are  therefore  points  of  L  to  the 
right  of  £0,  and  this  contradicts  the  definition  of  R.  It  is 
therefore  impossible  that  R  should  exist. 


1S8  CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  [v 

Thus  every  point  £  belongs  to  L.  Now  b  is  the  right-hand 
end  point  of  an  interval  of  i",  say  (61;  b),  and  bx  belongs  to  L. 
Hence  there  is  a  member  an  of  a  sequence  a1}  a.2>  az,  ...  such  that 
an  >  bi .  But  then  we  may  take  the  interval  (an't  an+i)  corre- 
sponding to  an  to  be  (b1}  b),  and  so  Ave  obtain  a  sequence  in  which 
the  term  after  the  nth  coincides  with  b,  and  therefore  a  finite  set 
of  intervals  having  the  properties  required.  Thus  the  theorem  is 
proved. 

It  is  instructive  to  consider  the  examples  of  p.  186  in  the  light  of  this 
theorem. 

(i)  Here  the  conditions  of  the  theorem  are  not  satisfied  the  points 
1/?!,  2/h,  3/n,  ...  do  not  lie  inside  any  interval  of  I 

(ii)  Here  the  conditions  of  the  theorem  are  satisfied.  The  set  of 
intervals 

(0,  20,  (e,3«),  (2f,  4C),  ...,  (1-2*,  1), 

associated  with  the  points  e,  2e,  3e,  ...,  1  - e,  possesses  the  properties  re- 
quired. 

(iii)  In  this  case  we  can  prove,  by  using  the  theorem,  that  there  are, 
if  «  is  small  enough,  points  of  (0,  1)  which  do  not  lie  in  any  interval  of  I. 

If  every  point  of  (0,  1)  lay  inside  an  interval  of  /  (with  the  obvious 
reservation  as  to  the  end  points),  then  we  could  find  a  finite  number  of  intervals 
of  /  possessing  the  same  property  and  having  therefore  a  total  length  greater 
than  1.  Now  there  are  two  intervals, of  total  length  2e,  for  which  g,=  l,  and 
q-\  intervals,  of  total  length  2e(q—l)/q?',  associated  with  any  other  value 
of  q.  The  sum  of  any  finite  number  of  intervals  of  /  can  therefore  not  be 
greater  than  2e  times  that  of  the  series 

,      1       2       3 
1+23  +  33  +  43+  .». 

which  will  be  shown  to  be  convergent  in  Ch.  VIII.  Hence  it  follows  that,  if 
e  is  small  enough,  the  supposition  that  every  point  of  (0,  1)  lies  inside  an 
interval  of  /  leads  to  a  contradiction. 

The  reader  may  be  tempted  to  think  that  this  proof  is  needlessly 
elaborate,  and  that  the  existence  of  points  of  the  interval,  not  in  any  interval 
of  /,  follows  at  once  from  the  fact  that  the  sum  of  all  these  intervals  is  less 
than  1.  But  the  theorem  to  which  he  would  be  appealing  is  (when  the  set  of 
intervals  is  infinite)  far  from  obvious,  and  can  only  be  proved  rigorously  by 
some  such  use  of  the  Heine-Borel  Theorem  as  is  made  in  the  text. 

106.  We  shall  now  apply  the  Heine-Borel  Theorem  to  the 
proof  of  two  important  theorems  concerning  the  oscillation  of  a 
continuous  function. 


105,  106]      CONTINUOUS    AND    DISCONTINUOUS    FUNCTIONS  189 

Theorem  I.  If  <j>  (as)  is  continuous  throughout  the  interval 
(a,  b),  then  we  can  divide  (a,  b)  into  a  finite  number  of  sub-intervals 
(a,  #,),  (#!,  :c2),  ...  (ccn,  b),  in  each  of  which  the  oscillation  of  <j>(x)  is 
less  than  an  assigned  positive  number  8. 

Let  £  be  any  number  between  a  and  b.  Since  <£  (#)  is  con- 
tinuous for  x  =  £,  we  can  determine  an  interval  (f  —  e,  £  +  e)  such 
that  the  oscillation  of  <£  (#)  in  this  interval  is  less  than  8.  It  is 
indeed  obvious  that  there  are  an  infinity  of  such  intervals  corre- 
sponding to  every  £  and  every  8,  for  if  the  condition  is  satisfied  for 
any  particular  value  of  e,  then  it  is  satisfied  a  fortiori  for  any  smaller 
value.  What  values  of  e  are  admissible  will  naturally  depend  upon 
£;  we  have  at  present  no  reason  for  supposing  that  a  value  of  e 
admissible  for  one  value  of  £  will  be  admissible  for  another.  We 
shall  call  the  intervals  thus  associated  with  £  the  8-intervals  of  %. 

If  f-  =  a  then  we  can  determine  an  interval  (a,  a  +  e).  and  so  an 
infinity  of  such  intervals,  having  the  same  property.  These  we 
call  the  8-intervals  of  a,  and  we  can  define  in  a  similar  manner  the 
S-intervals  of  b. 

Consider  now  the  set  I  of  intervals  formed  by  taking  all  the 
S-intervals  of  all  points  of  (a,  b).  It  is  plain  that  this  set  satisfies 
the  conditions  of  the  Heine-Borel  Theorem ;  every  point  interior 
to  the  interval  is  interior  to  at  least  one  interval  of  /,  and  a  and  b 
are  end  points  of  at  least  one  such  interval.  We  can  therefore 
determine  a  set  /'  which  is  formed  by  a  finite  number  of  intervals 
of  /,  and  which  possesses  the  same  property  as  /  itself. 

The  intervals  which  compose  the  set  /'  will  in  general  overlap 

as  in  Fig.  34.     But  their  end  

points    obviously    divide    up         -  =Z^ZI — ^ZZ. 

(a,  b)  into  a  finite  set  of  in-     a  b 

tervals   I"   each  of   which  is 

included  in  an  interval  of  /',  and  in  each  of  which  the  oscillation 
of  (f>  (x)  is  less  than  8.     Thus  Theorem  I  is  proved. 

Theorem  II.  Given  any  positive  number  8,  we  can  find  a 
number  rj  such  that,  if  the  interval  (a,  b)  is  divided  in  any  manner 
into  sub-intervals  of  length  less  than  v,  then  the  oscillation  of  <j>(x) 
in  each  of  them  will  be  less  than  8. 


190  CONTINUOUS   AND   DISCONTINUOUS    FUNCTIONS  [V 

Take  8V  <  £8,  and  construct,  as  in  Theorem  I,  a  finite  set  of  sub- 
intervals  j  in  each  of  which  the  oscillation  of  <f>  (x)  is  less  than  8X. 
Let  77  be  the  length  of  the  least  of  these  sub-intervals  j.  If 
now  we  divide  (a,  b)  into  parts  each  of  length  less  than  77,  then  any 
such  part  must  lie  entirely  within  at  most  two  successive  sub- 
intervals  j.  Hence,  in  virtue  of  (3)  of  §  103,  the  oscillation  of  <f>  (x), 
in  one  of  the  parts  of  length  less  than  77,  cannot  exceed  twice  the 
greatest  oscillation  of  </>  (x)  in  a  sub-interval  j,  and  is  therefore 
less  than  281}  and  therefore  than  8. 

This  theorem  is  of  fundamental  importance  in  the  theory  of 
definite  integrals  (Ch.  VII).  It  is  impossible,  without  the  use  of 
this  or  some  similar  theorem,  to  prove  that  a  function  continuous 
throughout  an  interval  necessarily  possesses  an  integral  over  that 
interval. 

107.     Continuous   functions  of  several   variables.     The 

notions  of  continuity  and  discontinuity  may  be  extended  to 
functions  of  several  independent  variables  (Ch.  II,  §§  31  et  seg.). 
Their  application  to  such  functions,  however,  raises  questions 
much  more  complicated  and  difficult  than  those  which  we  have 
considered  in  this  chapter.  It  would  be  impossible  for  us  to 
discuss  these  questions  in  any  detail  here  ;  but  we  shall,  in  the 
sequel,  require  to  know  what  is  meant  by  a  continuous  function  of 
two  variables,  and  we  accordingly  give  the  following  definition. 
It  is  a  straightforward  generalisation  of  the  last  form  of  the  de- 
finition of  §  98. 

The  function  (j>(x,  y)  of  the  two  variables  x  and  y  in  said  to  be 
continuous  for  x=  £,  y  =  77  if  given  any  positive  number  8,  how- 
ever small,  we  can  choose  e  (8)  so  that 

\(f>(x,y)-(f)  (|,  77)  j  <  8 

when  0  ^  |  x  —  f  |  ^  e  (8)  and  0  g  |  y  —  77 1  s  e  (8);  that  is  to  say  if  we 
can  draw  a  square,  ivhose  sides  are  parallel  to  the  axes  of  coordinates 
and  of  length  2e  (8),  whose  centre  is  the  point  (£,  77),  and  which  is  such 
that  the  value  of  c}>  (x,  y)  at  any  point  inside  it  or  on  its  boundary 
differs  from  <$>  (£,  77)  by  less  than  8* 

This  definition  of  course  presupposes  that  (f>  (x,  y)  is  defined  at 
all  points  of  the  square  in  question,  and  in  particular  at  the  point 

*  The  reader  should  draw  a  figure  to  illustrate  the  definition. 


106-108]     CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  191 

(£  77).  Another  method  of  stating  the  definition  is  this :  <f>  (x,  y)  is 
continuous  for  x=£,  y=v  if  <f>(x,y)  ^</>(£,  n)  ivhen  x+\,  y^v 
in  any  manner.  This  statement  is  apparently  simpler;  but  it 
contains  phrases  the  precise  meaning  of  which  has  not  yet  been 
explained  and  can  only  be  explained  by  the  help  of  inequalities 
like  those  which  occur  in  our  original  statement. 

It  is  easy  to  prove  that  the  sums,  the  products,  and  in  general 
the  quotients  of  continuous  functions  of  two  variables  are  them- 
selves continuous.  A  polynomial  in  two  variables  is  continuous  for 
all  values  of  the  variables ;  and  the  ordinary  functions  of  x  and  y 
which  occur  in  every-day  analysis  are  generally  continuous,  i.e. 
are  continuous  except  for  pairs  of  values  of  x  and  y  connected  by 
special  relations. 

The  reader  should  observe  carefully  that  to  assert  the  continuity  of 
<f>  (x,  y)  with  respect  to  the  two  variables  x  and  y  is  to  assert  much  more 
than  its  continuity  with  respect  to  each  variable  considered  separately.  It  is 
plain  that  if  <£  (x,  y)  is  continuous  with  respect  to  x  and  y  then  it  is  certainly 
continuous  with  respect  to  x  (or  y)  when  any  fixed  value  is  assigned  to  y 
(or  .r).     But  the  converse  is  by  no  means  true.     Suppose,  for  example,  that 

(b  (x,  y)  =    ., 

when  neither  x  nor  y  is  zero,  and  (p  (x,  y)=0  when  either  x  or  y  is  zero.  Then 
if  y  has  any  fixed  value,  zero  or  not,  <f>  (x,  y)  is  a  continuous  function  of  x, 
and  in  particular  continuous  for  x  =  0;  for  its  value  when  x=0  is  zero,  and  it 
tends  to  the  limit  zero  as  x-*-0.  In  the  same  way  it  may  be  shown  that 
(p  (x,  y)  is  a  continuous  function  of  y.  But  q>  (x,  y)  is  not  a  continuous  function 
of  x  and  y  for  aj=0,  y  =  0.  Its  value  when  x=0,  y  =  0  is  zero  ;  but  if  x  and 
y  tend  to  zero  along  the  straight  line  y  =  ax,  then 

0  (*i  y) = T-r-t »    lim  4>  to  y) = rrn;  • 

J.  +  it-  1  +  a" 

which  may  have  any  value  between  —  1  and  1. 

108.  Implicit  functions.  We  have  already,  in  Ch.  II,  met  with 
the  idea  of  an  implicit  function.     Thus,  if  x  and  y  are  connected   by  the 

relation 

yr'-xy-y-x  =  0 (1), 

then  y  is  an  'implicit  function'  of  x. 

But  it  is  far  from  obvious  that  such  an  equation  as  this  does  really  define 
a  function  y  of  x,  or  several  such  functions.  In  Ch.  II  we  were  content  to 
take  this  for  granted.  We  are  now  in  a  position  to  consider  whether  the 
assumption  we  made  then  was  justified. 


192 


CONTINUOUS    AND    DISCONTINUOUS    FUNCTIONS 


[V 


We  shall  find  the  following  terminology  useful.  Suppose  that  it  is  possible 
to  surround  a  point  («,  b),  as  in  §  107,  with  a  square  throughout  which 
a  certain  condition  is  satisfied.  We  shall  call  such  a  square  a  neighbourhood 
of  (a,  6),  and  say  that  the  condition  in  question  is  satisfied  in  the  neighbour- 
hood of  {a,  b),  or  near  (a,  b),  meaning  by  this  simply  that  it  is  possible  to  find 
some  square  throughout  which  the  condition  is  satisfied.  It  is  obvious  that 
similar  language  may  be  used  when  we  are  dealing  with  a  single  variable,  the 
square  being  replaced  by  an  interval  on  a  line. 

Theorem.  If  (i)  f(x,  y)  is  a  continuous  function  of  x  and  y  in  the 
neighbourhood  of  (a,  b), 

(ii)    f{a,b)=0, 

(iii)  fix,  y)  is,  for  all  values  of  x  in  the  neighbourhood  of  a,  a  steadily 
increasing  function  of  y,  in  the  stricter  sense  of  %  95, 

then  (1)  there  is  a  unique  function  y  =  cf>  (x)  which,  when  substituted  in  the 
equation  f(x,  y)=0,  satisfies  it  identically  for  all  values  of  x  in  the  neighbour- 
hood of  a, 

(2)  <p  (x)  is  continuous  for  all  values  of  x  in  the  neighbourhood  of  a. 

In  the  figure  the  square  represents  a  '  neighbourhood '  of  (a,  b)  through- 
out which  the  conditions  (i)  and  (iii)  are 
satisfied,  and  P  the  point  (a,  b).  If  we 
take  Q  and  R  as  in  the  figure,  it  follows  from 
(iii)  that/(#,  y)  is  positive  at  Q  and  negative 
at  R.  This  being  so,  and  f(x,  y)  being  con- 
tinuous at  Q  and  at  R,  we  can  draw  lines  Qf/ 
and  RR!  parallel  to  OX,  so  that  R'Q'  is  parallel 
to  OY  and  f(x,  y)  is  positive  at  all  points  of 
QQ'  and  negative  at  all  points  of  RR'.  In  par- 
ticular f(x,  y)  is  positive  at  Q'  and  negative  at 
R',  and  therefore,  in  virtue  of  (iii)  and  §  100, 
vanishes  once  and  only  once  at  a  point  P'  on 
/<"(/.  The  same  construction  gives  us  a  unique  point  at  which  f(x,y)  =  0 
on  each  ordinate  between  RQ  and  R'Q'.  It  is  obvious,  moreover,  that  the 
same  construction  can  be  carried  out  to  the  left  of  RQ.  The  aggregate  of 
points  such  as  P'  gives  us  the  graph  of  the  required  function  y  =  (f>(x). 

It  remains  to  prove  that  <£  (x)  is  continuous.  This  is  most  simply  effected 
by  using  the  idea  of  the  'limits  of  indetermination  '  of  (f>  (x)  as  x-^a  (J)  96). 
Suppose  that  x-*-a,  and  let  X  and  A  be  the  limits  of  indetermination  of  $  (x) 
as  x-*-a.  It  is  evident  that  the  points  (a,  X)  and  (a,  A)  lie  on  QR.  Moreover, 
we  can  find  a  sequence  of  values  of  x  such  that  (f>  (.>•)-»- X  when  x-*-a  through 
the  values  of  the  sequence;  and  since  /  {x,  <f>  (x)\  =0,  and  f(x,y)  is  a  con- 
tinuous function  of  x  and  y,  we  have 

/(«,X)  =  0. 
Hence  X  =  7>;  and  similarly  A=6.     Thus  cf>  (x)  tends  to  the  limit  b  as  x-*-a, 
and  so  $>(x)  is  continuous  for  x=a.     It  is  evident  that  we  can  show  in 


Q 

Q.' 

^ 

P' 

(a.b) 

'? 

R' 

R 

Fig.  35. 


108-109]     CONTINUOUS   AND   DISCONTINUOUS    FUNCTIONS  193 

exactly  the  same  way  that  <j>  (x)  is  continuous  for  any  value  of  x  in  the 
neighbourhood  of  a. 

It  is  clear  that  the  truth  of  the  theorem  would  not  be  affected  if  we  were 
to  change  'increasing'  to  'decreasing5  in  condition  (iii). 

As  an  example,  let  us  consider  the  equation  (1),  taking  a  =  0,  6=0.  It  is 
evident  that  the  conditions  (i)  and  (ii)  are  satisfied.     Moreover 

/(^y)-/(^y')=(y-y)(3/4+3/¥+^y2+yy3+y4-x-i) 

has,  when  x,  y,  and  y'  are  sufficiently  small,  the  sign  opposite  to  that  of 
y-y'.  Hence  condition  (iii)  (with  'decreasing'  for  'increasing')  is  satisfied. 
It  follows  that  there  is  one  and  only  one  continuous  function  y  which 
satisfies  the  equation  (1)  identically  and  vanishes  with  x. 

The  same  conclusion  would  follow  if  the  equation  were 
y2-xy-y-x  =  0. 
The  function  in  question  is  in  this  case 

y  =  h  {1  +x-  s/(l  +  6x+x2)}, 
where  the  square  root  is  positive.    The  second  root,  in  which  the  sign  of  the 
square  root  is  changed,  does  not  satisfy  the  condition  of  vanishing  with  x. 

There  is  one  point  in  the  proof  which  the  reader  should  be  careful  to  ob- 
serve. We  supposed  that  the  hypotheses  of  the  theorem  were  satisfied  'in 
the  neighbourhood  of  (a,  b)\  that  is  to  say  throughout  a  certain  square 
^-f  =  ^  =  ^  +  f]  ^-(gj/<?)  +  e.  The  conclusion  holds  'in  the  neighbourhood 
of  a: = a',  that  is  to  say  throughout  a  certain  interval  £  —  et  <  x  <  £  +  e x .  There 
is  nothing  to  show  that  the  el  of  the  conclusion  is  the  e  of  the  hypotheses,  and 
indeed  this  is  generally  untrue. 

109.  Inverse  Functions.  Suppose  in  particular  that/(#,  y)  is  of  the 
form  F(y)  -  x.     We  then  obtain  the  following  theorem. 

If  F(y)  is  a  function  of y,  continuous  and  steadily  increasing  {or  decreasing), 
in  the  stricter  sense  of  §  95,  in  the  neighbourhood  of  y  =  b,  and  F(b)  =  a,  then 
there  is  a  unique  continuous  function  y  =  <P  (x)  which  is  equal  to  b  when  x  =  a 
and  satisfies  the  equation  F(y)  =  x  identically  in  the  neighbourhood  of  x=a. 

The  function  thus  defined  is  called  the  inverse  function  of  F{y). 

Suppose  for  example  that  y3  =  x,  a  =  0,  b  =  0.  Then  all  the  conditions  of 
the  theorem  are  satisfied.     The  inverse  function  is  x=f/y. 

If  we  had  supposed  that  y2  =  x  then  the  conditions  of  the  theorem  would 
not  have  been  satisfied,  for  y2  is  not  a  steadily  increasing  function  of  y  in  any 
interval  which  includes  y  =  0:  it  decreases  when  y  is  negative  and  increases 
when  y  is  positive.  And  in  this  case  the  conclusion  of  the  theorem  does  not 
hold,  for  y2=x  defines  two  functions  of  x,  viz.  y=*Jx  and  y=  -Jx,  both  of 
which  vanish  when  x—0,  and  each  of  which  is  defined  only  for  positive  values 
of  x,  so  that  the  equation  has  sometimes  two  solutions  and  sometimes  none. 
The  reader  should  consider  the  more  general  equations 

yin=x,    y2n  +  1  =  x, 
H.  13 


194  CONTINUOUS   AND   DISCONTINUOUS    FUNCTIONS  [V 

in  the  same  way.    Another  interesting  example  is  given  by  the  equation 

yh  —y  -  .x'=0, 

already  considered  in  Ex.  xiv.  7. 

Similarly  the  equation  siny  =  x 

has  just  one  solution  which  vanishes  with  x,  viz.  the  value  of  arc  sin  a?  which 
vanishes  with  x.  There  are  of  course  an  infinity  of  solutions,  given  by  the 
other  values  of  arc  sin  x  (cf.  Ex.  xv.  10),  which  do  not  satisfy  this  condition. 

So  far  we  have  considered  only  what  happens  in  the  neighbourhood  of  a 
particular  value  of  x.  Let  us  suppose  now  that  F(y)  is  positive  and  steadily 
increasing  (or  decreasing)  throughout  an  interval  (a,  b).  Given  any  point  £ 
of  (a,  b),  we  can  determine  an  interval  i  including  £,  and  a  unique  and  con- 
tinuous inverse  function  0,-  (x)  defined  throughout  i. 

From  the  set  I  of  intervals  i  we  can,  in  virtue  of  the  Heine-Borel  Theorem, 
pick  out  a  finite  sub-set  covering  up  the  whole  interval  (a,  b)  ;  and  it  is  plain 
that  the  finite  set  of  functions  <£;  (x),  corresponding  to  the  sub-set  of  intervals  i 
thus  selected,  define  together  a  unique  inverse  function  (f>  (x)  continuous 
throughout  (a,  b). 

We  thus  obtain  the  theorem  :  if  x  =  F(y),  where  F(y)  is  continuous  and 
increases  steadily  and  strictly  from  A  to  B  as  x  increases  from  a  to  6,  then  there 
is  a  unique  inverse  function  y  =  (j)  (x)  which  is  continuous  and  increases  steadily 
and  strictly  from  a  to  b  as  x  increases  from  A  to  B. 

It  is  worth  while  to  show  how  this  theorem  can  be  obtained  directly  with- 
out the  help  of  the  more  difficult  theorem  of  §  108.  Suppose  that  A  <£<B, 
and  consider  the  class  of  values  of  y  such  that  (i)  a  <y  <  b  and  (ii)  F(y)  5£  £. 
This  class  has  an  upper  bound  rj,  and  plainly  F(rj)2k£.  If  F(rj)  were  less 
than  £,  we  could  find  a  value  of  y  such  that  y  >  r)  and  F(y)  <  £,  and  rj  would 
not  be  the  upper  bound  of  the  class  considered.  Hence  F(j))  =  g.  The 
equation  F{y)  =  £  has  therefore  a  unique  solution  y  =  T)  =  <p(£),  say;  and 
plainly  rj  increases  steadily  and  continuously  with  £,  which  proves  the  theorem. 

MISCELLANEOUS  EXAMPLES  ON  CHAPTER  V. 

1.  Show  that,  if  neither  a  nor  b  is  zero,  then 

axn  +  bxn  ~x  + . . .  +k = axn  (1  +  fx), 
where  ex  is  of  the  first  order  of  smallness  when  x  is  large. 

2.  If  P  (x)  =  ax11  +  bxn  ~ 1  + . . .  +  fc,  and  a  is  not  zero,  then  as  x  increases 
P(x)  has  ultimately  the  sign  of  a;  and  so  has  P  (x  +  \)-P(x),  where  X  is 
any  constant. 

3.  Show  that  in  general 

(axn  +  bxn~1  +  ...+k)/{Axn  +  Bxn~l  + ...  +  K)=a  +  ((3/x)  (I  +  ex), 

where  a  =  a/A,  ji  =  {bA  -aB)/A2,  and  ex  is  of  the  first  order  of  smallness  when 
v  is  large.     Indicate  any  exceptional  cases. 


CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  195 

4.  Express  (ax2+bx  + c)/(Ax2  +  Bx+C) 
in  the  form  a  +  (fi/s)  +  (y/x2)  (1  +  ex), 
where  ex  is  of  the  first  order  of  smallness  when  x  is  large. 

5.  Show  that  li  m  s?x  {sl(x + a)  -  »Jx}  =  ^a. 

[Use  the  formula  ^(x  +  a)  —  K/x  =  al{S/'(x  +  a)  +  s,fx}.] 

6.  Show  that  J(x+a)  =  sfx  +  ^  (a/s/x)  (1  +  ex),  where  ex  is  of  the  first  order 
of  smallness  when  x  is  large. 

7.  Find  values  of  a  and  j3  such  that  J  (ax-  +  2bx+c)  -  ax  -  ft  has  the  limit 
zero  as  x-*-cc  ;  and  prove  that  liuxx{^/(ax2  +  2bx  +  c)  —  ax— /3}  =  (ac  -  62)/2a. 

8.  Evaluate  lim  .t'{s/[lr2  +  N^(.f4  +  l)]-.rx/2}. 

9.  Prove  that  (sec  x  —  tan  x)  -»-  0  as  x  -»-  \  «-. 

10.  Prove  that  $  (x)  =  1  -  cos  (1  -  cos  x)  is  of  the  fourth  order  of  smallness 
when  x  is  small ;  and  find  the  limit  of  (j>  (x)^  as  x-*~Q. 

11.  Prove  that  <f>  (x)—x sin  (sin x)  -  sin2 x  is  of  the  sixth  order  of  smallness 
when  x  is  small  ;  and  find  the  limit  of  $  ^jx6  as  x-*~Q. 

12.  From  a  point  P  on  a  radius  OA  of  a  circle,  produced  beyond  the  circle, 
a  tangent  PT  is  drawn  to  the  circle,  touching  it  in  T,  and  TN  is  drawn  per- 
pendicular to  OA.     Show  that  NA  jAP-*~l  as  P  moves  up  to  A. 

13.  Tangents  are  drawn  to  a  circular  arc  at  its  middle  point  and  its 
extremities ;  A  is  the  area  of  the  triangle  formed  by  the  chord  of  the  arc  and 
the  two  tangents  at  the  extremities,  and  A'  the  area  of  that  formed  by  the 
three  tangents.     Show  that  A/A'-*-4  as  the  length  of  the  arc  tends  to  zero. 

14.  For  what  values  of  a  does  {a  +  sin  (l/x)}/x  tend  to  (1)  oc ,  (2)  -  oc , 
as  x-*-01  [To  qo  if  a>\,  to  -00  if  «<-l:  the  function  oscillates  if 
-1  <«<1.] 

15.  If  <p(x)  =  l/q  when  x=plq,  and  <j)(x)  =  0  when  x  is  irrational,  then 
(f)  (x)  is  continuous  for  all  irrational  and  discontinuous  for  all  rational  values 
of  x. 

16.  Show  that  the  function  whose  graph  is  drawn  in  Fig.  32  may  be  repre- 
sented by  either  of  the  formulae 

1  - x  +  [x~\- [1  - x],     1  - x -  lim  (cos2n  + 1  irx). 

71-*-  00 

17.  Show  that  the  function  </>(#)  which  is  equal  to  0  when  #=0,  to  \  —  x 
when  0<x<^,  toi  when  x=^,  to  f  —  x  when  ^<x<\,  and  to  1  when 
,r  =  l,  assumes  every  value  between  0  and  1  once  and  once  only  as  x  increases 
from  0  to  1,  but  is  discontinuous  for  x=0,  #=£,  and  x=l.  Show  also  that 
the  function  may  be  represented  by  the  formula 

^-x  +  i[2x]-h[l-±r]. 

13—2 


196  CONTINUOUS   AND   DISCONTINUOUS   FUNCTIONS  [V 

18.  Let  <f>(x)  =  x  when  x  is  rational  and  cf>  (x)  =  1  -  x  when  x  is  irrational. 
Show  that  <f)  (x)  assumes  every  value  between  0  and  1  once  and  once  only  as  x 
increases  from  0  to  1,  but  is  discontinuous  for  every  value  of  x  except  %=\. 

19.  As  x  increases  from  —  \rt  to  \ir,  ?/  =  sin  x  is  continuous  and  steadily 
increases,  in  the  stricter  sense,  from  -  1  to  1.  Deduce  the  existence  of  a 
function  #=arc  sinj/  which  is  a  continuous  and  steadily  increasing  function 
of  y  from  y  =  -  1  to  y  =  \. 

20.  Show  that  the  numerically  least  value  of  arc  tan  y  is  continuous  for 
all  values  of  y  and  increases  steadily  from  —\n  to  \n  as  y  varies  through  all 
real  values. 

21.  Discuss,  on  the  lines  of  §§  108 — 109,  the  solution  of  the  equations 

y2-y-x=0,    y*-y*-x*=0,    tf-y*+x*=0 
in  the  neighbourhood  of  x=0,  y  =  0. 

22.  If  ax2  +  2bxy  +  cy2  +  2dx  +  2ey=0  and  A  =  2bde-ae2-cd2,  then  one 
value  of  y  is  given  by  y  =  ax  +  fix2  +  (y  +  cx)  a'3j  where 

a=-d/e,     0  =  A/2e3,     y  =  (cd -  be)  A/2e5, 

and  ex  is  of  the  first  order  of  smallness  when  x  is  small. 

[If  y-ax  =  rj  then 

—  2ex]  =  ax2  +  2bx{r)-\- ax)  +  C  (rj  +  ax)2  =  Ax2  +  2Bxrj  +  Crj2, 

say.  It  is  evident  that  rj  is  of  the  second  order  of  smallness,  xrj  of  the  third, 
and  r]2  of  the  fourth  ;  and  -  2erj  =  Ax2  —  (AB/e)  x3,  the  error  being  of  the  fourth 
order.] 

23.  If  x=ay  +  by2  +  cy3  then  one  value  of  y  is  given  by 

y  =  ax + fix2  +  (y  +  ex)  x3, 

where  a=l/a,  /3=  —b/a3,  y=(2b2-ac)/ab,  and  ex  is  of  the  first  order  of  small- 
ness when  x  is  small. 

24.  If  x  =  ay  +  byn,  where  n  is  an  integer  greater  than  unity,  then  one 
value  of  y  is  given  by  y  =  ax  +  fix11  +  (y  +  ex)  x2n~l,  where  a=l/a,  /9*=  —  6/an  +  1, 
y  =  nb2/a2,l  +  1,  and  ex  is  of  the  (n—  l)th  order  of  smallness  when  x  is  small. 

25.  Show  that  the  least  positive  root  of  the  equation  xy  =  sinx  is  a  con- 
tinuous function  of  y  throughout  the  interval  (0,  1),  and  decreases  steadily 
from  7r  to  0  as  y  increases  from  0  to  1.  [The  function  is  the  inverse  of 
(sin  x)jx :  apply  §  109.] 

26.  The  least  positive  root  of  xy  —  tan  x  is  a  continuous  function  of  y 
throughout  the  interval  (1,  cc ),  and  increases  steadily  from  0  to  |?r  as  y 
increases  from  1  towards  oo . 


CHAPTER    VI 

DERIVATIVES    AND    INTEGRALS 

110.  Derivatives  or  Differential  Coefficients.  Let  us  return 
to  the  consideration  of  the  properties  which  we  naturally  associate 
with  the  notion  of  a  curve.  The  first  and  most  obvious  property 
is,  as  we  saw  in  the  last  chapter,  that  which  gives  a  curve  its 
appearance  of  connectedness,  and  which  we  embodied  in  our  defini- 
tion of  a  continuous  function. 

The  ordinary  curves  which  occur  in  elementary  geometry,  such 
as  straight  lines,  circles  and  conic  sections,  have  of  course  many 
other  properties  of  a  general  character.  The  simplest  and  most 
noteworthy  of  these  is  perhaps  that  they  have  a  definite  direction 
at  every  point,  or  what  is  the  same  thing,  that  at  every  point  of 
the  curve  we  can  draw  a  tangent  to  it.  The  reader  will  probably 
remember  that  in  elementary  geometry  the  tangent  to  a  curve  at 
P  is  defined  to  be  '  the  limiting  position  of  the  chord  PQ,  when  Q 
moves  up  towards  coincidence  with  P '.  Let  us  consider  what  is 
implied  in  the  assumption  of  the  existence  of  such  a  limiting- 
position. 

In  the  figure  (Fig.  36)  P  is  a  fixed  point  on  the  curve,  and  Q 
a  variable  point;  PM,  QN  are  parallel  to  0Y  and  PR  to  OX. 
We  denote  the  coordinates  of  P  by  x,  y  and  those  of  Q  by 
%  +  h,  y  +  k:  h  will  be  positive  or  negative  according  as  N  lies  to 
the  right  or  left  of  M. 

We  have  assumed  that  there  is  a  tangent  to  the  curve  at  P, 
or  that  there  is  a  definite  '  limiting  position '  of  the  chord  PQ. 
Suppose  that  PT,  the  tangent  at  P,  makes  an  angle  yjr  with  OX. 
Then  to  say  that  PT  is  the  limiting  position  of  PQ  is  equivalent 
to  saying  that  the  limit  of  the  angle  QPR  is  yp;  when  Q  approaches 


198 


DERIVATIVES   AND   INTEGRALS 


[VI 


P  along  the  curve  from  either  side.     We  have  now  to  distinguish 
two  cases,  a  general  case  and  an  exceptional  one. 


N  M 

Fig.  36. 

The  general  case  is  that  in  which  yjr  is  not  equal  to  |tt,  so  that 
PT  is  not  parallel  to  OY.     In  this  case  RPQ  tends  to  the  limit 

yjr,  and 

RQ/ PR  =  tan  RPQ 

tends  to  the  limit  tan  yfr.     Now 

RQ/PR  =  (NQ  -  MP)jMN  =  {$  (x  +  h)  -  <£  (x)}fh ; 

..      <\> (x  +  h) -  <f) O)      ,        ,  ,_. 

and  so  lim -^ ^ — 7-  \  /  _  tan  ^    ^^ 

a  ^-  0  'i 

The  reader  should  be  careful  to  note  that  in  all  these  equa- 
tions all  lengths  are  regarded  as  affected  with  the  proper  sign, 
so  that  (e.g.)  RQ  is  negative  in  the  figure  when  Q  lies  to  the  left 
of  P ;  and  that  the  convergence  to  the  limit  is  unaffected  by  the 
sign  of  h. 

Thus  the  assumption  that  the  curve  which  is  the  graph  of 
$  (x)  has  a  tangent  at  P,  which  is  not  perpendicular  to  the  axis  of 
x,  implies  that  cj>  (x)  has,  for  the  particular  value  of  x  corresponding 
to  P,  the  property  that  [(f)  (x  +  h)  —  <£  (x))jh  tends  to  a  limit  when 
h  tends  to  zero. 

This  of  course  implies  that  both  of 

{<(>  (x  +  h)-4>  {x))lh,     {0  (.r -  h)  -  0  (tf)}/(  -  h) 
tend  to  limits  when  h-*-d  by  positive  values  only,  and  that  the  two  limits 
are  equal.     If  these  limits  exist  but  are  not  equal,  then  the  curve  y  =  <$>  (x) 
has  an  angle  at  the  particular  point  considered,  as  in  Fig.  37. 

Now  let  us  suppose  that  the  curve  has  (like  the  circle  or 
ellipse)  a  tangent  at  every  point  of  its  length,  or  at  any  rate  every 


110,  111]  DERIVATIVES   AND   INTEGRALS  199 

portion  of  its  length  which  corresponds  to  a  certain  range  of 
variation  of  x.  Further  let  us  suppose  this  tangent  never  per- 
pendicular to  the  axis  of  x  :  in  the  case  of  a  circle  this  would  of 
course  restrict  us  to  considering  an  arc  less  than  a  semicircle. 
Then  an  equation  such  as  (1)  holds  for  all  values  of  x  which  fall 
inside  this  range.  To  each  such  value  of  x  corresponds  a  value  of 
tan  -v/r :  tan  ty  is  a  function  of  x,  which  is  defined  for  all  values  of 
x  in  the  range  of  values  under  consideration,  and  which  may  be 
calculated  or  derived  from  the  original  function  (f>(x).  We  shall 
call  this  function  the  derivative  or  derived  function  of  <£  (x),  and 
we  shall  denote  it  by 

<f>'(x). 

Another  name  for  the  derived  function  of  $  (x)  is  the  differ- 
ential coefficient  of  <£  (x) ;  and  the  operation  of  calculating 
<£'  (x)  from  cf>  (x)  is  generally  known  as  differentiation.  This 
terminology  ■  is  firmly  established  for  historical  reasons :  see 
§  115. 

Before  we  proceed  to  consider  the  special  case  mentioned 
above,  in  which  yjr  —  \tt,  we  shall  illustrate  our  definition  by  some 
general  remarks  and  particular  illustrations. 

111.  Some  general  remarks.  (1)  The  existence  of  a  derived 
function  <f>'  (x)  for  all  values  of  x  in  the  interval  a  ^  x  £b  implies 
that  <j6  (x)  is  continuous  at  every  point  of  this  interval.  For  it  is 
evident  that  {<£  (x  +  h)  —  <f>  (x)}/h  cannot  tend  to  a  limit  unless 
lim  </>  (x  +  h)  =  <$>(x),  and  it  is  this  which  is  the  property  denoted 
by  continuity. 

(2)  It  is  natural  to  ask  whether  the  converse  is  true,  i.e. 
whether  every  continuous  curve  has  a 
definite  tangent  at  every  point,  and 
every  function  a  differential  coefficient 
for  every  value  of  x  for  which  it  is 
continuous.*  The  answer  is  obviously 
No :  it  is  sufficient  to  consider  the 
curve    formed    by   two    straight    lines  pig   37# 

meeting  to  form  an   angle   (Fig.  37). 

*  We  leave  out  of  account  the  exceptional  case  (which  we  have  still  to  examine) 
in  which  the  curve  is  supposed  to  have  a  tangent  perpendicular  to  OX:  apart  from 
this  possibility  the  two  forms  of  the  question  stated  above  are  ecpivalent. 


200  DERIVATIVES   AND   INTEGRALS  [VI 

The  reader  will  see  at  once  that  in  this  case  {<£  (x  +  li)  —  <$>  (x)}/h 
has  the  limit  tan/3  when  h^>-0  by  positive  values  and  the  limit 
tan  a  when  h-*~0  by  negative  values. 

This  is  of  course  a  case  in  which  a  curve  might  reasonably  be  said  to  have 
two  directions  at  a  point.  But  the  following  example,  although  a  little  more 
difficult,  shows  conclusively  that  there  are  cases  in  which  a  continuous  curve 
cannot  be  said  to  have  either  one  direction  or  several  directions  at  one  of  its 
points.  Draw  the  graph  (Fig.  14,  p.  53)  of  the  function  a?sin(l/#).  The 
function  is  not  denned  for  x=0,  and  so  is  discontinuous  for  x=0.  On 
the  other  hand  the  function  defined  by  the  equations 

<£(#)=a?sin(l/#)    U'4=0),  0(.r)  =  O    (x=0) 

is  continuous  for  x=0  (Exs.  xxxvu.  14,  15),  and  the  graph  of  this  function 
is  a  continuous  curve. 

But  <f)(%)  has  no  derivative  for  x  =  0.  For  $'  (0)  would  be,  by  definition, 
lim  {(p  (h)  -  <f>  (0)}/A  or  lim  sin  (l//t) ;  and  no  such  limit  exists. 

It  has  even  been  shown  that  a  function  of  x  may  be  continuous  and  yet 
have  no  derivative  for  any  value  of  x,  but  the  proof  of  this  is  much  more 
difficult.  The  reader  who  is  interested  in  the  question  may  be  referred  to 
Bromwich's  Infinite  Series,  pp.  490-1,  or  Hobson's  Theory  of  Functions 
of  a  Real   Variable,  pp.  620-5. 

(3)  The  notion  of  a  derivative  or  differential  coefficient  was 
suggested  to  us  by  geometrical  considerations.  But  there  is 
nothing  geometrical  in  the  notion  itself.  The  derivative  </>'  (x)  of 
a  function  (j>  (x)  may  be  defined,  without  any  reference  to  any  kind 
of  geometrical  representation  of  (/>  (x),  by  the  equation 


^(w)silim^±E^M.t 


h 

and  <£  (x)  has  or  has  not  a  derivative,  for  any  particular  value  of  x, 
according  as  this  limit  does  or  does  not  exist.  The  geometry  of 
curves  is  merely  one  of  many  departments  of  mathematics  in  which 
the  idea  of  a  derivative  finds  an  application. 

Another  important  application  is  in  dynamics.  Suppose  that  a  particle  is 
moving  in  a  straight  line  in  such  a  way  that  at  time  t  its  distance  from  a  fixed 
point  on  the  line  is  s  =  0  (t).     Then  the  'velocity  of  the  particle  at  time  t'  is 

by  definition  the  limit  of 

<t>(t  +  h)-(f)(t) 
h 

as  h-*~0.  The  notion  of  '  velocity  '  is  in  fact  merely  a  special  case  of  that  of 
the  derivative  of  a  function. 


Ill,  112]  DERIVATIVES   AND    INTEGRALS  201 

Examples  XXXIX.  1.  If  <£  (a>)  is  a  constant  then  $'  (a?)=0.  Interpret 
this  result  geometrically. 

2.  If  <f)(x)  =  ax+b  then  <f)'(x)  =  a.  Prove  this  (i)  from  the  formal  de- 
finition and  (ii)  by  geometrical  considerations. 

3.  If  $  (x)=xm,  where  m  is  a  positive  integer,  then  <£'  (x)=mxm~\ 
[For  <£'  (.i-)  =  lim  (*  +  hT-xm 

=  lim  lm.vm-1+171  ^m~   \vm-2h  +  ...  +  hm-1\  . 

The  reader  should  observe  that  this  method  cannot  be  applied  to  apto, 
where  pjq  is  a  rational  fraction,  as  we  have  no  means  of  expressing  {x  +  ICfii 
as  a  finite  series  of  powers  of  h.  We  shall  show  later  on  (§  118)  that  the  result 
of  this  example  holds  for  all  rational  values  of  m.  Meanwhile  the  reader 
will  find  it  instructive  to  determine  <f>'  (x)  when  m  has  some  special  fractional 
value  {e.g.  £),  by  means  of  some  special  device.] 

4.  If  (f>  (x)  =  sin  x,  then  <£'  (x)  =  cos  x  ;  and  if  <j>  (x)  =  cos  x,  then 
<j)'  (x)  =  -  sin  x. 

[For  example,  if  0  (o?)  =  sin  x,  we  have 

{0  (x+h)  -  (f>  (x)}/h  =  {2  sin  \h  cos  (x  +  lh)}/h, 
the  limit  of  which,  when  A-»-0,  is  cos  x,  since  lim  cos  (#  +  M)  =  cos  x  (the  cosine 
being  a  continuous  function)  and  lim  {(sin  |A)/M}  =  1  (Ex.  xxxvi.  13).] 

5.  Equations  of  the  tangent  and  normal  to  a  curve  y=(f>(x).  The 
tangent  to  the  curve  at  the  point  (.r0,  y0)  is  the  line  through  (.r0,  y0)  which 
makes  with  OX  an  angle  ^,  where  tan  ^  =  $'  (x0).     Its  equation  is  therefore 

y-#o=(#-#o)0'(#o) ; 

and  the  equation  of  the  normal  (the  perpendicular  to  the  tangent  at  the 
point  of  contact)  is 

We  have  assumed  that  the  tangent  is  not  parallel  to  the  axis  of  y.  In 
this  special  case  it  is  obvious  that  the  tangent  and  normal  are  x=x0  and 
y=y0  respectively. 

6.  Write  down  the  equations  of  the  tangent  and  normal  at  any  point  of 
the  parabola  x2  =  4ay.  Show  that  if  x0  =  2a/m,  y0  =  a/m2,  then  the  tangent 
at  (xQ,  y0)  is  x=my  +  (a/m). 

112.  We  have  seen  that  if  </>  (x)  is  not  continuous  for  a  value 
of  x  then  it  cannot  possibly  have  a  derivative  for  that  value  of  x. 
Thus  such  functions  as  lfx  or  sin  (l/x),  which  are  not  denaed  for 
x  =  0,  and  so  necessarily  discontinuous  for  x  =  0,  cannot  have 
derivatives  for  x  =  0.  Or  again  the  function  [x],  which  is  discon- 
tinuous for  every  integral  value  of  x,  has  no  derivative  for  any 
such  value  of  x. 


202 


DERIVATIVES   AND   INTEGRALS 


[VI 


Example.  Since  [.r]  is  constant  between  every  two  integral  values  of  x, 
its  derivative,  whenever  it  exists,  has  the  value  zero.  Thus  the  deriva- 
tive of  fa?],  which  we  may  represent  by  [x]',  is  a  function  equal  to  zero  for 
all  values  of  x  save  integral  values  and  undefined  for  integral  values.     It 

is  interesting  to   note  that  the  function   1  - ■  has  exactly  the  same 

properties. 

We  saw  also  in  Ex.  xxxvil.  7  that  the  types  of  discontinuity 
which  occur  most  commonly,  when  we  are  dealing  with  the  very 
simplest  and  most  obvious  kinds  of  functions,  such  as  polynomials 
or  rational  or  trigonometrical  functions,  are  associated  with  a 
relation  of  the  type 

</>  (x)  -»  +  oo 

or  <j>  (x)  -*-  —  oo  .     In  all  these  cases,  as  in  such  cases  as  those  con- 
sidered above,  there  is  no  derivative  for  certain  special  values  of  x. 


Q 


Q.         Q 


R        R 


,Q         Q 


(*) 


^R 


(&) 


Fig.  38. 


In  fact,  as  was  pointed  out  in  §  111,  (1),  all  discontinuities  of<f)(x)  are 
also  discontinuities  of  cf)' (x).  But  the  converse  is  not  true,  as  we 
may  easily  see  if  we  return  to  the  geometrical  point  of  view  of  §  110 
and  consider  the  special  case,  hitherto  left  aside,  in  which  the  graph 
of  <f)  (x)  has  a  tangent  parallel  to  OY.  This  case  may  be  subdivided 
into  a  number  of  cases,  of  which  the  most  typical  are  shown  in 
Fig.  3<8.  In  cases  (c)  and  (d)  the  function  is  two  valued  on  one  side 
of  P  and  not  denned  on  the  other.  In  such  cases  we  may  consider 
the  two  sets  of  values  of  $  (x),  which  occur  on  one  side  of  P  or  the 
other,  as  defining  distinct  functions  <$>i{x)  and  <$>2(x),  the  upper 
part  of  the  curve  corresponding  to  fa  (x). 


112,  113]  DERIVATIVES   AND   INTEGRALS  203 

The  reader  will  easily  convince  himself  that  in  (a) 
{$  (x  +  h)-cf>  (x)}/h  —  +  cc  , 
as  h-*-0,  and  in  (6) 

{<f>(x  +  h)  —  <j)(x)}(h-*-—cG  ; 
while  in  (c) 

{fa  (x  +  h)  -  fa  (x)}/h-^  +  co  ,     {<f>2  (x  +  h)  -  fa  (x)}/h -■ —  cd  , 

and  in  (d) 

{(f),  (x  +  h)  -  fa  (x)}/h-* —  oo  ,     {<f>2  (x  +  h)  -  fa  (x)}/k-*~+  oo  , 

though  of  course  in  (c)  only  positive  and  in  (d)  only  negative 
values  of  h  can  be  considered,  a  fact  which  by  itself  would  preclude 
the  existence  of  a  derivative. 

We  can  obtain  examples  of  these  four  cases  by  considering  the 
functions  defined  by  the  equations 

(a)    y3  =  x,       (b)    y*=-x,      (c)     y"  =  x,       (d)    y"  =  -x, 

the  special  value  of  x  under  consideration  being  x  =  0. 

113.  Some  general  rules  for  differentiation.  Through- 
out the  theorems  which  follow  we  assume  that  the  functions 
f{x)  and  F(x)  have  derivatives /'(#)  and  F'{x)  for  the  values  of 
x  considered. 

(1)  If  cf>  (x)  =f(x)  +  F  (x),  then  0  (x)  has  a  derivative 

fa(x)=f(x)  +  F'(x). 

(2)  If  (f>  (x)  =  kf  (x),  where  h  is  a  constant,  then  <£  (x)  has  a 
derivative 

fa(x)  =  lcf(x). 

We  leave  it  as  an  exercise  to  the  reader  to  deduce  these  results 
from  the  general  theorems  stated  in  Ex.  xxxv.  1. 

(3)  If  <f>  (x)  =f(x)  F(x),  then  cf>  (x)  has  a  derivative 

4,\x)=f(x)F\x)+f(x)F(x). 

For         fW-lim/(«  +  *)J'<«+^-/WJ'<*) 

.to  |/(, + ;,)  *■(«+» -*■(«>  +  >  (./»+*)-/W[ 


204  DERIVATIVES   AND   INTEGRALS  [VI 

(4)  If  0  (x)  =  >--t  ,  then  <f>  (x)  has  a  derivative 

J\x) 

In  this  theorem  we  of  course  suppose  that  f(x)  is  not  equal  to 
zero  for  the  particular  value  of  x  under  consideration.     Then 

*(*>-1,m  A  !/(.  +  *)/(«)  J-    {/(5)}«- 

(5)  If  (ji(x)=  ^rr-{  ,  then  <$>  (x)  has  a  derivative 

Jf  (x) 

.,,  ,     f(x)F(x)-f(x)F'(x) 
(PW_"  {F(x)\* 

This  follows  at  once  from  (3)  and  (4). 

(6)  If  4>(x)  =  F  {/(&")}>  then  4>  (x)  h(ts  a  derivative 

<j>'(x)=F'{f(x)}f'(x). 

For  let  /0)=?/,    f(x  +  h)=y  +  k. 

Then  k-*0  as  h-*-0,  and  kjh-^f(x).     And 

=  F'(y)f(x). 

This  theorem  includes  (2)  and  (4)  as  special  cases,  as  we  see  on 
taking  F(x)  =  kx  or  F(x)  =  ljx.  Another  interesting  special  case 
is  that  in  which  f(x)  =  ax  +  b  :  the  theorem  then  shows  that  the 
derivative  of  F  (ax  +  b)  is  aF'  (ax  +  b). 

Our  last  theorem  requires  a  few  words  of  preliminary  explana- 
tion. Suppose  that  x  =  yjr  (y),  where  yjr  (y)  is  continuous  and 
steadily  increasing  (or  decreasing),  in  the  stricter  sense  of  §  95,  in 
a  certain  interval  of  values  of  y.  Then  we  may  write  y  =  </>  (x), 
where  <f>  is  the  '  inverse '  function  (§  109)  of  ifr. 

(7)  If  y  =  cf>(x),  ivhere  (j>  is  the  inverse  f  miction  of  yjr,  so  that 
x  =  yfr  (y),  and  -ty  (y)  has  a  derivative  i/r'  (y)  which  is  not  equal  to 
zero,  then  <£  (x)  has  a  derivative 

d,'(a-)  =  _J_. 


113-115]  DERIVATIVES   AND   INTEGRALS  205 

For  if  <£  (as  +  hi)  =  y  +  k,  then  k  -*■  0  as  /t-*»0,  and 

#  w  =  Hm  »(«  +  *)-»(«)  =  lim      (y+k)-y     =  _1_ 

A^o       (x  +  h)-x  k+0  ^r  (y  +  k)  -  f  (y)      ^  (y) 

The  last  function  may  now  be  expressed  in  terms  of  x  by  means 
of  the  relation  y  =  </>  (x),  so  that  <$>{x)  is  the  reciprocal  of  yjr'{<j>  (x)}. 
This  theorem  enables  us  to  differentiate  any  function  if  we  know 
the  derivative  of  the  inverse  function. 

114.  Derivatives  of  complex  functions.  So  far  we  have 
supposed  that  y  =  <f>  (x)  is  a  purely  real  function  of  x.  If  y  is  a 
complex  function  <f>  (x)  +  i-^r  (x),  then  we  define  the  derivative  of  y 
as  being  <f>' (x)  +  iyfr'  (x).  The  reader  will  have  no  difficulty  in 
seeing  that  Theorems  (1) — (5)  above  retain  their  validity  when 
cf>(x)  is  complex.  Theorems  (6)  and  (7)  have  also  analogues  for 
complex  functions,  but  these  depend  upon  the  general  notion  of 
a  '  function  of  a  complex  variable ',  a  notion  which  we  have  en- 
countered at  present  only  in  a  few  particular  cases. 

115.  The  notation  of  the  differential  calculus.  We  have 
already  explained  that  what  we  call  a  derivative  is  often  called  a 
differential  coefficient.  Not  only  a  different  name  but  a  different 
notation  is  often  used ;  the  derivative  of  the  function  y  =  cf)  (x) 
is  often  denoted  by  one  or  other  of  the  expressions 

«*  & 

Of  these  the  last  is  the  most  usual  and  convenient :  the  reader 
must  however  be  careful  to  remember  that  dyjdx  does  not  mean 
'  a  certain  number  dy  divided  by  another  number  dx ' :  it  means 
'  the  result  of  a  certain  operation  Dx  or  d/dx  applied  to  y  =  <£  (x) ', 
the  operation  being  that  of  forming  the  quotient  {(f>  (x  +  h)  —  <f>  (x)}/h 
and  making  /i-*0. 

Of  course  a  notation  at  first  sight  so  peculiar  would  not  have  been 
adopted  without  some  reason,  and  the  reason  was  as  follows.  The  denomi- 
nator h  of  the  fraction  {<fi  (x  +  h)  -  (p  (x)}/h  is  the  difference  of  the  values  x  +  h, 
x  of  the  independent  variable  x  ;  similarly  the  numerator  is  the  difference  of 
the  corresponding  values  <p(x  +  h),  0  (x)  of  the  dependent  variable  y.  These 
differences  may  be  called  the  increments  of  x  and  y  respectively,  and  denoted 
by  hx  and  by.  Then  the  fraction  is  8y/8x,  and  it  is  for  many  purj^oses 
convenient  to  denote  the  limit  of  the  fraction,  which  is  the  same  thing  as 


206  DERIVATIVES   AND   INTEGRALS  [VI 

cf)'  (x),  by  dy/dx.  But  this  notation  must  for  the  present  be  regarded  as 
purely  symbolical.  The  dy  and  dx  which  occur  in  it  cannot  be  separated, 
and  standing  by  themselves  they  would  mean  nothing  :  in  particular  dy  and 
dx  do  not  mean  lim  Sy  and  lim  8x,  these  limits  being  simply  equal  to  zero. 
The  reader  will  have  to  become  familiar  with  this  notation,  but  so  long  as  it 
puzzles  him  he  will  be  wise  to  avoid  it  by  writing  the  differential  coefficient  in 
the  form  Dxy,  or  using  the  notation  (f>  (x),  (f>'(x),  as  we  have  done  in  the 
preceding  sections  of  this  chapter. 

In  Ch.  VII,  however,  we  shall  show  how  it  is  possible  to  define  the  symbols 
dx  and  dy  in  such  a  way  that  they  have  an  independent  meaning  and  that 
the  derivative  dy/dx  is  actually  their  quotient. 

The  theorems  of  §  113  may  of  course  at  once  be  translated  into 
this  notation.     They  may  be  stated  as  follows : 

(i)   ify  =  y,+y„  then  £-§£+{£; 

(2)  if  y=  ley,,  then  ^  =  *^i 

,«v  t  dv         dy.,         dy, 

(3)  »/y  =  y,y„  then        £  =  y,±  +  y>±i 

(5)  ify^then  %-{»%-&)/* 

(6)  if  y  is  a  function  of  x,  and  z  a  function  of  y,  then 

dz  _dz  dy 
dx     dy  dx ' 

m  £-»/©■ 

Examples  XL.     1.    If  y =3/12/23/3  then 


dy  dyx  dy2  ,  dy, 

dx=^3dx+^dx+^di 


and  if  3/ =3/13/2 --^n  then 

dy       »  dyr 

In    particular,   if-?/  =  2",    then   d,y/dx=nzn~1(dz/dx)  :    and    if    y=xn,   then 
dyjdx=nxn~l,  as  was  proved  otherwise  in  Ex.  xxxix.  3. 


115,  116]  DERIVATIVES   AND    INTEGRALS  207 

y  dx     yx  dx     y2  dx  yn  dx  ' 

In  particular,  if  y  =  zn,  then  -  -~  =  -  —  . 
1  '     3       '  y  dx     z  dx 

116.  Standard  forms.  We  shall  now  investigate  more 
systematically  the  forms  of  the  derivatives  of  a  few  of  the 
simplest  types  of  functions. 

A.     Polynomials.     If  0  (x)  =  a0xn  +  aixn~l  +  . . .  +  an,  then 

<f>'  (x)  =  nctoX11-1  +  (n  -  1)  a.x71-2  +  ...+  an^. 

It  is  sometimes  more  convenient  to  use  for  the  standard  form  of  a 
polynomial  of  degree  n  in  x  what  is  known  as  the  binomial  form, 
viz. 

ciox11  +  (Jj  alaP~1+  (fj  a,xn->+  ...  +  an. 
In  this  case 

The  binomial  form  of  <f>(x)  is  often  written  symbolically  as 

(o0,  Qii, ...,  an^x,  i.)  j 

and  then  (f>'(x)  =  n(a0,  a1}  ...,  «„_!"$#,  l)n_1. 

We  shall  see  later  that  <f>  (x)  can  always  be  expressed  as  the 
product  of  n  factors  in  the  form 

<£  (a?)  =a0(x-  ax)  (x  -  a2)  . ..  (a?  -  orn), 

where  the  a's  are  real  or  complex  numbers.     Then 

</>'  (a;)  =  a0  2  (a?  -  or2)  (x  -  as)  . . .  (x  -  an), 

the  notation  implying  that  we  form  all  possible  products  of  n  —  1 
factors,  and  add  them  all  together.  This  form  of  the  result  holds 
even  if  several  of  the  numbers  a  are  equal ;  but  of  course  then 
some  of  the  terms  on  the  right-hand  side  are  repeated.  The 
reader  will  easily  verify  that  if 

<£  (x)  =  a0  (x -  cO'"'  (x -  ar3)m'  ...(x-  at)m", 
then         0' (»=  aoXtih  (x  -  a,)**"1  (x  ~  a*)"h  ...(*-  «,)'""• 


208  DERIVATIVES    AND    INTEGRALS  [VI 

Examples  XLI.  1.  Show  that  if  0  (x)  is  a  polynomial  then  <f>'  (x)  is 
the  coefficient  of  h  in  the  expansion  of  $  (x  +  h)  in  powers  of  h. 

2.  If  0  (x)  is  divisible  by  (x  —  a)2,  then  <£'  (a?)  is  divisible  by  x  —  a  :  and 
generally,  if  cp  (x)  is  divisible  by  (x  —  a)m,  then  <j>' (x)  is  divisible  by  (x  —  a)m~K 

3.  Conversely,  if  0  (x)  and  0'  (#)  are  both  divisible  by  x  —  a,  then  $  (#)  is 
divisible  by  (%-  a)2 ;  and  if  0  (a-)  is  divisible  by  x  -  a  and  cf>'  (x)  by  (#-  a)m~1> 
then  0  (a-)  is  divisible  by  (x  -  a)"1. 

4.  Show  how  to  determine  as  completely  as  possible  the  multiple  roots 
of  P(x)  =  0,  where  P  (x)  is  a  polynomial,  with  their  degrees  of  multiplicity, 
by  means  of  the  elementary  algebraical  operations. 

[If  Hx  is  the  highest  common  factor  of  P  and  P',  H2  the  highest  common 
factor  of  Hx  and  P",  H3  that  of  H2  and  P"',  and  so  on,  then  the  roots  of 
ff1H3/E22=0  are  the  double  roots  of  P=0,  the  roots  of  H2IIi/H32  =  0  the  treble 
roots,  and  so  on.  But  it  may  not  be  possible  to  complete  the  solution  of 
HiHslHf^Q,  H2HiIH32  =  0,  ....  Thus  if  P(x)  =  (x-l)3(x*-x-7)2  then 
Hi  R3/ H22  =  x6  —  x -7  and  If2FIijII32  =  x—l  ;  and  we  cannot  solve  the  first 
equation.] 

5.  Find  all  the  roots,  with  their  degrees  of  multiplicity,  of 

xi  +  3x3  -  3x2  - 1  Ix  -  6  =  0,     afi  +  2a5  -  8x*  -  1 4x*  + 1  lx2  +  28j?  + 1 2  =  0. 

6.  If  ax2  +  2bx+c  has  a  double  root,  i.e.  is  of  the  form  a(x-a)2,  then 
2  (ax  +  b)  must  be  divisible  by  x  -  a,  so  that  n=  -  fe/a.  This  value  of  #  must 
satisfy  ax2  +  2bx  +  c  =  0.  Verify  that  the  condition  thus  arrived  at  is 
ac  —  b2—0. 

7.  The  equation  ll(x-  a)  +  lj(x-b)  +  \/(x-c)  =  0  can  have  a  pair  of 
equal  roots  only  if  a  =  b  =  c.  (Math.    Trip.    1905.) 

8.  Show  that  ax3  +  Sbx2  +  3cx  +  d=0 

has  a  double  root  if  £2  +  4#3=0,  where  II=ac-b2,  O  =  a2d - Sabc  +  2b\ 

[Put  ax+b=y,  when  the  equation  reduces  to  y3  +  3ffy  +  G=Q.  This 
must  have  a  root  in  common  with  y2  +  II=0.~\ 

9.  The  reader  may  verify  that  if  a,  /3,  y,  8  are  the  roots  of 

ax*  +  Abx3  +  Qcx2  +  4dx + e = 0, 
then  the  equation  whose  roots  are 

iW(°-i3)(y-8)-(y~a)(/3-8)}, 

and  two  similar  expressions  formed  by  permuting  a,  $,  y  cyclically,  is 

453-^-^=0, 

where  g2  =  ae  -  Abd  +  3c2,     g3  =  ace  +  2bcd  -  ad2  -  eb2  -  c3. 

It  is  clear  that  if  two  of  a,  /3,  y,  8  are  equal  then  two  of  the  roots  of  this  cubic 
will  be  equal.     Using  the  result  of  Ex.  8  we  deduce  that  g23-27g32=0. 


116,  117]  DERIVATIVES   AND    INTEGRALS  209 

10.  Rolle's  Theorem  for  polynomials.  If  <f>  (x)  is  any  polynomial, 
then  between  any  pair  of  roots  of(f>(x)=0  lies  a  root  o/ <£'(#)  =  0. 

A  general  proof  of  this  theorem,  applying  not  only  to  polynomials  but  to 
other  classes  of  functions,  will  be  given  later.  The  following  is  an  algebraical 
proof  valid  for  polynomials  only.  We  suppose  that  a,  /3  are  two  successive 
roots,  repeated  respectively  m  and  n  times,  so  that 

0  (x)  =  (x  -  a)m  (x  -  /3)»  8  (x), 

where  6  (x)  is  a  polynomial  which  has  the  same  sign,  say  the  positive  sign,  for 
a?gx?gft.     Then 

<t>'  (x)  =  (x  -  a)m  (x  -  /3)n  &  (x)  +  {m  (x  -  a)m  -  >  (x  -  j3)» + n  (x  -  a)m  (x  -  /3)»  " »}  6  (x) 
=  (x  -  a^-^x-py-^ix-  a)(x-p)ff \x)  +  {7n(x-p)  +  n(x-a)}d(x)] 
=  (x-a)m-1(x-^t)n-1F  (x), 

say.  Now  F(a)  =  vi(a-&)  6  (a)  and  F  (/3)  =  n  (/3 -  a)  0(/3),  which  have  opposite 
signs.  Hence  F(x),  and  so  <p'  (x),  vanishes  for  some  value  of  x  between 
a  and  /3 

117.     B.     Rational  Functions.     If 

where  P  and  Q  are  polynomials,  it  follows  at  once  from  §  113,  (5)  that 
_  P'(x)Q(x)-P(x)Q'{x) 

and  this  formula  enables  us  to  write  down  the  derivative  of  any- 
rational  function.  The  form  in  which  we  obtain  it,  however,  may  or 
may  not  be  the  simplest  possible.  It  will  be  the  simplest  possible  if 
Q  (x)  and  Q'  (x)  have  no  common  factor,  i.e.  if  Q  (x)  has  no  repeated 
factor.  But  if  Q(x)  has  a  repeated  factor  then  the  expression 
which  we  obtain  for  R'  (x)  will  be  capable  of  further  reduction. 

It  is  very  often  convenient,  in  differentiating  a  rational 
function,  to  employ  the  method  of  partial  fractions.  We  shall 
suppose  that  Q(x),  as  in  §  116,  is  expressed  in  the  form 

a0 (x  -  aO'"1  (x  -  a2)m*  ...(x-  <xv)mv. 
Then  it  is  proved  in  treatises  on  Algebra*   that  R(x)  can  be 
expressed  in  the  form 

nl      \     ,        -^1,1        .           -"-1,2           .                   .           -"■] ,  TO, 
(x)  +  - — '■ f-  -. —  +  ...  4-  , —~— 

x  —  ax      (x  —  aj'  {x  —  ax)"h 

x  —  a2     (x  —  a2)-  (x  —  a2)'"2 

*  See,  e.g.,  Chrystal's  Algebra,  vol.  i,  pp.  151  et  seq. 
n.  H 


210  DERIVATIVES   AND    INTEGRALS  [VI 

where  II  (x)  is  a  polynomial;  i.e.  as  the  sum  of  a  polynomial  and 
the  sum  of  a  number  of  terms  of  the  type 

A 

where  a  is  a  root  of  Q  (x)  =  0.  We  know  already  how  to  find  the 
derivative  of  the  polynomial:  and  it  follows  at  once  from  Theorem  (4) 
of  §  113,  or,  if  a  is  complex,  from  its  extension  indicated  in  §  114, 
that  the  derivative  of  the  rational  function  last  written  is 

pA  (x  -  ay-1  _  pA 

' ~Xx^ajw~ ~     («-a)*+1* 
We  are  now  able  to  write  down  the  derivative  of  the  general 
rational  function  R  (x),  in  the  form 

TT'  (    \  —  '*  "-"1,2  Ao-i ^A2>2     _ 

W     (a; -a,)2     (a-^)3      '"      (x-a2f     (x-a2V       "" 
Incidentally  we  have  proved  that  the  derivative  of  xm  is  ma;1'1-1, 
for  all  integral  values  of  m  positive  or  negative. 

The  method  explained  in  this  section  is  particularly  useful 
when  we  have  to  differentiate  a  rational  function  several  times 
(see  Exs.  XLV). 


Examples  XLII.     1.     Prove  that 

d  (     x    \  —x1  d  f\—xl\ 

dx 


(    x   \        1-m?         d  (l-x^\  4x 


2.     Prove  that 

d   (  ax2  +  2bx + c  \      (ax  +  6)  (Bx  +  0)-  (bx  +  c)(Ax+B) 


\x  \Ax*+\ 


2Bx  +  CJ  (A  a? + 2Bx  +  Vf 

3.  If  Q  has  a  factor  (x-a)m,  then  the  denominator  of  R'  (when  R'  is 
reduced  to  its  lowest  terms)  is  divisible  by  (x  —  a)"'+1  but  by  no  higher  power 
of  x  —  a. 

4.  In  no  case  can  the  denominator  of  R'  have  a  simple  factor  x-a. 
Hence  no  rational  function  (such  as  1/x)  whose  denominator  contains  any 
simple  factor  can  be  the  derivative  of  another  rational  function. 

118.  C.  Algebraical  Functions.  The  results  of  the  pre- 
ceding sections,  together  with  Theorem  (6)  of  §  113,  enable  us  to 
obtain  the  derivative  of  any  explicit  algebraical  function  whatsoever. 

The  most  important  such  function  is  xm,  where  m  is  a  rational 
number.     We  have  seen  already  (§  117)  that  the  derivative  of  this 


117,  118]  DERIVATIVES    AND    INTEGRALS  211 

function  is  mas"*-1  when  m  is  an  integer  positive  or  negative ;  and 
we  shall  now  prove  that  this  result  is  true  for  all  rational  values 
of  m.  Suppose  that  y  =  xm  =  xi)l(i,  where  p  and  q  are  integers  and 
q  positive  ;  and  let  z  =  xx,i,  so  that  x  =  zi  and  y  =  zp.     Then 


2-sy©.-*"- 


mx' 
2 

This  result  may  also  be  deduced  as  a  corollary  from  Ex.  xxxvi. 
3.     For,  if  0  (x)  =  xm,  we  have 

..        Pn-Xm 

=  hm  ^-tt =  ma;"1-1. 

?— a;      ?  —  * 

It  is  clear  that  the  more  general  formula 

j-  (ax  +  b)m  =  ma  (ax  +  b)m~1 

holds  also  for  all  rational  values  of  m. 

The  differentiation  of  implicit  algebraical  functions  involves 
certain  theoretical  difficulties  to  which  we  shall  return  in  Ch.  VII. 
But  there  is  no  practical  difficulty  in  the  actual  calculation  of  the 
derivative  of  such  a  function :  the  method  to  be  adopted  will  be 
illustrated  sufficiently  by  an  example.  Suppose  that  y  is  given  by 
the  equation 

Xs  +  y?j  —  Saxy  =  0. 

Differentiating  with  respect  to  x  we  find 

x'+y"-d£-a{«+xd£t° 

,  dy         x1  —  ay 

and  so  -r-  =  —    „ . 

ax         y2  —  ax 

•Examples  XLIII      1.     Find  the  derivatives  of 

2.     Prove  that 

d_  (  :v_ } 


dx  p (a2  +  fyj      (a?  +  .x'2)3'2 '     dx  [J {a*  -  *2)j      {a-  -  a?2)3  a " 
3.     Find  the  differential  coefficient  of?/  when 

(i)     ax'2  +  2hxy + by-  +  %gx  +  %fy  +  c = 0,         (ii)     a'5 + #5  -  5a.<%2 = 0. 

14—2 


212  DERIVATIVES    AND    INTEGRALS  [VI 

119.  D.  Transcendental  Functions.  We  have  already 
proved  (Ex.  xxxix.  4)  that 

Dx  sin  x  =  cos  x,     Dx  cos  x  =  —  sin  x. 
By  means  of  Theorems  (4)  and  (5)  of  §  113,  the  reader  will 
easily  verify  that 

Dx  tan  x  =  sec2  x,  Dx  cot  x  —  —  cosec2  x, 

Dx  sec  x  =  tan  x  sec  x,     Dx  cosec  x  =  —  cot  x  cosec  x. 
And  by  means  of  Theorem  (7)  we  can  determine  the  derivatives 
of  the    ordinary  inverse  trigonometrical    functions.     The   reader 
should  verify  the  following  formulae : 

Dx  arc  sin  x  =  ±  1/V(1  —  oc"),  Dx  arc  cos  x  =  +  1/V(1  -  &)> 

Dx  arc  tana;  =  1/(1  +  a;2),  Dx  arc  cot  x  =  —  1/(1  +  x-), 

Dx  arc  sec  a;  =  ±  l/{x^/(x2  —  1)},     D^  arc  cosec  a;  =  +  l/{x\/(xi  —  1)}. 

In  the  case  of  the  inverse  sine  and  cosecant  the  ambiguous  sign 
is  the  same  as  that  of  cos  (arc  sin  a;),  in  the  case  of  the  inverse 
cosine  and  secant  the  same  as  that  of  sin  (arc  cos  a;). 

The  more  general  formulae 

Dx  arc  sin  (xja)  =  ±  l/\/(a2  —  x2),     Dx  arc  tan  (x/a)  =  a/(x2  +  a2), 

which  are  also  easily  derived  from  Theorem  (7)  of  §  113,  are  also 
of  considerable  importance.  In  the  first  of  them  the  ambiguous 
sign  is  the  same  as  that  of  a  cos  {arc  sin  (x/a)},  since 

a V{1  -  (a;2/a2)}  =  ±  V(a2  -  O 

according  as  a  is  positive  or  negative. 

Finally,  by  means  of  Theorem  (6)  of  §  113,  we  are  enabled  to 
differentiate  composite  functions  involving  symbols  both  of  alge- 
braical and  trigonometrical  functionality,  and  so  to  write  down 
the  derivative  of  any  such  function  as  occurs  in  the  following 
examples. 

Examples  XLIV.*     1.     Find  the  derivatives  of 

cosm#,     sin"1*,     cos*™,     sin  a;"*,     cos  (sin*),     sin(cos.r), 

It     n  n  .     19       •      9        \  COS  *  Sin  * 

x/(a2  cos-*  *  +  b2  si  n2  *),     -r— —  , 

v  (a  cos  x  +  b  sm  x) 

x  arc  sin  * + ,J(1  —  x2),     (1  +  *)  arc  tan  *Jx  —  stx. 

*  In  these  examples  m  is  a  rational  number  and  a,  b,  ...  ,  a,  /3,  ...  hr-ve  such 
values  that  the  functions  which  involve  them  are  real. 


119]  DERIVATIVES    AND    INTEGRALS  213 

2.  Verify  by  differentiation  that  arc  sin  x  +  arc  cos  x  is  constant  for  all 
values  of  x  between  0  and  1,  and  arc  tan  u;+arc  cot  x  for  all  positive  values 
of  x. 

3.  Find  the  derivatives  of 

arc  sin  N/(  1  -  x2),     arc  sin  {2x  %/( 1  -  #2)},     arc  tan  (  ■ ) . 

How  do  you  explain  the  simplicity  of  the  results  1 

4.  Differentiate 

1                ,          ax+b                   1  .       ax+b 

arc  tan  -T. j^r ,     — ;- s  arc  sin 


sj(ac  -  b2) "     —  J(ac  _  b2) '         V(  -  «)  V(&2  -  ™) ' 

5.     Show  that  each  of  the  functions 

has  the  derivative 


V{(«-*)(*-/3)}' 

6  Prove  that 

3S  r°  C°S  v(^)|  =  V  (cos  6  cos  3d)  * 

(J/aM.  Trip.  1904.) 

7  Show  that 


1_      _d_r  /(Q(ax2+c)} 

J{G{Ac-aa)}dx[_&rCGm\J  \c(Ax2+C)j 


(Ax2+C)J(ax2+c) 


8.     Each  of  the  functions 


1  (a 

arc  cos 


— r- 1 1    -77-5 — ttn  arc  tan  \ x  /  ( T  )  tan  kv[ 

+  bcosx/'    sl(a2-b2)  \\    \a  +  bj        2   J 


s!{a2-b2) 
has  the  derivative  l/(a  +  bcosx). 

9.  If  X=  a  +  b  cos  x+c  sin  a?,  and 

1  aX-a2  +  b2  +  c2 

V-  v'(a2_fe2_c2)arc cos     JTVC^  +  c3)     ' 
then  dy/dx=l/X. 

10.  Prove  that  the  derivative  of  F[/{0(a?)}]is  ^/ {$<>)}]/' (00*0}  <$>'{x), 
and  extend  the  result  to  still  more  complicated  cases. 

11.  If  u  and  v  are  functions  of  x,  then 

Dx arc  tan  (%/■»)  =  {vDx u  —  uDxv)j(u2 -f v2). 

12.  The  derivative  of  y  =  (tan  x  +  sec  x)m  is  my  sec  x. 

13.  The  derivative  of  y = cos  &■ + 1  sin  #  is  ly. 

14.  Differentiate  x  cos  a?,  (sin#)/.r.  Show  that  the  values  of  #  for  which 
the  tangents  to  the  curves  y=x  cos  x,  y=(sin  #)/#  are  parallel  to  the  axis  of  x 
are  roots  of  cot#=.r,  tan#=.r  respectively. 


214  DERIVATIVES   AND   INTEGRALS  [VI 

15.  It  is  easy  to  see  (cf.  Ex.  xvn.  5)  that  the  equation  sin  x—ax,  where  a 
is  positive,  has  no  real  roots  except  #=0  if  a  >  1,  and  if  a<  1  a  finite  number  of 
roots  which  increases  as  a  diminishes.  Prove  that  the  values  of  a  for  which 
the  number  of  roots  changes  are  the  values  of  cos  £,  where  £  is  a  positive  root 
of  the  equation  tan  £  =  £.  [The  values  required  are  the  values  of  a  for  which 
y  =  ax  touches  y=sin  #.] 

16.  If  <£(.z)=#2sin(l/#)  when  j; +  0,  and  </>  (0)=0,  then 

<f>'  (x)  =  2x  sin  (1  \x)  -  cos  (1/x) 
when  #=#0,  and  <j>'(0)-0.     And  <£'  (x)  is  discontinuous  for  x  =  0  (cf.  §  111, 
(2))- 

17.  Find  the  equations  of  the  tangent  and  normal  at  the  point  (x0,  y0) 
of  the  circle  x2+y'*=a2. 

[Here  y=/J{a2-x2),  dyjdx=  —x/sf(a2-x2),  and  the  tangent  is 

y  -y*= (-p  -  #o)  { -  -^o/\/(«2  -  V)}> 

which  maybe  reduced  to  the  form  xx0+yy0  =  a2.     The  normal  is  xy0-yx0=0, 
which  of  course  passes  through  the  origin.] 

18.  Find  the  equations  of  the  tangent  and  normal  at  any  point  of  the 
ellipse  (x/a)2  +  (ylb)2=l  and  the  hyperbola  (xja)2  -  (y/6)2  =  1. 

19.  The  equations  of  the  tangent  and  normal  to  the  curve  x  =  <f>(t)T 
y  =  \jf  (t),  at  the  point  whose  parameter  is  t,  are 

^'to°  =  7(if '  {x " * (0} *' (0 + [y ~ * (0} *' w = °- 

120.  Repeated  differentiation.  We  may  form  a  new  function 
</>"(#)  from  <£'(#)  just  as  we  formed  <//  (#)  from  <j>  (x).  This 
function  is  called  the  second  derivative  or  second  differential 
coefficient  of  </>  (#).  The  second  derivative  of  y  =  <j)  (x)  may  also 
be  written  in  any  of  the  forms 


^  .         /  d  y        d2u 


In  exactly  the  same  way  we  may  define  the  nth  derivative  or 
nth  differential  coefficient  of  y  =  cf>  (x),  which  may  be  written  in  any 
of  the  forms 


r<-i  D*"*  (s)V  g- 


But  it  is  only  in  a  few  cases  that  it  is  easy  to  write  down  a 
general  formula  for  the  nth  differential  coefficient  of  a  given 
function.  Some  of  these  cases  will  be  found  in  the  examples 
which  follow. 


119,  120]  DERIVATIVES   AND   INTEGRALS  215 

Examples  XL V.     1.    If  4>(x)=xm  then 

(j)(n)  (x)  s=m  (»i  - 1) ...  (m— n + 1)  xm  -  ". 

This  result  enables  us  to  write  down  the  nth  derivative  of  any  polynomial. 

2.  If  $  (x)  =  (ax+b)m  then 

<£(>•)  (a?)  =  m  (m  - 1 ) . . .  (m  -  n  + 1 )  an  (ax  +  b)m  ~  ". 

In  these  two  examples  m  may  have  any  rational  value.     If  m  is  a  positive 
integer,  and  n>m,  then  <£(n)  (x)  =  0. 

3.  The  formula 

(AY      A      =(     1YP(P  +  V-(P+n-VA 
\dx)    (x-a)P     K       '  (x-a)P  +  n 

enables  us  to  write  down  the  nth  derivative  of  any  rational  function  expressed 
in  the  standard  form  as  a  sum  of  partial  fractions. 

4.  Prove  that  the  nth  derivative  of  1/(1  —x2)  is 

£(»  !){(1 -#)-»-i +(-l)»(l +#)-»- J}. 

5.  Leibniz'  Theorem.  If  y  is  a  product  uv,  and  we  can  form  the 
first  n  derivatives  of  u  and  v,  then  we  can  form  the  nth  derivative  of  y  by 
means  of  Leibni£  Theorem,  which  gives  the  rule 


(uv)n = unv + (J)  un  _  ,  v1  +  r  M 


v2  +  ...  +  [     )un_rvr  +  ...+uvn, 


where  suffixes  indicate  differentiations,  so  that  un,  for  example,  denotes  the 
?ith  derivative  of  u.     To  prove  the  theorem  we  observe  that 

(uv)i=ulv+uv1, 

(itv)2=zu2v  +  2ulv1  +  uv2, 
and  so  on.     It  is  obvious  that  by  repeating  this   process  we  arrive   at  a 
formula  of  the  type 

Let  us  assume  that  an  ,.=  (     )  for  r=l,  2, ...  n  —  1,  and  show  that  if  this 


is  so  then  a,i  +  i,  ,•  =  (  )    f°r  r— lj  2,  ...  n.     It  will   then   follow   by  the 

principle  of  mathematical  induction  that  aiu  r  =  (     )  for  all  values  of  n  and  r 
in  question. 

When  we  form  («v)n  +  1by  differentiating  (uv)n  it  is  clear  that  the  coefficient 
of  un  +  l_rvr  is 

«».,+«>,  ,-1=Q  +  (r!1)=('^1). 

This  establishes  the  theorem. 


216  DERIVATIVES   AND   INTEGRALS  [VI 

G.     The  nth  derivative  of  xmf{x)  is 

_J!Ll_^-»/(*)+.»r-4^fi>-*+1/(*) 

(m-n)  !  J  K  (m-n  +  1)  ! 

+  ~1.2     (m-»+2)!*  7   W  +  "" 

the  series  being  continued  for  n  + 1  terms  or  until  it  terminates. 

7.  Prove  that  ZV  cos  a-  =  cos  (a'+^wjt),  ZV  sin  o?=sin  (.r  +  inrr) 

8.  If  #  =  A  cos  ma-  +  B  sin  m#  then  ZVy + m2y = 0.     And  if 

y=A  cos  wi#  +  B  sin  »i.a? + Pn  (x), 
where  PM(#)  is  a  polynomial  of  degree  »,  then  Dxn  +  3  y  +  m2  Dxn  +  1  y  =  0. 

9.  If  x2Dx2y  +  xDxy+y  =  0  then 

.z2Z)sn  +  2y  +  (2;i  +  l):rZV  +  1y  +  (>j2  +  l)Z)/y=0. 
[Differentiate  ?i  times  by  Leibnitz'  Theorem.] 

10.  If  Un  denotes  the  «th  derivative  of  ( Lx  +  M  )/(x2  -  ~2Bx  +  C),  then 

[First  obtain  the  equation  when  n  =  0  ;   then  differentiate  n  times  by 
Leibnitz'  Theorem.] 

11.  The  nth  derivatives  of  aj{a2+x2)  and  xj(a2+x2).    Since 
1/1  1    \         x  1/1  1 


a2  +  x2     1i\x  —  cti     x  +  ai)''    a2  +  x2     "Z\x  —  ai     x  +  ai 
we  have 

/     a     \      (-l)»»lf         1 1         1 

*    V  +  a?V  2t        \(a?-at)"  +  1     (#  +  ai)n  +  ,j' 

and  a  similar  formula  for  Dx"  {x/(a2  +  x2)}.  If  p  =  J(x2  +  a2),  and  #  is  the 
numerically  smallest  angle  whot>e  cosine  and  sine  are  x/p  and  a/p,  then 
x  +  ai=p  Cis  0  and  x  -  ai=p  Cis  ( -  0),  and  so 

ZV  {a/(a2  +  x2)}  =  {( -  l)»w  !/2i}  p"""1  [Cis  {(»+ 1)0}-  Cis  {-  (n  +  Y)  6}] 

=  (-l)"n!  (x2  +  a2)-in  +  1)'2  sin  {(»  +  l)  arc  tan  (a/a?)}. 
Similarly 

ZV  W(«2  +  -r 2)}  =  ( -  1 )"  n !  (x2  +  a2)  ~  («  +  iV2  cos  {(»  + 1 )  arc  tan  (a/a?)}. 

12.  Prove  that 

Z»x"  {(cos  a?)/a?}  =  {Pw  cos  (x  +  \  mr)  +  Qn  sin  (x  +  ^nn)}  jxn  +  \ 
Dxn  {(sin  a?)/a?}  =  {Pn  sin  (a?  +  ^  mr )  -  Qn  cos  (x + \  mr)}/xn  + 1, 
where  Pn  and  Qn  are  polynomials  in  x  of  degree  n  and  ?i  -  1  respectively. 

13.  Establish  the  formulae 

^£_i  /7^\      ****-     <%  /(dyY     <Px__(<Py  dy_     fd2y\]    I (dyV> 
dy~  l\dx)'    dy2~     dx2/  \dx)  '    oy»~      \dV>  tfa;        \&)j/\di)  m 


z 

2\ 

22 

=  1 !  y*    y-i 

2\ 

22 

23 

y2 '  y-i    yA' 

22 

z3 

24 

{Math.  Trip.  1905.) 

y 

Z 

U 

,  dashes  denoting  differentiations  with 

y' 

2' 

u' 

y" 

z" 

u" 

120,    121]  DERIVATIVES.  AND    INTEGRALS  217 

14.     If yz=l  andyr=(l/r!)  Djy,  *,«(!/«  I)  Dx»z,  then 


15.     If  W(y,  z,  u)  = 


respect  to  x,  then  W (y,  z,  u)=yiW  i\,  -,   -). 

16.     If  ax2  +  2kxy  +  by2  +  2gx  +  2fy  +  c  =  0, 

then  dy\dx  =  -  (ax  +  hy  +g)/(hx  +  by  +/ ) 

and  d2y\dx2  =  (abc+2fgh  -  a/2  -  bg2  -  ch2)/(hx  +  by+f)\ 

121.  Some  general  theorems  concerning  derived  func- 
tions. In  all  that  follows  we  suppose  that  </>  (%)  is  a  function  of  x 
which  has  a  derivative  <£'  (x)  for  all  values  of  x  in  question.  This 
assumption  of  course  involves  the  continuity  of  <f)(x). 

The  meaning  of  the  sign  of  <£  (x).  Theorem  A.  If 
<£'  (x0)  >  0  then  (f>  (x)  <  (f>  (x0)  for  all  values  of  x  less  than  x0  but 
sufficiently  near  to  x0,  and  cf)(x)>  <f)(x0)  for  all  values  of  x  greater 
than  x0  but  sufficiently  near  to  oc0. 

For  {<£  (x0  +  h)  —  0  (x0)}/h  converges  to  a  positive  limit  <£'  (#•„)  as 
h-*-0.  This  can  only  be  the  case  if  </>  (x0  +  h)  —  (f>  (x0)  and  h  have 
the  same  sign  for  sufficiently  small  values  of  h,  and  this  is  precisely 
what  the  theorem  states.  Of  course  from  a  geometrical  point  of 
view  the  result  is  intuitive,  the  inequality  0'  (x)  >  0  expressing 
the  fact  that  the  tangent  to  the  curve  y  =  </>  (x)  makes  a  positive 
acute  angle  with  the  axis  of  x.  The  reader  should  formulate  for 
himself  the  corresponding  theorem  for  the  case  in  which  <f>'  (x)  <  0. 

An  immediate  deduction  from  Theorem  A  is  the  following 
important  theorem,  generally  known  as  Rolle's  Theorem.  In  view 
of  the  great  importance  of  this  theorem  it  may  be  well  to  repeat 
that  its  truth  depends  on  the  assumption  of  the  existence  of  the 
derivative  <£'  (x)  for  all  values  of  x  in  question. 

Theorem  B.  If  0  (a)  =  0  and  cf>  (b)  =  0,  then  there  must  be  at 
least  one  value  of  x  which  lies  between  a  and  b  and  for  which 
<f>'  (x)  =  0. 

There  are  two  possibilities :  the  first  is  that  <£  (x)  is  equal  to 


218  DERIVATIVES   AND   INTEGRALS  [VI 

zero  throughout  the  whole  interval  (a,  b).  In  this  case  <f>'  (x)  is 
also  equal  to  zero  throughout  the  interval.  If  on  the  other  hand 
cf>  (x)  is  not  always  equal  to  zero,  then  there  must  be  values  of 
x  for  which  <£  (x)  is  positive  or  negative.  Let  us  suppose,  for 
example,  that  <f>  (as)  is  sometimes  positive.  Then,  by  Theorem  2  of 
§  102,  there  is  a  value  £  of  x,  not  equal  to  a  or  b,  and  such  that  (/>  (£) 
is  at  least  as  great  as  the  value  of  <f>  (x)  at  any  other  point  in 
the  interval.  And  <f>'  (£)  must  be  equal  to  zero.  For  if  it  were 
positive  then  <f>(x)  would,  by  Theorem  A,  be  greater  than  <f)  (£)  for 
values  of  x  greater  than  £  but  sufficiently  near  to  f,  so  that  there 
would  certainly  be  values  of  <f>  (x)  greater  than  (f)  (f ).  Similarly  we 
can  show  that  <f>'  (£)  cannot  be  negative. 

Cor.  1.  If  <£  (a)  —  <fr(b)  =  k,  then  there  must  be  a  value  of  x 
between  a  and  b  such  that  <f>  (x)  =  0. 

We  have  only  to  put  <f>  (x)  —  k  =  -ty  (x)  and  apply  Theorem  B 
to  ty  (x). 

Cor.  2.  If  §'  (x)  >  0  for  all  values  of  x  in  a  certain  interval, 
then  <f>  (x)  is  an  increasing  function  of  x,  in  the  stricter  sense  o/§  95, 
throughout  that  interval. 

Let  #!  and  x2  be  two  values  of  x  in  the  interval  in  question, 
and  xx  <  x2.  We  have  to  show  that  <f>  (x^)  <  0  (#.,).  In  the  first 
place  </>  (a-j)  cannot  be  equal  to  (f>  (x2) ;  for,  if  this  were  so,  there 
would,  by  Theorem  B,  be  a  value  of  x  between  x1  and  x2  for  which 
<£'  (x)  =  0.  Nor  can  <f)  (x^  be  greater  than  <f>  (x2).  For,  since  0'  (x^ 
is  positive,  <f>  (x)  is,  by  Theorem  A,  greater  than  <f>  (a^)  when  x  is 
greater  than  xx  and  sufficiently  near  to  xx.  It  follows  that  there  is 
a  value  x3  of  x  between  x1  and  x2  such  that  <f)  (x3)  =  $  (x^) ;  and  so, 
by  Theorem  B,  that  there  is  a  value  of  as  between  xx  and  x3  for 
which  <f>'  (x)  =  0. 

Cor.  3.  The  conclusion  of  Cor.  2  still  holds  if  the  interval 
(a,  b)  considered  includes  a  finite  number  of  exceptional  values  of  x 
for  which  <f>'  (x)  does  not  exist,  or  is  not  positive,  provided  <£  (x)  is 
continuous  even  for  these  exceptional  values  of  x. 

It  is  plainly  sufficient  to  consider  the  case  in  which  there  is 
one  exceptional  value  of  x  only,  and  that  corresponding  to  an  end 
of  the  interval,  say  to  a.  If  a  <  x1  <  x2  <  b,  we  can  choose  a  +  e 
so  that  a  +  e  <  aslt  and  <£'  (x)  >  0  throughout  (a  +  e,  b),  so  that 
<f>  (#i)  <  <f>  (x2),    by   Cor.   2.      All    that  remains    is   to  prove   that 


121,  122] 


DERIVATIVES   AND    INTEGRALS 


219 


<f>(a)<  <f>  (xj).     Now  <f>  (xj)  decreases  steadily,  and  in  the  stricter 
sense,  as  xx  decreases  towards  a,  and  so 

<f>  (a)  =  <f>  (a  +  0)  =    Km    <£  (xx)  <  <£  (<i\). 


■rt+0 


Cor.  4.  If  (}>'  (x)  >  0  throughout  the  interval  (a,  b),  and  cf>  (a)  ^  0, 
then  <f>  (x)  is  positive  throughout  the  interval  (a,  b). 

The  reader  should  compare  the  second  of  these  corollaries  very  carefully 
with  Theorem  A.  If,  as  in  Theorem  A,  we  assume  only  that  <f>'(x)  is  positive 
at  a  single  point  x  =  x0,  then  we  can  prove  that  <j>  (xi)<(f>(x2)  when  X\  and  x2 
are  sufficiently  near  to  x0  and  x1<x0<x2.  For  0(.r1)<<^)(^o)  and  (p(x2)xp(x0), 
by  Theorem  A.  But  this  does  not  prove  that  there  is  any  interval  including 
x0  throughout  which  $  (x)  is  a  steadily  increasing  function,  for  the  assumption 
that  X\  and  x2  lie  on  opposite  sides  of  x0  is  essential  to  our  conclusion.  AVe 
shall  return  to  this  point,  and  illustrate  it  by  an  actual  example,  in  a  moment 
(§  124). 

122.  Maxima  and  Minima.  We  shall  say  that  the  value  </>(£) 
assumed  by  <f>  (x)  when  x  =  £  is  a  maximum  if  </>  (£)  is  greater  than 
any  other  value  assumed  by  <f>  (x)  in  the  immediate  neighbourhood 
of  x  =  £,  i.e.  if  we  can  find  an  interval  (£  —  e,  £  +  e)  of  values  of 
x  such  that  </>  (£)  >  <f>  (x)  when  £  —  e  <  x  <  £  and  when  £  <  x  <  l~+e: 
and  we  define  a  minimum  in  a  similar  manner.  Thus  in  the  figure 
the  points  A  correspond  to  maxima,  the  points  B  to  minima  of 


Bo 
Fig.  39. 

the  function  whose  graph  is  there  shown.  It  is  to  be  observed  that 
the  fact  that  A3  corresponds  to  a  maximum  and  Bx  to  a  minimum 
is  in  no  way  inconsistent  with  the  fact  that  the  value  of  the 
function  is  greater  at  Bl  than  at  A3. 

Theorem   C.     A    necessary   condition  for   a   maximum   or 
minimum  value  of  <f)(x)  at  x=i;  is  that  <£'(£)  =  0.* 

*  A  function  which  is  continuous  but  has  no  derivative  may  have  maxima  and 
minima.     We  are  of  course  assuming  the  existence  of  the  deiivative. 


220  DERIVATIVES    AND    INTEGRALS  [VI 

This  follows  at  once  from  Theorem  A.  That  the  condition  is  not 
sufficient  is  evident  from  a  glance  at  the  point  G  in  the  figure. 
Thus  if  y  =  x3  then  4>'(x)  =  3x2,  which  vanishes  when  x  =  0.  But 
x  =  0  does  not  give  either  a  maximum  or  a  minimum  of  xz,  as  is 
obvious  from  the  form  of  the  graph  of  x3  (Fig.  10,  p.  45). 

But  there  will  certainly  be  a  maximum  at  x  =  g  if  $'(£)  =  0, 
<f>'  (x)  >  0  for  all  values  of  x  less  than  but  near  to  f,  and  <$>'  (x)  <  0 
for  all  values  of  x  greater  than  but  near  to  f :  and  if  the  signs 
of  these  two  inequalities  are  reversed  there  will  certainly  be  a 
minimum.  For  then  we  can  (by  Cor.  3  of  §  121)  determine  an 
interval  (£  —  e,  £)  throughout  which  <f>  (x)  increases  with  x,  and  an 
interval  (£,  £  +  e)  throughout  which  it  decreases  as  x  increases: 
and  obviously  this  ensures  that  </>  (£)  shall  be  a  maximum. 

This  result  may  also  be  stated  thus.  If  the  sign  of  <f>  (x) 
changes  at  x  =  £  from  positive  to  negative,  then  x  =  f  gives 
a  maximum  of  </>  (x) :  and  if  the  sign  of  <£'  (x)  changes  in  the 
opposite  sense,  then  x  =  £  gives  a  minimum. 

123.  There  is  another  way  of  stating  the  conditions  for  a 
maximum  or  minimum  which  is  often  useful.  Let  us  assume 
that  (f>  (x)  has  a  second  derivative  <f>"  (x) :  this  of  course  does  not 
follow  from  the  existence  of  </>'  (x),  any  more  than  the  existence  of 
</>'  (x)  follows  from  that'  of  </>  (x).  But  in  such  cases  as  we  are 
likely  to  meet  with  at  present  the  condition  is  generally  satisfied. 

Theorem  D.  If  </>'(£)=0  and  <J>"(!)  +  0,  then  <f>(x)  has  a 
maximum  or  minimum  at  x=£,  a  maximum  if  <£"(f)<0,  a 
minimum  if  <£"  (£)  >  0. 

Suppose,  e.g.,  that  <f>"  (£)  <  0.  Then,  by  Theorem  A,  <tf  (x)  is 
negative  when  x  is  less  than  £  but  sufficiently  near  to  £,  and 
positive  when  x  is  greater  than  £  but  sufficiently  near  to  |.  Thus 
x  =  £  gives  a  maximum. 

124.  In  what  has  preceded  (apart  from  the  last  paragraph)  we  have 
assumed  simply  that  (f>(x)  has  a  derivative  for  all  values  of  x  in  the  interval 
under  consideration.  If  this  condition  is  not  fulfilled  the  theorems  cease  to 
be  true.     Thus  Theorem  B  fails  in  the  case  of  the  function 

y  =  \-J{x% 


122-124]  DERIVATIVES   AND   INTEGRALS  221 

where  the  square  root  is  to  be  taken  positive.  The  graph  of  this  function  is 
shown  in  Fig.  40.  Here  <£(-l)  =  <Ml)  =  0:  but  #'  (a?),  as  is  evident  from  the 
figure,  is  equal  to  1  if  x  is  negative  and  to  - 1  if  x  is  positive,  and  never 
vanishes.  There  is  no  derivative  for  x=0,  and  no  tangent  to  the  graph 
at  P.  And  in  this  case  x=0  obviously  gives  a  maximum  of  <£(#),  but 
<f>'  (0),  as  it  does  not  exist,  cannot  be 
equal  to  zero,  so  that  the  test  for  a 
maximum  fails. 

The  bare  existence  of  the  derivative 
(j>'  (x),  however,  is  all  that  we  have  as- 
sumed. And  there  is  one  assumption 
in  particular  that  we  have  not  made, 
and  that  is  that  0'  (x)  itself  is  a  con- 


tinuous function.     This  raises  a  rather     -1  O 

subtle  but  still  a  very  interesting  point.  pj„   ^ 

Can  a  function  <f>  (x)  have  a  derivative 

for  all  values  of  x  which  is  not  itself  continuous  ?     In  other  words  can  a 

curve  have  a  tangent  at  every  point,  and  yet  the  direction  of  the  tangent 

not  vary  continuously  ?     The  reader,  if  he  considers  what  the  question  means 

and  tries  to  answer  it  in  the  light  of  common  sense,  will  probably  incline 

to  the  answer  No.     It  is,  however,  not  difficult  to  see  that  this  answer  is 

wrong. 

Consider  the  function  $  (x)  defined,  when  x  4=  0,  by  the  equation 

<f)(x)=x2sm(\/x) ; 

and  suppose  that  0(0) =0.  Then  (f){x)  is  continuous  for  all  values  of  x. 
If  #4=0  then 

<£'  (x)  =  2x  sin  (1/x)  —  cos  (l/x) ; 

while  <f>  (0)  =  lira 7^— ;  =  0. 

7t-*o         ll 

Thus  (f>'  (x)  exists  for  all  values  of  x.  But  cf>'  (x)  is  discontinuous  for  x=0; 
for  2#sin  (l/x)  tends  to  0  as  x-*-Q,  and  cos  (l/x)  oscillates  between  the  limits 
of  indetermination  —1  and  1,  so  that  <£'  (x)  oscillates  between  the  same 
limits. 

What  is  practically  the  same  example  enables  us  also  to  illustrate  the 
point  referred  to  at  the  end  of  §  121.     Let 

<j)  (x)  =x2  sin  (l/.r)  +ax, 

where  0  <  a  <  1 ,  when  x  4=  0,  and  0  (0)  =  0.  Then  0'  (0)  =  a  >  0.  Thus  the 
conditions  of  Theorem  A  of  §  121  are  satisfied.     But  if  x  4=0  then 

$'  (x)  =  2x  sin  (l/x)  —  cos  (l/#)  +  a, 

which  oscillates  between  the  limits  of  indetermination  a  —  1  and  a  +  las  x-^0. 
As  a-l<0,  we  can  find  values  of  x,  as  near  to  0  as  we  like,  for  which 
<))'(x)<0;  and  it  is  therefore  impossible  to  find  any  interval,  including  x=0, 
throughout  which  0  (a?)  is  a  steadily  increasing  function  of  x. 


222  DERIVATIVES   AND    INTEGRALS  [VI 

It  i.s,  however,  impossible  that  <f)'(x)  should  have  what  was  called  in 
Ch.  V  (Ex.  xxxvu.  18)  a  'simple'  discontinuity;  e.g.  that  (])'(x)-*-a  when 
,r^-  +  0,  (f)'(.v)-*-b  when  %-*--0,  and  $'(0)  =  c,  unless  a  =  b  =  c,  in  which  case 
4>'(x)  is  continuous  for  x  =  0.     For  a  proof  see  §  125,  Ex.  xlvii.  3. 

Examples  XL VI.     1.     Verify  Theorem  B  when  0  (x)  =  (x-  a)m  (x  -  b)n  or 

(f)  (x)  =  (x-a)m(x-b)n(x-  c)p,  where  m,  n,  p  are  positive  integers  and  a<b<c. 
[The  first  function  vanishes  for  so=a  and  x  =  6.     And 

<j)'(x)  =  (x- a)m~l  (x-  b)"-1  {{m  +  n) x-mb-na) 
vanishes  for  x=(mb  +  na)l(m  +  n),   which  lies   between   a  and   b.      In   the 
second  case  we  have  to  verify  that  the  quadratic  equation 

(m  +  n+p)  xn-  -  {m  (6  +  c)  +  n  (c  +  a)  +p  (a  +  b)}  x  +  mbc  +  nca  +pab  =  0 
has  roots  between  a  and  b  and  between  b  and  c] 

2.  Show  that  the  polynomials 

2x-3  +  3.r2  -  12a;  +  7,     3.r4  +  8a-3  -  6#2  -  24r  + 19 
are  positive  when  x>l. 

3.  Show  that  x  -  sin  #  is  an  increasing  function  throughout  any  interval 
of  values  of  x,  and  that  tan.r-.r  increases  as  x  increases  from  -\ir  to  \tt. 
For  what  values  of  a  is  ax  —  amx  a  steadily  increasing  or  decreasing  function 
of  xl 

4.  Show  that  tana?  —  on  also  increases  from  x  =  \ir  to  x=Sir,  from  a?= In- 
to ,r  =  ^7r,  and  so  on,  and  deduce  that  there  is  one  and  only  one  root  of  the 
equation  t;u\x—x  in  each  of  these  intervals  (cf.  Ex.  xvn.  4). 

5.  Deduce    from    Ex.    3    that    sin#  —  x<0    if    x>0,   from    this    that 

cos  x-  1  +  2 x- >0,  and   from   this   that   &mx-x+ ^x3>0.     And,  generally, 

prove  that  if 

v-  r2m 

CW  =  cos*-l+'2!-...-(-l)><'  — , 


X 


.2m  +  1 


^2m  +  1  =  sin.r-^+--...-(-ir(2— 1):, 

and  .r>0,  then  C2m  and  #2m+i  are  positive  or  negative  according  as  m  is  odd 
or  even. 

6.  If  f(x)  and  /"  (x)  are  continuous  and  have  the  same  sign  at  every 
point  of  an  interval  (a,  b),  then  this  ii  terval  can  include  at  most  one  root  of 
either  of  the  equations /(.r)=0,/'  (.r)  =  0. 

7.  The  functions  ??,  v  and  their  derivatives  ?t',  v'  are  continuous 
throughout  a  certain  interval  of  values  of  x,  and  uv'  —  u'v  never  vanishes 
at  any  point  of  the  interval.  Show  that  between  any  two  roots  of  «=0 
lies  one  of  v  =  0,  and  conversely.     Verify  the  theorem  when  u  =  cosx,  i>=sin:r. 

[If  v  does  not  vanish  between  two  roots  of  u  =  0,  say  a  and  /3,  then  the 
function  ujv  is  continuous  throughout  the  interval  (a,  /3)  and  vanishes  at  its 
extremities.  Hence  (ujv)'  =  (u'v  —  uv')lv2  must  vanish  between  a  and  /3,  which 
contradicts  our  hypothesis.] 


124]  DERIVATIVES   AND   INTEGRALS  223 

8.  Determine  the  maxima  and  minima  (if  any)  of  (cc—  l)2(.r  +  2),  a?-Z% 
2.c3-3.£2-36.r+10,  4^-18a;2  +  27.>,'-7,  3xi-4xs+l,  afi-  15u;:i  +  3.  In  each 
case  sketch  the  form  of  the  graph  of  the  function. 

[Consider  the  last  function,  for  example.  Here  <£'  (x)  =  bx2  (x2  -  9),  which 
vanishes  for  x=  -3,  x=0,  and  x  =  Z.  It  is  easy  to  see  that  x=  -3  gives  a 
maximum  and  x=3  a  minimum,  while  x  =  0  gives  neither,  as  <f>'(x)  is  negative 
on  both  sides  of  #=0.] 

9.  Discuss  the  maxima  and  minima  of  the  function  (x  -  a)m  (x  -  b)n,  where 
m  and  n  are  any  positive  integers,  considering  the  different  cases  which  occur 
according  as  m  and  n  are  odd  or  even.     Sketch  the  graph  of  the  function. 

10.  Discuss  similarly  the  function  (x  -  a)  (x  -  b)2  (x  -  c)s,  distinguishing 
the  different  forms  of  the  graph  which  correspond  to  different  hypotheses  as 
to  the  relative  magnitudes  of  a,  b,  c. 

11.  Show  that  (ax  +  b)/(cx  +  d)  has  no  maxima  or  minima,  whatever 
values  a,  b,  c,  d  may  have.     Draw  a  graph  of  the  function. 

12.  Discuss  the  maxima  and  minima  of  the  function 

y  =  (ax2  +  2bx  +  c)/(Ax2  +  2Bx  +  c), 

when  the  denominator  has  complex  roots. 

[We  may  suppose  a  and  A  positive.     The  derivative  vanishes  if 

(ax  +  b)(Bx  +  C)-(Ax  +  B)(bx  +  c)  =  0 (1). 

This  equation  must  have  real  roots.  For  if  not  the  derivative  would  always 
have  the  same  sign,  and  this  is  impossible,  since  y  is  continuous  for  all  values 
of  x,  and  y->-a/A  as  x->-  +  qo  or  x-*~—  oo .  It  is  easy  to  verify  that  the  curve 
cuts  the  line  y  =  a/A  in  one  and  only  one  point,  and  that  it  lies  above  this 
line  for  large  positive  values  of  x,  and  below  it  for  large  negative  values,  or 
vice  versa,  according  as  bja>  BjA  or  b/a<BjA.  Thus  the  algebraically 
greater  root  of  (1)  gives  a  maximum  if  bja>BjA,  a  minimum  in  the  contrary 
case.] 

13.  The  maximum  and  minimum  values  themselves  are  the  values  of  A 
for  which  ax2  +  2bx  +  c-\(Ax2  +  2Bx  +  C)  is  a  perfect  square.  [This  is  the 
condition  that  y  =  \  should  touch  the  curve.] 

14.  In  general  the  maxima  and  minima  of  R(x)  =  P(x)jQ(x)  are  among 
the  values  of  X  obtained  by  expressing  the  condition  that  P(x)  —  \Q(x)  =  0 
should  have  a  pair  of  equal  roots. 

15.  If  Ax2~[-2Bx  +  C=0  has  real  roots  then  it  is  convenient  to  proceed  as 
follows.     We  have 

y  -  (a/A  )  =  (Xv+n)/{A(Ax2  +  2nx  +  C)}, 

where    \=bA—  aB,    p=cA—aC.     Writing   further   £   for   X.r  +  /x   and    r)    for 

(A/X2)(Ay-a),  we  obtain  an  equation  of  the  form 

9=l/{(|-i>)(|-?)}. 


224 


DERIVATIVES   AND   INTEGRALS 


[VI 


This  transformation  from  (x,  y)  to  (£,  rj)  amounts  only  to  a  shifting  of  the 
origin,  keeping  the  axes  parallel  to  themselves,  a  change  of  scale  along  each 
axis,  and  (if  X  <0)  a  reversal  in  direction  of  the  axis  of  abscissae;  and  so  a 
minimum  of  y,  considered  as  a  function  of  x,  corresponds  to  a  minimum  of  7 
considered  as  a  function  of  |,  and  vice  versa,  and  similarly  for  a  maximum. 

The  derivative  of  r\  with  respect  to  £  vanishes  if 

(£-p)(|-?)-K£-p)-£(£-?)=o, 

or  if  £2=pq.  Thus  there  are  two  roots  of  the  derivative  if  p  and  q  have  the 
same  sign,  none  if  they  have  opposite  signs.  In  the  latter  case  the  form  of 
the  graph  of  r\  is  as  shown  in  Fig.  41  a. 


Fig.  41a. 


Fig.  41c. 


Y 

v__ 

O 

*2          X 

Fiir.  42. 


When  p  and  q  are  positive  the  general  form  of  the  graph  is  as  shown  in 
Fig.  41  b,  and  it  is  easy  to  see  that  i-  =  *J(pq)  gives  a  maximum  and  |=  —  *J{pq) 
a  minimum.* 

In  the  particular  case  in  which  p  —  q  the 
function  is 

v=§K£-p)\ 

and  its  graph  is  of  the  form  shown  in  Fig.  41c. 
The  preceding  discussion  fails  if  X  =  0,  i.e. 
if  a/A  =  b/B.     But  in  this  case  we  have 
y-{alA)=pl{A  (Ax2  +  2Bx+C)} 

=  ^l{A2(x-x1)(x-x2)}, 

say,  and  dy/dx  =  0  gives  the  single  value  x  =  ^(xi  +  x.>).  On  drawing  a  graph 
it  becomes  clear  that  this  value  gives  a  maximum  or  minimum  according  as 
\i  is  positive  or  negative.  The  graph  shown  in  Fig.  42  corresponds  to  the 
former  case. 

[A  full  discussion  of  the  general  function  y  =  (ax2  +  2bx  +  c)/(A x2  +  2Bx  +  C)r 
by  purely  algebraical  methods,  will  be  found  in  Chrystal's  Algebra,  vol  i, 
pp.  464-7.] 

16.  Show  that  (x-a)  (x  —  ft)/(x-y)  assumes  all  real  values  as  x  varies,  if 
y  lies  between  a  and  /3,  and  otherwise  assumes  all  values  except  those  included 
in  an  interval  of  length  4v/(|a-y|  |/3--y|). 

*  The  maximum  is  -ll(\fp- Jq)\  the  minimum  -ll(Jp  +  Jq)2,  of  which  the 
latter  is  the  greater. 


124]  DERIVATIVES    AND    INTEGRALS  225 

17.  Show  that 

_x2  +  2x+e 

can  assume  any  real  value  if  0  <c  <  1,  and  draw  a  graph  of  the  function  in 
this  case.  (Math.  Trip.  1910.) 

18.  Determine  the  function  of  the  form  (ax2  +  2bx  +  c),(Ax2+2Bx  +  C) 
which  has  turning  values  (i.e.  maxima  or  minima)  2  and  3  when  x=\  and 
tt=  - 1  respectively,  and  has  the  value  2-5  when  x  =  0.        (Math.  Trip.  1908.) 

19.  The  maximum  and  minimum  of  (x  +  a)(x  +  b)j(x-a)  (x-b),  where  a 
and  b  are  positive,  are 

Wa+Jby  /Ja-Jby 


20.  The  maximum  value  of  (x  -  l)2/(#  + 1)3  is  -^. 

21.  Discuss  the  maxima  and  minima  of 

x  (x -  \)\(x2  +  3x  +  3),     **/(* - 1 )  (x -  3)3, 
(x-Yf-  (3x2-  2x-31)l(x  +  o)2(3x2-  Ux-  1). 

(Math.  Trip.  1898.) 
[If  the  last  function  be  denoted  by  P(x)/Q(x),  it  will  be  found  that 
P'Q-PQ'  =  -2  (x -  7)  (x  -3)(x-l)(x  +  l)  (x  +  2)  (x  +  5).] 

22.  Find  the  maxima  and  minima  of  a  cos  x  +  b  sin  x.  Verify  the  result 
by  expressing  the  function  in  the  form  A  cos  (x  —  a). 

23.  Find  the  maxima  and  minima  of 

a2  cos2  x  +  b2  si  n2  x,     A  cos2  x  +  2  H  cos  x  si  n  x  +  B  si  n2  x. 

24.  Show  that  sin  (x  +  a)/sin  (a-  +  b)  has  no  maxima  or  minima.  Draw 
a  graph  of  the  function. 

25.  Show  that  the  function 

sin2#  .„  . 

sin  (a  +  a)  sin  (# +6) 

has  an  infinity  of  minima  equal  to  0  and  of  maxima  equal  to 

-  4  sin  a  sin  6/sin2(a  -  b).  (Math.  Trip.  1909.) 

26.  The  least  value  of  a2  sec2  x  +  b2  cosec2  x  is  (a  +  b)2. 

27.  Show  that  tan  3x  cot  2a;  cannot  lie  between  ^  and  f. 

28.  Show  that,  if  the  sum  of  the  lengths  of  the  hypothenuse  and  another 
side  of  a  right-angled  triangle  is  given,  then  the  area  of  the  triangle  is  a 
maximum  when  the  angle  between  those  sides  is  60°.  (Math.  Trip.  1909.) 

29.  A  line  is  drawn  through  a  fixed  point  (a,  b)  to  meet  the  axes  OX,  OY 
in  P  and  Q.  Show  that  the  minimum  values  of  PQ,  OP+OQ,  and  OP.  OQ 
are  respectively  (a2/3  +  b2/3f12,  (*Ja  +  Jb)\  and  Aab. 

H.  15 


226 


DERIVATIVES    AND    INTEGRALS 


[VI 


30.  A  tangent  to  an  ellipse  meets  the  axes  in  P  and  Q.     Show  that  the 
least  value  of  PQ  is  equal  to  the  sum  of  the  semiaxes  of  the  ellipse. 

31.  Find  the  lengths  and  directions  of  the  axes  of  the  conic 

ax2  +  2hxy  +  by2 = 1 . 

[The  length  r  of  the  semidiameter  which  makes  an  angle  6  with  the  axis 

of  x  is  given  by 

\\r2  =  a  cos2  6  +  2A  cos  6  sin  6  +  b  sin2  6. 

The  condition  for  a  maximum  or  minimum  value  of  r  is  tan2#  =  2/t/(«  —  b). 
Eliminating  6  between  these  two  equations  we  find 

{a-(l/!*)}{&-(l/^}=A«] 

32.  The    greatest    value    of  xmyn,   where  x  and    v  are    positive  and 
x+y  =  k,  is 

mm nn lm  +  n/(m  +  n)m  +  n.  . 

33.  The   greatest   value   of  ax  +  by,  where  x  and  y   are  positive  and 
x2  +  xy+yi  =  3K2,  is 

2k  VO2  -ab  +  b2). 

[If  ax-Vby  is  a  maximum  then  a  +  b(dy/dx)  =  0.     The  relation  between  x 
and  y  gives  (2x+y)  +  (x+2y)  (dy!dx)  =  Q.     Equate  the  two  values  of  dy/dx.] 

34.  If  6  and  0  are  acute  angles  connected  by  the  relation  asecd  +  bsec<j>  =  c, 
where  a,  b,  c  are  positive,  then  a  cos  6  +  b  cos  ^  is  a  minimum  when  6  =  (f>. 

125.  The  Mean  Value  Theorem.  We  can  proceed  now  to 
the  proof  of  another  general  theorem  of  extreme  importance,  a 
theorem  commonly  known  as  '  The  Mean  Value  Theorem  or  '  The 
Theorem  of  the  Mean'. 

Theorem.     If  <f>  (x)  has  a  derivative  for  all  values  of  x  in  the 
interval    (a,    b),  then  there   is  a 
value  i;  of  x  between  a  and  b, 
such  that 


m 


<p(a) 


*(&)-*  («*)-(& -a)*' (ft 

Before  we  give  a  strict  proof 
of  this  theorem,  which  is  perhaps 
the  most  important  theorem  in 
the  Differential  Calculus,  it  will 
be  well  to  point  out  its  obvious 
geometrical  meaning.  This  is 
simply  (see  Fig.  43)  that  if  the 
curve  APB  has  a  tangent  at  all  points  of  its  length  then  there 


Fig.  43. 


124,  125]  DERIVATIVES    AND    INTEGRALS  227 

must  be  a  point,  such  as  P,  where  the  tangent  is  parallel  to  All. 
For  </>'(£)  is  the  tangent  of  the  angle  which  the  tangent  at  P 
makes  with  OX,  and  {<f>  (b)  -  </>  («)}/(&  -  a)  the  tangent  of  the  angle 
which  AB  makes  with  OX. 

It  is  easy  to  give    a   strict   analytical    proof.     Consider   the 
function 

*(&)-*(*)-|^  {*(&)-*  (a)}, 

which  vanishes  when  x  =  a  and  x  =  b.  It  follows  from  Theorem  B 
of  §  121  that  there  is  a  value  £  for  which  its  derivative  vanishes. 
But  this  derivative  is 


jz^r    -f(#); 


0(ft)-0(Q) 

a 

which  proves  the  theorem.  It  should  be  observed  that  it  has  not 
been  assumed  in  this  proof  that  <f>'  (x)  is  continuous. 

It  is  often  convenient  to  express  the  Mean  Value  Theorem  in 
the  form 

<j>(b)  =  cf>  (a)  +  (b-a)  4>'  {a  +  d(b-  a)}, 

where  0  is  a  number  lying  between  0  and  1.  Of  course  a  +  6(b—a) 
is  merely  another  way  of  writing  'some  number  £  between  a  and  b\ 
If  we  put  b  =  a  +  h  we  obtain 

</>  (a  +  li)  =  <j>  (a)  +  h(f>'  (a  +  6h), 

which  is  the  form  in  which  the  theorem  is  most  often  quoted. 

Examples  XL VII.     1.    Show  that 

.0  (6)  -0(#)-|^&  (&)-*(«)} 

is  the  difference  between  the  ordinates  of  a  point  on  the  curve  and  the 
corresponding  point  on  the  chord. 

2.  Verify  the  theorem  when  cf>  (,v)  =  x2  and  when  <£  (x)  =  x3. 

[In  the  latter  case  we  have  to  prove  that  (63  —  a3)/(b  —  a)  =  3£2,  where 
a<£<b;  i.e.  that  if  J  (b2  +  ab  +  a2)  =  £2  then  £  lies  between  a  and  b.] 

3.  Establish  the  theorem  stated  at  the  end  of  §  1 24  by  means  of  the  Mean 
Value  Theorem. 

[Since  <f>'(0)  =  c,  we  can  find  a  small  positive  value  of  .r  such  that 
{(f>(x)-(f>(0)}/x  is  nearly  equal  to  c;  and  therefore,  by  the  theorem,  a  small 
positive  value  of  £  such  that  $  (£)  is  nearly  equal  to  c,  which  is  inconsistent 
with    lim   (f>'(x)  =  a,  unless  a  =  c.     Similarly  6  =  c] 

15—2 


228  DERIVATIVES   AND    INTEGRALS  [VI    | 

4.     Use  the  Mean  Value  Theorem  to  prove  Theorem  (6)  of  §  113,  assuming 
that  the  derivatives  which  occur  are  continuous. 
[The  derivative  of  F{f(x)}  is  by  definition 

F{f(x  +  h)}-F{f(x)} 
nm  h  ' 

But,  by  the  Mean  Value  Theorem,  f(x+h)=f(x)  +  hf '(g),  where  £  is  a  number 
lying  between  x  and  x  +  h.     And 

F{f(.v)+hf'(g)}=F{f(x)}+hf  (£)*"  (£), 
where  £j  is  a  number  lying  between/(o;)  and  f(x)  +  hf  (£).    Hence  the  deriva- 
tive of  F{f(x))  is 

lim/'  (|)  F'  «,)=/'  (x)F'{f(x)}, 

since  £-»-#  and  £i-*~f(x)  as  A-»-0.] 

126.  The  Mean  Value  Theorem  furnishes  us  with  a  proof  of  a 
result  which  is  of  great  importance  in  what  follows :  if  <f>'  (x)  =  0, 
throughout  a  certain  interval  of  values  of  x,  then  </>  (x)  is  constant 
throughout  that  interval. 

For,  if  a  and  b  are  any  two  values  of  x  in  the  interval,  then 
</>  (b)  -  (/>  (a)  =  (b  -  a)  <£'  {a  +  d  (b  -  a)}  =  0. 
An   immediate  corollary  is  that  if  <f>'  (x)  =  y\r'  (x),  throughout  a 
certain  interval,  then  the  functions  cf>  (x)  and  -^  (x)  differ  through- 
out that  interval  by  a  constant. 

127.  Integration.  We  have  in  this  chapter  seen  how  we  can 
find  the  derivative  of  a  given  function  (f>  (x)  in  a  variety  of  cases, 
including  all  those  of  the  commonest  occurrence.  It  is  natural  to 
consider  the  converse  question,  that  of  determining  a  function 
whose  derivative  is  a  given  function. 

Suppose  that  yfr  (x)  is  the  given  function.  Then  we  wish  to 
determine  a  function  such  that  <f>'  (x)  =  \jr  (x).  A  little  reflection 
shows  us  that  this  question  may  really  be  analysed  into  three 
parts. 

(1)  In   the   first   place  we  want  to  know   whether   such    a 
function  as  <f>  (x)  actually  exists.     This  question  must  be  carefully 
distinguished  from  the  question  as  to   whether  (supposing  that* 
there  is  such  a  function)  we  can  find   any   simple    formula    to 

express  it. 

(2)  We  want  to  know  whether  it  is  possible  that  more  than 
one  such  function  should  exist,  i.e.  we  want  to  know  whether  our 


125-127]  DERIVATIVES    AND   INTEGRALS  229 

problem  is  one  which  admits  of  a  unique  solution  or  not ;  and 
if  not,  we  want  to  know  whether  there  is  any  simple  relation 
between  the  different  solutions  which  will  enable  us  to  express  all 
of  them  in  terms  of  any  particular  one. 

(3)  If  there  is  a  solution,  we  want  to  know  how  to  find  an 
actual  expression  for  it. 

It  will  throw  light  on  the  nature  of  these  three  distinct  ques- 
tions if  we  compare  them  with  the  three  corresponding  questions 
which  arise  with  regard  to  the  differentiation  of  functions. 

(1)  A  function  <f>  (x)  may  have  a  derivative  for  all  values  of  x, 
like  xm,  where  m  is  a  positive  integer,  or  sin  x.  It  may  generally, 
but  not  always  have  one,  like  \/x  or  tan#  or  sec  a;.  Or  again 
it  may  never  have  one  :  for  example,  the  function  considered  in 
Ex.  xxxvii.  20,  which  is  nowhere  continuous,  has  obviously  no 
derivative  for  any  value  of  x.  Of  course  during  this  chapter  we 
have  confined  ourselves  to  functions  which  are  continuous  except  for 
some  special  values  of  x.  The  example  of  the  function  ZJx,  how- 
ever, shows  that  a  continuous  function  may  not  have  a  derivative 
for  some  special  value  of  x,  in  this  case  x  =  0.  Whether  there 
are  continuous  functions  which  never  have  derivatives,  or  con- 
tinuous curves  which  never  have  tangents,  is  a  further  question 
which  is  at  present  beyond  us.  Common-sense  says  No :  but,  as 
we  have  already  stated  in  §  111,  this  is  one  of  the  cases  in  which 
higher  mathematics  has  proved  common-sense  to  be  mistaken. 

But  at  any  rate  it  is  clear  enough  that  the  question  '  has  (j>  (x) 
a  derivative  <£'  (x)  ? '  is  one  which  has  to  be  answered  differently 
in  different  circumstances.  And  we  may  expect  that  the  converse 
question  '  is  there  a  function  <f>  (x)  of  which  y{r  (x)  is  the  deriva- 
tive ? '  will  have  different  answers  too.  We  have  already  seen 
.that  there  are  cases  in  which  the  answer  is  No :  thus  if  yfr  (x)  is 
the  function  which  is  equal  to  a,  b,  or  c  according  as  x  is  less  than, 
equal  to,  or  greater  than  0,  then  the  answer  is  No  (Ex.  xlvii.  3), 
unless  a  =  b  =  c. 

This  is  a  case  in  which  the  given  function  is  discontinuous. 
In  what  follows,  however,  we  shall  always  suppose  yjr(x)  continuous. 
And  then  the  answer  is  Yes :  if  ^r{x)  is  continuous  then  there  is 
always  a  function  <f>  (x)  such  that  <f>'  (x)  =  yjr  (x).  The  proof  of  this 
will  be  given  in  Ch.  VII. 


230  DERIVATIVES   AND   INTEGRALS  [VI 

(2)  The  second  question  presents  no  difficulties.  In  the  case 
of  differentiation  we  have  a  direct  definition  of  the  derivative 
which  makes  it  clear  from  the  beginning  that  there  cannot 
possibly  be  more  than  one.  In  the  case  of  the  converse  problem 
the  answer  is  almost  equally  simple.  It  is  that  if  <f>  (x)  is  one 
solution  of  the  problem  then  <f>  (x)  +  G  is  another,  for  any  value  of 
the  constant  C,  and  that  all  possible  solutions  are  comprised  in 
the  form  <f>  (x)  +  C.     This  follows  at  once  from  §  126. 

(3)  The  practical  problem  of  actually  finding  <£'  (x)  is  a  fairly 
simple  one  in  the  case  of  any  function  defined  by  some  finite  com- 
bination of  the  ordinary  functional  symbols.  The  converse  problem 
is  much  more  difficult.  The  nature  of  the  difficulties  will  appear 
more  clearly  later  on. 

Definitions.  If  \jr  (x)  is  the  derivative  of  <£  (x),  then  we  call 
(f>  (x)  an  integral  or  integral  function  of  -^  (./').  The  operation 
of  forming  yjr  (x)  from  </>  (x)  ive  call  integration. 

We  shall  use  the  notation 

<f)(x)=  I  yjr  (x)  dx. 

It  is  hardly  necessary  to  point  out  that  \...dx  like  djdx  must,  at 
present  at  any  rate,  be  regarded  purely  as  a  symbol  of  operation : 
the  I  and  the  dx  no  more  mean  anything  when  taken  by  them- 
selves than  do  the  d  and  dx  of  the  other  operative  symbol  djdx. 

128.  The  practical  problem  of  integration.  The  results 
of  the  earlier  part  of  this  chapter  enable  us  to  write  down  at  once 
the  integrals  of  some  of  the  commonest  functions.     Thus 

r  xm+1         [  .  f  . 

I  xmdx  = ,      I  cos  xdx  =  sin  x,     I  sin  xdx  =  —  cos  x. .  .(1). 

These  formulae  must  be  understood  as  meaning  that  the 
function  on  the  right-hand  side  is  one  integral  of  that  under 
the  sign  of  integration.  The  most  general  integral  is  of  course 
obtained  by  adding  to  the  former  a  constant  C,  known  as  the 
arbitrary  constant  of  integration. 


127,  128]  DERIVATIVES   AND   INTEGRALS  231 

There  is  however  one  case  of  exception  to  the  first  formula,  that 
in  which  m=  —  1.  In  this  case  the  formula  becomes  meaningless, 
as  is  only  to  be  expected,  since  we  have  seen  already  (Ex.  xlii.  4) 
that  l/x  cannot  be  the  derivative  of  any  polynomial  or  rational 
fraction. 

That  there  really  is  a  function  F(x)  such  that  DxF(x)  =  1/sc 
will  be  proved  in  the  next  chapter.  For  the  present  we  shall  be 
content  to  assume  its  existence.  This  function  F(x)  is  certainly 
not  a  polynomial  or  rational  function ;  and  it  can  be  proved  that 
it  is  not  an  algebraical  function.  It  can  indeed  be  proved  that 
F(x)  is  an  essentially  new  function,  independent  of  any  of  the 
classes  of  functions  which  we  have  considered  yet,  that  is  to  sav 
incapable  of  expression  by  means  of  any  finite  combination  of  the 
functional  symbols  corresponding  to  them.  The  proof  of  this  is 
unfortunately  too  detailed  and  tedious  to  be  inserted  in  this  book; 
but  some  further  discussion  of  the  subject  will  be  found  in  Ch.  IX, 
where  the  properties  of  F(x)  are  investigated  systematically. 

Suppose  first  that  x  is  positive.     Then  we  shall  write 

'dx 


i 


=  lo^# (2), 

x         °  w' 


and  Ave  shall  call  the  function  on  the  right-hand  side  of  this 
equation  the  logarithmic  function :  it  is  defined  so  far  only  for 
positive  values  of  x. 

Next  suppose  x  negative.    Then  —  x  is  positive,  and  so  log  (— x) 
is  defined  by  what  precedes.     Also 

dx     °  —  x     x' 

so  that,  when  x  is  negative, 


/ 


^  =  log(-*) (3). 


The  formulae  (2)  and  (3)  may  be  united  in  the  formulae 
'dx 


j—  =  \og(±x)  =  \og\x\    (4), 

J     00 


where  the  ambiguous  sign  is  to  be  chosen  so  that  ±  x  is  positive 
these  formulae  hold  for  all  real  values  of  x  other  than  x  =  0. 


232  DERIVATIVES   AND   INTEGRALS  [VI 

The  most  fundamental  of  the  properties  of  log  x  which  will  be  proved  in 
Ch.  IX  are  expressed  by  the  equations 

log  1  =  0,     log  (1/x)  =  -  log  x,     log  xy  =  log  re  +  log  y, 

of  which  the  second  is  an  obvious  deduction  from  the  first  and  third.  It  is 
not  really  necessary,  for  the  purposes  of  this  chapter,  to  assume  the  truth  of 
any  of  these  formulae ;  but  they  sometimes  enable  us  to  write  our  formulae 
in  a  more  compact  form  than  would  otherwise  be  possible. 

It  follows  from  the  last  of  the  formulae  that  log.r2  is  equal  to  21og.^  if 
x  >  0  and  to  2  log  ( -  x)  if  x  <  0,  and  in  either  case  to  2  log  |  x  \ .  Either  of  the 
formulae  (4)  is  therefore  equivalent  to  the  formula 


I 


/' 


_L  =  i  log  x2  (5). 


The  five  formulae  (1) — (3)  are  the  five  most  fundamental 
standard  forms  of  the  Integral  Calculus.  To  them  should  be 
added  two  more,  viz. 

I =  arc  tan  x,     \  -jyz -.  =  +  arc  sin  x* (6). 

129.  Polynomials.  All  the  general  theorems  of  §  113  may  of 
course  also  be  stated  as  theorems  in  integration.  Thus  we  have, 
to  begin  with,  the  formulae 

!{f(x)  +  F(x)}  dx=  ff(x)dx  +  JF{x)dx (1), 

!kf(x)dx  =  kjf(x)dx (2). 

Here  it  is  assumed,  of  course,  that  the  arbitrary  constants  are 
adjusted  properly.  Thus  the  formula  (1)  asserts  that  the  sum  of 
any  integral  of  f(x)  and  any  integral  of  F  (x)  is  an  integral  of 

f{x)  +  F(x). 

These  theorems  enable  us  to  write  down  at  once  the  integral 
of  any  function  of  the  form  £  A  „/„(#),  the  sum  of  a  finite  number 
of  constant  multiples  of  functions  whose  integrals  are  known.  In 
particular  we  can  write  down  the  integral  of  any  polynomial: 
thus 


/ 


x  7       a0a"+1     a,x11 

(fl(/t''1  +  ttiX71-1  +  ...  +  an) dx  —  —  -—-  H +  . . .  +  anx. 

n  +  l         n 

*  See  §  119  for  the  rule  for  determining  the  ambiguous  sign. 


128-130]  DERIVATIVES    AND    INTEGRALS  233 

130.     Rational  Functions.     After  integrating  polynomials 

it  is  natural  to  turn  our  attention  next  to  rational  functions. 
Let  us  suppose  R  (x)  to  be  any  rational  function  expressed  in  the 
standard  form  of  §  117,  viz.  as  the  sum  of  a  polynomial  II  (x)  and 
a  number  of  terms  of  the  form  A/(x  —  ol)p. 

We  can  at  once  write  down  the  integrals  of  the  polynomial 
and  of  all  the  other  terms  except  those  for  which  p  =  1,  since 


/i 


A       ,  A  1 

ax  =  — 


(x  -  a)P  p  -  1  {x  -  a)^1 ' 

whether  a  be  real  or  complex  (§  117). 

The    terms   for   which  p  =  1   present   rather   more   difficulty. 
It  follows  immediately  from  Theorem  (6)  of  §  113  that 


\F'{f{x)}f'{x)alx  =  F{f(x)} (3). 


In  particular,  if  we  take  f(x)  =  ax  +  b,  where  a  and  b  are  real, 
and  write  <p  (x)  for  F  (x)  and  i/r  (x)  for  F'  (x),  so  that  </>  (x)  is  an 
integral  of  yfr  (x),  we  obtain 


/ 


■fr(ax  +  b)  dx  =  -<f)  (ax  +  b)    (4). 


Thus,  for  example, 

dx         1,,  7 

—v  =  -  log  \ax-\-  b 
ax  +  b     a 

and  in  particular,  if  a  is  real, 

dx 


/, 


/ 


a  =  \og\x-a\. 


We  can  therefore  write  down  the  integrals  of  all  the  terms  in 
i?  (as)  for  which  p  =  1  and  a.  is  real.  There  remain  the  terms  for 
which  p  =  1  and  a  is  complex. 

In  order  to  deal  with  these  we  shall  introduce  a  restrictive 
hypothesis,  viz.  that  all  the  coefficients  in  R  (x)  are  real.  Then  if 
a  =  ry  +  8i  is  a  root  of  Q  (x)  =  0,  of  multiplicity  m,  so  is  its  con- 
jugate a  =  y  —  Bi;  and  if  a  partial  fraction  Ap/(x—a)p  occurs  in 
the  expression  of  R  (x),  so  does  Ap/(x  —  a)P,  where  Ap  is  conjugate 
to  Ap.  This  follows  from  the  nature  of  the  algebraical  processes 
by  means  of  which  the  partial  fractions  can  be  found,  and  which 
are  explained  at  length  in  treatises  on  Algebra*. 

*  See,  for  example,  Ckrystal's  Algebra,  vol.  i,  pp.  151-9. 


23-i  DERIVATIVES   AND    INTEGRALS  [VI 

Thus,  if  a  term  (A.  +  fii)/(x  —  y  —  Si)  occurs  in  the  expression 
of  R  (x)  in  partial  fractions,  so  will  a  term  (A  —  fii)/(x  —  y  +  Si)  ; 
and  the  sum  of  these  two  terms  is 

(x-y)2+B2 

This  fraction  is  in  reality  the  most  general  fraction  of  the  form 

Ax  +  B 
ax2  +  2bx  +  c ' 

where  b2  <  ac.  The  reader  will  easily  verify  the  equivalence  of 
the  two  forms,  the  formulae  which  express  A,  /a,  y,  8  in  terms  of 
A,  B,  a,  b,  c  being 

A  =  A/2a,     (x  =  -  D/(2ch/A),     y  =  -bja,     S  =  \/A/a, 
where  A  =  ac  —  b2,  and  D=  aB—bA. 

If  in  (3)  we  suppose  j^  {f{x)\  to  be  log  \f(x)  |,  we  obtain 

|^^  =  log|/(.r)i    (5); 

and  if  we  further  suppose  that/(#)  =  (x  —  X)2  +  /a2,  we  obtain 

And,  in  virtue  of  the  equations  (6)  of  §  128  and  (4)  above,  we 
have 

f,    ~2^    ,  ^ 2S  arc  tan  (^±)  . 

These  two  formulae  enable  us  to  integrate  the  sum  of  the  two 
terms  which  we  have  been  considering  in  the  expression  of  R  (x) ; 
and  we  are  thus  enabled  to  write  down  the  integral  of  any  real 
rational  function,  if  all  the  factors  of  its  denominator  can  be  deter- 
mined. The  integral  of  any  such  function  is  composed  of  the  sum 
of  a  polynomial,  a  number  of  rational  functions  of  the  type 

A  1 

p  - 1  (x  -  ay-1 ' 

a  number  of  logarithmic  functions,  and  a  number  of  inverse  tangents. 

It  only  remains  to  add  that  if  a.  is  complex  then  the  rational 
function  just  written  always  occurs  in  conjunction  with  another  in 
which  A  and  a  are  replaced  by  the  complex  numbers  conjugate  to 
them,  and  that  the  sum  of  the  two  functions  is  a  real  rational  function. 


130]  DERIVATIVES   AND    INTEGRALS  235 

Examples  XL VIII.     1.     Prove  that 

[_^±B_      _A  .  D  \ax  +  b-s!{-A)\ 

(where  X=ax2+2bx  +  c)  if  A<0,  and 

{,  iXZB+    ^9-log|2-|  +  -4-arcten  ("+*) 
J  ax*+2bx+c  2a    ° '      '     ajA  \  Ja  J 

if  A  >  0,  A  and  D  having  the  same  meanings  as  on  p.  234. 
2.     In  the  particular  case  in  which  ac  =  b2  the  integral  is 

D  A, 

-\ — log   ax  +  b  \. 


a  (ax  +  b)      a 

3.     Show  that  if  the  roots  of  Q(x)=0  are  all  real  and  distinct,  and  P(x) 
is  of  lower  degree  than  Q  (,r),  then 

Il(x)dx  =  ^^\og\x-a\, 

the  summation  applying  to  all  the  roots  a  of  Q  (x)  =  0. 

[The  form  of  the  fraction  corresponding  to  a  may  be  deduced  from  the 
facts  that 

Q(x)      rut  \     /         \  »/  \      P(a) 


I 


as  #-»-a.] 

4.  If  all  the  roots  of  Q  (x)  are  real  and  a  is  a  double  root,  the  other  roots 
being  simple  roots,  and  P  (x)  is  of  lower  degree  than  Q  (x),  then  the  integral 
is  Aj(x-a)  + A' \og\  x-  a  |  +  2  B  log  |  x-fi  |,  where 

2PW  2{3P'(a)Q"(a)-P(a)Q"'(a)}  P(/3) 

"«"(«)'    '  3{§"(«)}2  '    ^~e(/3)' 

and  the  summation  applies  to  all  roots  13  of  Q  (x)  =  0  other  than  a. 

dx 


5.     Calculate 


[ 


L)(^2+l)}2' 
[The  expression  in  partial  fractions  is 

1  _1 i  2-t  t  2  +  i 

4(^-l)2  ~  2~(>-l)  "  8(a?-i)*  +  8(*-i)  +  8  (# +i)*  +  8 (#+i) ' 

and  the  integral  is 
1 


4(^-1)      4(^+1) 


-  \  log  |  x  —  1 1  + 1  log  (.t-  + 1) + j  arc  tan  x.] 


6.     Integrate 
x 


{x-a)(x-b)(x-cY      {x-af(x-by      (x-a)2  (x-b)2'      (x-a)3' 
x  x2  x2  —  a2  x2  —  a2 


(x2  +  a2)  (x2  +  b2) '     (x2  +  a2)  (x2 + b)2 '     x2  {x2  +  a2) '    x  {x2  +  a-)2  * 


h 


236  DERIVATIVES    AND   INTEGRALS  [VI 

7.     Prove  the  formulae : 

l+#*      4V2\  B\1-*V2+*"/  \l-«*/J 

131.     Note  on   the   practical   integration   of  rational   functions. 

The  analysis  of  §  130  gives  us  a  general  method  by  which  we  can  find  the 
integral  of  any  real  rational  function  R  (a?),  provided  ice  can  solve  the  equation 
Q{x)  =  0.  In  simple  cases  (as  in  Ex.  5  above)  the  application  of  the  method 
is  fairly  simple.  In  more  complicated  cases  the  labour  involved  is  some- 
times prohibitive,  and  other  devices  have  to  be  used.  It  is  not  part  of  the 
purpose  of  this  book  to  go  into  practical  problems  of  integration  in  detail. 
The  reader  who  desires  fuller  information  may  be  referred  to  Goursat's  Cows 
d?  Analyse,  second  ed.,  vol.  i,  pp.  246  et  seq.,  Bertrand's  Calcid  Integral,  and 
Dr  Bromwich's  tract  Elementary  Integrals  (Bowes  and  Bowes,  1911). 

If  the  equation  Q(.v)=0  cannot  be  solved  algebraically,  then  the  method  of 
partial  fractions  naturally  fails  and  recourse  must  be  had  to  other  methods'*. 

132.  Algebraical  Functions.  We  naturally  pass  on  next  to 
the  question  of  the  integration  of  algebraical  functions.  We  have 
to  consider  the  problem  of  integrating  y,  where  y  is  an  algebraical 
function  of  x.  It  is  however  convenient  to  consider  an  apparently 
more  general  integral,  viz. 


/ 


R  (x,  y)  dx, 


where  R  (x,  y)  is  any  rational  function  of  x  and  y.  The  greater 
generality  of  this  form  is  only  apparent,  since  (Ex.  XIV.  6)  the 
function  R  (x,  y)  is  itself  an  algebraical  function  of  x.  The  choice 
of  this  form  is  in  fact  dictated  simply  by  motives  of  convenience : 
such  a  function  as 

px  +  q  +  \f(ax2  +  2bx  +  c) 
px  +  q  —  \/(ax2  +  2bx  +  c) 

is  far  more  conveniently  regarded  as  a  rational  function  of  x  and 
the  simple  algebraical  function  *J(ax2  +  2bx  +  c),  than  directly  as 
itself  an  algebraical  function  of  x. 

*  See  the  author's  tract  "The  integration  of  functions  of  a  single  variable" 
{Cambridge  Tracts  in  Mathematics,  No.  2,  second  edition,  1915).  This  does  not 
often  happen  in  practice. 


130-134]  DERIVATIVES    AND    INTEGRALS  237 

133.     Integration   by   substitution    and   rationalisation. 

It  follows  from  equation  (3)  of  §  130  that  if  I  ty  (a?)  das  =  <£(#)  then 

*  {/(*)}/ (t)dt  =  <}>{f(t)} (1). 


/ 


This  equation  supplies  us  with  a  method  for  determining  the 
integral  of  yjr  (x)  in  a  large  number  of  cases  in  which  the  form  of 
the  integral  is  not  directly  obvious.  It  may  be  stated  as  a  rule  as 
follows:  put  x=f(t),  where  f (t)  is  any  function  of  a  new  variable 
t  which  it  may  be  convenient  to  choose ;  multiply  by  f  it),  and 
determine  (if  possible)  the  integral  of  -v/r  {f(t)}f  (t);  express  the 
result  in  terms  of  x.  It  will  often  be  found  that  the  function  of  t 
to  which  we  are  led  by  the  application  of  this  rule  is  one  whose 
integral  can  easily  be  calculated.  This  is  always  so,  for  example, 
if  it  is  a  rational  function,  and  it  is  very  often  possible  to  choose 
the  relation  between  x  and  t  so  that  this  shall  be  the  case.  Thus 
the  integral  of  R  (\Jx),  where  R  denotes  a  rational  function,  is 
reduced  by  the  substitution  x=t2  to  the  integral  of  2tR(t2), 
i.e.  to  the  integral  of  a  rational  function  of  t.  This  method  of 
integration  is  called  integration  by  rationalisation,  and  is  of 
extremely  wide  application. 

Its  application  to  the  problem  immediately  under  consideration 
is  obvious.  If  we  can  find  a  valuable  t  such  that  x  and  y  are  both 
rational  functions  oft,  say  x  =  R^t),  y  =  R2(t),  then 

f  R(x,  y)  dx  =  JR  {R^t),  R2(t)}  Ri(t)dt, 

and  the  latter  integral,  being  that  of  a  rational  function  of  t,  can  be 
calculated  by  the  methods  o/§  130. 

It  would  carry  us  beyond  our  present  range  to  enter  upon  any 
general  discussion  as  to  when  it  is  and  when  it  is  not  possible  to 
find  an  auxiliary  variable  t  connected  with  x  and  y  in  the  manner 
indicated  above.  We  shall  consider  only  a  few  simple  and  inter- 
esting special  cases. 

134.  Integrals  connected  with  conies.  Let  us  suppose 
that  x  and  y  are  connected  by  an  equation  of  the  form 

ax2  +  2hxy  +  by2  +  2gx  +  2fy  +  c  =  0; 

in  other  words  that  the  graph  of  y,  considered  as  a  function  of  x 


238  DERIVATIVES   AND    INTEGRALS  [VI 

is  a  conic.  Suppose  that  (£,  rj)  is  any  point  on  the  conic,  and 
let  x  —  £  =  X,  y  —  t]  =  Y.  If  the  relation  between  x  and  y  is 
expressed  in  terms  of  X  and  Y,  it  assumes  the  form 

aX2  +  2hXY  +  bY2  +  2GX  +  2FY  =  0, 
where    F=hi;  +  br}+f,  G  =  ag  +  hrj  +  g.     In   this    equation   put 
Y  =  tX.     It  will  then  be   found   that   X   and    Y  can   both   be 
expressed  as  rational  functions  of  t,  and  therefore  x  and  y  can 
be  so  expressed,  the  actual  formulae  being 

2  (G  +  Ft)  _  2t(G  +  Ft) 

x~s-      a+2ht  +  bt2  '     y     V  a  +  2ht  +  bt*' 

Hence  the  process  of  rationalisation  described  in  the  last  section 
can  be  carried  out. 

The  reader  should  verify  that 

hx  +  by+f=-%(a+2ht  +  to)  ~ , 

S0  that  J  hxTbf+f  ~      " J  a+2ht  +  to ' 

When  h-  >  ab  it  is  in  some  ways  advantageous  to  proceed  as 

follows      The  conic  is  a  hyperbola  whose  asymptotes  are  parallel 

to  the  lines 

ax2  +  2hxy  +  by2  =  0, 

or  b(y-  fix)  (y  -  fi'x)  =  0, 

say      If  we  put  y  —  fxx  =  t,  we  obtain 

?rriB  +  2/V  +  c 

and  it  is  clear  that  x  and  y  can  be  calculated  from  these  equations 
as  rational  functions  of  t.  We  shall  illustrate  this  process  by  an 
application  to  an  important  special  case. 


135.      The  integral 


/  -77 — o  .  nt — ; — c«      Suppose    in    particular    that 
J  J  (ax1  +  2bx  +  c)  rr  r 

y2  =  ax2  +  2bx  +  c,  where  a>0.     It  will  be  found  that,  if  we  put  y+x s/a  =  t, 

we  obtain 

dx  _(ti  +  c)Ja  +  2bt  _  (t*  +  c)  Ja  +  2bt 

dt~      (tja  +  b?     '       y~    ~tja  +  b    "' 
and  so 

fdx       f     dt            1    .             ,               b   \ 
J  J  =  JT7a^b  =  7alOSr*/a  +  9J+;ra\    ^ 


/; 


134-136]  DERIVATIVES    AND    INTEGRALS  239 

If  in  particular  a  =  l,  6  =  0,  c=a2,  or  a=l,  b  =  0,  c=  -a2,  we  obtain 

equations  whose  truth  may  be  verified  immediately  by  differentiation.  With 
these  formulae  should  be  associated  the  third  formula 

v/(a2-^)  =  arCSin(-y/a)    <3)' 

which  corresponds  to  a  case  of  the  general  integral  of  this  section  in  which 
a  <  0.  In  (3)  it  is  supposed  that  a  > 0 ;  if  a  <  0  then  the  integral  is  arc  sin  (.v/ 1  a  j) 
(cf.  §  119).  In  practice  we  should  evaluate  the  general  integral  by  reducing  it 
(as  in  the  next  section)  to  one  or  other  of  these  standard  forms. 

The  formula  (3)  appears  very  different  from  the  formulae  (2) :  the  reader 
will  hardly  be  in  a  position  to  appreciate  the  connection  between  them  until 
he  has  read  Ch.  X. 

/*  X«£  -f-  Lt 

136.     The  integral      — — - — -£ dx.     This  integral  can 

J  *J(ax-  +  Ibx+c)  ° 

be  integrated  in  all  cases  by  means  of  the  results  of  the  preceding 
sections.     It  is  most  convenient  to  proceed  as  follows.     Since 

Xx  +  fj,  =  (X/a)  (ax  +  b)  +  /x  -  (Xb/a), 
dx  =  *J(ax-  +  2bx  +  c), 


I 


V(a#2  +  2bx  +  c) 
we  have 

t     (Xx  +  fx)dx         X  ,  ,  oi     ,     \  .  (        ^\  f  dx 

—rr £r "v  =  -  *J(ax-  +  Zbx  +  C)  +  [  a -,-, —^ r  . 

)  ^(ax*  +  2bx  +c)     a Y  v  ;      V         aJJ  s/(ax-  +  2bx+c) 

In  the  last  integral  a  may  be  positive  or  negative.     If  a  is 
positive  we  put  x  *Ja  +  (b/\Ja)  =  t,  when  we  obtain 

dt 


if. 

J  a  J  i 


y/a  J  i\/(t%  +  k)  ' 

where  k  =  (ac  —  b2)/a.     If  a  is  negative  we  write  A  for  —  a  and 
put  x\JA  —(b/\/A)  =  t,  when  we  obtain 

1        f         dt 
*J{-a)}  sj{-K-Vy 

It  thus  appears  that  in  any  case  the  calculation  of  the  integral 
may  be  made  to  depend  on  that  of  the  integral  considered  in 
§  135,  and  that  this  integral  may  be  reduced  to  one  or  other 
of  the  three  forms 

f       dt  f       dt  f       dt 

J  vV3  +  a2) '      1  V(«s  ~  a2) '      J  v^2  - 12) ' 


240  DERIVATIVES    AND    INTEGRALS  [VI 

137.  The  integral  l(\x  +  tJ.)*J(ax2  +  2bx  +  c)dx.  In  exactly  the  same 
way  we  find 

f(Xx + h)  <J(ax2  +  %>x  +  c) dx  =  (£\  (ax2  +  2bx  +  c)3'2 

+  (     '—)    J  (<**2 +2bx  +  c)  dx  ; 

and  the  last  integral  may  be  reduced  to  one  or  other  of  the  three  forms 

fj(P  +  a*)dt,      U(f-a*)dt,      fj(a2-P)dt. 

In  order  to  obtain  these  integrals  it  is  convenient  to  introduce  at  this  point 
another  general  theorem  in  integration. 

138.  Integration  by  parts.  The  theorem  of  integration  by 
parts  is  merely  another  way  of  stating  the  rule  for  the  differentia- 
tion of  a  product  proved  in  §  113.  It  follows  at  once  from 
Theorem  (3)  of  §  113  that 

jf  (0)  F(x)  dx  =f(x)  F  (x)  -  ff(x)  F'  (0)  dx. 

It  may  happen  that  the  function  which  we  wish  to  integrate  is 
expressible  in  the  form  f'(x)  F (x),  and  that  f  (x)  F'  (x)  can  be 
integrated.  Suppose,  for  example,  that  $  (x)  =  xty  (x),  where  yfr  (x) 
is  the  second  derivative  of  a  known  function  x  (x)-     Then 

J  <j>  (x)  dx=  J  xX"  (0)  dx  =  xX'  (0)  -jx  0)  dx  =  xX'  (0)  -  %  (0). 

We  can  illustrate  the  working  of  this  method  of  integration  by  applying 
it  to  the  integrals  of  the  last  section.     Taking 

/  (x)  =  ax +b,     F  (a?)  =  J  (ax2  +  2bx +c)=y, 
we  obtain 

a  h/dx=(a.v  +  b)y-  J  — '-  dx=(ax  +  b)y  -  a  J  ydx+(ac-b2)  /— , 

,,    ,  [    ,      (ax  +  b)y  ,  ac-b2  [dx 

so  that  J^=L__il+__J_. 

and  we  have  seen  already  (§  135)  how  to  determine  the  last  integral 
Examples  XLIX.     1      Prove  that  if  a>0  then 

J  slix*  +  a2) dx =\x  sjix1  +  a2)  +| a2  log  {x  +  v/(.r2  +  n2)} , 

J  v/(.f2  -  a2)  dx=hx  s!(x2  -  a2)  -  \  a2  log  |  x  +  sl(x2  -  a2)  | , 

I  s/(a2  —  x2)  dx = \x  *J(a2  -  x2) + 1  a2  arc  sin  (xja). 


137,  138]  DERIVATIVES    AND    INTEGRALS  241 

2.  Calculate  the  integrals    /  ~--2*_  /  v/(a2  _  xt}  jx  by  moans  of  the 

substitution  x=asm6,  and  verify  that  the  results  agree  with  those  obtained 
in  §  135  and  Ex.  1. 

3.  Calculate  /  x  (x + a)m dx,  where  m  is  any  rational  number,  in  three 

ways,  viz.  (i)  by  integration  by  parts,  (ii)  by  the  substitution  (x+a)m  =  t,  and 
(iii)  by  writing  (x+a)-a  for  x  ;  and  verify  that  the  results  agree. 

4.  Prove,  by  means  of  the  substitutions  ax  +  b=l/t  and  x=l/u,  that  (in 
the  notation  of  §§  130  and  138) 


[dx  _  ax  +  b  fxdx  _      b.r 

}f~     Ay     '       J'y1"      '~t 


+  G 

Ay 


f  dx 

5.  Calculate    I  -j- — ^-,  where  b>a,  in  three  ways,  viz.  (i)  by 

J  \!  ((X       a)  (U       X)) 

the  methods  of  the  preceding  sections,  (ii)  by  the  substitution  (b  -  x)/(x  —  a)  —  t2, 
and  (iii)  by  the  substitution  x=a  cos2d  +  b  sin2d ;  and  verify  that  the  results 
agree. 

6.  Integrate  s/{(x  -a)(b-  x)}  and  s/{(b  -  x)/(x  -  a)} . 

7.  Show,  by  means  of  the  substitution  2x  +  a  +  b  =  \  (a  —  b)  {t2  +  (l/t)2}, 
or  by  multiplying  numerator  and  denominator  by  J{x  +  a)  -  *J(x  +  b),  that  if 
a  >  b  then 

L+g)t/(^rK/(3"6)  ('+s?):- 

f  dx 

8.  Find  a  substitution  which  will   reduce    I  , cjs— -, c^s   to   the 

J  (x  +  a)6ii  +  (x  —  a)312 

integral  of  a  rational  function.  (Math.  Trip.  1899.) 

9.  Show  that  JR{x,  %/(ax  +  b)}dx  is  reduced,  by  the  substitution 
ax  +  b=yn,  to  the  integral  of  a  rational  function. 

10.  Prove  that 

ff"  {x)F(x)dx=f  (x)  F(x)-f(x)  F1.  (x)  +  ff(x)  F"(x)  dx 
and  generally 
f /(»)  (x)  F(x)  dx=f(n  ~  "  (x)  F{x)  -  /(»  -  2>  (x)  F'  (x)  +  ..:  +  (- 1 )"  f  f(x)  FW  (x)  dx. 

11.  The  integral  I  (1  +x)p  xP  dx,  where  p  and  q  are  rational,  can  be  found 

in  three  cases,  viz.  (i)  if  p  is  an  integer,  (ii)  if  q  is  an  integer,  and  (iii)  if 
p  +  q  is  an  integer.  [In  case  (i)  put  x=us,  where  s  is  the  denominator  of  q  ; 
in  case  (ii)  put  1  +x=ts,  where  s  is  the  denominator  of  p ;  and  in  case  (iii)  put 
1  +x=xt8,  where  s  is  the  denominator  of  p.] 

n.  16 


242  DERIVATIVES   AND   INTEGRALS  [VI 

12.  The  integral  /  xm(axn  +  b)"dx    can   be  reduced  to    the    preceding 

integral  by  the  substitution  axn=bt.  [In  practice  it  is  often  most  con- 
venient to  calculate  a  particular  integral  of  this  kind  by  a  'formula  of 
reduction'  (cf.  Misc.  Ex.  39).] 

13.  The  integral  I R  {x,  ^(ax  +  b),  sj(cx  +  d)}  dx  can  be  reduced  to  that  of 
a  rational  function  by  the  substitution 

4x=  -  (5/o)  {t+mr  -  (die)  {t-  (l/o}2. 

14.  Reduce  I R  (x,  y)  dx,  where  y2  (x  —  y)  =  x2,  to  the  integral  of  a  rational 
function.     [Putting  y=tx  we  obtain  x=l/{t2(l  —  t)},  y  =  l/{t(l  -t)}.] 

15.  Reduce  the  integral  in  the  same  way  when  (a)  y(x  —  y)2=x, 
(b)  (x2+y2)2=a2  (x2-y2).  [In  case  (a)  put  x  —  y  =  t:  in  case  (6)  put 
x2 +y2=t(x-  y),  when  we  obtain  x = a2t  (t2  +  a2)/(*4  +  a4),  y  =  a2t  (t2  -  a2)/(*4  +  a4).  ] 

f    dx 

16.  If  y  (x-y)2 =x  then  I  — —  =  J  log  {(x  —  y)2  —  1}. 

J  x—3y 

17.  If  {x2+y2)2=2c2  (x2-y2)  then  \—*L—~  =  -  -9log  (^±2*)  . 

a  '  *  J y(xi+y2  +  c2)  c2    °\x-yj 


139.      The  general  integral 


I  .ft  (a?,  y)  dx,  where  y2=ax2  +  2bx+c. 


The  most  general  integral,  of  the  type  considered  in  §  134,  and  associated  with 
the  special  conic  y2=ax2  +  2bx +c,  is 


/ 


R(x,*JX)dx  (1), 

where  X=y2—ax2  +  2bx  +  c.     We  suppose  that  R  is  a  real  function. 

The  subject  of  integration  is  of  the  form  P/Q,  where  P  and  Q  are  poly- 
nomials in  x  and  *JX.     It  may  therefore  be  reduced  to  the  form 

A  +  BJX_{A  +  BJX){C-DJX)_1Pl  „  /y 
C+DJX  C2  -  D2X  -&  +  !<  v a, 

where  A,  B, ...  are  rational  functions  of  x.  The  only  new  problem  which 
arises  is  that  of  the  integration  of  a  function  of  the  form  F  JX,  or,  what  is 
the  same  thing,  G/JX,  where  G  is  a  rational  function  of  x.     And  the  integral 

G 


I 


xdx  (2) 

can  always  be  evaluated  by  splitting  up  G  into  partial  fractions.     When  we 
do  this,  integrals  of  three  different  types  may  arise. 

(i)    In  the  first  place  there  may  be  integrals  of  the  type 


/ 


-p*  dx    (3), 


138,  139]  DERIVATIVES   AND   INTEGRALS  243 

where  m  is  a  positive  integer.  The  cases  in  which  m=Oorm=l  have  been 
disposed  of  in  §  136.  In  order  to  calculate  the  integrals  corresponding  to 
larger  values  of  m  we  observe  that 

where  a,  /3,  y  are  constants  whose  values  may  be  easily  calculated.  It  is  clear 
that,  when  we  integrate  this  equation,  we  obtain  a  relation  between  three 
successive  integrals  of  the  type  (3).  As  we  know  the  values  of  the  integral 
for  m  =  0  and  m  =  1,  we  can  calculate  in  turn  its  values  for  all  other  values  of  m. 


(ii)     In  the  second  place  there  may  be  integrals  of  the  type 

dx 
(x-p)m,JX  ^  )' 

where  p  is  real.     If  we  make  the  substitution  x  —  p  =  \jt  then  this  integral  is 
reduced  to  an  integral  in  t  of  the  type  (3). 


/ 


(iii)  Finally,  there  may  be  integrals  corresponding  to  complex  roots  of  the 
denominator  of  G.  We  shall  confine  ourselves  to  the  simplest  case,  that  in 
which  all  such  roots  are  simple  roots.  In  this  case  (cf.  §  130)  a  pair  of  con- 
jugate complex  roots  of  G  gives  rise  to  an  integral  of  the  type 

Lx+M 


Si 


dx (5). 


(Ax2  +  2Bx  +  C)  V  (ax2  +  2bx  +  c 
In  order  to  evaluate  this  integral  we  put 

_ftt+V 

X~t+\  ' 

where  p  and  v  are  so  chosen  that 

apv  +  b(jj.+v)+c=0,     A^v  +  B(fj.  +  v)  +  C=0; 

so  that  fi  and  v  are  the  roots  of  the  equation 

(aB-bA)£2-(cA-a(7)$  +  (bC-cB)  =  0. 

This  equation  has  certainly  real  roots,  for  it  is  the  same  equation  as 
equation  (1)  of  Ex.  xlvi.  12  ;  and  it  is  therefore  certainly  possible  to  find 
real  values  of  /x  and  v  fulfilling  our  requirements. 

It  will  be  found,  on  carrying  out  the  substitution,  that  the  integral  (5) 
assumes  the  form 

tdt „ /" dt_ 

(at2+®  J(yt*+8y     J  (at2 +(3)  J(yt2  +  8) 

The  second  of  these  integrals  is  rationalised  by  the  substitution 

t 


*k\wt-° 


which  srives 


/ 


J(yt'  +  8)    w' 
dt  f  die 


(at2  +  0)  J(yt2  +  8)      J  0  +  (ad  -  /Sy)  u2 


16- 


244  DERIVATIVES   AND   INTEGRALS  [VI 

Finally,  if  we  put  t=l/u  in  the  first  of  the  integrals  (6),  it  is  transformed  into 
an  integral  of  the  second  type,  and  may  therefore  be  calculated  in  the  manner 
just  explained,  viz.  by  putting  uU(y+8u2)  =  u,  i.e.  l//J(yt2  +  8)  =  v.* 

Examples  L.     1.     Evaluate 

f  dx  f  dx  f dx 

JxJ(a?+2x  +  3y    ]{x~-ijj(x*  +  l)'    ](x  +  l)<J(l  +  2x~^~xT)- 

2.  Prove  that 

f dx _     2  //x-g\ 

J  (x-P)<s/{(z-P)(x-Q)}  ~  2  ~P  V  \x-p)  ' 

3.  If  ag2  +  ch2  =  -  v  <  0  then 

( *L_     _  =  -JL  arc  tan  rVM^2  +  c)H  _ 

J  (Jix+g)  ,J(ax2  +  c)        Jv  [_     ch-agx    J 

f      dv 

4.  Show  that  I ; ;    ,  where  ?/2  =  a.v2  +  2bx  +  c,  may  be  expressed  in  one 

J(x-x0)y 


or  other  of  the  forms 

axx0  +  b(x+x0)+c  +  y ?/0 


1  (axx0  +  b(x  +  xn)  +  c 

-  arc  tan  ' 


log 

#o  oc-ooq  z0  (  yzQ 

according  as  ax02  +  2bxo  +  c  is  positive  and  equal  to  y0-  or  negative  and  equal 
to  —  z02. 

5.  Show  by  means  of  the  substitution  y  =  Sff(ax2  +  2bx+c)/(x-p)  that 

f  dx  _  f      dy 

J  (* - P)  V(«*2  +  2bx  +  c)~~  J  J(\y2 - fi) ' 
where  \  =  ap2  +  2bp  +  c,  /j.  =  ac-b2.      [This  method  of  reduction  is  elegant  but 
less  straightforward  than  that  explained  in  §  139.] 

6.  Show  that  the  integral 

dx 


I 


'  xj(3x2  +  2x+l) 

is  rationalised  by  the  substitution  x=(l+y2)/(3-y2).  (Math.  Trip.  1911.) 

7.     Calculate 

(x  +  l)dx 


I 


(x2  +  4)J(.e2  +  9)' 

*  The  method  of  integration  explained  here  fails  if  ajA  =  bjB;  but  then  the 
integral  may  be  reduced  by  the  substitution  ax  +  b  —  t.  For  further  information 
concerning  the  integration  of  algebraical  functions  see  Stolz,  Grundziige  der 
Differential-und-integralrechnung,  vol.  i,  pp.  331  et  seq.;  Bromwicb,  Elementary 
Integrals  (Bowes  and  Bowes,  1911).  An  alternative  method  of  reduction  has  been 
given  by  Sir  G.  GreenhiJl:  see  his  A  Chapter  in  the  Integral  Calculus,  pp.  12  et 
seq  ,  and  the  author's  tract  quoted  on  p.  236. 


139-141]  DERIVATIVES   AND   INTEGRALS  245 

8.     Calculate 


i 


dx 


(5x2  +  Ux  +  8)  st(5a?  +  2%  -  7)  * 

[Apply  the  method  of  §  139.  The  equation  satisfied  by  M  and  v  is 
£2  +  3£  +  2  =  0,  so  that  /x=-2,  i/=-l,  and  the  appropriate  substitution  is 
x=  -(2t+l)/(t  +  l).     This  reduces  the  integral  to 

_  f dt r tdt 

J  (4*3  + 1-)  N/(9*2  -  4)      J  (4f~  +  l)7(9^-~4)  * 

The  first  of  these  integrals  may  be  rationalised  by  putting  */V(9*2-4)  =  m  and 
the  second  by  putting  l/v/(9<2-  4)  =  v.] 

9.  Calculate 

f (*+l)«fc f (x-\)dx 

J  (2a,-2  - 2x+ 1)  N/(3o;2  -  2*+ 1).'     J  (2a,-2  -  Or  +  5)  sl(7x*  -  22.r  +19)* 

(JfeA.  Trip.  1911.) 

10.  Show  that  the  integral  I  ^  (.r,  y)  dx,  where  y2 = ax1  +  2bx + c,  is  ration- 
alised by  the  substitution  t  =  (x-p)j(y  +  q),  where  (p,  j)  is  any  point  on  the 
conic  y2=a.r2  +  26.r+c.  [The  integral  is  of  course  also  rationalised  by  the 
substitution  t  =  {x—p)l(y  —  q)  :  cf.  §  134.] 

140.  Transcendental  Functions.  Owing  to  the  immense 
variety  of  the  different  classes  of  transcendental  functions,  the 
theory  of  their  integration  is  a  good  deal  less  systematic  than 
that  of  the  integration  of  rational  or  algebraical  functions.  We 
shall  consider  in  order  a  few  classes  of  transcendental  functions 
whose  integrals  can  always  be  found. 

141.  Polynomials  in  cosines  and  sines  of  multiples  of  x. 

We  can  always  integrate  any  function  which  is  the  sum  of  a 
finite  number  of  terms  such  as 

A  cosm  ax  sinm  ax  cos1lbx  smn'bx..., 

where  m,  m,  n,  n,  ...  are  positive  integers  and  a,  b,  ...  any  real 
numbers  whatever.  For  such  a  term  can  be  expressed  as  the 
sum  of  a  finite  number  of  terms  of  the  types 

a  cos  {(pa  +  qb  +  ...)  x\,     /3sin  {(pa  +  qb  +  ...)x] 

and  the  integrals  of  these  terms  can  be  written  down  at  once. 


246  DERIVATIVES   AND    INTEGRALS  [VI 

Examples  LI.  1.  Integrate  sin3  x  cos2  2x.  In  this  case  we  use  the 
formulae 

sin3  x  =  \  (3  sin  x  -  sin  3x),     cos2  2x= |(1  +  cos  4a?). 

Multiplying  these  two  expressions  and  replacing  sin  x  cos  4x,  for  example, 
by  \  (sin  bx  —  sin  3d;),  we  obtain 

Jg  I  (7  sin  a;  -  5  sin  3a?  +  3  sin  bx  —  sin  7a?)  dx 

=  -  /g-  cos  x + -fg  cos  3a;  —  £%  c°s  5a;  +i^2  cos  7x. 

The  integral  may  of  course  be  obtained  in  different  forms  by  different 
methods.     For  example 

I  sin3  x  cos2  2xdx  —  I  (4  cos4  x  -  4  cos2  x+ 1)  (1  —  cos2  a?)  sin  xdx, 

which  reduces,  on  making  the  substitution  cosx=t,  to 

/  (4l6 -8ti  +  5t2-l)dt  =  f-  cos7 x -  f  cos5 a;  +  §  cos3 a; -  cos x. 

It  may  he  verified  that  this  expression  and  that  obtained  above  differ  only  by 
a  constant. 

2.  Integrate  by  any  method  cos  ax  cos  bx,  sin  ax  sin  bx,  cos  ax  sin  fea?, 
cos2.?;,  sin3.*;,  cos4 a?,  cos  x  cos  2x  cos  3a?,  cos3  2a?  sin2  3.r,  cos6  a;  sin7  a;.  [In  cases  of 
this  kind  it  is  sometimes  convenient  to  use  a  formula  of  reduction  (Misc. 
Ex.  39).] 

142.     The  integrals  I  x11  cos  x  dx,  J  xn  sin  x  dx  and  associated 

integrals.     The  method   of  integration   by  parts  enables  us  to 
generalise  the  preceding  results.     For 

I  xn  cos  x  dx  =     x11  sin  x  —  n  I  a;71-1  sin  x  dx, 

I  xn  sin  x  dx  =  —  xn  cos  x  +  n  I  a;'1-1  cos  a?  <fo, 

and    clearly    the    integrals    can   be    calculated    completely   by   a 
repetition  of  this  process  whenever  n  is  a  positive  integer.     It 


follows  that  we  can  always  calculate  I  xn  cos  ax  dx  and  I  x11  sin  axdx 

if  n  is  a  positive  integer;  and  so,  by  a  process  similar  to  that  of 
the  preceding  paragraph,  we  can  calculate 


/ 


P  (x,  cos  ax,  sin  ax,  cos  bx,  sin  bx,  . . .)  dx, 
where  P  is  any  polynomial. 


141-143]  DERIVATIVES    AND    INTEGRALS  247 

Examples  LII.  1.  Integrate  a;sin#,  a? coax,  ^cos2^,  x2  sin2 x sin2 2x, 
X  sin2  x  cos4  x,  x3  sin3  \x. 

2.  Find  polynomials  P  and  Q  such  that 

/ {(3.r- 1)  cos x+(l  - 2x)  sin  x)  dx=  P  cos x+Q  sin  x. 

3.  Prove  that  I  ^,l  cos  .zofo;=  Pw  cos  x  +  Qn  sin  x,  where 

P,i  =  n.i-»-1-?i(«-l)(»-2)^»-3+...,     &«*»-ri(n-l)  «»-«+.... 

143.  Rational  Functions  of  cos  x  and  sin  a*.  The  integral 
of  any  rational  function  of  cos  x  and  sin  x  may  be  calculated  by 
the  substitution  tan  \x  =  t.     For 

1  - 12      .  It        dx         2 

cos#=- -.  sina;= ,    — -  = , 

l+t2'  l+t2'    dt      l  +  t*' 

so  that  the  substitution  reduces  the  integral  to  that  of  a  rational 
function  of  t. 

Examples  LIII.     1.     Prove  that 

I sec xdx = log  |  sec  x + tan  x\,    J  cosec x dx = log  |  tan  hx |. 

[Another  form  of  the  first  integral  is  log  |  tan  {\^+\x)\'t  a  third  form  is 
h  log  |  (1  +  sin  x)j(  1  -  sin  x)  | .] 

2.  I  tan  xdx  =  -log  |  cos  x  |,    /  cot . v  dx= log  |  sin  x\,    /sec2  xdx=t&n  x, 

J  cosec2  a'  dx  =  —  cot  #,       /  tan  #  sec  .r  dx = sec  a?,      I  cot  x  cosec  #  efcc =  —  cosec .  v. 

[These  integrals  are  included  in  the  general  form,  but  there  is  no  need  to 
Tise  a  substitution,  as  the  results  follow  at  once  from  §  119  and  equation  (5) 
of  §  130.] 

3.  Show  that  the  integral  of  l/(a  +  bcosx),  where  a  +  b  is  positive,  may 
be  expressed  in  one  or  other  of  the  forms 

J(b  +  a)  +  tJ(b-a) 


tanf\/(^)}'      ^2W)l0g 


J{b  +  a)-ts!{b-a) 


where  <  =  tan^.r,  according  as  a2>b2  or  a2  <  b2.  If  a2  =  ft2 then  the  integral 
reduces  to  a  constant  multiple  of  that  of  sec2  ix  or  cosec2  J.r,  and  its  value 
may  at  once  be  written  down.  Deduce  the  forms  of  the  integral  when  a  +  b 
is  negative. 

4.     Show  that  if  y  is  defined  in  terms  of  x  by  means  of  the  equation 
{a  +  b  cos  x)  (a-b  cos  y)  =  a2  —  b2, 
where  a  is  positive  and  a?>b2,  then  as  x  varies  from  0  to  n  one  value  of  y 
also  varies  from  0  to  n.     Show  also  that 

sj{a2  -  b2)  sin  y  sin  x     dx BJny 


sino.'= 


a-bcosy     '     a  +  bcosx  dy     a—bcosy' 


248  DERIVATIVES   AND    INTEGRALS  [VI 

and  deduce  that  if  0  <  x  <  n  then 


I 


dx  1  (a  cos  x  +  b 

arc  cos 


a  +  b  cos  x     V  (a  ~  "  )  \a  +  b  cos  .r/ 

Show  that  this  result  agrees  with  that  of  Ex.  3. 

5.  Show  how  to  integrate  1  j(a  +  b  cos  x  +  c  sin  x).  [E xpress  b  cos  x  +  c  sin  # 
in  the  form  *J(b2  +  c2)  cos  (x  —  a).] 

6.  Integrate  (a  +  6  cos  x  +  c  sin  x)/(a+j3  cos  #  +  7  sin  .r) 
[Determine  X,  ju,  i>  so  that 

a  +  6  cos  #  +  c  sin  x= X +/x  (a +/3  cos  #  +  7  sin  a;)  +  v  ( —  /3  sin  x+y  cos  #). 

Then  the  integral  is 

11  -in/"  <^-r  -1 

u^  +  i'  log  a  4-/3  cos  .r  4  7  sin  #  +\  I  — ■ : —  .1 

r  °  '        '  '  '        J  a  4  /3  cos  a?  +  7  sin  X  J 

7.  Integrate  l/(acos2a7  +  26cos^sin.r  +  c  sin2.^).  [The  subject  of  inte- 
gration may  be  expressed  in  the  form  l/(A+Bco$2x  +  Csm2x),  where 
A=\(a+c\  B=\(a  —  c),  C=b  :  but  the  integral  may  be  calculated  more 
simply  by  putting  tnnx=t,  when  we  obtain 

f  sec2  x  dx  _  f        dt  , 

J  a  +  2bt&nx  +  cta,ri*x  ~  J  a  +  2bt  +  ct2  '* 

144.     Integrals  involving  arc  sin  x,  arc  tan  x,  and  log  x.     The 

integrals  of  the  inverse  sine  and  tangent  and  of  the  logarithm  can 
easily  be  calculated  by  integration  by  parts.     Thus 

f  C      oc  dx 

I  arc  sin  xdx  —  x  arc  sin  x  —  I    77^ -.  =  x  arc  sin  x  +  J  (I  —  x2). 

J  J^(l-x-) 

f  f  xdx 

I  arc  tan  xdx  =  x  arc  tan  x  —  I  — n  =  x  arc  tan  x  —  \  log  (1  +  x2), 

J  J    -I  4"  X" 

I  log  xdx  =  x  log  x  —  I  dx  =  x  (log x  —  1). 

It  is  easy  to  see  that  if  we  can  find  the  integral  of  y  —f{x) 
then  we  can  always  find  that  of  x  =  <j>  (y),  where  <f>  is  the  function 
inverse  to/     For  on  making  the  substitution  y  =f(x)  we  obtain 

Jcf>(y)dy=j  xf  (x)  dx  =  xf{x)  -  jf(x)  dx. 

The  reader  should  evaluate  the  integrals  of  arc  sin  y  and  arc  tan  y 
in  this  way. 

Integrals  of  the  form 

I  P  (x,  arc  sin  x)  dx,     I  P  (x,  log  x)  dx, 


143-145] 


DERIVATIVES    AND    INTEGRALS 


249 


where  P  is  a  polynomial,  can  always  be  calculated.     Take  the 
first  form,  for  example.    We  have  to  calculate  a  number  of  integrals 

of  the  type  I  xm  (arc  sin  x)n  dx.     Making  the  substitution  as  =  sin  y, 


we  obtain  yn  sinm  y  cos  ydy,  which  can  be  found  by  the  method  of 
§  142.  In  the  case  of  the  second  form  we  have  to  calculate  a  number 
of  integrals  of  the  type  I  x'n  (log  x)n  dx.  Integrating  by  parts  we 
obtain 

I  xm  (log  x)n  dx  =  afm+1(loSx)n  _  _  JL_  j  xm  (log  xy-i  dx> 

and  it  is  evident  that  by  repeating  this  process  often  enough  we 
shall  always  arrive  finally  at  the  complete  value  of  the  integral. 

145.  Areas  of  plane  curves.  One  of  the  most  important 
applications  of  the  processes  of  integration  which  have  been 
explained  in  the  preceding  sections  is  to  the  calculation  of  areas 
of  plane  curves.  Suppose  that  P^PP'  (Fig.  44)  is  the  graph  of 
a  continuous  curve  y  =  </>  (x)  which  lies  wholly  above  the  axis  of  x, 
P  being  the  point  (x,  y)  and  P'  the  point  (x  +  h,  y  +  k),  and  h  being 
either  positive  or  negative  (positive  in  the  figure). 

P' 


N, 


N         N' 
Fig.  44. 


The  reader  is  of  course  familiar  with  the  idea  of  an  'area',  and 
in  particular  with  that  of  an  area  such  as  ONPP0.  This  idea  we 
shall  at  present  take  for  granted.  It  is  indeed  one  which  needs 
and  has  received  the  most  careful  mathematical  analysis :  later  on 
we  shall  return  to  it  and  explain  precisely  what  is  meant  by 


250  DERIVATIVES   AND   INTEGRALS  [VI 

ascribing  an  '  area'  to  such  a  region  of  space  as  ONPP0.  For  the 
present  we  shall  simply  assume  that  any  such  region  has  associated 
with  it  a  definite  positive  number  (0NPP0)  which  we  call  its 
area,  and  that  these  areas  possess  the  obvious  properties  indicated 
by  common  sense,  e.g.  that 

{PRP')  +  (NN'RP)  =  (NN'P'P),    (N,  NPP, )  <  ( ONPP0), 
and  so  on. 

Taking  all  this  for  granted  it  is  obvious  that  the  area  0NPP0 
is  a  function  of  x ;  we  denote  it  by  <&  (x).  Also  O  (x)  is  a 
continuous  function.     For 

®(x  +  h)-<P  (x)  =  (NN'P'P) 

=  (NN'RP)  +  (PRP')  =  h<f>  (x)  +  (PRP'). 

As  the  figure  is  drawn,  the  area  PRP'  is  less  than  hk.  This  is 
not  however  necessarily  true  in  general,  because  it  is  not  neces- 
sarily the  case  (see  for  example  Fig.  44  a)  that  the  arc  PP' 
should  rise  or  fall  steadily  from  P  to  P'.  But  the  area  PRP' 
is  always  less  than  \h\\  (h),  where  \  (h)  is  the  greatest  distance  of 
any  point  of  the  arc  PP'  from  PR.  Moreover,  since  <£  (x)  is  a 
continuous  function,  \(h)-^-0\s  h-^~0.    Thus  we  have 

3>  (X  +  h)-&  (X)  =  h  {(f)  (X)  +  fi  (h)}, 

where  \fi  (h)\  <  X  (h)  and  X  (h)  -*0  as  h  -*0.  From  this  it  follows 
at  once  that  <3>  (x)  is  continuous.     Moreover 

<J>  (x)  =  hm  — ' —  =  hm  {cf>(x)  +  fi  (h)}  =  <£  (x). 

Thus  the  ordinate  of  the  curve  is  the  derivative  of  the  area,  and  the 
area  is  the  integral  of  the  ordinate. 

We  are  thus  able  to  formulate  a  rule  for  determining  the 
area  ONPP0.  Calculate  <&(x),the  integral  of<f>(x).  This  involves 
an  arbitrary  constant,  which  we  suppose  so  chosen  that  <I>  (0)  =  0. 
Then  the  area  required  is  <t>  (x). 

If  it  were  the  area  N1NPP1  which  was  wanted,  we  should  of  course  deter- 
mine the  constant  so  that  *  (#,)  =  0,  where  xx  is  the  abscissa  of  Px.  If  the 
curve  lay  below  the  axis  of  x,  *  (x)  would  be  negative,  and  the  area  would  be 
the  .absolute  value  of  *  (x). 


145,  146]  DERIVATIVES    AND    INTEGRALS  251 

146.  Lengths  of  plane  curves.  The  notion  of  the  length 
of  a  curve,  other  than  a  straight  line,  is  in  reality  a  more  difficult 
one  even  than  that  of  an  area.  In  fact  the  assumption  that  P0P 
(Fig.  44)  has  a  definite  length,  which  we  may  denote  by  S(x), 
does  not  suffice  for  our  purposes,  as  did  the  corresponding  as- 
sumption about  areas.  We  cannot  even  prove  that  8  (x)  is  con- 
tinuous, i.e.  that  lim  {S  (P')  -  S  (P)}  =  0.  This  looks  obvious 
enough  in  the  larger  figure,  but  less  so  in  such  a  case  as  is  shown 
in  the  smaller  figure.  Indeed  it  is  not  possible  to  proceed  further, 
with  any  degree  of  rigour,  without  a  careful  analysis  of  precisely 
what  is  meant  by  the  length  of  a  curve. 

It  is  however  easy  to  see  what  the  formula  must  be.  Let 
us  suppose  that  the  curve  has  a  tangent  whose  direction  varies 
continuously,  so  that  <$>'  (x)  is  continuous.  Then  the  assumption 
that  the  curve  has  a  length  leads  to  the  equation 

{S(x+h)-S(x)}/h  =  {PP'}lh  =  (PP'lh)  x  ({PP'}/PP'), 

where  {PP'\  is  the  arc  whose  chord  is  PF".     Now 

PP'  =  V(PP*  +  RF*)  =  h  ^/(l  + 1) , 

and  &  =  (/>(#  +  /i)  —  </>  {x)  =  h<f)'  (£), 

where  £  lies  between  x  and  x  +  h.     Hence 

lim  (PP'jh)  =  lim  V{1  +  [</>' (£)?}  =  V{  1  +  W  (*)?}. 
If  also  we  assume  that 

lim{PP'}/PP'  =  l, 
we  obtain  the  result 

S'  (x)  =  lim  {S  (x  +  h)-S  (x)}/h  =  V{1  +  W  O)]2} 

and  so  S(.v)  =  j^{l  +  [(f>'(x)J}dx. 

Examples  LIV.  1.  Calculate  the  area  of  the  segment  exit  off  from  the 
parabola  y  =  x~\A.a  by  the  ordinate  #=|,  and  the  length  of  the  arc  which 
bounds  it. 

2.  Answer  the  same  questions  for  the  curve  ay2=z3,  showing  that  the 
length  of  the  arc  is 

3.  Calculate  the  areas  and  lengths  of  the  circles  xi+y-  =  a1,  x~+y2=2ax 
by  means  of  the  formulae  of  §§  145 — 146. 


252  DERIVATIVES    AND    INTEGRALS  [\'I 

4.  Show  that  the  area  of  the  ellipse  (#2/a2)  4-  (j/2/62)  =  1  is  nab. 

5.  Find  the  area  bounded  by  the  curve  y=sinx  and  the  segment  of  the 
axis  of  x  from  x=0  to  x=<2nr.  [Here  *  (x)  =  -  cos x,  and  the  difference 
between  the  values  of  -  cos  x  for  x=0  and  #=27r  is  zero.  The  explanation  of 
this  is  of  course  that  between  x=tt  and  ^=27r  the  curve  lies  below  the  axis 
of  x,  and  so  the  corresponding  part  of  the  area  is  counted  negative  in  applying 
the  method.  The  area  from  x=0  to  x=ir  is  -costt+cos0=2;  and  the 
whole  area  required,  when  every  part  is  counted  positive,  is  twice  this, 
i.e.  is  4.] 

6.  Suppose  that  the  coordinates  of  any  point  on  a  curve  are  expressed 
as  functions  of  a  parameter  t  by  equations  of  the  type  x=cj)(t),  y=^r{t), 
<j)  and  \\r  being  functions  of  t  with  continuous  derivatives.  Prove  that 
if  x  steadily  increases  as  t  varies  from  t0  to  tit  then  the  area  of  the  region 
bounded  by  the  corresponding  portion  of  the  curve,  the  axis  of  x,  and  the  two 
ordinates  corresponding  to  t0  and  ti,  is,  apart  from  sign,  A  (t{)  —  A  (t0),  where 


A(t)  =  j^{t)<t>'{t)dt=jyd£tdt. 


7.  Suppose  that  C  is  a  closed  curve  formed  of  a  single  loop  and  not 
met  by  any  parallel  to  either  axis  in  more  than  two  points.  And  suppose 
that  the  coordinates  of  any  point  P  on  the  curve  can  be  expressed  as  in  Ex.  6 
in  terms  of  t,  and  that,  as  t  varies  from  t0  to  tx-,  P  moves  in  the  same 
direction  round  the  curve  and  returns  after  a  single  circuit  to  its  original 
position.  Show  that  the  area  of  the  loop  is  equal  to  the  difference  of  the 
initial  and  final  values  of  any  one  of  the  integrals 

this  difference  being  of  course  taken  positively. 

8.  Apply  the  result  of  Ex.  7  to  determine  the  areas  of  the  curves 
given  by 

...     x      1—  t2      y        2t  ....  „  ,    .   „ 

9.  Find  the  area  of  the  loop  of  the  curve  x3+y3=3axy.  [Putting 
y  =  tx  we  obtain  x=Satj(l  +t3),  y  =  3at2/(l+t3).  As  t  varies  from  0  towards 
oo  the  loop  is  described  once.     Also 

.  [(    dx        dy\  .  f  9dfy\.  .  f  9a?t*     ,t         3«2 

which  tends  to  0  as  t^~x> .     Thus  the  area  of  the  loop  is  ::a2.] 

10.  Find  the  area  of  the  loop  of  the  curve  x5+yf'  =  5ax2y2. 

11.  Prove  that  the  area  of  a  loop  of  the  curve  x=asu\2t,  y  =  asint  is 
fa2.  {Math.  Trip.  1908.) 


146]  DERIVATIVES    AND    INTEGRALS  253 

12.-    The  arc  of  the  ellipse  given  by  a>=acost,  g  =  h>imt,  between  the 
points  t  =  ti  and  t  =  t2,  is  F(t2)  —  F(ti),  where 


F(t)  =  a  fj(l-e2  ski2  t)dt, 


e  being  the   eccentricity.      [This  integral  cannot  however  be  evaluated  in 
terms  of  such  functions  as  are  at  present  at  our  disposal.] 

13.     Polar   coordinates.      Show  that  the  area  bounded  by  the  curve 
r=f(8),  where  f(8)  is  a  one-valued  function  of  8,  and  the  radii  8  =  8X,  8  —  82,  is 

F{82)-F(8l\  where  F(8)  =  ±lr2d8.     And  the  length  of  the  corresponding 

arc  of  the  curve  is  <!>  (82)  -  $  (8t),  where 


^A/WITH- 


Hence  determine   (i)  the  area  and  perimeter  of  the  circle  r=2ask\8; 

(ii)  the  area  between  the  parabola  r=^lsec2^8  and  its  latus  rectum,  and  the 

length  of  the  corresponding  arc  of  the  parabola ;  (iii)  the  area  of  the  liniacon 

r  =  a  +  bcos8,   distinguishing  the    cases  in   which    a>b,  a=b,   and  a<b ; 

and  (iv)  the  areas  of  the  ellipses  l/7-2=acos20  +  2Acos0sin0  +  &sin2#  and 

f        d8 

llr=\  +  ecosd.     [In  the  last  case  we  are  led  to  the  integral  I- — , , 

'  L  °      J(l+ecos0)2 

which  may  be  calculated  (cf.  Ex.  liii.  4)  by  the  help  of  the  substitution 
(l+ecos<9)(l-ecos$)  =  l-e2.] 

14.  Trace  the  curve  2#=(a/r) +  (?•/«)>  and  show  that  the  area  bounded 
by  the  radius  vector  8  =  /3,  and  the  two  branches  which  touch  at  the  point 
r  =  a,  8=1,  is  %  a2  (£2  -  1  )3'2.  {Math.  Trip.  1900.) 

15.  A  curve  is  given  by  an  equation  p  =  f(r),  r  being  the  radius  vector 
and  p  the  perpendicular  from  the  origin  on  to  the  tangent.  Show  that  the 
calculation  of  the  area  of  the  region  bounded  by  an  arc  of  the  curve  and  two 


radii  vectores  depends  upon  thai/  of  the  integral  ^  I  —~ — 


P2) 


MISCELLANEOUS    EXAMPLES   ON    CHAPTER  VI. 

1.  A  function  f{x)  is  defined  as  being  equal  to  1  -\-x  when  .r^O,  to  x  when 
0<#<1,  to  2  —  x  when  \S.x?L%  and  to  3x  —  x2  when  x>2.  Discuss  the 
continuity  of  f{x)  and  the  existence  and  continuity  of  f'(x)  for  x=0,  x=l, 
and  x  =  2.  (Math.  Trip.  1908.) 

2.  Denoting  a,  ax  +  b,  ax2  +  2bx  +  c,  ...  by  u0,  u:,  u2,  ...,  show  that 
u02u3  —  <Su()u1U2  +  2ul3  and  u0  ;<4  —  4w: u3  +  3«22  are  independent  of  x. 


254  DERIVATIVES    AND    INTEGRALS  [VI 

3.  If  a0,  au  ...,  a2n  are  constants  and  Ur=(au,  au  ...,  ar\x,  l)r,  then 

V0U*~2nU1U*-l+  2nf*-V  U2U,„_2-...+  UinU0 

is  independent  of  x.  {Math.  Trip.  1896.) 

[Differentiate  and  use  the  relation  Ur'  =  rUr^i.] 

4.  The  first  three  derivatives  of  the  function  arc  sin  (ji  sin  x)  -  sn,  where 
H>1,  are  positive  when  0£.x^Itt. 

5.  The  constituents  of  a  determinant  are  functions  of  x.  Show  that  its 
differential  coefficient  is  the  sum  of  the  determinants  formed  by  differentiating 
the  constituents  of  one  row  only,  leaving  the  rest  unaltered. 

6.  If  fiffiffs,  A  are  polynomials  of  degree  not  greater  than  4,  then 

A      A      A      A 
A'     fi      A'     A' 

A"   A"   A"   A" 
A"  A'"  A'"  A'" 

is  also  a  polynomial  of  degree  not  greater  than  4.     [Differentiate  five  times, 
using  the  result  of  Ex.  5,  and  rejecting  vanishing  determinants.] 

7.  liy3  +  3yx  +  2x3  =  0  then  x^il+x^y" -%xy' +y=0.    (Math.  Trip.  1903.) 

8.  Verify  that  the  differential  equation  y  =  <f>{^(yi)}  +  4>{:v-yP,(j/\yh 
where  yx  is  the  derivative  of  y,  and  ^  is  the  function  inverse  to  cj)',  is 
satisfied  by  y = (j>  (c)  +  <f>  (x  -  c)  or  by  y  =  2<f>(%x). 

9.  Verify  that  the  differential  equation  y  =  {xj-^r  {yx)}  <f>  {\p-  (y{)},  where  the 
notation  is  the  same  as  that  of  Ex.  8,  is  satisfied  by  y  =  ccf)(xlc)  or  by  y=fix, 
where  /3=0(a)/a  and  a  is  any  root  of  the  equation  4>(a)-  a$'(a)  =  0. 

10.  If  ax  +  by  +  c=0  then  y^=0  (suffixes  denoting  differentiations  with 
respect  to  x).  We  may  express  this  by  saying  that  the  general  differential 
equation  of  all  straight  lines  is  yo  =  0.  Find  the  general  differential  equations 
of  (i)  all  circles  with  their  centres  on  the  axis  of  x,  (ii)  all  parabolas  with 
their  axes  along  the  axis  of  x,  (iii)  all  parabolas  with  their  axes  parallel  to 
the  axis  of  y,  (iv)  all  circles,  (v)  all  parabolas,  (vi)  all  conies. 

[The  equations  are  (i)  \+y*+yy2=0,  (ii)  yx2+yy2=0,  (iii)  y3  =  0, 
(iv)  (l+#i2)j/3=3#iy22,  (v)  5?/32  =  3y2y4,  (vi)  ^y22i/6-^5y2y3yi  +  i0y.i3  =  0. 
In  each  case  we  have  only  to  write  down  the  general  equation  of  the  curves 
in  question,  and  differentiate  until  we  have  enough  equations  to  eliminate  all 
the  arbitrary  constants.] 

11.  Show  that  the  general  differential  equations  of  all  parabolas  and  of 
all  conies  are  respectively 

£*2(3/2-2;3)  =  0,     A;3(y2-2<'3)  =  0. 


DERIVATIVES   AND    INTEGRALS  255 

[The  equation  of  a  conic  may  be  put  in  the  form 

y=ax  +  b±sf(px2  +  2qx  +  r). 
From  this  we  deduce 

y%~  ±  (pr  -  q2)/(p.v2  +  2qx  +  r)3'2. 
If  the  conic  is  a  parabola  then  p  =  0.] 

._      _.       ,.        dy      1   d2y       1    d3y       1    diu  , 

12.  Denoting  ^,    -  ^,    -  -Jf„    -  -£,   ...    by  t,  a,  b,  c,  ...    and 

dx      1  d2x      1   oRr      1  d*x 

dy>    2!  o^ '    3!  dy3'    4!  %5'  '"  by  T'  a'  &  *  "'  show  that 

Aac  -  5&2  =  (4ay  -  5/32)/r8,     bt-a2=-  (/3r  -  a2)/r6. 

Establish  similar  formulae  for  the  functions  a2d-3abc  —  2b3,  (l+t2)b  —  2a2t 
2ct  -  bob. 

13.  Prove  that,  if  yk  is  the  £th  derivative  of  y=sin  (n  are  sin  x),  then 

(1  -x2)yk  +  2-  (2k+ 1)  xyk  + ,  +  (?t2-  k2)yk=Q. 
[Prove  first  when  £=0,  and  differentiate  k  times  by  Leibniz'  Theorem.] 

14.  Prove  the  formula 

vDx»  u = zy  («v)  -  »zy  - »  (mz>^) + w(f  ~1}  zy  -  -  ( nzy »)  - ■ 

where  »  is  any  positive  integer.     [Use  the  method  of  induction.] 

15.  A  curve  is  given  by 

,r=a(2cosi  +  cos2£),     y  =  a  (2  sin  t~sh\2t). 

Prove  (i)  that  the  equations  of  the  tangent  and  normal,  at  the  point  P 
whose  parameter  is  t,  are 

xsin^t+ycos^t=asm^t,  xcos^t-y  sin|£=3aco,s  :}t ; 
(ii)  that  the  tangent  at  P  meets  the  curve  in  the  points  Q,  R  whose  para- 
meters are  —  \t  and  ir  —  \t;  (iii)  that  QR  =  4a;  (iv)  that  the  tangents  at  Q 
and  R  are  at  right  angles  and  intersect  on  the  circle  x2+y2  =  a2;  (v)  that  the 
normals  at  P,  Q,  and  R  are  concurrent  and  intersect  on  the  circle  x2+y2  =  9a2; 
(vi)  that  the  equation  of  the  curve  is 

(x2 +3/2+l  lax + 9a2)2 = 4a  (2x  +  3a)3. 
Sketch  the  form  of  the  curve. 

16.  Show  that  the  equations  which  define  the  curve  of  Ex.  15  may 
be  replaced  by  £/a  =  2w  +  (l/«2),  rjja  =  (2/w)  +  u2,  where  £=x+yi,  rj=x-yi, 
«=Cis  t.     Show  that  the  tangent  and  normal,  at  the  point  defined  by  u,  are 

u2£  -ur\  =  a  (u3  - 1),     u2£  +  urj  =  2a  (u3  + 1 ), 

and  deduce  the  properties  (ii)— (v)  of  Ex.  15. 

17.  Show  that  the  condition  that  xi+4px3-4qx-l=0  should  have 
equal  roots  may  be  expressed  in  the  form  (p  +  q)2l3-(p-qfl3=l: 

(Math.  Trip.  1898.) 


256  DERIVATIVES   AND   INTEGRALS  [VI 

18.  The  roots  of  a  cubic  f(x)  =  0  are  a,  /3,  y  in  ascending  order  of  magni- 
tude. Show  that  if  (a,  /3)  and  (/3,  y)  are  each  divided  into  six  equal  sub-intervals, 
then  a  root  of  f'(x)  =  0  will  fall  in  the  fourth  interval  from  /3  on  each  side. 
What  will  be  the  nature  of  the  cubic  in  the  two  cases  when  a  root  of/'  (x)  =  0 
falls  at  a  point  of  division?  (Math.  Trip.  1907.) 

19.  Investigate  the  maxima  and  minima  of  f(x),  and  the  real  roots  of 
f(x)  =  0,  f(x)  being  either  of  the  functions 

x  —  sin  x  —  tan  a  (1  —  cos  x),     x  —  sin  x  —  (a  —  sin  a)  —  tan  \a  (cos  a  -  cos  x), 

and  a  an  angle  between  0  and  tt.     Show  that  in  the  first  case  the  condition  for 
a  double  root  is  that  tan  a  — a  should  be  a  multiple  of  w. 

20.  Show  that  by  choice  of  the  ratio  X  :  /j.  we  can  make  the  roots  of 
\  (ax2  +  bx+c) + /x  (a'  x2  +  b'x  +  c')=0  real  and  having  a  difference  of  any  mag- 
nitude, unless  the  roots  of  the  two  quadratics  are  all  real  and  interlace ;  and 
that  in  the  excepted  case  the  roots  are  always  real,  but  there  is  a  lower  limit 
for  the  magnitude  of  their  difference.  (Math.  Trip.  1895.) 

[Consider  the  form  of  the  graph  of  the  function  (ax2  +  bx+c)J(a'x2  +  b'x  +  c'): 
cf.  Exs.  xlvi.  12  et  seq.] 

t-.  ,ii  sin  irx 

21.  Prove  that  n  <— —     <4 

x(l  —  x) 

when  0  <  x  <  1,  and  draw  the  graph  of  the  function. 

22.  Draw  the  graph  of  the  function 

1  l 

7T  COt  TTX -  . 

X      x-l 

23.  Sketch  the  general  form  of  the  graph  of  y,  given  that 

*  t«*+.-i)C.-iyc+iy     (M^n^uoii 

ax  xi  r  ' 

24.  A  sheet  of  paper  is  folded  over  so  that  one  corner  just  reaches  the 
opposite  side.  Show  how  the  paper  must  be  folded  to  make  the  length  of  the 
crease  a  maximum. 

25.  The  greatest  acute  angle  at  which  the  ellipse  (x2ja2)  +  (y2/^2)  =  1  can 
be  cut  by  a  concentric  circle  is  arc  tan  {(a2  —  b-)j-2nb}.  (Math.  Trip.  1900.) 

26.  In  a  triangle  the  area  A  and  the  semi-perimeter  s  are  fixed.  Show  that 
any  maximum  or  minimum  of  one  of  the  sides  is  a  root  of  the  equation 
s(x  —  s)x2  +  4A2  =  0.  Discuss  the  reality  of  the  roots  of  this  equation,  and 
whether  they  correspond  to  maxima  or  minima. 

[The  equations  a  +  b  +  c  =  2s,  s(s-  a)(s-b)  (s— c)  =  A2  determine  a  and  b 
as  functions  of  c.  Differentiate  with  respect  to  c,  and  suppose  that  da/dc  =  0. 
It  will  be  found    that    b = c,  s-b  =  s-c  =  ^a,  from   which   we   deduce   that 

«(a-s)«2  +  4A2  =  0. 


DERIVATIVES    AND   INTEGRALS  2o7 

This  equation  has  three  real  roots  if  s4>27a2,  and  one  in  the  contrary 
case.  In  an  equilateral  triangle  (the  triangle  of  minimum  perimeter  for  a 
given  area)  s4  =  27A2;  thus  it  is  impossible  that  s4<27A2.  Hence  the 
equation  in  a  has  three  real  roots,  and,  since  their  sum  is  positive  and  their 
product  negative,  two  roots  are  positive  and  the  third  negative.  Of  the  two 
positive  roots  one  corresponds  to  a  maximum  and  one  to  a  minimum.] 

27.  The  area  of  the  greatest  equilateral  triangle  which  can  be  drawn 
with  its  sides  passing  through  three  given  points  A,  B,  C  is 

a2  +  b2  +  c2 

a,  6,  c  being  the  sides  and  A  the  area  of  A  BC.  {Math.  Trip.  1809.) 

28.  If  A,  A'  are  the  areas  of  the  two  maximum  isosceles  triangles  which 
can  be  described  with  their  vertices  at  the  origin  and  their  base  angles  on  the 
cardioid  r=a  (1  -f-cos  6),  then  256AA'  =  25a4 y/5.  (Math.  Trip.  1907.) 

29.  Find  the  limiting  values  which  (x2  -  4y  +  8)/(y2  -  6.r  +  3)  approaches 
as  the  point  (x,  y)  on  the  curve  x2y  —  4r2  —  4xy+y2  +  16x  —  2y—  7  =  0  ap- 
proaches the  position  (2,  3).  (Math.  Trip.  1903.) 

[If  we  take  (2,  3)  as  a  new  origin,  the  equation  of  the  curve  becomes 
£21  —  £2+q2  =  0>  and  the  function  given  becomes  (£2  +  4£-  4i;)/(jj2  +  6j7  -6£).  If 
we  put  Tj  =  tg,  we  obtain  £  =  (1  —  t2)jt,  r)  =  l  —  t2.  The  curve  has  a  loop  branching 
at  the  origin,  which  corresponds  to  the  two  values  t=  —  1  and  t=l.  Expressing 
the  given  function  in  terms  of  t,  and  making  t  tend  to  •  1  or  1,  we  obtain  the 
limiting  values  -  f,  -  §.] 

30.  If  /(*)  =  -■ —. . \ , 

sin  x  —  sin  a       (x  —  a)  cos  a 

then  —  {lim  f(x)}—  lim  /'  (x)  =  f  sec3  a  —  TSo  sec  a. 

da,  x-*-a  x-*-a 

(Math.  Trip.  1896.) 

31.  Show  that  if  0  (a?)  =  l/(l  +x2)  then  0<">  (x)  =  Qn  (x)j(l+x2)n  +  \  where 
Qn(x)  is  a  polynomial  of  degree  n.     Show  also  that 

(i)      Qn  +  i  =  (l+x2)Qn'-2(n  +  l)xQn, 

(ii)     Qn  +  2  +  ^(n  +  2)xQ)l  +  1  +  (n  +  2)  (n+l)  (l+x2)  Qn  =  0, 

(iii)     (l+x2)Qn"-2nxQn'  +  ?i(n  +  l)Qn=0, 

(iv)     ^  =  (-l)»n!J(n  +  l)^-(?l  +  1)3W,(?l~1)^-2+..j, 

(v)     all  the  roots  of  Qn  =  0  are  real  and  separated  by  those  of  0„_i  =  O. 

32.  If  f(x),  <fr  (x),  >Jr  (x)  have  derivatives  when  a  ^  x  £.b,  then  there  is 
a  value  of  £  lying  between  a  and  b  and  such  that 

f(a)         <j>(a)         +(a)     =0. 
f(b)         </>(&)         +(b) 

/'(£)      *'(£)       *'(£) 
h.  17 


258  DERIVATIVES   AND    INTEGRALS  [VI 

[Consider  the  function  formed  by  replacing  the  constituents  of  the  third 
row  hyf(x),  0  (x),  -<//■  (x).  This  theorem  reduces  to  the  Mean  Value  Theorem 
(§  125)  when  <f>(x)  =  x  and  ^(x)  =  l.] 

33.  Deduce  from  Ex.  32  the  formula 

f(b)-f(a)  _/'(£) 
0(6) -0(a)      0'(£) 

34.  If  $' (x)-*-a  as  x-*-co,  then  0  (x)/x-*-a.  If  0' (#)-*-<»  then 
0  (#)-*-  cc .    [Use  the  formula  0  (a?)  -  0  (#0)  =(af  —  #0)  0'  (£),  where  ^0  <  ^  <  -r-] 

35.  If  0  {x)-*-a  as  #-»-  oo ,  then  0'  (x)  cannot  tend  to  any  limit  other  than 
zero. 

36.  If  0  (.r)  +  0'  (x)-*-a  as  ar-*-oo,  then  <f>(x)-*-a  and  0'  (x)-»-0. 

[Let  (f)(x)  =  a  +  \^(x),  so  that  \//-  (:r)  +  ^;  (#)-*-0.  If  ^r'  (,r)  is  of  constant 
sign,  say  positive,  for  all  sufficiently  large  values  of  x,  then  ^  (x)  steadily 
increases  and  must  tend  to  a  limit  I  or  to  oo  .  If  -^  (x)-*~  oo  then  \//  (a:)  -»-  —  oo , 
which  contradicts  our  hypothesis.  If  y\r(x)-*-l  then  yjs' (x)-» — I,  and  this 
is  impossible  (Ex.  35)  unless  £=0.  Similarly  we  may  dispose  of  the  case  in 
which  \|/  (x)  is  ultimately  negative.  If  \js  (x)  changes  sign  for  values  of  x  which 
surpass  all  limit,  then  these  are  the  maxima  and  minima  of  yfr  (x).  If  x  has 
a  large  value  corresponding  to  a  maximum  or  minimum  of  y^(x),  then 
\}s  (x)  +  yjf'  (x)  is  small  and  i//  (x)=0,  so  that  \|/-  (.r)  is  small.  A  fortiori  are  the 
other  values  of  xfr  (x)  small  when  x  is  large. 

For  generalisations  of  this  theorem,  and  alternative  lines  of  proof,  see  a 
paper  by  the  author  entitled  "Generalisations  of  a  limit  theorem  of  Mr  Mercer," 
in  volume  43  of  the  Quarterly  Journal  of  Mathematics.  The  simple  proof 
sketched  above  was  suggested  by  Prof.  E.  W.  Hobson.] 

37.  Show  how  *  reauoe  )r  {*,  Jg$.  J(^fjdxu> 
the  integral  of  a  rational  function.     [Put  7nx+n=\jt  and  use  Ex.  xlix.  13.1 

38.  Calculate  the  integrals : 

(     dx  f      /(x-\\dx         f  xdx 

](l+afir       J\/\x+l)x>      ]s/(l+x)->/(l+x)> 

dx  /"cos x sin  xdx  f 

(2-sin^)(2+sin*-sin^)'       J  cos*  .r  +  sin*  a; '       J  cosec  *V(sec  2*)  <&, 

f  dx  fx+sinx  .  f  f 

M(l+sin*)(2+sin*)}'       JT+^Txd^      j^'^^dx,      J  (arc sin •)><*« 

[  j  [%  arc  sin  x  7  /"arc  sin  x  ,  /"arc  sin  x 

jx.rcvnxdx,      J^^^,      ./— ,-^,     j-— §  dx, 

fare  tan  a?  /"aretena?  /"log(a2+/3%2)  ,  Hog  (a+0*)  , 


DERIVATIVES    AND    INTEGRALS  259 

39.    Formulae  of  reduction,    (i)    Show  that 
a/       iw       i   n  (        dx  x  +  ^p 


«*»-*)  j$r. 


[Put  x  +  ^p  =  t,  q  —  \p2  —  \:  then  we  obtain 
[    dt  1   [_dt__    _1   [  J?dt 

j  (««+x)»  ~  x  j  (^+x)»-1    x  J  (*q-xy 

X  J  (<2  +  X)»-1  +  2X  (n-1)  J    rf<  U*2  +  X)»-_1J     '' 
and  the  result  follows  on  integrating  by  parts. 

A  formula  such  as  this  is  called  a  formula  of  reduction.     It  is  most  useful 

f         dx 

when  n  is  a  positive  integer.     We  can  then  express  l-r-s r  in  terms 

1  °  r         J  (x'+px  +  q)n 

f            dx 
of  /  7-s — ; — r,  and  so  evaluate  the  integral  for  every  value  of  n  in 

turn.] 


(ii)     Show  that  if  /,,,  7=  j xp  (1  +x)q  dx  then 


{p  +  l)lp>q  =  xr>^(\+xy-qlp  +  l>q_x, 

and  obtain  a  similar  formula  connecting  lp>q  with  ip_i,a+j..     Show  also,  by 
means  of  the  substitution  x=  —yj{\  +y),  that 


4«=(-  l)p+1/^p  (l  +y)-p-q-2dy. 

(iii)     Show  that  if  X=  a  +  bx  then 
i  xX~1'3  dx=  -  3  (3a  -  Zbx)  AT^/lOfc2, 


I 


/• 


a?X~  !/3  dtf  =  3  (9a2  -  6afcr  +  SbW)  X  2''3/40&3. 
.>;A'  -  V4  cfo=  -  4  (4a  -  3&.r)  A'3'4/21&2, 

v2X-Wdx=4:  (32a2-  24a&z  +  21&V!)  A'3/*/23]  63. 


I 


(iv)     If  7m,  „  =  |('l+J)W  then 

2(»-l)4Iitt=-a?B»-1(l+«2)-(n-1)+(wi-l)/w_2,B_1. 

(v)     If  /„  =  I  #n  cos  fix  dx   and    Jn  =  I  #"  sin  £.£  dx  then 

/3/n=a;nsin/3^-?it7(t_i,     /3-/,t=  -#ncos/&i*  +  ttin_i. 


17—2 


260 


DERIVATIVES    AND    INTEGRALS 


[VI 


(vi)     If  Zra  =  I  cos"  .rote  and  Jn=  \sinnxdx  then 
nln  =  sin  x  cosn~1x+{n-l)  4_2,    nJn=  -  cos  x  sin'1-1  x+(n  - 1)  Jn-i- 

(vii)     If  4=  \\&VLnxdx  then  (n-  1)  (4  +  /n_2)  =  tan"-1.r. 

(viii)     If  Inh  „  =  I  cosm  x  sin"  #  c&p  then 

(?n  +  7i)I„hn=  —  cosm+1#sinw-1.£  +  (?i-  1)  /,„,,; _2 

=     cosTO-1^sinn  +  1A,  +  (m-l)  lm-%n> 
[We  have 

(m  + 1 )  Im<  n  =  —  /  sin™ ~ x  j?  t-  (cosm  + 1  x)  dx 

=  -eosm+ixsinn~1  x  +  (n  —  \)   lcosm+2 xsh\n~'2xdx 

=  _  cos"1  + »  a-  sin"  -  *  ar + (to  - 1)  (/TO,„_2  -  4,  n), 
which  leads  to  the  first  reduction  formula.] 

(ix)   Connect  /„,,„=  J  sinm#  sin  nxdx  with  Jw_2,n-       {Math.  Trip.  1897.) 

(x)     If  Im<  n=  I  x™ cosecn x d.v  then 

(n-l)'(n-  2)  /,„,«  =  (»  -2)2/min_2  +  «i(m-l)/OT_2,n_2 

—  #"•  -  i  cosecn  ~ x  #  {wi  sin  x  +  (n  —  2)  x  cos  #} .  (Math.  Trip.  1 896. ) 

(xi)     If  In=  J  (a  +  b  cosx)~udx  then 
(«-l)(a2-62)/n=-fesin.r(a  +  6cos^)-("-])  +  (2H,-3)a/n_1-(n-2)/n_2. 

(xii)    If  In  =  l(acos2x+2hcosxsmx+bsm'ix)~ndx  then 

An  (»+  l)(o6  -  A2)  4+2-  2n  (2m  + 1)  (a  +  o)  In+l+4m?In=  -  ~  , 

(Math.  Trip.  1898.) 
(xiii)    If  Im< n  =  /  xm  (log  x)n  dx  then  (m  + 1 )  /,„,  „ = xm  + » (log  x)n  -  nJm<  „  _  , . 

40.  If  »  is  a  positive  integer  then  the  value  of  /  xm  (log  x)n  dx  is 

xm + 1  Kl?g  * )"  _  ttGog^)"-1      w(n-l)  (log  .*)"-» _  (-!)"» :  \ 

\m  +  l  (m  +  lf  (m  +  1)3  '"  +  (?h,  +  1)"  +  1J  ' 

41.  Show  that  the  most  general  function  (p(x),  such  that  <p"  +  a2(p  =  0  for 
all  values  of  x,  may  be  expressed  in  either  of  the  forms  A  cos  ax +B  sin  ax, 
p  cos(ax+e),  where  A,  B,  p,  c  are   constants.      [Multiplying  by   2cf>'  and 


d2In 


DERIVATIVES    AND    INTEGRALS  261 

integrating  we  obtain  #'2+a2$2  =  a2&2,  where  b  is  a  constant,  from  which  we 

deduce  that  ax=  \    .....      .„, .] 

./x/(62-0-) 

42.  Determine  the  most  general  functions  y  and  z  such  that  y'  +  a>2  =  0 
and  z'-a>y  =  0,  where  w  is  a  constant  and  dashes  denote  differentiation  with 
respect  to  x. 

43.  The  area  of  the  curve  given  by 

sin  a  sin  0  .  sinacosd) 

#=cos0+  2 — ^rr,  y  =  smd)-- — —1, 

l-cos-asm-<£     J  r     1  -  cos2a  sm2</> ' 

where  a  is  a  positive  acute  angle,  is  Jtt  (1  +sin  a)2/sin  a.        (i/a^.  Trip,  1904.) 

44.  The  projection  of  a  chord  of  a  circle  of  radius  a  on  a  diameter  is  of 
constant  length  2a  cos  j3 ;  show  that  the  locus  of  the  middle  point  of  the  chord 
consists  of  two  loops,  and  that  the  area  of  either  is  a2  (/3  —  cos  /3  sin  /3). 

{Math.  Trip.  1903.) 

45.  Show  that  the  length  of  a  quadrant  of  the  curve  (x/a)i  +  (yjb)s  =  l  is 
(a2  +  ab  +  b2)l(a  +  b).  (Math.  Trip.  1911.) 

46.  A  point  A  is  inside  a  circle  of  radius  a,  at  a  distance  b  from  the 
centre.  Show  that  the  locus  of  the  foot  of  the  perpendicular  drawn  from 
J.  to  a  tangent  to  the  circle  encloses  an  area  7r  (a2 +  |62).         (Math.  Trip.  1909.) 

47.  Prove  that  if  (a,  b,  c,  f,  g,  h~$x,  y,  1)2  =  0  is  the  equation  of  a  conic,  then 


A 


dx  .      PT     n 


(Lv  +  my  +  n)(hx  +  by+f)  a  PT 

where  PT,  PT'  are  the  perpendiculars  from  a  point  P  of  the  conic  on  the 
tangents  at  the  ends  of  the  chord  Ix  +  my  +  n=0,  and  a,  /3  are  constants. 

(Math.  Trip.  1902.) 

,.    .  ax2  +  2bx  +  c       7 

48.     Show  that  t-t-  „  — a -_ — -  ~  „  dx 

(Ax1  +  2Bx  +  C)2 

will  be  a  rational  function  of  x  if  and  only  if  one  or  other  of  AC—B2  and 
aC+cA  -  2bB  is  zero.* 


/ 


B<&* 


49.  Show  that  the  necessary  and  sufficient  condition  that 

m 

{F(x)}2 

where  /  and  F  are  polynomials  of  which  the  latter  has  no  repeated  factor, 
should  be  a  rational  function  of  x,  is  that  f'F'  -fF"  should  be  divisible  by  F. 

(Math.  Trip.  1910.) 

,.     .  fa  cos  .r+/3  sin  .r  +  -y   , 

50.  Show  that  I  — .,        K r^  dx 

J       (1  -  e  cos  x)- 

is  a  rational  function  of  cos  x  and  sin  x  if  and  only  if  ae  +  y=0 ;  and  determine 
the  integral  when  this  condition  is  satisfied.  (Math.  Trip.  1910.) 

*  See  the  author's  tract  quoted  on  p.  236. 


CHAPTER    VII 

ADDITIONAL  THEOREMS  IN  THE  DIFFERENTIAL  AND 
INTEGRAL  CALCULUS 

147.  Higher  Mean  Value  Theorems.  In  the  preceding 
chapter  (§  125)  we  proved  that  if  f(x)  has  a  derivative  f'(x) 
throughout  the  interval  (a,  b)  then 

/(&) -/(a)  =  (&-«)/' (a 

where  a<i;<b;    or  that,  if  f(x)  has  a   derivative   throughout 
(a,  a  +  h),  then 

f(a  +  h)-f(a)  =  hf'(a  +  eih)  (1)3 

where  0  <  6X  <  1.     This  we  proved  by  considering  the  function 

which  vanishes  when  x  =  a  and  when  x  =  b. 

Let  us  now  suppose  that  /  (a?)  has  also  a  second  derivative 
f"  (x)  throughout  (a,  b),  an  assumption  which  of  course  involves 
the  continuity  of  the  first  derivative  /'  (x),  and  consider  the 
function 

f(b)  -  f(x)  -  (b  -  x)f  (x)  -  (|^|J  {fib)  -f(a)  -(b-  a)f  («)}. 

This  function  also  vanishes  when  x  =  a  and  when  x  =  b;  and  its 
derivative  is 

^pjg?  {/(b)  -/(«) "  (fi  "  «)/  («)  "  i  (&  -  «)2/"  («». 

and  this  must  vanish  (§  121)  for  some  value  of  a;  between  a  and  6 
(exclusive  of  a  and  6).     Hence  there  is  a  value  £  of  x,  between 


147]  ADDITIONAL  THEOREMS   IN   THE   CALCULUS  203 

a  and  b,  and    therefore    capable    of  representation    in    the    form 
a  +  62  (b  -  a),  where  0  <  62  <  1,  for  which 

f(b)  -/(a)  +  (b  -  a)f  (a)  + \  (6  -  a)'/"  (f). 

If  we  put  b  =  a  +  h  we  obtain  the  equation 

f(a  +  h)=f(a)  +  hf'(a)  +  $h*f'(a  +  02h)  (2), 

which  is  the  standard  form  of  what  may  be  called  the  Mean  Value 
Theorem  of  the  second  order. 

The  analogy  suggested   by  (1)  and  (2)  at  once  leads  us  to 
formulate  the  following  theorem : 

Taylor's    or   the  General   Mean   Value    Theorem.      If 

f{x)  is  a  function  of  x  which  has  derivatives  of  the  first  n  orders 
throughout  the  interval  (a,  b),  then 

f(b)=f(a)  +  (b-a)f(a)+(t^^f'(a)  +  ... 

where  a<  %<b;  and  ifb  =  a  +  h  then 

f(a  +  h)=f(a)  +  hf'(a)  +  W"(a)+... 


(n-l)l 
where  0  <  6n  <  1. 

The  proof  proceeds  on  precisely  the  same  lines  as  were  adopted 
before  in  the  special  cases  in  which  n  =  1  and  n  =  2.  We  consider 
the  function 

Fn(x)-(^jFn{a), 
where       Fn{x)=f(b)  -f{x)~ (b-x)f  (x) -^-^ f" {x) -  ... 

This  function  vanishes  for  x  =  a  and  x  =  b;  its  derivative  is 

and  there  must  be  some  value  of  x  between  a  and  b  for  which 
the  derivative  vanishes.     This  leads  at  once  to  the  desired  result. 


264  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

In  view  of  the  great  importance  of  this  theorem  we  shall  give 
at  the  end  of  this  chapter  another  proof,  not  essentially  distinct 
from  that  given  above,  but  different  in  form  and  depending  on 
the  method  of  integration  by  parts. 

Examples  LV.  1.  Suppose  that  f{x)  is  a  polynomial  of  degree  r. 
Then  /(")  (x)  is  identically  zero  when  n  >  r,  and  the  theorem  leads  to  the 
algebraical  identity 

f{a+h)=f{a)+hf  (a)  +  ~  f"  (a)  +  ...+£j  /W  (a). 

2.  By  applying  the  theorem  to  f{x)  =  l/x,  and  supposing  x  and  x  +  h 
positive,  obtain  the  result 

1         1      A      A2  (-I)'1-1/*"-1         {-l)nhn 


x  +  h      x     x2     x3      '"  xn  {x  +  6nh)n  +  1' 

1        1     h      IP  ( _  i)n-i^«-i       r-iyhn 

[Since  — - r  = 2  +  ^3-—+- ~ZZ +     .,     ,  ,x, 

1  x  +  h      x      or      x6  x'1  xn  (x  +  h) 

we  can  verify  the  result  by  showing  that  at*  {x  +  h)  can  be  put  in  the  form 
{x  +  0Hh)n+1,  or  that  xn  + 1  <xn  {x+h)  <  (#  +  A)n  +  1,  as  is  evidently  the  case.] 

3.  Obtain  the  formula 

•    ,.      n       .  ,  A2    .  h3 

sin  {x  +  h)  =  sm  x+h  cos  x—  —  sin  x— ^-j  cos  #  +  ... 

+  (~  1)n_1  (2^Tj~!C°S'r  +  ("1)n  £  Shl  (*+M)> 
the  corresponding  formula  for  cos  {x+h),  and   similar  formulae  involving 
powers  of  h  extending  up  to  A'~"  +  1. 

4.  Show  that  if  m  is  a  positive  integer,  and  n  a  positive  integer  not 
greater  than  m,  then 

{x+hy»=?xm+  (™)  xm-ih  +  ...+  (j"\xm-n+1hn-l+(™){x+dji)m-»h». 

Show  also  that,  if  the  interval  {x,  x  +  h)  does  not  include  x=0,  the  formula 
holds  for  all  real  values  of  m  and  all  positive  integral  values  of  n  ;  and  that, 
even  if  x  <0<x  +  h  or  x  +  h<0  <x,  the  formula  still  holds  if  m-n  is 
positive. 

5.  The  formula  /  {x  +  h)  =/  {x)  +  hf  {x  +  8xh)  is  not  true  if  /  {x)  =  1  \x  and 
x<0<x+h.  [For  f(x+k)-f{x)>0  and  A/'  {x  +  d1h)=  -/</(#  + V)2<0  :  it 
is  evident  that  the  conditions  for  the  truth  of  the  Mean  Value  Theorem  arc 
not  satisfied.] 

6.  If  x=-a,  A  =  2a,  f{x)  =  x113,  then  the  equation 

f{x  +  h)=f{x)  +  hf{x  +  8lh) 
is  satisfied  by  ^i=^.±iV\/3.     [This  example  shows  that  the  result  of  the 
theorem  may  hold  even  if  the  conditions  under  which  it  was  proved  are 
not  satisfied.] 


147]  ADDITIONAL   THEOREMS    IN   THE    CALCULUS  205 

7.  Newton's  method  of  approximation  to  the  roots  of  equations.  Let 
£  be  an  approximation  to  a  root  of  an  algebraical  equation/ (a?)  =  0,  the  actual 
root  being  £  +  h.    Then 

0  =/(£  +  h)  =/(£)  +  hf  (£)  +  \Wf  (|  4-  62k), 
so  that  k--m-w£t§+W, 

It  follows  that  in  general  a  better  approximation  than  x  =  £  is 

If  the  root  is  a  simple  root,  so  that  /'  (|  +  A)  #=  0,  we  can,  when  A  is  small 
enough,  find  a  positive  constant  A'  such  that  \f  (x)  |  >  K  for  all  the  values  of 
x  which  we  are  considering,  and  then,  if  h  is  regarded  as  of  the  first  order  of 
smallness,  /(£)  is  of  the  first  order  of  smallness,  and  the  error  in  taking 
£  —  {/(£)//'  (£)}  as  the  root  is  of  the  second  order. 

8.  Apply  this  process  to  the  equation  x2=Z,  taking  £  =  3/2  as  the  first 
approximation.  [We  find  h—  -1/12,  £  +  A  =  17/12=l-417...,  which  is  quite  a 
good  approximation,  in  spite  of  the  roughness  of  the  first.  If  now  we  repeat 
the  process,  taking  £=17/12,  we  obtain  £  +  A  =  577/408  =  L414215...,  which 
is  correct  to  5  places  of  decimals. 

9.  By  considering  in  this  way  the  equation  .r2  —  l-y  =  0,  where  y  is 
small,  show  that  ^/(l  +y)  =  1  +\y  -  (j3/2/(2+y)}  approximately,  the  error  being 
of  the  fourth  order. 

10.  Show  that  the  error  in  taking  the  root  to  be  £  —  (///')  -  \  iff" If3)-, 
where  £  is  the  argument  of  every  function,  is  in  general  of  the  third  order. 

11.  The  equation  &mx=ax,  where  a  is  small,  has  a  root  nearly  equal  to 
7r.  Show  that  (1  —  a)  ir  is  a  better  approximation,  and  (l-a  +  a2)ir  a  better 
still.  [The  method  of  Exs.  7 — 10  does  not  depend  on  /(^)  =  0  being  an 
algebraical  equation,  so  long  as/'  and/"  are  continuous.] 

12.  Show  that  the  limit  when  A-»-0  of  the  number  6n  which  occurs  in 
the  general  Mean  Value  Theorem  is  l/(?t-+-l),  provided  that  fn  +  1)(x)  is 
continuous. 

[Vorf(x  +  h)  is  equal  to  each  of 

/(*)  +  ...+  £/W  (*+<U),   f{x)  + ...+  £/>)  (*)  +  ^^/<»+»(*+0B+1/4), 

where  #„  +  1  as  well  as  6n  lies  between  0  and  1.     Hence 

,,    w  ,    n  „    w     v        /lf<-n+1)(x  +  dn+lh) 

/(»)  (*  +  6nh)  =/(»)  (x)  +  -I A_^ +U 

But  if  we  apply  the  original  Mean  Value  Theorem  to  the  function  /(")  (x), 
taking  6nh  in  place  of  A,  we  find 

/w  (x + 6nh)  =/w  (x) + enhf(" +v(x+ eejt), 


2G6  ADDITIONAL   THEOREMS   IN  THE   CALCULUS  [VII 

where  6  also  lies  between  0  and  I.    Hence 

from  which  the  result  follows,  since /(»  +  ]> (x  +  6dnh)  and  /<•» +  x>  (a?  +  0„  + 1  A)  tend 
to  the  same  limit /<n+1)(ar)  as  /i -*►().] 

1 3.  Prove  that  {f(x  +  2A)  -  2f(x  +  k)  +f(x)}/h2+f"  (x)  as  A-^0,  provided 
that/"  (#)  is  continuous.     [Use  equation  (2)  of  §  147.] 

14.  Show  that,  if  the/(")  (x)  is  continuous  for  a?=0,  then 

f(x)  =  a0  +  a1x  +  a2x'i  +  ...  +  (an  +  ex)xn, 
where  ar  =  /(''>  (0)/r !  and  t^-^O  as  #-9-0.* 

15.  Show  that  if 

a0+a1x+a2x2  +  ...  +  (an+ex)xn=b0  +  b1x  +  b2x'2+...+(bn  +  rix)xn, 

where  ex  and  t]x  tend  to  zero  as.r-»-0,  then  a0=60,  ai  =  &i,  ...,  an=bn.  [Making 
x-^0  we  see  that  a0=b0.  Now  divide  by  x  and  afterwards  make  x-*-0. 
We  thus  obtain  a^—bi;  and  this  process  may  be  repeated  as  often  as  is 
necessary.  It  follows  that  if  f(x)  =  a0  +  a1x+a2x2+...  +  (an  +  ex)xn,  and  the 
first  n  derivatives  oif(x)  are  continuous,  then  ar—f(r)(0)/r  !.] 

148.  Taylor's  Series.  Suppose  that  f{x)  is  a  function  all 
of  whose  differential  coefficients  are  continuous  in  an  interval 
{a  —  i],  a +  7))  surrounding  the  point  x  =  a.  Then,  if  h  is  numeri- 
cally less  than  77,  we  have 

f(a  +  h)  =f{a)  +  hf  (a)  +  . . .  +  —^y,/^  (a)  +  £/«  (a  +  Bnh), 
where  0  <  6n<  1,  for  all  values  of  n.     Or,  if 

Sn  =  2  -,/»  (a),     K  =  ^/<»>(a  +  enh). 

0    V  '.  71  : 

we  have  f(a  +  h)  —  Sn  =  Rn. 

Now  let  us  suppose,  in  addition,  that  we  can  prove  that 
Rn-*~0  as  n^co  .     Then 

/(a  +  h)  =  lim  $,  =/(a)  +  hf  (a)  +  ~-f"  (a)  +  . . . . 

This  expansion  of  f(a+h)  is  known  as  Taylor's  Series. 
When  a  =  0  the  formula  reduces  to 

f(h)=f(0)  +  hf'(0)+£f'{Q)+..., 

*  It  is  in  fact  sufficient  to  suppose  that/(n'  (0)  exists.  See  R.  H.  Fowler,  "The 
elementary  differential  geometry  of  plane  curves "  (Cambridge  Tracts  in  Mathe- 
matics, No.  20,  p.  104). 


147,  148]      ADDITIONAL   THEOREMS   IN   THE   CALCULUS  2(17 

which  is  known  as  Maclaurin's  Series.    The  function  Rn  is  known 
as  Lagrange's  form  of  the  remainder. 

The  reader  should  be  careful  to  guard  himself  against  supposing  that  the 
continuity  of  all  the  derivatives  of  f(x)  is  a  sufficient  condition  for  the  validity 
of  Taylor's  series.  A  direct  discussion  of  the  behaviour  of  Rn  is  always 
essential. 

Examples  LVI.  1.  Let/(#)=siii  x.  Then  all  the  derivatives  of/(.e) 
are  continuous  for  all  values  of  x.  Also  \fn  (x)  |  ^  1  for  all  values  of  x  and  n. 
Hence  in  this  case  |  Rn  |  <  hnjn  !,  which  tends  to  zero  as»->-Qo  (Ex.  xxvn.  12) 
whatever  value  h  may  have.     It  follows  that 

.....        ,  .  li-    .  h?  W    . 

sin  (x+h)  =  sm  x+h  cos  x  -  —-.  am  x - -- cos x + -—,  sin^+..., 
2  !  3 !  41 

for  all  values  of  x  and  h.     In  particular 

.    .      .     h?      ¥ 

for  all  values  of  h.     Similarly  we  can  prove  that 

cos  (x  +  a)  =  cos  x  —  k  sin  x  —  —  cos  x  +  —  sin  x+ ...,     cos  A  =  1  —  — -  -\ :-.... 

2*  1  3  .  2  !      4 ! 

2.  The  Binomial  Series.  Let  f(x)  =  (l+x)m,  where  m  is  any  rational 
number,  positive  or  negative.  Then /(") (x)  =  m  (m  —  l)...(m  — ?i4-l)(l+a-')m-n 
and  Maclaurin's  Series  takes  the  form 


(i +xr= i + W  x+  M  z*+... 


When  m  is  a  positive  integer  the  series  terminates,  and  we  obtain  the 
ordinary  formula  for  the  Binomial  Theorem  with  a  positive  integral  exponent. 
In  the  general  case 


/?n=S/(n)(^)=C0'rn(i+^r) 


and  in  order  to  show  that  Maclaurin's  Series  really  represents  (l+#)m  for 
any  range  of  values  of  x  when  m  is  not  a  positive  integer,  we  must  show  that 
Rn-*-0  for  every  value  of  x  in  that  range.  This  is  so  in  fact  if  -  1  <.r<l, 
and  may  be  proved,  when  0^.f<l,  by  means  of  the  expression  given  above 

for  Rn,  since  (1  +  6nx)m~n  <  1  if  ?i >  m,  and  (      )  xn-^0  a,sn-*~oo  (Ex. xxvn.  13). 

But  a  difficulty  arises  if  -K.r<0,  since  \+6nx<\  and  (l  +  0„.r)"i-n>  1  if 
n  >  m  ;  knowing  only  that  0  <  6n  <  l,we  cannot  be  assured  that  1  +6ri.c  is  not 
quite  small  and  (l  +  ^)m_n  quite  large. 

In  fact,  in  order  to  prove  the  Binomial  Theorem  by  means  of  Taylor's 
Theorem,  we  need  some  different  form  for  Rn,  such  as  will  be  given  later 
(§  162). 


268  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

149.  Applications  of  Taylor's  Theorem.  A.  Maxima 
and  minima.  Taylor's  Theorem  may  be  applied  to  give  greater 
theoretical  completeness  to  the  tests  of  Ch.  VI,  §§  122 — 123, 
though  the  results  are  not  of  much  practical  importance.  It 
will  be  remembered  that,  assuming  that  <j>  (x)  has  derivatives  of 
the  first  two  orders,  we  stated  the  following  as  being  sufficient 
conditions  for  a  maximum  or  minimum  of  </>  (x)  at  x  =  ^:fora 
maximum,  0'(£)  =  O,  <£"(£)< 0;  for  a  minimum,  <f>'(%)  =  0,  <£"(f)>0. 
It  is  evident  that  these  tests  fail  if  cf>"  (£)  as  well  as  </>'(£)  is  zero. 

Let  us  suppose  that  the  first  n  derivatives 
<\>'{x),     <f>"(x),     ....     0«(«) 
are  continuous,  and  that  all  save  the  last  vanish  when  x  =  f .    Then, 
for  sufficiently  small  values  of  h, 

In  order  that  there  should  be  a  maximum  or  a  minimum  this 
expression  must  be  of  constant  sign  for  all  sufficiently  small 
values  of  h,  positive  or  negative.  This  evidently  requires  that  n 
should  be  even.  And  if  n  is  even  there  will  be  a  maximum  or  a 
minimum  according  as  <$>{n)  (£)  is  negative  or  positive. 

Thus  we  obtain  the  test :  if  there  is  to  be  a  maximum  or 
minimum  the  first  derivative  which  does  not  vanish  must  be  an  even 
derivative,  and  there  will  be  a  maximum  if  it  is  negative,  a  minimum 
if  it  is  positive. 

Examples  LVII.  1.  Verify  the  result  when  $  (x)  =  (x  —  a)m,  m  being  a 
positive  integer,  and  £=a. 

2.  Test  the  function  {x  -  a)m  (x  -  b)n,  where  m  and  n  are  positive  integers, 
for  maxima  and  minima  at  the  points  x=a,  x=b.  Draw  graphs  of  the 
different  possible  forms  of  the  curve  y  =  (x  —  a)m(x-b)n. 

3.  Test  the  functions  sin^— x,  $inx  —  x  +  '—,   sin  x  —  x-\—^  —  — — ,  ..., 

o  o       120 

cos  x  —  1,  cos  x  —  1  +  — ,  cos  x  -  1+ '-- — '— ,  ...  for  maxima  or  minima  at  x  =  0. 

150.  B.  The  calculation  of  certain  limits.  Suppose 
that /(a?)  and  <j>  (x)  are  two  functions  of  x  whose  derivatives  f  (x) 
and  <f>'  (x)  are  continuous  for  x  =  £  •  and  that  /'(£)  and  <£  (£)  are 
both  equal  to  zero.     Then  the  function 

^(x)=f{x)i<j>(x) 


149,  150]      ADDITIONAL    THEOREMS    IN    THE    CALCULUS  209 

is  not  denned  when  x  =  £.  But  of  course  it  may  well  tend  to  a 
limit  as  x-*-%. 

Now  f(x)  =f(x)  -/(£)  =  {x-  £)/' (Xl), 

where  x1  lies  between  £  and  x ;  and  similarly  <£  (x)  =  (x  —  %)  </>'  (x^), 
where  x2  also  lies  between  £  and  x.     Thus 

yjr  (x)  =/'  (a$W  («?,). 
We  must  now  distinguish  four  cases. 

(1)  If  neither/'  (£)  nor  (£'  (£)  is  zero,  then 

f{*JlH")'-+-fWifo 

(2)  If/'(£)  =  0,f(£)4=0,then 

(3)  lf/(f)  +  0,f  (f)=O,then/(aO/0(a;)  becomes  numerically 
very  large  as  x-^-%:  but  whether  f(x)/<p  (x)  tends  to  oo  or  —  oo  , 
or  is  sometimes  large  and  positive  and  sometimes  large  and 
negative,  we  cannot  say,  without  further  information  as  to  the  way 
in  which  <//  (x)-*~0  as  x-^-t;. 

(4)  If/'(£)  =  0,  <f>'(i;)  =  0,  then  we  can  as  yet  say  nothing  about 
the  behaviour  of /(#)/</>  (x)  as  x-*~0. 

But  in  either  of  the  last  two  cases  it  may  happen  that  f(x} 
and  </>  (x)  have  continuous  second  derivatives.     And  then 

f{x)  =  f{x)  -fit)  -ix-  g)/  (£)  =  i  (*  -  f)2/"  «), 

*(«)-*(*)-*(f)-(*-f)f  (f)  =  M*-£)af  (**), 

where  again  a?x  and  a;3  lie  between  £  and  a; ;  so  that 

We  can  now  distinguish  a  variety  of  cases  similar  to  those 
considered  above.  In  particular,  if  neither  second  derivative 
vanishes  for  x  =  £,  we  have 

/<*)/*(*)-/'(*)/*"  (0- 
It  is  obvious  that  this  argument  can  be  repeated  indefinitely, 
and  we  obtain  the  following  theorem:  suppose  that  fix)  and  <j>ix) 
and  their  derivatives,  so  far  as  may  be  wanted,  are  continuous  for 
x  =  %.  Suppose  further  that  f{p)ix)  and  <£(<?)  (x)  are  the  first 
derivatives  of  fix)  and  <f>ix)  which  do  not  vanish  when  x  —  g.    Then 

(i)   ifp  =  q,  /(«)/*<*Ww  (©/*««); 

(2)     ifp>q,    /(«)/*  (»)*0; 


270  ADDITIONAL   THEOREMS    IN   THE    CALCULUS  [VII 

(3)  if  p<q,  and  q—p  is  even,  either  f(x)/<j>  (x)-^+oc  or 
f(x)/(f>  (x)  -*-—  oo ,  the  sign  being  the  same  as  that  off{p)  (£)/<£ (9)  (£)', 

(4)  if  p<q  and  q—p  is  odd,  either  f(x)j(f>(x)^+cc  or 
f(x)/<f>(x)  -*  — oo  as  #^£  +  0,  the  sign  being  the  same  as  that  of 
fip)  (£)/<£ w  (IX  while  if  x-*-tj  —  0  the  sign  must  be  reversed. 

This  theorem  is  in  fact  an  immediate  corollary  from  the 
equations 

f(X)  =  fc=^/  «  (-0.    <f>  (*)  =  (fYr'  <P>  <«£ 

Examples  LVIII.     1.     Find  the  limit  of 

{x  -  (n  + 1)  xu  + » +  na;»  +  2}/(l  -  a)2, 

as  ,r-*-l.     [Here  the  functions  and  their  first  derivatives  vanish  for  x=l, 
and/"(l)=»(n+l),    <£"(1)=2.] 

2.  Find  the  limits  as  x-*-0  of 

(tan  x  -  x)/(x  —  sin  x),     (tan  H.r  -  n  tan  .r)/(%  sin  x—  sin  ?«;). 

3.  Find  the  limit  of  x {>J(x2  +  a2) - x)  as  #-*-<»  .     [Put  x=\jy!\ 

4.  Prove  that 

hm (x  —  n) cosec^77r  =  - ,     lim ^cosec#7r  —  .— -r—  }-=.— --  ' —  , 

»-*.»  7T  0-»*£  — n  ^  (#-«)7rJ  6 

n  being  any  integer ;  and  evaluate  the  corresponding  limits  involving  cot  xtt. 

5.  Find  the  limits  as  x-*~0  of 

1  /  1      x\         1  /  1      x\ 

-3^coscc^----j,       -(^cot.r--  +  3J. 

G.     (sin  x  arc  sin  x  —  x2)lx6-*-  ^g,  (tan  .r  arc  tan  a-  -  x'^/x6-*- §,  as  a.'-*-  0. 

151.  C.  The  contact  of  plane  curves.  Two  curves  are 
said  to  intersect  (or  cut)  at  a  point  if  the  point  lies  on  each  of  them. 
They  are  said  to  touch  at  the  point  if  they  have  the  same  tangent 
at  the  point. 

Let  us  suppose  now  that  f(x),  (j>  {x)  are  two  functions  which 
possess  derivatives  of  all  orders  continuous  for  x  =  i;,  and  let  us 
consider  the  curves  y  =f(x),  y  =  <f>  (x)-  In  general  /" (^)  and  (f>  (f) 
will  not  be  equal.  In  this  case  the  abscissa  x  =  £  does  not  corre- 
spond   to   a   point    of  intersection    of  the    curves.      If  however 


150,  151]      ADDITIONAL  THEOREMS   IN   THE   CALCULUS 


271 


/(£)  =  <£  (£)>  the  curves  intersect  in  the  point  x=%,  2/=/(£)=<K£). 

Let  us  suppose  this  to  be  the  case.     Then 

in  order  that  the  curves  should  not  only 

cut  but  touch  at  this  point  it  is  obviously 

necessary  and  sufficient    that    the  first 

derivatives  f  (x),  <f>'  (a?)  should  also  have 

the  same  value  when  x  =  £. 

The  contact  of  the  curves  in  this 
case  may  be  regarded  from  a  different 
point  of  view.  In  the  figure  the  two 
curves  are  drawn  touching  at  P,  and  QR 
is  equal  to  <f>  (f  +  h)  -f(£  +  h),  or,  since  <£(£)=/(£),  <£'(£)=/' <B  to 

where  6  lies  between  0  and  1.     Hence 

hm  £?  =  £{*"  (£)-/"(£)}, 

when  h^O.  In  other  words,  when  the  curves  touch  at  the  point 
whose  abscissa  is  £,  the  difference  of  their  ordinates  at  the  point 
whose  abscissa  is  £  +  h  is  at  least  of  the  second  order  of  smallness 
when  h  is  small. 

The  reader  will  easily  verify  that  lira  (QR/h)  =  <f)'  (£)  -/'  (£)  when  the  curves 
cut  and  do  not  touch,  so  that  QR  is  then  of  the  first  order  of  smallness  only. 

It  is  evident  that  the  degree  of  smallness  of  QR  may  be  taken 
as  a  kind  of  measure  of  the  closeness  of  the  contact  of  the  curves. 
It  is  at  once  suggested  that  if  the  first  n  —  1  derivatives  of  / 
and  <f>  have  equal  values  when  x  =  %,  then  QR  will  be  of  the 
?ith  order  of  smallness;  and  the  reader  will  have  no  difficulty 
in  proving  that  this  is  so  and  that 

lim^  =  l^'«)(0-/w(f)}. 

We  are  therefore  led  to  frame  the  following  definition : 

Contact  of  the  nth   order.     //  /(£)  =  <£  (£),  /'  (£)  =  <£'  (£), 

...,    y(«)(^)  =  (^)(n)(^)    iut  /"^(D^^f'+^d),    then    the    curves 

y  =f(x),  y  =  <f>  (x)  will  be  said  to  have  contact  of  the  nth  order 

at  the  point  whose  abscissa  is  %. 

The  preceding  discussion  makes  the  notion  of  contact  of  the 

nth    order  dependent  on   the  choice   of  axes,  and  fails  entirely 


272  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

when  the  tangent  to  the  curves  is  parallel  to  the  axis  of  y.  We  can 
deal  with  this  case  by  taking  y  as  the  independent  and  x  as  the 
dependent  variable.  It  is  better,  however,  to  consider  x  and  y  as 
functions  of  a  parameter  t.  An  excellent  account  of  the  theory  will 
be  found  in  Mr  Fowler's  tract  referred  to  on  p.  266,  or  in  de  la 
Vallee  Poussin's  Cours  d' Analyse,  vol.  ii,  pp.  396  et  seq. 

Examples  LIX.  1.  Let  <£  (x)  =  ax+b,  so  that  y  =  <f>  (x)  is  a  straight  line. 
The  conditions  for  contact  at  the  point  for  which  x=£  are  f(£)  =  ag  +  b, 
f  (£)  =  a.  If  we  determine  a  and  b  so  as  to  satisfy  these  equations  we  find 
a=f  (£),  &=/(£)  -  £/'  (£),  and  the  equation  of  the  tangent  to  y=f(x)  at  the 
point  x=£  is 

?=*/'(£)+{/<£-)-£/' (I)h 

or  y  -f  (0  =  (x  -  ©/.'  (|).     Cf.  Ex.  xxxix.  5. 

2.  The  fact  that  the  line  is  to  have  simple  contact  with  the  curve 
completely  determines  the  line.  In  order  that  the  tangent  should  have 
contact  of  the  second  order  with  the  curve  we  must  have  /"  (£)  =  <f>"  (£),  i.e. 
f"  (£)  =  0.  A  point  at  which  the  tangent  to  a  curve  has  contact  of  the 
second  order  is  called  a  point  of  inflexion. 

3.  Find  the  points  of  inflexion  on  the  graphs  of  the  functions  3a,4  -  6.rs  -f  1,. 
2o-7(l+  x2),  sin  a?,  a  cos2  a; +6  sin2  a?,  tana?,  arc  tan  x. 

4.  Show  that  the  conic  ax2  +  2hxy  +  by2  +  2g.v+2fy+c=0  cannot  have  a 
point  of  inflexion.     [Here  ax  +  hy +g  +  (hx  +  by+f)yl  =  0  and 

a  +  2hy,  +  tyi2  +  {hx  +  by  +f )  y.z  =  0, 
suffixes  denoting  differentiations.     Thus  at  a  point  of  inflexion 

a  +  2hy1  +  by{i  =  0, 
or         a  {hx+  by  +/)2  -  2h  (ax  +  hy  +g)  (hx  +  by  +/)  +  b  (ax  +  hy+gf=Qy 
or  (ab  -  h2)  {ax2  +  2hxy  +  by2+ 2gx  +  2fy}  +  af2-  2fgh  +  bg2 = 0. 

But  this  is  inconsistent  with  the  equation  of  the  conic  unless 

af2  -  2fgh  +  bg2=c(ab-  h2) 
or  abc  +  2fgh-af'2-bg2-ch2=0 ;    and  this  is  the  condition  that  the  conic 
should  degenerate  into  two  straight  lines.] 

5.  The  curve  y  =  (ax2+2bx  +  c)/(ax2  +  2@x  +  y)  has  one  or  three  points  of 
inflexion  according  as   the   roots   of  ax2  +  2j3x  +  y  =  0   are   real   or   complex. 

[The  equation  of  the  curve  can,  by  a  change  of  origin  (cf.  Ex.  XLVi.  15),  be 
reduced  to  the  form 

where  p,  q  are  real  or  conjugate.  The  condition  for  a  point  of  inflexion  will 
be  found  to  be  £3  —  Spq^+pq  (p-\-q}  =  0,  which  has  one  or  three  real  roots 
according  as  {pq  (p  —  q)\  is  positive  or  negative,  i.e.  according  as  p  and  q  are: 
real  or  conjugate.] 


151]  ADDITIONAL   THEOREMS    IN   THE    CALCULUS  273 

6.  Discuss  in  particular  the  curves  y  =  (l  -  x)/(l+x2),  y=(l  —  x2)/(l+x2), 
y=(l  +  s*)/(l-a*). 

7.  Show  that  when  the  curve  of  Ex.  5  has  three  points  of  inflexion,  they 
lie  on  a  straight  line.  [The  equation  £3  —  3pq$  +pq  ( p  +  q)  =  0  can  be  put  in 
the  form  (£—£>)  (£-<?)  (H+P  +  q)  +  (p  —  902£  =  O>  so  that  the  points  of  inflexion 
lie  on  the  line  £  +  A  (p-q)2r)+p  +  q=0  or  A£- 4 (AC- B2)r)  =  2B.] 

8.  Show  that  the  curves  y=xsinx,  y  =  (smx)/x  have  each  infinitely 
many  points  of  inflexion. 

9.  Contact  of  a  circle  with  a  curve.  Curvature*.  The  general 
equation  of  a  circle,  viz. 

(x-a)2  +  (y-b)2=r2 (1), 

contains  three  arbitrary  constants.  Let  us  attempt  to  determine  them  so 
that  the  circle  has  contact  of  as  high  an  order  as  possible  with  the  curve 
y=f(x)  at  the  point  (£,  r,),  where  r,  =/($).  We  write  m ,  m  for  /' (£),  f" (£). 
Differentiating  the  equation  of  the  circle  twice  we  obtain 

(*  -  a) +(y-  6)^1=0 (2), 

i+yi2+(y-^)y2=o (3). 

If  the  circle  touches  the  curve  then  the  equations  (1)  and  (2)  are  satisfied 
when  #=£  y=t),  y1=ij1.  This  gives  (|-a)/ij,=  -(r]-b)  =  rlJ(l  +  r]l2).  If 
the  contact  is  of  the  second  order  then  the  equation  (3)  must  also  be  satisfied 
when  y2  —  f]2-     Thus  b=r]  +  {(\  +r)12)/r)2} ;  and  hence  we  find 

The  circle  which  has  contact  of  the  second  order  with  the  curve  at  the  point 
(£,  tj)  is  called  the  circle  of  curvature,  and  its  radius  the  radius  of  curvature. 
The  measure  of  curvature  (or  simply  the  curvature)  is  the  reciprocal  of  the 
radius  :  thus  the  measure  of  curvature  is  /"  (£)/  {1  +  [/'  (£)]2} 3/2,  or 


del  x 


d£ 


10.  Verify  that  the  curvature  of  a  circle  is  constant  and  equal  to  the 
reciprocal  of  the  radius  ;  and  show  that  the  circle  is  the  only  curve  whose 
curvature  is  constant. 

> 

11.  Find  the  centre  and  radius  of  curvature  at  any  point  of  the  conies 

y2—Aax,  (xla)2  +  (ylb)2=l. 

12.  In  an  ellipse  the  radius  of  curvature  at  P  is  CD3 jab,  where  CD  is 
the  semi-diameter  conjugate  to  CP. 

*  A  much  fuller  discussion  of  the  theory  of  curvature  will  be  found  in  Mr  Fowler's 
tract  referred  to  on  p.  272. 

H.  18 


274  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

13.  Show  that  in  general  a  conic  can  be  drawn  to  have  contact  of  the 
fourth  order  with  the  curve  y=/(#)  at  a  given  point  P. 

[Take  the  general  equation  of  a  conic,  viz. 

ax2  +  %hxy  +  by2  +  2gx  +  %fy  +  c = 0, 
and  differentiate  four  times  with  respect  to  x.     Using  suffixes  to  denote 
differentiation  we  obtain 

ax  +  fy+g  +  {hx  +  by  +/)  y,  =  0, 
a  +  2hyl  +  by? + {has  +  by  +/)  y.2  =  0, 
3  (A  +  byx)  y2  +  {hx  +  by  +/)  y3  =  0, 
4  (A  +  by{)  y3  +  Zbyi2  +  (hx+by  +/)  y4  =  0. 
If  the  conic  has  contact  of  the  fourth  order,  then  these  five  equations  must 
be  satisfied  by  writing  £,  17,  171,  172*  *7s>  ^  for  a?,  y,  yu  y-2,  ys,  yi-     We  have  thus 
just  enough  equations  to  determine  the  ratios  a  :  b  :  c  :  f :  g  :  h.~] 

14.  An  infinity  of  conies  can  be  drawn  having  contact  of  the  third  order 
with  the  curve  at  P.     Show  that  their  centres  all  lie  on  a  straight  line. 

[Take  the  tangent  and  normal  as  axes.  Then  the  equation  of  the  conic  is 
of  the  form  2y  =  ax2  +  2hxy+by2,  and  when  x  is  small  one  value  of  y  may  be 
expressed  (Ch.  V,  Misc.  Ex.  22)  in  the  form 

y = \ax2  +  {hah  +  ex)  xs, 
where  ex->-0  with  x.     But  this  expression  must  be  the  same  as 

y-i/w(0)*»+{*/w(0)+«,'}*» 
where  ex'-*-0  with  x,  and  so  a=f"  (0),  h=f"  (0)/3/"  (0),  in  virtue  of  the  result 
of  Ex.  lv.  15.     But  the  centre  lies  on  the  line  ax  +  hy  =  0.] 

15.  Determine  a  parabola  which  has  contact  of  the  third  order  with  the 
ellipse  {x/a)'i  +  {y/b)'i=l  at  the  extremity  of  the  major  axis. 

16.  The  locus  of  the  centres  of  conies  which  have  contact  of  the  third 
order  with  the  ellipse  (.r/a)2  +  (y/6)2  =  l  at  the  point  (a  cos  a,  b  sin  a)  is  the 
diameter  xj{a  cos  a)  =yj{b  sin  a).     [For  the  ellipse  itself  is  one  such  conic] 

152.     Differentiation  of  functions  of  several  variables. 

So  far  we  have  been  concerned  exclusively  with  functions  of  a 
single  variable  x,  but  there  is  nothing  to  prevent  us  applying  the 
notion  of  differentiation  to  functions  of  several  variables  x,  y, .... 

Suppose  then  that/(#,  y)  is  a  function  of  two*  real  variables 
x  and  y,  and  that  the  limits 

]im  f(x+h>y)-f(x>y)     lim  f(x>y  +  k)-f(x,y) 

*  The  new  points  which  arise  when  we  consider  functions  of  several  variables 
are  illustrated  sufficiently  when  there  are  two  variables  only.  The  generalisations 
of  our  theorems  for  three  or  more  variables  are  in  general  of  an  obvious  character. 


151,  152]      ADDITIONAL   THEOREMS   IN   THE   CALCULUS  275 

exist  for  all  values  of  x  and  y  in  question,  that  is  to  say  that 
f(x,  y)  possesses  a  derivative  df/dx  or  Dxf(x,y)  with  respect  to  x 
and  a  derivative  df/dy  or  Dyf(x,  y)  with  respect  to  y.  It  is  usual 
to  call  these  derivatives  the  partial  differential  coefficients  of/,  and 
to  denote  them  by 

y  v 

dx'    dy 
OT  fxip,y),    fy{os,y) 

or  simply  fj,fy  orfx,  fy.  The  reader  must  not  suppose,  .however, 
that  these  new  notations  imply  any  essential  novelty  of  idea: 
'  partial  differentiation '  with  respect  to  x  is  exactly  the  same 
process  as  ordinary  differentiation,  the  only  novelty  lying  in  the 
presence  in  /  of  a  second  variable  y  independent  of  x. 

In  what  precedes  we  have  supposed  x  and  y  to  be  two  real 
variables  entirely  independent  of  one  another.  If  x  and  y  were 
connected  by  a  relation  the  state  of  affairs  would  be  very  different. 
In  this  case  our  definition  of  fx  would  fail  entirely,  as  we  could 
not  change  x  into  x  +  h  without  at  the  same  time  changing  y. 
But  then  f{oc,y)  would  not  really  be  a  function  of  two  variables 
at  all.  A  function  of  two  variables,  as  we  defined  it  in  Ch.  II, 
is  essentially  a  function  of  two  independent  variables.  If  y  depends 
on  x,  y  is  a  function  of  x,  say  y  =  <£  (x)  ;  and  then 

/<>>  2/)  =./>>  </><>)} 
is  really  a  function  of  the  single  variable  x.  Of  course  we  may  also 
represent  it  as  a  function  of  the  single  variable  y.  Or,  as  is  often 
most  convenient,  we  may  regard  x  and  y  as  functions  of  a  third 
variable  t,  and  then  f  (x,  y),  which  is  of  the  form  f{cj>  (t),  ijr(t)}, 
is  a  function  of  the  single  variable  t. 

ExamplesLX.  1.  Provethatif  x=rcos$,  y=rsind,  so  that  r=v/(.r2+j/2), 
#  =  arc  tan  (yjx),  then 

dr  _        x  dr  _        y  dd  _         y  dd  _      x 

dx     \/(^'"+3/2)'      dy     v '(x2  +#2)'      dx        x2+y2'      dy     x^+y21 

COG  C2J  v3C  OU 

-~  =  cos#,  ^  =  sin#,  ^=  —  rsind,       ^~  =  rcosd. 

dr  dr  dd  dd 

2.     Account  for  the  fact  that  ^-=|=l/(^j  and  5-, 4=1  / (ia)  •     [When 

we  were  considering  a  function  y  of  one  variable  x  it  followed  from  the 
definitions  that  dy/dx  and  dx/dy  were  reciprocals.     This  is  no  longer  the 

18—2 


276 


ADDITIONAL   THEOREMS   IN   THE    CALCULUS 


[VII 


Fig.  46. 


case  when  we  are  dealing  with  functions  of  two  variables.  Let  P  (Fig.  46) 
\>e  the  point  (x,  y)  or  (r,  6).  To  find  dr/dx  we  must  increase  x,  say  by  an 
increment  MMx  =  bx,  while  keeping  y  constant.  This  brings  P  to  Px.  If 
along  0PX  we  take  OP'  =  OP,  the  increment  of  r  is  P'P1  =  Sr,  say;  and 
dr/dx =lim(8r/dtf).  If  on  the  other  hand  we  want  to  calculate  3a?/3r,  a;  and 
y  being  now  regarded  as  functions  of  r 
and  0,  we  must  increase  r  by  Ar,  say, 
keeping  0  constant.  This  brings  P  to 
P2,  where  PP2  =  Ar:  the  corresponding 
increment  of  x  is  JO/i  =  Ax,  say  ;   and 

•    dxjdr =lim.  (Ax/ Ar). 
Now  A.z=&r*  :  but  Ar=¥8r.     Indeed  it  is 
easy  to  see  from  the  figure  that 

lim  (8r/8x)  =  lim  (P'A  /  PPX)  =  cos  6, 
but  lim  (Ar/Aa?)  =  lim  (PP2/  ^A)  =  sec  6, 
so  that        lim  (Sr/Ar)  =  cos2  0. 

"The  fact  is  of  course  that  dxjdr  and 
dr/dx  are  not  formed  upon  the  same  hypothesis  as  to  the  variation  of  P.] 

3.  Prove  that  if  z  =f(ax  +  by)  then  b  (dz/dx)  =  a  (dz/dy). 

4.  Find  dX/dx,  dX/dy,  ...  when  X+Y—x,  Y=xy.  Express  x,  y  as 
functions  of  X,   Y  and  find  3#/3X,  3^/3 7,  .... 

5.  Find  dX/dx,  ...  when  X+7"+Z=a',  Y+Z=xy,  Z=xyz;  express 
x,  y,  z  in  terms  of  X,   Y,  Z  and  find  dx/dX,  .... 

[There  is  of  course  no  difficulty  in  extending  the  ideas  of  the  last  section 
to  functions  of  any  number  of  variables.  But  the  reader  must  be  careful  to 
impress  on  his  mind  that  the  notion  of  the  partial  derivative  of  a  function  of 
several  variables  is  only  determinate  when  all  the  independent  variables  are 
specified.  Thus  if  u  =  x+y  +  z,  x,  y,  and  z  being  the  independent  variables, 
then  du/dx—1.  But  if  we  regard  u  as  a  function  of  the  variables  x,  x+y  —  rj, 
and  x+y  +  z  =  {,  so  that  u  =  (,  then  du/dx=0.~\ 

153.     Differentiation   of  a   function   of  two   functions. 

There  is  a  theorem  concerning  the  differentiation  of  a  function 
of  one  variable,  known  generally  as  the  Theorem  of  the  Total 
Differential  Coefficient,  which  is  of  very  great  importance  and 
depends  on  the  notions  explained  in  the  preceding  section  re- 
garding functions  of  two  variables.  This  theorem  gives  us  a  rule 
for  differentiating 

with  respect  to  t. 

*  Of  course  the  fact  that  Ax  =  8x  is  due  merely  to  the  particular  value  of  Ar 
that  we  have  chosen  (viz.  PP2).  Any  other  choice  would  give  us  values  of  Ax,  Ar 
proportional  to  those  used  here. 


152,  153]      ADDITIONAL   THEOREMS   IN   THE   CALCULUS  277 

Let  us  suppose,  in  the  first  instance,  that  f(x,  y)  is  a  function 
of  the  two  variables  x  and  y,  and  that  fx',  fy'  are  continuous 
functions  of  both  variables  (§  107)  for  all  of  their  values  which 
come  in  question.  And  now  let  us  suppose  that  the  variation  of 
x  and  y  is  restricted  in  that  (x,  y)  lies  on  a  curve 

x  =  <f>  (t),     y  =  yfr  (t), 

where  cf>  and  ty  are  functions  of  t  with  continuous  differential 
coefficients  <f>'  (t),  yjr'  (t).  Then/O,  y)  reduces  to  a  function  of  the 
single  variable  t,  say  F(t).     The  problem  is  to  determine  F' (t). 

Suppose  that,  when  t  changes  to  t  +  r,  x  and  y  change  to 
x  +  £  and  y  +  y.     Then  by  definition 

^T  =  Si  r [/  {<j3  (t  +  T)'  * {t  +  T)]  ~f  tt  (t)>  +  OH 
= lim  ;  I/O  +  ly  +  v)  -/(*,  y)} 
= Yim[f(x+^y+v)-f(x>y+v)  I ,  /(*»  y  +  v)  -/(*,  y)  5 

L  £  T  7;  T 

But,  by  the  Mean  Value  Theorem, 

{/(«  + &  y  +  v)  -/(*,  y  +  vM  -/.'  (*  +  0£  y  +  *), 
{/(«.  y  +  v)  -/(*.  y)}A?  =//  («,  y  +  ^)> 

where  0  and  0'  each  lie  between  0  and  1.  As  t-*-0,  £-^0  and 
r)-*-0,  and  Z/T-*-<f)'  (t),  rj/T-^y{r'(t):  also 

/.'  (*  +  0$  y  +  *  W.'  fe  y),   /,'  (^  y  +  0^  W*  (*  y). 

Hence 

J"  (0  =  A/{*  (0,  *  (0}  -/.'  (*  y)  f  (0  +/*'  te  y)  t'  (0, 

where  we  are  to  put  x  =  <£  (£),  y  =  ty(t)  after  carrying  out  the 
differentiations  with  respect  to  #  and  y.  This  result  may  also  be 
expressed  in  the  form 

df  =dfdx +  d/dy 
dt     dx  dt      dy  dt 

Examples  LXI.     1.     Suppose  <f>  (t)  =  (1  -  f-)l(l  +  fi),  ^  (t)  =  2t/(l  + 1%  so 
that  the  locus  of  (x,  y)  is  the  circle  x2+y2=l.     Then 

4>'  (o  =  -  4*/(i + *2)2,   ^'  W = 2  (i  -  W + '2)2> 

i?»  (o={-4*/(i+m/.'+{2  a  -  w+ w»'. 

where  #  and  y  are  to  he  put  equal  to  (1  -  «2)/(l  + 12)  and  2t/(l  +  t2)  after 
carrying  out  the  differentiations. 


278  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

We  can  easily  verify  this  formula  in  particular  cases.  Suppose,  e.g., 
that  /(#,  y)  =  x2  +  y1.  Then  fx  =  2x,  /„'  =  2y,  and  it  is  easily  verified  that 
F' (t)  =  2x (f)' (t)  +  2i/^'  (t)  =  0,  which  is  obviously  correct,  since  F(t)  =  l. 

2.  Verify  the  theorem  in  the  same  way  when  (a)  x—tm,  y=l  —  t"\ 
f(x,y)  =  x+y;   (b)  %=acoat,  y  =  asmt,  f(x,  y)  =  x2+y2. 

3.  One  of  the  most  important  cases  is  that  in  which  t  is  x  itself.  We 
then  obtain 

Dxf{x,  ^{x))=Dxf(x,  y)  +  DJ(x,  y)  *'(#)• 

where  y  is  to  be  replaced  by  \fr  (x)  after  differentiation. 

It  was  this  case  which  led  to  the  introduction  of  the  notation  df/dx,  df/dy. 
For  it  would  seem  natural  to  use  the  notation  df/dx  for  either  of  the  functions 
Dxf{x,  ty  (x)\  and  Dxf(x,  y),  in  one  of  which  y  is  put  equal  to  ^  (x)  before 
and  in  the  other  after  differentiation.  Suppose  for  example  that  y=\—x 
and  f(x,  y)=x+y.     Then  Dxf(x,  l-x)  =  DxI  =  0,  but  Bxf(x,  y)=\. 

The  distinction  between  the  two  functions  is  adequately  shown  by 
denoting  the  first  by  dfjdx  and  the  second  by  df/dx,  in  which  case  the 
theorem  takes  the  form 

df  =  df     df  dy  m 

dx     dx     dy  dx ' 

though  this  notation  is  also  open  to  objection,  in  that  it  is  a  little  misleading 

to  denote  the  functions  f{x,  -v//-  (x)}  and  f(x,  y),  whose  forms  as  functions  of  x 

are  quite  different  from  one  another,  by  the  same  letter /in  df/dx  and  df/dx. 

4.  If  the  result  of  eliminating  t  between  x=4>  (t),  y  =  ^r(t)  is  f(x,  y)  =  0, 
then 

dfdx  +  dfdy  =  Q 
dx  dt      dy  dt 

5.  If  x  and  y  are  functions  of  t,  and  r  and  6  are  the  polar  coordinates  of 
(x,  y),  then  r'=(xx' +yy')/r,  &  =  (xy'  -  yx')/r2,  dashes  denoting  differentiations 
with  respect  to  t. 

154.  The  Mean  Value  Theorem  for  functions  of  two 
variables.  Many  of  the  results  of  the  last  chapter  dejoended 
upon  the  Mean  Value  Theorem,  expressed  by  the  equation 

<£  (x  +  h)-cf>  O)  =  /;/'  O  +  6h), 
or  as  it  may  be  written,  if  y  =  (p  (%), 

8y=f(x  +  08x)8x. 

Now  suppose  that  z  =f(x,  y)  is  a  function  of  the  two  inde- 
pendent variables  x  and  y,  and  that  x  and  y  receive  increments 
h,  k  or  8x,  8y  respectively :  and  let  us  attempt  to  express  the 
corresponding  increment  of  z,  viz. 

hz=f{x  +  h,y  +  k)-f{m,y), 

in  terms  of  h,  k  and  the  derivatives  of  z  with  respect  to  x  and  y. 


153,  154]      ADDITIONAL   THEOREMS    IN   THE   CALCULUS  279 

Let  f(x  +  ht,y  +  kt)  =  F(t).     Then 

f(v  +  h,y  +  k)-f(x,y)  =  F(l)-F(0)  =  F'(e), 
where  0  <  0  <  1.     But,  by  §  153, 

F'(t)  =  Dtf(x  +  ht,y  +  M) 

=  hfx  O  +  ht,y  +  kt)  +  hfy  (x.  +  ht,y  +  kt). 
Hence  finally 

8z  =f(x  +  h,  y+k)  -f(x,  y)=hfx'(x  +  0h,  y+6k)  +  kf;(x  +  0h,y  +  6k), 
which  is  the  formula  desired.  Since  fx,  fy'  are  supposed  to  be 
continuous  functions  of  x  and  y,  we  have 

fx (x  +  eh,  y  +  6k)  =fx' (x,  y)  +  eh)k, 

fy'  (x  +  6h,  y  +  ek)  =//  (as,  y)  +  r}h)k, 

where  e^h  and  r)hik  tend  to  zero  as  h  and  k  tend  to  zero.  Hence 
the  theorem  may  be  written  in  the  form 

8z  =  (fx'  +  e)8x  +  (f;  +  v)8y     (1), 

where  e  and  i)  are  small  when  8x  and  8y  are  small. 

The  result  embodied  in  (1)  may  be  expressed  by  saying  that  the 
equation 

8z  =fx'8x  +fy'8y 

is  approximately  true ;  i.e.  that  the  difference  between  the  two 
sides  of  the  equation  is  small  in  comparison  with  the  larger  of  8x 
and  By*.  We  must  say  'the  larger  of  8x  and  8y'  because  one  of 
them  might  be  small  in  comparison  with  the  other;  we  might 
indeed  have  8x  —  0  or  8y  =  0. 

It  should  be  observed  that  if  any  equation  of  the  form  8z=\8x+p8i/ 
is  'approximately  true'  in  this  sense,  we  must  have  X  =fx',  p=fy'-  For  we 
have 

&? -/*' &v -fv'fy=*Sx  +  rlty,     8z-\8x-p.8y  =  e  8x  +  rj' 8y 

where  e,  rj,  c',  rj'  all  tend  to  zero  as  8x  and  8y  tend  to  zero  ;  and  so 

(X -U)  8x  +  (ji-fvf)  8y  =  P  8x  +p'fy 
where  p  and  p  tend  to  zero.     Hence,  if  £  is  any  assigned  positive  number,  we 
can  choose  o-  so  that 

I  (\  -/«*)  &* + 0*  -//)  fy  I  <  C  ( I  8*  I  + 1 8v  I ) 

for  all  values  of  8x  and  83/  numerically  less  than  o\  Taking  8y=0  we  obtain 
1  (\-/J)  8x  I  <  f  I  &r  I ,  or  I X  -/.'  |  <  £,  and,  as  f  may  be  as  small  as  we  please, 
this  can  only  be  the  case  if  X  =  fx'.     Similarly  p—fy'. 

*  Or  with  \8x\  +  \8y\  or  ,J  {dx*  +  8yn-). 


280  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

155.  Differentials.  In  the  applications  of  the  Calculus, 
especially  in  geometry,  it  is  usually  most  convenient  to  work  with 
equations  expressed  not,  like  equation  (1)  of  §  154,  in  terms  of  the 
increments  Bx,  By,  Bz  of  the  functions  x,  y,  z,  but  in  terms  of  what 
are  called  their  differentials  dx,  dy,  dz. 

Let  us  return  for  a  moment  to  a  function  y=f(x)  of  a  single 
variable  x.     Iff  (x)  is  continuous  then 

By={f(x)+e}Bx    (1), 

where  e-^0  as  Bx-^0:  in  other  words  the  equation 

By=f'{x)Bx (2) 

is 'approximately '  true.  We  have  up  to  the  present  attributed 
no  meaning  of  any  kind  to  the  symbol  dy  standing  by  itself.  We 
now  agree  to  define  dy  by  the  equation 

dy=f'(x)Bx (3). 

If  we  choose  for  y  the  particular  function  x,  we  obtain 

dx  =  Bx    (4), 

so  that  dy=f'(x)dx (5). 

If  we  divide  both  sides  of  (5)  by  dx  we  obtain 

t-fto ^ 

where  dyfdx  denotes  not,  as  heretofore,  the  differential  coefficient 
of  y,  but  the  quotient  of  the  differentials  dy,  dx.  The  symbol 
dy/dx  thus  acquires  a  double  meaning ;  but  there  is  no  incon- 
venience in  this,  since  (6)  is  true  whichever  meaning  we  choose. 

The  equation  (5)  has  two  apparent  advantages  over  (2).  It  is  exact  and 
not  merely  approximate,  and  its  truth  does  not  depend  on  any  assumption  as 
to  the  continuity  of/'  (x).  On  the  other  hand  it  is  precisely  the  fact  that  we 
can,  under  certain  conditions,  pass  from  the  exact  equation  (5)  to  the  approxi- 
mate equation  (2),  which  gives  the  former  its  importance.  The  advantages  of 
the  '  differential '  notation  are  in  reality  of  a  purely  technical  character.  These 
technical  advantages  are  however  so  great,  especially  when  we  come  to  deal 
with  functions  of  several  variables,  that  the  use  of  the  notation  is  almost 
inevitable. 

When  /'  (x)  is  continuous,  we  have 

limfUl 

when  S.r-^0.  This  is  sometimes  expressed  by  saying  that  dy  is  the  principal 
part  of  by  when  bx  is  small,  just  as  we  might  say  that  ax  is  the  '  principal 
part '  of  ax  +  bx2  when  x  is  small. 


155]  ADDITIONAL  THEOREMS   IN   THE   CALCULUS  281 

We  pass  now  to  the  corresponding  definitions  connected  with 
a  function  z  of  two  independent  variables  x  and  y.  We  define  the 
differential  dz  by  the  equation 

dz=fx'$a;+fy'8y    (7). 

Putting  z  =  x  and  z  =  y  in  turn,  we  obtain 

dx  =  Bx,     dy  =  By (8), 

so  that  dz  =fx'dx  +f,fdy (9), 

which  is  the  exact  equation  corresponding  to  the  approximate 
equation  (1)  of  §  154.  Here  again  it  is  to  be  observed  that  the 
former  is  of  importance  only  for  reasons  of  practical  convenience 
in  working  and  because  the  latter  can  in  certain  circumstances  be 
deduced  from  it. 

One  property  of  the  equation  (9)  deserves  special  remark.  "We  saw  in 
§  153  that  if  z=f(x,  y),  x  and  y  being  not  independent  but  functions  of  a 
single  variable  t,  so  that  z  is  also  a  function  of  t  alone,  then 

dz  _  df  dx      df  dy 
dt     dx  dt      dy  dt  ' 

Multiplying  this  equation  by  dt  and  observing  that 

.       dx  .  7       dy   .  7      dz  . 

dx=didt>  dy=ttdt>  dz=dtdt> 

we  obtain  dz=fx'dx+fy'dy, 

which  is  the  same  in  form  as  (9).  Thus  the  formula  which  expresses  dz  in  terms 
of  dx  and  dy  is  the  same  whether  the  variables  x  and  y  are  independent  or  not. 
This  remark  is  of  great  importance  in  applications. 

It  should  also  be  observed  that  if  z  is  a  function  of  the  two  independent 
variables  x  and  y,  and 

dz=\dx  +  fidy, 

then  \=fx,  p—fy-     This  follows  at  once  from  the  last  paragraph  of  §  154. 

It  is  obvious  that  the  theorems  and  definitions  of  the  last  three  sections 
are  capable  of  immediate  extension  to  functions  of  any  number  of  variables. 

Examples  LXII.  1.  The  area  of  an  ellipse  is  given  by  A  =  trab,  where 
a,  b  are  the  semiaxes.     Prove  that 

dA  _  da     db 

and  state  the  corresponding  approximate  equation  connecting  the  increments 
of  the  axes  and  the  area. 


282  ADDITIONAL   THEOREMS   IN    THE   CALCULUS  [VII 

2.  Express  A,  the  area  of  a  triangle  ABC,  as  a  function  of  (i)  a,  B,  C, 
(ii)  A,  b,  c,  and  (iii)  a,  b,  c,  and  establish  the  formulae 

dA     „  da        cdB  bdC  dA         ,    .   ,  A     db     do 

=  2 1 : =H : ~,         =  COtAdA  +  ^--\ , 

A  a      asmB     asmC         A  be7 

dA  =  R  (cos  Ada  +  cos  Bdb  +  cos  Cdc), 
where  R  is  the  radius  of  the  circumcircle. 

3.  The  sides  of  a  triangle  vary  in  such  a  way  that  the  area  remains 
constant,  so  that  a  may  be  regarded  as  a  function  of  b  and  c.     Prove  that 

da         cos  B        da  _     cos  C 
db  ~     cos  A '      dc  cos  A ' 

[This  follows  from  the  equations 

da  =  ^rdb  -»-  =-  dc,   cos  Ada  +  cos Bdb  + cos  Cdc =0. 
db  dc 

4.  If  a,  b,  c  vary  so  that  R  remains  constant,  then 

da  db  dc 


and  so 


cos  A      cos  B     cos  C 

da  _      cos  A        da  _      cos  A 

db  ~~      cos  B '      dc  ct  is  C 


[Use  the  formulae  a=2i2sin.4j  ...,  and  the  facts  that  R  and  A  +  B  +  C  &re 
constant.] 

5.  If  2  is  a  function  of  u  and  v,  which  are  functions  of  x  and  y,  then 

dz  _  9,3  du      dz  dv        dz  _  dz  du      dz  dv 
dx~dudx     3v3#'      dy     dudy     dvdy' 

[We  have 

7      dz   7      dz  ,  7      du  7      du  7  ,      3d  7       3d   7 

cfo = k-  du  +  x-  dv,       du  =  *—dx  +  7r-  dy,       dv=^r-dx  +  K-  dy. 

ou  ov  ox  dy  ox  oy 

Substitute  for  du  and  dv  in  the  first  equation  and  compare  the  result  with 
the  equation 

7      dz  ,       dz  7   n 
ds=dxdx  +  dyd^ 

6.  Let  z  be  a  function  of  x  and  y,  and  let  X,  Y,  Z  be  defined  by  the 
equations 

x=a1X+blY+c1Z,    y  =  a2X+b2Y+c2Z,     z=a3X+b3Y+c3Z. 

Then  Z  may  be  expressed  as  a  function  of  Jf  and  Y.  Express  dZ/dX, 
dZ/dY  in  terms  of  82/3*,  dzjdy.  [Let  these  differential  coefficients  be  denoted 
by  P,  Q  and  p,  q.     Then  dz  -  pdx  -  qdy  —  0,  or 

(cxp  +  c2q- c3)  dZ+  (axp  +  a2q-  a3)  dX+  {btf  +  b^q-  b3)  d Y=  0. 


155,  156]      ADDITIONAL   THEOREMS   IN   THE   CALCULUS 


283 


Comparing  this  equation  with  dZ—PdX-  QdY=Q  we  see  that 


P=- 


a1p  +  a2q-a3 


q_       6lff  +  &2g-&3   -i 


7. 
tlien 


Cip  +  c2q-c3'      v         cip  +  c2q-c3 

If         (ax  x  +  bxy  +  Cj  z)  p  +  (a.2x  +  b2y  +  c2z)q  =  a3x+b3y  +  c3z, 
(a^r+M'+CjZ)  P+(a2X+b2Y+c2Z)  Q  =  a3X+b3Y+c3Z. 

(Math.   Trip.  1899.) 


8.  Differentiation  of  implicit  functions.  Suppose  that/(#,  y)  and  its 
derivative  fy'  (x,  y)  are  continuous  in  the  neighbourhood  of  the  point  (a,  b), 
and  that 

f(a,b)  =  0,      ft(a,b)*0. 

Then  we  can  find  a  neighbourhood  of  (or,  b)  throughout  which  fv'  (x,  y)  has 
always  the  same  sign.  Let  us  suppose,  for  example,  that  fj  (x,  y)  is  positive 
near  (a,  b).  Then  f(x,  y)  is,  for  any  value  of  x  sufficiently  near  to  a,  and  for 
values  of  y  sufficiently  near  to  b,  an  increasing  function  of  y  in  the  stricter 
sense  of  §  95.  It  follows,  by  the  theorem  of  §  108,  that  there  is  a  unique 
continuous  function  y  which  is  equal  to  b  when  x= a  and  which  satisfies  the 
equation /(.r,  y)  =  0  for  all  values  of  x  sufficiently  near  to  a. 

Let  us  now  suppose  that  f(x,y)  possesses  a  derivative  fx'  (x,y)  which  is 
also  continuous  near  (a,  b).     If  f(x,  y)=0,  x=a  +  h,  y  =  b  +  k,  we  have 

where     and  ~\  tend  to  zero  with  h  and  k.     Thus 

h  fh'  +  r)~*      /ft" 


fb'+ 

1 

dy 
dx  ~  ~ 

fa 
7» 

9.     The  equation  of  the  tangent  to  the  curve  /  (x,  y)  =  0,  at  the  point 

(x -  x0)  fXo'  (x0,y0)  +  (2/-y0)  fy'  (x0 ,  y0) = 0. 

156.  Definite  Integrals  and  Areas.  It  will  be  remembered 
that,  in  Ch.  VI,  §  145,  we  assumed  that,  if  f{x)  is  a  continuous 
function  of  x,  and  PQ  is  the 
graph  of  y=f{x),  then  the 
region  PpqQ  shown  in  Fig.  47 
has  associated  with  it  a  definite 
number  which  we  call  its  area. 
It  is  clear  that,  if  we  denote 
Op  and  Oq  by  a  and  x,  and 
allow  x  to  vary,  this  area  is  a 
function  of  x,  which  we  denote 
byF(x). 


Fig.  47. 


284  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

Making  this  assumption,  we  proved  in  §  145  that  F'  (x)  =/(#), 
and  we  showed  how  this  result  might  be  used  in  the  calculation 
of  the  areas  of  particular  curves.  But  we  have  still  to  justify 
the  fundamental  assumption  that  there  is  such  a  number  as  the 
area  F(x). 

We  know  indeed  what  is  meant  by  the  area  of  a  rectangle, 
and  that  it  is  measured  by  the  product  of  its  sides.  Also  the 
properties  of  triangles,  parallelograms,  and  polygons  proved  by 
Euclid  enable  us  to  attach  a  definite  meaning  to  the  areas  of 
such  figures.  But  nothing  which  we  know  so  far  provides  us  with 
a  direct  definition  of  the  area  of  a  figure  bounded  by  curved  lines. 
We  shall  now  show  how  to  give  a  definition  of  F{x)  which  will 
enable  us  to  prove  its  existence.* 

Let  us  suppose  f{x)  continuous  throughout  the  interval  (a,  b), 
and  let  us  divide  up  the  interval  into  a  number  of  sub-intervals 
by  means  of  the  points  of  division  x0,  xu  x2, ...,  xn,  where 

Oj  -—  OCq  "V.  #1  \  . .  •  ^  OC^i — i  \  Ou)i  —  0, 

Further,  let  us  denote  by  S„  the  interval  {xv,  xv+1),  and  by  mv  the 
lower  bound  (§  102)  of/(#)  in  8„,  and  let  us  write 

s  =  m0S0  +  m181  + ...  +mllSn  =  'Zmv8v> 
say. 

It  is  evident  that,  if  M  is  the  upper  bound  of  f(x)  in  (a,  b),  then 
s^  M  (b  —  a).  The  aggregate  of  values  of  s  is  therefore,  in  the 
language  of  §  80,  bounded  above,  and  possesses  an  upper  bound 
which  we  will  denote  by  j.  No  value  of  s  exceeds  j,  but  there  are 
values  of  s  which  exceed  any  number  less  thanj. 

In  the  same  way,  if  Mv  is  the  upper  bound  off(x)  in  8„,  we  can 
define  the  sum 

S=1M„K> 

It  is  evident  that,  if  m  is  the  lower  bound  of  f{x)  in  (a,  b),  then 
S  ^  m  (b  —  a).  The  aggregate  of  values  of  S  is  therefore  bounded 
below,  and  possesses  a  lower  bound  which  we  will  denote  by  J. 
No  value  of  S  is  less  than  J,  but  there  are  values  of  S  less  than  any 
number  greater  than  J. 

*  The  argument  which  follows  is  modelled  on  that  given  in  Goursat's  Cours 
d' Analyse  (second  edition),  vol.  i,  pp.  171  et  seq. ;  but  Goursat's  treatment  is  much 
more  general. 


156] 


ADDITIONAL    THEOREMS    IN    THE    CALCULUS 


285 


It  will  help  to  make  clear  the  significance  of  the  sums  s  and  S  if 
we  observe  that,  in  the  simple  case 
in  which  f(x)  increases  steadily 
from  x=a  to  x  —  b,  mv  is  f(xv) 
and  Mv  is  f(xv  +  {).  In  this  case  s 
is  the  total  area  of  the  rectangles 
shaded  in  Fig.  48,  and  S  is  the 
area  bounded  by  a  thick  line.  In 
general  s  and  S  will  still  be  areas, 
composed  of  rectangles,  respectively 
included  in  and  including  the  curvi- 
linear region  whose  area  we  are 
trying  to  define. 

We  shall  now  show  that  no 
sum  such  as  s  can  exceed  any 


Fig.  48. 


sum  such  as  S.  Let  s,  S  be  the  sums  corresponding  to  one  mode  of 
subdivision,  and  s',  S'  those  corresponding  to  another.  We  have 
to  show  that  s  ^  S'  and  s  ^  S. 

We  can  form  a  third  mode  of  subdivision  by  taking  as  dividing 
points  all  points  which  are  such  for  either  s,  S  or  s,  S'.  Let  s,  S 
be  the  sums  corresponding  to  this  third  mode  of  subdivision. 
Then  it  is  easy  to  see  that 

s^s,     s^s',     S^S,     S^S' (1). 

For  example,  s  differs  from  s  in  that  at  least  one  interval  S„  which 
occurs  in  s  is  divided  into  a  number  of  smaller  intervals 

so  that  a  term  m„S„  of  s  is  replaced  in  s  by  a  sum 

mvA  £„,!  +  w„]2  S„>2  +  ...  +  mv p  hvp, 

where  m,A,  m„>2,  ...  are  the  lower  bounds  of  f(x)  in  8Vjl,  S„>2,  .... 
But  evidently  m„a  =?/i„,m„j2  =  mv, ...,  so  that  the  sum  just  written 
is  not  less  than  mv8„.  Hence  s.=  s;  and  the  otherinequalities  (1) 
can  be  established  in  the  same  way.  But,  since  s  ^  S,  it  follows 
that 

s^s^S^S', 

which  is  what  we  wanted  to  prove. 

It  also  follows  that  j  ^  J.  For  we  can  find  an  s  as  near  to  j 
as  we  please  and  an  S  as  near  to  J  as  we  please  *,  and  so  j  >J 
would  involve  the  existence  of  an  s  and  an  S  for  which  s  >  S. 


The  s  and  the  S  do  not  in  general  correspond  to  the  same  mode  of  subdivision. 


286  ADDITIONAL   THEOREMS   IN   THE   CALCULUS  [VII 

So  far  we  have  made  no  use  of  the  fact  that/(#)  is  continuous. 
We  shall  now  show  that  j  =  J,  and  that  the  sums  s,  S  tend  to  the 
limit  J  when  the  points  of  division  xv  are  multiplied  indefinitely 
in  such  a  way  that  all  the  intervals  8V  tend  to  zero.  More  pre- 
cisely, we  shall  show  that,  given  any  positive  number  e,  it  is  possible 
to  find  8  so  that 

0^J-s<e,     0^S-J<e 

whenever  8V  <  8  for  all  values  of  v. 

There  is,  by  Theorem  II  of  §  106,  a  number  8  such  that 
Mv  —  mv<  ej(b-a), 
whenever  every  8„  is  less  than  8.     Hence 

S-s  =  t(Mv-m„)8v<e. 
But  S-s  =  (S-J)  +  (J-j)  +  (j  -  s) ; 

and  all  the  three  terms  on  the  right-hand  side  are  positive,  and 
therefore  all  less  than  e.  As  J  —  j  is  a  constant,  it  must  be  zero. 
Hence  j  =  J  and  0  ^  j  —  s<  e,  0^S  —  J<e,  as  was  to  be  proved. 

We  define  the  area  of  PpqQ  as  being  the  common  limit  of  s  and 
S,  that  is  to  say  J.  It  is  easy  to  give  a  more  general  form  to  this 
definition.     Consider  the  sum 

<T  =  $f,8l/ 

where  /„  denotes  the  value  of  f(x)  at  any  point  in  8V.  Then  <r 
plainly  lies  between  s  and  S,  and  so  tends  to  the  limit  J  when  the 
intervals  8V  tend  to  zero.  We  may  therefore  define  the  area  as 
the  limit  of  o\ 

157.  The  definite  integral.  Let  us  now  suppose  that  f{x) 
is  a  continuous  function,  so  that  the  region  bounded  by  the  curve 
y  =f(x),  the  ordinates  x  =  a  and  x  =  b,  and  the  axis  of  x,  has  a 
definite  area.  We  proved  in  Ch.  VI,  §  145,  that  if  F(x)  is  an 
'  integral  function  '  oif(x),  i.e.  if 

F'(x)=f(x),     F(x)=\f(x)dx, 

then  the  area  in  question  is  F(b)  —  F(a). 

As  it  is  not  always  practicable  actually  to  determine  the  form 
of  F  (x),  it  is  convenient  to  have  a  formula  which  represents  the 
area  PpqQ  and  contains  no  explicit  reference  to  F  (x).    We  shall 

(PpqQ)=ff(*)dx. 


156,  157]      ADDITIONAL   THEOREMS   IN    THE    CALCULUS  287 

The  expression  on  the  right-hand  side  of  this  equation  may 

then  be  regarded  as  being  defined  in  either  of  two  ways.    We 

may  regard  it  as  simply  an  abbreviation  for  F(b)  —  F(a),  where 

F(x)  is  some  integral  function  of  f(x),  whether  an  actual  formula 

expressing  it  is  known  or  not ;  or  we  may  regard  it  as  the  value  of 

the  area  PpqQ,  as  directly  defined  in  §  156. 

rb 
The  number  f{x)dx 

J  a 

is  called  a  definite  integral;  a  and  b  are  called  its  lower  and 
upper  limits;  f{%)  is  called  the  subject  of  integration  or 
integrand;  and  the  interval  (a,  b)  the  range  of  integration. 

The  definite  integral  depends  on  a  and  b  and  the  form  of  the 
function  f{x)  only,  and  is  not  a  function  of  x.  On  the  other  hand 
the  integral  function 

F(x)  =  ff(x)dx 

is  sometimes  called  the  indefinite  integral  off(x). 

The  distinction  between  the  definite  and  the  indefinite  integral  is  merely 

fb 
one  of  point  of  view.     The  definite  integral    /    f(x)dx=F(b)-F(a)  is  a 

function  of  b,  and  may  be  regarded  as  a  particular  integral  function  of  f(b). 
On  the  other  hand  the  indefinite  integral  F(x)  can  always  be  expressed  by 
means  of  a  definite  integral,  since 


F(.v)  =  F(a)+fXf(t)dt. 

J  a 


But  when  we  are  considering  '  indefinite  integrals '  or  { integral  functions ' 
we  are  usually  thinking  of  a  relation  between  two  functions,  in  virtue  of  which 
one  is  the  derivative  of  the  other.  And  when  we  are  considering  a  '  definite 
integral '  we  are  not  as  a  rule  concerned  with  any  possible  variation  of  the 
limits.     Usually  the  limits  are  constants  such  as  0  and  1 ;  and 

lf(x)dx  =  F(l)-F(0) 


i: 


is  not  a  function  at  all,  but  a  mere  number. 

It  should  be  observed  that  the  integral   /   f(t)dt,  having  a  differential 

J  a 

coefficient  f(x),  is  a  fortiori  a  continuous  function  of  a*. 

Since  1/x  is  continuous  for  all  positive  values  of  x,  the  investigations  of 
the  preceding  paragraphs  supply  us  with  a  proof  of  the  actual  existence  of  the 
function  logx,  which  we  agreed  to  assume  provisionally  in  §  128. 


288 


ADDITIONAL   THEOREMS   IN   THE    CALCULUS 


[VII 


Fig.  49. 


158.    Area  of  a  sector  of  a  circle.  The  circular  functions. 

The  theory  of  the  trigonometrical  functions  cos  x,  sin  x,  etc.,  as 
usually  presented  in  text-books  of  elementary  trigonometry,  rests 
on  an  unproved  assumption.  An  angle  is  the  configuration  formed 
by  two  straight  lines  OA,  OP;  there  is  no  particular  difficulty  in 
translating  this  '  geometrical '  definition  into  purely  analytical 
terms.  The  assumption  comes  at  the  next  stage,  when  it  is  assumed 
that  angles  are  capable  of  numerical  measurement,  that  is  to  say 
that  there  is  a  real  number  x  associated 
with  the  configuration,  just  as  there  is 
a  real  number  associated  with  the  region 
PpqQ  of  Fig.  47.  This  point  once  ad- 
mitted, cos  x  and  sin#  may  be  defined 
in  the  ordinary  way,  and  there  is  no 
further  difficulty  of  principle  in  the 
elaboration  of  the  theory.  The  whole 
difficulty  lies  in  the  question,  what  is  the 
x  which  occurs  in  cos  x  and  sin  x  ?  To  answer  this  question,  we 
must  define  the  measure  of  an  angle,  and  we  are  now  in  a  position 
to  do  so.  The  most  natural  definition  would  be  this:  suppose  that 
AP  is  an  arc  of  a  circle  whose  centre  is  0  and  whose  radius  is 
unity,  so  that  OA  =  OP  =  1.  Then  x,  the  measure  of  the  angle,  is 
the  length  of  the  arc  A  P.  This  is,  in  substance,  the  definition 
adopted  in  the  text-books,  in  the  accounts  which  they  give  of  the 
theory  of  '  circular  measure  '.  It  has  however,  for  our  present  pur- 
pose, a  fatal  defect;  for  we  have  not  proved  that  the  arc  of  a  curve, 
even  of  a  circle,  possesses  a  length.  The  notion  of  the  length  of  a 
curve  is  capable  of  precise  mathematical  analysis  just  as  much  as 
that  of  an  area;  but  the  analysis,  although  of  the  same  general 
character  as  that  of  the  preceding  sections,  is  decidedly  more 
difficult,  and  it  is  impossible  that  we  should  give  any  general 
treatment  of  the  subject  here. 

We  must  therefore  found  our  definition  on  the  notion  not  of 
length  but  of  area.  We  define  the  measure  of  the  angle  AOP  as 
twice  the  area  of  the  sector  AOP  of  the  unit  circle. 

Suppose,  in  particular,  that  OA  is  y  =  0  and  that  OP  is  y  =  mx, 
where  m  >  0.  The  area  is  a  function  of  m,  which  we  may  denote 
by  (f>  (m)..  If  we  write  //,  for  (1  +m2)--,  P  is  the  point  (/x,  m/u,),  and 


158]  AND   INTEGRAL   CALCULUS  289 

we  have 

</>  (m)  =  \m^  +      V(l  -  «2)  dec. 
Differentiating  with  respect  to  m,  we  find 

1  Cm      fa 

*(M)=2o+^)'  *(")ta*j.r+?- 

Thus  the  analytical  equivalent  of  our  definition  would  be  to  define 

arc  tan  m  by  the  equation 

[m    dt 
arc  tan  m  = 


oi+r 

and  the  whole  theory  of  the  circular  functions  could  be  worked  out 
from  this  starting  point,  just  as  the  theory  of  the  logarithm  is 
worked  out  from  a  similar  definition  in  Ch.  IX.   See  Appendix  III. 

Examples  LXIII.    Calculation  of  the  definite  from  the  indefinite 
integral.     1.    Show  that 


/, 


6      7       bn  * l  -  a"  + 1 

xn  dx  = , 

t  n  +  1 


[i  1 

and  in  particular  that  /    xndx  = 

.    F  Jo  »+l 

fb  .      sin  mb-  sin  ma         fb  . 

2.  /    cos  mxdx= .        j    si 

J  a  m  J  a 

3.  /     — -L- 7,  =  arc  tan  b  —  arc  tan  a,  \    - — — s=i7r. 


&  .       sin  m&- sin  ma  /"ft  .  .      cos  ma- cos m& 

sin  mx  ax = — , 


/: 


[There  is  an  apparent  difficulty  here  owing  to  the  fact  that  arc  tan  x  is  a 
many  valued  function.  The  difficulty  may  be  avoided  by  observing  that,  in 
the  equation 

"*    dt 

„  =  arc  tan  x, 
ol  +  «- 

arc  tan  x  must  denote  an  angle  lying  between  -\n  and  \n.  For  the  integral 
vanishes  when  x  =  0  and  increases  steadily  and  continuously  as  x  increases. 
Thus  the  same  is  true  of  arc  tan  x,  which  therefore  tends  to  \n  as  x-*-x>. 
In  the  same  way  we  can  show  that  arc  tan  a?-*-— J  w  as  x-*-  —  cc  .  Similarly, 
in  the  equation 

dt 


i: 


-^  =  arc  sin  x, 


ox/(l-6 

where   —  1<jk<1,  arc  sin  x  denotes  an  angle  lying  between   -\ir  and  \n. 
Thus,  if  a  and  b  are  both  numerically  less  than  unity,  we  have 

'b      dx 


/. 


„.  =  arc  sin  b  —  arc  sin  a.l 
aJ(l-x2) 


fl       dx        _    2tt  f1       dx  it 

J  0  l-#+u;2-373'      j0l+x+xT2~3s/3 


u,  19 


290  ADDITIONAL   THEOREMS   IN   THE   DIFFERENTIAL  [VII 

5       J —  -  =  -—. —  if  —  7r<a<7r,  except  when  a  =  0,  when  the 

J  0l+2xcosa  +  x2      2  sin  a 

value  of  the  integral  is  ^,  which  is  the  limit  of  \a  coseca  as  a^~0. 

6.      f  J(\-x2)dx=liv,      j^(a2-x2)dx=^ira2     (a>0). 
Jo  Jo 


Jo  a- 


,  if  a  >  |  b  | .     [For  the  form  of  the  indefinite 


i  +  b  cos  x  s/(a2  -  b2) 
integral  see  Exs.  till.  3,  4.  If  |  a  |<|  b  |  then  the  subject  of  integration  has  an 
infinity  between  0  and  ir.  What  is  the  value  of  the  integral  when  a  is 
negative  and  -  a>\  b  \  ?] 

if  a  and  b  are  positive.      What  is  the 


J  0    a2  cos2  x  +  b2  sin2  x      2ab ' 
value  of  the  integral  when  a  and  b  have  opposite  signs,  or  when  both  are 
negative  ? 

9.  Fourier's  integrals.     Prove  that  if  m  and  n  are  positive  integers  then 

I     cos  rax  sin  ?j.i-  dx 
Jo 

is  always  equal  to  zero,  and 

/"27T  r2TT 

I      cos  wi.r  cos  «A*  dx,       I      sin  m.r  sin  nx  dx 

Jo  Jo 

are  equal  to  zero  unless  m  =  n,  when  each  is  equal  to  n. 

10.  Prove  that  I    cos  mx  cos  jza  cfo;  and  I    sin  )m  sin  n#  dx  are  each  equal 

Jo  Jo 

to  zero  except  when  m  =  n,  when  each  is  equal  to  £n- ;  and  that 

/""■  •  t  2ra  /"^  . 

|    cos  ??i.r  sin  nx  dx  =  -= r, ,  cos  m#  sm  nx  dx = 0, 

Jo  ?i2-»i-'       J0 

according  as  %  —  m  is  odd  or  even. 

159.  Calculation  of  the  definite  integral  from  its  defini- 
tion as  the  limit  of  a  sum.  In  a  few  cases  we  can  evaluate  a 
definite  integral  by  direct  calculation,  starting  from  the  definitions 
of  §§  156  and  157.  As  a  rule  it  is  much  simpler  to  use  the 
indefinite  integral,  but  the  reader  will  find  it  instructive  to  work 


f6 
Examples  LXIV,     1.     Evaluate  I    xdx  by  dividing  (a,  b)  into  n  equal 

J  a 

rta  by  the  points  of  division  a=x0,  xu  x2,  ...,  xn=b,  and  calculating  the 
lit  as  ft-*- oo  of 

fo  -  a?o)/(tfo)  +  (*2-  *i)/0*i)  +  ..;  +  (4?»-#»_1)/0f»_i). 


158-160]  AND   INTEGRAL   CALCULUS  291 

[This  sum  is 

^[a+(.+^)  +  (a+2^)+...+{(l+()(.1)^j] 

=  ^-['",+^f<1+2-,---+("-1K]=(i'-«){«+(''-<»)^li}. 

which  tends  to  the  limit  | (b2- a2)  as»-*x.     Verify  the  result  by  graphical 
reasoning.] 

2.  Calculate  /    x2dx  in  the  same  way. 

J  a 

3.  Calculate    /   xdx,  where  0<a<  b,  by  dividing  (a,  6)  into  %  parts  by 

J  a 
the  points  of  division  a,  ar,  ar2,  ...  a?-"-1,  arn,  where  rn=b\a.     Apply  the  same 

method  to  the  more  general  integral  /   xm  dx. 

J  a, 

4.  Calculate    /    cos  mx  dx  and    /    sin  mx  dx  by  the  method  of  Ex.  1. 

J  a  J  a 

»— 1         1 

5.  Prove  that  n  2    —. „  -*-irr  as  n-*-cc  . 

[This  follows  from  the  fact  that 

n  n  n  n~1     (I  In) 


n2     n2  +  l2  n2  +  (n-l)2    r=0  l+(?-/?t)2' 

r\    dx 

which  tends  to  the  limit  I „  as  n  -*•  oo  ,  in  virtue  of  the  direct  definition 

Jo 1+ff- 
of  the  integral.] 

6.     Prove  that  -*sV(»'- r2)-*-^.     [The  limit  is  /"V( I-*'2) dx."\ 

n"  r=0  J  0 

160.     General  properties  of  the  definite   integral.     The 

definite   integral    possesses    the   important   properties    expressed 
by  the  following  equations.* 

(1)  \bf(x)dx  =  -\af{x)dx. 

J  a  J  b 

This  follows  at  once  from  the  definition  of  the  integral  by  means  of  the 
integral  function  F(x),  since  F(b)- F(a)= -{F(a)-F(b)}.  It  should  be 
observed  that  in  the  direct  definition  it  was  presupposed  that  the  upper 
limit  is    greater    than    the    lower ;    thus   this   method   of    definition  does 

not  apply  to  the  integral  |    f(x)dx  when  a<b.     If  we  adopt  this  definition 

Jb' 
as  fundamental  we  must  extend  it  to  such  cases  by  regarding  the  equation  (1) 
as  a  definition  of  its  right-hand  side. 

*  All  functions  mentioned  in  these  equations  are  of  course  continuous,  as  the 
definite  integral  has  been  defined  for  continuous  functions  only. 

19—2 


292  ADDITIONAL   THEOREMS   IN   THE   DIFFERENTIAL  [VII 

(2)  \af(x)dx=0. 
•  a 

(3)  jf(x)dx+j   f(x)dx  =      f(x)dx. 

J  a  J  b  -'a 

(4)  (    kf(x)dx  =  k(  f{x)dx. 

(5)  I    {/OO  +  <£  0)}  dx=\  f(x)  dx+f<f>  (x)  dx. 

Ja  J  a  J  a 

The  reader  will  find  it  an  instructive  exercise  to  write  out  formal  proofs 
of  these  properties,  in  each  case  giving  a  proof  starting  from  (a)  the  definition 
by  means  of  the  integral  function  and  (3)  the  direct  definition. 

The  following  theorems  are  also  important. 

(6)  Iff(x)  ^  0  when  a^x^b,  then  j   f(x)  dx  ^  0. 

J  a 

We  have  only  to  observe  that  the  sum  s  of  §  156  cannot  be  negative.  It 
will  be  shown  later  (Misc.  Ex.  41)  that  the  value  of  the  integral  cannot  be 
zero  unless  f(x)  is  always  equal  to  zero  :  this  may  also  be  deduced  from  the 
second  corollary  of  §  121. 

(7)  If  H  ^  f{x)  ^  K  when  a^x^b,  then 

H(b-a)^f  f(x)dx^K(b-a). 

J  a 

This  follows  at  once  if  we  apply  (6)  to/(.r)  -if  and  K-f(x). 

(8)  [/(„)  dx  =  (b-a)f(Z), 

J  a 

where  £  lies  between  a  and  b. 

This  follows  from  (7).  For  we  can  take  H  to  be  the  least  and  K  the 
greatest  value  of  f(x)  in  (a,  b).  Then  the  integral  is  equal  to  tj  (b  -  a),  where 
»?  lies  between  H  and  K.  But,  since  fix)  is  continuous,  there  must  be  a 
value  of  £  for  which  /(£)  =  >;  (§  100). 

If  F(x)  is  the  integral  function,  we  can  write  the  result  of  (8)  in  the  form 

F(b)-F(a)  =  (b-a)F'(£), 

so  that  (8)  appears  now  to  be  only  another  way  of  stating  the  Mean  Value 
Theorem  of  §  125.  We  may  call  (8)  the  First  Mean  Value  Theorem  for 
Integrals. 


[160]  AND   INTEGRAL   CALCULUS  293 

(9)  The  Generalised  Mean  Value  Theorem   for   inte- 
grals.   If  <f>  (x)  is  positive,  and  H  and  K  are  defined  as  in  (7),  then 

H  !    $  (x)  dx  S  i   f{x)  $  (x)  dx^xf   <£  (x)  dx; 

•la  J  a  J  a 

and  J   f(x)(f)(x)dx=f(^)  j    <f>(x)dx, 

J  a  J  a 

where  f  is  defined  as  in  (8). 

This  follows  at  once  by  applying  Theorem  (6)  to  the  integrals 

/    {f(x)-H}$(x)dx,      f  {K-f{x)}<\>{x)dx. 
J  a  J  a 

The  reader   should  formulate  for  himself  the  corresponding  result   which 
holds  when  <£  (x)  is  always  negative. 

(10)  The  Fundamental  Theorem  of  the  Integral  Cal- 
culus .     Th  e  function 

F(x)=(Xf(t)dt 

J  a 

has  a  derivative  equal  to  f{%). 

This  has  been  proved  already  in  §  145,  but  it  is  convenient  to 
restate  the  result  here  as  a  formal  theorem.  It  follows  as  a 
corollary,  as  was  pointed  out  in  §  157,  that  F  (x)  is  a  continuous 
function  of  x. 

Examples  LXV.      1.      Show,  by  means  of  the  direct  definition  of  the 
definite  integral,  and  equations  (1) — (5)  above,  that 

(i)       fa  <f>(x2)dx=2  [a(t>(x*)dx,       fa  x<j>(x2)dx=0; 

J  -a  JO  J   -a 

(ii)       I      0  (cos  x)  dx  =  I      (j)  (sin  x)  dx  =  ^  I    (f>  (sin  x)  dx ; 
Jo  Jo  "Jo 

(iii)       I       (f>  (cos23-)  dx=m  |     cf>  (cos2  x)  dx, 
Jo  Jo 

m  being  an  integer.     [The  truth  of  these  equations  will  appear  geometrically 

intuitive,  if  the  graphs  of  the  functions  under  the  sign  of  integration  are 

sketched.] 


f  7T  si  ji  73,/r 

2.  Prove  that  /      — dx  is  equal  to  rr  or  to  0  according  as  n  is  odd  or 

J  o    sin  x 

sn.   [Use  the  formula  (smnx)/(smx)  =  2  cos{(«-  1)  x}  +  2  cos{(?i-3).r}  +  ..., 
i  last  term  being  1  or  2  cos  x.~\ 

3.  Prove  that  I    sin  nx  cot  xdx  is  equal  to  0  or  Lo  it  according  as  n  is  odd 

Jo 


or  even. 


294  ADDITIONAL   THEOREMS    IN    THE    DIFFERENTIAL  [VII 

4.  If  (f>  (x)  =  a0  +  ai cos  x  +  b\ s™  x  +  a2 cos  2#  + . ,.  +  a„  cos  nx  +  6„  sin  nx, 
and  £  is  a  positive  integer  not  greater  than  n,  then 

f  2tt  /"  2n-  /"  2ff 

I      <f)(x)  dx=2Tra0,     I      coskx(f>  (x)  dx=nak,    I      sin«# 0  {x)dx=irbk. 
Jo  Jo  Jo 

If  k>n  then  the  value  of  each  of  the  last  two  integrals   is  zero.     [Use 

Ex.  lxiii.  9.] 

5.  If  (j)  (x) = a0  +  ax  cos  x  +  a2  cos  2x+  ...+an  cos  nx,  and  h  is  a  positive 
integer  not  greater  than  n,  then 

/(f)  (x)  dx=  7ra0,       I    cos  Lv(f)  (x)  dx—\tvak. 
o  Jo 

It  k>n  then  the  value  of  the  last  integral  is  zero.     [Use  Ex.  lxiii.  IO.j 

6.  Prove  that  if  a  and  b  are  positive  then  |      --5 5 ,„  .  _     =  — =-. 

J  0    er  cos^  x+ 6J  sin*  a;      a© 

[Use  Ex.  lxiii.  8  and  Ex.  1  above.] 

fb  fb 

7.  If/(.r)  ^<f>(x)  when  a  <  .r  <  6,  then  /    fdx  ^  I    <£efo. 

J  a  J  a 

8.  Prove  that 

/"*»■     .  ,  /"Jt  /"It  /"It 

0<|       smn  +  1xdx<         sinnxdx,       0<l       tan"  +  2  .r  c£r  <  I      tan".rrf#. 
Jo  Jo  Jo  Jo 

/1/2        dx 
— ^T  <  -524.     [The  first  inequality  follows 
0    v  ( 1  -  #   j 

from  the  fact  that  ,/(l  —  x2n)<l,  the  second  from  the  fact  that 

x/(i-^»)>v/(i-4] 

0v/(4-.^2  +  ^3)      b 

11.  Prove  that  (3^  +  8)/16<  1/^/(4 -»*+*•)<  1/^/(4 -3a?)   if  0<#<1, 
and  hence  that  A?<|    —r. ~<#. 

12.  Prove  that  -573 <  \      Ul    f 'g r-<-595.     [Put  x=l+u:  then  re- 

place  i  +  3u2  +  u3  by  2  +  4w2  and  by  2  + 3m2.] 

13.  If  a  and  <p  are  positive  acute  angles  then 


[*  dx 

Jo  V(l  -sin2 a 


0<  /n      „;^.^8^<- 


sin2.^)      >/(l  —  sin2  a  sin2  0)* 
If  a=(f)=^7r,  then  the  integral  lies  between  -523  and  •541. 

1    fb  I         f> 

14.  Prove  that        f(x)dx\  ^  /    |/(*)  |  dx. 

\  J  a  I         /» 

[If  a  is  the  sum  considered  at  the  end  of  §  156,  and  a  the  corresponding 
sum  formed  from  the  function  \f(x)  \,  then  |  <r  |  ^  cr'  ] 

15.  If  \f(x)  I  <  M,  then    f  f(x)  0  (*)  dx  \  ^  M  P  |  <f>  (us)  \  dx. 

J  a  J  a 

*  Exs.  9— la  are  taken  from  Prof.  Gibson's  Elementary  Treatise  on  the  Calculus. 


160,  161]  AND   INTEGRAL   CALCULUS  295 

161.     Integration    by    parts    and    by    substitution.     It 

follows  from  §  138  that 

\  "fix)  #  (x)  dx  =f(b)  <f>  (b)  -/(a)  0  (a)  -  [  /'  (x)  <f>  (*)  dx. 

This  formula  is   known  as  the    formula   for   integration    of  a 
definite  integral  by  parts. 

Again,  we  know  (§  133)  that  if  F(i)  is  the  integral  function  of 
/(«),  then 


J; 


Hence,  if  <£  (a)  =  c,  <f>  (b)  =  d,  we  have 

[7(0  dt=F(d)-F(c)  =  F{<f>  (b)}  -F{<f>  (a)}  =  [7(0  (*)}  </>'  («)  *d; 

which  is  the  formula  for  the  transformation  of  a  definite  integral 
by  substitution. 

The  formulae  for  integration  by  parts  and  for  transformation 
often  enable  us  to  evaluate  a  definite  integral  without  the  labour 
of  actually  finding  the  integral  function  of  the  subject  of  integra- 
tion, and  sometimes  even  when  the  integral  function  cannot  be 
found.  Some  instances  of  this  will  be  found  in  the  following 
examples.  That  the  value  of  a  definite  integral  may  sometimes 
be  found  without  a  knowledge  of  the  integral  function  is  only  to 
be  expected,  for  the  fact  that  we  cannot  determine  the  general 
form  of  a  function  F(x)  in  no  way  precludes  the  possibility  that 
we  may  be  able  to  determine  the  difference  F(b)  —  F(a)  between 
two  of  its  particular  values.  But  as  a  rule  this  can  only  be 
effected  by  the  use  of  more  advanced  methods  than  are  at 
present  at  our  disposal. 

Examples  LXVI.     1.     Prove  that 

fb  xf"  (x)  dx={bf  (b)  -f(b)}  -  {af  (a)  -/(a)}. 

fb 

2.  More  generally,  /    xm/(m  +  1)  (x)dx=F(b)  —  F(a),  where 

J  a 

F(x)  =  xmf(m)  (x)  -  mxm  ~  1pn  ~  l)x + m  (m  - 1)  xm  -2f(m-2)x-...  +  (-\)mm !/(.?). 

3.  Prove  that 

I   arc  sm.xdx=\TT  —  1,       /    a?aarctana;cfo?=j7r—  \. 

Jo  Jo 


296  ADDITIONAL   THEOREMS    IN    THE    DIFFERENTIAL  [Vll 

4.     Prove  that  if  a  and  b  are  positive  then 

i'r      xcosxs'mxdx  n 


l„ 


0    (a2  cos2  x  +  b'A  sin2  x)%      \ab'1  (a  +  b)' 
[Integrate  by  parts  and  use  Ex.  lxiii.  8.] 

5.  If   Mx)=[Xf(t)dt,    f3{x)=  fXfi(t)dt,...,Mx)=( xfk-,{t)dt, 

Jo  Jo  Jo 

then  /*  (*)  =  (1  ~  iy]  //  (0  (*  -  *)*" 1  dt. 

[Integrate  repeatedly  by  parts.] 

6.  Prove  by  integration  by  parts  that  if  um>n=  I    xm(l  —  x)ndx,  where  m 

J  o 
and  n  are  positive  integers,  then  (m+n  +  l)  ftm,n:=?lMm,n~i>  and  deduce  that 

m  !  ft  ! 

7.  Prove  that  if  un=  \      ta,nnxdx  then    «„  +  wm_2  =  l/(ft  -  1).      Hence 

evaluate  the  integral  for  all  positive  integral  values  of  ft. 
[Put  tann#=tann-2a?(sec2#—  1)  and  integrate  by  parts.] 

8.  Deduce  from  the  last  example  that  un  lies  between  l/{2(n  — 1)}  and 
l/{2(n+l)}. 

9.  Prove  that  if  un=  I      sinn  x  dx   then   un=  {(ft-  l)/ft}«„_2-      [Write 

J  o 
sin™-1  x  sin  x  for  sinn.r  and  integrate  by  parts.] 

10.  Deduce  that  un  is  equal  to 

2.4.6..(ft-l)       ,     1.3.5..(ft-l) 
3.5. 7. .ft      '     *"      2. 4. 6.. ft      ' 
according  as  n  is  odd  or  even. 

11.  The  Second  Mean  Value  Theorem.  If  f(x)  is  a  function  of  x 
which  has  a  differential  coefficient  of  constant  sign  for  all  values  of  x  from 
x = a  to  x=b,  then  there  is  a  number  £  between  a  and  b  such  that 

f  b  /(#)  0  (*')  dx=f(a)  f  *  0  (j?)  dx+f{b)  t b  0  (a.-)  efc. 

Jo  .'a  J  f 

[Let  fX(f)(t)dt  =  ^{x).     Then 

f  ^  /(*)  0  (*)  dx=  f   f{x)  *'  (#)  <&=/(&)  $  (6)  -  [  b  f  (x)  *  (a;)  dx 
J  a  J  a  J  a 

=/(&)*(&)-*(£)  fV(*)«ki 

J  a 

by  the  generalised  Mean  Value  Theorem  of  §  160  :  i.e. 


1 


bf(x)cj>(x)dx=f(b)<l>(b)  +  {f(a)-f(b)}4>(£), 


which  is  equivalent  to  the  result  given.] 


161]  AND  INTEGRAL  CALCULUS  297 

12.    Bonnet's  form  of  the  Second  Mean  Value  Theorem.     If/'  (x)  is 
of  constant  sign,  and/(6)  and /(a) -/(ft)  have  the  same  sign,  then 


I   f(x)<$>(x)dx=f(a)  I     cji(x)dx, 
J  a  J  a 


where  X  lies  between  a  and  b.  [For  f(b)$(b)  +  {f(a)-f(b)}4>(£)  =  pf(a), 
where  fi  lies  between  <J>  (£)  and  *  (5),  and  so  is  the  value  of  *  (x)  for  a  value 
of  x  such  as  X.     The  important  case  is  that  in  which  0?gf(b)^f(x)g^f(a).] 

Prove  similarly  that  if  f(a)  and/ (6) -/(a)  have  the  same  sign,  then 

J  V(a-)  <£  (*)  dW(&)  f  V  (*)  <**. 

where  X  lies  between  a  and  6.  [Use  the  function  ^  (£)  =  /  <£  (x)  dx.  It 
will  be  found  that  the  integral  can  be  expressed  in  the  form 

The  important  case  is  that  in  which  0^/(a)^/(.r)^/(6).] 

13.  Prove   that  j  /       -dx\<-jr    if  A">A~>0.      [Apply  the    first 

I J  X      x  A. 

formula  of  Ex.  12,  and  note  that  the  integral  of  sin  x  over  any  interval  what- 
ever is  numerically  less  than  2.] 

14.  Establish  the  results  of  Ex.  lxv.  1  by  means  of  the  rule  for  sub- 
stitution. [In  (i)  divide  the  range  of  integration  into  the  two  parts  ( -  a,  0), 
(0,  a),  and  put  x=  —y  in  the  first.  In  (ii)  use  the  substitution  x  =  \tt  —  y  to 
obtain  the  first  equation :  to  obtain  the  second  divide  the  range  (0,  n)  into 
two  equal  parts  and  use  the  substitution  x  =  ^rr+y.  In  (iii)  divide  the  range 
into  m  equal  parts  and  use  the  substitutions  x=?r+y,  x=2n+y,  ....] 

15.  Prove  that  I   F  (x)  dx  =  I    F(a  +  b-  x)  dx. 

J  a  J  a 

16.  Prove  that  I     cosm  xs\i\mxdx=2~m  j      cosmxdx. 

Jo  Jo 

17.  Prove  that   I    x<p  (siu.x)dx=^n  I     <ft  (sin  x)dx.     [Put  x=7r—y.] 

Jo  Jo 


I 


18.  Prove  that  /     — „-  dx=\n'1. 

Jo    l+COS^ 

19.  Show    by   means   of   the  transformation  x  =  a  cos2  6  +  b  sin2  6  that 
b 

y/{(x  —  a)(b  —  x)}  dx  =  \rt(b  —  df. 


20.     Show  by  means  of  the  substitution  (a  +  b  cos  x)  (a -b  cos  y)  =  a2  —  b2 
that 

r (a  +  b cos x) -» dx=  (a2 - 62) " <n ~ »  J    (a- b  cos y)n~l dy, 

when   n   is  a  positive  integer  and  a>\b\,  and  evaluate  the  integral  when 
?i  =  l,  2,  3. 


298  ADDITIONAL   THEOREMS   IN    THE    DIFFERENTIAL  [VII 

21.     If  m  and  n  are  positive  integers  then 

f h {x - a)m (b - x)« dx={b- «)«  +  »  + »  .    OT^'      . 
J  a  (m+»  +  l)l 

[Put  #= a  +  (&  -  a)  y,  and  use  Ex.  6.] 

162.  Proof  of  Taylor's  Theorem  by  Integration  by 
Parts.  We  shall  now  give  the  alternative  form  of  the  proof  of 
Taylor's  Theorem  to  which  we  alluded  in  §  147. 

Let/(#)  be  a  function  whose  first  n  derivatives  are  continuous, 
and  let 

Fn  («)  -/<&)  -/(•)  -  (b  -  *)f  («)-.»-  (\~*i)if{n71]  (•> 


and 


Then  *V  («)  -  -  (-^^/(n)  (4 


so 


^  (a)  =  Fn  (b)  -  \h  Fn  (a)  dx  =  -—!— !  f\b-  •)«/»  (x)  dx. 

.'a  \"        *■)  -J  a 

If  now  we  write  a  +  h  for  b,  and  transform  the  integral  by  putting 
cc  =  a  +  th,  we  obtain 

f(a  +  h)  =/(a)  +  hf  (a)  +  ...  +  £L.{f*-V,  (a)  +  Rn  . . .(1), 

where  Rn  =,    h\,j\l- t)n-*f™  (a  +  th) dt     (2). 

Now,  if  p  is  any  positive  integer  not  greater  than  n,  we  have, 
by  Theorem  (9)  of  §  160, 

I"1  (1  -  0n-1/(n)  (a  +  th)  dt=  [  (1  -  t)n~P(l  -  ty~lfw  (a  +  th)  dt 
Jo  -o 

=  (i  -  ey-pp*  (a  +  eh)  (\i-  ty-*dt, 

Jo 
where  0  <  6  <  1.     Hence 

En  = p(n-l)l (3)" 

If  we  take  p  =  n  we  obtain  Lagrange's  form  of  JB„  (§  148).     If 
on  the  other  hand  we  take  p  =  1  we  obtain  Cauchy's  form,  viz. 

B»" 5TTi)l (4)  • 

*  The  method   used   in   §   147   can   also  be  modified  so  as  to   obtain  these 
alternative  forms  of  the  remainder. 


161-164]  AND   INTEGRAL   CALCULUS  299 

163.    Application  of   Cauchy's  form  to  the  Binomial    Series.     If 

f(x)  =  (l  +  x)m,  where  m  is  not  a  positive  integer,  then  Cauchy's  form  of  the 
remainder  is 

_  m(m-I)...(m-n+l)  (l-0)"-1;rn 

n~  1.2...(n-l)  (l  +  8x)n~m  ' 

Now  (1  — 0)1(1 +  dx)  is  less  than  unity,  so  long  as   —  \<x<l,  whether 

x  is  positive   or  negative;  and  (\  +  8x)m~1  is  less  than  a  constant  K  for 

all  values  of  n,  being  in  fact  less  than  (14-|#  |)"1-1   if  ?/i>l   and   than 

(1  -  |  x  |  )m_  x  if  m  <  1      Hence 

I    r>      i  tt  ,  I   I  fin  —  lM  I  I 

\Rn\<A\m\^n_l^\x"\  =  Pn, 

say  But  pn-*-0  as  n-^cc  ,  by  Ex.  xxvu.  13,  and  so  Rn-*-0.  The  truth  of  the 
Binomial  Theorem  is  thus  established  for  all  rational  values  of  m  and  all 
values  of  x  between  —  1  and  1.  It  will  be  remembered  that  the  difficulty  in 
using  Lagrange's  form,  in  Ex.  lvi.  2,  arose  in  connection  with  negative 
values  of  x. 

164.     Integrals  of  complex  functions  of  a  real  variable. 

So  far  we  have  always  supposed  that  the  subject  of  integration  in 
a  definite  integral  is  real.  We  define  the  integral  of  a  complex 
function  f(x)  =  ^r  (x)  +  ity  (x)  of  the  real  variable  x,  between  the 
limits  a  and  b,  by  the  equations 

rb  rb  rb  rb 

f{x)dx=-\    {</>(#)  +  ity  (x)}  dx  =  I   (f)(x)dx  +  il   yjr(x)dx; 

J  a  J  a  J  a  J  a 

and  it  is  evident  that  the  properties  of  such  integrals  may  be 
deduced  from  those  of  the  real  integrals  already  considered. 

There  is  one  of  these  properties  that  we  shall  make  use  of 
later  on.     It  is  expressed  by  the  inequality 

\hf(x)dx    ^!b\f(x)\dx     (1)*. 

J  a  J  a 

This  inequality  may  be  deduced  without  difficulty  from  the 
definitions  of  §§  156  and  157.  If  Bv  has  the  same  meaning  as  in 
§  156,  <f>v  and  yfrv  are  the  values  of  <f>  and  ^  at  a  point  of  8V,  and 
fv  =  (f)v  +  iyjrv,  then  we  have 

rb  rb  '    rb 

fdx  =  I    (pdx  +  i  I    tydx  =  lim  S  <£„  8V  +  i  lim  S  y\rv  Sv 

J  a  J  a  J  a 

=  lim  2  ((/>„  +  iyfrv)  S„  =  lim  £/„§„, 
and  so  I   fdx  =  |  lim  2/„o\,  |  =  lim  |  2/„o\,  | ; 

J  a 
*  The  corresponding  inequality  for  a  real  integral  was  proved  in  Ex.  lxv.  14. 


300  ADDITIONAL   THEOREMS    IN    THE    DIFFERENTIAL  [VII. 

while  I    \f\dx  =*\im  t\fv  \  K- 

J  a 

The  result  now  follows  at  once  from  the  inequality 

|2/.8,|rf|/,|a.. 

It  is  evident  that  the  formulae  (1)  and  (2)  of  §  162  remain 
true  when  f  is  a  complex  function  <f>  +  iy}r. 

MISCELLANEOUS   EXAMPLES   ON   CHAPTER  VII. 

1.  Verify  the  terms  given  of  the  following  Taylor's  Series : 

(1)  ta,nx  =  x  +  ^x3+^sx5  +  ..., 

(2)  secz  =  l+|#2  +  Wj.rt  +  ..., 

(3)  xcosecx=l+^x2  +  5^xi+ ..., 

(4)  xcotx=l  —  ^x2  —  ^sxA—.... 

2.  Show  that  if  f(x)  and  its  first  n  +  2  derivatives  are  continuous,  and 
f(n  + !)  (0)  =t=  0,  and  6n  is  the  value  of  6  which  occurs  in  Lagrange's  form  of  the 
remainder  after  n  terms  of  Taylor's  Series,  then 

_J_      n__  f/(»  +  2>(0)        } 

n ~  n  + 1  +  2  (»  + 1  )2  (n  +  2)  |/(n  +  J)  (0) +  **}  ^ 

where  fx-a-0  as  z-»-0.     [Follow  the  method  of  Ex.  lv.  12.] 

3.  Verify  the  last  result  when  / O)  =  1/(1+  x).    [Here  ( 1  +  8nx)n  + 1  =  1  +.r. ] 

4.  Show  that  if  f{x)  has  derivatives  of  the  first  three  orders  then 

fib)  =/(«)  +|(6-  a)  {/'  (a)  +/'  (6)}  -  &  (6  -  a)*/'"  (a), 
where  a  <  a  <  b.     [Apply  to  the  function 
/ (*)  -/(«)  -*(*-«)  {/'  (a)  +/"  (*)} 


^-aJ  [/ W  -/(«)  -* (6 - «)  {/'  («)  +/'  (&)}] 
arguments  similar  to  those  of  §•  147.] 

5.  Show  that  under  the  same  conditions 

f(b)  =/  (a)  +  (b  -  a)f  ft  (a  +  6)}  +  ^  (6  -  a)*/'"  (a). 

6.  Show  that  if/(#)  has  derivatives  of  the  first  five  orders  then 
f(b)=f(a)  +  Ub-a)[f'(a)+f'(b)  +  4f'{h(a  +  b)}]-^\u(b-aff^(a). 

7.  Show  that  under  the  same  conditions 

/(&)=/(«)  +  l(&  "«){/(«)  +/(&)}  -  A  (b  -  af{f'\b)  -/" (a)}  +  7Js (6  -  a)«/(5)(a) 


AND    INTEGRAL    CALCULUS 


301 


(i) 


Establish  the  formulae 

9  (a)    g(b) 
where  /3  lies  between  a  and  6,  and 

!./(«)    /(&)    /(c) 

(ii)     |  #(a)    #(&)    gr(c) 

|  A  (a)    A  (b)    A(c) 


■■{b~a) 


/(«)    /'(/3) 


=  £(6-c)(c-a)(a-6) 


/(«)   /'(/3)    /"(y) 

.</(«)     ^'(/3)     <?"(y) 
A  (a)     A' (/3)     A"(>) 


where  /3  and  y  lie  between  the  least  and  greatest  of  a,  b,  c.     [To  prove  (ii) 
consider  the  function 


<{>(x)  = 


#(«)     0(6)     ^(*) 
A  (a)     A  (6)    A(j?) 


(a— a)  (x  —  b) 

(c-a)  (c—  b) 


/(a)    /(&)    f(c) 

9(<*)  9(b)   g(fi) 

A  (a)     A  (6)     A  (c) 


which  vanishes  when  x=a,x=b,  and  x=c.  Its  first  derivative,  by  Theorem  B 
of  §  121,  must  vanish  for  two  distinct  values  of  x  lying  between  the  least  and 
greatest  of  a,  b,  c;  and  its  second  derivative  must  therefore  vanish  for  a  value 
y  of  x  satisfying  the  same  condition.     We  thus  obtain  the  formula 


/(«)     fd>)    /(«) 

9(a)    9(b)    9(c) 
h  (a)     A  (6)     A  (c) 


■■%(c-a)(c-rb) 


/(«)     f(P)   f"(y) 

gifl)  g(b)  g"{y) 

A  (a)     A  (6)     A"(y) 


The  reader  will  now  complete  the  proof  without  difficulty.] 

9.  If  F{x)  is  a  function  which  has  continuous  derivatives  of  the  first  n 
orders,  of  which  the  first  n—  1  vanish  when  x  =  0,  and  A  ^  F(")  (x)  g  B  when 
0  <  #  ^  A,  then  A  {xn\n  \)^F{x)^B  (xnjn  !)  when  0^x<  A. 

Apply  this  result  to 

f(x)-f(0)-xf(0)-...- 

and  deduce  Taylor's  Theorem. 


(n-1) 


i/^KO), 


10.     If  Ah(f>(x)  =  (j)(x)-(j)(x+h),  Ah2(f>(x)  =  Ah{Ah(f)(x)},  and  so  on,  and 
0  (a?)  has  derivatives  of  the  first  n  orders,  then 


Ahncf>(x)  =  2(-i; 


0(*  +  M)  =  (-A)»0<»)(& 


where  £  lies  between  x  and  x  +  nh.  Deduce  that  if  0(")  (#)  is  continuous  then 
{Ahn<i>{x)}lhn^-{  —  l)n  <£(")(#)  as  A^»0.  [This  result  has  been  stated  already 
when  »  =  2,  in  Ex.  lv.  13.] 


11.     Deduce  from  Ex.  10  that 


'  Ahnxm  -a-  m  (to  —  1) . . .  (to  -  n  + 1)  hn  as 


x  -*-  oo ,   m  being  any  rational   number  and  n  any   positive   integer, 
particular  prove  that 

X  *JX  Wx-2  J(x  +  I)  +  sJ(x  +  2)}-: J. 


In 


302  ADDITIONAL  THEOREMS   IN   THE   DIFFERENTIAL  [VII 

12.  Suppose  that  y  =  4>(x)  is  a  function  of  x  with  continuous  derivatives 
of  at  least  the  first  four  orders,  and  that  <f>  (0)  =  0,  cf>'  (0)  =  1,  so  that 

y  =  <p  (x)  =x  +  a2x2  +  a3x3  +  (a4  +  ex)  x*, 
where  ex->0as  x^-0.     Establish  the  formula 

x  =  f  (y)  =y  -  a2y2  +  (2a22  -  a3)  y3  -  (5a23  -  5tx2a3 +a4  +  fy)  2/1, 
where  ftf-*-0  &s  y-*-0,  for  that  value  of  x  which  vanishes  with  y;  and  prove 
that 

<t>(x)^(x)-x2  a      . 
#*  2 

as  x  -*-  0. 

13.  The  coordinates  (|,  q)  of  the  centre  of  curvature  of  the  curve  x=f(t), 
y=F(t),  at  the  point  (x,  y),  are  given  by 

-  (-1  -  *)/y  -  (■?  -y)/*' = -X2  +y'2)Kxy  -  x"y')  \ 

and  the  radius  of  curvature  of  the  curve  is 

(x'2+y'Zfi2l(x'y"  -x"y'), 
dashes  denoting  differentiations  with  respect  to  t. 

14.  The  coordinates  (£,  »;)  of  the  centre  of  curvature  of  the  curve 
27a?/2  =  4a*3,  at  the  point  (x,  y),  are  given  by 

3a(£+x)  +  2x2=0,     T)=4y  +  (day)lx.  (Math.  Trip.  1899.) 

15.  Prove  that  the  circle  of  curvature  at  a  point  (x,  y)  will  have  contact 
of  the  third  order  with  the  curve  if  (l+yi2)y3  =  3yiy22  at  that  point.  Prove 
also  that  the  circle  is  the  only  curve  which  possesses  this  property  at  every 
point ;  and  that  the  only  points  on  a  conic  which  possess  the  property 
are  the  extremities  of  the  axes.     [Cf.  Ch.  VI,  Misc.  Ex.  10  (iv).] 

16.  The  conic  of  closest  contact  with  the  curve  y = ax2  +  bx3  +  cxA  +  ...+  kxn, 
at  the  origin,  is  a3y  —  aix2  +  a2bxy+(ac  —  b2)y2.  Deduce  that  the  conic  of 
closest  contact  at  the  point  (£,  rj)  of  the  curve  y=f(x)  is 

l8r,j>T=9.T,2*(x-$)*+Gto%(x-$)  T+^^-Ak2)  T2, 

where  T=(^-t])-Vl(x-$).  (Math.  Trip.  1907.) 

17.  Homogeneous  functions.*  If  u  =  xnf(yjx,  z/x,  ...)  then  u  is  un- 
altered, save  for  a  factor  X™,  when  x, y,  z,  ...  are  all  increasedain  the  ratio  X  :  1. 
In  these  circumstances  u  is  called  a  homogeneous  function  of  degree  n  in  the 
variables  x,  y,  z,  ....     Prove  that  if  u  is  homogeneous  and  of  degree  n  then 

du        du       du 

ox    J  oy        oz 

This  result  is  known  as  Euler's  Theorem  on  homogeneous  functions. 

18.  If  u  is  homogeneous  and  of  degree  n  then  dujdx,  du/dy,  ...  are 
homogeneous  and  of  degree  n—1. 

*  In  this  and  the  following  examples  the  reader  is  to  assume  the  continuity  of 
all  the  derivatives  which  occur. 


AND  INTEGRAL  CALCULUS  303 

19.  Let  f(x,  y)  =  0  be  an  equation  in  x  and  y  {e.g.  xn+yn-x=0),  and  let 
F{x,  y,  z)=0  be  the  form  it  assumes  when  made  homogeneous  by  the  intro- 
duction of  a  third  variable  z  in  place  of  unity  (e.g.  xn+yn-xzn-1  =  0).  Show 
that  the  equation  of  the  tangent  at  the  point  (£,  t))  of  the  curve  f(x,  y)  =  0  is 

xFs+yFr,  +  zFi=0, 

where  F%,  Fv,  F$  denote  the  values  of  Fx,  Fy,  Fz  when  x=$,y  =  r,,  z=(=\. 

20.  Dependent  and  independent  functions.  Jacobians  or  functional 
determinants.  Suppose  that  u  and  v  are  functions  of  x  and  y  connected  by 
an  identical  relation 

<t>(u,  v)=0 (1). 

Differentiating  (1)  with  respect  to  x  and  y,  we  obtain 

3<£  du     d(f)  dv  30  du     d<f>  dv 

du  dx      dv  dx~   '      du  dy      dv  dy~       *"'' 


and,  eliminating  the  derivatives  of  <£, 


J= 


Vr 


=uxvy-uyvx=0 (3), 


where  ux,  uy,  vx,  vy  are  the  derivatives  of  u  and  v  with  respect  to  x  and  y. 
This  condition  is  therefore  necessary  for  the  existence  of  a  relation  such 
as  (1).  It  can  be  proved  that  the  condition  is  also  sufficient ;  for  this  we  must 
refer  to  Goursat's  Cows  d" 'Analyse,  vol.  i,  pp.  125  et  seq. 

Two  functions  n  and  v  are  said  to  be  dependent  or  independent  accoi'dino- 
as  they  are  or  are  not  connected  by  such  a  relation  as  (1).  It  is  usual  to  call 
J"  the  Jacobian  or  functional  determinant  of  u  and  v  with  respect  to  x  and  y, 
and  to  write 

J=d  (u,  v) 
9  (x,  y)  ' 

Similar  results  hold  for  functions  of  any  number  of  variables.  Thus  three 
functions  u,  v,  w  of  three  variables  x,  y,  z  are  or  are  not  connected  by  a 
relation  0  (u,  v,  w)  =  0  according  as 

ux     uu     «2 


«/= 


vr.      v„      V,. 


d  (u,  v,  id) 
d  (x,  y,  z) 


does  or  does  not  vanish  for  all  values  of  x,  y,  z. 

21.  Show  that  ax2  +  2hxy  +  by2  and  Ax2  +  2Hxy  +  By2  are   independent 
unless  aj A  =  h\R=  b/B. 

22.  Show  that  ax2 +by2+cz2  +  2fyz  +  2gzx+2hxy  can  be  expressed  as  a 
product  of  two  linear  functions  of  x,  y,  and  z  if  and  only  if 

abc  +  2fgh  -  af2  -  bg2  -  ch2 = 0. 

[Write  down  the  condition  that  px+qy+rz  and  p'x+q'y  +  r'z  should  be 
connected  with  the  given  function  by  a  functional  relation.] 


301  ADDITIONAL    THEOREMS   IN    THE    DIFFERENTIAL  [VII 

23.  If  u  and  v  are  functions  of  £  and  tj,  which  are  themselves  functions 
of  x  and  y,  then 

8  («,  *Q  =  8  (m,  *0  8  (&  q) 

9  to  y)    8  (I.  >?) 8  to  y) ' 

Extend  the  result  to  any  number  of  variables. 

24.  Let/ (#)  be  a  function  of  x  whose  derivative  is  \\x  and  which  vanishes 
when  x=  1.  Show  that  if  u=f(x)+f(y),  v  =  xy,  then  uxvy  —  uyvx=0,  and  hence 
that  u  and  «  are  connected  by  a  functional  relation.  By  putting  y  =  l,  show 
that  this  relation  must  hef(x)+f(y)  -f(xy).  Prove  in  a  similar  manner  that 
if  the  derivative  of  f(x)  is  1/(1  +x2),  and  /(0)=0,  then  f(x)  must  satisfy  the 
equation 

/W+/<f>-/(£*). 

25.  Prove  that  if  f(x)  =  J    —rj- — -^  then 

26.  Show  that  if  a  functional  relation  exists  between 
«-/(*)+/<*)+/(*),     r=/(y)/(«)+/(*)/(*)+/(*)/(y),     w=f(x)f(y)f(z), 

then  /  must  be  a  constant.  [The  condition  for  a  functional  relation  will  be 
found  to  be 

/'  to/'  (y)/'  (^)  {/(y)  -/(»))  {/(»)  -/to}  {/to  -/(y)} = o.] 

27.  If /(y,  2),  /(z,  #)>  and  /(#,  y)  are  connected  by  a  functional  relation 
then  /(a?,  #)  is  independent  of  x.  (Math.  Trip.  1909.) 

28.  If  u—0,  v  =  0,  w  =  0  are  the  equations  of  three  circles,  rendered 
homogeneous  as  in  Ex.  19,  then  the  equation 

d  (u,  v,  w)  _n 

8  to  y,  2)  ~ 

represents  the  circle  which  cuts  them  all  orthogonally.        (Math.  Trip.  1900.) 

29.  If  A,  B,  C  are  three  functions  of  x  such  that 

A  A'  A" 
B  B  B" 
G     C     C" 

vanishes  identically,  then  we  can  find  constants  A,  p,  v  such  that  \A  +fiB  +  vC 
vanishes  identically ;  and  conversely.  [The  converse  is  almost  obvious.  To 
prove  the  direct  theorem  let  a  =  BC - B'C,  ....  Then  a'  =  BC"  —  B"C,  ..., 
and  it  follows  from  the  vanishing  of  the  determinant  that  /3y'  —  /3'y  =  0,  ...  ; 
and  so  that  the  ratios  a  :  j3  :  y  are  constant.     But  a  A  +j3B  +  yC=0.] 

30.  Suppose  that  three  variables  x,  y,  z  are  connected  by  a  relation  in 
virtue  of  which  (i)  z  is  a  function  of  x  and  y,  with  derivatives  zx  zu,  and  (ii)  x 
is  a  function  of  y  and  z,  with  derivatives  xy ,  x„ .     Prove  that 

Xy=—ZyjZx,  Xz=  LjZX. 


AND  INTEGRAL  CALCULUS  305 

[We  have  dz  =  zx  dx + zy  dy,     dx = xy  dy  +  xz  dz 

The  result  of  substituting  for  dx  in  the  first  equation  is 

dz=(zxxy+zy)dy+zxxzdz, 
which  can  be  true  only  if  zxxy+zy  =  0,  zxxz=l.~] 

31.  Four  variables  x,  y,  z,  u  are  connected  by  two  relations  in  virtue  of 
which  any  two  can  be  expressed  as  functions  of  the  others.     Show  that 

where  yzn  denotes  the  derivative  of  y,  when  expressed  as  a  function  of  z  and  u, 
with  respect  to  z.  {Math.  Trip.  1897.) 

32.  Find  A ,  B,  C,  X  so  that  the  first  four  derivatives  of 
/(*)  dt  -  x  [Af(a)  +  Bf(a  + \x)  +  Cf(a  +  x)] 


/: 


vanish  when  x=0  ;  and  A,  B,  C,  D,  X,  p  so  that  the  first  six  derivatives  of 

(       f(f)  dt-x[A/(a)+B/(a  +  \x)  +  Cf(a  +  lix)  +  J)/(a  +  x)] 
J  a 

vanish  when  x=0. 

33.     If  a  >  0,  ac  —  b2>  0,  and  xx  >  x0,  then 


J  a 


<fo  =         !         arc  tan  (  (^-■r0)N/(ac-62)  \ 

«#2  +  2  6#  +  e     ^(ac  —  b2)  \ax\Xo  +  6  (#x  +  #0)  +  cj  ' 


the  inverse  tangent  lying  between  0  and  tj-.* 

I  Sill  n  fr  7" 

34.  Evaluate  the  integral  /       = — - ■ n- .     For  what  values  of  a  is 

°       J  _x  1  —  2xcosa  +  x- 

the  integral  a  discontinuous  function  of  a  ?  (Math.  Trip.  1904.) 

[The  value   of  the  integral  is  \n   if  2nn  <  a  <  (2n  + 1 )  tt,  and    -\n   if 
(2?i  -  1)  iv  <  a  <  2iitt,  n  being  any  integer  ;  and  0  if  a  is  a  multiple  of  7r.] 

35.  If  ax2  +  2bx  +  c>0  when  x0^x  ^x^ ,  f  (x)  =  ^  (ax2  +  2bx + c),  and 

#=/»>    yo=/(#o)>    yi=/(*i)>     ^=(*i-^o)/(3/i+yo)> 
.,  pWx-       1  1  +  X^a  2 

then       J,07  =  ^losr^Tv«'    ^^)arctan^V(-°)}. 

according  as   a  is  positive   or  negative.     In    the    latter    case    the  inverse 
tangent  lies  between  0  and  \ir.     [It  will  be   found   that  the   substitution 

X  —  Xn  f  ^       dt 

t— —  reduces  the  integral  to  the  form  2  I      - :,  .1 

y+Vo  Jo    l~at~ 

fa  f]T 

3.     Prove  that  :=&.  (Math.  Trip.  1913.) 

J  o  •t/~t~  \'\a  ~  ^  ) 

If  a  >  1  then  f     ^"^  efce  =  it  {a  -  V(«2  -  1 )} . 

J  -i      a  —  x 


36. 


37. 

*  In  connection  with  Exs.  33 — 35,  38,  and  40  see  a  paper  by  Dr  Bromwich 
in  vol.  xxxv  of  the  Messenger  of  Mathematics. 

h.  20 


306  ADDITIONAL   THEOREMS   IN   THE   DIFFERENTIAL  [VII 

38.     If>>l,  0<q<l,  then 

n  dx  2cc 


J  o  v'[{l  +  (P*  ~  1)  4  {1  -  (1  -  ?2)  *}]      (i*  +  2)  si"  » ' 
where  w  is  the  positive  acute  angle  whose  cosine  is  (l+jogO/^  +  gO- 

39.  If«>6>0,thenJo    -^-^^^{a-^-6% 

(Jfa&  2Wp.  1904.) 

40.  Prove  that  if  a  >  »/(62+c2)  then 

r        <**        _      2 arc  tan  K^2-62-c2)] 

J0  a  +  &cos0  +  Csin0-^(a*-&2-c2)arCtant  c  J' 

the  inverse  tangent  lying  between  0  and  it.. 

fb 

41.  If /(a?)  is  continuous  and  never  negative,  and    I    f(x)dx  =  0,  then 

./  « 
f(x)  =  0  for  all  values  of  #  between  a  and  &.    [If/ (a;)  were  equal  to  a  positive 
number  &  when  #  =  £,  say,  then  we  could,  in  virtue  of  the  continuity  of  /(*•), 
find  an  interval  (£-8,  £  +  8)  throughout  which  /(a?)  >  \k  ;  and  then  the  value 
of  the  integral  would  be  greater  than  8L] 

42.  Schwarz's  inequality  for  integrals.     Prove  that 


(JW  due)  ^  (  <£2  dx  f  \p  dx. 


[Use  the  definitions  of  §§  156  and  157,  and  the  inequality 

(Ch.  I,  Misc.  Ex.  10).] 

43.     If  Pn(x)  =  {p_\)nn[  (dx)n{(x-a)(P-*Vn>   then   P«(4  fs  a  P()1y- 
nomial  of  degree  n,  which  possesses  the  property  that 


/: 


Pn(x)6(x)dx=0 


if  6(x)  is  any  polynomial  of  degree  less  than  n.     [Integrate  by  parts  m  +  1 
times,  where  m  is  the  degree  of  6  (x),  and  observe  that  0(m  +  1)(.r)  =  O.] 

44.  Prove  that  I     Pm{x)  Pn(x)dx  =  0  if  m  4=  n,  but  that  if  m=«  then  the 
value  of  the  integral  is  (0  -  a)j(2n  +  1). 

45.  If  Qn  (a?)  is  a  polynomial  of  degree  n,  which  possesses  the  property 

that  I     §«  (a;)  6  (x)  dx  =  0  if  0  (x)  is  any  polynomial  of  degree  less  than  n,  then 

Qn  {%)  is  a  constant  multiple  of  Pn  (x). 

[We  can  choose  k  so  that  (?n  — «PM  is  of  degree  %- 1  :  then 

j      Qn(Qn-*Pn)dx  =  0,       j^  Pn  (Qn  -  KPn)  dx  =  0, 


AND   INTEGRAL   CALCULUS  307 

and  so  I     (Qn  -  <Pn)2  dx  =  0. 

Now  apply  Ex.  41. J 

46.    Approximate  Values  of  definite  integrals.     Show  that  the  error 

fb 

in  taking  \{b  —  a)  {0  (a)  +  0  (b)}  as  the  value  of  the  integral  I    0 {x)  dx  is  less 


than  ^3/{b  —  a)3,  where  J/ is  the  maximum  of  |0"(.r)|  in  the  interval  (a,  b); 
and  that  the  error  in  taking  (o  —  a)  0  {-|  (a  +  b)}  is  less  than  -^M  (b  —  a)3.  [Write 
f  (j£)  =  0  (x)  in  Exs.  4  and  5.]     Show  that  the  error  in  taking 

1(6  -a)  [0(a)  +  0(6)  +  40  {£(«  +  &)}] 
as  the  value  is  less  than  ogy/^-a)5,  where  1/  is  the  maximum  of  0(4)(:r). 
[Use  Ex.  6.     This  rule,  which  gives  a  very  good  approximation,  is  known  as 
Simpson's  Rule.     It  amounts  to  taking  one-third  of  the  first  approximation 
given  above  and  two-thirds  of  the  second.] 

Show  that  the  approximation  assigned  by  Simpson's  Eule  is  the  area 
bounded  by  the  lines  x=a,  x=b,  y  =  Q,  and  a  parabola  with  its  axis  parallel 
to  0  Y  and  passing  through  the  three  points  on  the  curve  3/  =  0  (#)  whose 
abscissae  are  a,  \  (a  +  b),  b. 

It  should  be  observed  that  if  0  (x)  is  any  cubic  polynomial  then  0(4)  (x)  =  0, 
and  Simpson's  Rule  is  exact.  That  is  to  say,  given  three  points  whose 
abscissae  are  a,  ^(a  +  b),  b,  we  can  draw  through  them  an  infinity  of  curves 
of  the  type  y=a  +  (3x+yx2  +  8x3 ;  and  all  such  ci;rves  give  the  same  area.  For 
one  curve  8  =  0,  and  this  curve  is  a  parabola. 

47.  If  0  (x)  is  a  polynomial  of  the  fifth  degree,  then 

•l 

0  Or)  dx  =  J5  {50  (a)  +  80  (h)  +  50  (£)}, 

a  and  /3  being  the  roots  of  the  equation  x2-x  +  ^  =  0.         {Math.  Trip.  1909.) 

48.  Apply  Simpson's   Rule  to  the   calculation   of  ir  from  the  formula 
r  i   dx 

j7r  =  I    -       2.     [The  result  is  '7833....     If  we  divide  the  integral  into  two, 
J  o  l+x" 

from  0  to  |  and  ^  to  1,  and  apply  Simpson's  Rule  to  the  two  integrals 

separately,  we  obtain  -7853916....     The  correct  value  is  -7853981....] 

49.  Show  that  8-9  <  f^(4:  +  x2)dx<9.  {Math.  Trip.  1903.) 

50.  Calculate  the  integrals 

0ifs'  A'vcfer  />"*><**■  fa™** 

to  two  places  of  decimals.  [In  the  last  integral  the  subject  of  integration  is 
not  defined  when  &'=0:  but  if  we  assign  to  it,  when  x  =  0,  the  value  1,  it 
becomes  continuous  throughout  the  range  of  integration.] 

20—2 


CHAPTER   VIII 

THE   CONVERGENCE   OF   INFINITE   SERIES  AND 
INFINITE   INTEGRALS 

165.  In  Ch.  IV  we  explained  what  was  meant  by  saying 
that  an  infinite  series  is  convergent,  divergent,  or  oscillatory,  and 
illustrated  our  definitions  by  a  few  simple  examples,  mainly 
derived  from  the  geometrical  series 

1  +x  +  w-  +  ... 

and  other  series  closely  connected  with  it.  In  this  chapter  we 
shall  pursue  the  subject  in  a  more  systematic  manner,  and  prove 
a  number  of  theorems  which  enable  us  to  determine  when  the 
simplest  series  which  commonly  occur  in  analysis  are  convergent. 

We  shall  often  use  the  notation 

n 
um  +  um+1  +  ...+  un  =  X 0  (v), 

m 

00 

and  write  Xun,  or  simply  2t*„,  for  the  infinite  series  w0+Wi+«2  + * 

o 

166.  Series  of  Positive  Terms.  The  theory  of  the  con- 
vergence of  series  is  comparatively  simple  when  all  the  terms  of 
the  series  considered  are  positive  f.     We  shall  consider  such  series 

*  It  is  of  course  a  matter  of  indifference  whether  we  denote  our  series  by 
wi  +  «2  +  •  •  •  (as  m  Ch.  IV)  or  by  u0  +  u1  +  ...  (as  here).  Later  in  this  chapter  we 
shall  be  concerned  with  series  of  the  type  a0  +  a1x  +  a2x2+ ...:  for  these  the  latter 
notation  is  clearly  more  convenient.  We  shall  therefore  adopt  this  as  our  standard 
notation.  But  we  shall  not  adhere  to  it  systematically,  and  we  shall  suppose  that  «x 
is  the  first  term  whenever  this  course  is  more  convenient.  It  is  more  convenient, 
for  example,  when  dealing  with  the  series  1  +  %  +  ^  + ...  ,  to  suppose  that  un=ljn 
and  that  the  series  begins  with  w1(  than  to  suppose  that  u„=l/(ra+l)  and  that  the 
series  begins  with  u0.     This  remark  applies,  e.g.,  to  Ex.  lxvii.  4. 

t  Here  and  in  what  follows  '  positive  '  is  to  be  regarded  as  including  zero. 


165-168]      THE   CONVERGENCE   OF   INFINITE    SERIES,    ETC.  309 

first,  not  only  because  they  are  the  easiest  to  deal  with,  but  also 
because  the  discussion  of  the  convergence  of  a  series  containing 
negative  or  complex  terms  can  often  be  made  to  depend  upon 
a  similar  discussion  of  a  series  of  positive  terms  only. 

When  we  are  discussing  the  convergence  or  divergence  of  a 
series  we  may  disregard  any  finite  number  of  terms.  Thus,  when 
a  series  contains  a  finite  number  only  of  negative  or  complex  terms, 
we  may  omit  them  and  apply  the  theorems  which  follow  to  the 
remainder. 

167  It  will  be  well  to  recall  the  following  fundamental 
theorems  established  in  §  77. 

A.  A  series  of  positive  terms  must  be  convergent  or  diverge 
to  oo  ,  and  cannot  oscillate. 

B.  The  necessary  and  sufficient  condition  that  %un  should  be 
convergent  is  that  there  shoidd  be  a  number  K  such  that 

k0  +  ux  +  ...+  un <  K 
for  all  values  of  n. 

C.  The  comparison  theorem.  If  2?/n  is  convergent,  and 
vn  =  Un  for  all  values  of  n,  then  Xvn  is  convergent,  and  Svn^  2«M. 
More  generally,  if  vn^Kun,  where  K  is  a  constant,  then  %vn 
is  convergent  and  2vn  ^  K%un.  And  if  Xun  is  divergent,  and 
vn  ^  Kun,  then  2vn  is  divergent.* 

Moreover,  in  inferring  the  convergence  or  divergence  of  %vn 
by  means  of  one  of  these  tests,  it  is  sufficient  to  know  that  the 
test  is  satisfied  for  sufficiently  large  values  of  n,  i.e.  for  all  values 
of  n  greater  than  a  definite  value  n0.  But  of  course  the  con- 
clusion that  %vn^KXun  does  not  necessarily  hold  in  this  case. 

A  particularly  useful  case  of  this  theorem  is 

D.  If  ^un  is  convergent  (divergent)  and  un/vn  tends  to  a  limit 
other  than  zero  as  n  -*-  oo ,  then  Xv„  is  convergent  (divergent). 

168.  First  applications  of  these  tests.  The  one  important 
fact  which  we  know  at  present,  as  regards  the  convergence  of  any 

*  The  last  part  of  this  theorem  was  not  actually  stated  in  §  77,  but  the  reader 
will  have  no  difficulty  in  supplying  the  proof. 


310  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

special  class  of  series,  is  that  trn  is  convergent  if  r  <  1  and 
divergent  if  rSL*  It  is  therefore  natural  to  try  to  apply 
Theorem  C,  taking  un  =  rn.     We  at  once  find 

1.  The  series  %vn  is  convergent  if  vn  g  Kr11,  where  r  <  I,  for  all 
sufficiently  large  values  of  n. 

When  K  =  1,  this  condition  may  be  written  in  the  form  vnl!n  ^  r, 
Hence  we  obtain  what  is  known  as  Cauchy's  test  for  the  con- 
vergence of  a  series  of  positive  terms ;  viz. 

2.  The  series  Xvn  is  convergent  if  vnVn^r,  where  r  <  1,  for 
all  sufficiently  large  values  of  n. 

There  is  a  corresponding  test  for  divergence,  viz. 

2a.  The  series  Xvn  is  divergent  if  vnvn  =  1  for  an  infinity  of 
values  of  n. 

This  hardly  requires  proof,  for  vn1/n  ^  1  involves  vn  ^  1.  The 
two  theorems  2  and  2a  are  of  very  wide  application,  but  for 
some  purposes  it  is  more  convenient  to  use  a  different  test  of 
convergence,  viz. 

3.  The  series  Xvn  is  convergent  if  vn+1fvn  ^  r,  where  r  <  1,/or 
all  sufficiently  large  values  of  n. 

To  prove  this  we  observe  that  if  vn+i/vn  =  r  when  n  =  n0  then 

vn_,  vn-2        vno  r'!« 

and  the  result  follows  by  comparison  with  the  convergent  series  %rn. 
This  test  is  known  as  d'Alembert's  test.  We  shall  see  later  that 
it  is  less  general,  theoretically,  than  Cauchy's,  in  that  Cauchy's 
test  can  be  applied  whenever  d'Alembert's  can,  and  sometimes 
when  the  latter  cannot.  Moreover  the  test  for  divergence  which 
corresponds  to  d'Alembert's  test  for  convergence  is  much  less 
general  than  the  test  given  by  Theorem  2a.  It  is  true,  as  the 
reader  will  easily  prove  for  himself,  that  if  vn+1/vn  =  r  =  1  for  all 
values  of  n,  or  all  sufficiently  large  values,  then  %vn  is  divergent. 
But  it  is  not  true  (see  Ex.  lxvii.  9)  that  this  is  so  if  only 
vn+J/vn  —  r  —  1  for  an  infinity  q/ values  of  n,  whereas  in  Theorem  2a 

*  We  shall  use  r  in  this  chapter  to  denote  a  number  which  is  always  positive 
or  zero. 


168]  AND   INFINITE   INTEGRALS  311 

our  test  had  only  to  be  satisfied  for  such  an  infinity  of  values. 
None  the  less  d'Alembert's  test  is  very  useful  in  practice,  because 
when  vn  is  a  complicated  function  vn+Jvn  is  often  much  less 
complicated  and  so  easier  to  work  with. 

In  the  simplest  cases  which  occur  in  analysis  it  often  happens 
that  vn+1/vn  or  vnVn  tends  to  a  limit  as  n  ->• oo  .*  When  this  limit 
is  less  than  1,  it  is  evident  that  the  conditions  of  Theorems  2  or 
3  above  are  satisfied.     Thus 

4.  If  v,iln  or  vn+]/yn  tends  to  a  limit  less  than  unity  as  ?i->-  oo  , 
then  the  series  1vn  is  convergent. 

It  is  almost  obvious  that  if  either  function  tend  to  a  limit 
greater  than  unity,  then  %vn  is  divergent.  We  leave  the  formal 
proof  of  this  as  an  exercise  to  the  reader.  But  when  vnvn  or 
Vn+i/Vn  tends  to  1  these  tests  generally  fail  completely,  and  they 
fail  also  when  vn1,n  or  vn+l/vn  oscillates  in  such  a  way  that,  while 
always  less  than  1,  it  assumes  for  an  infinity  of  values  of  n  values 
approaching  indefinitely  near  to  1.  And  the  tests  which  involve 
vn+1/vn  fail  even  when  that  ratio  oscillates  so  as  to  be  sometimes 
less  than  and  sometimes  greater  than  1.  When  vnVn  behaves  in 
this  way  Theorem  2a  is  sufficient  to  prove  the  divergence  of  the 
series.  But  it  is  clear  that  there  is  a  wide  margin  of  cases  in 
which  some  more  subtle  tests  will  be  needed. 

Examples  LXVII.  1.  Apply  Cauchy's  and  d'Alembert's  tests  (as 
specialised  in  4  above)  to  the  series  SjiV1,  where  k  is  a  positive  rational 
number. 

[Here  vn  +  1/ vn=  {('>i  +  l)ln}k r-*-?',  so  that  d'Alembert's  test  shows  at  once 
that  the  series  is  convergent  if  r  <  1  and  divergent  if  /•  >  1.  The  test  fails  if 
r=  1  :  but  the  series  is  then  obviously  divergent.  Since  limw1/re  =  l  (Ex.  xxvir. 
11),  Cauchy's  test  leads  at  once  to  the  same  conclusions.] 

2.  Consider  the  series  2(A7ik  +  Buk~1  + ...  +K)rn.  [We  may  supposed 
positive.  If  the  coefficient  of  rH  is  denoted  by  P(n),  then  P{n)jnk^-A  and, 
by  D  of  §  167,  the  series  behaves  like  2?iV\] 

3.  Consider  2  ^L±^1±L^±A  ,.»    (J>0,a>0). 

an1  +  fin1    1+  ...+  k 

[The  series  behaves  like  2nJc-lrn.  The  case  in  which  r=l,  k<l  requires 
further  consideration.] 

*  It  will  be  proved  in  Ch.  IX  (Ex.  lxxxvii.  36)  that  if  vn+1lvn-*-l  then  v^ln -*  I. 
That  the  converse  is  not  true  may  be  seen  by  supposing  that  vn=l  when  n  is  odd 
and  v,,-2  when  ?i  is  even. 


312  THE   CONVERGENCE   OF    INFINITE   SERIES  [VIII 

4.  We  have  seen  (Ch.  IV,  Misc.  Ex.  17)  that  the  series 

s        1  S I 

w(»i  +  l)'         n{n  +  \)...(n+p) 

are  convergent.     Show  that  Cauchy's  and  d'Alembert's  tests  both  fail  when 
applied  to  them.     [For  lim  unlln  =  lim  (un  + 1 \un)  =  1.] 

5.  Show  that  the  series  2?i-J>,  where  p  is  an  integer  not  less  than  2,  is 
convergent.  [Since  lim  {n(n+l)...(n+p-l)}jnp=l,  this  follows  from  the 
convergence  of  the  series  considered  in  Ex.  4.  It  has  already  been  shown 
in  §  77,  (7)  that  the  series  is  divergent  if  p  —  1,  and  it  is  obviously  divergent  if 

6.  Show  that  the  series 

An*  +  Bnk~l  +  ...  +  K 

an1  +  finl~l  +  ...  +  < 

is  convergent  if  I >  k  + 1  and  divergent  if  l<k  + 1. 

7.  If  mn  is  a  positive  integer,  and  »nB+1>mn,  then  the  series  2rm«  is  con- 
vergent if  r<l  and  divergent  if  r  >1.  For  example  the  series  l+r+?"4+?'9  +  ... 
is  convergent  if  r  <  1  and  divergent  if  ?-Sl. 

8.  Sum  the  series  l  +  2r+2>A+ ...  to  24  places  of  decimals  when  r='l 
and  to  2  places  when  r=49.  [If  »'="1,  then  the  first  5  terms  give  the 
sum  1-2002000020000002,  and  the  error  is 

2r25  +  2r36 +mi  <2r25  +  2?.36  +  £r47  + ...  = 2?-25/(l  - rll)<3/1025. 

If  r='9,  then  the  first  8  terms  give  the  sum  5'458...,  and  the  error  is  less 
than  2r6i/(l  -r«)<  003.] 

9  If  0<a<6<l,  then  the  series  a  +  b  +  a2  +  b2-\-a3  +  ...  is  convergent. 
Show  that  Cauchy's  test  may  be  applied  to  this  series,  but  that  d'Alembert's 
test  fails.     [For 

%+ik=(W"+1-*».     v2n  +  2lv2n  +  1  =  b  («/&)>'  +  2^0.] 

10.  The  series  1+?-+^  +  ^  +  ...  and  1  +  r  +  ~2  +  —  + . . .  are  convergent 
for  all  positive  values  of  r. 

11.  If  2w„  is  convergent  then  so  are  2m„2  and  2w„/(l  +  un). 

12.  If  2w„2  is  convergent  then  so  is  2unjn.  [For  2un/n^un2  +  (l/n2)  and 
2  (I/ft2)  is  convergent.] 

13.  Show  that  l+-2  +  p+...  =  -  M +-2  + 1  +  ...J  and 

1       1       1       11       1  _15  /       1       1 

1  +  22  +  32  +  52  +  g2  +  72  +  Q2  +"  •  •  •  —  TR  1  1  +  02  +  Q2  +  "  • 


168-170]  AND    INFINITE    INTEGRALS  313 

[To  prove  the  first  result  we  note  that 

1+p  +  S+'«-(i+J)  +  (p+J)+». 

,     11  i  /,    l      l        \ 

-l+p  +  Bi+,«  +  p(l+p  +  p  +  ...), 

by  Theorems  (8)  and  (6)  of  §  77.] 

14.     Prove  by  a  reductio  ad  absurdum  that  2  (l/«)  is  divergent.     [If  the 
series  were  convergent  we  should  have,  by  the  argument  used  in  Ex.  13, 

i+Hl+.«=a+i+*+.»)+*(i+i+i+-)» 

or  x+i  +  i  +  ...  =  i+i  +  i  +  ... 

which  is  obviously  absurd,  since  every  term  of  the  first  series  is  less  than  the 
corresponding  term  of  the  second.] 

169.  Before  proceeding  further  in  the  investigation  of  tests 
of  convergence  and  divergence,  Ave  shall  prove  an  important  general 
theorem  concerning  series  of  positive  terms. 

Dirichlet's  Theorem.*  The  sum  of  a  series  of  positive 
terms  is  the  same  in  whatever  order  the  terms  are  taken. 

This  theorem  asserts  that  if  we  have  a  convergent  series  of 
positive  terms,  u0  +  ux  +  u2  +  . . .  say,  and  form  any  other  series 

V0  +  V1  +  V,  +  . . . 

out  of  the  same  terms,  by  taking  them  in  any  new  order,  then  the 
second  series  is  convergent  and  has  the  same  sum  as  the  first. 
Of  course  no  terms  must  be  omitted :  every  u  must  come  some- 
where among  the  v's,  and  vice  versa. 

The  proof  is  extremely  simple.  Let  s  be  the  sum  of  the  series 
of  its.  Then  the  sum  of  any  number  of  terms,  selected  from  the 
u's,  is  not  greater  than  s.  But  every  v  is  a  u,  and  therefore  the 
sum  of  airy  number  of  terms  selected  from  the  v's  is  not  greater 
than  s.  Hence  Xvn  is  convergent,  and  its  sum  t  is  not  greater 
than  s.  But  we  can  show  in  exactly  the  same  way  that  sli. 
Thus  s  =  t. 

170.  Multiplication  of  Series  of  Positive  Terms.      An 

immediate  corollary  from  Dirichlet's  Theorem  is  the  following 
theorem  :ifu0+u1  +  ui  +  ...andv0  +  v1  +  v2+...  are  two  convergent 

*  This  theorem  seems  to  have  first  been  stated  explicitly  by  Dirichlet  in  1837. 
It  was  no  doubt  known  to  earlier  writers,  and  in  particular  to  Cauchy. 


314 


THE    CONVERGENCE    OF    INFINITE    SERIES 


[VIII 


series  of  positive  terms,  and  s  and  t  are  their  respective  sums, 
then  the  series 

U0V0  +  («!«,, +  MoWi)  +  (u2V0  +  U1V1  +  U0Va)+  ... 

is  convergent  and  has  the  sum  st. 

Arrange  all  the  possible  products  of  pairs  umvn  in  the  form  of 
a  doubly  infinite  array 


uxv0 

U2V0 

u-2v1 

U2V2 

U3V0 
U3V2 

uzv3 

tl0V2 

Wl  V2 

U0VS 

■UiV-i 

U2V3 

We  can  rearrange  these  terms  in  the  form  of  a  simply  infinite 
series  in  a  variety  of  ways.     Among  these  are  the  following. 

(1)  We  begin  with  the  single  term  u0v0  for  which  m  +  n  =  0; 
then  we  take  the  two  terms  u^,  u^  for  which  m  +  n  =  1 ;  then 
the  three  terms  u2v0,  u^,  u0v2  for  which  m  +  n  =  2 ;  and  so  on. 
We  thus  obtain  the  series 

Ua  VQ  +  (ll^o  +  U9  Vj)  +  (u2v0  +  Ux  Uj  +  u0v2)  +  ... 
of  the  theorem. 

(2)  We  begin  with  the  single  term  u0v0  for  which  both 
suffixes  are  zero;  then  we  take  the  terms  u^o,  ulv1,  w^  which 
involve  a  suffix  1  but  no  higher  suffix ;  then  the  terms  u2v0>  u2v1} 
u2v2,  UiV2,  it0v2  which  involve  a  suffix  2  but  no  higher  suffix;  and 
so  on.    The  sums  of  these  groups  of  terms  are  respectively  equal  to 

U0V0,      Oq  +  Mj)  Oo  +  Vy)  -  U0V0, 

(«0  +  «l  +  U2)  (VQ  +  Vi  +  V.2)  -  (ll0  +  Ux)  (V0  +  Vj),      . . . 

and  the  sum  of  the  first  n  +  1  groups  is 

Oo  +»i+...+  un)  (v0  +  v1  +  ...  +  vn), 
and  tends  to  st  as  n  -»■  oo .     When  the  sum  of  the  series  is  formed 
in  this  manner  the  sum  of  the  first  one,  two,  three,   ...  groups 
comprises  all  the  terms  in  the  first,  second,  third,  ...  rectangles 
indicated  in  the  diagram  above. 

The  sum  of  the  series  formed  in  the  second  manner  is  st. 
But  the  first  series  is  (when  the  brackets  are  removed)  a  rearrange- 
ment of  the  second;  and  therefore,  by  Dirichlet's  Theorem,  it  con- 
verges to  the  sum  st.   Thus  the  theorem  is  proved. 


170,  171]  AND   INFINITE   INTEGRALS  315 

Examples  LXVIII.     1     Verify  that  if  r  <  1  then 

l  +  r2  +  r  +  ri  +  r6  +  r3  +  ...  =  l+r  +  r3+r-  +  r5  +  ri  +  ...  =  l/(l-r). 

2.*  If  either  of  the  series  n0+Ux  + ...,  v0  +  vi+...  is  divergent,  then  so  is 
the  series  «oi'o  +  (?'il'o  +  "ovi)  +  (?i2vo  +  wivi+Moy2)  +  ---)  except  in  the  trivial 
case  in  which  every  term  of  one  series  is  zero. 

3.  If  the  series  uQ  +  ux  +  ...,  v0+»i  +  ...,  w$+Wx+..,  converge  to  sums 
r,  s,  t,  then  the  series  2\k,  where  \k  =  ~2,umvnwp,  the  summation  being  extended 
to  all  sets  of  values  of  m,  n,  p  such  that  m+n+p=&t  convei-ges  to  the 
sum  rst. 

4.  If  2«,t  and  2v»  converge  to  sums  s  and  t,  then  the  series  2%,  where 
wn  =  2uiVm,  the  summation  extending  to  all  pairs  I,  m  for  which  lm  =  n, 
converges  to  the  sum  st. 

171.      Further   tests  for   convergence    and   divergence. 

The  examples  on  pp.  311 — 313  suffice  to  show  that  there  are 
simple  and  interesting  types  of  series  of  positive  terms  which 
cannot  be  dealt  with  by  the  general  tests  of  §  168.  In  fact,  if 
we  consider  the  simplest  type  of  series,  in  which  un+1/un  tends 
to  a  limit  as  n ■-*■  oo  ,  the  tests  q/§  168  generally  fail  when  this  limit 
is  1.  Thus  in  Ex.  lxvii.  5  these  tests  failed,  and  we  had  to  fall 
back  upon  a  special  device,  which  was  in  essence  that  of  using 
the  series  of  Ex.  lxvii.  4  as  our  comparison  series,  instead  of 
the  geometric  series. 

The  fact  is  that  the  geometric  series,  by  comparison  with  which  the  tests 
of  §  168  were  obtained,  is  not  only  convergent  but  very  rapidly  convergent, 
far  more  rapidly  than  is  necessary  in  order  to  ensure  convergence.  The  tests 
derived  from  comparison  with  it  are  therefore  naturally  very  crude,  and  much 
more  delicate  tests  are  often  wanted. 

We  proved  in  Ex.  xxvil.  7  that  nftrn-»-0  as  n-s-ao,  provided  r<\,  what- 
ever value  k  may  have;  and  in  Ex.  lxvii.  1  we  proved  more  than  this, 
viz.  that  the  series  2nkrn  is  convergent.  It  follows  that  the  sequence 
r,  r2,  7'3,  ...,  r",  ...,  where  r<l,  diminishes  more  rapidly  than  the  sequence 
1-^2"*,  3 ~*,... •,»""*,  ....  This  seems  at  first  paradoxical  if  r  is  not  much  less 
than  unity,  and  k  is  large.     Thus  of  the  two  sequences 

2      4        8  .1 ] 1 

3»   95    2T>    •••  '       x)   409(5'    53I441J    ■■• 

whose  general  terms  are  (f)n  and  w-12,  the  second  seems  at  first  sight  to 
decrease  far  more  rapidly.  But  this  is  far  from  being  the  case  ;  if  only  we 
go  far  enough  into  the  sequences  we  shall  find  the  terms  of  the  first  sequence 
very  much  the  smaller.     For  example, 

(2/3)4  =  16/81<l/5,    (2/3)12<(l/5)3<(l/10)2,    (2/3)1000<(l/10)166, 
while  1000-12  =  10"36; 

*  In  Exs.  2 — i  the  series  considered  are  of  course  series  of  positive  terms. 


31G  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

so  that  the  1000th  term  of  the  first  sequence  is  less  than  the  10130th  part  of 
the  corresponding  term  of  the  second  sequence.  Thus  the  series  2  (2/3)"  is 
far  more  rapidly  convergent  than  the  series  2?i~12,  and  even  this  series  is 
very  much  more  rapidly  convergent  than  2«-2.* 

172.  We  shall  proceed  to  establish  two  further  tests  for  the 
convergence  or  divergence  of  series  of  positive  terms,  Maclaurin's 
(or  Cauchy's)  Integral  Test  and  Cauchy's  Condensation 
Test,  which,  though  very  far  from  being  completely  general,  are 
sufficiently  general  for  our  needs  in  this  chapter. 

In  applying  either  of  these  tests  we  make  a  further  assumption 
as  to  the  nature  of  the  function  un,  about  which  we  have  so  far 
assumed  only  that  it  is  positive.  We  assume  that  un  decreases 
steadily  with  n:  i.e.  that  un+1  g  un  for  all  values  of  n.  or  at  any  rate 
all  sufficiently  large  values. 

This  condition  is  satisfied  in  all  the  most  important  cases.  From  one 
point  of  view  it  may  be  regarded  as  no  restriction  at  all,  so  long  as  we  are 
dealing  with  series  of  positive  terms  :  for  in  virtue  of  Dirichlet's  theorem 
above  we  may  rearrange  the  terms  without  affecting  the  question  of  con- 
vergence or  divergence ;  and  there  is  nothing  to  prevent  us  rearranging  the 
terms  in  descending  order  of  magnitude,  and  applying  our  tests  to  the  series  of 
decreasing  terms  thus  obtained. 

But  before  we  proceed  to  the  statement  of  these  two  tests, 
we  shall  state  and  prove  a  simple  and  important  theorem,  which 
we  shall  call  Abel's  Theorem f.  This  is  a  one-sided  theorem  in 
that  it  gives  a  sufficient  test  for  divergence  only  and  not  for 
convergence,  but  it  is  essentially  of  a  more  elementary  character 
than  the  two  theorems  mentioned  above. 

173.  Abel's  (or  Pringsheim's)  Theorem.  If  2w„  is  a  convergent  series  of 
positive  and  decreasing  terms,  then  lim  nnn  =  0. 

Suppose  that  nun  does  not  tend  to  zero.  Then  it  is  possible  to  find  a 
positive  number  8  such  that  nun^.8  for  an  infinity  of  values  of  n.  Let  nx  be 
the  first  such  value  of  n ;  n2  the  next  such  value  of  n  which  is  more  than 

*  Five  terms  suffice  to  give  the  sum  of  2«-12  correctly  to  7  places  of  decimals, 
whereas  some  10,000,000  are  needed  to  give  an  equally  good  approximation  to  2h~2. 
A  large  number  of  numerical  results  of  this  character  will  be  found  in  Appendix  III 
(compiled  by  Mr  J.  Jackson)  to  the  author's  tract  '  Orders  of  Infinity'  (Cambridge 
Math.  Tracts,  No.  12). 

t  This  theorem  was  discovered  by  Abel  but  forgotten,  and  rediscovered  by 
Pringsheim. 


171-173]  AND   INFINITE   INTEGRALS  317 

twice  as  large  as  «j ;  ??3  the  next  such  value  of  n  which  is  more  than  twice 
as  large  as  n2 ;  and  so  on.  Then  we  have  a  sequence  of  numbers  nu  n2,  n3,  ... 
such  that  n2>2)i1,  n3>2?i2,  ...  and  so  n2-  ni>\n2,  nz  -  n1>^n3i  ... ; 
and  also  n^u^  g8,  n2um  riS, ....  But,  since  un  decreases  as  n  increases, 
we  have 

M0  +  ul  +  •  •  •  +  «n,  - 1=  1h  Un,  ~  §, 

«»,  +  ...  +««2-l  ^(?^2-«l)Wn2>  i^2«n2=iS, 

«)l2  +  ...  +«n3_i^(%-w2)  w„3>|»3M„3>^S, 
and  so  on.     Thus  we  can  bracket  the  terms  of  the  series  2  un  so  as  to  obtain 
a  new  series  whose  terms  are  severally  greater  than  those  of  the  divergent 
series 

8  +  ^8+^8  +  ...; 

and  therefore  2un  is  divergent. 

Examples  LXIX.  1.  Use  Abel's  theorem  to  show  that  2  (1/n)  and 
2{l/(cm  +  6)}  are  divergent.     [Here  nuH-*-\  or  nun^-l/a.] 

2.  Show  that  Abel's  theorem  is  not  true  if  we  omit  the  condition  that  un 
decreases  as  n  increases.     [The  series 

1      1111111        1 

1  +  2^  +  32+4+ 52 +  62  +  72+82  +  9  + iq2  +  "-' 

in  which  v.,n  =  \\n  or  l/»2,  according  as  n  is  or  is  not  a  perfect  square,  is 
convergent,  since  it  may  be  rearranged  in  the  form 

1111111  /       1      1 

22  +  32  +  52+^  +  72 +82  +  Io,+  -+(1+4  +  9  +  ' 

and  each  of  these  series  is  convergent.  But,  since  nuH=l  whenever  u  is  a 
perfect  square,  it  is  clearly  not  true  that  nun-*-0.] 

3.  The  converse  of  Abel's  theorem  is  not  true,  i.e.  it  is  not  true  that,  if  zin 
decreases  with  11  and  lim  nun  =  0,  then  1un  is  convergent. 

[Take  the  series  2  (l/n)  and  multiply  the  first  term  by  1,  the  second  by  ^, 
the  next  two  by  $,  the  next  four  by  J,  the  next  eight  by  £,  and  so  on.  On 
grouping  in  brackets  the  terms  of  the  new  series  thus  formed  we  obtain 

i+i-f +£(£+£)+£(£+*+*+£)+•■•; 

and  this  series  is  divergent,  since  its  terms  are  greater  than  those  of 
1  ±1   I4.1    ill    1 4. 

which  is  divergent.     But  it  is  easy  to  see  that  the  terms  of  the  series 


satisfy  the  condition  that  m<«-^0.     In  fact  nun—\jv  if  2"  2<?t<2"  \  and 
v-z-oz  as  ?i-»-co.] 


318  THE   CONVERGENCE   OF   INFINITE   SERIES  [VlII 

174.      Maclaurin's  (or  Cauchy's)  Integral  Test.*     If  wn 

decreases  steadily  as  n  increases,  we  can  write  un  =  (f>  (n)  and 
suppose  that  <f>  (n)  is  the  value  assumed,  when  x  =  n,  by  a  con- 
tinuous and  steadily  decreasing  function  $  (x)  of  the  continuous 
variable  x.     Then,  if  v  is  any  positive  integer,  we  have 

<£0-l)^</><»^0) 
when  v  —  1  ^  x  £  v.     Let 

v„  =  (f>  (v  -  1)  -  I       c})(x)dx=l       {</>  (v  -  1)  -  (/>  0)}  dx, 

J v—l  J v-1 

so  that  0  =S  vv  g  $  (v  -  1)  -  <£  (v). 

Then  2v„  is  a  series  of  positive  terms,  and 

vs  +  v3  +  ...  +  vn  sS  <f>(l)  -  <£  (w)  ^  <£  (1). 
Hence  2i>„  is  convergent,  and  so  v2  +  v3  +  . . .  +  vn  or 

M-l  r'* 

2  (f>  (v)  —       </>  (#)  dx 
i  Ji 

tends  to  a  positive  limit  as  7i-*-<x> . 

Let  us  write  <£  (£)  =  I   $  (#)  <&», 

so  that  3?  (f)  is  a  continuous  and  steadily  increasing  function  of  £. 

Then 

Wj  +  u„  +  . . .  +  «„_i  -  <£  O) 

tends  to  a  positive  limit,  not  greater  than  <£  (1),  as  n  -*-  oo  .  Hence 
2w„  is  convergent  or  divergent  according  as  <J>  (n)  tends  to  a  limit 
or  to  infinity  as  n-^oo  ,  and  therefore,  since  <E>  (n)  increases  steadily, 
according  as  <£>  (£)  tends  to  a  limit  or  to  infinity  as  £-»-  oo  .  Hence 
i/"  (/>  (a;)  is  a  function  of  x  which  is  positive  and  continuous  for  all 
values  of  x  greater  than  unity,  and  decreases  steadily  as  x  increases, 
then  the  series 

£(l)  +  0(2)  +  ... 
does  or  does  not  converge  according  as 

<£(£)=[  <f>(x)dx 

does  or  does  not  tend  to  a  limit  I  as  £-*-oo ;  and,  in  the  first  case, 
the  sum  of  the  series  is  not  greater  than  <fi  (1)  + 1. 

*  The  test  was  discovered  by  Maclaurin  and  rediscovered  by  Cauchy,  to  whom 
it  is  usually  attributed. 


174,  175]  AND   INFINITE   INTEGRALS  319 

The  sum  must  in  fact  be  less  than  cp  (l)  +  l.  For  it  follows  from  (6)  of 
§  160,  and  Ch.  VII,  Misc.  Ex.  41,  that  vv<(f>(v-l)-<f>(v),  unless  <f>  (x)  =  <j>  (v) 
throughout  the  interval  (v  - 1,  v)  ;  and  this  cannot  be  true  for  all  values  of  v. 

Examples  LXX.     1.    Prove  that 

2.  Prove  that  -i7r<2  -^— -^hn.  (Math.  Trip.  1909.) 

3.  Prove  that  if  to  >  0  then 

J_  _1 _1  TO  +  1 

TO2      (to  +  1)2      (to  +  2)2      '"  TO 

175.     The  series  S«_s.     By  far  the  most  important  applica- 
tion of  the  Integral  Test  is  to  the  series 

1-8  +  2-*  +  3-*  +  . . .  +  n-s  +  . .  i; 

where  s  is  any  rational  number.     We  have  seen  already  (§77  and 
Exs.  lxvii.  14,  lxix.  1)  that  the  series  is  divergent  when  s  =  1. 

If  s  ^  0  then  it  is  obvious  that  the  series  is  divergent.     If 
s  >  0  then  un  decreases  as  n  increases,  and  we  can  apply  the  test. 

Here 

'idx     f-s-l 


*(*)=/' 


l  X8  1-5     ' 

unless  s=l.     If  5  >  1  then  £1-s-^  0  as  £-*-  oo  ,  and 

say.  And  if  5  <  1  then  £1_s-»-  oo  as  £  -^  x  ,  and  so  <I>  (£)  -^  oc  . 
Thus  the  series  %n~s  is  convergent  if  s>\,  divergent  if  s  ^  1,  and  in 
the  first  case  its  sum  is  less  than  s/(s  —  1). 

So  far  as  divergence  for  s<l  is  concerned,  this  result  might  have 
been  derived  at  once  from  comparison  with  2(l/«),  which  we  already  know 
to  be  divergent. 

It  is  however  interesting  to  see  how  the  Integral  Test  may  be  applied  to 
the  series  2  (l/»),  when  the  preceding  analysis  fails.     In  this  case 

'  dx 


•(0 


and  it  is  easy  to  see  that  <E>  (£)-»-  x>  as  £-*-  o> .     For  if  |>2"  then 
'°~"clv       [*dx  .    [*dx  .        .    [^ld_x 

X 


*($)>/   —  =    — +    —+...+ 

J  i    *     7 1  A"     j  2  #  y  2 


320  THE    CONVERGENCE    OF    INFINITE    SERIES  [VIII 

But  by  putting  x  =  2ru  we  obtain 

f*+1dx=  f*du 

J  2'        X    ~  J  1    U  ' 

[2du 
and  so  <t>  (£)>»  I    —  5  which  shows  that  *(£)-»-  co  as  £  -*-oo . 

Examples  LXXI.     1.     Prove  by  an  argument  similar  to  that  used  above, 

[idx 
and  without  integration,  that  $  (£)  =  /     -^-,  where  s<  1,  tends  to  infinity  with  £. 

2.  The  series  2w-2,  2ft-3/2,  2?i_11/1°  are  convergent,  and  their  sums  are 
not  greater  than  2,  3,  1-1  respectively.  The  series  2?i-1/2,  2n~wln  are 
divergent. 

3.  The  series  2  na '/(«.'+ a),  where  a>0,  is  convergent  or  divergent  accord- 
ng  as  t>\+s  or  i  <  1  +s.     [Compare  with  2  m.8-'.] 

4.  Discuss  the  convergence  or  divergence  of  the  series 

2  {axn*i  +  a2ns2+  ,..+aknSk)/(b1nt^  +  b2nt2+  ...+bintl)_ 

where  all  the  letters  denote  positive  numbers  and  the  s's  and  fs  are  rational 
and  arranged  in  descending  order  of  magnitude. 

5.  Prove  that 

^<27l  +  372  +  473  +  -<^7r  +  1)- 

{Math.  Trip.  1911.) 

6.  If  cf>  (n)  -*-l>l  then  the  series  2  n  ~  $  \n)  is  convergent.  If  (p  (n)  -»- 1  <  1 
then  it  is  divergent. 

176.  Cauchy's  Condensation  Test.  The  second  of  the 
two  tests  mentioned  in  §172  is  as  follows:  ifun  =  <p(n)  is  a 
decreasing  function  of  n,  then  the  series  20(«)  is  convergent  or 
divergent  according  as  22n  0  (2n)  is  convergent  or  divergent. 

We  can  prove  this  by  an  argument  which  we  have  used 
already  (§  77)  in  the  special  case  of  the  series  2(l/n).  In  the 
first  place 

0(3)  +  0(4)  =£20  (4), 

0(5) +  0(6) +...+0(8)2;  40  (8), 


0  (2M  +  1)  +  0  (2"  +  2)  +  . . .  +  0  (2»+1)  ^  2n  0  (2n+1). 
If  £2w0(2n)  diverges  then  so  do  22n+10(2n+1)  and  2  2"  0  (2n+1), 
and  then  the  inequalities  just  obtained  show  that  2<f)(n)  diverges. 


175-177]  AND   INFINITE   INTEGRALS  321 

On  the  other  hand 

0(2) +  0(3)  ^20  (2),     0(4) +  0(5)+...+ 0(7)  ^40  (4), 

and  so  on.  And  from  this  set  of  inequalities  it  follows  that 
if  2  2n  0  (2'1)  converges  then  so  does  2  0  (n).  Thus  the  theorem  is 
established. 

For  our  present  purposes  the  field  of  application  of  this  test  is 
practically  the  same  as  that  of  the  Integral  Test.  It  enables  us 
to  discuss  the  series  2  n~s  with  equal  ease.  For  2  n~s  will  converge 
or  diverge  according  as  2  2"  2~ns  converges  or  diverges,  i.e.  ac- 
cording as  s  >  1  or  s  £  1. 

Examples  LXXII.  1.  Show  that  if  a  is  any  positive  integer  greater 
than  1  then  2  0(?i)  is  convergent  or  divergent  according  as  2aw0(an)  is 
convergent  or  divergent.  [Use  the  same  arguments  as  above,  taking  groups 
of  a,  a2,  a3,  ...  terms.] 

2.  If  22n<p  (2™)  converges  then  it  is  obvious  that  lim  2"0  (2")  =  0.  Hence 
deduce  Abel's  Theorem  of  §  173. 

177.  Infinite  Integrals.  The  Integral  Test  of  §  174  shows 
that,  if  0  (x)  is  a  positive  and  decreasing  function  of  x,  then  the 
series  2  0  (n)  is  convergent  or  divergent  according  as  the  integral 
function  <I>  (#)  does  or  does  not  tend  to  a  limit  as  x  -*-  oo .  Let 
us  suppose  that  it  does  tend  to  a  limit,  and  that 

lim   [    <p(t)dt  =  l. 

x-*-*>  J  1 

Then  we  shall  say  that  the  integral 

0  (0  dt 


/; 


ts  convergent,    awe?   has   the   value   I;    and   we   shall   call   the 
integral  an  infinite  integral. 

So  far  we  have  supposed  0  (t)  positive  and  decreasing.  But  it 
is  natural  to  extend  our  definition  to  other  cases.  Nor  is  there 
any  special  point  in  supposing  the  lower  limit  to  be  unity.  We 
are  accordingly  led  to  formulate  the  following  definition : 

If<p  (t)  is  a  function  oft  continuous  when  t  =  a,  and 

lim   I"  <f>(t)dt  =1, 

x-^x>  J  a 

n.  21 


322  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

then  we  shall  say  that  the  infinite  integral 

r  ${t)dt a) 

J  a 

is  convergent  and  has  the  value  I. 

The  ordinary  integral  between  limits  a  and  A,  as  denned  in 
Ch.  VII,  we  shall  sometimes  call  in  contrast  a  finite  integral. 

On  the  other  hand,  when 


\ cj>(t)dt 

J  a 


•CO, 


we  shall  say  that  the  integral  diverges  to  co ,  and  we  can  give  a 
similar  definition  of  divergence  to  —  oo .  Finally,  when  none  of 
these  alternatives  occur,  we  shall  say  that  the  integral  oscillates, 
finitely  or  infinitely,  as  x  •*■  oo  . 

These  definitions  suggest  the  following  remarks. 


(i)     If  we  write  j    4>{t)dt  =  $>  (x), 

J  a 


then  the  integral  converges,  diverges,  or  oscillates  according  as  *  (.r)  tends  to 
a  limit,  tends  to  oo  (or  to  -  oo  ),  or  oscillates,  as  #->-  oo  .  If  4>  (x)  tends  to  a 
limit,  which  we  may  denote  by  3>  (oo  ),  then  the  value  of  the  integral  is  *  (oo  ). 
More  generally,  if  $  (x)  is  any  integral  function  of  <f>  (x),  then  the  value  of  the 
integral  is  4>  (oo  )  —  $  (a). 

(ii)  In  the  special  case  in  which  <j>  (t)  is  always  positive  it  is  clear 
that  *(#)  is  an  increasing  function  of  x.  Hence  the  only  alternatives  are 
convergence  and  divergence  to  oo . 

(hi)  The  integral  (1)  of  course  depends  on  a,  but  is  quite  independent  of  t, 
and  is  in  no  way  altered  by  the  substitution  of  any  other  letter  for  t  (cf. 

§  157). 

(iv)  Of  course  the  reader  will  not  be  puzzled  by  the  use  of  the  term 
infinite  integral  to  denote  something  which  has  a  definite  value  such  as 
2  or  \ir.  The  distinction  between  an  infinite  integral  and  a  finite  integral 
is  similar  to  that  between  an  infinite  series  and  a  finite  series :  no  one  supposes 
that  an  infinite  series  is  necessarily  divergent. 

(v)  The  integral  I  $  (t)  dt  was  defined  in  §§  156  and  157  as  a  simple 
J  a 
limit,  i.e.  the  limit  of  a  certain  finite  sum.  The  infinite  integral  is  therefore 
the  limit  of  a  limit,  or  what  is  known  as  a  repeated  limit.  The  notion  of  the 
infinite  integral  is  in  fact  essentially  more  complex  than  that  of  the  finite 
integral,  of  which  it  is  a  development. 


177,  178]  AND   INFINITE    INTEGRALS  323 

(vi)     The  Integral  Test  of  §  174  may  now  be  stated  in  the  form  :  if  <fi  (x)  is 
positive  and  steadily  decreases  as  x  increases,  then  the  infinite  series  2$  (n)  and  the 

infinite  integral  I     <£  (x)  dx  converge  or  diverge  together. 

(vii)    The  reader  will  find  no  difficulty  in  formulating  and  proving  theorems 
for  infinite  integrals  analogous  to  those  stated  in  (1) — (6)  of  §  77.     Thus  the 

result  analogous  to  (2)  is  that  if  j     cf>  (x)  dx  is  convergent,  and  b>a,  then 

J  a 

J     (f>  (x)  dx  is  convergent  and 

■b 


I    <f>(x)dx  =      4>(.v)dx+       cf)(x)dx. 
J  a  J  a  J  b 


178.  The  case  in  which  <£  (V)  is  positive.  It  is  natural 
to  consider  what  are  the  general  theorems,  concerning  the  con- 
vergence or  divergence  of  the  infinite  integral  (1)  of  §  177, 
analogous  to  theorems  A — D  of  §  167.  That  A  is  true  of  integrals 
as  well  as  of  series  we  have  already  seen  in  §  177,  (ii).  Corre- 
sponding to  B  we  have  the  theorem  that  the  necessary  and  sufficient 
condition  for  the  convergence  of  the  integral  (1)  is  that  it  should  be 
possible  to  find  a  constant  K  such  that 

t'%(f>(t)dt<K 

J  a 

for  all  values  of  x  greater  than  a. 

Similarly,   corresponding   to    C,    we    have    the    theorem :    if 
<f)  (x)  dx  is  convergent,  and  y{r  (x)  g  Kcf>  (x)  for  all  values  of  x 

greater  than  a,  then  I     y}r  (x)  dx  is  convergent  and 


yjr  (x)  dx  ^  K        (f)  (x)  dx. 

J  a 


We  leave  it  to  the  reader  to  formulate  the  corresponding  test  for 
divergence. 

We  may  observe  that  D'Alembert's  test  (§  168),  depending 
as  it  does  on  the  notion  of  successive  terms,  has  no  analogue  for 
integrals ;  and  that  the  analogue  of  Cauchy's  test  is  not  of  much 
importance,  and  in  any  case  could  only  be  formulated  when  we 
have   investigated  in  greater  detail   the  theory  of  the    function 

21—2 


324 


THE    CONVERGENCE    OF    INFINITE    SERIES 


[VIII 


(f)  (#)  =  rx,  as  we  shall  do  in  Ch.  IX.  The  most  important  special 
tests  are  obtained  by  comparison  with  the  integral 

f'dx         .       AN 

L  *    (a>0)- 

whose  convergence  or  divergence  we  have  investigated  in  §  175, 
and  are  as  follows :  if  </>  (x)  <  Kx~s,  where  s  >1,  when  x  =  a,  then 

I     (f)  (x)  dx  is  convergent ;  and  if  </>  (x)  >  Kx~s,  where  s  ^1,  when 

•  a 

x  =  a,  then  the  integral  is  divergent ;  and  in  particular,  if 
lim  Xs  (f>  (x)  =  I,  where  I  >  0,  then  the  integral  is  convergent  or 
divergent  according  as  s  >  1  or  s  g  1. 

There  is  one  fundamental  property  of  a  convergent  infinite  series  in 
regard  to  which  the  analogy  between  infinite  series  and  infinite  integrals 
breaks  down.     If  2<p(n)  is  convergent  then  $(?i)->-0;  but  it  is  not  always 

true,  even  when  <£  (x)  is  always  positive,  that  if   I     (f>  (x)  dx  is  convergent 

then  4>(x)-*-0. 

Consider  for  example  the  function  $  (x)  whose  graph  is  indicated  by  the 
thick  line  in  the  figure.  Here  the  height  of  the  peaks  corresponding  to  the 
points  x=l,  2,  3, ...  is  in  each  case  unity,  and  the  breadth  of  the  peak  corre- 


Fig.  50. 

sponding  to  x=n  is  2/(?i  +  l)2.     The  area  of  the  peak  is  l/(n+l)2,  and  it  is 
evident  that,  for  any  value  of  £, 


/; 


cj>(x)dx<2 


o(«  +  l)2' 

so  that  /    <[>  (x)  dx  is  convergent ;  but  it  is  not  true  that  <£  (x)  -=-0 
J  o 

Examples  LXXIII.     1.     The  integral 

r  axr+pxr-1  +  ...+\ 
J  a  Ax*+Bx*-i  +  ...  +  LclV> 

where  a  and  A  are  positive  and  a  is  greater  than  the  greatest  root  of  the 
denominator,  is  convergent  if  s>r  +  l  and  otherwise  divergent. 


178]  AND   INFINITE   INTEGRALS  325 

2.  Which  of  the  integrals 

fxdx        I'00  dx        fx     dx_        f"  xdx         Cx  x2dx         f"         x2dx 
J  a   «/•'     J  a  X**'      la   C2  +  X2'      J  a   C^  +  X2'      )a   C2+X*'      Ja    a  +  2^2  +  y^ 
are  convergent  ?     In  the  first  two  integrals  it  is  supposed  that  a  >  0,  and 
in  the  last  that  a  is  greater  than  the  greatest  root   (if  any)  of  the  de- 
nominator. 

3.  The  integrals 

I   cos  xdx,       /    am  xdx,       I    cos(ax  +  (3)dx 
J  a  J  a  J  a 

oscillate  finitely  as  £-»-  oo . 

4.  The  integrals 

I    A'cos^efo,       I    x2  sin  xdx,      I   xn  cos  (ax  +  ft)  dx, 
J  a  J  a  J  a 

where  n  is  any  positive  integer,  oscillate  infinitely  as  £^-qo . 

ra 

5.  Integrals  to  -  co  .     If  I    $  (#)  efo  tends  to  a  limits  as  £-*--  co ,  then  we 

say  that  I       <£  (.r)  cfe-  is  convergent  and  equal  to  I.     Such  integrals  possess 

properties  in  every  respect  analogous  to  those  of  the  integrals  discussed  in  the 
preceding  sections  :  the  reader  will  find  no  difficulty  in  formulating  them. 

<5.     Integrals  from  -  co  to  +  co  .     If  the  integrals 

ra  r<* 

I       <j>  (x)  dx,      I    0  (x)  dx 

are  both  convergent,  and  have  the  values  k,  I  respectively,  then  we  say  that 

f  °° 

I      <f>  (x)  dx 

is  convergent  and  has  the  value  k  + 1. 

7.  Prove  that 

f  °      dx    _  [  °"    dx   _  1    f  °*     dx 

8.  Prove  generally  that 

t     (f>(x*)dx=2  (   (j>(x2)dx, 

provided  that  the  integral  I     <f>  (x2)  dx  is  convergent. 

9.  Prove  that  if  /     xcf)  (x2)  dx  is  convergent  then  J      xcf)(x2)dx  =  0. 

Jo  ./-<*> 


326  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

10.  Analogue  of  Abel's  Theorem  of  §  173.     If  ${x)  is  positive  and 

steadily  decreases,  and  I     cp(x)dx  is  convergent,  then  x<j>(x)-a-0.     Prove  this 

J  a 
(a)  by  means  of  Abel's  Theorem  and  the  Integral  Test  and  (6)  directly,  by 
arguments  analogous  to  those  of  §  1 73. 

11.  If  a  =  x0<Xi<x^<  ...  and  ,rn-9-ao,  and  un=  I         <f> (x)  dx, then  the 

n 
convergence  of  I     $  (x)  dx  involves  that  of  2  un.     If  cf>  (x)  is  always  positive 

J  a 

the  converse  statement  is  also  true.  [That  the  converse  is  not  true  in 
general  is  shown  by  the  example  in  which  <f>(x)  =  coax,  xn=mr.~\ 

179.  Application  to  infinite  integrals  of  the  rules  for 
substitution  and  integration  by  parts.  The  rules  for  the 
transformation  of  a  definite  integral  which  were  discussed  in 
|  161  may  be  extended  so  as  to  apply  to  infinite  integrals. 

(1)      Transformation  by  substitution.     Suppose  that 

I™  4>{x)dx    (1) 

J  a 

is  convergent.  Further  suppose  that,  for  any  value  of  £  greater 
than  a,  we  have,  as  in  §  161, 

\* 4>(x)dx=\T 4>{f(t))f{t)dt (2), 

J  a  J  b 

where  a=f(b),  £=/(t).  Finally  suppose  that  the  functional 
relation  x=f{t)  is  such  that  x^cc  ast-^oo.  Then,  making  t 
and  so  £  tend  to  go  in  (2),  we  see  that  the  integral 

r4>{f(t))f(t)dt (3) 

J  b 

is  convergent  and  equal  to  the  integral  (1). 

On  the  other  hand  it  may  happen  that  £-*■  go  as  t --* — oo 
or  as  t-*-  c.     In  the  first  case  we  obtain 

lX (f>(x)dx=   lim   \T^{f{t)}f'{t)dt 

J  a  t-*--xJ  b 

=  ~  lim   fb^{f(t)}f(t)dt  =  -fb    ${f{t)}f(t)dl 

T- »■  —00  J  T  J    —CD 

In  the  second  case  we  obtain 

/     <f>(x)dx  =  \im  W{fit)}f'(t)dt (4X 

We  shall  return  to  this  equation  in  §  181. 


178,  179]  AND   INFINITE   INTEGRALS  327 

There  are  of  course  corresponding  results  for  the  integrals 

fa  r<*> 

J       <f>  (x)  dx,       I       <fi  (x)  dx, 

J    —  CO  J    —00 

which  it  is  not  worth  while  to  set  out  in  detail :  the  reader  will 
be  able  to  formulate  them  for  himself. 

Examples  LXXIV.  1.  Show,  by  means  of  the  substitution  x=ta, 
that  if  s  >  1  and  a  >  0  then 

f     x~»dx=a  f     ta(1~$)-1dt; 

and  verify  the  result  by  calculating  the  value  of  each  integral  directly. 

f° 

2.  If  /  0  (x)  dx  is  convergent  then  it  is  equal  to  one  or  other  of 

J  a 

a  4>(at  +  P)dt,       -a  \  4>(at  +  p)dt, 

J  (a-j8)/a  J  -« 

according  as  a  is  positive  or  negative. 

3.  If  (f>  (x)  is  a  positive  and  steadily  decreasing  function  of  x,  and  a  and 
/3  are  any  positive  numbers,  then  the  convergence  of  the  series  2  0  (n)  implies 
and  is  implied  by  that  of  the  series  2$(an  +  /3). 

[It  follows  at  once,  on  making  .  the  substitution  x=at-\-(3,  that  the 
integrals 

I     d>(x)dx,      I  (f>(at  +  @)dt 

J  a  J  («-0)/a 

converge  or  diverge  together.     Now  use  the  Integral  Test.] 

f  °°        rfr 

4.  Show  that         .-        .    ,   =  U.     [Put*=*2.] 

5.  Show  that  |     tt— — r»  dx  =  \ir. 

Jo   (>■+%) 

[Put  x  =  t2  aud  integrate  by  parts.] 

6.  If  <£  (x)->~k  as  .r-s-oo ,  and  cf>  (x)-*-fr  as  x^-  -  oo  ,  then 

[      {0  (x -  a)  -  0  (x -  b)}  dx=  -(a-b)  (h  -  h). 

[For  I  *    {<f>(x-a)-c}>(x-b)}dx=  \       <$>{x-a)dx-  \       <f>(x-b)dx 

J  -i>  J  -t  J  -r 

=  1  *~a  <}>(t)dt-f  '  ~b<t>(t)dt=  f~i~b<i>(t)dt-  yr'4>{t)dt. 


328 


THE   CONVERGENCE   OF   INFINITE   SERIES 


[viu 


The  first  of  these  two  integrals  may  be  expressed  in  the  form 

r-t'-b 
(a-b)k+  I  pdt, 

J   -f'-a 

where  p->-0  as  |'-^oo,  and  the  modulus  of  the  last  integral  is  less  than  or 
equal  to  |  a  -  b  \  k,  where  k  is  the  greatest  value  of  p  throughout  the  interval 
(-tj'-a,  -£'-&).     Hence 

<p{t)dt+(a-b)Jc. 

•e-a 

The  second  integral  may  be  discussed  similarly.] 

(2)  Integration  by  parts.  The  formula  for  integration  by 
parts  (§  161)  is 

*/(«)  </>'  (as)  dx  =/(£)  0  (f)  -/(a)  tf>  (a)  -  f  */'  (*)  *  (•)  dx- 

z  •>  a 

Suppose  now  that  f-^-oo  .  Then  if  any  two  of  the  three  terms 
in  the  above  equation  which  involve  £  tend  to  limits,  so  does  the 
third,  and  we  obtain  the  result 

f  "/(«)  f  («)  cfe  =  Km  /(f)  £  ( j)  -/(a)  <£  (a)  -  f  °°/  («)  </>  (*)  dx. 

There  are  of  course  similar  results  for  integrals  to  —  oo ,  or  from 

—  oo   to  oo  . 


Examples  LXXV.     1.     Show  that 


(1+0!} 


dx=\ 


dx 


{\+x) 


—r.dx- 


,dx  =  l. 


'o   (1+^)4    '      3Jo  (l+#)3 

3.  If  m  and  n  are  positive  integers,   and  Im,n—  j     n  ,     \m  +  w>   then 
Im,n  —  {mftm+n  —  l)}/m_1-n.     Hence  prove  that  Jm,n=m  !  (»-2)!/(ra-Hi  — 1)!. 

4.  Show  similarly  that  if    /m,B=  I      n+a^m'+n    tnen 

4i,n=W(TO  +  'l-1)}/m-i,n,     2/m,n=»i!(»-2)  !/(m  +  »-l)!. 
Verify  the  result  by  applying  the  substitution  x  =  t2  to  the  result  of  Ex.  3. 

180.  Other  types  of  infinite  integrals.  It  was  assumed, 
in  the  definition  of  the  ordinary  or  finite  integral  given  in 
Ch.  VII,  that  (1)  the  range  of  integration  is  finite  and  (2)  the 

subject  of  integration  is  continuous. 

It  is  possible,  however,  to  extend  the  notion  of  the  'definite 
integral '  so  as  to  apply  to  many  cases  in  which  these  conditions 


179,  180]  AND   INFINITE   INTEGRALS  329 

are  not  satisfied.  The  'infinite'  integrals  which  we  have  discussed 
in  the  preceding  sections,  for  example,  differ  from  those  of  Ch.  VII 
in  that  the  range  of  integration  is  infinite.  We  shall  now  suppose 
that  it  is  the  second  of  the  conditions  (1),  (2)  that  is  not  satisfied. 
It  is  natural  to  try  to  frame  definitions  applicable  to  some  such 
cases  at  any  rate.  There  is  only  one  such  case  which  we  shall 
consider  here.  We  shall  suppose  that  cj)(x)  is  continuous  throughout 
the  range  of  integration  (a,  A)  except  for  a  finite  number  of  values 
of  x,  say  a;  =  £1}  f2,  ...,and  that  <f>(x)  -^cc  or  <£  (a?)-*-—  oo  as  x  tends 
to  any  of  these  exceptional  values  from  either  side. 

It  is  evident  that  we  need  only  consider  the  case  in  which 
(a,  A)  contains  one  such  point  £.  When  there  is  more  than  one  such 
point  we  can  divide  up  {a,  A)  into  a  finite  number  of  sub-intervals 
each  of  which  contains  only  one ;  and,  if  the  value  of  the  integral 
over  each  of  these  sub-intervals  has  been  defined,  we  can  then 
define  the  integral  over  the  whole  interval  as  being  the  sum  of 
the  integrals  over  each  sub-interval.  Further,  we  can  suppose 
that  the  one  point  f  in  {a,  A)  comes  at  one  or  other  of  the 
limits  a,  A.     For,  if  it  comes  between  a  and  A,  we  can  then 

[A 

define  I     </>  (x)  dx  as 

J  a 

I    (/>  (x)  dx  +  I     cf>  (x)  dx, 

J  a  J  £ 

assuming  each  of  these  integrals  to  have  been  satisfactorily  de- 
fined. We  shall  suppose,  then,  that  £  =  a ;  it  is  evident  that  the 
definitions  to  which  we  are  led  will  apply,  with  trifling  changes,  to 
the  case  in  which  £=  A. 

Let  us  then  suppose  </>  (x)  to  be  continuous  throughout  (a,  A) 
except  for  x  =  a,  while  <f)(x)-*~oc  as  x^-a  through  values  greater 
than  a.     A  typical  example  of  such  a  function  is  given  by 

<f>  (x)  =  (x  —  a)~s, 

where  s  >  0  ;  or,  in  particular,  if  a  =  0,  by  <£  (x)  =  x~s.  Let  us 
therefore  consider  how  we  can  define 

fAd4 ax 

Jo   &" 
when  s  >  0. 


330  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

The  integral  |     y*~-  dy  is  convergent  if  s  <  1  (§  175)  and  means 

J  i/A 
I'V 

lim  I      ys~2  dy.     But  if  we   make  the  substitution  y  =  1/x,  we 


obtain 

rr,  fA 


r 


ys~'~  dy  =  I      x  s  dx. 
Thus    lim  /      x~s  dx,  or,  what    is    the  same  thing, 

«-»■»  J  lln 


lim    I    x  s  dx, 


exists  provided  that  s  <  1 ;  and  it  is  natural  to  define  the  value  of 
the  integral  (1)  as  being  equal  to  this  limit.    Similar  considerations 

fA 
lead  us  to  define  I     (x  —  a)~s  dx  by  the  equation 

\     (x  —  a  )~s  dx  =  lim    I       (x  —  a)~s  dx. 

Ja  e-»»+oia+e 

We  are  thus  led  to  the  following  general  definition :  if  the  integral 

e'A 

I       <f>  (x)  dx 

J  a+e 

tends  to  a  limit  I  as  e^+  0,  we  shall  say  that  the  integral 

ca 
I    (f>  (x)  dx 

•  a 

is  convergent  and  has  the  value  I. 

Similarly,  when  <f>(x)^oc  as  x  tends  to  the  upper  limit  A,  we 

[A 
define       <j>(x)dx  as  being 


rA-e 
lim    /        <f)  (x)  dx : 

(-*-  +  oJ  a 

and  then,  as  we  explained  above,  we  can  extend  our  definitions  to 
cover  the  case  in  which  the  interval  (a,  A)  contains  any  finite 
number  of  infinities  of  </>  (x). 

An  integral  in  which  the  subject  of  integration  tends  to  oo 
or  to  —  oo  as  x  tends  to  some  value  or  values  included  in  the  range 
of  integration  will  be  called  an  infinite  integral  of  the  second  kind : 
the  first  kind  of  infinite  integrals  being  the  class  discussed  in 
§§  177  et  seq.  Nearly  all  the  remarks  (i) — (vii)  made  at  the  end  of 
§  177  apply  to  infinite  integrals  of  the  second  kind  as  well  as  to 
those  of  the  first. 


180,  181]  AND   INFINITE   INTEGRALS  331 

181.     We  may  now  write  the  equation  (4)  of  §  179  in  the  form 

r<t>{X)dx=\c<t>{f(t)}f(t)dt (i). 

J  a  J  b 

The  integral  on  the  right-hand  side  is  denned  as  the  limit,  as  t-»-c,  of  the 
corresponding  integral  over  the  range  (6,  r),  i.e.  as  an  infinite  integral  of  the 
second  kind.  And  when  $  {/(*)}  /'  (0  has  an  infinity  at  t  =  c  the  integral  is 
essentially  an  infinite  integral.  Suppose  for  example,  that  0  (x)  =  (l+x)~m, 
where  Km<2,  and  a  =  0,  and  that/(0  =  */(l -*)•  Then  6  =  0,  c=l,  and  (1) 
becomes 

("/i—  ^=  (\l-()m-2dt    (2); 

Jo  (l  +  *')m      V 

and  the  integral  on  the  right-hand  side  is  an  infinite  integral  of  the  second 

kind. 

On  the  other  hand  it  may  happen  that  </>  {f(t)}f  (t)  is  continuous  for  t=c. 

In  this  case 

jCb<p{f(t)}f'(t)dt 

is  a  finite  integral,  and 

lim  fT<f>{f(t)}f'(t)dt=[C<p{/(t)}f'(t)dt, 

t-*-c  J  b  J  b 

in  virtue  of  the  corollary  to  Theorem  (10)  of  §  160.  In  this  case  the 
substitution  x=f{t)  transforms  an  infinite  into  a  finite  integral.  This  case 
arises  if  m  2.  2  in  the  example  considered  a  moment  ago. 

Examples  LXXVI.     1.     If  <j>(x)  is  continuous  except  for  x=a,  while 

fA 
<j>(x)-*-cc  &sx-*-a,  then  the  necessary  and  sufficient  condition  that  I      0  (%)  dx 


should  be  convergent  is  that  we  can  find  a  constant  K  such  that 

I       (f>  (x)  dx  <  K 

J  a+c 

for  all  values  of  e,  however  small  (cf.  §  178). 

It  is  clear  that  we  can  choose  a  number  A'  between  a  and  A,  such  that 
<f>{x)  is  positive  throughout  (a,  A').  If  cf>(x)  is  positive  throughout  the 
whole  interval  (a,  A)  then  we  can  of  course  identify  A'  and  A.     Now 

fA  fA'  fA 

I        <j>  (x)  dx  =  I        <f>  (x)  dx  +  I      cf)  (x)  dx. 

J  a-e  J  a-*e  J  A' 

The  first  integral  on  the  right-hand  side  of  the  above  equation  increases 
as  e  decreases,  and  therefore  tends  to  a  limit  or  to  oo  ;  and  the  truth  of  the 
result  stated  becomes  evident. 

If  the  condition  is  not  satisfied  then  I        (jj(x)dx-*-cc  .    We  shall  then  say 

J  a—e 

that  the  integral  I    0  (x)  dx  diverges  to  oo  .     It  is  clear  that,  if  0  (x)  -*-  x 

J  a 
as  x-*-a+0,  then  convergence  and  divergence  to  oo  are  the  only  alternatives 

for  the  integral.     We  may  discuss  similarly  the  case  in  which  0  (x)  -*--  oo . 


332  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

2.  Prove  that 

fA,         s       7       U-a)1-* 
Jj.v-a)sdx  =  K     l_1s 

if  5  <  1,  while  the  integral  is  divergent  if  s  S  1. 

3.  If  <f>(x)-»-ao  as  x->-a  +  Q  and  rp  (x)  <  K  (x  —  a)  ~ s,  where  s<l,  then 


A 

cji(x)dx  is  convergent;  and  if  cp(x)>  K(x  —  a)~*,  where  s  >l,  then  the 


/. 

integral  is  divergent.     [This  is  merely  a  particular  case  of  a  general  com- 
parison theorem  analogous  to  that  stated  in  §  178.] 

4.  Are  the  integrals 

l'A  dx  fA  dx  fA  dx 

j  a  J{(x-a)(A-x)}  '      J  a  (A-x)*/(x-a) '      J  a  (4-tf)^(4 -J?) ' 

fA       dx  fA       dx  fA    dx  fA    dx 

jasj^-a'y     JaV(A3-a?y     Ja&^d*     J  a  A*^ 

convergent  or  divergent  ? 

fl   dx      fa+1       dr 

5.  The  integrals  /      r>~  >  J        ajr~~~\  are  convergent,  and  the  value  of 

J  -i  \fx   J  a -iv (•*•'-«) 
each  is  zero. 

C  IT  ($£ 

6.  The  integral  /     — — -  is  convergent.     [The  subject  of  integration 

J  o  v(sm  x) 

tends  to  qo  as  x  tends  to  either  limit.] 

7.  The  integral  /    —. r-  is  convergent  if  and  only  if  s  <  1. 

,— — r,  dx  is  convergent  if  t  <  s  + 1. 

o    (sin*)' 

/       SI  11  1* 

9.     Show  that  I    — ir^x,  where  h  >0,  is  convergent  if  p<  2.     Show  also 
that,  if  0  <p  <  2,  the  integrals 

f*smx  .        f^"s\nx  ,        f^sinx  7 

I     — r  dx,     I     - dx,     I      — -—  dx,  ... 

JO     #P  J*       X"  '      Jsrr    -Vp 

alternate  in  sign  and  steadily  decrease  in  absolute  value.     [Transform  the 
integral  whose  limits  are  kn  and  (k  +  \)w  by  the  substitution  x  =  kir+y.~\ 

— —dx,  where  0</?<2,  attains  its  greatest  value 
o    xV 
when  h=ir.  (Math.  Trip.  1911.) 

11.  The  integral  I      (cos  x)1  (sin  x)mdx  is  convergent  if  and  only  if  l>  -  1, 

Jo 
m>  - 1. 

f  °°  x8~*dv 

12.  Such  an  integral  as  /    — : ,  where  s<l,  does  not  fall  directly 

under  any  of  our  previous  definitions.   For  the  range  of  integration  is  infinite 


181]  AND   INFINITE   INTEGRALS  333 

and  the  subject  of  integration  tends  to  oo    &sx^>-  +  0.      It  is  natural  to 
define  this  integral  as  being  equal  to  the  sum 


P  xs~1dx       fx  xs~1dx 
Jo"l"+^T  +  J1     1+x    ' 


provided  that  these  two  integrals  are  both  convergent. 

The  first  integral  is  a  convergent  infinite  integral  of  the  second  kind 
if  0<s<l.  The  second  is  a  convergent  infinite  integral  of  the  first  kind  if 
s  <1.  It  should  be  noted  that  when  s  >  1  the  first  integral  is  an  ordinary 
finite  integral ;  but  then  the  second  is  divergent.  Thus  the  integral  from  0  to 
oo  is  convergent  if  and  only  if  0  <  s  <  1. 

f  *  Xs'1 

13.     Prove  that  I dx  is  convergent  if  and  only  if  0  <  5  <  t. 

Jo  l+x 

/'  X£8  —  l  _  yA  —  1 
- — - — dx  is  convergent  if  and  only  if  0  <  s<  1, 

0<t<\.  [It  should  be  noticed  that  the  subject  of  integration  is  undefined 
when  x=\  ;  but  (xs~1-xt~1)/(l  -a?)-*- 1  —  s  as  x-^-l  from  either  side;  so  that 
the  subject  of  integration  becomes  a  continuous  function  of  x  if  we  assign  to  it 
the  value  t  —  s  when  x—  1. 

It  often  happens  that  the  subject  of  integration  has  a  discontinuity  which 
is  due  simply  to  a  failure  in  its  definition  at  a  particular  point  in  the  range 
of  integration,  and  can  be  removed  by  attaching  a  particular  value  to  it  at 
that  point.  In  this  case  it  is  usual  to  suppose  the  definition  of  the  subject 
of  integration  completed  in  this  way.     Thus  the  integrals 


f  \*  sin  mx  7        f  i*  sin  mx  , 

I      ax,     I      —. dx 

Jo        no  Jo     sin  x 


are  ordinary  finite  integrals,  if  the  subjects  of  integration  are  regarded  as 
having  the  value  m  when  x=0.] 

15.  Substitution  and  integration  by  parts.  The  formulae  for  trans- 
formation by  substitution  and  integration  by  parts  may  of  course  be  extended 
to  infinite  integrals  of  the  second  as  well  as  of  the  first  kind.  The  reader 
should  formulate  the  general  theorems  for  himself,  on  the  lines  of  §  179. 

16.  Prove  by  integration  by  parts  that  if  s  >  0,  t  >  1,  then 

[1z»-1Q.-xjt-1da;  =  —  /"\b«(1 -*)*-»<** 

/i  rs-i,7r        r  °°  f-srff 
±_^:=       »«     [Put*=l/<.] 
0     1+X         Ji    l+t  '    J 

18.     IfO<s<lthen        ^-—^—dx^i    -1—^=        f! 
Jo       1+J?  jo    l  +  «       Jo     J 


dt 


l  +  t 

19.     Ifa  +  6>0then 

,     ,       'Vu j*  -    „  "",  ,-, .  (Math.  Trip.  1909.) 

6  (x+a)J(x-b)     >J{a  +  b)  r  ' 


/; 


334 


THE   CONVERGENCE   OF   INFINITE   SERIES 


[VIII 


20.  Show,  by  means  of  the  substitution  x=t/(l  -  t),  that  if  I  and  m  are 

both  positive  then 

x  doe=       tl-^{\-t)m-ldt. 

Jo   (l  +  ^  +  m  jo        V 

21.  Show,  by  means  of  the  substitution  x=pt/(p  +  \  -t),  that  if  I,  m,  and 

p  are  all  positive  then 

dr                   IP 
—  = j    fl-i  (l  -  i)™-1  dt 

ClJ-p)lprn  J  Ql         \l       l'  al- 

xdx 

^a){b-x)\ 

(i)  by  means  of  the  substitution  x  =  a  +  (b  —  a)  t2,  (ii)  by  means  of  the  substitu- 
tion {b—x)j{x  —  a)=t,  and  (iii)  by  means  of  the  substitution  x=a  cos2  t  +  b  sin2 1. 


xl~x  (\-x)m-1  , 


_  ,,    ,  fb  dx  ,  fb  xdx  .     .       .  x 

22.     Prove  that  I     -/77 ryr r-,  =  *"  ancl  I     7  u \7I vi  =  t*  \a  +  b)-> 

J  aSJ{{x-a){b-x))  J  as/{(x-a)(b-x)} 


23.     If  s>-l  then 


Jo  JoV(l  -a")        Jo   V(I-a-)  Jo  >/a? 

24.  Establish  the  formulae 

P     f(x)dx  P"'     ,,     ■        AS     ,A 

/^ra^)r2/.i'y(;cos2"+65i"!S)^ 

/:/{\As?f>}  rf'l'=4a  /."/fton *> cos  *  8i" *  * 

25.  Pi-ovq  that 

P dx /l       j_\ 

Mi+*)(2+*M*(i -•*)}"*  w»    s'6/ 

[Put  .?;=sin2<9  and  use  Ex.  lxiii.  8.J  (vJ/a*/i   TWp.  1912.) 

182.  Some  care  has  occasionally  to  be  exercised  in  applying  the  rule 
for  transformation  by  substitution.  The  following  example  affords  a  good 
illustration  of  this. 


Let 


J=  /  '  (a?-Gx+lS)dx. 


We  find  by  direct  integration  that  J =  48.    Now  let  us  apply  the  substitution 

?/  =  .r2-6.r+13, 
which  gives  x=Z±.-J(y  —  £).    Since  y  =  8  when  x=l  and  #  =  20  when  x  =  7,  we 
appear  to  be  led  to  the  result 

The  indefinite  integral  is 

l(y_  4)3/2+4  (y- 4)1/2, 

arid  so  we  obtain  the  value  ±i%1,  which  is  certainly  wrong  whichever  sign  we 
choose. 


181-183]  AND   INFINITE   INTEGRALS  335 

The  explanation  is  to  be  found  in  a  closer  consideration  of  the  relation 
between  x  and  y.  The  function  #2-6a'+13  has  a  minimum  for  .-£=3,  when 
y  =  ±.  As  x  increases  from  1  to  3,  y  decreases  from  8  to  4,  and  dxjdy  is 
negative,  so  that 

dx_  1 

dy~     2v/(y-4)- 

As  x  increases  from  3  to  7,  y  increases  from  4  to  20,  and  the  other  sign  must 
be  chosen.     Thus 

a  formula  which  will  be  found  to  lead  to  the  correct  result. 

Similarly,  if  we  transform   the   integral    I     dx  =  n   by  the   substitution 

#  =  arc  sin  y,  we  must  observe  that  dxjdy  =  \jj{\  —  y2)  or  dxjdy  =  — 1/^(1  —  y2) 
according  as  0  ^  x  <  \n  or  in-  <  x  g.  w. 

Example.     Verify  the  results  of  transforming  the  integrals 

/( ix2  —  x  +  t^  )  dx,      I     cos2  x  dx 
o  Jo 

by  the  substitutions  4x2 -x  +  ^s=y,  .r=arc  sin y  respectively. 

183.  Series  of  positive  and  negative  terms.  Our  defini- 
tions of  the  sum  of  an  infinite  series,  and  the  value  of  an  infinite 
integral,  whether  of  the  first  or  the  second  kind,  apply  to  series 
of  terms  or  integrals  of  functions  whose  values  may  be  either 
positive  or  negative.  But  the  special  tests  for  convergence  or 
divergence  which  we  have  established  in  this  chapter,  and  the 
examples  by  which  we  have  illustrated  them,  have  had  reference 
almost  entirely  to  the  case  in  which  all  these  values  are  positive. 
Of  course  the  case  in  which  they  are  all  negative  is  not  essentially 
different,  as  it  can  be  reduced  to  the  former  by  changing  un  into 
—  un  or  (f>  (x)  into  —  <f>  (x). 

In  the  case  of  a  series  it  has  always  been  explicitly  or  tacitly 
assumed  that  any  conditions  imposed  upon  un  may  be  violated  for 
a  finite  number  of  terms  :  all  that  is  necessary  is  that  such  a 
condition  {e.g.  that  all  the  terms  are  positive)  should  be  satisfied 
from  some  definite  term  onwards.  Similarly  in  the  case  of  an 
infinite  integral  the  conditions  have  been  supposed  to  be  satisfied 
for  all  values  of  x  greater  than  some  definite  value,  or  for  all  values 
of  x  within   some  definite  interval  (a,  a  +  5)  which  includes  the 


336  THE    CONVERGENCE    OF    INFINITE    SERIES  [VIII 

value  a  near  which  the  subject  of  integration  tends  to  infinity. 
Thus  our  tests  apply  to  such  a  series  as 

v  n2  -  10 

since  rc2  —  10  >  0  when  n  =  4,  and  to  such  integrals  as 


r  Sx-7  ,      rn-zx  , 


since  3a;  —  7  >  0  when  a;  >  §,  and  1  —  2x  >  0  when  0  <  x  <  \. 

But  when  the  changes  of  sign  of  un  persist  throughout  the  series, 
i.e.  when  the  number  of  both  positive  and  negative  terms  is  in- 
finite, as  in  the  series  1—  \  +  %  —  £+...;  or  when  </>  (x)  continually 
changes  sign  as  x  -*-  oo  ,  as  in  the  integral 

f  °°  sin  x 


i      x" 


dx, 


or  as  x-*-a,  where  a  is  a  point  of  discontinuity  of  <£  (x),  as  in 
the  integral 

[A   .    f    1    \     dx 
sin ; 

J  a         \x  —  a/  x  —  a 

then  the  problem  of  discussing  convergence  or  divergence  becomes 
more  difficult.  For  now  we  have  to  consider  the  possibility  of 
oscillation  as  well  as  of  convergence  or  divergence. 

We  shall  not,  in  this  volume,  have  to  consider  the  more 
general  problem  for  integrals.  But  we  shall,  in  the  ensuing 
chapters,  have  to  consider  certain  simple  examples  of  series  con- 
taining an  infinite  number  of  both  positive  and  negative  terms. 

184.  Absolutely  Convergent  Series.  Let  us  then  consider 
a  series  %un  in  which  any  term  may  be  either  positive  or 
negative.     Let 

|«n|=«n, 
so  that  an  =  un  if  un  is  positive  and  ctn=  -  un  if  un  is  negative. 
Further,  let  vn  =  vn  or  vn  =  0,  according  as  un  is  positive  or  negative, 
and  wn  =  —  vn  or  wn  =  0,  according  as  un  is  negative  or  positive ; 
or,  what  is  the  same  thing,  let  vn  or  wn  be  equal  to  an  according 
as  un  is  positive  or  negative,  the  other  being  in  either  case  equal 
to  zero.  Then  it  is  evident  that  vn  and  wn  are  always  positive,  and 
that 

un  =  vn  —  wn)    an  =  vn  +  wn. 


183,  184]  AND   INFINITE   INTEGRALS  337 

If,  for  example,  our  series  is  1 -(l,'2)2  +  (l/3)2- ...,  then  «„  =  (  -  l)""1/**2 
and  a,t  =  l//i2,  while  vn  =  ljn2  or  vn=0  according  as  n  is  odd  or  even  and 
wn  =  \jrfi  or  u',h  =  0  according  as  n  is  even  or  odd. 

We  can  now  distinguish  two  cases. 

A.  Suppose  that  the  series  1an  is  convergent.  This  is  the 
case,  for  instance,  in  the  example  above,  where  2an  is 

1 +(1/2)*+ (1/3)*  +  .... 

Then  both  %vn  and  %wn  are  convergent :  for  (Ex.  xxx.  18)  any 
series  selected  from  the  terms  of  a  convergent  series  of  positive 
terms  is  convergent.  And  hence,  by  theorem  (6)  of  §  77,  £wn  or 
%  (vn  —  wn)  is  convergent  and  equal  to  £vn  _  Sw». 

We  are  thus  led  to  formulate  the  following  definition. 

Definition.  When  San  or  S  un  j  is  convergent,  the  series  Xun 
is  said  to  be  absolutely  convergent. 

And  what  we  have  proved  above  amounts  to  this  :  if  "Eun  is 
absolutely  convergent  then  it  is  convergent;  so  are  the  series  formed 
by  its  positive  and  negative  terms  taken  separately ;  and  the  sum  of 
the  series  is  equal  to  the  sum  of  the  positive  terms  plus  the  sum 
of  the  negative  terms. 

The  reader  should  carefully  guard  himself  against  supposing  that  the 
statement  '  an  absolutely  convergent  series  is  convergent ;  is  a  mere  tautology. 
When  we  say  that  2  un  is  '  absolutely  convergent '  we  do  not  assert  directly 
that  2«„  is  convergent:  we  assert  the  convergence  of  another  series  2|zf„|, 
and  it  is  by  no  means  evident  a  priori  that  this  precludes  oscillation  on 
the  part  of  2?</(. 

Examples  LXXVII.  1.  Employ  the  'general  principle  of  convergence' 
(§  84)  to  prove  the  theorem  that  an  absolutely  convergent  series  is  con- 
vergent. [Since  2  |  un  |  is  convergent,  we  can,  when  any  positive  number  b  is 
assigned,  choose  n0  so  that 

K»1+i|+K>i+2[  +  ---+  l"»J<8 

when  n2  >  «i  =  «o-     -^  fortiori 

| «»,  +1+'«»i  +  2+"  .+Uni\<8, 

and  therefore  2?<n  is  convergent.] 

2.  If  2o,t  is  a  convergent  series  of  positive  terms,  and  |  bn  |  g.  Kan ,  then 
2o„  is  absolutely  convergent. 

3.  If  2an  is  a  convergent  series  of  positive  terms,  then  the  series  "2anxn  is 
absolutely  convergent  when  — 1<#<1. 

h.  22 


338  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

4.  If  2  a„  is  a  convergent  series  of  positive  terms,  then  the  series  2  a,,  cos  n  0, 
2"„sinH0  are  absolutely  convergent  for  all  values  of  6.  [Examples  are 
afforded  by  the  series  2r"cos«0,  2?"'  sin  n6  of  §  88.] 

5.  Anv  series  selected  from  the  terms  of  an  absolutely  convergent  series 
is  absolutely  convergent.  [For  the  series  of  the  moduli  of  its  terms  is  a 
selection  from  the  series  of  the  moduli  of  the  terms  of  the  original  series.] 

6.  Prove  that  if  2  |  un  |  is  convergent  then 

|2u*|  g2|w»|, 
and  that  the  only  case  to  which  the  sign  of  equality  can  apply  is  that  in 
which  every  term  has  the  same  sign. 

185.  Extension  of  Dirichlet's  Theorem  to  absolutely- 
convergent  series.  Dirichlet's  Theorem  (§  169)  shows  that  the 
terms  of  a  series  of  positive  terms  may  be  rearranged  in  any  way 
without  affecting  its  sum.  It  is  now  easy  to  see  that  any  abso- 
lutely convergent  series  has  the  same  property.  For  let  %un  be 
so  rearranged  as  to  become  2t<n',  and  let  an',  vn',  wn'  be  formed 
from  un'  as  an,  vn,  wn  were  formed  from  un.  Then  2a,/  is  con- 
vergent, as  it  is  a  rearrangement  of  2a»,  and  so  are  2vn',  %wn', 
which  are  rearrangements  of  2vm,  2«>n.  Also,  by  Dirichlet's 
Theorem,  %vn'  =  2vn  and  %wn'  =  2wn,  and  so 

186.  Conditionally  convergent  series.  B.  We  have 
now  to  consider  the  second  case  indicated  above,  viz.  that  in 
which   the  series  of  moduli  2a»  diverges  to  oo . 

Definition.  If  %un  is  convergent,  but  2  |  un  \  divergent,  the 
original  series  is  said  to  be  conditionally  convergent. 

In  the  first  place  we  note  that,  if  %un  is  conditionally  con- 
vergent, then  the  series  %vn,  Xwn  of  §  184  must  both  diverge  to  oc  . 
For  they  obviously  cannot  both  converge,  as  this  would  involve 
the  convergence  of  %(vn  +  wn)  or  2an.  And  if  one  of  them,  say 
%wn,  is  convergent,  and  2vn  divergent,  then 

N  N  N 

Xun=  2v„-2w„ (1), 

0  0  0 

and   therefore  tends    to    cc    with  N,    which    is    contrary   to    the 
hypothesis  that  tun  is  convergent. 

Hence  2vn,  2wn  are  both  divergent.  It  is  clear  from  equa- 
tion (1)  above  that  the  sum  of  a  conditionally  convergent  series 


184-187]  AND   INFINITE   INTEGRALS  339 

is  the  limit  of  the  difference  of  two  functions  each  of  which  tends 
to  co  with  n.  It  is  obvious  too  that  Xun  no  longer  possesses  the 
property  of  convergent  series  of  positive  terms  (Ex.  xxx.  18),  and 
all  absolutely  convergent  series  (Ex.  lxxvii.  5),  that  any  selection 
from  the  terms  itself  forms  a  convergent  series.  And  it  seems  more 
than  likely  that  the  property  prescribed  by  Dirichlet's  Theorem 
will  not  be  possessed  by  conditionally  convergent  series ;  at  any 
rate  the  proof  of  §  185  fails  completely,  as  it  depended  essentially 
on  the  convergence  of  Xvn  and  %wn  separately.  We  shall  see  in  a 
moment  that  this  conjecture  is  well  founded,  and  that  the  theorem 
is  not  true  for  series  such  as  we  are  now  considering. 

187.  Tests  of  convergence  for  conditionally  convergent 
series.  It  is  not  to  be  expected  that  we  should  be  able  to  find 
tests  for  conditional  convergence  as  simple  and  general  as  those 
of  §§  167  et  seq.  It  is  naturally  a  much  more  difficult  matter  to 
formulate  tests  of  convergence  for  series  whose  convergence,  as  is 
shown  by  equation  (1)  above,  depends  essentially  on  the  cancelling 
of  the  positive  by  the  negative  terms.  In  the  first  instance  there 
are  no  comparison  tests  for  convergence  of  conditionally  convergent 
series. 

For  suppose  we  wish  to  infer  the  convergence  of  %vn  from 
that  of  Xun      We  have  to  compare 

v0+vx+  ...  +  vn,      110  +  1^  +  ...  +  un. 

If  every  u  and  every  v  were  positive,  and  every  v  less  than  the 
corresponding  u,  we  could  at  once  infer  that 

v0+  vx  +  ...  +vn<  v0  +  ...+  un, 

and  so  that  2,vn  is  convergent.  If  the  us  only  were  positive  and 
every  v  numerically  less  than  the  corresponding  u,  we  could  infer 

that 

\v0\  +  \vl\  +  ...  +  \vn\<u0+  ...+  un, 

and  so  that  Svn  is  absolutely  convergent.  But  in  the  general  case, 
when  the  u's  and  vs  are  both  unrestricted  as  to  sign,  all  that  we 
can  infer  is  that 

|  l'o|  +\Vi  |  +  ...  +  \vn\  <|«0|+  ...  +!  Un\. 

This  would  enable  us  to  infer  the  absolute  convergence  of  2vn 
from  the  absolute  convergence  of  Xun;  but  if  Xun  is  only  con- 
ditionally convergent  we  can  draw  no  inference  at  all. 


340  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

Example.  We  shall  see  shortly  that  the  series  1  —  \+\-\  +  ...  is  con- 
vergent. But  the  series  £+£+£+£+...  is  divergent,  although  each  of  its 
terms  is  numerically  less  than  the  corresponding  term  of  the  former  series. 

It  is  therefore  only  natural  that  such  tests  as  we  can  obtain 
should  be  of  a  much  more  special  character  than  those  given  in 
the  early  part  of  this  chapter. 

188.  Alternating  Series.  The  simplest  and  most  common 
conditionally  convergent  series  are  what  is  known  as  alternating 
series,  series  whose  terms  are  alternately  positive  and  negative. 
The  convergence  of  the  most  important  series  of  this  type  is 
established  by  the  following  theorem. 

If  (f>  (n)  is  a  positive  function  of  n  which  tends  steadily  to 
zero  as  n-^oo ,  then  the  series 

*(0)-$(l)+*(2)- 

is  convergent,  and  its  sum  lies  hetween  <f>  (0)  and  <£  (0)  —  (f>  (1). 

Let  us  write  fa,  <£j, ...  for  (f)(0),  <f)  (1),  ... ;  and  let 

sn=  fa-fa  +  fa-  •••  +(-  1)H  fai- 
Then 

S«n+i  —  S2„-i  =  fam  —  #2»+l  =  0>     sm  ~  Sm-2  =  —  (fan-i  ~  fan)  =  0. 

Hence  s0,  s2,  s4, ...,  s2n,  ...  is  a  decreasing  sequence,  and  therefore 
tends  to  a  limit  or  to  -co,  and  slf  ss,  s5, ...,  s2n+1,  ...  is  an  in- 
creasing sequence,  and  therefore  tends  to  a  limit  or  to  oo .  But 
lim  (sm+1  —  sm)  =  lim  (—  l)2'l+1  fan+1  =  0,  from  which  it  follows  that 
both  sequences  must  tend  to  limits,  and  that  the  two  limits  must 
be  the  same.  That  is  to  say,  the  sequence  s0,  slt  ...,sn,  ...  tends  to 
a  limit.  Since  s0  =  fa,  $i  =  fa  —  fa,  it  is  clear  that  this  limit  lies 
between  fa  and  fa  —  <£x. 

Examples  LXXVIII.     1.    The  series 

111  1        1        1 

2  +  3      4  +  '"'     1~72+73~V4"K"' 
s(_l)»        „    (-1)«         „     (-1)»  „     (-1)« 


(7i  + a)'         V(»  +  «)'         (Jn+Ja)'         (Jn  +  Ja)*' 
where  a>0,  are  conditionally  convergent. 

2.     The  series  2  (-1)™  (%  +  «)~s,  where  a>0,  is  absolutely  convergent  if 
s>l,  conditionally  convergent  if  0<5^1,  and  oscillatory  if  s^0. 


187,  188]  AND    INFINITE    INTEGRALS  341 

3.  The  sum  of  the  series  of  §  188  lies  between  sn  and  sH  +  1  for  all  values 
of  n  ;  and  the  error  committed  by  taking  the  sum  of  the  first  n  terms  instead 
of  the  sum  of  the  whole  series  is  numerically  not  greater  than  the  modulus  of 
the  (?i  +  l)th  term. 

4.  Consider  the  series 

(-1)" 
N/rc+(-l)»' 

which  we  suppose  to  begin  with  the  term  for  which  n  =  2,  to  avoid  any 
difficulty  as  to  the  definitions  of  the  first  few  terms.  This  series  may  be 
written  in  the  form 

yrr  (-i)n     i-m  ,  (-ot 

LU/»+(-l)»        s'«>  J         s.'n   J 


r(-i)n        i      i 


say.  The  series  2  \jrn  is  convergent ;  but  2  xn  is  divergent,  as  all  its  terms  are 
positive,  and  lim  ?^H=1.  Hence  the  original  series  is  divergent,  although  it 
is  of  the  form  02  —  $3  +  04-  •••>  where  (pH-^0.  This  example  shows  that  the 
condition  that  (f>a  should  tend  steadily  to  zero  is  essential  to  the  truth  of  the 
theorem.  The  reader  will  easily  verify  that  x/(2w  +  l)  — 1  <s/(2n)  +  l,  so  that 
this  condition  is  not  satisfied. 

5.  If  the  conditions  of  §  188  are  satisfied  except  that  <f>n  tends  steadily 
to  a  positive  limit  I,  then  the  series  2  ( —  1)"  <j>n  oscillates  finitely. 

6.  Alteration  of  the  sum  of  a  conditionally  convergent  series  by- 
rearrangement  of  the  terms.  Let  s  be  the  sum  of  the  scries  1—  |  +  ;^-j  +  ..., 
and  s2n  the  sum  of  its  first  2n  terms,  so  that  lim  s2ll=s. 

Now  consider  the  series 

1+W+HW+ a) 

in  which  two  positive  terms  are  followed  by  one  negative  term,  and  let  t3n 
denote  the  sum  of  the  first  3n  terms.     Then 

1  1  1      1  _1_ 


_1  1  1 

~S2n  +  2n  +  l  +  2n  +  3+'"  +  4»-l 


Now      iim[-Iri-^  +  2;-L^-...+4-LT-l 


=0, 


since    the    sum    of    the    terms  inside    the    bracket    is    clearly  less    than 
ra/(2n  +  l)(2»+2);  and 

,.      /     1              1  1\     ...     1  n         1           ,  [%dx 

\2«  +  2      2m  +  4  An/     *        nr=i\  +  {rjn)       Ji  % 

by  §§  156  and  158.    Hence 

..  ,   [*dx 

J  i  •* 


342  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

and  it  follows  that  the  sum  of  the  series  (1)  is  not  s,  but  the  right-hand  side  of 
the  last  equation.  Later  on  we  shall  give  the  actual  values  of  the  sums  of  the 
two  series  :  see  §  213  and  Ch.  IX,  Misc.  Ex.  19. 

It  can  indeed  be  proved  that  a  conditionally  convergent  series  can  always 
be  so  rearranged  as  to  converge  to  any  sum  whatever,  or  to  diverge  to  co  or 
to  —  oo .     For  a  proof  we  may  refer  to  Bromwich's  Infinite  Series,  p.  68. 

7.    The  series  1  +  7o-72  +  75  +  77-74  +  ■"  diver§es  to  °°  ■     [Here 

1  1  1  n 

'3»-*2»  +  7(2^  +  1)  +  J(2n  +  Sy'"  +  N/(4/i-l)>S2n  +  V(4»-l)' 

where  s2)l=  1  — #s  +  •••  — j^~  >  which  tends  to  a  limit  as  n->-  oo  .] 

189.  Abel's  and  Dirichlet's  Tests  of  Convergence.  A  more  general 
test,  which  includes  the  test  of  §  188  as  a  particular  test  case,  is  the  following. 

Dirichlet's  Test.  If  (j>n  satisfies  the  same  conditions  as  in  §  188,  and  2an 
is  any  series  which  converges  or  oscillates  finitely,  then  the  series 

is  convergent. 

The  reader  will  easily  verify  the  identity 

«O0O  +  «101  +  '--  +  «7l^>re  =  SO(0O-0l)  +  Sl(01-02)  +  ---+S„-l(0„-1-^i))  +  S«^n-. 

where  sn=an  +  ai  +  ...+aa.     Now  the  series  (</>0  —  0i)  +  (</>i  —  <£2)  +  ---  is  con 
vergent,  since  the  sum  to  n  terms  is  0o -</>«.  and  lim<£„  =  0;    and  all  its 
terms  are  positive.     Also  since  2an,  if  not  actually  convergent,  at  any  rate 
oscillates  finitely,  we  can  determine  a  constant  K  so  that   |  sv  |  <  K  for  all 
values  of  v.     Hence  the  series 

2s„(#„-#„+i) 
is  absolutely  convergent,  and  so 

tends  to  a  limit  as  n  -»-  oo .  Also  <f>n,  and  therefore  sH(pn,  tends  to  zero 
And  therefore 

tends  to  a  limit,  i.e.  the  series  2a„$„  is  convergent. 

Abel's  Test.  There  is  another  test,  due  to  Abel,  which,  though  of  less- 
frequent  application  than  Dirichlet's,  is  sometimes  useful. 

Suppose  that  <f)n>  as  in  Dirichlet's  Test,  is  a  positive  and  decreasing 
function  of  n,  but  that  its  limit  as  n  -»-  oo  is  not  necessarily  zero.  Thus  we 
postulate  less  about  <£„,  but  to  make  up  for  this  we  postulate  more  about 
2«n,  viz.  that  it  is  convergent.  Then  we  have  the  theorem:  if  (j>n  is  a  positive 
and  decreasing  function  of  n,  and  "2an  is  convergent,  then  2«„0n  is  convergent. 

For  <f)n  has  a  limit  as  n  -»-  oo ,  say  I :  and  lim  ($„  - 1)  =  0.  Hence,  by 
Dirichlet's  Test,  2an((pn-l)  is  convergent;  and  as  2a„  is  convergent  it 
follows  that  2ancj)n  is  convergent. 


188,  189]  AND   INFINITE   INTEGRALS  343 

This  theorem  may  be  stated  as  follows :  a  convergent  series  remains  con- 
vergent if  we  multiply  its  terms  by  any  sequence  of  positive  and  decreasing 
facto7's. 

Examples  LXXIX.  1.  Dirichlet's  and  Abel's  Tests  may  also  be  established 
by  means  of  the  general  principle  of  convergence  (§  84).  Let  us  suppose, 
for  example,  that  the  conditions  of  Abel's  Test  are  satisfied.  We  have 
identically 

fflj» $»»  +  «m  +  1  </>m  +  1  +  •  •  •  +  an <Pn  =  V  m  (<£ro  ~  </>m  +  l)  +  V,  m  + 1  ($m  + 1  ~  4>m  +  Si) 
+  ---+sm,n-l(4>n-l-<l>n)  +  sm,n<t>n (1)> 

where  sm<  v  =  am  +  am  +  1  +  ...+av. 

The  left-hand  side  of  (1)  therefore  lies  between  h(f)m  and  H<pm,  where  h  and 
H  are  the  algebraically  least  and  greatest  of  sm,m,  snum  +  1,  ...,  sm<n.  But, 
given  any  positive  number  8,  we  can  choose  »i0  so  that  |  sm<v  \<8  when  m  ^  m0, 
and  so 

I  «m</>m  +  «»i  +  l#m  +  l  +  ...+«„ <f>n\<8(f)m  ^  S$l 

when  n>»i>m0.     Thus  the  series  2ancf)n  is  convergent. 

2.  The  series  2  cos  ?i#  and  2  sin  nd  oscillate  finitely  when  6  is  not  a 
multiple  of  tt.  For,  if  we  denote  the  sums  of  the  first  n  terms  of  the  two 
series  by  s„  and  tn,  and  write  z= Cis  0,  so  that  |  z  |=1  and  z  4=  1,  we  have 

1-g" 
1-2 


>»  +  «*J  = 


1  +  U"I 

1 !  < 


and  so  |  sn  |  and  |  tn  |  are  also  not  greater  than  2/\l-z\.  That  the  series  are 
not  actually  convergent  follows  from  the  fact  that  their  nth  terms  do  not  tend 
to  zero  (Exs.  xxiv.  7,  8). 

The  sine  series  converges  to  zero  if  8  is  a  multiple  of  it.  The  cosine 
series  oscillates  finitely  if  6  is  an  odd  multiple  of  n  and  diverges  if  6  is  an 
even  multiple  of  n. 

It  follows  that  if  (f>n  is  a  positive  function  of  n  which  tends  steadily  to 
zero  as  /£->•  co  ,  then  the  series 

2(f>ncosnd,     2(j>nsmnd 

are  convergent^  except  perhaps  the  first  series  when  6  is  a  multiple  of  2n.  In 
this  case  the  first  series  reduces  to  2$,„  which  may  or  may  not  be  conver- 
gent: the  second  series  vanishes  identically.  If  2$„  is  convergent  then  both 
series  are  absolutely  convergent  (Ex.  lxxvii.  4)  for  all  values  of  6,  and  the 
whole  interest  of  the  result  lies  in  its  application  to  the  case  in  which 
2$n  is  divergent.  And  in  this  case  the  series  above  written  are  con- 
ditionally and  not  absolutely  convergent,  as  will  be  proved  in  Ex.  lxxix.  6. 
If  we  put  0  =  tt  in  the  cosine  series  we  are  led  back  to  the  result  of  §  188, 
since  cos  nir  =  (  -  l)n. 

3.  The  series  2?i-scosm#,  2«~ssin?i0  are  convergent  if  s>0,  unless  (in 
the  case  of  the  first  series)  0  is  a  multiple  of  2n  and  0  <s^l. 


344  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

4.  The  series  of  Ex.  3  are  in  general  absolutely  convergent  if  s>l, 
conditionally  convergent  if  0<s  <  1,  and  oscillatory  if  s  <  0  (finitely  if  s=0 
and  infinitely  if  s<0).     Mention  any  exceptional  cases 

5.  If  1ann~a  is  convergent  or  oscillates  finitely,  then  2ann~l  is  convergent 
when  t>s. 

6.  If  0«  is  a  positive  function  of  n  which  tends  steadily  to  0  as  n-°~- oo , 

and  2(f)n  is  divergent,  then  the  series  2<£n  cosnd,  2<£nsin?i#  are  not  absolutely 

convergent,  except  the  sine-series  when  6  is  a  multiple  of  it.     [For  suppose, 

e.g.,  that  2(pn  |  cos  nd  |  is  convergent.     Since  cos2  nd  g  |  cos  nO  | ,  it  follows  that 

2<£„cos2«0  or 

|2<^),((l  +  cos2?i(9) 

is  convergent.  But  this  is  impossible,  since  20„  is  divergent  and  2(f>n  cos  2nd, 
by  Dirichlet's  Test,  convergent,  unless  6  is  a  multiple  of  tt.  And  in  this 
case  it  is  obvious  that  2$»|cos??#|  is  divergent.  The  reader  should  write 
out  the  corresponding  argument  for  the  sine-series,  noting  where  it  fails 
when  6  is  a  multiple  of  it.] 

190.  Series  of  complex  terms.  So  far  we  have  confined 
ourselves  to  series  all  of  whose  terms  are  real.  We  shall  now 
consider  the  series 

XUn^ZiVn  +  nVn), 

where  vn  and  wn  are  real.  The  consideration  of  such  series  does 
not,  of  course,  introduce  anything  really  novel.  The  series  is 
convergent  if,  and  only  if,  the  series 

are  separately  convergent.  There  is  however  one  class  of  such 
series  so  important  as  to  require  special  treatment.  Accordingly 
we  give  the  following  definition,  which  is  an  obvious  extension  of 
that  of  §184. 

Definition.  The  series  2wn,  where  un  =  vn  +  iivn,  is  said  to  be 
absolutely  convergent  if  the  series  %vn  and  %tvn  are  absolutely 
convergent. 

Theorem.  The  necessary  and  sufficient  condition  for  the  absolute 
convergence  of  %an  is  the  convergence  of  2  |  un  \  or  2  \f{v,?  +  «/na). 

For  if  2?<n  is  absolutely  convergent,  then  both  of  the  series 
2  |  vn  | ,  2  |  wn  |  are  convergent,  and  so  2  {  |  vn  \  +  |  wn  \ )  is  con- 
vergent :  but 

|  un  |  =  s/(vn-  4-  w^)  ^  |  vn  |  +  |  wn  | , 


189-191]  AND   INFINITE   INTEGRALS  345 

and  therefore  2  |  un  \  is  convergent.      On  the  other  hand 

|  Vn  I  =  VOn2  +  wn%     |  wn  |  =  V<>»2  +  <), 

so  that  S  | «»  |  and  2  |  wM  |  are  convergent  whenever  2  |  ww  |  is  con- 
vergent. 

It  is  obvious  that  an  absolutely  convergent  series  is  convergent, 
since  its  real  and  imaginary  parts  converge  separately.  And 
Dirichlet's  Theorem  (§§  169,  185)  may  be  extended  at  once  to 
absolutely  convergent  complex  series  by  applying  it  to  the 
separate  series  2vM  and  2ww. 

The  convergence  of  an  absolutely  convergent  series  may  also  be  deduced 
directly  from  the  general  principle  of  convergence  (cf.  Ex.  lxxvii.  1).  We  leave 
this  as  an  exercise  to  the  reader. 

191.  Power  Series.  One  of  the  most  important  parts  of 
the  theory  of  the  ordinary  functions  which  occur  in  elementary 
analysis  (such  as  the  sine  and  cosine,  and  the  logarithm  and 
exponential,  which  will  be  discussed  in  the  next  chapter)  is  that 
which  is  concerned  with  their  expansion  in  series  of  the  form 
"S,an0Bn.  Such  a  series  is  called  a  power  series  in  x.  We  have 
already  come  across  some  cases  of  expansion  in  series  of  this  kind 
in  connection  with  Taylor's  and  Maclaurin's  series  (§  148).  There, 
however,  we  were  concerned  only  with  a  real  variable  x.  We  shall 
now  consider  a  few  general  properties  of  power  series  in  z,  where 
z  is  a  complex  variable. 

A.  A  'power  series  Xanzn  may  be  convergent  for  all  values  of  z, 
for  a  certain  region  of  values,  or  for  no  values  except  z  =  0. 

It  is  sufficient  to  give  an  example  of  each  possibility. 

1.  The  series  2  — r  is  converqent  for  all  values  of  x.     For  if  ult  =  — ;  then 

K+i|/KH»l/(»+i)--*o 

as  n  -*-  oo ,  whatever  value  z  may  have.  Hence,  by  d'Alembert's  Test,  2  |  un  \  is 
convergent  for  all  values  of  s,  and  the  original  series  is  absolutely  con- 
vergent for  all  values  of  z.  We  shall  see  later  on  that  a  power  series,  when 
convergent,  is  generally  absolutely  convergent. 

2.  The  scries  2n!sn  is  not  convergent  for  any  value  of  z  except  2=0. 
For  if  un=n !  zn  then  |  un  +  1 1/|  un  |  =  («  + 1)  |  z  \ ,  which  tends  to  oo  with  n,  unless 
2  =  0.  Hence  (cf.  Exs.  xxvn.  1,  2,  5)  the  modulus  of  the  ?ith  term  tends  to  oo 
with  n;  and  so  the  series  cannot  converge,  except  when  2  =  0.  It  is  obvious 
that  any  power  series  converges  when  2  =  0. 


346  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 

3.  The  series  2zn  is  always  convergent  when  |s|<l,  and  never  convergent 
when  |  z  |  ^  1.  This  was  proved  in  §  88.  Thus  we  have  an  actual  example  of 
each  of  the  three  possibilities. 

192.  B.  If  a  power  series  2  anzn  is  convergent  for  a  par- 
ticular value  of  z,  say  z1  =  i\  (cos  61  +  i  sin  0^,  then  it  is  absolutely 
convergent  for  all  values  of  z  such  that  \z\  <i\. 

For  lim  anzxn  =  0,  since  2ams1n  is  convergent,  and  therefore  we 
can  certainly  find  a  constant  K  such  that  |  anzyn  \  <  K  for  all 
values  of  n.     But,  if  |  z  \  —  r  <  r1}  we  have 

/r\n  I r\n 

K*B|H«»*iBl(-)    <K{-)' 

and  the  result  follows  at  once  by  comparison  with  the  convergent 
geometrical  series  2  (r/?'1)n. 

In  other  words,  if  the  series  converges  at  P  then  it  converges 
absolutely  at  all  points  nearer  to  the  origin  than  P. 

Example.  Show  that  the  result  is  true  even  if  the  series  oscillates 
finitely  when  z=zv  [If  sn=a0+a1z1  +  ...  +  anz1n  then  we  can  find  K  so  that 
|  sn  |  <  K  for  all  values  of  n.  But  |  an zf1 1  =  |  «„ — *B_1 1  g  |  s„ _i  |  + 1 «» |  <  2 K, 
and  the  argument  can  be  completed  as  before.] 

193.  The  region  of  convergence  of  a  power  series. 
The  circle  of  convergence.  Let  z  —  r  be  any  point  on  the 
positive  real  axis.  If  the  power  series  converges  when  z  =  r  then 
it  converges  absolutely  at  all  points  inside  the  circle  |  z  \  =  r.  In 
particular  it  converges  for  all  real  values  of  z  less  than  r. 

Now  let  us  divide  the  points  r  of  the  positive  real  axis  into 
two  classes,  the  class  at  which  the  series  converges  and  the  class 
at  which  it  does  not.  The  first  class  must  contain  at  least  the 
one  point  z  =  0.  The  second  class,  on  the  other  hand,  need  not 
exist,  as  the  series  may  converge  for  all  values  of  z.  Suppose 
however  that  it  does  exist,  and  that  the  first  class  of  points 
does  include  points  besides  z  =  0.  Then  it  is  clear  that  every 
point  of  the  first  class  lies  to  the  left  of  every  point  of  the  second 
class.  Hence  there  is  a  point,  say  the  point  z  =  R,  which  divides 
the  two  classes,  and  may  itself  belong  to  either  one  or  the  other. 
Then  the  series  is  absolutely  convergent  at  all  points  inside  the 
circle  \z\  =  R. 


191-193]  AND   INFINITE   INTEGRALS  347 

For  let  P  be  any  such  point.  We  can  draw  a  circle,  whose 
centre  is  0  and  whose  radius  is 
less  than  R,  so  as  to  include  P 
inside  it.  Let  this  circle  cut  OA 
in  Q.  Then  the  series  is  con- 
vergent at  Q,  and  therefore,  by 

Theorem  B,   absolutely   conver-      ° 
gent  at  P. 

On  the  other  hand  the  series 
cannot  converge  at  any  point  P'  g'  81" 

outside  the  circle.  For  if  it  converged  at  P'  it  would  converge 
absolutely  at  all  points  nearer  to  0  than  P ;  and  this  is  absurd, 
as  it  does  not  converge  at  any  point  between  A  and  Q'  (Fig.  51). 

So  far  we  have  excepted  the  cases  in  which  the  power  series 
(1)  does  not  converge  at  any  point  on  the  positive  real  axis 
except  z  =  0  or  (2)  converges  at  all  points  on  the  positive  real 
axis.  It  is  clear  that  in  case  (1)  the  power  series  converges 
nowhere  except  when  z  =  0,  and  that  in  case  (2)  it  is  absolutely 
convergent  everywhere.  Thus  we  obtain  the  following  result:  a 
'power  series  either 

(1)  converges  for  z  =  0  and  for  no  other  value  of  z;  or 

(2)  converges  absolutely  for  all  values  of  z ;  or 

(3)  converges  absolutely  for  all  values  of  z  ivithin  a  certain 
circle  of  radius  R,  and  does  not  converge  for  any  value 
of  z  outside  this  circle. 

In  case  (3)  the  circle  is  called  the  circle  of  convergence 
and  its  radius  the  radius  of  convergence  of  the  power  series. 

It  should  be  observed  that  this  general  result  gives  absolutely 
no  information  about  the  behaviour  of  the  series  on  the  circle  of 
convergence.  The  examples  which  follow  show  that  as  a  matter 
of  fact  there  are  very  diverse  possibilities  as  to  this. 

Examples  LXXX.     1.     The  series  l  +  az  +  a?z2+... ,  where  a  >  0,  has  a 

radius  of  convergence  equal  to  \ja.     It  does  not  converge  anywhere  on  its 

circle  of  convei"gence,  diverging  when  z  —  l/a  and  oscillating  finitely  at  all  other 

points  on  the  circle. 

z       z2      z3 
2.     The  series  r^-f  ^  +  ^>  +  ...  has  its  radius  of  convergence  equal  to  1 ; 

it  converges  absolutely  at  all  points  on  its  circle  of  convergence. 


348  THE    CONVERGENCE    OF    INFINITE    SERIES  [VIII 

3.  More  generally,  if  |  an  + 1  |  /|  an  |  -*»  X,  or  |  aH  \1,n  -^  X,  as  n  -*-  cc  ,  then  the 
series  a0  +  a1z  +  a2z2  +  ...  has  1/X  as  its  radius  of  convergence.    In  the  first  case 

Iim  |  an+izn+i  |/j  anzu  |  =  X|  z  | , 

which  is  less  or  greater  than  unity  according  as  \z\  is  less  or  greater  than 
1/X,  so  that  we  can  use  D'Alenihert's  Test  (§  168,  3).  In  the  second  case  we 
can  use  Cauchy's  Test  (§  168,  2)  similarly. 

4.  The  logarithmic  series.     The  series 

is  called  (for  reasons  which  will  appear  later)  the  'logarithmic'  series.  It 
follows  from  Ex.  3  that  its  radius  of  convergence  is  unity. 

"When  z  is  on  the  circle  of  convergence  we  may  write  2  =  cos#+i  sin#, 
and  the  series  assumes  the  form 

cos  6  -  J  cos  2(9  +  J  cos  3(9  -...+{ (sin  6  —  h  sin  2(9  +  J  sin  3(9  -...). 

The  real  and  imaginary  parts  are  both  convergent,  though  not  absolutely 
convergent,  unless  6  is  an  odd  multiple  of  it  (Exs.  lxxix.  3,  4).  If  6  is  an  odd 
multiple  of  it  then  z=  —  \,  and  the  series  assumes  the  form  — 1  — £  —  J  — ..., 
and  so  diverges  to  —  oo  .  Thus  the  logarithmic  series  converges  at  all  points 
of  its  circle  of  convergence  except  the  point  z=  —  1. 

5.  The  binomial  series.     Consider  the  series 

m  (m  —  1)   „     m  (m  —  1)  (w?  —  2)   „ 
l+ms+^2!      ;22+_i ^ V  +  ... 

If  m  is  a  positive  integer  then  the  series  terminates.     In  general 

I  an  +  1  I  _  I  m  —  n  I 

\a„\    "    n  +  1  ' 

so  that  the  radius  of  convergence  is  unity.  We  shall  not  discuss  here  the 
question  of  its  convergence  on  the  circle,  which  is  a  little  more  difficult.* 

194.  Uniqueness  of  a  power  series.  If  Sa„f  is  a  power  series  which 
is  convergent  for  some  values  of  z  at  any  rate  besides  2=0,  and  f(z)  is  its 
sum,  then  it  is  easy  to  see  that/ (z)  can  be  expressed  in  the  form 

O0  +  ax  Z  +  Oo  £2  +  •  •  •  +  («n  +  ea)  zn, 

where  t2-*-0  as  |  z  |-»-0.  For  if  /x  is  any  number  less  than  the  radius  of  con- 
vergence of  the  series,  and  |2|<ii,  then  \aH\fj.n<K,  where  A' is  a  constant 
(cf.  §  192),  and  so 

1/(2) -2a„2"    ^!«„  +  ilk,  +  1l  +  K  +  2||2n  +  2l  +  ... 
I  o  ' 

/I  2  l\m  +  1   /  I  2  I  I  2  i2  \  A"  I 


/x  /      \      it      it2        y    ti"  (M  - 1 2 1) ' 

*  See  Bromwich,  Infinite  Series,  pp.  225  et  seq.  ;  Hobson,  Plane  Trigonometry 
(3rd  edition),  pp.  268  et  seq. 


193-195]  AND   INFINITE   INTEGRALS  349 

where  K  is  a  number  independent  of  z.  It  follows  from  Ex.  lv.  15  that 
if  2a,izn  =  2b)lzn  for  all  values  of  z  whose  modulus  is  less  than  some 
number  /x,  then  an  —  bn  for  all  values  of  n.  This  result  is  capable  of  considerable 
generalisations  into  which  we  cannot  enter  now.  It  shows  that  the  same 
function  f{z)  cannot  be  represented  by  two  different  power  series. 

195.  Multiplication  of  Series.  We  saw  in  §  170  that  if 
%un  and  1vn  are  two  convergent  series  of  positive  terms,  then 
2«n  x  Xvn=  ~ivn,  where 

Wn  =  U0Vn  +  IhVn^  +  . . .  +  UnV0. 

We  can  now  extend  this  result  to  all  cases  in  which  %un  and  2v„ 
are  absolutely  convergent ;  for  our  proof  was  merely  a  simple 
application  of  Dirichlet's  Theorem,  which  we  have  already  ex- 
tended to  all  absolutely  convergent  series. 

Examples  LXXXI.  1.  If  \z\  is  less  than  the  radius  of  convergence 
of  either  of  the  series  2aazH,  "2bnzn,  then  the  product  of  the  two  series  is 
2<v'\  vvhere  cn=a0bn+a1b,l_1  +  ...  +  anbQ. 

2.  If  the  radius  of  convergence  of  2«„«"  is  E,  and  f(z)  is  the  sum  of 
the  series  when  | z |  < E,  and  \z\  is  less  than  either  E  or  unity,  then 
f(z)/(l-z)  =  2snzn,  where  sn=a0  +  a1  +  ...  +  an. 

3.  Prove,  by  squaring  the  series  for  1/(1  -  z),  that  1/(1  —z)2=  1  +2z  +  3z2  + ... 
if|*|<l. 

4.  Prove  similarly  that  1/(1  -z)3=l+3z+6z2+...,  the  general  term 
being  |(n  +  l)(w  +  2)s". 

5.  The  Binomial  Theorem  for  a  negative  integral  exponent.     If 

|s|<l,  and  in  is  a  positive  integer,  then 

(l_z)m  -*+«*+      1>2  +-+  1.2..  .n  Z  +-' 

[Assume  the  truth  of  the  theorem  for  all  indices  up  to  m.   Then,  by  Ex.  2, 

1/(1-  z)m  + 1  =  2  snzn,  where 

m(m  +  l)            m(m  +  l)...(m  +  n-l)      (m  +  l)(m  +  2)...(m+n) 
sn=l+m+      1<2     +...+ ^g— —  = ^ — , 

as  is  easily  proved  by  induction.] 

6.  Prove  by  multiplication  of  series  that  if 

/6M-i+(7)  .+£)*•-.-, 

and  |  z  |  <  1,  then/ (mi,  z)f(m',  z)  =f(m+m',  z).  [This  equation  forms  the  basis  of 
Euler's  proof  of  the  Binomial  Theorem.     The  coefficient  of  zn  in  the  product 


CO 


'iJU-ij+un»-2j+-+u-iHij+C 


350  THE    CONVERGENCE   OF   INFINITE    SERIES  [VIII 

This  is  a  polynomial  in  m  and  m!  :    but  when  m  and  m'  are  positive 

integers  this  polynomial  must  reduce  to  (      ,       J ,  in  virtue  of  the  Binomial 

Theorem  for  a  positive  integral  exponent,  and  if  two  such  polynomials  are 
equal  for  all  positive  integral  values  of  m  and  ml  then  they  must  be  equal 
identically.] 

7.     If  f(z)=l  +  z+  ol+...  then  f(z)f(z')=f(z  +  z').     [For  the  series  for 

f(z)  is  absolutely  convergent  for  all  values  of  z  :  and  it  is  easy  to  see  that  if 
(z  +  z')n 
nl'    "    n\ 


«»-— i  >  vn=—< .  then  wn=  — -y-  .] 


8.  If  C(,)  =  l-^  +  ^-...,      0(,)=.-*  +  *-.„, 

then     'C{z+z')  =  C{z)C(z')-S(z)S{z'),     S(z+*)=S(z)  CW  +  C^S  &), 

and  {<7(,-)}2  +  {>S'(^)}'2  =  l. 

9.  Failure  of  the  Multiplication  Theorem.  That  the  theorem  is  not 
always  true  when  2%  and  2vn  are  not  absolutely  convergent  may  be  seen  by 
considering  the  case  in  which 

(-1)" 

Then 

n  1 

Wn=(-  1)»  j^———. 

But  V{(r  +  1)(m  +  1  -r)]^|(?i  +  2),  and  so  |ww|>  (2n+2)/(»i  +  2),  which  tends 
to  2  ;  so  that  2  wn  is  certainly  not  convergent. 


MISCELLANEOUS  EXAMPLES  ON  CHAPTER  VIII. 

1.  Discuss  the  convergence  of  the  series  2nk{J(n  +  l) -2Khi+J(n-l)\ , 
where  k  is  real.  {Math.  Trip.  1890.) 

2.  Show  that  2?irAfc(n8), 

where  At*„  =  «„  —  «„  + 1 ,     A2('„  =  A  (Am,,), 

and  so  on,  is  convergent  if  and  only  if  k">r+s+l,  except  when  s  is  a  positive 
integer  less  than  k,  when  every  term  of  the  series  is  zero. 

[The  result  of  Ch.  VII,  Misc.  Ex.  11,  shows  that  Ak  (na)  is  in  general  of 
order  ns~k.] 

3.  Show  that 

00  n2  +  9?i  +  5  _  5 

?  (n+l)  (2n  +  3)  (2w  +  5)  (»  +  4)  ~  36' 

(1/a?/;.  2W]p.  1912.) 
[Resolve  the  general  term  into  partial  fractions.] 


AND   INFINITE    INTEGRALS  351 

4.  Show  that,  if  R(n)  is  any  rational  function  of  n,  we  can  determine 
a  polynomial  P  (n)  and  a  constant  A  such  that  2{R  (n)-  P(n)  —  (A/n)}  is 
convergent.  Consider  in  particular  the  cases  in  which  R(n)  is  one  of 
the  functions  l/(an  +  b),     (an2  +  2bii  +  c)/(an'2  +  2^?i  +  y). 

5.  Show  that  the  series 

1         1         1         1 1_ 

1+s      2      2  +  z+3      3  +  z+"' 
is  convergent  provided  only  that  s  is  not  a  negative  integer. 

6.  Investigate  the  convergence  or  divergence  of  the  series 

2  sin-,     2 -sin-,      2(-l)nsin-,     2(1- cos-),     2  (-  l)nn  ( 1  -cos  -  ), 
n  n        n  n         \  nj'  \  nj 

where  a  is  real. 

7.  Discuss  the  convergence  of  the  series 


2      3  nj         n 

where  6  and  a  are  real.  (Math.  Trip.  1899.) 

8.  Prove  that  the  series 

l_I_l4-Ij.l4.1_I_l_l_    1    _|_ 

x        2        3~4~5    '    6        7        8        9        10    '    ■■•» 

in  which  successive  terms  of  the  same  sign  form  groups  of  1,  2,  3,  4,  ...  terms, 
is  convergent ;  but  that  the  corresponding  series  in  which  the  groups  contain 
1,  2,  4,  8,  ...  terms  oscillates  finitely.  (Math.  Trip.  1908.) 

9.  If  «i,  u-i,  M3,  ...  is  a  decreasing  sequence  of  positive  numbers  whose 
limit  is  zero,  then  the  series 

Mi-i(Mi  +  M2)  +  HMl  +  M2+M3)-"-l      Ml-i(«l  +  «3)+i(Ml  +  M3  +  M5)-... 

are  convergent.  [For  if  (mi  +  m2 +...+«„)/«=»„  then  vly  v2,  v3,  ...  is  also  a 
decreasing  sequence  whose  limit  is  zero  (Ch.  IV,  Misc.  Exs.  8,  27).  This 
shows  that  the  first  series  is  convergent ;  the  second  we  leave  to  the  reader. 
In  particular  the  series 

i-*(i+i)+Mi+i+*)-->    W(i+i)+*a+i+*)-- 

ai'e  convergent.] 

10.  If  Uo  +  Ui  +  ic2-\-...  is  a  divergent  series  of  positive  and  decreasing 
terms,  then 

(M0  +  «2  +  ---+«2!»)/Cwl  +  M3  +  ."  +  «&t  +  l)-9\L 

11.  Prove  that  if  a>0  then     lim     2  (p  +  7i)~1~a  =  0. 

12.  Prove  that     lim  a  2  «-1-a=l.     [It  follows  from  §  174  that 


0<l-1-a  +  2-1-a+...  +  («-ir1_a-  I    orx~ada:^l, 
and  it  is  easy  to  deduce  that  2m_1-°  lies  between  1/a  and  (l/a)  +  l.] 


352  THE   CONVERGENCE   OF   INFINITE   SERIES  [VIII 


13.     Find  the  sum  of  the  series  2  un,  where 

i 

xn_x-n-l  1        / 


l"n~(xn  +  x-n)(xn  +  ^  +  x-n-1)     x-l\xn  +  x~n     xn  +  1+x~n 

for  all  real  values  of  x  for  which  the  series  is  convergent.     {Math.  Trip.  1901.) 

[If  |  x  |  is  not  equal  to  unity  then  the  series  has  the  sum  #/{(#  —  1)  (a;2+l)}. 
If  x  =  l  then  «„  =  0  and  the  sum  is  0.  If  x=-l  then  Mn=§(-l)n+1  and 
the  series  oscillates  finitely.] 

14.  Find  the  sums  of  the  series 

2         _2z2_        4a4  s  a2  z4 

(in  which  all  the  indices  are  powers  of  2),  whenever  they  are  convergent. 
[The  first  series  converges  only  if  |  z  |  <  1,  its  sum  then  being  zj(\  —  z) ;  the 
second  series  converges  to  zj{\  -z)  if  |  z  \  <  1  and  to  1/(1  —z)  if  |  z  |  >  1.] 

15.  If  |«ti|^1  for  all  values  of  n  then  the  equation 

Q=\+alz  +  a2zi  +  ... 

cannot  have  a  root  whose  modulus  is  less  than  •£,  and  the  only  case  in  which 
it  can  have  a  root  whose  modulus  is  equal  to  ^  is  that  in  which  a„=  -  Cis(n0), 
when  2=i  Cis  ( -  6)  is  a  root. 

16.  Recurring  Series.  A  power  series  2aKsn  is  said  to  be  a  recurring 
series  if  its  coefficients  satisfy  a  relation  of  the  type 

an+Pian-i+P2an-2+-~+Pk«n-k=Q    (1), 

where  n^h  and  px,  p2,  ...,  pu  are  independent  of  n.  Any  recurring  series  is 
the  expansion  of  a  rational  function  of  z.  To  prove  this  we  observe  in  the 
first  place  that  the  series  is  certainly  convergent  for  values  of  z  whose  modulus 
is  sufficiently  small.     For  let  O  be  the  greater  of  the  two  numbers 

1)     \lh\  +  \P2\  +  —  +  \pk\- 
Then   it  follows  from  the  equation   (1)   that   \an\^Gan,  where   a„  is  the 
modulus  of  the  numerically  greatest  of  the  preceding  coefficients ;  and  from 
this  that  |  an  \  <  KGn,  where  K  is  independent  of  n.     Thus  the  recurring  series 
is  certainly  convergent  for  values  of  z  whose  modulus  is  less  t"han  l/(r. 

But  if  we  multiply  the  series  f(z)  =  2anzn  by  ptz,  p2z2,  ...pk^,  and  add 
the  results,  we  obtain  a  new  series  in  which  all  the  coefficients  after  the 
(k-  l)th  vanish  in  virtue  of  the  relation  (1),  so  that 

(l+p1z+p2z°~  +  ...+pkzk)f(z)  =  P0  +  P1z+...+Pk_1zk-\ 
where  P0,  Plt ..., Pk-i are  constants.     The  polynomial  1  +piz+p2zz+...+pkzk 
is  called  the  scale  of  relation  of  the  series. 

Conversely,  it  follows  from  the  known  results  as  to  the  expression  of  any 
rational  function  as  the  sum  of  a  polynomial  and  certain  partial  fractions  of 
the  type  Aj(z  —  a)p,  and  from  the  Binomial  Theorem  for  a  negative  integral 


AND   INFINITE   INTEGRALS  353 

exponent,  that  any  rational  function  whose  denominator  is  not  divisible  by  z 
can  be  expanded  in  a  power  series  convergent  for  values  of  z  Whose  modulus  is 
sufficiently  small,  in  fact  if  \z  |  <  p,  where  p  is  the  least  of  the  moduli  of  the  roots 
of  the  denominator  (cf.  Ch.  IV,  Misc.  Exs.  18  et  seq.).  And  it  is  easy  to  see, 
by  reversing  the  argument  above,  that  the  series  is  a  recurring  series.  Thus 
the  necessary  and  sufficient  condition  that  a  power  series  shoidd  be  a  recurring 
series  is  that  it  shoidd  be  the  expansion  of  such  a  rational  function  of  z. 

17.  Solution  of  Difference-Equations.  A  relation  of  the  type  of  (1) 
in  Ex.  16  is  called  a  linear  difference-equation  in  an  with  constant  coefficients. 
Such  equations  may  be  solved  by  a  method  which  will  be  sufficiently  ex- 
plained by  an  example.     Suppose  that  the  equation  is 

an-«n-i-8a»-2  +  12aB_3=0. 

Consider  the  recurring  power  series  2anzn.  We  find,  as  in  Ex.  16,  that  its 
sum  is 

ao  +  (ai-a0)z  +  (a2-ai-8a0)z2  __     Ax  A2  B 

l-z-8^  +  1223  ~  1  - 2z  +  (1  - 2z)2      T+3z  ' 

where  A\t  A2,  and  B  are  numbers  easily  expressible  in  terms  of  a0,  cq,  and  a2. 
Expanding  each  fraction  separately  we  see  that  the  coefficient  of  zn  is 

an=2«{A1  +  (n+l)A2}  +  (-S)»B. 

The  values  of  Ax,  A2,  B  depend  upon  the  first  three  coefficients  a0,  au  a2, 
which  may  of  course  be  chosen  arbitrarily. 

18.  The  solution  of  the  difference-equation  uH  —  2  cosd  un_1  +  un_2  =  Q  is 
un  —  A  cos?j#  +  .Ssin  nd,  where  A  and  B  are  arbitrary  constants. 

19.  If  un  is  a  polynomial  in  n  of  degree  I;  then  2unzn  is  a  recurring 
series  whose  scale  of  relation  is  (1  -z)k  +  1.  (Math.  Trip.  1904.) 

20.  Expand  9/{(z  —  1)  (s  +  2)2}  in  ascending  powers  of  z. 

(Math.  Trip.  1913.) 

21.  Prove  that  if/(%)  is  the  coefficient  of  zn  in  the  expansion  of  2/(1  +  z  +  z2) 
in  powers  of  z,  then 

(1)    /(»)+/(»-l)+/(n-2)=0,  (2)    /(m)  =  (a,3»-a,32'0/(co3-co32), 

where  a>3  is  a  complex  cube  root  of  unity.  Deduce  that  f(n)  is  equal  to  0 
or  1  or  - 1  according  as  n  is  of  the  form  3£  or  3£  +  l  or  3£  +  2,  and  verify 
this  by  means  of  the  identity  zj(\  +  z  +  z2)=z  (l-z)/(l  -z3). 

22.  A  player  tossing  a  coin  is  to  score  one  point  for  every  head  he  turns 
up  and  two  for  every  tail,  and  is  to  play  on  until  his  score  reaches  or  passes 
a  total  n.     Show  that  his  chance  of  making  exactly  the  total  n  is  J  {2  -f-  (  —  I )»}. 

(Math.  Trip.  1898.) 
[If  pn  is  the  probability  thenpn=J  (Pn-i+Pn-z)  •  also  p0=l,  Pi  =  i-] 
H.  23 


/: 


354  THE    CONVERGENCE    OF    INFINITE    SERIES  [VIII 

23.  Prove  that 

1  1  1         /»\  J_  _  /n\  1! 

a+l+a+2+"       a+n      \l/a  +  l      W  («  +  l)  («  +  2)        * 
if  n  is  a  positive  integer  and  a  is  not  one  of  the  numbers  -1,  -2,  ...,  —  n. 
[This  follows  from  splitting  up  each  term  on  the  right-hand  side  into  partial 
fractions.     When  a  >  - 1,  the  result  may  be  deduced  very  simply  from  the 
equation 

1   ™n  f  1  (Jiy 

^-d.V=         (1  _,;)«{!  _(l-^)«}^ 
1  —  X  Jo  X 

by  expanding  (1  - xn)/(l  -x)  and  l-(l-.r)'1  in  powers  of  x  and  integrating 
each  term  separately.  The  result,  being  merely  an  algebraical  identity,  must 
be  true  for  all  values  of  a  save  —  1,  —2,  ...,  —  n.] 

24.  Prove  by  multiplication  of  series  that 

oo  zn    cc  l_\\n-lzn         cc    /  i        l  Jv    2n 

2  —,2{ '—r-  =2    l+H  +  o  +  --  ■  +  -)—,  • 

o  n  !  i      ?i .  n  !  t  \       2      3  »/ » l 

[The  coefficient  of  zn  will  be  found  to  be 

.M©-i©+S©-4 

Now  use  Ex.  23,  taking  a  =  0.] 

25.  If  An-*-A  and  Bn^~B  as  n-s-oo ,  then 

(^£n + J  2JBn  _  x  + . . .  +  J  ^j)/^  .4  £. 
[Let  An=A  +  en.     Then  the  expression  given  is  equal  to 

AB1  +  B2  +  ...  +  Bn     e1Bn+e2Bn_1  +  .  ..+€%Bl 

n  n 

The  first  term  tends  to  AB  (Ch.  IV,  Misc.  Ex.  27).  The  modulus  of 
the  second  is  less  than  /3  { |  ex  \  +  |  e2 1  + . .  •  + 1  *n  \  }/n,  where  ft  is  any  number 
greater  than  the  greatest  value  of  |  Bv  | :  and  this  expression  tends  to  zero.] 

26.  Prove  that  if  cn=a1bn  +  a2 &„_i  +  ...+«n^i  and 

An  =  a1  +  a2  +  ...+an,     Bn  =  b1  +  b2+  ...  +  bn,     Cn  =  c1  +  c2  +  ...+c„, 
then 

Cn  =  a1BH+a2Bn_1  +  ...  +  anB1  =  b1An  +  b2An_1  +  ...  +  bnA1 

and  C1  +  C2  +  ...  +  Cn=AlBn  +  A2Bn_1  +  ...  +  AnB1. 

Hence  prove  that  if  the  series  1an,  2bn  are  convergent  and  have  the  sums 
A,  B,  so  that  An-*-A,  Bn-*-B,  then 

(C1  +  C2  +  ...  +  Cn)/n+AB. 
Deduce  that  if  2c„  is  convergent  then  its  sum  is  AB.  This  result  is  known  as 
Abel's  Theorem  on  the  multiplication  of  Series.  We  have  already  seen 
that  we  can  multiply  the  series  2a„,  s6n  in  this  way  if  both  series  are 
absolutely  convergent  :  Abel's  Theorem  shows  that  we  can  do  so  even  if 
one  or  both  are  not  absolutely  convergent,  provided  only  that  the  product  series 
is  convergent. 


AND   INFINITE    INTEGRALS  355 

27.  Prove  that 

l(i-i+l-...)'-4-l(i+i)+i(i+Hi)-.»i 

[Use  Ex.  9  to  establish  the  convergence  of  the  series.] 

28.  For  what  values  of  m  and  n  is  the  integral  /    sin"'  x  (1  -  cos  x)n  dx 

.   Jo 
convergent  ?     [If  m  + 1  and  m  +  2n  +  l  are  positive.] 

29.  Prove  that  if  a  >  1  then 

f1  dx  it 

J  _!  (a-*)v'(I^) =  ^(a2-l)  * 

30.  Establish  the  formulae 

j"  Fy{x*  +  \)-x)dx  =  hj^  (l  +  j^F{y)dy. 
In  particular,  prove  that  if  w  >  1  then 

f  °°  ,  „  /* r  =  f  "  V(*2  +  1)  -  *}n  dx  =  -^  . 

Jo   {N/(^2  +  l)+^}n     Jo  X  '  v?~\ 

[In  this  and  the  succeeding  examples  it  is  of  course  supposed  that  the 
arbitrary  functions  which  occur  are  such  that  the  integrals  considered  have  a 
meaning  in  accordance  with  the  definitions  of  §§  177  et  seq.] 

31.  Show  that  if  2y  =  ax—(bfx),  where  a  and  b  are  positive,  then  y  in- 
creases steadily  from  —  oo  to  oc  as  x  increases  from  0  to  oc  .    Hence  show  that 

/."4  (-  ¥i)}  <b>-\r_jw+<*»  I'+vT^h 

2 


f{J{y*  +  ab))dy. 
a  J  o 

32.  Show  that  if  2y  =  ax  +  (b/x),  where  a  and  b  are  positive,  then  two 
values  of  x  correspond  to  any  value  of  y  greater  than  s/(ab).  Denoting  the 
greater  of  these  by  x\  and  the  less  by  x2,  show  that,  as  y  increases  from 
J(ab)  towards  oo,  xx  increases  from  Jibja)  towards  oo,  and  a?2  decreases 
from  \/(b/a)  to  0.     Hence  show  that 

f  °°        f(y)  dxx  =  -  {^        f(y)[ -irv-^jx  + 1\  dy, 
J  V(6/«)  a  J  y/(ab)  W(f ' ab)         J 

and  that 

("/^(W^l^-  I"*       -/{^rdy  =  l  (™ fy{z>  +  ab))dz. 
Jo'  V\         -WJ  a]^ab)sl{y'i-ab)    '      a  J  0 

23—2 


356  INFINITE   SERIES    AND    INTEGRALS  [VIII 

33.     Prove  the  formula 

fn  .,       ,              iv       dx           f17 ','              ,      dx 
I    /  (sec  * x + tan  \  x)  -rz-. r  =  /    /  (cosec  x)  -r—. . , 


34.     If  a  and  b  are  positive,  then 

f  °°  dx  it  f  °°  x2  dx 


J  0  (x2  +  a2)  (x2  +  b2)      2ab  (a  +  b) '     J  0  (x2  +  a2)  (x2  +  b2)      2  (a  +  6) ' 

Deduce  that  if  a,  /3,  and  y  are  positive,  and  /32  ^  ay,  then 

fx  dx  ir  f°°        x2dx  it 

J  0  axi  +  2l3x*  +  y  =  2  V(2y4) '      J  0  a>  +  2/3.V2  +  y  =  2,J{2aA) ' 

where  A  =  /3  +  J  {ay).  Also  deduce  the  last  result  from  Ex.  31,  by  putting 
f(y)  =  ll(c2+y2).  The  last  two  results  remain  true  when  /32<ay,  but  their 
proof  is  then  not  quite  so  simple. 

35.     Prove  that  if  b  is  positive  then 

/"  °°  x2dx  n        f  °°  .r4c/.r  7r 


Jo   (^2-a:!)2  +  62^2     26'    J0   {(^2-a2)2  +  62.^2}2     463' 

36.     Extend   Schwarz's  inequality   (Ch.    VII,   Misc.    Ex.  42)  to  infinite 
integrals  of  the  first  and  second  kinds. 


37.  Prove  that  if  <f>(x)  is  the  function  considered  at  the  end  of  §  178 
then 

Jc>      W  o^n+iy 

38.  Prove  that 

Establish  similar  results  in  which  the  limits  of  integration  are  0  and  1. 

(Math.  Trip.  1913.) 


CHAPTER  IX 

THE  LOGARITHMIC  AND  EXPONENTIAL  FUNCTIONS 
OF  A  REAL  VARIABLE 

196.  The  number  of  essentially  different  types  of  functions 
with  which  we  have  been  concerned  in  the  foregoing  chapters 
is  not  very  large.  Among  those  which  have  occurred  the  most 
important  for  ordinary  purposes  are  polynomials,  rational  functions, 
algebraical  functions,  explicit  or  implicit,  and  trigonometrical 
functions,  direct  or  inverse. 

We  are  however  far  from  having  exhausted  the  list  of  functions 
which  are  important  in  mathematics.  The  gradual  expansion  of 
the  range  of  mathematical  knowledge  has  been  accompanied  by 
the  introduction  into  analysis  of  one  new  class  of  function  after 
another.  These  new  functions  have  generally  been  introduced 
because  it  appeared  that  some  problem  which  was  occupying  the 
attention  of  mathematicians  was  incapable  of  solution  by  means  of 
the  functions  already  known.  The  process  may  fairly  be  compared 
with  that  by  which  the  irrational  and  complex  numbers  were  first 
introduced,  when  it  was  found  that  certain  algebraical  equations 
could  not  be  solved  by  means  of  the  numbers  already  recognised. 
One  of  the  most  fruitful  sources  of  new  functions  has  been  the 
problem  of  integration.  Attempts  have  been  made  to  integrate 
some  function  fix)  in  terms  of  functions  already  known.  These 
attempts  have  failed ;  and  after  a  certain  number  of  failures  it 
has  begun  to  appear  probable  that  the  problem  is  insoluble. 
Sometimes  it  has  been  proved  that  this  is  so ;  but  as  a  rule  such 
a  strict  proof  has  not  been  forthcoming  until  later  on.  Generally 
it  has  happened  that  mathematicians  have  taken  the  impossibility 
for  granted  as  soon  as  they  have  become  reasonably  convinced 
of  it,  and  have  introduced  a  new   function  F  (x)  defined  by  its 


358  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

possessing  the  required  property,  viz.  that  F'  (x)  =/(#).    Starting 

from    this   definition,   they  have    investigated    the   properties   of 

F(x);and  it  has  then  appeared  that  F  (x)  has  properties  which 

no  finite  combination  of  the  functions   previously  known  could 

possibly  have ;  and  thus  the  correctness  of  the  assumption  that 

the    original    problem    could    not    possibly   be    solved  has  been 

established.     One  such   case    occurred   in    the   preceding   pages, 

when  in  Ch.  VI  we  defined  the  function  log  a;  by  means  of  the 

equation 

,  fdx 

log«-J-. 

Let  us  consider  what  grounds  we  have  for  supposing  log.r  to  be  a  really 
new  function.  We  have  seen  already  (Ex.  xlii.  4)  that  it  cannot  be  a  rational 
function,  since  the  derivative  of  a  rational  function  is  a  rational  function 
whose  denominator  contains  only  repeated  factors.  The  question  whether  it 
can  be  an  algebraical  or  trigonometrical  function  is  more  difficult.  But  it  is 
very  easy  to  become  convinced  by  a  few  experiments  that  differentiation  will 
never  get  rid  of  algebraical  irrationalities.  For  example,  the  result  of 
differentiating  v/(l  +x)  any  number  of  times  is  always  the  product  of  ,J(l+x) 
by  a  rational  function,  and  so  generally.  The  reader  should  test  the 
correctness  of  the  statement  by  experimenting  with  a  number  of  examples. 
Similarly,  if  we  differentiate  a  function  which  involves  sin.*?  or  cos.r,  one 
or  other  of  these  functions  persists  in  the  result. 

"We  have,  therefore,  not  indeed  a  strict  proof  that  logx  is  anew  function — 
that  we  do  not  profess  to  give* — but  a  reasonable  presumption  that  it  is. 
We  shall  therefore  treat  it  as  such,  and  we  shall  find  on  examination  that  its 
properties  are  quite  unlike  those  of  any  function  which  we  have  as  yet 
encountered. 

197.     Definition  of  log  x.    We  define  log  x,  the  logarithm  of  x, 
by  the  equation 

We  must  suppose  that  x  is  positive,  since  (Ex.  lxxvi.  2)  the 
integral  has  no  meaning  if  the  range  of  integration  includes 
the  point  x  =  0.  We  might  have  chosen  a  lower  limit  other 
than  1 ;  but  1  proves  to  be  the  most  convenient.  With  this 
definition  log  1=0. 

We  shall  now  consider  how  log  x  behaves  as  x  varies  from  0 
towards  go  .     It  follows  at  once  from  the  definition  that  log  #  is  a 

*  For  such  a  proof  see  the  author's  tract  quoted  on  p.  236. 


196,  197] 


OF    A   REAL   VARIABLE 


359 


continuous  function  of  x  which  increases  steadily  with  x  and  has 
a  derivative 

Dx  log  x  =  1/x; 

and  it  follows  from  §  175  that  log  x  tends  to  oo  as  x  ->-  oo  . 

If  a;  is  positive  but  less  than  1,  then  logx  is  negative.     For 


i  [*dt  [ldt     a 

l0gX  =  ]1l  =  -]xJ<°- 


Moreover,  if  we  make  the  substitution  t=l/u  in  the  integral,  we 
obtain 

.  f'dt        [V*du        ,     ,_.. 

Iog"=i1T  =  -j1    — -log(V*)- 

Thus  log  x  tends  steadily  to  —  oo  as  x  decreases  from  1  to  0. 

The  general  form  of  the  graph  of  the  logarithmic  function  is 
shown  in  Fig.  52.     Since  the  derivative  of  log  x  is  ljx,  the  slope  of 

Y 


Fig.  52. 
the  curve  is  very  gentle   when  x  is  very  large,  and  very  steep 
Avhen  x  is  very  small. 

Examples  LXXXII.     1.     Prove  from  the  definition  that  if  u  >  0  then 

ttl(l+u)  <  log  (1+u)  <  u. 

/u  clf 
-— ,  and  the  subject  of  integration  lies  between  1  and 
o  1  +  ' 

2.  Prove  that  log(l  +  2t)  lies  between  u  —  —  and  u—  — when  a  is 

/u  tdt 
.] 

3.  If  0  <  u  <  1  then  u  <  -  log  (1  -  u)  <  «/(l  -  u). 

4.  Prove  that 

..      log„r     ..      log  (1  +  0     , 
lira  — £-=  lim  -?  v        ;  =  l. 

[Use  Ex.  1.] 


SCO  THE   LOGARITHMIC    AND   EXPONENTIAL   FUNCTIONS  [iX 

198.     The  functional  equation    satisfied  by  logic.      The 
function  \ogx  satisfies  the  functional  equation 

/(*>/)= /(*) +f(y) (i). 

For,  making  the  substitution  t  =  yu,  we  see  that 

_  f*»  dt  _  fx  du  _  fxdu  _  Plvdu 

Jl        t        Jljy   u        Ji     u        J  i       u 

=  log  x  -  log  (1/y)  =  log  x  +  log  y, 
which  proves  the  theorem. 


Examples  LXXXIII.  1.  It  can  be  shown  that  there  is  no  solution  of 
the  equation  (1)  which  possesses  a  differential  coefficient  and  is  fundamentally 
distinct  from  log  x.  For  when  we  differentiate  the  functional  equation,  first 
with  respect  to  x  and  then  with  respect  to  y,  we  obtain  the  two  equations 

yf '  (xy)  =/'  (*)>   xf  (xy)  =/'  (y) ; 

and  so,  eliminating  /'  (xy\  xf'i(x)=yf  (y).  But  if  this  is  true  for  every  pair 
of  values  of  x  and  y,  then  we  must  have  xf  (x)  =  C,  or  f  (x)  =  Cjx,  where  C 
is  a  constant.     Hence 

/(.,;)=   /  ^dx+C'  =  C\0gX+C, 

and  it  is  easy  to  see  that  C"  =  0.  Thus  there  is  no  solution  fundamentally 
distinct  from  log.r,  except  the  trivial  solution  f(x)  =  0,  obtained  by  taking 

2.     Show  in  the  same  way  that  there  is  no  solution  of  the  equation 

/w+/w-/(;_+£) 

which  possesses  a  differential  coefficient  and  is  fundamentally  distinct  from 
arc  tan  x. 

199.    The  manner  in  which  log  x  tends  to  infinity  with  x. 

It  will  be  remembered  that  in  Ex.  xxxvi.  6  we  defined  certain 
different  ways  in  which  a  function  of  x  may  tend  to  infinity  with  x, 
distinguishing  between  functions  which,  when  x  is  large,  are  of 
the  first,  second,  third, . . .  orders  of  greatness.  A  function  f  (x) 
was  said  to  be  of  the  A'th  order  of  greatness  when  f  (x)/xk  tends  to 
a  limit  different  from  zero  as  x  tends  to  infinity. 

It  is  easy  to  define  a  whole  series  of  functions  which  tend  to 
infinity  with  x,  but  whose  order  of  greatness  is  smaller  than  the  first. 
Thus  >Jx,  tyx,  \/x,  . . .  are  such  functions.  We  may  say  generally 
that  xa,  where  a  is  any  positive  rational  number,  is  of  the  ath 
order  of  greatness  when  x  is  large.     We  may  suppose  a  as  small 


198-201]  OF    A    REAL   VARIABLE  3G1 

as  we  please,  e.g.  less  than  "0000001.  And  it  might  be  thought 
that  by  giving  a  all  possible  values  we  should  exhaust  the 
possible  '  orders  of  infinity '  of  f  (x).  At  any  rate  it  might  be 
supposed  that  if  f  (x)  tends  to  infinity  with  x,  however  slowly,  we 
could  always  find  a  value  of  a.  so  small  that  xa  would  tend  to 
infinity  more  slowly  still ;  and,  conversely,  that  if  f(x)  tends  to 
infinity  with  x,  however  rapidly,  we  could  always  find  a  value 
of  a  so  great  that  xa  would  tend  to  infinity  more  rapidly  still. 

Perhaps  the  most  interesting  feature  of  the  function  log  x  is  its 
behaviour  as  x  tends  to  infinity.  It  shows  that  the  presupposition 
stated  above,  which  seems  so  natural,  is  unfounded.  -The  logarithm 
of  x  tends  to  infinity  with  x,  but  more  slowly  than  any  positive  'power 
of  x,   integral    or  fractional.      In    other   words    loga?-*-  oo    but 

log^0 
of- 

for  all  positive  values  of  a.  This  fact  is  sometimes  expressed 
loosely  by  saying  that  '  the  order  of  infinity  of  log  x  is  infinitely 
small ';  but  the  reader  will  hardly  require  at  this  stage  to  be  warned 
against  such  modes  of  expression. 

200.  Proof  that  (logx)/xa-+0  as  a?-*-oo.  Let  yS  be  any 
positive  number.     Then  \Jt<  l/t1^  when  t  >  1,  and  so 

fxdt      fx  dt 

l08W=]1i<!1t^' 

or  log  x  <  (a>a  —  l)//3  <  oP/ft, 

when  x  >  1.     Now  if  a  is  any  positive  number  we  can  choose  a 

smaller  positive  value  of  (3.     And  then 

0  <  (log  x)/xa  <  #3_a//3        (x  >  1). 
But,  since  a  >  /3,  #'8-a//3^0  as  x  ^  oo  ,  and  therefore 

(log®)/ oc* -^0. 

201.  The  behaviour  of  log  x  as  x  -»-  +  0.     Since 

(log x)fxa  =  -ya  logy 
if  x  =  1/y,  it  follows  from  the  theorem  proved  above  that 
lim  ya  log  y  =  —  lim    (log  x)/xa  =  0. 

Thus  logx  tends  to  —  oo  and  log(l/x)  =  —  logx  to  co  as  x  tends 
to  zero  by  positive  values,  but  log  (l/x)  tends  to  oo  more  slowly 
than  any  positive  power  of  1/x,  integral  or  fractional. 


:"     :  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

202.     Scales  of  inrrnity.     The  logarithmic  scale.    Let  us  coumdar 
:       -    -       -        -      ::.:.:::  :.? 

•  -  sses   -  that,  if  / (x)  and  d»(*]  are  any  twe 
functions  contained  in  it,  then/(x)  and  <f>  (*]  both  tend  to  x  as  x-»-x ,  while 
-'  -   6(x)  tends  to  0  or  to  ao  according     -  oeurs  to  the  right  or  the 

r        in  the  ~:r:es.     We  can  now  ;  this  aeries  by  ib        -ertion 

of  new  terms  to  the  right  of  all  those  already  written  i.^wn.     We  c-jin  begin 
with  log*,  which  tends  to  infinity  more  .in  any  of  the  old  terms. 

Then  N   I  g      tends  tc   ddhr   •  -an  log*,  jTflog*]  than  N'(log. 

so  on.    Thus  we  obtain  a  s 

;-••••   -  -  -•    %'--•?   ■   <  •  r  •• ;  -  - 

formed  of  two  aim]  '"  inffa  rxanged  :ne  after  the  other.     But  this 

is  not  aH  sides  the  function  log  log*,  the  logarithm  of  log  x.     Since 

(kjgx)  xa-*-0,  for  all  positive  values  of  a,  it  follows  on  putting  x=logy  that 

(loglogy    togj  -=  log*    ■•  =  —  (). 

Thus  leg  logy  tends  to  x  with  y,  but  more  slowly  than  any  power  or 
Hence  we  may  continue  c  :"  jrm 

,    N   .    log    ,  I  _     J  log  . '  .  .  N  :  > :  ■         N  :  T-  :  . 

and  it  will  by  now  be  obTioos  that  by  introducing  the  functions  log  log  log  *, 
log  log  log  log  *,  ...  we  can  prolong  the  series  to  any  :  like.     By 

putti:.  -     =  I      ~e  obtain  a  similar  scale  of  infinity  for  functions  of  y  which 
tend:  -rvalues.* 

Examples  LSX5IV.    1.   E~:-r^n  any  two  terms /(x  .  F  %  of  the  series 
n.  insert  a  new  I  ten  Is  to  x  more  slowly  than 

•  :  .   :■:■  .  F  bet  ^  N   "  we  could  insert 

N       .  14       -      ve  could  insert  log*      -.    A:.  1,  generally, 

:        =  s  F  '  ited.] 

Find  a  1  Ib  1     x  more  slowly  than  srx,  but  more 

rapidly  than  *",  where  a  is  any  rational  number  less  than  11     [N 
~  :;1  .:";:_.-       .     rN       '   .      ".  """.. --/-:-  :  >  iny  positive rational  mnnber.] 

3.     Find  a  function  wn  to  ao  m  re  -1  "  .;■"  than  N    .  '   .-  mora 

rapid.  ..  a      *,  where  a  is  any  rational  number.      [The  fin 

3,  function.    It  will  be  gathered  from  these  example  - 
{RCompZeteness  is  an  inherent  characteristic  of  the  logarithmic  scale  of  infinity.] 

■L     How  does  the  function 

=  ■  logx^aogiog  -     -  iog*)^aogiog*)^> 

behave  as  *  ten  -i  "If  a±$  then  the  behaviour  of 

f      =    -i  tog     '":       -     -     '  ~r 

*  For  fuller  information  as  to  '  scales  of  ^infinity  '  see  the  author 
of  Infr.    7         .nib.  Math.  1  12. 


202.  203]  A   KEAL  VARIABLE  Ho 

is  i:r_;~i-oi  -v  :i^:    .i  ..'-'-      li  -.=  z  ':.--.  zl-.  -._•  "  -:     :'       ;'_-   yy-:-  '--i 
the  behaviour  of  /  x;  is  dominated  by  that  of  (logx  *  _y,  Trnkss  a —3".  wmm 

■  ■    ii  -  --.•-.  _ ■      .       -       -  T..  _-    "         -~  r.  : 

-=  ~        ^  -       r  a  =  5.  a  =  z  ^  z        z.  i  -         -~  '.  -J.  z  <  :  =:         "  : 

;=  r.   i  =  :      i     <  :     | 

5.    .  s  "'.'gj  "•.•:*:•;  - 

- " ."  -   ?-'-■?       -------      -.:•:   riir-g  v.   :!•=  riT-ii."   ~.\z  -_.::.  :ir;    --i; 


kgkgx/(xlogxV,     (kgx)/x,    xlogkgxV(x*+  ,      -; 

_         :_t  :  ^  r. 

7.     Arrange 

.  •  1:  g  1 :  g   1  ~  .     N       I:-g   I        .     %    .r  -iz.  -  I  :-z  1  .r   .       1  -  :-:>=   ■   L:>g   1 /x 
.:.:-:.: ...y  :•:  -._-  z-.y^rr — r.'z.  —  :_:-i.  -±tJ  :~i.i  :•:  ziz-:   -.-      -■ — 

-       -      -  -      - 

2>xloglogx=:  Z>.logloglogx=l -rlogxlogk^ 

-    ■  -    - 

203.       The  number  £ 
usually  der  -    :  f  immense  importance  in  higher 

mathematics.       I:    is.  like  —     :.r     :    :Le    :n.  _  .l_t:  :   .    ;._--       - 
. :  :.: 

We   define  e  as  the  number  whose  logarithm 

■wor  -       dned  bv  the  equs  - 

Since  logrj*  is  an  ir.  .  - 

rass  once  through  the  value  1.     Hence  oar 
lefinition  does  in  fact  deni  edniie  r 

N ."""._       =  -  _"     -     _ 

-    -=--_  _      ='-_  -_       = 

whri-      is    :  satire  im    z  Hence 

-      -    :       =         -     i    :    = 


364  THE    LOGARITHMIC    AND    EXPONENTIAL    FUNCTIONS  [iX 

Again,  if  p  and  q  are  any  positive  integers,  and  ep/q  denotes  the 
positive  qth.  root  of  ep,  we  have 

p  =  log  e?  =  log  (<?2>/*)9  =  q  log  e*/«, 

so  that  \ogep,q  =  p/q.  Thus,  if  3/  has  any  positive  rational  value, 
and  ey  denotes  the  positive  ytih  power  of  e,  we  have 

\ogev  =  y   (1), 

and  log  e~y  =  —  log  ey  =  —  y.     Hence  the  equation  (1)  is  true  for 

all  rational  values  of  y,  positive  or  negative.     In  other  words  the 

equations 

y  =  \ogx,     x=ey  (2) 

are  consequences  of  one  another  so  long  as  y  is  rational  and  ey 
has  its  positive  value.  At  present  we  have  not  given  any  definition 
of  a  power  such  as  ey  in  which  the  index  is  irrational,  and  the 
function  ey  is  defined  for  rational  values  of  y  only. 

Example.     Prove  that  2  <  e  <  3.     [In  the  first  place  it  is  evident  that 

'2  eft 


/ 


t     <lj 

1      I 


and  so  2  <  e.     Also 

[3dt_f2dt      [3dt_(1    du        f1   du  P    du 

)i  T~  Jl  7  +  J,   J~  Jo  2-uJo  2  +  u~    Jo  ^-Ti 
so  that  e  <  3.] 


5>1, 


204.  The  exponential  function.  We  now  define  the  ex- 
ponential function  ey  for  all  real  values  of  y  as  the  inverse  of 
the  logarithmic  function.     In  other  words  we  write 

x  —  ey 
if  y  =  log  x. 

We  saw  that,  as  x  varies  from  0  towards  00 ,  y  increases 
steadily}  in  the  stricter  sense,  from  —  00  towards  00 .  Thus  to 
one  value  of  x  corresponds  one  value  of  y,  and  conversely.  Also  y 
is  a  continuous  function  of  x,  and  it  follows  from  §  109  that  x  is 
likewise  a  continuous  function  of  y. 

It  is  easy  to  give  a  direct  proof  of  the  continuity  of  the  exponential  function. 
For  if  x=ev  and  z  +  £  =  ey+r>  then 

■x+Sdt, 
t 

Thus  hi  is  greater  than  £/(#  +  £)  if  £>0,  and  than  |£|/#  if  £<0;  and  if  rj  is 
very  small  £  must  also  be  very  small. 


203-205] 


OF    A    REAL    VARIABLE 


365 


Thus  ey  is  a  positive  and  continuous  function  of  y  which 
increases  steadily  from  0  towards  oo  as  y  increases  from  —  oo 
towards  oo  .  Moreover  ey  is  the  positive  yih.  power  of  the  number 
e,  in  accordance  with  the  elementary  definitions,  whenever  y  is 
a  rational  number.  In  particular  ey  =  1  when  y  =  0.  The  general 
form  of  the  graph  of  ey  is  as  shown  in  Fig.  53. 


0  X 

Fig.  53. 

205.  The  principal  properties  of  the  exponential 
function.  (1)  If  x  =  ey,  so  that  y  =  log  x,  then  dy/dx  =  l/x 
and 

dx 


dy 


=  x=  ey. 


Thus  the  derivative  of  the  exponential  function   is   equal   to  the 
function  itself     More  generally,  if  x  =  eay  then  dxjdy  =  aeay. 

(2)     The  exponential  function  satisfies  the  functional  equation 

This  follows,  when  y  and  z  are  rational,  from  the  ordinary  rules 
of  indices.  If  y  or  z,  or  both,  are  irrational  then  we  can  choose  two 
sequences  2/j, 2/2,  •••>  y^  •••  and  zlt  z2,  ..-,zn,  ...  of  rational  numbers 
such  that  lim  yn  =  y,  Km  zn  =  z.  Then,  since  the  exponential 
function  is  continuous,  we  have 

ey  x  ez  =  lim  eyn  x  lim  ezn  =  lim  eyn+zn  =  ey+z. 

In  particular  ey  x  e~y  =  e°  =  1,  or  e~y  =  l/ey. 

We  may  also  deduce  the  functional  equation  satisfied  by  ey 
from  that  satisfied  by  \ogx.  For  if  y1  =  \ogx1,  y2  =  \ogx.2,  so  that 
^  =  ey> ,  %.2  =  ey2,  then  yx  +  y2=  log x1  +  log x2  —  log xxx2  and 

ej/i+y>  =  glogx.x,  _  XiXz  —  ey,  x  eyK 


366  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [lX 

Examples  LXXXV.  1.  If  dx/dy  =  ax  then  x  =  Keay,  where  K  is  a 
constant. 

2.  There  is  no  solution  of  the  equation  f(y  +  z)=f(y)f(z)  fundamentally 
distinct  from  the  exponential  function.  [We  assume  that  / (3/)  has  a  differential 
coefficient.  Differentiating  the  equation  with  respect  to  y  and  z  in  turn,  we 
obtain 

/'(y+*)=/'(y)/(»).  /'(y+»)=/(y)/'0 

and  so  f  Q/)lf(y)=f  (z)lf(z),  and  therefore  each  is  constant.     Thus  if  x=f{y) 
then  dxjdy  =  ax,  where  a  is  a  constant,  so  that  x  =  Keay  (Ex.  1).] 

3.  Prove  that  (eay  -  1  )/y  -*-  a  as  y  ->-  0.  [Applying  the  Mean  Value 
Theorem,  we  obtain  eay  -  1  =  aye«1),  where  0  <  1 17 1  <  \y  | .] 

206.  (3)  The  function  ey  tends  to  infinity  with  y  more  rapidly 
than  any  power  of  y,  or 

lim  yaJey  =  lim  e~yya  =  0 

as  y  ->■  00  ,  for  all  values  of  a  however  great. 

We  saw  that  (\ogx)/x^-»~0  as  #-*-oo,  for  any  positive  value 
of  /3  however  small.  Writing  a  for  1//3,  we  see  that  (\ogx)a/x-*-0 
for  any  value  of  a  however  large.  The  result  follows  on  putting 
x  =  ey.  It  is  clear  also  that  e"*y  tends  to  co  if  7  >  0,  and  to  0  if 
7  <  0,  and  in  each  case  more  rapidly  than  any  power  of  y. 

From  this  result  it  follows  that  we  can  construct  a  '  scale  of  infinity ' 
similar  to  that  constructed  in  §  202,  but  extending  in  the  opposite  direction  ; 
i.e.  a  scale  of  functions  which  tend  to  00  more  and  more  rapidly  as  a;-*- so.* 
The  scale  is 

/v»        /y>l        /yO  pX       pZX  /)£"  /)3j3  pG 

where  of  course  ex°,  ...,  eeI,  ...  denote  e(*'),  ...,  e(e*),   .... 

The  reader  should  try  to  apply  the  remarks  about  the  logarithmic  scale, 
made  in  §  202  and  Exs.  lxxxiv,  to  this  'exponential  scale'  also.   The  two  scales 
may  of  course  (if  the  order  of  one  is  reversed)  be  combined  into  one  scale 
...loglog:r,   ...     log.?,   ...     x,  ...     ex,  ...     ee\  .... 

207.  The  general  power  ax.  The  function  ax  has  been 
defined  only  for  rational  values  of  x,  except  in  the  particular  case 

*  The  exponential  function  was  introduced  by  inverting  the  equation  y  =  logx 
into  x  =  ey ;  and  we  have  accordingly,  up  to  the  present,  used  ?/  as  the  independent 
and  x  as  the  dependent  variable  in  discussing  its  properties.  We  shall  now  revert 
to  the  more  natural  plan  of  taking  x  as  the  independent  variable,  except  when  it  is 
necessary  to  consider  a  pair  of  equations  of  the  type  y  —  log  x,  x  =  ey  simultaneously, 
or  when  there  is  some  other  special  reason  to  the  contrary. 


205-207]  OF   A   REAL   VARIABLE  367 

when  a  =  e.  We  shall  now  consider  the  case  in  which  a  is  any 
positive  number.  Suppose  that  a;  is  a  positive  rational  number 
p/q.  Then  the  positive  value  y  of  the  power  avlq  is  given  by 
yi  =  av\  from  which  it  follows  that 

q\ogy=p  log  a,     log y  =  (p/q)  \oga  =  x  log  a, 
and  so  y  =  exloga. 

We  take  this  as  our  definition  of  ax  when  x  is  irrational.  Thus 
10v/2  =  e^2log10.  It  is  to  be  observed  that  ax,  when  x  is  irrational, 
is  defined  only  for  positive  values  of  a,  and  is  itself  essentially 
positive;  and  that  log  ax  =  x  log  a.  The  most  important  properties 
of  the  function  ax  are  as  follows. 

(1)  Whatever  value  a  may  have,  ax  x  ay  =  ax+y  and  (ax)y  =  axy. 
In  other  words  the  laws  of  indices  hold  for  irrational  no  less  than 
for  rational  indices.     For,  in  the  first  place, 

clx  x  ay  =  exlosa  x  eyl°Ba  =  e^x+y> losa  =  ax+y  ■ 

and  in  the  second 

(2)  If  a  >  1  then  ax  =  exloga  =  eaX,  where  a  is  positive.  The 
graph  of  ax  is  in  this  case  similar  to  that  of  ex,  and  a^-^x 
as  x  -*■  oo  ,  more  rapidly  than  any  power  of  x. 

If  a  <  1  then  aa;  =  ea;loga=  e-^,  where  /3  is  positive.  The  graph 
of  ax  is  then  similar  in  shape  to  that  of  ex,  but  reversed  as  regards 
right  and  left,  and  ax-^0  as  x^-cc ,  more  rapidly  than  any 
power  of  1/x. 

(3)  ax  is  a  continuous  function  of  x,  and 

Dx  ax  =  Dxexloea  =  exlosa  log  a  =  ax  log  a. 

(4)  a*  is  also  a  continuous  function  of  a,  and 

Da  ax  =  Da  exlosa  =  exlosa  (x/a)  =  xax'\ 

(5)  (ax  —  l)/x  ^-loga  as  x-^0.  This  of  course  is  a  mere 
corollary  from  the  fact  that  Dxax  =  ax  log  a,  but  the  particular 
form  of  the  result  is  often  useful ;  it  is  of  course  equivalent  to  the 
result  (Ex.  LXXXV.  3)  that  (e"*  —  l)/x  ^aasa'-*0. 

In  the  course  of  the  preceding  chapters  a  great  many  results  involving 
the  function  ax  have  been  stated  with  the  limitation  that  x  is  rational.  The 
definition  and  theorems  given  in  this  section  enable  us  to  remove  this 
restriction. 


368  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

208.     The  representation  of  ex   as  a  limit.     In  Ch.  IV, 

§73,  we  proved  that  {l+(l/n)\n  tends,  as  n-*-cc,  to  a  limit 
which  we  denoted  provisionally  by  e.  We  shall  now  identify  this 
limit  with  the  number  e  of  the  preceding  sections.  We  can 
however  establish  a  more  general  result,  viz.  that  expressed  by 
the  equations 

lim  (l  +  -)l  =  \im(l--)~a  =  ex    (1). 

n+*  \         n/      »*»\         »/ 

As  the  result  is  of  very  great  importance,  we  shall  indicate  alter- 
native lines  of  proof. 

(1)     Since 

Ct    1  ,-  -v  CO 

^a  +  aO-TTrt' 

it  follows  that 

log(l  +  xh) 
lim     °     , =  x. 

If  we  put  h  =  \j%,  we  see  chat 

(        x^ 
lim   £  log  ( 1  +  -s  ■  =  w 

as  f  -^  oo  or  £  -^  —  oo  .  Since  the  exponential  function  is  con- 
tinuous it  follows  that 


( 


&* 


1  +    ]  =  eeiog{i+(*/f)}^.g» 
as  £  -^-  go  or  %-*•—  °o  :  i.e.  that 

lim(l  +  C)=  lim  (l  +  f\=&  (2). 


If  we  suppose  that  £  -*-  oo  or  £-^> —  oo  through  integral  values 
only,  we  obtain  the  result  expressed  by  the  equations  (1). 

(2)     If  n  is  any  positive  integer,  however  large,  and  x  >  1,  we  have 

fx     dt  fxdt       fx     dt 

]l  fiT^)<]1  7<J1<I^(i/n)» 

or         i  n(\-x~'lln)  <\ogx  <n{xlin-\)     (3). 

Writing  y  for  log.r,  so  that  y  is  positive  and  x  =  ev,  wc  obtain,  after  some 
simple  transformations, 

H)"<*<nr «• 

Now  let 

V  ,     V      1 

l+-  =  '?i,      1--  =  -. 
n  n     t]2 


208-210]  OF   A   REAL   VARIABLE  369 

Then  0<?;1<^2,  at  any  rate  for  sufficiently  large  values  of  n ;  and,  by 
(9)  of  §74, 

V2n ~  7l"  < W«?2n~ l  (t?2 -  171)  =  fn2n/n> 

which  evidently  tends  to  0  as  n-*-<x> .  The  result  now  follows  from  the 
inequalities  (4).  The  more  general  result  (2)  may  be  proved  in  the  same  way, 
if  we  replace  ljn  by  a  continuous  variable  h. 

209.    The  representation  of  log*  as  a  limit.      We  can  also  prove 
(cf.  §  75)  that 

lim  n  (1  —  x~i,n) = lim  n  (x1!*  —  1 )  =  log  x. ' 

For  n (#V»  -l)-n(l-arW»)«»  (xl>n - 1)  (1  - x - V*), 

which  tends  to  zero  as  %-»-oo,  since  n(x1,'n  —  l)  tends  to  a  limit  (§  75)  and 
a?-V»  to  1  (Ex.  xxvil.  10).  The  result  now  follows  from  the  inequalities  (3)  of 
§  208. 

Examples  LXXXVI.     1.     Prove,  by  taking  y=l  and  n=6  in  the  in- 
equalities (4)  of  §  208,  that  2 .  5  <  e  <  2 .  9. 

2.  Prove  that  if  t>  1  then  {tlin-t-Vn)l(t-t-1)  <l}n,  and  so  that  if 
x  >  1  then 

/•*     dt  [»    dt         \[x(  _  ±\dA_l(       1       ' 

J  it1-  M        J  !  t1  +  (» ■'»)  <  M  J  1    V        «/    *    ~  »   V  +  *  _ 

Hence  deduce  the  results  of  §  209. 

3.  If  |B  is  a  function  of  n  such  that  n|n  ->•  I  as  w  -*-  qo  ,  then  (1  +  £»)"-»-  e\ 
[Writing  n  log  (1 +£„)  in  the  form 

and  using  Ex.  LXXXH.  4,  we  see  that  n  log  (1  +  £„)-»-£.] 

4.  If  «£„  -a-  00 ,  then  (1  +  |„)n  -»-  00  ;  and  if  1  +  £n  >  0  and  ngn  -*--»,  then 

(l  +  ln)"-0. 

5.  Deduce  from  (1)  of  §  208  the  theorem  that  ey  tends  to  infinity  more 
rapidly  than  any  power  of  y. 

210.  Common  logarithms.  The  reader  is  probably  familiar 
with  the  idea  of  a  logarithm  and  its  use  in  numerical  calculation. 
He  will  remember  that  in  elementary  algebra  \ogax,  the  logarithm 
of  x  to  the  base  a,  is  defined  by  the  equations 

x  =  ay,    y  =  \oga  x. 

This  definition  is  of  course  applicable  only  when  y  is  rational, 
though  this  point  is  often  passed  over  in  silence. 

h.  24 


370  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

Our  logarithms  are  therefore  logarithms  to  the  base  e.  For 
numerical  work  logarithms  to  the  base  10  are  used.     If 

y  —  log  x  =  logv  x,     z  =  log10  x, 
then  x  =  ev  and  also  x  =  10-'  =  ez  log,°,  so  that 
log10#  =  (loge#)/(loge10). 

Thus  it  is  easy  to  pass  from  one  system  to  the  other  when  once 
loge10  has  been  calculated. 

It  is  no  part  of  our  purpose  in  this  book  to  go  into  details 
concerning  the  practical  uses  of  logarithms.  If  the  reader  is 
not  familiar  with  them  he  should  consult  some  text-book  on 
Elementary  Algebra  or  Trigonometry.* 

Examples  LXXXVII.     1.     Show  that 

Dxe°*  cosbx=reax  cos  {bx  +  6),     Dxe(txsm  bx  =  reax  sin  (bx+6) 

where  r=v/(a2  +  62),  eos  6  =  ajr,  sin  6  =  bjr.  Hence  determine  the  nth.  deri- 
vatives of  the  functions  ettXcosbx,  eax&mbx,  and  show  in  particular  that 
Dxn  eux= a"  eax. 

2.  Trace  the  curve  y=e~ax  sin  bx,  whore  a  and  b  are  positive.  Show 
that  y  has  an  infinity  of  maxima  whose  values  form  a  geometrical  progression 
and  which  lie  on  the  curve 

y=  — _ — r^e-**.  (Math.  Trip.  1912.) 

3     sf{a2  f  b2)  y  f  ) 

3.  Integrals  containing  the  exponential  function.     Prove  that 

/7     ,       a  cos  bx  +  b  sin  bx  „„         f  „„   .     t     ,       «  sin  bx  —  b  cos  bx     „ 
eax  cos  bxdx  = 5 — r^ eax.        \  eax  sin  bxdx=  —     — 5 — r„ ew. 
a--\-bi                      J                                   a2  +  b- 

[Denoting  the  two  integrals  by  7,  J,  and  integrating  by  parts,  we  obtain 

al=  eax  cos  bx  +  bJ,     a  J  =  eax  sin  bx  -  bl. 

Solve  these  equations  for  /and  J.] 

4.  Prove  that  the  successive  areas  bounded  by  the  curve  of  Ex.  2  and  the 
positive  half  of  the  axis  of  x  form  a  geometrical  progression,  and  that  their 
sum  is 

b      \+e-aT-'b 
a*  +  b*l-e-an/b' 

5.  Prove  that  if  a  >  0  then 

I     e~ax  cos  bxdx=-^ — r„,     |     e~ax  sin  bxdx  =  — ■. — r^. 
Jo  «2  +  ^'2     Jo  "'  +  b2 


*  See  for  example  Chrystal's  Algebra,  vol.  i,  ch.  xxi.      The  value  of  loge  10  is 
2  302...  and  that  of  its  reciprocal  -434...  . 


210]  OF    A   REAL    VARIABLE  371 

6.  Ifln=leaxxndx  then  aln=eaxxn-nln_i.     [Integrate  by  parts.     It 
follows  that  In  can  be  calculated  for  all  positive  integral  values  of  n.] 

7.  Prove  that,  if  n  is  a  positive  integer,  then 

e~xxndx  =  n\  e~f  ( ef-  1  -£-  f-"-...-^-r ) 
\  2 !  n !/ 


/: 


and  /    e~xxndx  =  ?i !. 

o 


/; 


8.  Show  how  to  find  the  integral  of  any  rational  function  of  ex.  [Put 
x  —  logu,  when  ex=%(,  dx/du  =  l/u,  and  the  integral  is  transformed  into  that 
of  a  rational  function  of  u.] 

9.  Integrate 


(c  V + a2e  ~  x)  (cV  +  b2e  ~  x) ' 
distinguishing  the  cases  in  which  a  is  and  is  not  equal  to  b. 

10.  Prove  that  we  can  integrate  any  function  of  the  form  P(x,  eax,  ebx, ...), 
where  P  denotes  a  polynomial.  [This  follows  from  the  fact  that  P  can  be 
expressed  as  the  sum  of  a  number  of  terms  of  the  type  Axmekx,  where  m  is  a 
positive  integer.] 

11.  Show  how  to  integrate  any  function  of  the  form 

P  (x,  eax,  ebx,  ...,  cos  Ix,  cos?nx,  ...,  sinlx,  sinm#,  ...). 

12.  Prove  that  I    e-toR(x)dx,  where  A>0  and  a  is  greater  than  the 

greatest  root  of  the  denominator  of  R  (x),  is  convergent.  [This  follows  from 
the  fact  that  ekx  tends  to  infinity  more  rapidly  than  any  power  of  x.] 

13.  Prove  that  I      e-^+wdx,  where  X  >  0,  is  convergent  for  all  values  of 

fx,  and  that  the  same  is  true  of  /  e-^x2+iJ.xxndx,  where  n  is  any  positive 
integer. 

14.  Draw  the  graphs  of  e*2,  e'*2,  xex,  xe~x,  xe*2,  xe~x2,  and  .rlog^,  deter- 
mining any  maxima  and  minima  of  the  functions  and  any  points  of  inflexion 
on  their  graphs. 

15.  Show  that  the  equation  eax=bx,  where  a  and  b  are  positive,  has  two 
real  roots,  one,  or  none,  according  as  b>ae,  b  =  ae,  or  b  <  ae.  [The  tangent 
to  the  curve  y  =  eax  at  the  point  (£,  ea%)  is 

y  —  e°£=ae<£(x  —  £;), 

which  passes  through  the  origin  if  a|=l,  so  that  the  line  y=aex  touches  the 
curve  at  the  point  (Ija,  e).  The  result  now  becomes  obvious  when  we  draw 
the  line  y  =  bx.  The  reader  should  discuss  the  cases  in  which  a  or  6  or  both 
are  negative.] 

24—2 


372  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

16.  Show  that  the  equation  ex  =  l+x  has  no  real  root  except  x=0,  and 
that  ex=l+x+%x2  has  three  real  roots. 

1 7.  Draw  the  graphs  of  the  functions 

log  {x  +  s/(x2  +  1)},       log  [yZ^)  '       e~aX  cos2  hx> 
e-(W,     e-(i/*)V(l/#),     e~cotx,     e~cot2x. 

18.  Determine  roughly  the  positions  of  the  real  roots  of  the  equations 
log{*+V(*l+l)}-i*j6,     e*-|r|=io5oO.     e*sin*=7,     «*•  sin  tf=10000. 

19.  The  hyperbolic  functions.  The  hyperbolic  functions  cosh  x,* 
sinha?,  ...  are  denned  by  the  equations 

cosh  x= \  (ex  +  e  ~~  x),      sinh  a;  =  ^  (e*  —  e  ~  x), 

tanh  a; = (sinh  .r)/(cosh  a;),  coth  x=  (cosh  a;)/(sinh  a;), 

sech  x=  l/(cosh  x),  cosech  x  =  l/(sinh  a;). 

Draw  the  graphs  of  these  functions. 

20.  Establish  the  formulae 

cosh  ( —  x)  =  cosh  x,    sinh  ( —  x)  =  —  sinh  x,     tanh  ( —  x)  =  —  tanh  x, 

cosh2  x— sinh2  a?=l,     sech2  .r  +  tanh2  x=l,     coth2  a;  —  cosech2  a?=l, 

cosh  2x= cosh2  a?  +  sinh2  a,',     sinh  2a;  =  2  sinh  a?  cosh  #, 

cosh  (x +y)  =  cosh  x  cosh  3/  +  sinh  x  sinh  ?/, 

sinh  (a; + y)  =  sinh  #  cosh  y  -f  cosh  a?  sinh  y. 

21.  Verify  that  these  formulae  may  be  deduced  from  the  corresponding 
formulae  in  cos  x  and  sin  x,  by  writing  cosh  x  for  cos  x  and  i  sinh  x  for  sin  a;. 

[It  follows  that  the  same  is  true  of  all  the  formulae  involving  cos  nx  and 
sin  nx  which  are  deduced  from  the  corresponding  elementary  properties  of 
cos  a;  and  sin  a?.     The  reason  of  this  analogy  will  appear  in  Ch.  X.] 

22.  Express  cosh  x  and  sinh  x  in  terms  (a)  of  cosh  2x  (b)  of  sinh  2x. 
Discuss  any  ambiguities  of  sign  that  may  occur.  (Math.  Trip.  1908.) 

23.  Prove  that 

2)3 cosh #= sinh  a?,  Dx sinh  x  =  cosh x,  Dxt&nh x=  sech2a;,  Z)^cotha;=  — cosech2 a\ 

Dx  sech  x  =  -  sech  x  tanh  x,     Dx  cosech  x  =  —  cosech  x  coth  a;, 

Z>^  log  cosh  x = tanh  x,  Dx  log  |  sinh  a?  |  =  coth  a?, 

Dx  arc  tan  e* = J  sech  a-',     Z)^  log  |  tanh  \  x  \  =  cosech  x. 

[All  these  formulae  may  of  course  be  transformed  into  formulae  in  inte- 
gration.] 

*  '  Hyperbolic  cosine  '  :  for  an  explanation  of  this  phrase  see  Hobson's  Trigo- 
nometry,  ch.  xvi. 


210]  OF    A    REAL   VARIABLE  373 

24.  Prove  that  cosh  x  >  1  and  —  1  <  tanh  x  <  1. 

25.  Prove  that  if  y=coskx  then  x=\og{y±s?(y2  —  1)},  if  v/  =  sinh.r  then 
«F=log[y-rV(«/2  +  l)},  and  if  j/  =  tanh.r  then  x  =  ^\og  {(l+y)/(l  —  y)}.  Account 
for  the  ambiguity  of  sign  in  the  first  case. 

26.  We  shall  denote  the  functions  inverse  to  cosh  x,  sinh  x,  tanh  x  by 
arg  cosh  x,  arg  sinh  x,  arg  tanh  x.  Show  that  arg  cosh  x  is  defined  only  when 
.r^l,  and  is  in  general  two-valued,  while  arg  sinh  x  is  defined  for  all  real 
values  of  x,  and  arg  tanh  x  when  —  1<.v<1,  and  both  of  the  two  latter 
functions  are  one- valued.     Sketch  the  graphs  of  the  functions. 

27.  Show  that  if  -\ir  <x  <\n  and  y  is  positive,  and  cos x cosh y  =  \,  then 

y  =  log  (sec  x + tan  x),     Dxy  —  sec  x,     By  x  =  sech  y. 

f       dx  .  f      dx 

28.  Prove  that  if  a  >  0  then  I  — — r — 57  =  arg  sinh  (xla),  and  I  ,.     ' — 5.  is 

J  Jix'  +  a*)  Js/(x~-a~) 

equal  to  arg  cosh  (x/a)  or  to  —  arg  cosh  ( —  x/a),  according  as  x>  0  or  x  < 0. 

'    f    dx 

29.  Prove  that  if  a>0  then  I  — ^ 2  is  equal  to  —  (1/a)  arg  tanh  (.r/a)  or 

to  —  (1/a)  arg  coth  (x/a),  according  as  |  x  |  is  less  than  or  greater  than  a.  [The 
results  of  Exs.  28  and  29  furnish  us  with  an  alternative  method  of  writing 
a  good  many  of  the  formulae  of  Ch.  VI.] 

30.  Prove  that 


l^Mr2l0g{v'(""a)W(l"A)}        (a<b<x), 
=  - 2 log y(a - x)  +  sf(b - x)}  (x«z< b), 


sj{(a  —  x)(b-x)} 

/  -777 rrr r  =2arctan  .  /('^ )  (a<x<b). 

j  J{{x-a)(b-x)}  V   \b-xj  K 

31.  Prove  that 

I   ^log(l  +  i^)«T^=|-|log-o<if1.r2^  =  ^. 
/o  70 

(1/a^.  ZWjt>.  1913.) 

32.  Solve  the  equation  a  cosh  .r  +  b  sinh  .r  =  c,  where  c  >  0,  showing  that  it 
has  no  real  roots  if  b2  +  c2  —  a2  <  0,  while  if  62-f  c2  — a2>0  it  has  two,  one,  or 
no  real  roots  according  as  a +  6  and  a  —  b  are  both  positive,  of  opposite  signs, 
or  both  negative.     Discuss  the  case  in  which  b2  +  c1  —  a2  =  0. 

33.  Solve  the  simultaneous  equations  cosh  x  cosh  y  =  a,  sinh  x  sinh  y  =  b. 

34.  xV*  -^  1  as  x  -*-  00  .  [For  a£/*=gP<**)/*  and  (log  a?)/*-  -*-  0.  Cf. 
Ex.  xxvii.  11.]  Show  also  that  the  function  xllx  has  a  maximum  when 
x  =  e,  and  draw  the  graph  of  the  function  for  positive  values  of  x. 

35.  xx-*~l  as  x  +  +  0. 


374  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

36.  If  {f(n+l)}l{f(n)}-*-l,  where  l>0,  as  n-*-<x>,  then  #{/(n)}  +1. 
[For  log/(ji  +  l)-log/(»)-^l°g^  and  so  (l/«)log/(ra)-*-log*  (Ch.  IV,  Misc. 
Ex.  27).] 

37.  %(n  !)/n  -*-  l/e  as  »-*-<». 

[If/(n)=»"»»l  then  {/(n+l)}/{/(n)}  =  {l  +  (l/n)}-»-*l/«.  Now  use 
Ex.  36.] 

38.  #{(2w) !  j(n  !)2}  -»•  4  as  n  ->■  oo . 

39.  Discuss  the  approximate  solution  of  the  equation  ex  =  xmom. 

[It  is  easy  to  see  by  general  graphical  considerations  that  the  equation 
has  two  positive  roots,  one  a  little  greater  than  1  and  one  very  large*  and  one 
negative  root  a  little  greater  than  - 1.  To  determine  roughly  the  size  of  the 
large  positive  root  we  may  proceed  as  follows.    If  e* =#1000000  t^eil 

(\o<y  lo°'  x\ 
1  +    ^.gg     ) , 

roughly,  since  13-82  and  263  are  approximate  values  of  log  106  and  log  log  106 
respectively.  It  is  easy  to  see  from  these  equations  that  the  ratios  log  x  :  1382 
and  log  log  x  :  2-63  do  not  differ  greatly  from  unity,  and  that 

a?=10G  (13-82 +log log  «)  =  10°  (13-82 +  2-63)  =  16450000 
gives  a  tolerable  approximation  to  the  root,  the  error  involved  being  roughly 
measured  by  10°  (log  log  #-2-63)  or  (10°loglogx-)/13-82  or  (10ex  2-63)/13-82, 
which  is  less  than  200,000.     The  approximations  are  of  course  very  rough, 
but  suffice  to  give  us  a  good  idea  of  the  scale  of  magnitude  of  the  root.] 

40.  Discuss  similarly  the  equations  e*=  1000000 .£1000000,  e*5  =  #1000000000. 

211.     Logarithmic  tests   of  convergence  for  series  and 
integrals.     We  showed  in  Ch.  VIII  (§§  175  et  seq.)  that 


fff'j.if       <a>0) 


are  convergent  if  s  >  1  and  divergent  if  s  ^  1.     Thus  2(1/??)  is 
divergent,  but  2  n~x~a  is  convergent  for  all  positive  values  of  a. 

We  saw  however  in  §  200  that  with  the  aid  of  logarithms  we 
can  construct  functions  which  tend  to  zero,  as  n  -*■  00 ,  more 
rapidly  than  1/n,  yet  less  rapidly  than  n-1_a,  however  small  a  may 
be,  provided  of  course  that  it  is  positive.  For  example  l/(?ilog?i) 
is  such  a  function,  and  the  question  as  to  whether  the  series 

n  log  n 

*  The  phrase  '  very  large '  is  of  course  not  used  here  in  the  technical  sense 
explained  in  Ch.  IV.  It  means  '  a  good  deal  larger  than  the  roots  of  such  equations 
as  usually  occur  in  elementary  mathematics '.  The  phrase  '  a  little  greater  than  ' 
must  he  interpreted  similarly. 


210,  211]  OF    A   REAL   VARIABLE  375 

is  convergent  or  divergent  cannot  be  settled  by  comparison  with 
any  series  of  the  type  S  n~s. 

The  same  is  true  of  such  series  as 

v  _  J y    log  log  n 

~  n  (log  n)- '         n  \J(\og  n) ' 

It  is  a  question  of  some  interest  to  find  tests  which  shall  enable 
us  to  decide  whether  series  such  as  these  are  convergent  or 
divergent;  and  such  tests  are  easily  deduced  from  the  Integral 
Test  of  §  174. 

For  since 

1  —  s  1 

Dx  (log  x)l~s  =  —. r-,     Bx  log  log  x  =— : , 

v    6    ;  a: (log a?)8'  8     6        a; log  a;' 

we  have 

f *       d,       m  (logf).--(logn)-    [*     d^  _  _ 

J  a  a;  (log  a;)s  1  -  *  J    a?  log  a;         &     &  ^         &     &    » 

if  a  >  1.  The  first  integral  tends  to  the  limit  —  (\oga)l~s/(l—s) 
as  £  •-*•  go  ,  if  s  >  1,  and  to  co  if  s  <  1.  The  second  integral  tends 
to  so .     Hence  the  series  and  integral 


» i_      r dx__ 

non(lognjs'   Ja  x(logx)s' 


where  nQ  and  a  are  greater  than  unity,  are  convergent  if  s>l, 
divergent  if  s  S  1. 

It  follows,  of  course,  that  £<£(«)  is  convergent  if  <f>(n)  is 
positive  and  less  than  K '/ '{n  (log  n)s) ,  where  s  >  1,  for  all  values  of  n 
greater  than  some  definite  value,  and  divergent  if  <£  (n)  is  positive 
and  greater  than  Kj{n  log  n)  for  all  values  of  n  greater  than  some 
definite  value.  And  there  is  a  corresponding  theorem  for  integrals 
which  we  may  leave  to  the  reader. 

Examples  LXXXVIII.     1.     The  series 

1  ^(log%)100  ?i2-l         1 


71  (log  ft)2'  ^101/100     '  n2  +  l?i(l0g?l)7/6 

are  convergent.  [The  convergence  of  the  first  series  is  a  direct  consequence 
of  the  theorem  of  the  preceding  section.  That  of  the  second  follows  from 
the  fact  that  (log  n)m  is  less  than  n^  for  sufficiently  large  values  of  n,  how- 
ever small  /3  may  be,  provided  that  it  is  positive.  And  so,  taking  /3  =  1/200, 
(log ft)100  ji-ioi/ioo  js  ]ess  than  n~'m/m  for  sufficiently  large  values  of  n.  The 
convergence  of  the  third  series  follows  from  the  comparison  test  at  the  end  of 
the  last  section.] 


376  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

2.  The  series 

1  1  n  log  n 

2  n  (log  nf~ '      2 ,7100/101  (i0g  w)ioo '      2  (nlognf  +  1 
are  divergent. 

3.  The  series 

(logn)P  (log7t)P(loglog?Qg  (log  log  »)> 

n1+*    '  n1+*  '         %(log?i)1  +  8' 

where  s>0,  are  convergent  for  all  values  of  p  and  q  ;  similarly  the  series 

1  1  1 

n1  ~*  (log  n)p '         n1  ~ s  (log  n)p  (log  log  n)i '         w  (log  n)1  ~  *  (log  log  ?j)p 

are  divergent. 

4.  The  question  of  the  convergence  or  divergence  of  such  series  as 

1  log  log  log  n 

n  log  n  log  log  n '         11  log  n  ^/(log  log  ?i) 

cannot  be  settled  by  the  theorem  of  p.  375,  since  in  each  case  the  function 
under  the  sign  of  summation  tends  to  zero  more  rapidly  than  l/(?ilogn)  yet 
less  rapidly  than  n~l  (log?i)_1_a,  where  a  is  any  positive  number  however 
small.  For  such  series  we  need  a  still  more  delicate  test.  The  reader  should 
be  able,  starting  from  the  equations 

1  —  s 

x  (  OgkX)        - ^ l0g  -p  log2 x  ^ .logfc_!  X (lOgfc X)»  ' 

Dx  logfc+  i  X=  —, 7 ; ; , 

°K+l  #logX-log2A'...logi:_1.zlogjl;.'t' 

where  log2  #=loglog.r,  log3.v=logloglog  x,  ...,  to  prove  the  following 
theorem  :    the  series  and  integral 

<» 1  r dx 

n0  wlog»logaw...logt_1?i  (logfc  n)° '     J  a  x  log  x  log2  X  . . .  log*  _  1  X  (logjt  x)» 

are  convergent  if  s~>\  and  divergent  if  sgl,  n0  and  a  being  any  numbers 
sufficiently  great  to  ensure  that  logfc?i  and  log^A'  are  positive  when  n^.n() 
or  x^a.  These  values  of  n0  and  a  increase  very  rapidly  as  k  increases: 
thus  log#>0  requires  x>l,  log2.?'>0  requires  x>e,  loglog#>0  requires 
x  >  ee,  and  so  on  ;  and  it  is  easy  to  see  that  ee  >  10,  e<?  >  e10  >  20,000, 
gee<!  •->  e2o,ooo  ->  10s000. 

The  reader  should  observe  the  extreme  rapidity  with  which  the  higher 
exponential  functions,  such  as  &eX  and  ee  ,  increase  with  x.  The  same 
remark  of  course  applies  to  such  functions  as  daX  and  Cla  ,  where  a  has 
any  value  greater  than  unity  It  has  been  computed  that  999  has  369,693,100 
figures,  while  lOlO™  has  of  course  10,000,000,000.  Conversely,  the  rate  of 
increase  of  the  higher  logarithmic  functions  is  extremely  slow.  Thus  to  make 
log  log  log  log x  >  1  we  have  to  suppose  x  a  number  with  over  8000  figures.* 

*  See  the  footnote  to  p.  362. 


211]  OF    A    REAL   VARIABLE  377 

5.     Prove  that  the  integral    /     -  jlog  (  -U  dx,  where  0  <  a  <  1,  is  con- 
vergent if  s  <  —  1,  divergent  if  s  >  -  1.     [Consider  the  behaviour  of 


i:i{ 


X) 


as  e-*-+0.  This  result  also  may  be  refined  upon  by  the  introduction  of 
higher  logarithmic  factors.] 

6.  Prove  that    /    -<  log  ( - )  j-  das  has  no  meaning  for  any  value   of  s. 

[The  last  example  shows  that  s  <  —  1  is  a  necessary  condition  for  convergence 
at  the  lower  limit :  but  {log(l/.r)}8  tends  to  co  like  (1—  x)3,  as  a'-*-1-0,  if  s 
is  negative,  and  so  the  integral  diverges  at  the  upper  limit  when  s  <  -  1.] 

7.  The    necessary   and    sufficient    conditions    for    the    convergence   of 
/    x*-1  -jlog (  -  )[  dx  are  a>0,  s>  -1. 

Examples  LXXXIX.    1.    Euler's  limit.    Show  that 

*(w)=l  +  2  +  3  +  ...+^ri-log» 

tends  to  a  limit  y  as  ?i-=>-ao,  and  that  0<y^l.  [This  follows  at  once  from 
§  174.  The  value  of  y  is  in  fact  '577...,  and  y  is  usually  called  Euler's 
constant.] 

2 .  If  a  and  b  are  positive  then 

-  +  — rr+— T177+---  +  - T7 Tw.  _  t  log  (a +  «6 

a      a  +  b      a  +  2b  a  +  (n-l)o      b 

tends  to  a  limit  as  n-*-<x> . 

3.  If  0  <  s  <  1  then 

:U9-US-U       4-fm.—  ^-*_ 

l-S 


(«,)  =  l  +  2-«+3-8+...  +  (n-l)-s- 


tends  to  a  limit  as  n^-  oo  . 

4.  Show  that  the  series 

1  1  1 

l  +  2(l+|)  +  3(l  +  i  +  ^)  +  '" 
is  divergent.     [Compare  the  general   term   of  the   series   with    l/(%logn).] 
Show  also  that  the  series  derived  from  2  n~8,  in  the  same  way  that  the  above 
series  is  derived  from  2  (l/«),  is  convergent  if  s  >  1  and  otherwise  divergent. 

5.  Prove  generally  that  if  "2un  is  a  series  of  positive  terms,  and 

Sn  =  Ui  +  U2  +  ...+Un, 

then  2  («n/*n_i)  is  convergent  or  divergent  according  as  2un  is  convergent  or 


378  THE   LOGARITHMIC   AND   EXPONENTIAL    FUNCTIONS  [iX 

divergent.  [If  2wn  is  convergent  then  sn_i  tends  to  a  positive  limit  I,  and  so 
2(wn/sn-i)  is  convergent.     If  2un  is  divergent  then  sn_1-^aD,  and 

mw/s„_i > log  {1  +  («»/*»- 1)}= loS (s,Ai-i) 
(Ex.  lxxxii.  1) ;  and  it  is  evident  that 

log  («j/*i)  +  log  (s3/s2)  + . . .  +  log  (*B/*„_i)  =  log  (sjsj) 
tends  to  oo  as  n -»-  co  .] 

6.  Prove  that  the  same  result  holds  for  the  series  2  («n/s»)-  [The  proof 
is  the  same  in  the  case  of  convergence.  If  "2un  is  divergent,  and  un<sn_1 
from  a  certain  value  of  n  onwards,  then  sn<2sn_1,  and  the  divergence  of 
2  (ujsn)  follows  from  that  of  2  {ujsn_^).  If  on  the  other  hand  «„>s„_i  for 
an  infinity  of  values  of  n,  as  might  happen  with  a  rapidly  divergent  series, 
then  un/sn  ^  J  for  all  these  values  of  n.] 

7.  Sum  the  series  1  —  -|  +§■-... .     [We  have 

1+^+-  +  ^=log(2"  +  1)+r+f"'  2 G  +  i+-  +^l)=1°s(?i+1)+y+f«'> 

by  Ex.  1,  y  denoting  Euler's  constant,  and  en,  en'  being  numbers  which  tend 
to  zero  as  n-*-<x> .  Subtracting  and  making  «-*•<»  we  see  that  the  sum  of  the 
given  series  is  log  2.     See  also  §  213.] 

8.  Prove  that  the  series 

2(-l)"(l+l  +  ...  +  ?TlT-log,-0) 

oscillates  finitely  except  when  C=y,  when  it  converges. 

212.     Series  connected  with  the  exponential  and  log- 
arithmic functions.    Expansion  of  ex  by  Taylor's  Theorem. 

Since  all  the  derivatives  of  the  exponential  function  are  equal 
to  the  function  itself,  we  have 

2!  (n  —  1)!      nl 

where  0<  6<  1.  But  xnjn !  -*-9  as  n~>-  oo  ,  whatever  be  the  value  of  x 
(Ex.  xxvii.  12);  and  eex  <  ex.    Hence,  making  n  tend  to  oo  ,  we  have 

e*  =  i  +  tf  +  |:  +  ...  +  2  + (i). 

The  series  on  the  right-hand  side  of  this  equation  is  known  as 
the  exponential  series.     In  particular  we  have 


.-1  +  1+1+...  +  I  + (2); 


and  so 


,      ,       1                  1           V      _             of-                xn  /ox 

1  +  1  +  -  +  .. .  +  -  +  ...     =l+x  +  -+...+  -  + (3), 

2!  nl  I  2!  nl 


211,  212]  OF   A   REAL   VARIABLE  379 

a  result  known  as  the  exponential  theorem.     Also 

i                 •    t        \      (#loga)2  ... 

ax  =  ex]0-a  =  1  +  (w\og  a)  +  v — f^—L  + (4) 

for  all  positive  values  of  a. 

The  reader  will  observe  that  the  exponential  series  has  the  property  of 
reproducing  itself  when  every  term  is  differentiated,  and  that  no  other  series 
of  powers  of  x  would  possess  this  property  :  for  some  further  remarks  in  this 
connection  see  Appendix  II. 

The  power  series  for  ex  is  so  important  that  it  is  worth  while  to  investigate 
it  by  au  alternative  method  which  does  not  depend  upon  Taylor's  Theorem. 
Let 

En(x)=l+x+~ +  ...  +  -, 

and  suppose  that  x  >  0.     Then 

(1+-y_1+M^+»ifcii(£y+...+^i>^f?y 

\       ii  J  \nj         1.2       \nj  1.2...M      \nj 

which  is  less  than  En  (x).    And,  provided  n>x,  we  have  also,  by  the  binomial 
theorem  for  a  negative  integral  exponent, 


Thus 


(1+?)"<«,(*)<(i-£f. 


But  (§  208)  the  first  and  last  functions  tend  to  the  limit  ex  as  n-a-oo,  and 
therefore  En(x)  must  do  the  same.  From  this  the  equation  (1)  follows  when 
x  is  positive  ;  its  truth  when  x  is  negative  follows  from  the  fact  that  the 
exponential  series,  as  was  shown  in  Ex.  lxxxi.  7,  satisfies  the  functional 
equation  f{x)f{y)  =f(x+y),  so  that  /  (*)/  ( -  *)  =/(0)  =  1. 

Examples  XC.     1.     Show  that 

cosh x  =  1  +  '—  + '—}  + ...,  sinh x=x+'^{  +  '—  +  .... 

2.  If  x  is  positive  then  the  greatest  term  in  the  exponential  series  is  the 
([x]  +  l)-th,  unless  x  is  an  integer,  when  the  preceding  term  is  equal  to  it. 

3.  Show  that  n  !>(n/e)re.     [For  nnjn  !  is  one  term  in  the  series  for  en.] 

4.  Prove  that  en  =  {nn\n  !)  (2  +  Sx  +  S2\  where 

l  +  i/      (l+i>)  (l+2i/j 
and  v  =  l/n;  and  deduce  that  « !  lies  between  2  (?i/ti)re  and  2  (ii  + 1)  (n/e)B. 

5.  Employ  the  exponential  series  to  prove  that  ex  tends  to  infinity  more 
rapidly  than  any  power  of  x.     [Use  the  inequality  ex>xnjii !.] 


380  THE   LOGARITHMIC   AND   EXPONENTIAL    FUNCTIONS  [iX 

6.  Show  that  e  is  not  a  rational  number.     [If  e=pjq,  where  p  and  q  are 
integers,  we  must  have 

p  -  1       1  1    , 

q  2!      3 !  q\ 

or,  multiplying  up  by  q  ! , 

q'\q  2!  qlj      q  +  1  ^  (q  +  l)  (q  +  2)^ 

and  this  is  absurd,  since  the  left-hand  side  is  integral,  and  the  right-hand 
side  less  than  {l/(q  +  l)}  +  {l/(q  +  l)}2  +  ...  =  llq.] 

to  £.11 

7.  Sum  the  series  2  Pr  {%)  — ,  where  Pr  (n)  is  a  polynomial  of  degree  r 

o  n  • 

in  n.     [We  can  express  Pr  (n)  in  the  form 

A0  +  Ain  +  A2n(n-l)  +  ...  +  A  rn(n  —  \)...  (n-r  + 1), 
and 

CO  a*H  CO    /yiTl  OO  -v>H  00  /vVTl 

2  Pr  (n)  '—.  =  A0  2  —  +  At  2  -,       -.,  +  ...+iirS  , n 

=  (A0  +  A1x  +  A2x2  +  ...+Arxr)ex.] 

8.  Show  that 

oo  ^3  oo  *t4 

2  —  xn  =  (x  +  3x'i  +  x*)ex      2—.xn  =  (x  +  7x'1  +  6x:i  +  xi)ex', 
and  that  if  £„  =  l3  +  23  +  ...+n3  then 

2  Sn~  =  k  (4r  +  ltefi  +  toiP+x*)  e*. 

!  ?i  !        * 

In  particular  the  last  series  is  equal  to  zero  when  x  —  —2.    {Math.  Trip.  1904.) 

9.  Prove  that  2  (njn !)  =  e,  2  (tf/n !)  =  2e,  2  (w3/« !)  =  be,  and  that  2  (wfc/n !), 
where  k  is  any  positive  integer,  is  a  positive  integral  multiple  of  e. 

10.  Prove  that  2  ^"V/!  =  {(*2  -  3*  +  3)  ex  +  A.r2  -  3}/**. 

!  (w  +  2)w!     u  y         2  " 

[Multiply  numerator  and  denominator  by  n  +  1,  and  proceed  as  in  Ex.  7.] 

11.  Determine  a,  b,  c  so  that  {(x  +  a)ex  +  (bx  +  c)}/x3  tends  to  a  limit  as 

x-*-0,  evaluate  the  limit,  and  draw  the  graph  of  the  function  ex  +  — . 

°    r  x+a 

12.  Draw  the  graphs  of  1+x,  \+x+^x2,  1  +  x  +  \x2  +  ^x3,  and  compare 
them  with  that  of  ex. 

13.  Prove  that   e~x  — 1+x  — '—  +  ...  — (  —  l)n'—t   is    positive   or  negative 

Z  I  71  ! 

according  as  n  is  odd  or  even.     Deduce  the  exponential  theorem. 


212,  213]  OF    A   REAL   VARIABLE  381 

14.  If 

X0=e?,  A\  =  ex-1,  X2  =  ex-l-x,  X3  =  e*-l-x-(x2j2\),  ..., 
then  dXvjdx=Xv  _  x .     Hence  prove  that  if  t  >  0  then 

ft  ft  ft  ft  ,2 

A1(0=l    X^dx<te\  X3(t)=  I    Xldx<\   xe*dx<ef      xdx=—el, 
Jo  Jo  Jo  Jo  2 ! 

tv 
and  generally  Xv(t)<—  eK     Deduce  the  exponential  theorem. 
v ! 

15.  Show  that  the  expansion  in  powers  of  p   of  the  positive  root  of 
x2+p=a2  begins  with  the  terms 

a  { 1  -  \p  log  a  +  %p2  log  a  (2  +  log  a)} .       (Math.  Trip.  1 909. ) 

213.     The    logarithmic    series.      Another   very   important 
expansion  in  powers  of  x  is  that  for  log(l  +  x).     Since 

los(1 +*>=/!  if? 

and  1/(1  + 1)  =  1  —  t  4-  t2  —  . . .  if  £  is  numerically  less  than  unity,  it  is 
natural  to  expect*  that  log  (1  +  x)  will  be  equal,  when  —  1  <  x  <  1, 
to  the  series  obtained  by  integrating  each  term  of  the   series 

1  —  t  +  t2  — ...  from  £=0  to  t= x,  i.e.  to  the  series  x  —  \x2  +  ^ a?  — 

And  this  is  in  fact  the  case.     For 

1/(1  +  t)  =  1  -  t  +  t2  -  ...  +  (-  \yn-Hm-i  +  I         J  f 


+ 

and  so,  if  x  >  —  1, 

dt 

'x  tmdt 
+  t' 

We  require  to  show  that  the  limit  of  Rm>  when  m  tends  to  oo , 
is  zero.  This  is  almost  obvious  when  0  <  x  ^  1 ;  for  then  Rm  is 
positive  and  less  than 

xm+1 
tmdt  = 


fx      rJf  r2  Tm 

\og(l  +  x)=j^t  =  x-2  +  .--  +  (-l)m->-  +  (-l)mRm, 

fx  pt 

where  i2m  =  /    j- 

J  o  -1 


o  w  +  1  ' 

and  therefore  less  than  l/(m  +  1).    If  on  the  other  hand  — 1<#<0, 
we  put  t  =  —  u  and  x  =  —  £,  so  that 

*  See  Appendix  II  for  some  further  remarks  on  this  subject. 


382  THE   LOGARITHMIC    AND   EXPONENTIAL    FUNCTIONS  [iX 

which  shows  that  Rm  has  the  sign  of  (—  1)"\  Also,  since  the 
greatest  value  of  1/(1  —  u)  in  the  range  of  integration  is  1/(1  -  £), 
we  have 

0  <  I  Rm  I  <  = s      umdu  =  j-  -4m k  < 


1-fJo  (ro  +  l)(l-£)     (m+l)(l-f)' 

and  so  i2m  -*■  0. 

Hence  log  (1  +  x)  =  x  —  %a?  +  ^a?—  ..., 

provided  that  —  1  <oc  ^  1.  If  a;  lies  outside  these  limits  the  series 
is  not  convergent.     If  co  =  1  we  obtain 

loga-i-j+i-..., 

a  result  already  proved  otherwise  (Ex.  lxxxix.  7). 

214.     The  series  for  the  inverse  tangent.     It  is  easy  to 
prove  in  a  similar  manner  that 

fx     dt         fx 
arc  tan  x  =  I       ——  —  \    (1  -  t-  +  t*  —  . . .)  dt 

J  o  L  +  t~     Jo 

—  rp  __    1  /y>3   _!_.    1  /y.5  _ 

—  iAj  o  tt/       T^    X  •  •  •   i 

provided  that  -lS^gl.  The  only  difference  is  that  the  proof  is 
a  little  simpler ;  for,  since  arc  tan  x  is  an  odd  function  of  x,  we  need 
only  consider  positive  values  of  x.  And  the  series  is  convergent 
when  x  =  —  1  as  well  as  when  x  =  1.  We  leave  the  discussion  to  the 
reader.  The  value  of  arc  tan  x  which  is  represented  by  the  series 
is  of  course  that  which  lies  between  —  ^ir  and  \nr  when  —  1  g  x  ^  1, 
and  which  we  saw  in  Ch.  VII  (Ex.  lxiii.  3)  to  be  the  value 
represented  by  the  integral.     If  x=  1,  we  obtain  the  formula 

J-7T—  1    —1-4-1  — 
47T—  X         3  -T  5  

Examples  XCI.    1.    log(^-A=z+$x*+$x*  +  ...  if  -1  <.r<l. 

2.  arg  fcanh  x = \  log  h^jj  =#+ i  *" + fr*6  + . . .  if  - 1  <x <  1  ■ 

3.  Prove  that  if  x  is  positive  then 

i»g(i+*>=if!.+i(r^)2+i(i^)!+"" 

(i/aC/;.  TWp.  1911.) 

4.  Obtain  the  series  for  log(l+#)  and  arc  tan  a;  by  means  of  Taylor's 
theorem. 

[A  difficulty  presents  itself  in  the  discussion  of  the  remainder  in  the 


213,  214]  OF   A   REAL   VARIABLE  383 

first  series  when  x  is  negative,  if  Lagrange's  form  Rn  =  (  —  l)n~1xn/{n  (l  +  8x)n} 
is  used ;  Cauchy's  form,  viz. 

#„=(-i)*-1(i-0)B-1*,7(i+^)ns 

should  be  used  (cf.  the  corresponding  discussion  for  the  Binomial  Series, 
Ex.  lvi.  2  and  §  163). 

In  the  case  of  the  second  series  we  have 
Dxn  arc  tan  x  =  Dxn  ~ 1  {1  /( 1  +  .r2)} 

=  (  -  l)*-i(n-l) !  (x2  +  l)-'"'2  sin  {n  arc  tan(l/a?)} 
(Ex.  xlv.  11),  and  there  is  no  difficulty  about  the  remainder,  which  is  obviously 
not  greater  in  absolute  value  than  l/n.*] 

5.  If  y>0  then 

[Use  the  identity  y  =  ( 1  +  - — = )  /  ( 1  —  - — - J .     This  series  may  be  used  to 

calculate  log  2,  a  purpose  for  which  the  series  \  —  \  +  \  —  ...,  owing  to  the 
slowness  of  its  convergence,  is  practically  useless.  Put  y  =  2  and  find  log  2 
to  3  places  of  decimals.] 

6.  Find  log  10  to  3  places  of  decimals  from  the  formula 

log  10  =  3  log  2  +  log  (1+ 1). 

7.  Prove  that 

/a?+l\_g  f    1  1  1  \ 

gV   oo    )      '  Uf  +  l  +  3(2a,'  +  l)3  +  5(2^+l)5  +  "-J 

if  x  >  0,  and  that 

v(x-l)2(x  +  2)_     (     2  1/ 2     \3      1/     2 y 

g  (x  +  lf(x-2)~     \x^-3x  +  3\xs-3x)  +  5\x^-3x)  +"" 
if  x>2.     Given   that  log  2  = -6931471...    and   log  3  =  1-0986123...,  show,  by 
putting  #  =  10  in  the  second  formula,  that  log  11  =  2-397895.... 

(Math.  Trip.  1912.) 

8.  Show  that  if  log  2,  log  5,  and  log  11  are  known,  then  the  formula 

logl3  =  3logll+log5-91og2 
gives  log  13  with  an  error  practically  equal  to  -00015.  (Math.  Trip.  1910.) 

9.  Show  that 

^log2  =  7<x  +  56  +  3c,     h\og3  =  lla  +  8b+5c,     £  log5  =  16a  +  126+7c, 
where  a  =  argtanh  (1/31),  6=arg  tanh  (1/49),  c  =  arg  tanh  (1/161). 

[These  formulae  enable  us  to  find  log  2,  log  3,  and  log  5  rapidly  and  with 
any  degree  of  accuracy.] 

*  The  formula  for  Dxn  arc  tan  x  fails  when  x  =  0,  as  arctan(l/;r)  is  then 
undefined.  It  is  easy  to  see  (cf.  Ex.  xlv.  11)  that  arc  tan  (1/x)  must  then  be 
interpreted  as  meaning  \tt. 


38-i  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

10.  Show  that 

J7r=arctan  (l/2)  +  arctan  (1/3)  =  4  arc  tan  (1/5) -arc tan (1/239), 
and  calculate  it  to  6  places  of  decimals. 

11.  Show  that  the  expansion  of  (1  +  x)1  +  x  in  powers  of  x  begins  with  the 
terms  l+x  +  x-  + 1  afl.  (Math.  Trip.  1910.) 


12.     Show  that 

'1+A  =  logio  « 
x   J       24» 


\ogwe-y/{x(x+l)}  logic  v- 


approximately,  for  large  values  of  x.     Apply  the  formula,  when  #=10,  to 
obtain  an  approximate  value  of  log10  e,  and  estimate  the  accuracy  of  the  result. 

(Math.  Trip.  1910.) 

13.  Show  that     -^—  log  (-^—)  =  x  +  (l+%)x2  +  (l+%  +  ±)x*  +  ..., 

1.  —  00  \X       A/  J 

if  -  1<  x  <  1.     [Use  Ex.  lxxxi.  2.] 

14.  Using  the  logarithmic  series  and  the  facts  that  log102-3758  =  •3758099... 
and  log10e  =  '4343...,  show  that  an  approximate  solution  of  the  equation 
x=100  log10#  is  237-58121.  (Math.  Trip.  1910.) 

15.  Expand  log  cos  x  and  log  (sin  #//£)  in  powers  of  x  as  far  as  x\  and 
verify  that,  to  this  order, 

log  sin  x  =  log  x  —  ^g  log  cos  x + ff  log  cos  \ x. 

(Math.  Trip.  1908.) 

fx    dt 

16.  Show  that         — — T=ai-^sfi+\ofi- ...  if  -1  <a?^l.     Deduce  that 

y  o  i  +  ' 

l-^+J-...=  {»r+2log(>/2  +  l)}/4V2.    (J/a*A.  TWp.  1896.) 
[Proceed  as  in  §  214  and  use  the  result  of  Ex.  xlviii.  7.J 

17.  Prove  similarly  that 

i"Hl1r-.-  =  Jol1^={--21og(v/2  +  l)}/4N/2. 

18.  Prove  generally  that  if  a  and  b  are  positive  integers  then 

1 1_  1  _  nta-1dt 

a     a  +  b  + a  +  2b     "'~]0l  +  tb' 

and  so  that  the  sum  of  the  series  can  be  found.     Calculate  in  this  way  the 
sums  of  1 -£+}  —  ...  and  |-£+|--... 

215.     The    Binomial    Series.      We   have   already   (§   163) 
investigated  the  Binomial  Theorem 


214,  215]  OF   A    REAL    VARIABLE  385 

assuming  that  -  1  <  x  <  1  and  that  in  is  rational.  When  m  is 
irrational  we  have 

(1  +  x)m  =  eml0«  (1+*> , 

Dx  (1  +  x)m  =  {ml {I  +  x)}  eml°s  u+«9  =  OT  (1  +  x)m-i} 

so  that  the  rule  for  the  differentiation  of  (1  +  x)m  remains  the 
same,  and  the  proof  of  the  theorem  given  in  §  163  retains  its 
validity.  We  shall  not  discuss  the  question  of  the  convergence 
of  the  series  when  x  =  1  or  x  =  —  1.* 

Examples  XCII.     1.     Prove  that  if  -  Kx<  1  then 


x/(l+.tf2)  2'     '  2.4"      -'       v'(l-^2)       ^2~T2.4" 

2.    Approximation  to  quadratic  and  other  surds.    Let  SIM  be  a 

quadratic  surd  whose  numerical  value  is  required.  Let  N2  be  the  square 
nearest  to  M ;  and  let  M=  N2  +  x  or  M=N'2  —  x,  x  being  positive.  Since  x 
cannot  be  greater  than  JY,  x/N2  is  comparatively  small  and  the  surd 
<JM=N */{l±(x/JV2)}  can  be  expressed  in  a  series 


--H(£)-H(l-.)a±4 


which  is  at  any  rate  fairly  rapidly  convergent,  and  may  be  very  rapidly  so. 
Thus 

,e7^(«4+3)=8{1+i(|4)-^(iy+..j 

Let  us  consider  the  error  committed  in  taking  8fs  (the  value  given  by 
the  first  two  terms)  as  an  approximate  value.  After  the  second  term  the 
terms  alternate  in  sign  and  decrease.  Hence  the  error  is  one  of  excess,  and 
is  less  than  32/642,  which  is  less  than  '003. 

3      If  x  is  small  compared  with  iV2  then 

the  error  being  of  the  order  xi/N7.     Apply  the  process  to  x/9^7. 
[Expanding  by  the  binomial  theorem,  we  have 

the  error  being  less    than    the    numerical  value    of   the  next  term,   viz. 
5A-4/128iV7.     Also 

Nx  x    /        x   \~1       x         x2  x3 


2(2A2  +  .r)     4N\      2N2)  AN     8JV3      WN^ 

the  error  being  less  than  xil%2N'1.     The  result  follows.     The  same  method 
may  be  applied  to  surds  other  than  quadratic  surds,  e.g.  to  ^1031.] 

*  See  Bromwich,  Infinite  Series,  pp.  150  ct  seq. ;  Hobson,  Plane  Trigonometry 
(3rd  edition),  p.  271. 

h.  25 


386  THE    LOGARITHMIC    AND    EXPONENTIAL    FUNCTIONS  [iX 

4.  If  M  differs  from  AT3  by  less  than  1  per  cent,  of  either  then  HJM  differs 
from  \N+\  (MjN2)  by  less  than  ir/90000.  {Math.  Trip.  1882.) 

5.  If  M=Ni+x,  and  x  is  small  compared  with  A7,  then  a  good  approxi- 
mation for  $M  is 

51         |I  27Afo 

56      +56  N3+WjM+bW)' 

Show  that  when  iV=10,  x=l,  this  approximation  is  accurate  to  16  places 
of  decimals.  (Math.  Trip.  1886.) 

6.  Show  how  to  sum  the  series 

2Pr(n)( 

o 

where  Pr  (n)  is  a  polynomial  of  degree  r  in  n. 

[Express  Pr{n)  in  the  form  A0  +  Aln  +  A2n(n-\)  + ...  as  in  Ex.  xc.  7.] 

00        /7)l\  °0  /7}}\ 

7.  Sum  the  series  2n  (      I  xn,  2?i2  I      )  xn  and  prove  that 

o     W         o       W 

2?i3  ()  xn={m3x3+m(Sm-l)x2  +  mx}{l+x)m-3. 
0        \nJ 

216.  An  alternative  method  of  development  of  the  theory  of  the 
exponential  and  logarithmic  functions.  We  shall  now  give  an  outline  of 
a  method  of  investigation  of  the  properties  of  ex  and  log  a;  entirely  different 
in  logical  order  from  that  followed  in  the  preceding  pages.     This  method 

starts  from  the  exponential  series  1+X  +  —  + We  know  that  this  series 

is  convergent  for  all  values  of  x,  and  we  may  therefore  define  the  function 
exp  x  by  the  equation 

x 
ex])x  =  l+x  +  —+ (1). 

We  then  prove,  as  in  Ex.  lxxxi.  7,  that 

exp#x  exp  y  =  exp  (x+y)  (2). 

exp  A—  1  h       h2  ,         ... 

Again  -£j =  1  +  2-, +  3-, +  ...  =  !+/>(/<), 

where  p  (h)  is  numerically  less  than 

I  £A  |  +  ]  U  |2+ 1  ^  j3  + ...  =  1  |A  |/(1  - 1  M  |), 
so  that  p  (h)  -^OasA->0.     And  so 

exr>(x+h)-exy>x  /expA-l\ 

=  exp.r  I  — ±— 1  -9-exp# 


h  r     V       h 

as  h  ->-  0,  or 

Dx  exp  x= exp  x (3). 

Incidentally  we  have  proved  that  exp.r  is  a  continuous  function. 

We  have  now  a  choice  of  procedure.     Writing  y  =  es.ipx  and  observim 
that  exp  0  =  1,  we  have 

dy  fvdt 


215,  216]  OF   A   REAL   VARIABLE  387 

and,  if  we  define  the  logarithmic  function  as  the  function  inverse  to  the 
exponential  function,  we  are  brought  back  to  the  point  of  view  adopted  earlier 
in  this  chapter. 

But  we  may  proceed  differently.    From  (2)  it  follows  that  if  n  is  a  positive 
integer  then 

(exp  a?)n=exp  rue,    (exp  l)"=exp  n. 

If  x  is  a  positive  rational  fraction  m/n,  then 

{exp  (m/n)}n  =  exp  m  =  (exp  l)m, 

and  so  exp  (m/n)  is  equal  to  the  positive  value  of  (exp  1  )min.  This  result  may 
be  extended  to  negative  rational  values  of  x  by  means  of  the  equation 

exp  x  exp  ( —  x)  =  1  ; 
and  so  we  have 

exp  x = (exp  l)x = ex, 

say,  where  e=exp  1  =  1  +  1  +  — + -+..., 

for  all  rational  values  of  x.  Finally  we  define  ex,  when  x  is  irrational,  as 
being  equal  to  exp#.  The  logarithm  is  then  defined  as  the  function  inverse 
to  exp  x  or  ex. 

Example.     Develop  the  theory  of  the  binomial  series 


i+(J)«+(J)  «•+...-/<«,-* 


where  -  1  <#<1,  in  a  similar  manner,  starting  from  the  equation 

f  {m,  x)f(m',  se)=f(m-*-m'  x) 
(Ex.  LXXXI.  6). 


MISCELLANEOUS   EXAMPLES   ON   CHAPTER   IX*. 

1.  Given  that  log10  e  =  "4343  and  that  210  and  Z'il  are  nearly  equal  to  powers 
of  10,  calculate  log102  and  log103  to  four  places  of  decimals. 

(Math.  Trip.  1905.) 

2.  Determine  which  of  (\e)  •  and  (\/2)-'r  is  the  greater.  [Take  logarithms 
and  observe  that  -^3/(^3 +  £tt)  <  §  JS  <  -6929  <  log  2.  J 

3.  Show  that  log10w  cannot  be  a  rational  number  if  n  is  any  positive 
integer  not  a  power  of  10.  [If  n  is  not  divisible  by  10,  and  \og10n—p/q,  we 
have  I0p  =  nq,  which  is  impossible,  since  10p  ends  with  0  and  nq  does  not. 
If  n  =  10ai\r,  where  N  is  not  divisible  by  10,  then  log10  N  and  therefore 

\ogwn=a  +  \ogwN 
cannot  be  rational.] 

*  A  considerable  number  of  these  examples  are  taken  from  Bromwich's  Infinite  Series. 

25—2 


388  THE    LOGARITHMIC    AND    EXPONENTIAL    FUNCTIONS  [iX 

4.  For  what  values  of  x  are  the  functions  log x,  log  log x,  log  log  log.r,  ... 
(a)  equal  to  0  (b)  equal  to  1  (c)  not  denned  ?  Consider  also  the  same  question 
for  the  functions  Ix,  llx,  lllx,  ...,  where  lx=log  \x\. 

5.  Show  that 

log *'-(?)  loS (*+  1)+  (g)  loS (*+2) - ...+(- 1)" log  (*+*) 

is  negative  and  increases  steadily  towards  0  as  x  increases  from  0  towards  00 . 
[The  derivative  of  the  function  is 

0^       '    \r)x~+r~x(x+l)...(x  +  n)' 
as  is  easily  seen  by  splitting  up  the  right-hand  side  into  partial  fractions. 
This  expression  is  positive,  and  the  function  itself  tends  to  zero  as  x  -*-  00 , 
since 

log(a?+r)=loga?+ea., 


where  „-~0,  and  1  -  (")  +  Q)  -  ...  =  0.] 


6.  Prove  that 

-r        -2— =  1 /—T~  (logA'-l---...-- 

<ir/       X  xn  +  1      \    °  2  w 

(Math.  Trip.  1909.) 

7.  If  x>  - 1  then  .r2>  (1  +x)  {log  (1  +x)}2.  (Math.  Trip.  1906.) 
[Put  1  +#=e*,  and  use  the  fact  that  sinh  £>  £  when  £>  0.] 

8.  Show  that  {log  (1  +x)}/x  and  #/{(l  +  x)  log  (1  +#)} both  decrease  steadily 
as  x  increases  from  0  towards  00 . 

9.  Show  that,   as    x    increases    from    - 1     towards    oo ,    the    function 
(l+x)~1/x  assumes  once  and  only  once  every  value  between  0  and  1. 

(Math.  Trip.  1910.) 

10.  Show  that  -. — r >-  -  as  x  ->-  0. 

log  (l+x)      x       2 

11.  Show  that : <  —  decreases  steadily  from  1  to  0  as  x  increases 

log  (l  +  x)     x  J 

from  —1  towards  00.  [The  function  is  undefined  when  x=0,  but  if  we 
attribute  to  it  the  value  -|  when  x  =  0  it  becomes  continuous  for  x=0.  Use 
Ex.  7  to  show  that  the  derivative  is  negative.] 

12.  Show  that   the  function  (log  £  —  log  x)j(£  —  x),  where  £  is  positive, 
decreases  steadily  as  x  increases  from  0  to  £,  and  find  its  limit  as  #-*-£. 

13.  Show  that  ex  >  Mx  ,  where  M  and  N  are  large  positive  numbers,  f 
x  is  greater  than  the  greater  of  2  log  M  and  16  iV2. 

[It  is  easy  to  prove  that  log  x<%sJx  ;   and  so  the  inequality  given  is- 
certainly  satisfied  if 

x>\ogM+2N  Jx, 

and  therefore  certainly  satisfied  if  ^x>\og  M,  %x>  2N  ^/x."] 


OF    A    REAL    VARIABLE  389 

14.  If  f(x)  and  cf>(.v)  tend  to  infinity  as  x-*-ao,  and  /'(.£)/<£' (#)-*- oo , 
thenf(x)/(f)(x)^cc.  [Use  the  result  of  Ch.  VI,  Misc.  Ex.  33.]  By  taking 
f(x)  =  xa,  (f>(x)  =  logx,  prove  that  (log  x)jxa  -*■  0  for  all  positive  values  of  a. 


15.     If  p  and  q  are  positive  integers  then 

(S) 


11  1 


pn+1      pn  +  2  qn 

as  n  -*-  oo  .     [Cf.  Ex.  lxxviii.  6.] 

16.  Prove   that  if  x  is  positive   then   n  log  {h  (I  +  x1'11)} -* — h  log  x  as 
n  ->-  oo .     [We  have 

7il0g{^(l+.r1/»)}  =  ?ll0g  {l-A(l-A-l/»)}  =  i?i(l-^/»)  l0g  (1  "  ^ 

where  w=£  (1  -  xv'1).     Now  use  §  209  and  Ex.  lxxxii.  4.] 

17.  Prove  that  if  a  and  b  are  positive  then 

{£(aV»+&V~)}»-e^(a&). 

[Take  logarithms  and  use  Ex.  16.] 

18.  Show  that 

l  +  l  +  l  +  ...  +  ^-i  =  ilogn  +  \og2  +  ly  +  en, 
where  y  is  Euler's  constant  (Ex.  lxxxix.  1)  and  e n  ^-  0  as  n  -*-  x . 

19.  Show  that 

the  series  being  formed  from  the  series  1  —  \  +  A-  —  . . .  by  taking  alternately  two 
positive  terms  and  then  one  negative.     [The  sum  of  the  first  'in  terms  is 


,11  1  1/.     1  1 

3      5  An  —  1      2  \       2  n 


=  £log2H  +  log2  +  £y  +  fn-$(log»  +  y  +  en'), 
where  e„  and  en'  tend  toOasm->co.     (Cf.  Ex.  lxxviii.  6).] 

20.  Show  that  l-^-|+i-i-i  +  i-Jo-...=|log2. 

21.  Prove  that 

n  i 


tv(36.>2-l) 


=  ~  3  +  3  23n  + 1  —  2n  -  Sn 


where  SH=l+~  +  ...  +  -,  2„=l  +  =  +  ,..  +  = r.     Hence  prove  that  the  sum 

of  the  series  when  continued  to  infinity  is 

-  3 +§  log  3  +  2  log  2.  {Math.  Trip.  1905.) 

22.     Show  that 

S     -,\     ,x=2log2-l,     I      /f>  *      ,    =  a  (log 3-1). 


390  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [IX 

23.     Prove  that  the  sums  of  the  four  series 

co          1  oc  (_l)n-l         »  l  oo       (_1>-1 

2\   J  ,    ,    2,- —a — r,    2 


,4b2-1'     i   4w2-l    '    i(2»+l)a-l'     1(2»+1)2-1 

are  \,  \ir  - |,  j,  \  log  2  -£  respectively. 

24.  Prove  that  « !  {ajn)n  tends  to  0  or  to  oo  according  as  a  <  e  or  a  >  e. 

[If  un  =  n\{ajn)n  then  Mn  +  1/w„  =  a{l  +  (l/tt)}-n-»-a/e.  It  can  be  shown 
that  the  function  tends  to  oo  when  a  =  e:  for  a  proof,  which  is  rather  beyond 
the  scope  of  the  theorems  of  this  chapter,  see  Bromwich's  Infinite  Series, 
pp.  461  et  seq.] 

25.  Find  the  limit  as  x-*-  oo  of 

/a0+a1x  +  ...  +  a,.xr\Xo+KlX 


V>o  +  ^i x  +  •  •  •  +  hrxr  / 
distinguishing  the  different  cases  which  may  arise.  {Math.  Trip.  1886.) 

26.  Prove  that 

Slog(l+!)  (*>0) 

diverges  to  oo .     [Compare  with  2  (x/n).]     Deduce  that  if  x  is  positive  then 
(1  +x)  (2+.r) ...  (n+x)/n  !  -*  oo 

as  «->■  oo  .     [The  logarithm  of  the  function  is  2  log  ( 1  +  '- ).] 

27.  Prove  that  if  x  >  -  1  then 

1  1  1! 

+ 


(x+1)2      {x+l){x  +  2)^  {x+l){x  +  2){x+Z) 

2! 


(x+1)  {x  +  2)  (x  +  S)  (ff+4)  ^     ' 

{Math,  Trip.  1908.) 
[The  difference  between  l/(.r  +  l)2  and  the  sum  of  the  first  n  terms  of  the 
series  is 

1 n\ , 

{x+l)2  {x+2){x  +  3)  ,..{x  +  n  +  l)  'J 

28.  No  equation  of  the  type 

Aeax  +  Befix+...  =  0, 
where  A,  B,  ...  are  polynomials  and  a,  /3, ...  different  real  numbers,  can  hold 
for  all  values  of  x.    [If  a  is  the  algebraically  greatest  of  a,  /3,  . . . ,  then  the  term 
Aeax  outweighs  all  the  rest  as  .r->-  oo  .] 

29.  Show  that  the  sequence 

a1  =  e,    a2=e  ,     «3=ec  ,  ... 
tends  to  infinity  more  rapidly  than  any  member  of  the  exponential  scale. 

[Let  ex  {x)  —  ex,  e<!{x)  =  eeiW,  and  so  on.    Then,  if  ek {x)  is  any  member  of  the 
exponential  scale,  an>  ek{n)  when  n  >  k.~\ 


OF    A    REAL   VARIABLE  391 

30.  Prove  that 

where  a  is  to  be  put  equal  to  ^  (x)  and   /3  to  <£  (.r)  after  differentiation. 
Establish  a  similar  rule  for  the  differentiation  of  0  (j;)Lir  (x)r      J_ 

31.  Prove  that  if  Bxne~x2  =  e-x2(fxn(x)  then  (i)  <£„(#)  is  a  polynomial  of 
degree  to,  (ii)  <£„  +  i=  -2.r  $„+$„',  and  (iii)  all  the  roots  of  0„=O  are  real  and 
distinct,  and  separated  by  those  of  cj)n_1  =  0.  [To  prove  (iii)  assume  the  truth 
of  the  result  for  to=  1,  2,  ...  k,  and  consider  the  signs  of  <£K  ,  1  for  the  n  values 
of  x  for  which  4>K  =  0  and  for  large  (positive  or  negative)  values  of  x.] 

32.  The  general  solution  of  f(xy)=f(x)f(y),  where  /  is  a  differentiable 
function,  is  xa,  where  a  is  a  constant :  and  that  of 

f{x+y)+f{x-y)=2f{x)f{y) 
is  cosh  ax  or  cosaa',  according  as/"(0)  is  positive  or  negative.     [In  proving 
the  second  result  assume  that  /  has  derivatives  of  the  first  three  orders. 
Then 

»/(*)  +f  if"  (*)  +  '  v)  =  2/W  [/  (°)  +nf  (°)  +  if  if"  (0)  +  *  /}] , 
where  ey  and  *y'  tend  to  zero  with  y.      It  follows  that  /(0)  =  1,  /'(0)  =  0, 
f"(x)=f"(0)f(x),  so  that  a  =  v/{/"(0)}  or  a=J{-f"(0)}.] 

33.  How  do  the  functions  xs'm  (l/x\  xsin*  im,  xcos™^  behave  as  x+  +  0  ? 

34.  Trace  the  curves  #  =  tana; etan,,;.  y  =  sin ^ log  tan  £#. 

35.  The  equation  ex=ax  +  b  has  one  real  root  if  a<0  or  a=0,  6>0.  If 
a  >  0  then  it  has  two  real  roots  or  none,  according  as  a  log  a  >b  —  a  or 
a  log  a  <  b  —  a. 

36.  Show  by  graphical  considerations  that  the  equation  ex  =  ax2  +  2bx+c 
has  one,  two,  or  three  real  roots  if  a  >  0,  none,  one,  or  two  if  a  <0;  and  show 
how  to  distinguish  between  the  different  cases. 

37.  Trace  the  curve  y  =  -  log  (  — —  J ,  showing  that  the  point  (0,  \)  is 

a  centre  of  symmetry,  and  that  as  x  increases  through  all  real  values,  y 
steadily  increases  from  0  to  1.     Deduce  that  the  equation 

1.      /e*-l\ 

-xloA —)=a 

has  no  real  root  unless  0<o<l,  and  then  one,  whose  sign  is  the  same  as 
that  of  a-i-     [In  the  first  place 

is  clearly  an  odd  function  of  x.     Also 


392  THE    LOGARITHMIC    AND   EXPONENTIAL    FUNCTIONS  [iX 

The  function  inside   the  large   bracket  tends  to  zero  as  x-*-0;    and  its 
derivative  is 


x  \      \sinh  %x)  J  ' 


which  has  the  sign  of  x.     Hence  di//dx>  0  for  all  values  of  x.~\ 

3S.     Trace  the  curve  i/  —  e^x *J(x2+2x),  and  show  that  the  equation 

eVxs/(x2  +  2x)  =  a 

has  no  real  roots  if  nis  negative,  one  negative  root  if  0<a<a  =  e1y'v'2»/(2  +  2v/2), 
and  two  positive  roots  and  one  negative  if  a  >  a. 

39.  Show  that  the  equation  fn(x)  =  l  +x  +  —  +  ...  +  — -=0  has  one  real 

2!  n\ 

root  if  n  is  odd  and  none  if  n  is  even. 

[Assume  this  proved  for  n  =  \,  2,  ...  2k.  Then  /2ft  + 1  (#)  =  0  has  at  least 
one  real  root,  since  its  degree  is  odd,  and  it  cannot  have  more  since,  if  it 
had,  fik  + 1  (x)  or/2i.  (x)  would  have  to  vanish  once  at  least.  Hence  /2jt  + 1  (x)  =  0 
has  just  one  root,  and  so  fu  +  2  {x)  =  0  cannot  have  more  than  two.  If  it  has 
two,  say  a  and  /3,  then  /'2J.  +  2  (x)  or/at+i  (x)  must  vanish  once  at  least  between 
a  and  /3,  say  at  y.     And 

/a + 2  (y)  =U  + 1  (y)  +  p^gji  >  °- 

But  f%c+2(x)  is  also  positive  when  x  is  large  (positively  or  negatively),  and 
a  glance  at  a  figure  will  show  that  these  results  are  contradictory.  Hence 
/2t  +  2(.v)=0  has  no  real  roots.] 

40.  Prove  that  if  a  and  b  are  positive  and  nearly  equal  then 

.      a      1  ,       LN  /l      P 

approximately,  the  error  being  about  ${(a  —  b)/a}3.  [Use  the  logarithmic 
series.  This  formula  is  interesting  historically  as  having  been  employed  by 
Napier  for  the  numerical  calculation  of  logarithms.] 

41.  Prove  by  multiplication  of  series  that  if  —  1  <  x  <  1  then 

A{iog(i+.r)}2=iA'2--Hi+*)^+J(i+Hi)^---. 

|(arctan.r)2  =  *.i'2-i(l  +  i)^  +  Kl  +  -J  +  ^)^--- 

42.  Prove  that 

(1  +axfx  =  ea  {1  -^n2.r  + Jj  (8  +  3a)  a*x°-  (1  +  f*)}, 

where  nx  ->-  0  with  x. 

/  x2  xn \ 

43.  The  first  n  +  2  terms  in  the  expansion  of  log  f  1  +  x  +  '—  +  . . .  -\ -J  in 

powers  of  x  are 

X       n\     [n  +  1      l!(»  +  2)+2!(n  +  3)         XV       Jnl(2n  +  l)j 

(Math.  Trip.  1899.) 


OF    A   REAL   VARIABLE  393 

44.  Show  that  the  expansion  of 

in  powers  of  x  begins  with  the  terms 

1  -x  + T  -  2  ; ^— .     (Math.  Trip.  1909.) 

n  +  1     s=i(»+s)(n+s+l)       v  e 

45.  Show  that  if  —  1  <x  <  1  then 

1       ,  ]  -402    2  ,  !-4-"o9    ,  ,  x(x  +  3) 

lv+Li23^,Lli_7  *(*J+ite+o) 

3i  +  3.624  +3.6.9d^  +-_    27(l-x)10/3    " 

[Use  the  method  of  Ex.  xcn.  6.  The  results  are  more  easily  obtained  by 
differentiation;  but  the  problem  of  the  differentiation  of  an  infinite  series  is 
beyond  our  range.] 

dx 


46.     Prove  that      P° ff =  _2_  W  (f\ , 

Jo  (x+a)(x  +  b)     a-b     °  W 


0  (.r  +  a^  +  ^^V*  f-6-6  l0g  (!)}' 


xaA-  1        f    ,       /a\  ,  I 

0  ixT^K^w^{a^w\al0S{b)-a+br 

J  0  (x+a) («*+&)  =  {a*+V)b  \^a-b  l0g (!)/  ' 

/o  (^4^^)  =  ^{^6  +  al0g(?)}' 
provided  that  a  and  b  are  positive.     Deduce,  and  verify  independently,  that 
each  of  the  functions 

a  — 1  — log  a,  aloga  — a-fl,  ina  —  log  a,  itr+aloga 
is  positive  for  all  positive  values  of  a. 

47.     Prove  that  if  a,  /3,  y  are  all  positive,  and  /32>  ay,  then 
/•-  dx         =         1  ^  +  s/(^_ay)^  _ 

J0ax*  +  2!3x  +  y      J  (ft* -ay)   °°  \         J  {ay)         J' 
while  if  a  is  positive  and  ay>j82  the  value  of  the  integral  is 

1 — -T-  arc  tan  J v  ^ay~     ' 


that  value  of  the  inverse  tangent  being  chosen  which  lies  between  0  and  «■• 
Are  there  any  other  really  different  cases  in  which  the  integral  is  convergent? 

48.     Prove  that  if  a  >  -  1  then 

du 


r      dx  =  r dt   =2 

J  i  (x  +  a)s,f{x"  —  l)     J0  coslW+a 


i  v?  +  2au-tl  ' 


394  THE   LOGARITHMIC   AND   EXPONENTIAL   FUNCTIONS  [iX 

al  is 


and  deduce  that  the  value  of  the  integral  is 

2 


—  arc  ta, 
if  — l<a<l,  and 

if  a  >  1.     Discuss  the  case  in  which  a  =  1. 

49.     Transform  the  integral   I     . .    ..  „ — — . ,  where  a>0,  in  the  same 

Jo  (x+a)s!(x-  +  l) 

ways,  showing  that  its  value  is 

x7^TT) log  c^WCa2^!)  =  ^mT)  ar§ tanh  ~5+r 


50. 


Prove  that  I   arc  tan  xdx= \tt  -  \  log  2. 

jo 


51.     If  0<a<l,  0</3<l,  then 


/' 


rf.r 


1       .       1  +  J(a8) 


.lv/{(l-2a^  +  aa)(l-2/3^  +  32)}       ^(a/S)      &l-^(a/j)' 

52.  Prove  that  if  a>  b>  0  then 

J  _  a,  a  cosh  0  +  &  sinh  0  ~  sJ{ai  -  62) 

53.  Prove  that 


and  deduce  that  if  a>0  then 

losr.r 


Jo  «2 


dx.—  —  log  a. 


[Use  the  substitutions  x=\jt  and  .»=£ra.] 


54.     Prove  that  I     log  ( 1  +  —  )  dx  =  na  if  a>0.     [Integrate  by  parts.] 


CHAPTER  X 

THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,  EXPONENTIAL, 
AND  CIRCULAR  FUNCTIONS 

217.     Functions   of  a  complex  variable.     In  Ch.  Ill  we 

defined  the  complex  variable 

z  —  x  +  iy  *, 

and  we  considered  a  few  simple  properties  of  some  classes  of 
expressions  involving  z,  such  as  the  polynomial  P(z).  It  is 
natural  to  describe  such  expressions  as  functions  of  z,  and  in 
fact  we  did  describe  the  quotient  P  (z)/Q  (z),  where  P  (z)  and  Q  (z) 
are  polynomials,  as  a  '  rational  function  '.  We  have  however  given 
no  general  definition  of  what  is  meant  by  a  function  of  z. 

It  might  seem  natural  to  define  a  function  of  z  in  the  same 
way  as  that  in  which  we  defined  a  function  of  the  real  variable 
x,  i.e.  to  say  that  Z  is  a  function  of  z  if  any  relation  subsists 
between  z  and  Z  in  virtue  of  which  a  value  or  values  of  Z  corre- 
sponds to  some  or  all  values  of  z.  But  it  will  be  found,  on  closer 
examination,  that  this  definition  is  not  one  from  which  any  profit 
can  be  derived.  For  if  z  is  given,  so  are  x  and  y,  and  conversely  : 
to  assign  a  value  of  z  is  precisely  the  same  thing  as  to  assign  a 
pair  of  values  of  x  and  y.  Thus  a  '  function  of  z ',  according  to 
the  definition  suggested,  is  precisely  the  same  thing  as  a  complex 
function 

f  (^  y)  +  *9  (k  y), 
of  the  two  real  variables  x  and  y.     For  example 

x  —  iy,     xy,     \z\  =  ^(x2  +  y2),     am  z  =  arc  tan  (y/x) 
are  'functions  of  z\     The  definition,  although  perfectly  legitimate, 

*  In  this  chapter  we  shall  generally  find  it  convenient  to  write  x  +  iy  rather 
than  x  +  yi. 


396      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [X 

is  futile  because  it  does  not  really  define  a  new  idea  at  all.  It  is 
therefore  more  convenient  to  use  the  expression  '  function  of  the 
complex  variable  z '  in  a  more  restricted  sense,  or  in  other  words 
to  pick  out,  from  the  general  class  of  complex  functions  of  the 
two  real  variables  x  and  y,  a  special  class  to  which  the  expression 
shall  be  restricted.  But  if  we  were  to  attempt  to  explain  how 
this  selection  is  made,  and  what  are  the  characteristic  properties 
of  the  special  class  of  functions  selected,  we  should  be  led  far 
beyond  the  limits  of  this  book.  We  shall  therefore  not  attempt 
to  give  any  general  definitions,  but  shall  confine  ourselves  entirely 
to  special  functions  defined  directly. 

218.  We  have  already  defined  polynomials  in  z  (§  39), 
rational  functions  of  z  (§  46),  and  roots  of  z  (§  47).  There  is 
no  difficulty  in  extending  to  the  complex  variable  the  definitions 
of  algebraical  functions,  explicit  and  implicit,  which  we  gave 
'(§§  26 — 27)  in  the  case  of  the  real  variable  x.  In  all  these  cases 
we  shall  call  the  complex  number  z,  the  argument  (§  44)  of  the 
point  z,  the  argument  of  the  function  f  (z)  under  consideration. 
The  question  which  will  occupy  us  in  this  chapter  is  that  of  defining 
and  determining  the  principal  properties  of  the  logarithmic,  ex- 
ponential, and  trigonometrical  or  circular  functions  of  z.  These 
functions  are  of  course  so  far  defined  for  real  values  of  z  only,  the 
logarithm  indeed  for  positive  values  only. 

We  shall  begin  with  the  logarithmic  function.  It  is  natural 
to  attempt  to  define  it  by  means  of  some  extension  of  the  definition 


log  « =  J    j         (x>0): 


and  in  order  to  do  this  we  shall  find  it  necessary  to  consider 
briefly  some  extensions  of  the  notion  of  an  integral. 

219.     Real   and  complex  curvilinear  integrals.     Let  AB 

be  an  arc  G  of  a  curve  defined  by  the  equations 

where  <£  and  yfr  are  functions  of  t  with  continuous  differential 
coefficients  </>'  and  yfr' ;  and  suppose  that,  as  t  varies  from  t0  to  t1} 
the  point  (x,  y)  moves  along  the  curve,  in  the  same  direction,  from 
A  to  B. 


/. 


217-220]        EXPONENTIAL,    AND   CIRCULAR   FUNCTIONS  397 

Then  we  define  che  curvilinear  integral 

{g(x,y)dx  +  h(x,y)dy} (1), 

c 

where  g  and  h  are  continuous  functions  of  x  and  y,  as  being  equi- 
valent to  the  ordinary  integral  obtained  by  effecting  the  formal 
substitutions  x  —  <p  (t),  y  =  ^r  (t),  i.e.  to 

I ' {gifrWV  +h(<f>)ir)y}dt. 

We  call  C  the  path  of  integration. 

Let  us  suppose  now  that 

z  =  x  +  iy  =  <f)  (t)  +  i-fy  (t), 

so  that  z  describes  the  curve  C  in  Argand's  diagram  as  t  varies. 
Further  let  us  suppose  that 

f{z)  —  u  +  iv 

is  a  polynomial  in  z  or  rational  function  of  z. 

Then  we  define 


L 


fi*)d* (2) 

J  V 

as  meaning 

I    (u  +  iv)  (dx  +  idy), 
J  c 

which  is  itself  defined  as  meaning 


L 


(udx—  vdy)  +  i\    (v  dx  +  u  dy). 
J  c 


220.  The  definition  of_Log.  £  Now  let  £=  f  +  it;  be  any- 
complex  number.  We  define  Log  £,  the  general  logarithm  of  £ 
by  the  equation 

dz 


Log£=( 


Jc  z 

where  C  is  a  curve  which  starts  from  1  and  ends  at  £  and  does 
not  pass  through  the  origin.  Thus  (Fig.  54)  the  paths  (a),  (b),  (c) 
are  paths  such  as  are  contemplated  in  the  definition.  The  value 
of  Log  z  is  thus  defined  when  the  particular  path  of  integration 
has  been  chosen.  But  at  present  it  is  not  clear  how  far  the  value 
of  Log  z  resulting  from  the  definition  depends  upon  what  path  is 
chosen.      Suppose  for  example  that  £  is  real  and  positive,   say 


398 


THE  GENERAL  THEORY  OF  THE  LOGARITHMIC, 


Lx 


equal  to  f.     Then  one  possible  path  of  integration  is  the  straight 
line  from  1  to  £,  a  path  which  we  may  suppose  to  be  defined  by 

Y 


Pig.  54. 


the  equations  x  =  t,  y  =  0.     In  this  case,  and  with  this  particular 
choice  of  the  path  of  integration,  we  have 


Los: 


*-/:* 


so  that  Log  f  is  equal  to  log  £,  the  logarithm  of  £  according  to  the 
definition  given  in  the  last  chapter.  Thus  one  value  at  any  rate 
of  Log  £,  when  £  is  real  and  positive,  is  log  £.  But  in  this  case,  as 
in  the  general  case,  the  path  of  integration  can  be  chosen  in  an 
infinite  variety  of  different  ways.  There  is  nothing  to  show  that 
every  value  of  Log  £  is  equal  to  log  £;  and  in  point  of  fact  we 
shall  see  that  this  is  not  the  case.  This  is  why  we  have  adopted 
the  notation  Log  £,  Log  £  instead  of  log  £",  log  f.  Log  £  is  (possibly 
at  any  rate)  a  many  valued  function,  and  log  f  is  only  one  of  its 
values.  And  in  the  general  case,  so  far  as  we  can  see  at  present, 
three  alternatives  are  equally  possible,  viz.  that 

(1)     we  may  always  get  the  same  value  of  Log  f,  by  whatever 

path'  we  go  from  1  to  £ ; 
we    may   get    a   different    value    corresponding  to  every 

different  path  ; 
we  may  get  a  number  of  different  values  each  of  which 

corresponds  to  a  whole  class  of  paths  : 

and  the  truth  or  falsehood  of  any  one  of  these  alternatives  is  in 
no  way  implied  by  our  definition. 


(2) 


(3) 


220,  221J        EXPONENTIAL,    AND   CIRCULAR   FUNCTIONS 


399 


221.     The  values  of  Log  £     Let  us  suppose  that  the  polar 
coordinates  of  the  point  z  =  £  are  p  and  </>,  so  that 

£=  p  (cos  <f>  +  i  sin  <£). 
We  suppose  for  the  present  that  —  ir  <  <f>  <  ir,  while  p  may  have 
any  positive  value.     Thus  £  may  have  any  value  other  than  zero 
or  a  real  negative  value. 

The  coordinates  (*,  y)  of  any  point  on  the  path  C  are  functions 
of  t,  and  so  also  are  its  polar  coordinates  (r,  6).     Also 


w^f  *-[  dx+idy 

Jo  z      Jc    x  +  iy 


=f-M 


dx      .  dy 
dt         dt 


~)dt, 


t„  x  +  ly 
in  virtue  of  the  definitions  of  §  219.    But  x  =  r  cos  6,y  —  r  sin  6,  and 


dx      .dy 

di  +  ldi 


adr  .       dd\       .  (  .     n  d? 

cos  a  -r  —  r  sin  6  -j-  J  -f  %  f  sin  6 


Qd6 

~dt+rcos6-dt 


/       n  ,    •    •    /ix  fdr      .    dd 

=(co*e+ism0Kdt+irdt 


so  that 


dt  +  i       ~j,dt  =  [log  r]  +  i  [0], 


°g  ?~Je„  r  eft  ""  '  "./«,  c& 
where  [log  r]  denotes  the  difference  between  the  values  of  log  r  at 
the  points  corresponding  to  t  =  ^  and  t  =  t0,  and  [#]  has  a  similar 
meaning. 

It  is  clear  that 

[log  r]  =  log  p  -  log  1  =  log  p ; 
but  the  value  of  [6]  requires  a  little  more  consideration.     Let  us 
suppose  first  that  the  path  of  integration  is  the  straight  line  from 
1   to   £     The  initial  value  of  6  is  the  amplitude  of  1,  or  rather 
one  of  the  amplitudes   of  1,  viz. 
Ikir,  where  k  is  any  integer.     Let 
us  suppose  that  initially  6  =  2kir. 
It  is  evident  from  the  figure  that 
6  increases  from  2kir  to  Ikir  +  <j> 
as  t  moves  along  the  line.     Thus 

[0]=(2Jbr+<£)-2fc7r=0, 
and,  when  the  path  of  integration 
is  a  straight  line,  Log  £  =  log  p  +  i<f). 


Fig.  55. 


400 


THE  GENERAL  THEORY  OF  THE  LOGARITHMIC, 


[* 


We  shall   call   this  particular  value   of  Log£  the  principal 

value.     When  £  is  real  and  positive,  £=  p  and  9  =  0,  so  that  the 

principal  value  of  Log  £  is  the  ordinary  logarithm  log  £.     Hence  it 

will   be  convenient  in  general   to  denote  the  principal  value  of 

Log  ?  by  log  £     Thus 

}ogZ  =  \ogp  +  i<f>, 

and    the   principal    value  is    characterised    by  the    fact   that   its 
imaginary  part  lies  between  —  it  and  it. 

Next  let  us  consider  any  path  (such  as  those  shown  in  Fig.  56) 
such  that  the  area  or  areas  included 
between  the  path  and  the  straight 
line  from  1  to  £  does  not  include 
the  origin.  It  is  easy  to  see  that 
[0]  is  still  equal  to  9.  Along  the 
curve  shown  in  the  figure  by  a 
continuous  line,  for  example,  0, 
initially  equal  to  2&tt,  first  de- 
creases to  the  value 

2ICTT-X0P  Fi8-56- 

and  then  increases  again,  being  equal  to  2kir  at  Q,  and  finally 
to  Ikir  +  9.  The  dotted  curve  shows  a  similar  but  slightly  more 
complicated  case  in  which  the  straight  line  and  the  curve  bound 
two  areas,  neither  of  which  includes  the  origin.  Thus  if  the  path 
of  integration  is  such  that  the  closed  curve  formed  by  it  and  the 
line  f rum  1  to  £  does  not  include  the  origin,  then 
Log  £=  log  £= log  p  +  i(f>. 

On  the  other  hand  it  is  easy 
to  construct  paths  of  integration 
such  that  [0]  is  not  equal  to  </>. 
Consider,  for  example,  the  curve 
indicated  by  a  continuous  line  in 
Fig.  57.  If  0  is  initially  equal 
to  2/b7r,  it  will  have  increased 
by  2tt  when  we  get  to  P  and 
by  4-7T  when  we  get  to  Q;  and  its 
final  value  will  be  2&7r  +  4nr  +  cp, 
so  that  [0]  =  47r  +  9  and  Fig.  57. 

Log  £  =  log  p  +  *  (47T  +  9). 


221]  EXPONENTIAL,   AND   CIRCULAR   FUNCTIONS  401 

In  this  case  the  path  of  integration  winds  twice  round  the 
origin  in  the  positive  sense.  If  we  had  taken  a  path  winding 
k  times  round  the  origin  we  should  have  found,  in  a  precisely 
similar  way,  that  [0]  =  2kir  +  </>  and 

Log  £=  log  p  +  i  (2k  jt  +  (/>). 

Here  k  is  positive.  By  making  the  path  wind  round  the  origin 
in  the  opposite  direction  (as  shown  in  the  dotted  path  in  Fig.  57), 
we  obtain  a  similar  series  of  values  in  which  k  is  negative. 
Since  |  £  |  =  p,  and  the  different  angles  2krr  +  (f>  are  the  different 
values  of  am  f,  we  conclude  that  every  value  of  log  |  £  |  +  i  am  £  is 
a  value  of  Log£;  and  it  is  clear  from  the  preceding  discussion 
that  every  value  of  Log  £  must  be  of  this  form. 

We  may  summarise  our  conclusions  as  follows :  the  general 
value  of  Log  £  is 

log  \£\+i  am  f  =  log  p  +  i  (2&jt  +  <f>), 

where  k  is  any  positive  or  negative  integer.  The  value  of  k  is 
determined  by  the  path  of  integration  chosen.  If  this  path  is  a 
straight  line  then  k  =  0  and 

Log  f  =  log  f  =  log  p  +  i<p. 

In  what  precedes  we  have  used  £  to  denote  the  argument  of 
the  function  Log  £,  and  (£,  77)  or  (p,  <£)  to  denote  the  coordinates  of 
£ ;  and  z,  (%,  y),  (r,  6)  to  denote  an  arbitrary  point  on  the  path  of 
integration  and  its  coordinates.  There  is  however  no  reason  now 
why  we  should  not  revert  to  the  natural  notation  in  which  z  is  used 
as  the  argument  of  the  function  Log  z,  and  we  shall  do  this  in 
the  following  examples. 

Examples  XCTII.  1.  We  supposed  above  that  —  n<6<n,  and  so 
excluded  the  case  in  which  z  is  real  and  negative.  In  this  case  the  straight 
line  from  1  to  z  passes  through  0,  and  is  therefore  not  admissible  as  a  path  of 
integration.  Both  n  and  —  n  are  values  of  am  z,  and  6  is  equal  to  one  or 
other  of  them:  also  r=—z.  The  values  of  Logz  are  still  the  values  of 
log|s|+iam2,  viz. 

\og(-z)  +  (2k  +  l)7n\ 

where  h  is  an  integer.  The  values  log  ( -  z)  +  ni  and  log  ( -  z)  -  iri  correspond 
to  paths  from  1  to  z  lying  respectively  entirely  above  and  entirely  below  the 
real  axis.  Either  of  them  may  be  taken  as  the  principal  value  of  Logz,  as 
convenience  dictates.  We  shall  choose  the  value  log  ( -  z)  +  ni  corresponding 
to  the  first  path. 

h.  26 


402      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [X 

2.  The  real  and  imaginary  parts  of  any  value  of  Logs  are  both  continuous 
functions  of  x  and  y,  except  for  x  —  0,  y  =  0. 

3.  The  functional  equation  satisfied  by  Logs.  The  function  Logs 
satisfies  the  equation 

Lbg«r1*a=Log21+Logai  (1), 

in  the  sense  that  every  value  of  either  side  of  this  equation  is  one  of  the  values 
of  the  other  side.     This  follows  at  once  by  putting 

zl  =  ri  (cos  di  +  i  sin  61),  z2  =  r2  (cos  82  +  * sin  ^2)) 
and  applying  the  formula  of  p.  401.     It  is  however  not  true  that 

log«1z2=logz1+16g«2    (2) 

in  all  circumstances.    If,  e.g., 

21  =  22  =  i  (-  1  + 1\/3)  =  cos  fjr+isin  k, 

then  log21  =  log22  =  |7ri,  and  \ogzl  +  \ogZ2  =  §ni,  which  is  one  of  the  values  of 
Log zlz2,  but  not  the  principal  value.     In  fact  log  2X22  =  —  § ni. 

An  equation  such  as  (1),  in  which  every  value  of  either  side  is  a  value 
of  the  other,  we  shall  call  a  complete  equation,  or  an  equation  which  is 
completely  true. 

4.  The  equation  Log2m=»iLog2,  where  m  is  an  integer,  is  not  completely 
true :  every  value  of  the  right-hand  side  is  a  value  of  the  left-hand  side,  but 
the  converse  is  not  true. 

5.  The  equation  Log  (1/2)= -Logs  is  completely  true.  It  is  also  true 
that  log (1/2)=  —log 2,  except  when  2  is  real  and  negative. 

6.  The  equation 

lo§  ( JZb)  =  loS  (*-  a)  -  log  («  -  b) 

is  true  if  2  lies  outside  the  region  bounded  by  the  line  joining  the  points  z  =  a, 
z=b,  and  lines  through  these  points  parallel  to  OX  and  extending  to  infinity 
in  the  negative  direction. 


iog(r!Wogfi-?Wog(i 


7.  The  equation 

'"[b-zJ—*\*--Z/ 
is  true  if  2  lies  outside  the  triangle  formed  by  the  three  points  0,  a,  b. 

8.  Draw  the  graph  of  the  function  I  (Logs-)  of  the  real  variable  x.  [The 
graph  consists  of  the  positive  halves  of  the  lines  y  =  2kir  and  the  negative 
halves  of  the  lines  y=(2k+l)  n.] 

9.  The  function /(.f)  of  the  real  variable  x,  denned  by 

«■/ (*)  =p»r + (q- p)  I  (log  *), 

is  equal  to  p  when  x  is  positive  and  to  q  when  a?-is  negative. 


221,  222]        EXPONENTIAL,   AND   CIRCULAR   FUNCTIONS  403 

10.  The  function /(#)  defined  by 

■/(*)  =prr+(q-p)l  {log {x -l)}+(r-q)l  (log x) 
is  equal  to  p  when  x>\,  to  q  when  0<.r<l,  and  to  r  when  x<0. 

11.  For  what  values  of  z  is  (i)  log 2  (ii)  any  value  of  Logs  (a)  real  or 
(b)  purely  imaginary  ? 

12.  If  z = a;  +  ly  then  Log  Log  z = log  7?  +  i  (9 + 2/fc'jr),  where 

i?2  =  (logr)2  +  (^  +  2y{;7r)2 
and  0  is  the  least  positive  angle  determined  by  the  equations 

cos  9  :  sin  0  : 1 :  :  log  r:6  +  2kir  :  >/{(log  rf  +  (d  +  2£ir)*}. 
Plot  roughly  the  doubly  infinite  set  of  values  of  Log  Log  (l  +  i*v/3),  indicating 
which  of  them  are  values  of  log  Log(l  +  i\/3)  and  which  of  Loglog(l  +  iv'3). 

222.  The  exponential  function.  In  Ch.  IX  we  denned 
a  function  ey  of  the  real  variable  y  as  the  inverse  of  the  function 
y  =  log  x.  It  is  naturally  suggested  that  we  should  define  a  function 
of  the  complex  variable  z  which  is  the  inverse  of  the  function 
Logs. 

Definition.  If  any  value  of  Log  z  is  equal  to  £  we  call  z  the 
exponential  of  £  and  lurite 

z  =  exp  £". 

Thus  2  =  exp  £  if  %=~Logz.  It  is  certain  that  to  any  given 
value  of  z  correspond  infinitely  many  different  values  of  £".  It 
would  not  be  unnatural  to  suppose  that,  conversely,  to  any  given 
value  of  £  correspond  infinitely  many  values  of  z,  or  in  other  words 
that  exp  £  is  an  infinitely  many-valued  function  of  £  This  is 
however  not  the  case,  as  is  proved  by  the  following  theorem. 

Theorem.      The  exponential  function   exp  £  is  a   one-valued 
function  of  £ 

For  suppose  that 

zx  =  i\  (cos  #i  +  i  sin  0X),  z2  =  r2  (cos  0.,  +  i  sin  02) 

are  both  values  of  exp  £     Then 

£=  Log  4  =  Log  4, 

and  so  log  i\  +  i  (0t  +  2»i7r)  =  log  r2  +  i  (02  +  2mr), 

where  m  and  n  are  integers.     This  involves 

log  1\  =  log  ?',,     0X  +  2mrr  =  02  +  2mr. 

Thus  7",  =  r2,  and  0X  and  02  differ  by  a  multiple  of  2tt.      Hence 

Z\  =  Z2' 

26—2 


404      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [X 

Corollary.  If  £is  real  then  exp  £=ef,  the  real  exponential 
function  of  £  defined  in  Ch.  IX. 

For  if  z  =  e^  then  logz  =  £,  i.e.  one  of  the  values  of  Log,?  is  f, 
Hence  z  =  exp  £ 

223.  The  value  of  exp  £     Let  f  =  £  +  ^  and 

z  =  exp  £  =  r  (cos  0  +  i  sin  0). 

Then  £  +  iw  =  Log  £  =  log  r  +  t  (0  +  2mir), 

where  m  is  an  integer.     Hence  £  =  log  r,  w  =  0  +  2mir,  or 

r  =  ef,    0  =  7]  —  2/mt  ; 
and  accordingly 

exp  (£  + 1^)  =  e^  (cos  rj  +  i  sin  ?;). 

If  »7  =  0  then  exp  £  =  ef,  as  we  have  already  inferred  in  §  222. 
It  is  clear  that  both  the  real  and  the  imaginary  parts  of  exp  (£  +  irj) 
are  continuous  functions  of  f  and  rj  for  all  values  of  £  and  ?;. 

224.  The   functional    equation  satisfied  by  exp  f,     Let 

gi  =  1 1  +  Hfc ,  §"a  =  f 2  + 1% .     Then 

exp  £  x  exp  £2  =  e&  (cos  ?;x  +  i  sin  97,)  x  e&  (cos  77,  +  i  sin  %) 

=  efi+f*  {cos  (77,  +  ?72)  + 1  sin  (%  +  ??2)} 

=  exp  (£+£,). 

The  exponential  function  therefore  satisfies  the  functional  relation 
/(S'i  +  Sb)=/(?i)  /(&),  an  equation  which  we  have  proved  already 
(§  205)  to  be  true  for  real  values  of  &  and  fa. 

225.  The  general  power  ai  It  might  seem  natural,  as 
exp  £=e*  when  £  is  real,  to  adopt  the  same  notation  when  £is 
complex  and  to  drop  the  notation  exp  £  altogether.  We  shall  not 
follow  this  course  because  we  shall  have  to  give  a  more  general 
definition  of  the  meaning  of  the  symbol  e$\  we  shall  find  then 
that  e$  represents  a  function  with  infinitely  many  values  of  which 
exp  £  is  only  one. 

We  have  already  defined  the  meaning  of  the  symbol  af  in  a 
considerable  variety  of  cases.  It  is  defined  in  elementary  Algebra 
in  the  case  in  which  a  is  real  and  positive  and  f  rational,  or  a  real 
and  negative  and  £  a  rational  fraction  whose  denominator  is  odd. 
According  to  the  definitions  there  given  a*  has  at  most  two  values. 


222-225]        EXPONENTIAL,   AND   CIRCULAR   FUNCTIONS  405 

In  Ch.  Ill  we  extended  our  definitions  to  cover  the  case  in  which 
a  is  any  real  or  complex  number  and  f  any  rational  number  p/q ; 
and  in  Ch.  IX  we  gave  a  new  definition,  expressed  by  the  equation 

a(  _  e<r  log  a> 

which  applies  whenever  £  is  real  and  a  real  and  positive. 

Thus  we  have,  in  one  way  or  another,  attached  a  meaning  to 
such  expressions  as 

3'/*,     (-I)1/3,     (v/3+i0-1/2,     (3-S)1^; 

but  we  have  as  yet  given  no  definitions  which  enable  us  to  attach 
any  meaning  to  such  expressions  as 

(l+iy\     2\     (d  +  2if+si. 

We  shall  now  give  a  general  definition  of  a$  which  applies  to  all 
values  of  a  and  f,  real  or  complex,  with  the  one  limitation  that 
a  must  not  be  equal  to  zero. 

Definition.     The  function  at  is  defined  by  the  equation 

a$  =  exp  (£Log  a) 

where  Log  a  is  any  value  of  the  logarithm  of  a. 

We  must  first  satisfy  ourselves  that  this  definition  is  consistent 
with  the  previous  definitions  and  includes  them  all  as  particular 
cases. 

(1)  If  a  is  positive  and  £  real,  then  one  value  of  £"Loga,  viz. 
£  log  a,  is  real:  and  exp  (£log  a)  =  e^los  a,  which  agrees  with  the 
definition  adopted  in  Ch.  IX.  The  definition  of  Ch.  IX  is,  as 
we  saw  then,  consistent  with  the  definition  given  in  elementary 
Algebra ;  and  so  our  new  definition  is  so  too. 

(2)  If  a  =  eT  (cos  yjr  +  i  sin  -i/r),  then 

Log  a  =  t  +  i  (yjr  +  2nnr), 

exp  {(p/q)  Log  a]  =  e^T/?  Cis  {(p/q)  (f  +  2nnr)}, 

where  m  may  have  any  integral  value.  It  is  easy  to  see  that  if  m 
assumes  all  possible  integral  values  then  this  expression  assumes  q 
and  only  q  different  values,  which  are  precisely  the  values  of  aPlq 
found  in  §  48.  Hence  our  new  definition  is  also  consistent  with 
that  of  Ch.  III. 


406      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [x 

226.     The  general  value  of  ai     Let 

£  =  £  +  ir],     a  =  a  (cos  ijr  +  i  sin  yfr) 
where  —  7r<^  =  7r,  so  that,  in    the   notation    of  §  225,  a  =  eT  or 
t  =  log  <r. 

Then 

£Log  a  =  (|  +  irj)  {log  a  +  i  (i/r  +  2m7r)}  =  Z  +  %M, 
where 

X  =  |  log  O-  —  77  (l|r  +  2OT7r),       M  =7}  log  O-  +  £  (\/r  +  2  J717T)  J 

and  af  =  exp  (f  Log  a)  =  eL  (cos  ilf  +  t  sin  M). 

Thus  the  general  value  of  a^  is 

efioga-,(*+2mir)  [cos  |^  \0ga.  +  j:ty  +  2nnr)} 

+  i  sin  {7?  log  <r+^(i|r  +  2«wr)}]. 
In  general  a4  is  an  infinitely  many-valued  function.     For 

I  ai  I  _  gf  logo— T)(i/(+2»njr) 

has  a  different  value  for  every  value  of  m,  unless  77  =  0.  If  on  the 
other  hand  77  =  0,  then  the  moduli  of  all  the  different  values  of  a* 
are  the  same.  But  any  two  values  differ  unless  their  amplitudes 
are  the  same  or  differ  by  a  multiple  of  2ir.  This  requires  that 
i;(yjr  +  Imir)  and  i;(\jr  +  2mr),  where  m  and  n  are  different  integers, 
shall  differ,  if  at  all,  by  a  multiple  of  2ir.     But  if 

then  £  =  k/(m  —  n)  is  rational.  We  conclude  that  a^  is  infinitely 
many-valued  unless  %is  real  and  rational.  On  the  other  hand  we 
have  already  seen  that,  when  £  is  real  and  rational,  a$  has  but  a 
finite  number  of  values. 

The  principal  value  of  a^  =  exp  ((Log  a)  is  obtained  by  giving  Log  a  its 
principal  value,  i.e.  by  supposing  m  =  0  in  the  general  formula.  Thus  the 
principal  value  of  or  is 

gf  log  o -  „*  ;cos  ^  ]og  o-  +  ^)  +  j-  sin  (^  iog  a  +  ^)}# 

Two  particular  cases  are  of  especial  interest.  If  a  is  real  and  positive 
and  f  real,  then  a-  =  a,  ^  =  0,  £==f,  »7  =  0,  and  the  principal  value  of  a>  is 
e^loga,  which  is  the  value  denned  in  the  last  chapter.  If  [  or  |  =  1  and  (  is 
real,  then  <r  =  l,  £  =  £  »?  =  0,  and  the  principal  value  of  (cos  >//■  +  z  sin  ^)^  is 
cos  £\}a  +  i  sin  fy.  This  is  a  further  generalisation  of  De  Moivre's  Theorem 
(§§45,49). 


226]  EXPONENTIAL,   AND   CIRCULAR   FUNCTIONS  407 

Examples  XCIV.     1.     Find  all  the  values  of  i\     [By  definition 
i'  =  exp  (i  Log  i). 
But  t  =  cos^n  +  ism%ir,     hogi—(2k  +  \)ni, 

where  k  is  any  integei*.     Hence 

i*=exp{-(2£+|)  n}  =e-(2t+i),r. 
All  the  values  of  il  are  therefore  real  and  positive.] 

2.  Find  all  the  values  of  (1  +  i)\  i1 +«,  (1  +  i)i  +  *. 

3.  The  values  of  a%  when  plotted  in  the  Argand  diagram,  are  the  vertices 
of  an  equiangular  polygon  inscribed  in  an  equiangular  spiral  whose  angle  is 
independent  of  a.  (Math.  Trip.  1899.) 

[If  a^=  r  (cos  6  +  i  sin  6)  we  have 

r^IogT-nW+am^     3=r,log<r+£U,+2mv)>, 

and  all  the  points  lie  on  the  spiral  i^a-^'^^e'^K] 

4.  The  function  e\  If  we  write  e  for  a  in  the  general  formula,  so  that 
log  o-=l,  ^  =  0,  we  obtain 

e<T=  el-2»wni  ;cog  ^  +  2wi?r ^  +  ■  gJn  ^  +  2wi7r|)}. 

The  principal  value  of  e>  is  e*  (cos  q-H'sin  77),  which  is  equal  to  exp  £  (§  223). 
In  particular,  if  £  is  real,  so  that  r]  =  0,  we  obtain 

e*  (cos  2mn£+i  sin  2hi7t^) 

as  the  general  and  &•  as  the  principal  value,  e4  denoting  here  the  positive 
value  of  the  exponential  defined  in  Ch.  IX. 

5.  Show  that  Log  e^=(l  +2wnri)  £+2ft7n',  where  m  and  n  are  any  integers, 
and  that  in  general  Log  a?  has  a  double  infinity  of  values. 

6.  The  equation  l/a*  =  a~'  is  completely  true  (Ex.  xcm.  3):  it  is  also  true 
of  the  principal  values. 

7.  The  equation  a^xb^=(aby  is  completely  true  but  not  always  true  of 
the  principal  values. 

8.  The  equation  ct^x  a^'  =  a^+*'  is  not  completely  true,  but  is  true  of  the 
principal  values.  [Every  value  of  the  right-hand  side  is  a  value  of  the  left- 
hand  side,  but  the  general  value  of  a^xa^ ,  viz. 

exp  {£  (log  a  +  2mni)  +  f  (log  a+2mri)} , 
is  not  as  a  rule  a  value  of  a^+>  unless  m=n.] 

9.  What  are  the  corresponding  results  as  regards  the  equations 

Log  J=  C  Log  a,     (off  =  (a?)*  =  a®  ? 

10.  For  what  values  of  £  is  (a)  any  value  (b)  the  principal  value  of  eb 
(i)  real  (ii)  purely  imaginary  (iii)  of  unit  modulus  ? 


408      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [X 

11.  The  necessary  and  sufficient  conditions  that  all  the  values  of  at  should 
be  real  are  that  2£  and  {rj  log  |  a  J  +£  am  aj/rr,  where  am  a  denotes  any  value  of 
the  amplitude,  should  both  be  integral.  What  are  the  corresponding  con- 
ditions that  all  the  values  should  be  of  unit  modulus  ? 

12.  The  general  value  oi\xi+x~i\,  where  x>0,  is 

e-(m-n)7r^2  |cosh  2  (m  +  n)  „-+COS  (2  logo;)}]. 

13.  Explain  the  fallacy  in  the  following  argument :  since  e2mwi=e2n*i=l, 
where  m  and  n  are  any  integers,  therefore,  raising  each  side  to  the  power  i, 
we  obtain  e-*"l"=e-*m. 

14.  In  what  circumstances  are  any  of  the  values  of  xx,  where  x  is  real, 
themselves  real  ?    [If  x>  0  then 

xx  =  exp  (x  Log  x)  =  exp  (x  log  x)  Cis  2vi7rx, 
the  first  factor  being  real.     The  principal  value,  for  which  m=0,  is  always 
real. 

If  x  is  a  rational  fraction  p/(2q  +  l),  or  is  irrational,  then  there  is  no  other 
real  value.     But  if  x  is  of  the  form  pj2q,  then  there  is  one  other  real  value, 
viz.  -  exp  (so  log  x\  given  by  m  -  q. 
If#=-£<0then 

^  =  exp{-£Log(-£)}  =  exp(-£log£)Cis{-(2m  +  l)7r!}. 
The  only  case  in  which  any  value  is  real  is  that  in  which  £=pj(2q  +  l),  when 
m=q  gives  the  real  value 

exp(-£log£)Cis(-^)  =  (-l)^. 
The  cases  of  reality  are  illustrated  by  the  examples 

a>*-M,  (*)*-±«A>  (-t)"*-n    (-*)"*-- w 

15.  Logarithms  to  any  base.  We  may  define  £  =  Loga2  in  two  different 
ways.  We  may  say  (i)  that  £=  Loga  z  if  the  principal  value  of  a?  is  equal  to  z ; 
or  we  may  say  (ii)  that  (  -  Loga  2  if  a»y  value  of  o.»  is  equal  to  2. 

Thus  if  a=e  then  f=Log,,2,  according  to  the  first  definition,  if  the 
principal  value  of  e>  is  equal  to  z,  or  if  exp  £=2;  and  so  Loge2  is  identical 
with  Log  z.     But,  according  to  the  second  definition,  ^  =  Loge2  if 

ef=exp(fLoge)  =  2,     £Loge  =  Log2, 

or  £=(Log  z)/(Log  e),  any  values  of  the  logarithms  being  taken.     Thus 

f     T  ~,  •     log  I  ^  I  +  (am  g  +  2?»ir)t 
f  =  L0g62= j-j^j-j , 

so  that  f  is  a  doubly  infinitely  many-valued  function  of  2.     And  generally, 
according  to  this  definition,  Loga2  =  (Log2)/(Loga). 

16.  Loggl=2m»ri/(l  +  2»7ri),  Loge(-l)  =  (2m  +  l)  iriftl  +  2%ni)1  where  m 
and  n  are  any  integers. 


226-228]        EXPONENTIAL,    AND   CIRCULAR   FUNCTIONS  409 

227.  The  exponential  values  of  the    sine    and  cosine. 

From  the  formula 

exp  (f  +  it))  =  exp  £  (cos  77  +  i  sin  77), 

we  can  deduce  a  number  of  extremely  important  subsidiary 
formulae.  Taking  £=0,  we  obtain  exp  (177)  =  cos  77  +  i  sin  77 ;  and, 
changing  the  sign  of  77,  exp  (—  irj)  =  cos  rj  —  i  sin  77.     Hence 

cos 77=     ^    {exp  (irj)  +  exp (  —ir))), 

sin  77  =  —  ^  i  {exp  (177)  —  exp  (  —  {77)}. 

We  can  of  course  deduce  expressions  for  any  of  the  trigonometrical 
ratios  of  77  in  terms  of  exp  (it]). 

228.  Definition   of  sin  f  and  cos  f  for   all  values  of f. 

We  saw  in  the  last  section  that,  when  f  is  real, 

cosf  =     \  {exp  (if)  +  exp  (-  if)} (la), 

sin  £  =  -  \i {exp  (if)  -  exp  (-  if)} (16). 

The  left-hand  sides  of  these  equations  are  defined,  by  the  ordinary 
geometrical  definitions  adopted  in  elementary  Trigonometry,  only 
for  real  values  of  f.  The  right-hand  sides  have,  on  the  other 
hand,  been  defined  for  all  values  of  f,  real  or  complex.  We  are 
therefore  naturally  led  to  adopt  the  formulae  (1)  as  the  definitions 
of  cos  f  and  sin  f  for  all  values  of  f.  These  definitions  agree,  in 
virtue  of  the  results  of  §  227,  with  the  elementary  definitions  for 
real  values  of  f. 

Having  defined  cos  f  and  sin  f,  we  define  the  other  trigono- 
metrical ratios  by  the  equations 

sin  f  cos  f  1  «.       1         /s>x 

tanf  = 7.,    cotf=- — y,    secf  = ^,    cosec  f=  -r— ■  ...(2). 

cos  f  sin  f  cos  f  sin  f      v 

It  is  evident  that  cos  f  and  sec  f  are  even  functions  of  f,  and 
sin  f,  tan  f,  cot  f,  and  cosec  f  odd  functions.  Also,  if  exp  (if)  =  t, 
we  have 

cos  f =  h  {t  +  (1/0},     sin  f  =  -£  i  {£  -  (1/0}, 

cos2f  +  sin2f=i[{«+(l/0]2-^-(l/0j2]=l (3). 

We  can  moreover  express  the  trigonometrical  functions  of 
f+  f  in  terms  of  those  of  f  and  f  by  precisely  the  same  formulae 


410      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC.        [X 

as  those  which  hold  in  elementary  trigonometry.    For  if  exp  (i%)  =  t, 
exp  (if)  =  t',  we  have 

coS(?+n=i(«'+^{K)(<'4)  +  K)(('-F 

=  cos  £cos  f  —  sin  £sin  £' (4) ; 

and  similarly  we  can  prove  that 

sin(£+0  =  sm£cos£'  +  cos£sin£' (5). 

In  particular 

cos(£  +  £7r)  =  -sin  £         sin(^+^7r)  =  cos  £ (6). 

All  the  ordinary  formulae  of  elementary  Trigonometry  are 
algebraical  corollaries  of  the  equations  (2) — (6) ;  and  so  all  such 
relations  hold  also  for  the  generalised  trigonometrical  functions 
defined  in  this  section. 

229.  The  generalised  hyperbolic  functions.  In  Ex.  lxxxvii.  19,  we 
denned  cosh  f  and  sinh  £,  for  real  values  of  £,  by  the  equations 

cosh  C=i {exp C + exp  (-£)},      sinh  (=h  {exp  f- exp  (-{)}  (1). 

We  can  now  extend  this  definition  to  complex  values  of  the  variable; 
i.e.  we  can  agree  that  the  equations  (1)  are  to  define  cosh  f  and  sinh£  for 
all  values  of  £  real  or  complex.  The  reader  will  easily  verify  the  following 
relations : 

cos  i£ = cosh  £,     sin  i£=  i  sinh  £,     cosh  i£= cos  f,     sinh  i£— i  sin  £. 

We  have  seen  that  any  elementary  trigonometrical  formula,  such  as 
the  formula  cos  2^"= cos2  £  —  sin2  f,  remains  true  when  £  is  allowed  to  assume 
complex  values.  It  remains  true  therefore  if  we  write  cos  i(  for  cos  £,  sin  i£ 
for  sin  £  and  cos  2^  for  cos  2£;  or,  in  other  words,  if  we  write  cosh  £  for  cos  (, 
isinh  £  for  sin  £,  and  cosh  2£  for  cos  2£.     Hence 

cosh  2^= cosh2  C+sinh2  £. 
The  same  process  of  transformation  may  be  applied  to  any  trigonometrical 
identity.     It  is  of  course  this  fact  which  explains  the  correspondence  noted 
in  Ex.  lxxxvii.  21  between  the  formulae  for  the  hyperbolic  and  those  for  the 
ordinary  trigonometrical  functions. 

230.  Formulae  for  cos(£+iij),  sin(£+ty),  etc.  It  follows  from  the 
addition  formulae  that 

cos  (£  +  *i?)  =  cos  £  cos  iff  —  sin  £  sin  it]  =  cos  |  cosh  ip—i  sin  £  sinh  rj, 

sin  (£  +  irj)  =  sin  |  cos  ir)  +  cos  £  sin  it]  =  sin  £  cosh  r/  +  i  cos  £  sinh  rj. 

These  formulae  are  true  for  all  values  of  £  and  rj.  The  interesting  case 
is  that  in  which  £  and  77  are  real.  They  then  give  expressions  for  the  real  and 
imaginary  parts  of  the  cosine  and  sine  of  a  complex  number. 


228-230]        EXPONENTIAL,   AND   CIRCULAR    FUNCTIONS  411 

Examples  XCV.  1.  Determine  the  values  of  £  for  which  cos  £  and  sin  £ 
are  (i)  real  (ii)  purely  imaginary.  [For  example  cos£  is  real  when  77=0  or 
when  £  is  any  multiple  of  77-.] 

2.  I  cos  (£  +  irf)  I  =  s/(cos2 £  +  sinh2  77)  =  J{\  (cosh  2r)  +  cos  2|)}, 
I  sin  (£  +  irj)  I  =  N/(sin2  £  +  sinh2  77)  =  J{$  (cosh  2rj  -  cos  2£)}. 

[Use  {e.g.)  the  equation     |  cos  (%+irj)  |  =  s/{cos  (£  +  ^77)  cos  (£  —  177)}.] 

_  /*  .  ■  \     sin  2^4- ^ sinh  2^  ..     sin2|-i'sinh2n 

3.  ten(g+t,)=cosh2)7+co^g>        cot(g+»,)  =  co8h2||_co82r 

[For  example 

tan  (g+^)=Sin(/f+^C°3/(f-^  =  si"2g  +  sin2^ 
v        "     cos  ($+ it))  cos  (£-ir))     cos  2§  +  cos  2i»; 

which  leads  at  once  to  the  result  given.] 

cos  £  cosh  77  +  i  sin  £  sinh  jj 


sec  (£  +  ^77) 
cosec  (£+^7?)  = 


£  (cosh  2»j+ cos  2£)         ' 
sin  £  cosh  7;  —  i  cos  £  sinh  77 
J  (cosh  2r;  —  cos  2£) 


5.  If  I  cos  (£  +  {77)  |  =  1  then  sin2 £  =  sinh2  77,  and  if  |  sin  (t-  +  ir))  |  =  1  then 
cos2  £  =  sinh2 17. 

6.  If  |cos(f  +  i!i)|  =  l,  then 

sin*{am  cos  (£  +  £77)}  =  ±  sin2 £  =  ±  sinh2  rj. 

7.  Prove  that  Log  cos  (£  + 17)  —  A-\-  iB,  where 

J.  =  \  log  {|  (cosh  2r/  +  cos  2£)} 
and  5  is  any  angle  such  that 

cos  B  sin  Z?  1 


cos  £  cosh  77         sin  £  sinh  77      s/{j  (cosh  277  +  cos  2£)}  * 
Find  a  similar  formula  for  Log  sin  (£  +  177). 

8.    Solution  of  the  equation  cos£=a,  where  a   is  real.     Putting 
£=£  +  177,  and  equating  real  and  imaginary  parts,  we  obtain 

cos  £  cosh  77  =  a,         sin  £  sinh  77  =  0. 
Hence  either  77  =  0  or  £  is  a  multiple  of  71-.     If  (i)  77  =  0  then  cos  £  =  a,  which  is 
impossible  unless  -l£a£l.     This  hypothesis  leads  to  the  solution 

£=2£n-±arc  cos  a, 
where  arccosa  lies  between  0  and  \n.    If  (ii)  £=mn  then  cosh  t/  =  (  —  l)ma,  so 
that  either  o^l  and  m  is  even,  or  a<  —  1  and  m  is  odd.    If  a=  ±1  then  77  =  0, 
and  we  are  led  back  to  our  first  case.     If  |  a  |  >  1  then  cosh  77  =  |  a  \ ,  and  we 
are  led  to  the  solutions 

£=  2for±i*log{      a  +  V(n2-l)}        (a>l), 

C=(2k  +  1)  ir±tlog{-  a  +  v/(a2-l)}        (a<-l). 
For  example,  the  general  solution  of  cos  f=  —  f  is  (=  (2k  + 1)  tt±  Hog  3. 


412      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,       [X 

9.  Solve  sin  f =  a,  where  a  is  real. 

10.  Solution  of  cos  ^= a +  i'/3,  where  04=0.  We  may  suppose  /3>0, 
since  the  results  when  /3  <0  may  be  deduced  by  merely  changing  the  sign  of  i. 
In  this  case 

cos  £  cosh  r)  =  a,       sin  £  sinhr;=  -/3 (1), 

and  (a/cosh  t])2  +  (/3/sinh  7)2=1. 

If  we  put  cosh2  jj  =  x  we  find  that 

a;2-(l+a2+/32j#4-a2=0 
■or  a;  =  (^  i  ±  A 2)2,  where 

vl,  =  W{(«  +  l)2+/32},         ^2  =  K/{(«-D2+32}. 
Suppose  a  >  0.     Then  Ax >  A2  >  0  and  cosh  ^  =  J  t  ±  42.     Also 

cos  |  =  a/(cosh  rj)  =  A1  +  A2, 
and  since  cosh  rj  >  cos  £  we  must  take 

cosh  r)  =  A1 +A2,        cos£  =  ^41-^2' 
The  general  solutions  of  these  equations  are 

£  =  2/br±arccos  J/,         v=  ±log  [Z  +  N/(Z2- 1)} (2), 

where  L  =  At+A2,  3f=A1-A2,  and  arc  cos  M lies  between  0  and  \n. 

The  values  of  r\  and  £  thus  found  above  include,  however,  the  solutions  of 
the  equations 

cos  |  cosh  r)  =  a,  sin  £  sinhr?  =  /3  (3), 

as  well  as  those  of  the  equations  (1),  since  we  have  only  used  the  second  of 
the  latter  equations  after  squaring  it.  To  distinguish  the  two  sets  of 
solutions  we  observe  that  the  sign  of  sin  £  is  the  same  as  the  ambiguous  sign 
in  the  first  of  the  equations  (2),  and  the  sign  of  sinh  r)  is  the  same  as  the 
ambiguous  sign  in  the  second.  Since  /3  >  0,  these  two  signs  must  be  different. 
Hence  the  general  solution  required  is 

f=2/(-7r±[arccosJ/-ilog{Z;  +  N/(Z2-l)}]. 

11.  Work  out  the  cases  in  which  a  <0  and  a  =  0  in  the  same  way. 

12.  If  /3  =  0  then  L  =  h\a  +  l\ +h\a-l\  and  M=h\  a  +  l  \-h\  a-l  \. 
Verify  that  the  results  thus  obtained  agree  with  those  of  Ex.  8. 

13.  Show  that  if  a  and  /3  arc  positive  then  the  general  solution  of 
sin  £=a  +  i'/3  is 

(=l-n  +  ( -  l)*  [arc  sin  M+ilog  {L+J(L2  - 1)}], 

where  arc  sin  M  lies  between  0  and  ^n.  Obtain  the  solution  in  the  other 
possible  cases. 

14.  Solve  tan  (=a,  where  a  is  real.     [All  the  roots  are  real.] 


230,  231]       EXPONENTIAL,   AND    CIRCULAR   FUNCTIONS  413 

15.  Show  that  the  general  solution  of  tan  f=a  +  z'/3,  where  fi=£0,  is 

>■     i     L l/i  ,i-i       («2  +  (l+/3)2l 

where  8  is  the  numerically  least  angle  such  that 

cos  6  :  sin  6  :  1 : :  1  -  a2  -  /32 :  2a  :  V{(1  -  a2  -  /32)2+ 4a2}. 

16.  If  2  =  £exp  (jn-i),  where  £  is  real,  and  c  is  also  real,  then  the  modulus 
of  cos  2nz  -  cos  2nc  is 

Vrjf  {1  +cos  4ttc  +  cos  (2tt|  V2)  +  cosh  (2tt£  ^2) 

-  4  cos  2n-c  cos  (»r^  v/2)  cosh  («■£  >/2)}]. 

1 7.  Prove  that         |  exp  exp  (£  +  iq)  |  =  exp  (exp  £  cos r?), 

E,  {cos  cos  ({■  + it))}  =  cos  (cos  £  cosh  77)  cosh  (sin  £  sinh  »?), 
I  {sin  sin  (£  +  iij)}  =  cos  (sin  £  cosh  77)  sinh  (cos  £  sinh  ??). 

18.  Prove  that  |exp  f  |  tends  to  00  if  £  moves  away  towards  infinity  along 
any  straight  line  through  the  origin  making  an  angle  less  than  \tt  with  OXr 
and  to  0  if  (  moves  away  along  a  similar  line  making  an  angle  greater  than. 
|tt  with  OX. 

19.  Prove  that  |cosf|  and  |sin£|  tend  to  co  if  f  moves  away  towards 
infinity  along  any  straight  line  through  the  origin  other  than  either  half  of 
the  real  axis. 

20.  Prove  that  tan  f  tends  to  —  i  or  to  i  if  f  moves  away  to  infinity 
along  the  straight  line  of  Ex.  19,  to  —  i  if  the  line  lies  above  the  real  axis  and 
to  i  if  it  lies  below. 

231.  The  connection  between  the  logarithmic  and  the  inverse 
trigonometrical  functions.  We  found  in  Ch.  VI  that  the  integral  of  a 
rational  or  algebraical  function  4>(x,  a,  /3,  ...),  where  a,  /3,  ...  are  constants, 
often  assumes  different  forms  according  to  the  values  of  a,  /3,  ... ;  sometimes 
it  can  be  expressed  by  means  of  logarithms,  and  sometimes  by  means  of 
inverse  trigonometrical  functions.     Thus,  for  example, 

rlv         1                  x  .  . 

— . —  =  —r-  arc  tan  — -    (1) 

x*  +  a      si  a  s!a 

if  a  >  0,  but 

doc  1        .      \x—J(  —  a)\  ,  . 

log ;  „  ,  ,;     ,     (2) 


/. 


/ 


2+a  2N/(-a)  6|*+V(-«») 
if  a  <  0.  These  facts  suggest  the  existence  of  some  functional  connection 
between  the  logarithmic  and  the  inverse  circular  functions.  That  there 
is  such  a  connection  may  also  be  inferred  from  the  facts  that  we  have  ex- 
pressed the  circular  functions  of  £  in  terms  of  exp  if,  and  that  the  logarithm 
is  the  inverse  of  the  exponential  function. 

Let  us  consider  more  particularly  the  equation 


f    doe         1  .      (x  —  a\ 


414      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,       [x 

which  holds  when  a  is  real  and  (x-a)/(x  +  a)  is  positive.  If  we  could  write 
ia  instead  of  a  in  this  equation,  we  should  be  led  to  the  formula 

arctan(-)  =  ilog(^-aW (3), 

\aj       2i      °  \X+laJ 

where  G  is  a  constant,  and  the  question  is  suggested  whether,  now  that  we 
have  defined  the  logarithm  of  a  complex  number,  this  equation  will  not  be 
found  to  be  actually  true. 

Now  (§  221) 

Log(#±i'a)  =  £log(.r2  +  a2)±i(0  +  2£7r), 

where  k  is  an  integer  and  <£  is  the  numerically  least  angle  such  that 
cos  0  =  A'/v/(^'2  +  a2)  and  sin  0  =  a/N/(.^2  +  a2).     Thus 

—.   Log  (  r  )  =  —  6  -  ?TTt 

2 J        °  \x  +  iaj  r  ' 

where  I  is  an  integer,  and  this  does  in  fact  differ  by  a  constant  from  any 
value  of  arc  tan  (xja). 

The  standard  formula  connecting  the  logarithmic  and  inverse  circular 
functions  is 

arctan^=-.Log(f±^j (4), 

where  x  is  real.  It  is  most  easily  verified  by  putting  #=tany,  when  the  right- 
hand  side  reduces  to 

1  T       /cos?/  +  isiny\       1  T       .        _.  .         ,  , 
— .  Log         " — — — -  )  =  -— .  Log  (exp  2iy)  =  v  +  Ictt, 
2i      °  \cosy-ismyJ      2t      oV     r     "'    * 

where  h  is  any  integer,  so  that  the  equation  (4)  is  '  completely '  true  (Ex.  xcin. 
3).     The  reader  should  also  verify  the  formulae 

arccos.r=  -^Log  {x±i*J(l  —  #2)},     arc  sin  #=  —  i' Log  {£f  +  ^/(l  —  x2)}...(5), 

where  —  \£.x£L\:  each  of  these  formulae  also  is  'completely'  true. 

Example.     Solving  the  equation 

cos  «=#  =  |{y +  (%)}, 

where  y  =  exp  (iu),  with  respect  to  ?/,  we  obtain  y  =  x  +  isJ(\  —  x"1).     Thus: 

u=  - i  Logy=  —  i  Log  {x±i,J(l  -.r2)}, 

which  is  equivalent  to  the  first  of  the  equations  (5).  Obtain  the  remaining 
equations  (4)  and  (5)  by  similar  reasoning. 

232.     The   power   series   for   exp  z*.      We   saw   in  §  212 

that  when  z  is  real 

exp  z  =  1  +  z  +o-,+ (1). 

Moreover  we  saw  in  §  191  that  the  series  on  the  right-hand  side 

*  It  will  be  convenient  now  to  use  z  instead  of  £  as  the  argument  of  the 
exponential  function. 


231,  232]       EXPONENTIAL,    AND    CIRCULAR    FUNCTIONS  415 

remains  convergent  (indeed  absolutely  convergent)  when  z  is  com- 
plex. It  is  naturally  suggested  that  the  equation  (1)  also  remains 
true,  and  we  shall  now  prove  that  this  is  the  case. 

Let  the  sum  of  the  series  (1)  be  denoted  by  F(z).  The  series 
being  absolutely  convergent,  it  follows  by  direct  multiplication  (as 
in  Ex.  lxxxi.  7)  that  F (z)  satisfies  the  functional  equation 

F(z)F{h)  =  F(z  +  h)     (2). 

Now  let  z  =  iy,  where  y  is  real,  and  F  (z)  =  f(y).     Then 

f(y)f(k)=f(y+k); 

and  so  /(y  +  *)-/(y)  =  /(y)f/(fc)-l 


k  J  XJ'  \      k 

^   .  f(k)-l      .  f-       ik     (iky-  ) 

and  so,  if  |  k  \  <  1, 


f(k)-l 
k 


1        1 


2  ! 


+  3T+...)|*|<(«-2)|* 


Hence  {f  (k)  -  l\/k-*~i  as&-*-0,  and  so 

f,(y)  =  limf(y  +  V-f(y)  =  {f(lj)  (3) 

Now 

f{y)  =  *»)  =  1  +  (iy)  +  {^+...  =  (fi{y)  +  if  (y), 

where  <f>  (y)  is  an  even  and  i/r  (^/)  an  odd  function  of  y,  and  so 

\f(y)\  =  */l{<!>(y)}*+{+(y)Y] 

-VW(y)  +  *t^)H*(y)-»+(y)}] 

and  therefore 

/  (y)  —  cos  ^r + *  sm  ^  > 
where  Y  is  a  function  of  y  such  that  —  it  <  Y  S  7r.  Since/  (y)  has 
a  differential  coefficient,  its  real  and  imaginary  parts  cos  Ya,nd  sin  Y 
have  differential  coefficients,  and  are  a  fortiori  continuous  functions 
of  y.  Hence  Y  is  a  continuous  function  of  y.  Suppose  that  Y 
changes  to  Y  +  K  when  y  changes  to  y  +  k.  Then  K  tends  to 
zero  with  k,  and 

K  _  fcos  ( Y  +  K)  -  cos  Y\  /  (cos  ( Y  +  K)  -  cos  T 


k       {  k  ]/  (  K 

Of  the   two  quotients  on  the  right-hand  side  the  first  tends  to  a 


416      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [x. 

limit  when  Jc-*~0,  since  cos  Y  has  a  differential  coefficient  with 

respect  to  y,  and  the  second  tends  to  the  limit  —  sin  Y.    Hence  K/fc 

tends  to  a  limit,  so  that  Y  has  a  differential  coefficient  with  respect 

to  y. 

dY 

Further  /'  (y)  -■=  (-  sin  Y+i  cos  Y)  -=-  . 

But  we  have  seen  already  that 

/'  (y)  =  if(y)  =  -  sin  F+  i  cos  Y. 

dY 
Hence  -j-  =  *>  Y  =  y  +  G, 

dy 

where  G  is  a  constant,  and 

/  (y)  =  cos  (y+C)  +  i  sin  (y  +  C). 

But  /(0)  =  1  when  y=0,  so  that  C  is  a  multiple  of  2tt,  and 

y  (y)  =  cos  y  +  i  sin  y.      Thus   F  (iy)  =  cos  ?/  +  i  sin  3/   for  all   real 

values  of  y.     And,  if  x  also  is  real,  we  have 

F  (x  +  iy)  =  F  (x)  F{iy)  =  exp  x  (cos  y  +  i  sin  y)  =  exp  (x  +  iy), 

or  exipz=l+z  +  ^]+  -••> 

for  all  values  of  z. 

233.     The    power   series  for  cos  z  and  sin  z.      From  the 
result  of  the  last  section  and  the  equations  (1)  of  §  228  it  follows 

at  once  that 

z"      zA  .  z3      z5 

cosz  =  l  -  2l  +  4!-"-'  smz==2:~'si  +  5''~'" 
for  all  values  of  z.     These  results  were  proved  for  real  values  of  z 
in  Ex.  lvi.  1. 

Examples  XCVI.     1.     Calculate  cos  i  and  sin  %  to  two  places  of  decimals 
by  means  of  the  power  series  for  cos  z  and  sin  z. 

2.  Prove  that  |  cos z  \  S.  cosh  |  z  \  and  |  sin  z  \  ^ sinh  \z\. 

3.  Prove  that  if  |  z  |  <  1  then  |  cos z  |  <  2  and  |  sin  z  \  < f  j  z  \. 

4.  Since  sin  %z—1  sin  z  cos  z  we  have 

\    )       3,    -r    5,  ^^     3!-r...^i     2,-r...y. 

Prove  by  multiplying  the  two  series  on  the  right-hand  side  (§  195)  and 
equating  coefficients  (§  194)  that 

(TK"3+>-+Gi::;H- 

Verify  the  result  by  means  of  the  binomial  theorem.    Derive  similar  identities 
from  the  equations 

cos2 z  +  sin2 2=1,     cos  22  =  2  cos2 z  — 1  =  1-2  sin2 2. 


232-234]        EXPONENTIAL,    AND    CIRCULAR    FUNCTIONS  417 

5.  Show  that        exp  {( 1  + 1)  z}  =  2  2-n  exp  (£ ?itu')  — . 

6.  Expand  cos  2  cosh  0  in  powers  of  2.     [We  have  ■ 

cos  z  cosh  s+i  sin  2  sinh  z— cos  {(1  - 1')  2}  =  |-  [exp  {(l+i)  2}  +  exp  {  -  (1  +  i)  2}] 

=|22*»{l  +  (-l)»}eXp(4W)!J, 

and  similarly  cos  2  cosh  s  —  i  sin  2  sinh  2  =  cos  (1  +  ?')  z 

=  |l2^{l  +  (-l)»}exp(-i^)^. 

0  »! 

Hence        cosscosh«=i2  2Jw{l  +  (-l)n}cosi7i7r  ^=1-  -7?  +  ^-; — 1 

"0  n !  4 !        8 !  J 

7.  Expand  sin  2  sinh  2,  cos  2  sinh  2,  and  sin  2  cosh  z  in  powers  of  2. 

8.  Expand  sin2  2  and  sin32  in  powers  of  2.     [Use  the  formulae 

sin22=i  (1  -cos  2z),     sin32=J  (3  sin  2 -sin  32),  .... 
It  is  clear  that  the  same  method  may  be  used  to  expand  cos™  2  and  sinn2, 
where  n  is  any  integer.] 

9.  Sum  the  series 

~     ,     cos  2     cos  2z     cos  32  _,    sin  2     sin  2z     sin  32 

[Here        0+iS=l+?&®  +  ?ffi£*>+...=«p  {exp(is)} 

=  exp  (cos  2)  {cos  (sin  z)  +  i  sin  (sin  2)}, 
and  similarly 

C—JS=  exp  {exp  ( -  iz)}  =  exp  (cos  2)  (cos  (sin  2)  -  »  sin  (sin  2)}. 

Hence  C=exp(cos2)  cos  (sin  2),     S= exp  (cos  2)  sin  (sin  2).] 

„  a  cos  2     a2  cos  22  a  sin  2     a2  sin  22 

10.  Sum  i  +  -Tr+_T]— +...,  ___  +  ___+.... 

_,         ,   COS  22   COS  42        cos  2   COS  32 

11.  Sum         1-—  r+  -4,     -...,     — 3j-+- 

and  the  corresponding  series  involving  sines. 

12.  Show  that 

,      cos  Az  ,   cos  82  ,  if/  s        i    ,  •       v 

1  -\ — — j — I — ^y-  + . . .  =  §  {cos  (cos  2)  cosh  (sin  2)  +  cos  (sin  2)  cosh  (cos  2)}. 

13.  Show  that  the  expansions  of  cos  (x+h)  and  sin  (x  +  h)  in  powers  of  h 
(Ex.  lvi.  1)  are  valid  for  all  values  of  x  and  h,  real  or  complex. 

234.     The  logarithmic  series.       We  found  in  §  213  that 

\og(l+z)  =  z-±z2  +  U3- (1) 

when  z  is  real  and  numerically  less  than  unity.     The  series  on  the 

right-hand  side  is  convergent,  indeed  absolutely  convergent,  when 

h.  27 


418      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [x 

z  has  any  complex  value  whose  modulus  is  less  than  unity.  It  is 
naturally  suggested  that  the  equation  (1)  remains  true  for  such 
complex  values  of  z.  That  this  is  true  may  be  proved  by  a 
modification  of  the  argument  of  §  213.  We  shall  in  fact  prove 
rather  more  than  this,  viz.  that  (1)  is  true  for  all  values  of  z  such 
that  |  « |  =  1,  with  the  exception  of  the  value  —  1. 

It  will  be  remembered  that  log  (1  +  z)  is  the  principal  value  of 
Log  (1  4-  z),  and  that 

i      /-i       x      f  du 

lo«(1+*>-.U' 

where  C  is  the  straight  line  joining  the  points  1  and  1  +  z  in  the 
plane  of  the  complex  variable  u.  We  may  suppose  that  z  is  not 
real,  as  the  formula  (1)  has  been  proved  already  for  real  values 
of  z. 

If  we  put 

z  =  r  (cos  9  +  i  sin  6)  =  £r, 
so  that  |  r  |  g  1,  and 

u  =  1  +  ft 

then  u  will  describe  G  as  t  increases  from  0  to  r.     And 


f  du_  FZdt 

J c  u       Jo  1 


+  & 

=  J    h-  ft  +  ?F-  ...  +  (-  l)™"1  £"*•«  +        1+ w       \dt 

-p         2    +    3        ",+(     L)         m    +Km 

=  ,__+__...+(_!).-._  +Rm    (2), 

where 

fr    fni/Jt 

^  =  (-l)-r+1Jol+^  (3). 

It  follows  from  (1)  of  §  164  that 

fr      Wflf 

'*»'a/o|Wl    ; <4>' 

Now  1 1  +  %t  |  or  |  u  |  is  never  less  than  ■or,  the  perpendicular  from 
0  on  to  the  line  C*     Hence 

1  fr                     rm+i                    n 
Rm  I  ^  "         tmdt  =  , — -  g  ; —-  , 

*  Since  z  is  not  real,  C  cannot  pass  through  0  when  produced.     The  reader  is 
recommended  to  draw  a  figure  to  illustrate  the  argument. 


234,  235]        EXPONENTIAL,   AND   CIRCULAR   FUNCTIONS  419 

and  so  Rm  -*•  0  as  m  -*■  oo  .     It  follows  from  (2)  that 

\og(l+z)  =  z-^z2  +  ^- (5). 

We  have  of  course  shown  in  the  course  of  our  proof  that  the 
series   is   convergent :    this    however    has    been    proved   already 
(Ex.  lxxx.  4).     The  series  is  in  fact  absolutely  convergent  when 
z  |  <  1  and  conditionally  convergent  when  |  z  \  =  1. 

Changing  z  into  —  z  we  obtain 

1°g(n^)=-lo8-(1-*)=*  +  ^2+^3+ («)■ 

235.     Now 

log  ( 1  +  z)  =  log  {(l+r  cos  0)  +  ir  sin  0} 

=  A  log  (1  +  2r  cos  0  +  r2)  +  i  arc  tan  ( -A  . 

\1  +  rcostv 

That  value  of  the  inverse  tangent  must  be  taken  which  lies 
between  —  \ir  and  \tt.  For,  since  1  +  z  is  the  vector  represented 
by  the  line  from  —  1  to  z>  the  principal  value  of  am  (1  4-  2)  always 
lies  between  these  limits  when  z  lies  within  the  circle  |  z\  =  L* 

Since  zm  =  rm  (cos  m0  +  i  sin  md),  we  obtain,  on  equating  the 
real  and  imaginary  parts  in  equation  (5)  of  §  234, 

I  log  (1  +  2r  cos  0  +  r2)  =  r  cos  0  -  |r2  cos  26  +  £r3  cos  30  -  . . ., 

(r  sin  0   \ 
^)  =  rsin  0  —  Ar2sin20  +  ir3sin30  —  .... 
l  +  r  cos  6/  *  6 

These  equations  hold  when  0  ^  r  ^  1,  and  for  all  values  of  0,  except 
that,  when  r  =  1,  0  must  not  be  equal  to  an  odd  multiple  of  ir. 
It  is  easy  to  see  that  they  also  hold  when  —  1  ^  r  ^  0,  except  that, 
when  r  —  —  1,  0  must  not  be  equal  to  an  even  multiple  of  ir. 

A   particularly  interesting  case  is  that  in   which  r  =  1.      In 
this  case  we  have 

log  (1  +  z)  =  log (1  +  Cis  0)  =\  log  (2  +  2  cos0)  +  i  arc  tan  ( — -r) 

=  |log(4cos2i0)  +  i;0, 
if  —ir<0<7r,  and  so 

cos0-,|cos20  +  icos30  -...  =  £  log  (4  cos2  J0), 

sin  0-  |  sin  20  +  £  sin  30  -  ...  =  |0. 

*  See  the  preceding  footnote. 

27—2 


420 


THE  GENERAL  THEORY  OF  THE  LOGARITHMIC, 


[X 


The  sums  of  the  series,  for  other  values  of  &,  are  easily  found  from 
the  consideration  that  they  are  periodic  functions  of  6  with  the 
period  2tt.  Thus  the  sum  of  the  cosine  series  is  ^  log  (4  cos2  \6)  for 
all  values  of  6  save  odd  multiples  of  ir  (for  which  values  the  series 
is  divergent),  while  the  sum  of  the  sine  series  is  ^  (6  —  2kir)  if 
(2k  —  1)  it  <  6  <  (2k  +  1)  7r,  and  zero  if  6  is  an  odd  multiple  of  it. 
The  graph  of  the  function  represented  by  the  sine  series  is  shown 
in  Fig.  58.     The  function  is  discontinuous  for  6  =  (2k  +  1)  it. 


Fig.  58. 
If  we  write  iz  and  -  iz  for  z  in  (5),  and  subtract,  we  obtain 

^i„g(L±|)-2-is»+}s5-.... 

If  z  is  real  and  numerically  less  than  unity,  we  are  led,  by  the  results  of 
§  231,  to  the  formula 

arc  tan  z=z  —  \zz  +  \£>-  ..., 

already  proved  in  a  different  manner  in  §  214. 

Examples  XCVII.     1.     Prove  that,  in  any  triangle  in  which  a>b, 

Jogc  =  loga — cos  C-  — ,cos2(7-.... 
a  2a- 

[Use  the  formula  log  c=-|  log  (a2+b2-2ab  cos  C).] 

2.  Prove  that  if  -  l<r<l  and  ~ln<0<\7r  then 

r  sin  20  -  \r2  sin  40  +  Jr3  sin  60 - ...  =  f-arc  tan  |(~)  tan  0 [ , 

the  inverse  tangent  lying  between  -\tv  and  |tt.     Determine  the  sum  of  the 
series  for  all  other  values  of  0. 

3.  Prove,  by  considering  the  expansions  of  log  (1 +  12)  and  log(l-;V)  in 
powers  of  z,  that  if  - l<r<  1  then 

rsm0  +  %r2cos20-§r3sm30-±?-*cos40+...  =  hlog(l  +  2rsm0  +  r2), 

r  cos  0  +  ± r2  sin  20  -  £ r3  cos  30  -  % r*  sin  40  +  ...=  arc  tan  (    rC0*6    )  , 

\]  —  r  sin  6 J 

6  *     O\l-2rsm0  +  r2)' 

r  cos  0  -  J?-3  cos  Z0  + . . .  =  *  arc  tan  &52L^ 

\  1-r- 

the  inverse  tangents  lying  between  -  £tt  and  \n. 


he* 


l-\hz  I ' 


235,  236]        EXPONENTIAL,   AND   CIRCULAR   FUNCTIONS  421 

4.     Prove  that 

cos  0  cos  0  -  \  cos  20  cos2  0  +  £  cos  38  cos3  0 - ...  =$  log  (1  +  3  cos2  0), 

sin  6 sin  #  -  ^ sin  2(9 sin2  0  +  J  sin  30  sin3  0  -  ...  =  arc  cot  (1  +  cot  0  +  cot2 0), 

the  inverse  cotangent  lying  between    -^tt  and  ^7r  ;    and  find  similar  ex- 
pressions for  the  sums  of  the  series 

cos  0 sin  0 -|  cos 20  sin2  0  + . ..,     sin  0  cos  0-J  sin  2(9  cos2  0  + . ... 

236.  Some  applications  of  the  logarithmic  series,  The 
exponential  limit.  Let  z  be  any  complex  number,  and  h  a  real 
number  small  enough  to  ensure  that  |  hz  \  <  1.     Then 

log  (1  +  hz)  =  hz-\  (he?  +  i  (hz)s 

and  so 

log(l  -f  hz)  ,  '       . 

-^—h ;  =  z+<f>(h,z), 

where 

<p (h,  z)  =  - \he*  +  Ihrz- -  \h*z*  +  ..., 

]  <j>  (h,  z)\<\  hz2  \(l  +  \hz\  +  \  h?e* 1  +  ...)  = 

so  that  <f>  (h,  z)^~0  as  h-^0.     It  follows  that 

,.     log  (1+ As) 

hm    8  v  .    — -  =  z    (1). 

If  in  particular  we  suppose  h  =  1/n,  where  n  is  a  positive  integer, 
we  obtain 

lim   n  log  ( 1  +  -  ]  -■=  z, 

Jl->-30  \  11/ 

Jsi  i1 + w)  =  i^ exp  r log  (x + w) j  exp  ^ (2)" 

This  is  a  generalisation  of  the  result  proved  in  §  208  for  real 
values  of  z. 

From  (1)  we  can  deduce  some  other  results  which  we  shall 
require  in  the  next  section.  If  t  and  h  are  real,  and  h  is  sufficiently- 
small,  we  have 

log(l+te  +  fo)-log(l+fe)  =  1  j      /  hz    \ 

h  h     S  V        1  +  tz) 

which  tends  to  the  limit  e/(l  +  tz)  as  h^O.     Hence 

sli«g(i^»  =  r^ (3> 


and  so 


422      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [X 

We  shall  also  require  a  formula  for  the  differentiation  of 
(1  +  tz)m,  where  m  is  any  number  real  or  complex,  with  respect 
to  t  We  observe  first  that,  if  <£  (t)  =  ^  (t)  +  i%  (t)  is  a  complex 
function  of  t,  whose  real  and  imaginary  parts  $  (t)  and  x  (t) 
possess  derivatives,  then 

d  d 

jt  (exp  <t> )  =  jt  {(cos  X  +  i  sin  x)  exp  yjr] 

=  {(cos  x  +  *  sm  %)  ^'  +  (—  sm  %  +  *'  cos  X)  %'i  exP  ^ 
=  (i//  +  i^')  (cos  %  +  *  sin  x)  exP  ^ 

=  W  +  *'%')  exP  (^  +  *%)  =  </>'  exP  0. 
so  that  the  rule  for  differentiating  exp  </>  is  the  same  as  when  <f>  is 
real.     This  being  so  we  have 

^(1  +  tzT  =  jt exp  [m log (1  +  tz)} 

=  f^  t~  exp  {wi  log  (1  +  tz)} 

=  mz(l  +tz)m~1  (4). 

Here  both  (1  +  tz)m  and  (1  +  tz)m~l  have  their  principal  values 

237.     The  general  form  of  the  Binomial  Theorem.     We 

have  proved  already  (§  215)  that  the  sum  of  the  series 

is  (1  +  z)m  =  exp  \m  log  (1  +  z)},  for  all  real  values  of  m  and  all  real 
values  of  z  between  —  1  and  1.     If  an  is  the  coefficient  of  zn  then 

1, 


n+1 


whether  m  is  real  or  complex.  Hence  (Ex.  lxxx.  3)  the  series 
is  always  convergent  if  the  modulus  of  z  is  less  than  unity,  and  we 
shall  now  prove  that  its  sum  is  still  exp  [m  log  (1  +  z)},  i.e.  the 
principal  value  of  (1  +  z)m. 

It  follows  from  §  236  that  if  t  is  real  then 
~ -  (I  +  tz)m  =  mz(l+  tzfl-\ 


236,  237]       EXPONENTIAL,   AND   CIRCULAR   FUNCTIONS  423 

z  and  m  having  any  real  or  complex  values  and  each  side  having 
its  principal  value.     Hence,  if  <j>  (t)  =  (1  +  tz)m,  we  have 

<£<«>  (t)=m(m-l)...(m-n  +  1)  zn  (1  +  tz)m~n. 
This  formula  still  holds  if  t  =  0,  so  that 

w !         \n  / 
Now,  in  virtue  of  the  remark  made  at  the  end  of  §  164,  we  have 

#(i)-#W+f(0)+«^  +  ...  +  g^+*. 

where  ij„  =  —L— j\l  -  ()»->  £M  («)  <fc. 

But  if  z  =  r  (cos  6  +  i  sin  0)  then 

1 1  +  tz  |  =  V(l  +  2£r  cos  0  +  £V)  ^  1  -  tr, 
and  therefore 

|ro(m-l)...(m-n  +  l)|       fi  (1-t)™    , 
|jKnl<  "  (n-l)l  '   JoO^W^ 

\m(m-  l)...(m-»  +  l)|  (l-fl)71-1?-" 
(n  -  1)  !  '   (1  -  tfr)"  ' 

where  0  <  0  <  1  ;  so  that  (cf.  §  163) 


jm(m-l)...(m-rc  +  l)j 


(11-1)1 

Pn+i  _ 

j  m  — w  | 

Pn 

say.     But 


and   so   (Ex.  xxvu.  6)  /9n^»0,  and  therefore  i2n-*-0,  as  w^-oo. 
Hence  we  arrive  at  the  following  theorem. 

Theorem.     The  sum  of  the  binomial  series 

_      fm\         fm\    „ 

is  exp  {m  log  (1  +  z)},  where  the  logarithm  has  its  principal  value, 
for  all  values  of  m,  real  or  complex,  and  all  values  of  z  such  that 
z\<  1. 

A  more  complete  discussion  of  the  binomial  series,  taking 
account  of  the  more  difficult  case  in  which  \z\=  1,  will  be  found 
on  pp.  225  et  seq.  of  Bromwich's  Infinite  Series. 


424  THE    GENERAL   THEORY   OF   THE   LOGARITHMIC,  [X 

Examples  XCVIII.     1.     Suppose  to  real.     Then  since 


we  obtain 


log  (1  +  z)  =  £  log  ( 1  +  2?-  cos  6  +  r2)  +  i  arc  tan 

\l  +r  cos 


00  f  iyi\  f  /    t  sin  $    \  1 

2  (      I  2n=cxp  {hn  log  (1  +  2r  cos  #+r2)}  Cis  -!m  arc  tan  ( )  V 

o  W  {  \l  +  r  cos  6J) 

=  (1  +  2r  cos  6  +  r2)im  Cis  (m  arc  tan  f-I^^-\  I , 
v  ;  1  \l+»-cos0/J' 

all  the  inverse  tangents  lying  between  —\ir  and  \v.     In  particular,  if  we 
suppose  0=571-,  z  =  iV,  and  equate  the  real  and  imaginary  parts,  we  obtain 

1  _  f™\  r2+(^\ri-...  =  (l  +  r2fm  cos  (to  arc  tan  r), 

(?)  r "  (I)  ^  +  (?)  ''5 "  •"  =  (1  +r2)h"1  Sin  (m  ar° tan  r)- 

2.     Verify  the  formulae  of  Ex.  1  when  wi=l,  2,  3.     [Of  course  when  to  is 
a  positive  integer  the  series  is  finite.] 


3.  Prove  that  if  0  £r  <  1  then 

,     1.3  ,     1.3.5.7    . 
1—   — r2+  »•*  — 

2.4         2.4.6.8 

1  1.3.5,     1.3.5.7.0. 

_  r )-3  J j.5  _ 

2  2.4.6         2.4.6.8.10 
[Take  m=  —  \  in  the  last  two  formulae  of  Ex.  1.] 

4.  Prove  that  if  —  \tt<6<^tt  then 


V  I    2(l  +  r2)    /' 

//V(l4-0-l| 

V  1    2(l+r2)    /• 


"I 


cos  m8  =  coiimB  \  1  -  (  9  j  tan2<9  +  f      )  tan*  5 


sinm0  =  cos"l0-U     J  tan 5-  (     j  tan3<9  +  ... 

for  all  real  values  of  to.     [These  results  follow  at  once  from  the  equations 
cos  md  +  i  sin  md  =  (cos  6  +  i  sin  0)m = cos"1 6  (1  +  *  tan  0)"1.] 

5.  We  proved  (Ex.  lxxxi.  6),  by  direct  multiplication  of  series,  that 
/(to,  2)  =  2  I      j  s",  where  |  z  |  <1,  satisfies  the  functional  equation 

/(to,  2) /(to',  z)=f(m  +  m',  z). 

Deduce,  by  an  argument  similar  to  that  of  §  216,  and  without  assuming  the 
general  result  of  p.  423,  that  if  m  is  real  and  rational  then 

/  ( m,  z)  =  exp  {m  log  ( 1+  z)} . 

6.  If  z  and  /x  are  real,  and  —  1  <  z  <  1,  then 
*n  a»=cos  {fx  log  (1+2)}+*  sin  {M  log  (1  +2)}. 


EXPONENTIAL,    AND    CIRCULAR    FUNCTIONS  425 

MISCELLANEOUS   EXAMPLES  .OX   CHAPTER   X. 

1.  Show  that  the  real  part  of  ilog  (1_H)  is 

e(4&+l)nS/8  cos  1 1  (4£  +  1)  ^  log  2}, 

■where  k  is  any  integer. 

2.  If  a  cos  8+b  sin  8+c=0,  where  a,  b,  c  are  real  and  c2>a2  +  62,  then 

a  L       .-.■,         |c|  +  v'(c2-«2-62) 

where  wi  is  any  odd  or  any  even  integer,  according  as  c  is  positive  or  negative, 
and  a  is  an  angle  whose  cosine  and  sine  are  a/N/(a2  +  62)  and  bj^(a2  +  b-). 

3.  Prove  that  if  8  is  real  and  sin  6  sin  <£  =  1  then 

(f>  =  (£  +  ^)  7T  +  i  log  COt  |  (for  +  8), 

where  k  is  any  even  or  any  odd  integer,  according  as  sin  8  is  positive  or 
negative. 

4.  Show  that  if  x  is  real  then 

-j-  exp  {(a  +  ib)  x)  =  (a  +  26)  exp  {(a  +  ib)  x) , 


Deduce  the  results  of  Ex.  lxxxvii.  3. 

/-»  j 

5.  Show  that  if  a>0  then  /    exp  { -  (a  +  26)  a;}  dx = ^ ,  and  deduce  the 

Jo  a  +  ib' 

results  of  Ex.  lxxxvii.  5. 

6.  Show  that  if  (x/a)2  +  (y/b)2  =  l  is  the  equation  of  an  ellipse,  and/(#,  y) 
denotes  the  terms  of  highest  degree  in  the  equation  of  any  other  algebraic 
curve,  then  the  sum  of  the  eccentric  angles  of  the  points  of  intersection  of  the 
ellipse  and  the  curve  differs  by  a  multiple  of  2n  from 

-2  {log/ (a,  ib)  -  log  f  (a,  -ib)}. 
[The  eccentric  angles  are  given  by/(acosa,  6  sin  a)  -f-...=0  or  by 


/{*«(«+*)'    -^(M"3}  +  -=0) 


where  u  =  expia  ;  and  2a  is  equal  to  one  of  the  values  of  —  i  LogP,  where  P  is 
the  product  of  the  roots  of  this  equation.] 

7.     Determine  the  number  and  approximate  positions  of  the  roots  of  the 
equation  tan  z= az,  where  a  is  real. 

[We  know  already  (Ex.  xvn.  4)  that  the  equation  has  infinitely  many  real 
roots.     Now  let  z  =  x  +  iy,  and  equate  real  and  imaginary  parts.     AVe  obtain 

sin  2a;/ (cos  2.r  +  cosh  2y)  =  ax,     sinh  2y/(cos  2x  +  cosh  2y)  =  ay, 
so  that,  unless  x  or  y  is  zero,  we  have 

(sin  2*)/2.f=(sinh  2y)/2y. 


426      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [X 

This  is  impossible,  the  left-hand  side  being  numerically  less,  and  the  right- 
hand  side  numerically  greater  than  unity.  Thus  x=0  or  y  =  0.  If  y  =  0  we 
come  back  to  the  real  roots  of  the  equation.  If  x  =  0  then  tanh  y=ay.  It  is 
easy  to  see  that  this  equation  has  no  real  root  other  than  zero  if  a  ^  0  or 
a 2.1,  and  two  such  roots  if  0 <a <  1.  Thus  there  are  two  purely  imaginary 
roots  if  0<a<l  ;   otherwise  all  the  roots  are  real.] 

8.  The  equation  ta.nz=az  +  b,  where  a  and  b  are  real  and  b  is  not  equal 
to  zero,  has  no  complex  roots  if  a^.0.  If  a>0  then  the  real  parts  of  all  the 
complex  roots  are  numerically  greater  than  |  b/2a  | . 

9.  The  equation  tanz=a/s,  where  a  is  real,  has  no  complex  roots,  but 
has  two  purely  imaginary  roots  if  a<0. 

10.  The  equation  tan  z  =  a  tanh  cz,  where  a  and  c  are  real,  has  an  infinity 
of  real  and  of  purely  imaginary  roots,  but  no  complex  roots. 

11.  Show  that  if  x  is  real  then 

e«*cos  6#=!^  \an-  rlj  an~2b2+  LJ  an~464-  ...1 , 

where  there  are  ^(n  +  l)  or  ^(n  +  2)  terms  inside  the  large  brackets.  Find 
a  similar  series  for  eax  sin  bx. 

12.  If  n  (f>  (z,  n)  -*-  z  as  n-^-cc ,  then  {l+(f>  (z,  n)}n  -*-  exp  z. 

13.  If  <f>  (t)  is  a  complex  function  of  the  real  variable  t,  then 

[Use  the  formulae 

0  =  ty  +  %    lo§  0  =  5  loS  (^2  +  X2)  +  »' arc  tan  (x/V)-1 

14.  Transformations.  In  Ch.  Ill  (Exs.  xxi.  21  et  seq.,  and  Misc.  Exs. 
22  ct  seq.)  we  considered  some  simple  examples  of  the  geometrical  relations 
between  figures  in  the  planes  of  two  variables  z,  Z  connected  by  a  relation 
z=f(Z).  We  shall  now  consider  some  cases  in  which  the  relation  involves 
logarithmic,  exponential,  or  circular  functions. 

Suppose  firstly  that 

z  =  exp  (irZja),     Z=  {ajtt)  Log  z 

where  a  is  positive.  To  one  value  of  Z  corresponds  one  of  z,  but  to  one  of  z 
infinitely  many  of  Z.  If  x,  y,  r,  6  are  the  coordinates  of  z  and  X,  V,  li,  Q 
those  of  Z,  we  have  the  relations 

x=enX/a  cos  («■  Y\a\        y  =  e*x/a  sin  (n-  V/a), 

X=  (ajn)  log  r,  Y=  (adfn)  +  2ka, 

where  h  is  any  integer.  If  we  suppose  that  -  n<6^n,  and  that  Logs  has  its 
principal  value  logs,  then  &=0,  and  Z  is  confined  to  a  strip  of  its  plane  parallel 
to  the  axis  OX  and  extending  to  a  distance  a  from  it  on  each  side,  one  point 


EXPONENTIAL,    AND    CIRCULAR    FUNCTIONS  427 

of  this  strip  corresponding  to  one  of  the  whole  s-plane,  and  conversely.  By 
taking  a  value  of  Logs  other  than  the  principal  value  we  obtain  a  similar 
relation  between  the  2-plane  and  another  strip  of  breadth  2a  in  the  Z-plane. 

To  the  lines  in  the  Z- plane  for  which  X  and  Y  are  constant  correspond  the 
circles  and  radii  vectores  in  the  z-plane  for  which  r  and  6  are  constant.  To 
one  of  the  latter  lines  corresponds  the  whole  of  a  parallel  to  OX,  but  to  a 
circle  for  which  r  is  constant  corresponds  only  a  part,  of  length  2a,  of  a 
parallel  to  OY.  To  make  Z  describe  the  whole  of  the  latter  line  we  must 
make  z  move  continually  round  and  round  the  circle. 

15.  Show  that  to  a  straight  line  in  the  Z-plane  corresponds  an  equi- 
angular spiral  in  the  2-plane. 

16.  Discuss  similarly  the  transformation  z=c  cosh  (nZj a),  showing  in 
particular  that  the  whole  2-plane  corresponds  to  any  one  of  an  infinite 
number  of  strips  in  the  Z-plane,  each  parallel  to  the  axis  OX  and  of 
breadth  2a.     Show  also  that  to  the  line  X—Xq  corresponds  the  ellipse 

2 


)j    +ir:"1-'- 


\c  cosh  (7j-A'0/a)J         [c  sinh  {irX^a) 

and  that  for  different  values  of  X0  these  ellipses  form  a  confocal  system  ;  and 
that  the  lines  Y  —  Y0  correspond  to  the  associated  system  of  confocal  hyper- 
bolas. Trace  the  variation  of  z  as  Z  describes  the  whole  of  a  line  X=X0  or 
Y=  Y0  How  does  Z  vary  as  z  describes  the  degenerate  ellipse  and  hyperbola 
formed  by  the  segment  between  the  foci  of  the  confocal  system  and  the 
remaining  segments  of  the  axis  of  xl 

17.  Verify  that  the  results  of  Ex.  16  are  in  agreement  with  those  of  Ex.  14 
and  those  of  Ch.  Ill,  Misc.  Ex.  25.  [The  transformation  z  =  ccoah{TrZja) 
may  be  regarded  as  compounded  from  the  transformations 

e=cz1,     2i  =  K^2  +  (l/22)},     z2  =  exY>(nZla).] 

18.  Discuss  similarly  the  transformation  z=cta,nh(nZja),  showing  that 
to  the  lines  X=X0  correspond  the  coaxal  circles 

{x-c  coth  (27rX0/a)}2  +y2  =  c2  cosech2  (2ir XJa), 

and  to  the  lines  Y=  Y0  the  orthogonal  system  of  coaxal  circles. 

19.  The  Stereographic  and  Mercator's  Projections.  The  points  of  a 
unit  sphere  whose  centre  is  the  origin  are  projected  from  the  south  pole  (whose 
coordinates  are  0,  0,  —  1)  on  to  the  tangent  plane  at  the  north  pole.  The 
coordinates  of  a  point  on  the  sphere  are  £,  r/,  £,  and  Cartesian  axes  OX,  OY 
are  taken  on  the  tangent  plane,  parallel  to  the  axes  of  £  and  tj.  Show  that 
the  coordinates  of  the  projection  of  the  point  are 

*=2£/(l+a    y  =  2,,/(l-r  £), 

and  that  x  +  iy  =  2  tan  \  6  Cis  <j>,  where  <fr  is  the  longitude  (measured  from  the 
plane  t]  =  0)  and  6  the  north  polar  distance  of  the  point  on  the  sphere. 


428      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC,        [X 

This  projection  gives  a  map  of  the  sphere  on  the  tangent  plane,  generally- 
known  as  the  Stereographic  Projection.     If  now  we  introduce  a  new  complex 

variable 

Z=X+iY=  —i\oghz=  —i\og\{x  +  iy) 

so  that  X=<f>,  Y=  log  cot  1 6,  we  obtain  another  map  in  the  plane  of  Z, 
usually  called  Mercator's  Projection.  In  this  map  parallels  of  latitude  and 
longitude  are  represented  by  straight  lines  parallel  to  the  axes  of  X  and  Y 
respectively. 

20.     Discuss  the  transformation  given  by  the  equation 

fZ- 


*\Z-b, 

showing  that  the  straight  lines  for  which  x  and  y  are  constant  correspond  to 
two  orthogonal  systems  of  coaxal  circles  in  the  iT-plane. 

21.     Discuss  the  transformation 


(J(Z-a)  +  J(Z-b)] 
LOgl 7(&~«)  }• 


showing  that  the  straight  lines  for  which  x  and  y  are  constant  correspond  to 
sets  of  confocal  ellipses  and  hyperbolas  whose  foci  are  the  points  Z=a  and 
Z=b. 

[AVehave  J(Z-a)  +  J(Z-b)  =  s/(b-a)ex]i(    x+iy), 

s!(Z-  a)  —  *J(Z-  b)  =  sl(b  -  a)  exp  ( -  x  -  iy) ; 

and  it  will  be  found  that 

\Z-a\  +  \Z-b\  =  \b-a\ cosh  2x,     \Z-a\-\Z-b\  =  \b-a\cos 2y.\ 

22.  The  transformation  z=Z\  If  z—Z\  where  the  imaginary  power 
has  its  principal  value,  we  have 

exp  (log  r  +  iff)  =z  =  exp  (i  log  Z) = exp  (i  log  R  -  e), 

so  that  log  r  =  —  0,  6  =  log  R  +  2kir,  where  k  is  an  integer.  As  all  values  of  k 
give  the  same  point  z,  we  shall  suppose  that  k  =  0,  so  that 

logr=-e,     6  =  \ogR (1). 

The  whole  plane  of  Z  is  covered  when  R  varies  through  all  positive 
values  and  6  from  -  it  to  n  :  then  r  has  the  range  exp  (-  rr)  to  exp  n  and  6 
ranges  through  all  real  values.  Thus  the  Z-plane  corresponds  to  the  ring 
bounded  by  the  circles  »-=exp(-7r),  r  —  expn-;  but  this  ring  is  covered 
infinitely  often.  If  however  6  is  allowed  to  vary  only  between  -  n  and  tt. 
so  that  the  ring  is  covered  only  once,  then  R  can  vary  only  from  exp  ( -  it)  to 
exp  7r,  so  that  the  variation  of  Z  is  restricted  to  a  ring  similar  in  all  respects 
to  that  within  which  z  varies.  Each  ring,  moreover,  must  be  regarded  as 
having  a  barrier  along  the  negative  real  axis  which  z  (or  Z)  must  not  cross,  as 
its  amplitude  must  not  transgress  the  limits  —  it  and  tt. 


EXPONENTIAL,   AND    CIRCULAR    FUNCTIONS  429 

We  thus  obtain  a  correspondence  between  two  rings,  given  by  the  pair  of 

equations 

z  =  Z\     Z=z~\ 

where  each  power  has  its  principal  value.     To  circles  whose  centre  is  the 
origin  in  one  plane  correspond  straight  lines  through  the  origin  in  the  other. 

23.  Trace  the  variation  of  z  when  Z,  starting  at  the  point  exp  n,  moves 
round  the  larger  circle  in  the  positive  direction  to  the  point  —  exp  71- ,  along 
the  barrier,  round  the  smaller  circle  in  the  negative  direction,  back  along  the 
barrier,  and  round  the  remainder  of  the  larger  circle  to  its  original  position. 

24.  Suppose  each  plane  to  be  divided  up  into  an  infinite  series  of  rings 
by  circles  of  radii 

-(2»+l)»  -n  n  Sn  g(2»+l)  *, 

Show  how  to  make  any  ring  in  one  plane  correspond  to  any  ring  in  the 
other,  by  taking  suitable  values  of  the  powers  in  the  equations  z  =  Zl,  Z=z~l. 

25.  If  z=Z{,  any  value  of  the  power  being  taken,  and  Z  moves  along  an 
equiangular  spiral  whose  pole  is  the  origin  in  its  plane,  then  z  moves  along  an 
equiangular  spiral  whose  pole  is  the  origin  in  its  plane. 

26.  How  does  Z-zai,  where  a  is  real,  behave  as  z  approaches  the  origin 
along  the  real  axis.  \Z  moves  round  and  round  a  circle  whose  centre  is  the 
origin  (the  unit  circle  if  zai  has  its  principal  value),  and  the  real  and  imaginary 
parts  of  Z  both  oscillate  finitely.] 

27.  Discuss  the  same  question  for  Z=za  +  bi,  where  a  and  b  are  any  real 
numbers. 

28.  Show  that  the  region  of  convergence  of  a  series  of  the  type  2  anz»aiT 

-  » 

where  a  is  real,  is  an  angle,  i.e.  a  region  bounded  by  inequalities  of  the  type 
60 <  am  z  <  61      [The  angle  may  reduce  to  a  line,  or  cover  the  whole  plane.] 

29.  Level  Curves.  It  f{z)  is  a  function  of  the  complex  variable  z,  we 
call  the  curves  for  which  1/(2)  |  is  constant  the  level  curves  of  f(z).  Sketch 
the  forms  of  the  level  curves  of 

z  —  a      (concentric  circles),         (2 -a)  (z- b)         {Cartesian  ovals), 

(z-a)/(z  —  b)     (coaxal  circles),  exp  2  (straight  lines). 

30.  Sketch  the  forms  of  the  .  level  carves  of  (z  —  a)(z-b)(z-c), 
(l+z*/3+z2)/z.  [Some  of  the  level  curves  of  the  latter  function  are  drawn  in 
Fig.  59,  the  curves  marked  i-vii  corresponding  to  the  values 

•10,  2-v/3  =  -27,  -40,   1-00,  2-00,  2+N/3  =  373,  4'53 

of  1/(2)  I .  The  reader  will  probably  find  but  little  difficulty  in  arriving  at  a 
general  idea  of  the  forms  of  the  level  curves  of  any  given  rational  function  ; 
but  to  enter  into  details  would  carry  us  into  the  general  theory  of  functions 
of  a  complex  variable.] 


430      THE  GENERAL  THEORY  OF  THE  LOGARITHMIC, 


[X 


Fig.  59. 


Fig.  60. 


Fig.  61. 


EXPONENTIAL,    AND    CIRCULAR    FUNCTIONS 


431 


31.  Sketch  the  forms  of  the  level  curves  of  (i)  zexpz,  (ii)  sinz.  [See 
Fig.  60,  which  represents  the  level  curves  of  sin  z.  The  carves  marked  i-vni 
correspond  to  £  =  -35,  -50,  71,  1-00,  1-41,  2-00,  2-83,  4-00.] 

32.  Sketch  the  forms  of  the  level  curves  of  exp  z-c,  where  c  is  a  real 
constant.  [Fig.  61  shows  the  level  curves  of  [exp 2  — 1|,  the  curves  i-vn 
corresponding  to  the  values  of  k  given  by  \ogk=  -1-00,  --20,  — '05,  000, 
•05,  -20,  1-00.] 

33.  The  level  curves  of  sin  z-c,  where  c  is  a  positive  constant,  are 
sketched  in  Figs.  62,  63.  [The  nature  of  the  curves  differs  according  as 
to  whether  c<\  or  c>l.  In  Fig.  62  we  have  taken  c=-5,  and  the  curves 
i-viii  correspond  to  £=-29,  -37,  *50,  -87,  1-50,  2-60,  4-50,  779.  In  Fig.  63 
we  have  taken  c  =  2,  and  the  curves  i-vii  correspond  to  £="58,  1-00,  1"73, 
3-00,  5-20,  9-00,  15-59.  If  c  =  l  then  the  curves  are  the  same  as  those  of 
Fig.  60,  except  that  the  origin  and  scale  are  different.] 


Fig.  62. 


Fig.  63. 


34. 


Prove  that  if  0<<9<tt  then 

cos  8  + 1  cos  30  +  i  cos  58  +  ...=£  log  cot2  \8, 

sin0  +  Jsin3#+isin50  +  ...  =  !7r, 

and  determine  the  sums  of  the  series  for  all  other  values  of  6  for  which  they 
are  convergent.     [Use  the  equation 

*+^3+!s5+...=!iog(j±0 

where  2  =  cos  8+i  sin  8.  When  8  is  increased  by  it  the  sum  of  each  series 
simply  changes  its  sign.  It  follows  that  the  first  formula  holds  for  all  values 
of  8  save  multiples  of  it  (for  which  the  series  diverges),  while  the  sum  of  the 
second  series  is  Jtt  if  2kn  <8  <{2k +  1)  «-,  -Jtt  if  (2k  +  l)  n<8<(2k+2)  it, 
and  0  if  8  is  a  multiple  of  71-.] 


432  THE    LOGARITHMIC    AND    EXPONENTIAL    FUNCTIONS  [X 

35.  Prove  that  if  0< 6 <% n  then 

COS  6-\  COS  30  +  i  COS  50  —  . . .  =  j  77, 

sin  6-1  sin  30  +  \  sin  50  -  . . .  =  |  log  (sec  0  +  tan  0)2 ; 
and  determine  the  sums  of  the  series  for  all  other  values  of  0  for  which  they 
are  convergent. 

36.  Prove  that 

cos  0  cos  a  +  J?  cos  20  cos  2a  +  J  cos  30  cos  3a  +...=  —  j  log  {4  (cos  0  -  cos  a)2}, 
unless  0-a  or  0+a  is  a  multiple  of  2n. 

37.  Prove  that  if  neither  a  nor  b  is  real  then 
dx  log  (-a)— log  (-b) 


/„ 


'o    (x-a)(x-b)  a-b 

each  logarithm  having  its  principal  value.  Verify  the  result  when  a  =  ci, 
b=  —ei,  where  c  is  positive.  Discuss  also  the  cases  in  which  a  or  b  or  both 
are  real  and  negative. 

38.     Prove  that  if  a  and  /3  are  real,  and  #>0,  then 

dx  TTl 


/. 


0    X2-(a  +  ip)2      2  (a  +  ifty 
What  is  the  value  of  the  integral  when  j3<0  ? 

39.     Prove  that,  if  the  roots  of  Ax2  +  2Bx  +  C=0  have  their  imaginary 
parts  of  opposite  signs,  then 

dx  TTl 


! 


_„  Ax2+ 2Bx  +  C     s,l{B2  -AC)' 

the  sign  of  S/{B2  —  AC)  being  so  chosen  that  the  real  part  of  is/(B2-AC)}/Ai 
is  positive. 


APPENDIX   I 


(To  Chapters  III,  IV,  V) 

The  Proof  that  every  Equation  has  a  Root 

Let  Z=P(z)  =  a0zn  +  a1zn~1  +  ...  +  an 

be  a  polynomial  in  z,  with  real  or  complex  coefficients.  We  can  represent 
the  values  of  z  and  Z  by  points  in  two  planes,  which  we  may  call  the  2-plane 
and  the  Z-plane  respectively.  It  is  evident  that  if  z  describes  a  closed  path  y 
in  the  z-plane,  then  Z  describes  a  corresponding  closed  path  r  in  the  Z-plane. 
We  shall  assume  for  the  present  that  the  path  T  does  not  pass  through  the 
origin. 

To  any  value  of  Z  correspond  an  infinity  of  values  of  am  Z,  differing  by 
multiples  of  2n,  and  each  of  these  values  varies  continuously  as  Z  describes 
r.*     We  can  select  a  particular  value  of  am  Z  corresponding  to  each  point 


Fig.  A.  Fig.  B. 

of  r,  by  first  selecting  a  particular  value  corresponding  to  the  initial  value 
of  Z,  and  then  following  the  continuous  variation  of  this  value  as  Z  moves 
along  r.  We  shall,  in  the  argument  which  follows,  use  the  phrase  'the 
amplitude  of  Z'  and  the  formula  am  Z  to  denote  the  particular  value  of  the 
amplitude  of  Z  thus  selected.  Thus  &mZ  denotes  a  one- valued  and  con- 
tinuous function  of  X  and  T,  the  real  and  imaginary  parts  of  Z. 

*  It  is  here  that  we  assume  that  T  does  not  pass  through  the  origin. 
n.  28 


434 


APPENDIX    I 


When  Z,  after  describing  T,  returns  to  its  original  position,  its  amplitude 
may  be  the  same  as  before,  as  will  certainly  be  the  case  if  r  does  not  enclose 
the  origin,  like  path  (a)  in  Fig.  B,  or  it  may  differ  from  its  original  value  by 
any  multiple  of  2tt.  Thus  if  its  path  is  like  (6)  in  Fig.  B,  winding  once  round 
the  origin  in  the  positive  direction,  then  its  amplitude  will  have  increased 
by  2?r.  These  remarks  apply,  not  merely  to  r,  but  to  any  closed  contour  in 
the  if-plane  which  does  not  pass  through  the  origin.  Associated  with  any 
such  contour  there  is  a  number  which  we  may  call  '  the  increment  of  am  Z 
when  Z  describes  the  contour ',  a  number  independent  of  the  initial  choice  of 
a  particular  value  of  the  amplitude  of  Z. 

We  shall  now  prove  that  if  the  amplitude  of  Z  is  not  the  same  when.  Z 
returns  to  its  original  position,  then  the  path  of  z  must  contain  inside  or  on 
it  at  least  one  point  at  ivhich  Z=0. 

We  can  divide  y  into  a  number  of  smaller  contours  by  drawing  parallels 
to  the  axes  at  a  distance  Si  from  one  another,  as  in  Fig.  C*  If  there  is, 
on  the  boundary  of  any  one  of  these  contours,  a  point  at  which  Z=0, 
what  we  wish  to  prove  is  already  established.     We  may  therefore  suppose 


ffi    . ' 

" 

r   * —  * 
"I 

w                                I 

> 

Q 

I 

1 

-£ 

Fig.  C. 


Fig.  D. 


that  this  is  not  the  case.  Then  the  increment  of  &mZ,  when  z  describes 
y,  is  equal  to  the  sum  of  all  the  increments  of  am  Z  obtained  by  supposing 
z  to  describe  each  of  these  smaller  contours  separately  in  the  same  sense  as  y. 
For  if  z  describes  each  of  the  smaller  contours  in  turn,  in  the  same  sense, 
it  will  ultimately  (see  Fig.  D)  have  described  the  boundary  of  y  once,  and 
each  part  of  each  of  the  dividing  parallels  twice  and  in  opposite  directions. 
Thus  PQ  will  have  been  described  twice,  once  from  P  to  Q  and  once  from  Q 
to  P.  As  z  moves  from  P  to  Q,  am  Z  varies  continuously,  since  Z  does  not 
pass  through  the  origin  ;  and  if  the  increment  of  am  Z  is  in  this  case  6,  then 
its  increment  when  z  moves  from  Q  to  P  is  -  6  ;  so  that,  when  we  add 
up  the  increments  of  am  Z  due  to  the  description  of  the  various  parts  of  the 
smaller  contours,  all  cancel  one  another,  save  the  increments  due  to  the 
description  of  parts  of  y  itself. 

*  There  is  no  difficulty  in  giving  a  definite  rule  for  the  construction  of  these 
parallels:  the  most  obvious  course  is  to  draw  all  the  lines  x  =  k51,  y  =  kd1,  where 
7c  is  an  integer  positive  or  negative. 


APPENDIX   I  435 

Hence,  if  am  Z  is  changed  when  z  describes  y,  there  must  be  at  least  one 
of  the  smaller  contours,  say  yt ,  such  that  am  Z  is  changed  when  z  describes 
y1.  This  contour  may  be  a  square  whose  sides  are  parts  of  the  auxiliary 
parallels,  or  may  be  composed  of  parts  of  these  parallels  and  parts  of  the 
boundary  of  y.  In  any  case  every  point  of  the  contour  lies  in  or  on  the 
boundary  of  a  square  &x  whose  sides  are  parts  of  the  auxiliary  parallels  and 
of  length  5j. 

We  can  now  further  subdivide  yt  by  the  help  of  parallels  to  the  axes  at  a 
smaller  distance  82  from  one  another,  and  we  can  find  a  contour  y2,  entirely 
included  in  a  square  A2,  of  side  82  and  itself  included  in  Al5  such  that  amZ 
is  changed  when  z  describes  the  contour. 

Now  let  us  take  an  infinite  sequence  of  decreasing  numbers  8it  82>  ..., 
8m,  ...,  whose  limit  is  zero.*  By  repeating  the  argument  used  above,  we  can 
determine  a  series  of  squares  A1}  A2,  ...,  Am,  ...  and  a  series  of  contours  y1} 
72>  •••)  y»M  •••  such  that  (i)  Am  +  1  lies  entirely  inside  Am,  (ii)  ym  lies  entirely 
inside  A,„,  (iii)  amZ  is  changed  when  z  describes  ym. 

If  {x-,n,  2/m)  and  (xm  +  8m,  ym  +  8m)  are  the  lower  left-hand  and  upper  right- 
hand  corners  of  Am,  it  is  clear  that  xx,  x2,  ...,  xm)  ...  is  an  increasing  and 
#i  +  &ij  #2+ #2)  •••'  xm  +  8m,  •••  a  decreasing  sequence,  and  that  they  have  a 
common  limit  x0.  Similarly  ym  and  ym+8m  have  a  common  limit  y0,  and 
(x0,  y0)  is  the  one  and  only  point  situated  inside  every  square  A)U.  How- 
ever small  8  may  be,  we  can  draw  a  square  which  includes  (x0 ,  y0),  and  whose 
sides  are  parallel  to  the  axes  and  of  length  8,  and  inside  this  square  a  closed 
contour  such  that  am  Z  is  changed  when  z  describes  the  contour. 

It  can  now  be  shown  that 

P(z0+iy0)=0. 

For  suppose  that  P(xQ  +  iy0)  =  a,  where  |  a|  =  p>0.  Since  P(x  +  ty)  is  a  con- 
tinuous function  of  x  and  y,  we  can  draw  a  square  whose  centre  is  (^0,  #o) 
and  whose  sides  are  parallel  to  the  axes,  and  which  is  such  that 

I P  (?+iy)-P  (x0  +  iy0)\<hp 

at  all  points  x  +  iy  inside  the  square  or  on  its  boundary.     At  all  such  points 

P(x  +  iy)  =  a  +  (f>, 

where  |  0  |  <  hp.  Now  let  us  take  any  closed  contour  lying  entirely  inside 
this  square.  As  z  describes  this  contour,  Z=a  +  cp  also  describes  a  closed 
contour.  But  the  latter  contour  evidently  lies  inside  the  circle  whose  centre 
is  a  and  whose  radius  is  \p,  and  this  circle  does  not  include  the  origin. 
Hence  the  amplitude  of  Z  is  unchanged. 

But  this  contradicts  what  was  proved  above,  viz.  that  inside  each  square  A,u 
we  can  find  a  closed  contour  the  description  of  which  by  z  changes  amZ 
Hence  P  (x0  +  iy0)  =  0. 

*  We  may,  e.g.,  take  5m=o1/2m_1. 

28—2 


436  APPENDIX   I 

All  that  remains  is  to  show  that  we  can  always  find  some  contour  such  that 
am  Z  is  changed  when  z  describes  y.     Now 

\      a0z      a0z*  a0znJ 

We  can  choose  R  so  that 

I  a\  I  I  a2  I  I  an | 

\o^fE  + \o^\W  +  "'  +  \a0\Rn<  ' 

where  §  is  any  positive  number,  however  small ;  and  then,  if  y  is  the  circle 
whose  centre  is  the  origin  and  whose  radius  is  R,  we  have 

Z=a0zn(l  +  p), 

where  |  p  |  <  8,  at  all  points  on  y.  "We  can  then  show,  by  an  argument 
similar  to  that  used  above,  that  am(l+p)  is  unchanged  as  z  describes 
y  in  the  positive  sense,  while  am  zn  on  the  other  hand  is  increased  by  2mr. 
Hence  a,mZ  is  increased  by  2mr,  and  the  proof  that  Z—0  has  a  root  is 
completed. 

We  have  assumed  throughout  the  argument  that  neither  r,  nor  any  of  the 
smaller  contours  into  which  it  is  resolved,  passes  through  the  origin.  This 
assumption  is  obviously  legitimate,  for  to  suppose  the  contrary,  at  any  stage 
of  the  argument,  is  to  admit  the  truth  of  the  theorem. 

We  leave  it  as  an  exercise  to  the  reader  to  infer,  from  the  discussion 
which  precedes  and  that  of  §  43,  that  when  z  describes  any  contour  y  in  the 
positive  sense  the  increment  of  amZ  is  2kir,  where  k  is  the  number  of  roots 
of  Z—0  inside  y,  multiple  roots  being  counted  multiply. 

There  is  another  proof,  proceeding  on  different  lines,  which  is  often  given. 
It  depends,  however,  on  an  extension  to  functions  of  two  or  more  variables  of 
the  results  of  §§  102  et  seq. 

We  define,  precisely  on  the  lines  of  §  102,  the  upper  and  lower  bounds  of  a 
function  f(x,  y),  for  all  pairs  of  values  of  x  and  y  corresponding  to  any  point 
of  any  region  in  the  plane  of  (x,  y)  bounded  by  a  closed  curve.  And  we 
can  prove,  much  as  in  §  102,  that  a  continuous  function  /  (x,  y)  attains  its- 
upper  and  lower  bounds  in  any  such  region. 

Now  \Z\  =  \P(x  +  iy)\ 

is  a  positive  and  continuous  function  of  x  and  y.  If  m  is  its  lower  bound  for 
points  on  and  inside  y,  then  there  must  be  a  point  z0  for  which  \Z\  =  m,  and 
this  must  be  the  least  value  assumed  by  \Z\.  If  m=0,  then  P(20)=0,  and 
we  have  proved  what  we  want.     We  may  therefore  suppose  that  m>0. 

The  point  z0  must  lie  either  inside  or  on  the  boundary  of  y :  but  if  y  is 
a  circle  whose  centre  is  the  origin,  and  whose  radius  R  is  large  enough,  then 
the  last  hypothesis  is  untenable,  since  |  P  (z)  \  -*-  oo  as  |  z  \  ~»-  go  .  We  may 
therefore  suppose  that  z0  lies  inside  y. 


APPENDIX    I  437 

If  we  put  8==2o  +  £,  and  rearrange  P{z)  according-  to  powers  of  f,  we  obtain 
P(z)  =  P(z0)  +  A1£+A2(*  +  ...  +  AnC'1, 
say.     Let  Ak  be  the  first  of  the  coefficients  which  does  not  vanish,  and  let 
\Ak\  =  fi,  |  ( |  =p.     We  can  choose  p  so  small  that 

l^  +  ilp  +  Ufc  +  2|p2  +  ...  +  Mn|p"-fc<ip. 
Then  \P(z)-P(z0)-Ak£*\<hH.p\ 

and  |P(2)|<|P(2o)  +  Jsfs|+^pfc. 

Now  suppose  that  z  moves  round  the  circle  whose  centre  is  z0  and  radius  p. 
Then 

P(z0)  +  AUk 
moves  k  times  round  the  circle  whose  centre  is  P(z0)  and  radius  |  Ak£k\  =  p.pk, 
and  passes  k  times  through  the  point  in  which  this  circle  is  intersected  by 
the  line  joining  P{zq)  to  the  origin.     Hence  there  are  k  points  on  the  circle 
described  by  z  at  which  \  P  (z0)  +  A  h  £ k  |  =  |  P  (z0)  \  —  ppk  and  so 

\P  (z)\<\P  (z(i)\-lxpk  +  \p.pk  =  m-\ppk<m; 
and  this  contradicts  the  hypothesis  that  m  is  the  lower  bound  of  |  P  (z)  | . 
It  follows  that  m  must  be  zero  and  that  P  (z0)  =  0. 


EXAMPLES   ON  APPENDIX  I 

1.  Show  that  the  number  of  roots  of  f(z)  =  0  which  lie  within  a  closed 
contour  which  does  not  pass  through  any  root  is  equal  to  the  increment  of 

{\ogf(z)}/2ni 

when  z  describes  the  contour. 

2.  Show  that  if  R  is  any  number  such  that 

L^li  j_  L?2 1  ,        .  |«nl  ^i 
R   "*"  B?  +,,,+  Rn  ^  ' 

then  all  the  roots  of  zn  +  alzn~1  +  ...  +  an=0  are  in  absolute  value  less  than 
R.  In  particular  show  that  all  the  roots  of  zh  -  \Zz  -7  =  0  are  in  absolute 
value  less  than  2^. 

3.  Determine  the  numbers  of  the  roots  of  the  equation  z2P  +  az  +  b  =  0 
where  a  and  b  are  real  and  p  odd,  which  have  their  real  parts  positive  and 
negative.  Show  that  if  a>0,  6>0  then  the  numbers  are  p  —  1  and  ^  +  1 ;  if 
a<0,  6>0  they  are  p  +  l  and  p  —  1 ;  and  if  6<0  they  are  p  and  p.  Discuss 
the  particular  cases  in  which  a=0  or  6  =  0.     Verify  the  results  when  jo  =  l. 

[Trace  the  variation  of  &m(z2p+az  +  b)  as  z  describes  the  contour  formed 
by  a  large  semicircle  whose  centre  is  the  origin  and  whose  radius  is  R,  and 
the  part  of  the  imaginary  axis  intercepted  by  the  semicircle.] 

4.  Consider  similarly  the  equations 

s^v  +  az  +  b^O,     zi*-1  +  az  +  b  =  0,     zii  +  1  +  az  +  b  =  0. 


438  APPENDIX   I 

5.  Show  that  if  a  and  /3  are  real  then  the  numbers  of  the  roots  of  the 
equation  z2n  +  a2z2n~1+^2=0  which  have  their  real  parts  positive  and 
negative  are  n-1  and  n+1,  or  n  and  n,  according  as  n  is  odd  or  even. 

{Math.  Trip.  1891.) 

6.  Show  that  when  z  moves  along  the  straight  line  joining  the  points 
z=zu  z  =  z2,  from  a  point  near  zt  to  a  point  near  z2,  the  increment  of 


1  1 

+  — 


is  nearly  equal  to  tt. 


7.  A  contour  enclosing  the  three  points  z=zv  z=z2,  z=z3  is  denned  by 
parts  of  the  sides  of  the  triangle  formed  by  zu  z2,  z3,  and  the  parts  exterior 
to  the  triangle  of  three  small  circles  with  their  centres  at  those  points. 
Show  that  when  z  describes  the  contour  the  increment  of 

Z  —  Zy        Z  —  Z2        Z  —  £3 

is  equal  to  -  2  jr. 

8.  Prove  that  a  closed  oval  path  which  surrounds  all  the  roots  of  a  cubic 
equation  /  (s)  =  0  also  surrounds  those  of  the  derived  equation  f'(z)  =  0.  [Use 
the  equation 

/'  (*)  =/(*)  (t-V  +  A  +  7V)  . 

\Z  -  i,1        i,  —  c2        Z-  43/ 

where  zx,  z2,  z3  are  the  roots  of  f(z)=0,  and  the  result  of  Ex.  7.] 

9.  Show  that  the  roots  of/'(s)  =  0  are  the  foci  of  the  ellipse  which  touches 
the  sides  of  the  triangle  («l5  z2,  z3)  at  their  middle  points.  [For  a  proof  see 
Cesaro's  Elementares  Lehrbuch  der  algebraischen  Analysis,  p.  352.] 

10.  Extend  the  result  of  Ex.  8  to  equations  of  any  degree. 

11.  If  f(z)  and  (f>  (z)  are  two  polynomials  in  z,  and  y  is  a  contour  which 
does  not  pass  through  any  root  of  f(z),  and  |  cf>  (z)  |<|/(z)  |  at  all  points  on  7, 
then  the  numbers  of  the  roots  of  the  equations 

/(*)=o,  /(2)+<M*)=o 

which  lie  inside  y  are  the  same. 

12.  Show  that  the  equations 

ez  =  az,     ez  =  az2,     es=az3, 

where  a>e,  have  respectively  (i)  one  positive  root  (ii)  one  positive  and  one 
negative  root  and  (iii)  one  positive  and  two  complex  roots  within  the  circle 
1 2 1  =  1.  (Math.  Trip.  1910.) 


APPENDIX   II 

(To  Chapters  IX,  X) 

A  Note  on  Double  Limit  Problems 

In  the  course  of  Chapters  IX  and  X  we  came  on  several  occasions  into 
contact  with  problems  of  a  kind  which  invariably  puzzle  beginners  and 
are  indeed,  when  treated  in  their  most  general  forms,  problems  of  great 
difficulty  and  of  the  utmost  interest  and  importance  in  higher  mathematics. 

Let  us  consider  some  special  instances.     In  §  213  we  proved  that 

log  (1  +  x) = x - \x2  +  \x3 -  ..., 

where  —l<x£.\,  by  integrating  the  equation 

l/{l  +  t)  =  l-t+t2-... 

between  the  limits  0  and  x.     What  we  proved  amounted  to  this,  that 


rx    dt  rx  rx  rx 

\    - — =       dt-       tdt+ \    t2dt-...; 
JoHf      Jo        Jo  Jo 


or  in  other  words  that  the  integral  of  the  sum  of  the  infinite  series  1  — 1+ 12  —  . . ., 
taken  between  the  limits  0  and  x,  is  equal  to  the  sum  of  the  integrals  of  its 
terms  taken  between  the  same  limits.  Another  way  of  expressing  this  fact  is  to 
say  that  the  operations  of  summation  from  0  to  oo ,  and  of  integration  from 
0  to  x,  are  commutative  when  applied  to  the  function  (  —  l)ntn,  i.e.  that  it  does 
not  matter  in  what  order  they  are  performed  on  the  function. 

Again,  in  §  216,  we  proved  that  the  differential  coefficient  of  the  ex- 
ponential function 

exp  x  =  1  +.t,  +  H_,  +  ... 

is  itself  equal  to  exp  x,  or  that 


Dx[ 


-x+^]  +  ..}j  =  Dxl  +  Dxx  +  Dx^l  +...; 


440  APPENDIX    II 

that  is  to  say  that  the  differential  coefficient  of  the  sum  of  the  series  is  equal 
to  the  sum  of  the  differential  coefficients  of  its  terms,  or  that  the  operations  of 
summation  from  0  to  oo  and  of  differentiation  with  respect  to  x  are  commu- 
tative when  applied  to  xMjn\. 

Finally  we  proved  incidentally  in  the  same  section  that  the  function 
exp  x  is  a  continuous  function  of  x,  or  in  other  words  that 

lim('l+a;+!7+...)=l+£+!j  +  ...=lim  l+lima'+lim  §-.  +  ...; 

i.e.  that  the  limit  of  the  sum  of  the  series  is  equal  to  the  sum  of  the  limits  of 
the  terms,  or  that  the  sum  of  the  series  is  continuous  for  x=$,  or  that  the 
operations  of  summation  from  0  to  oo  and  of  making  x  tend  to  £  are  com- 
mutative when  applied  to  xnjn !. 

In  each  of  these  cases  we  gave  a  special  proof  of  the  correctness  of  the 
result.  We  have  not  proved,  and  in  this  volume  shall  not  prove,  any  general 
theorem  from  which  the  truth  of  any  one  of  them  could  be  inferred  im- 
mediately. In  Ex.  xxxvn.  1  we  saw  that  the  sum  of  a  finite  number  of  con- 
tinuous terms  is  itself  continuous,  and  in  §  113  that  the  differential  coefficient 
of  the  sum  of  a  finite  number  of  terms  is  equal  to  the  sum  of  their  differential 
coefficients  ;  and  in  §  160  we  stated  the  corresponding  theorem  for  integrals. 
Thus  we  have  proved  that  in  certain  circumstances  the  operations  symbolised 

by 

lim...,     Dx...,     j...dx 

are  commutative  with  respect  to  the  operation  of  summation  of  a  finite  number 
of  terms.  And  it  is  natural  to  suppose  that,  in  certain  circumstances  which 
it  should  be  possible  to  define  precisely,  they  should  be  commutative  also  with 
respect  to  the  operation  of  summation  of  an  infinite  number.  It  is  natural  to 
suppose  so :  but  that  is  all  that  we  have  a  right  to  say  at  present. 

A  few  further  instances  of  commutative  and  non-commutative  operations 
may  help  to  elucidate  these  points. 

(1)  Multiplication  by  2  and  multiplication  by  3  are  always  commutative, 

for 

2x3x.r=3x2xA- 
for  all  values  of  x. 

(2)  The  operation  of  taking  the  real  part  of  z  is  never  commutative  with 
that  of  multiplication  by  i,  except  when  z=0  ;  for 

i  x  R  (x  +  iy)  =  ix,     R,  {i  x  (x  +  iy)}  ■=  -y. 

(3)  The  operations  of  proceeding  to  the  limit  zero  with  each  of  two 
variables  x  and  y  may  or  may  not  be  commutative  when  applied  to  a 
function  f(x,y).     Thus 

lim  {lim  (x+y)}  =  \im  x  =  0,     lim  { lim (.r  +  ?/)}  =  limy =0; 


APPENDIX    II  441 

but  on  the  other  hand 

lira' — "M  =  lim   -    =  lirnl  =  l, 

lim  (  lim  ^£)  =  lim  — ^  =  lim  ( - 1)=  - 1. 

(4)     The  operations  2...,  lim...  may  or  may  not  be  commutative.      Thus 
l        x-*-l 

if  .r-^1  through  values  less  than  1  then 

lim^  2V -#4=lim  log(l +o;)  =  log2, 

x+\  I  l      n  )      x^-l 

g  J  lim  ^__AL  £»L  =       2  ^— ^      =log  2  ; 


2 

1 

but  on  the  other  hand 


lim  J2  (^'l-a;'l  +  1)l=lim{(l-^)  +  (.r-^2)  +  ...}  =  liml  =  l, 

2  \lim(xn-xn  +  1)\  =  2  (1  -  l)  =  0  +  0  +  0-K..  =  0. 
l   \x-*-l  J       l 

The  preceding  examples  suggest  that  there  are  three  possibilities  with 
respect  to  the  commutation  of  two  given  operations,  viz. :  (1)  the  operations 
may  always  be  commutative  ;  (2)  they  may  never  be  commutative,  except  in 
very  special  circumstances ;  (3)  they  may  be  commutative  in  most  of  the  ordinary 
cases  which  occur  practically. 

The  really  important  case  (as  is  suggested  by  the  instances  which  we 
gave  from  Ch.  IX)  is  that  in  which  each  operation  is  one  which  involves 
a  passage  to  the  limit,  such  as  a  differentiation  or  the  summation  of  an 
infinite  series  :  such  operations  are  called  limit  operations.  The  general 
question  as  to  the  circumstances  in  which  two  given  limit  operations  are 
commutative  is  one  of  the  most  important  in  all  mathematics.  But  to 
attempt  to  deal  with  questions  of  this  character  by  means  of  general  theorems 
would  carry  us  far  beyond  the  scope  of  this  volume. 

"We  may  however  remark  that  the  answer  to  the  general  question  is  on 
the  lines  suggested  by  the  examples  above.  If  L  and  L'  are  two  limit 
operations  then  the  numbers  LL'z  and  L'Lz  are  not  generally  equal,  in  the 
strict  theoretical  sense  of  the  word  'general'.  We  can  always,  by  the  exercise 
of  a  little  ingenuity,  find  z  so  that  LL'z  and  L'Lz  shall  differ  from  one  another. 
But  they  are  equal  generally,  if  we  use  the  word  in  a  more  practical  sense, 
viz.  as  meaning  'in  a  great  majority  of  such  cases  as  are  likely  to  occur 
naturally ''  or  in  ordinary  cases. 


442  APPENDIX    II 

Of  course,  in  an  exact  science  like  pure  mathematics,  we  caunot  be  satisfied 
with  an  answer  of  this  kind  ;  and  in  the  higher  branches  of  mathematics  the 
detailed  investigation  of  these  questions  is  an  absolute  necessity.  But  for 
the  present  the  reader  may  be  content  if  he  realises  the  point  of  the  remarks 
which  we  have  just  made.  In  practice,  a  result  obtained  by  assuming  that 
two  limit-operations  are  commutative  is  probably  true  :  it  at  any  rate  affords 
a  valuable  suggestion  as  to  the  answer  to  the  problem  under  consideration. 
But  an  answer  thus  obtained  must,  in  default  of  a  further  study  of  the  general 
question  or  a  special  investigation  of  the  particular  problem,  such  as  we  gave 
in  the  instances  which  occurred  in  Ch.  IX,  be  regarded  as  suggested  only  and 
not  proved. 

Detailed  investigations  of  a  large  number  of  important  double  limit 
problems  will  be  found  in  Bromwich's  Infinite  Series. 


APPENDIX  III 
(To  §  158  and  Chapter  IX) 

The  circular  functions 

The  reader  will  find  it  an  instructive  exercise  to  work  out  the  theory  of 
the  circular  functions,  starting  from  the  definition 

dt 


(1)  y—y  (^)  =  arc  tan  x  —  \    - — 

.'o  1  + 


Df  * 
t2 


The  equation  (1)  defines  a  unique  value  of  y  corresponding  to  every  real 
value  of  as.  As  y  is  continuous  and  strictly  increasing,  there  is  an  inverse 
function  x=x  (y),  also  continuous  and  steadily  increasing.     We  write 

(2)  x = x  (y)  =  tan  y.  Df . 

If  we  define  it  by  the  equation 


Df. 


<3>  **-£&. 

then  this  function  is  defined  for  —  \tc  <.y  <\n. 

"We  write  further 

1  x 

(4)  cosy=—= ==,     siny=  — ==,  Df. 

S/l  +  x1  \/l  +  x2 

where  the  square  root  is  positive;  and  we  define  cosy  and  siny,  when  y  is  -  £  n 
or  ^7r,  so  that  the  functions  shall  remain  continuous  for  those  values  of  y. 
Finally  we  define  cosy  and  siny,  outside  the  interval  (  —  ^ir,  hir),  by 

(5)  tan  (y  +  7r)  =  tany,  cos  (y  +  n)=  -cosy,  sin  (y  +  n)=  —siny.    Df. 
We  have  thus  defined  cosy  and  siny  for  all  values  of  y,  and  tany  for  all 

values  of  y  other  than  odd  multiples  of  \tt.     The  cosine  and  sine  are  continuous 
for  all  values  of  y,  the  tangent  except  at  the  points  where  its  definition  fails. 
The  further  development  of  the  theory  depends  merely  on  the  addition 
formulae.     Write 

1—  X\X% 

and  transform  the  equation  (1)  by  the  substitution 

Xi+u               t-x\ 
t=i ,     u=- —J. 

1—  #!%'  l+Xit 

We  find 

Xi+.v2        f%2       du         fa>i    du         fa    du 

arc  tan =  / =  /      - h  +  / h 

1-^1  %2        J  -Xll  +  U-        JO      1  +  U*       JO     l+«2 

=  arc  tan  X\  +  arc  tan  x2. 

*  These  letters  at  the  end  of  a  line  indicate  that  the  formulae  which  it  contains 
are  definitions. 


444  APPENDIX   III 

From  this  we  deduce 

tanyi+tanyo  . 

(-6>  ta"^+^-l-tanyitanl> 

an  equation  proved  in  the  first  instance  only  when  yu  y2,  and  y\+y<i  lie  in 
(  — -^rr,  \n),  but  immediately  extensible  to  all  values  of  yx  and  y2  by  means  of 
the  equations  (5). 

From  (4)  and  (6)  we  deduce 

cos  (yi  +y2)  =  ±  (cos  y1  cos  y2  —  sin  y1  sin  y2). 
To  determine  the  sign  put  y2=0.  The  equation  reduces  to  cosy1=±cosyl, 
which  shows  that  the  positive  sign  must  be  chosen 'for  at  least  one  value  of  y2, 
viz.  #2  =  0-  It  follows  from  considerations  of  continuity  that  the  positive  sign 
must  be  chosen  in  all  cases.  The  corresponding  formula  for  sin(y1+y2)  may 
be  deduced  in  a  similar  manner. 

The  formulae  for  differentiation  of  the  circular  functions  may  now  be  de- 
duced in  the  ordinary  way,  and  the  power  series  derived  from  Taylor's 
Theorem. 

An  alternative  theory  of  the  circular  functions  is  based  on  the  theory  of 
infinite  series.  An  account  of  this  theory,  in  which,  for  example,  cos.v  is 
defined  by  the  equation 

COS x=l- -+--.., 

will  be  found  in  Whittaker  and  Watson's  Modern  Analysis  (Appendix  A). 


APPENDIX  IV 

The  infinite  in  analysis  and  geometry 

Some,  though  not  all,  systems  of  analytical  geometry  contain  'infinite' 
elements,  the  line  at  infinity,  the  circular  points  at  infinity,  and  so  on.  The 
object  of  this  brief  note  is  to  point  out  that  these  concepts  are  in  no  way 
dependent  upon  the  analytical  doctrine  of  limits. 

In  what  may  be  called  '  common  Cartesian  geometry ',  a  point  is  a  pair  of 
real  numbers  (x,  y).  A  line  is  the  class  of  points  which  satisfy  a  linear  relation 
ax  +  by  +  c = 0,  in  which  a  and  6  are  not  both  zero.  There  are  no  infinite  elements, 
and  two  lines  may  have  no  point  in  common. 

In  a  system  of  real  homogeneous  geometry  a  point  is  a  class  of  triads  of 
real  numbers  (x,  y,  z),  not  all  zero,  triads  being  classed  together  when  their 
constituents  are  proportional.  A  line  is  a  class  of  points  which  satisfy  a  linear 
relation  ax  +  by  +  cz  =  0,  where  a,  b,  c  are  not  all  zero.  In  some  systems  one 
point  or  line  is  on  exactly  the  same  footing  as  another.  In  others  certain 
'  special '  points  and  lines  are  regarded  as  peculiarly  distinguished,  and  it  is  on 
the  relations  of  other  elements  to  these  special  elements  that  emphasis  is  laid. 
Thus,  in  what  may  be  called  'real  homogeneous  Cartesian  geometry',  those 
points  are  special  for  which  2  =  0,  and  there  is  one  special  line,  viz.  the  line 
2=0.     This  special  line  is  called  'the  line  at  infinity'. 

This  is  not  a  treatise  on  geometry,  and  there  is  no  occasion  to  develop  the 
matter  in  detail.  The  point  of  importance  is  this.  The  infinite  of  analysis 
is  a  '  limiting '  and  not  an  '  actual '  infinite.  The  symbol  '  <x> '  has,  throughout 
this  book,  been  regarded  as  an  'incomplete  symbol',  a  symbol  to  which  no 
independent  meaning  has  been  attached,  though  one  has  been  attached  to 
certain  phrases  containing  it.  But  the  infinite  of  geometry  is  an  actual  and 
not  a  limiting  infinite.  The  'line  at  infinity'  is  a  line  in  precisely  the  same 
sense  in  which  other  lines  are  lines. 

It  is  possible  to  set  up  a  correlation  between  'homogeneous'  and  'common' 
Cartesian  geometry  in  which  all  elements  of  the  first  system,  the  special 
elements  excepted,  have  correlates  in  the  second.  The  line  ax  +  by  +  cz  —  0,  for 
example,  corresponds  to  the  line  ax  +  by  +  c  =  0.  Every  point  of  the  first  line 
has  a  correlate  on  the  second,  except  one,  viz.  the  point  for  which  2=0. 
When  (x,  y,  z)  varies  on  the  first  line,  in  such  a  manner  as  to  tend  in  the  limit 
to  the  special  point  for  which  2=0,  the  corresponding  point  on  the  second  line 
varies  so  that  its  distance  from  the  origin  tends  to  infinity.  This  correlation 
is  historically  important,  for  it  is  from  it  that  the  vocabulary  of  the  subject 
has  been  derived,  and  it  is  often  useful  for  purposes  of  illustration.  It  is  how- 
ever no  more  than  an  illustration,  and  no  rational  account  of  the  geometrical 
infinite  can  be  based  upon  it.  The  confusion  about  these  matters  so  prevalent 
among  students  arises  from  the  fact  that,  in  the  commonly  used  text  books  of 
analytical  geometry,  the  illustration  is  taken  for  the  reality. 


CAMBRIDGE  !   PRINTED  BY 

J.  B.  PEACE,  M.A., 
AT  THE  UNIVERSITY  PRESS 


*^*oD^4y 


VSE 


*Rom 


f04jrfS****nD 


U.  C.  BERKELEY  LIBRARIES 


CDbi3«n33a 


*t 


,-\ 


y 


THE  UNIVERSITY  OF  CALIFORNIA  LIBRARY