Skip to main content

Full text of "Elements of theoretical physics"

See other formats


THE  LIBRARY 

OF 

THE  UNIVERSITY 
OF  CALIFORNIA 

LOS  ANGELES 


DEC  11  1976 


THEORETICAL   PHYSICS 


ELEMENTS 

OF 


THEORETICAL    PHYSICS 


BY 

DR.  C.  CHRISTIANSEN 

PROFESSOR   OF   PHYSICS   IN  THE   UNIVERSITY   OF  COPENHAGEN 


TRANSLATED  INTO  ENGLISH  BY 

W.  F.  MAGIE,  PH.D. 

PROFESSOR   OF  PHYSICS   IN   PRINCETON   UNIVERSITY 


MACMILLAN    AND    CO,    LIMITED 

NEW  YORK  :    THE  MACMILLAN  COMPANY 

1897 

All  rights  resei~eed 


GLASGOW  :     PRINTED   AT  THE   UNIVERSITY   PRESS 
BY   ROBERT   MACLEHOSE  AND   CO. 


Physii 


TRANSLATOR'S  PREFACE. 

THE  treatise  of  Professor  Christiansen,  of  which  a  translation  is  here 
given,  presents  the  fundamental  principles  of  Theoretical  Physics, 
and  develops  them  so  far  as  to  bring  the  reader  in  touch  with 
much  of  the  new  work  that  is  being  done  in  that  subject.  It  is 
not  in  every  respect  exhaustive,  but  it  is  stimulating  and  informing, 
and  furnishes  a  view  of  the  whole  field,  which  will  facilitate  the 
reader's  subsequent  progress  in  special  parts  of  it.  The  need  of  such 
a  book,  in  which  the  various  branches  of  the  subject  are  developed 
in  connection  with  one  another  and  in  a  consistent  notation,  has 
been  long  felt  by  both  teachers  and  students. 

The  thanks  of  the  translator  are  due  to  Professor  Christiansen  for 
his  courtesy  in  permitting  the  use  of  his  book.  The  translation  was 
made  from  the  German  of  Miiller.  The  first  draft  of  it  was 
prepared  by  the  translator's  wife,  without  whose  aid  the  task  might 
never  have  been  accomplished. 

W.  F.  MAGIE. 

PRINCETON  UNIVERSITY,  September,  1896. 


TABLE   OF   CONTENTS. 


PAGE 

INTRODUCTION,       -  1 


CHAPTER  I. 
GENERAL  THEORY   OF   MOTION. 

SECTION 

I.     Freely  Falling  Bodies,    -  5 

II.     The  Motion  of  Projectiles,  7 

III.  Equations  of  Motion  for  a  Material  Point,   -  8 

IV.  The  Tangential  and  Normal  Forces,                                            -  13 
V.     Work  and  Kinetic  Energy,    -  14 

VI.     The  Work  Done  on  a  Body  during  its  Motion  in  a  Closed 

Path,    -  16 

VII.     The  Potential,          ....                                     -         -  21 

VIII.     Constrained  Motion,  24 

IX.     Kepler's  Laws,         -         -  27 

X.     Universal  Attraction,                                                                      -  30 

XI.     Universal  Attraction  (continued},   -  31 

XII.     The  Potential  of  a  System  of  Masses,  -                  ...  34 

XIII.  Examples.     Calculation  of  Potentials,    -                                    -  36 

XIV.  Gauss's  Theorem.     The  Equations  of  Laplace  and  Poisson,  -  41 
XV.     Examples    of    the    Application    of    Laplace's    and    Poisson's 

Equations,    -------         ---46 

XVI.     Action  and  Reaction.     On  the  Molecular  and  Atomic  Structure 

of  Bodies,.  - -         -  48 

XVII.     The  Centre  of  Gravity,  -        -.-                *        .        -        -        -  50 


CONTENTS. 


SECTION 

PAGE 

XVIII. 

A  Material  System, 

53 

XIX. 

Moment  of  Momentum,  -         -         -         -        -        - 

55 

XX. 

The  Energy  of  a  System  of  Masses, 

56 

XXI. 

Conditions  of  Equilibrium.     Rigid  Bodies,     -         -         - 

58 

XXII. 

Rotation  of  a  Rigid  Body.     The  Pendulum, 

60 

CHAPTER  II. 

THE  THEORY  OF  ELASTICITY. 

XXIII. 

62 

XXIV. 

Components  of  Stress,     -         -                  - 

64 

XXV. 

Relations  among  the  Components  of  Stress, 

67 

XXVI. 

The  Principal  Stresses,   - 

69 

XXVII. 

Faraday's  Views  on  the  Nature  of  Forces  Acting  at  a 

Distance,     -                                     

72 

XXVIII. 

Deformation,    --------- 

74 

XXIX. 

Relations  between  Stresses  and  Deformations, 

79 

XXX. 

Conditions  of  Equilibrium  of  an  Elastic  Body, 

82 

XXXI. 

Stresses  in  a  Spherical  Shell, 

83 

XXXII. 

Torsion,    

85 

XXXIII. 

Flexure,  ---------- 

87 

XXXIV. 

Equations  of  Motion  of  an  Elastic  Body,       - 

89 

XXXV. 

Plane  Waves  in  an  Infinitely  Extended  Body, 

90 

XXXVI. 

Other  Wave  Motions,      -         -         -         - 

93 

XXXVII. 

Vibrating  Strings,  -         -         -         .... 

95 

XXXVIII. 

Potential  Energy  of  an  Elastic  Body,    -         -        -        - 

96 

CHAPTER   III. 
EQUILIBRIUM   OF   FLUIDS. 

XXXIX.     Conditions  of  Equilibrium,      ... 
XL.     Examples  of  the  Equilibrium  of  Fluids, 


101 


CHAPTER  IV. 
MOTION   OF  FLUIDS. 

XLI.     Euler's  Equations  of  Motion, 
XLII.     Transformation  of  Euler's  Equations, 


103 
106 


CONTENTS. 


IX 


SECTION 

XLIII. 

XLIV. 

XLV. 

XL  VI. 


Vortex  Motions  and  Currents  in  a  Fluid, 
Steady  Motion  with  Velocity-Potential, 
Lagrange's  Equations  of  Motion,    - 
Wave  Motions, 


PAGE 

107 
109 
111 
112 


CHAPTER  V. 
INTERNAL   FRICTION. 

XL VII.     Internal  Forces, 

XLVIII.     Equations  of  Motion  of  a  Viscous  Fluid, 
XLIX.     Flow  through  a  Tube  of  Circular  Cross  Section, 


115 

118 
119 


CHAPTER  VI. 
CAPILLARITY. 

L.     Surface  Energy,       - 
LI.     Conditions  of  Equilibrium,      - 
LII.     Capillary  Tubes,      - 


121 
123 
125 


CHAPTER  VII. 
ELECTROSTATICS. 

LIIL     Fundamental  Phenomena  of  Electricity,         -         -         -  127 

LIV.     Electrical  Potential, 128 

LV.     The  Distribution  of  Electricity  on  a  Good  Conductor.  -  130 
LVI.     The  Distribution  of  Electricity  on  a  Sphere  and  on  an 

Ellipsoid,    -         -                                                                -  132 

LVII.     Electrical  Distribution,    -                  .....  135 

LVI  1 1.     Complete  Distribution, -         -  139 

LIX.     Mechanical  Force  Acting  on  a  Charged  Body,       -         -  141 

LX.     Lines  of  Electrical  Force, 143 

LXI.     Electrical  Energy,  -         -         -        -        -         -         -         -  145 

LXII.     A  System  of  Conductors, 147 

LXIII.     Mechanical  Forces,                    150 

LXIV.     The  Condenser  and  Electrometer, 151 

LXV.     The  Dielectric, 155- 


CONTENTS. 


SECTION 

PAGE 

LXVI.     Conditions  of  Equilibrium, 157 

LXVII.     Mechanical  Force  and  Electrical  Energy  in  the  Dielectric,     158 

CHAPTER   VIII. 
MAGNETISM. 

LXVIII.     General  Properties  of  Magnets, 163 

LXIX.     The  Magnetic  Potential,       ------  166 

LXX.     The  Potential  of  a  Magnetized  Sphere,        -         -         -  168 

LXXL     The  Forces  which  Act  on  a  Magnet,  -         -         -         -  169 

LXXII.     Potential  Energy  of  a  Magnet,    -         -        -        -         -  171 

.  LXXIII.     Magnetic  Distribution,          -         -         .        _        .'        -  173 

LXXIV.     Lines  of  Magnetic  Force, 174 

LXXV.     The  Equation  of  Lines  of  Force,          -         -         -         -  178 

LXXVI.     Magnetic  Induction,      - 179 

LXXVII.     Magnetic  Shells,    -         -         .         .        fc        .         .  180 

CHAPTEE  IX. 

ELECTEO-MAGNETISM. 

LXX VIII.     Biot  and  Savart's  Law,        -        .        „       _        .  j84 

LXXIX.     Systems  of  Currents,    -         -        .         ,        .        .         -  186 

LXXX.     The  Fundamental  Equations  of  Electro-Magnetism,   -  188 

LXXXI.     Systems  of  Currents  in  General,           -         -         .         -  190 

LXXXII.     The  Action  of  Electrical  Currents  on  each  other,       -  192 

LXXXIII.     The  Measurement  of  Current-Strength  on  the  Quantity 

of  Electricity,  -         ..        .        .        .         .        _        '-194 

LXXXI V.     Ohm's  Law  and  Joule's  Law,       .        „        .        .        -  197 


CHAPTER  X. 
INDUCTION. 

LXXXV.     Induction,     -         -        .... 
LXXXVI.     Coefficients  of  Induction,      -    '    . 
LXXXVII.     Measurement  of  Resistance, 
LXXXVIII.     Fundamental  Equations  of  Induction, 
LXXXIX.     Electro-Kinetic  Energy, 
XC.     Absolute  Units,    - 


202 
205 


-  210 

-  211 


CONTENTS. 


CHAPTER  XL 
ELECTRICAL  OSCILLATIONS. 


SECTION 

XCI. 


Oscillations  in  a  Conductor,     ------ 

XCII.     Calculation  of  the  Period,        ------ 

XCI II.     The    Fundamental    Equations   for   Electrical   Insulators 
or  Dielectrics,     -------- 

XCIV.     Plane  Waves  in  the  Dielectric,        - 

XCV.     The  Hertzian  Oscillations, 

XCVI.     Poynting's  Theorem, 224 


215 

217 

219 
221 
223 


CHAPTER  XII. 

REFRACTION  OF  LIGHT  IN  ISOTROPIC  AND  TRANSPARENT 
BODIES. 

XCVII.     Introduction,    -         - 

XCVIII.     Fresnel's  Formulas, 

XCIX.     The  Electro-Magnetic  Theory  of  Light, 

Equations  of  the  Electro-Magnetic  Theory  of  Light, 

Refraction  in  a  Plate, 242 

Double  Refraction,  ----....     246 
Discussion  of  the  Velocities  of  Propagation,  -         -         -     249 

CIV.     The  Wave  Surface, 251 

The  Wave  Surface  (continued),        -         -         ...     254 

The  Direction  of  the  Rays, 256 

Uniaxial  Crystals,    -         -         - 259 


C. 

or. 

GIL 
GUI. 


CV. 

CVI. 

CVII. 


229 
231 
235 
237 


CVIII.     Double  Refraction  at  the  Surface  of  a  Crystal,      -         -     261 
CIX.     Double  Refraction  in  Uniaxial  Crystals,          -         -         -     264 


CHAPTER   XIII. 
THERMODYNAMICS. 

CX.     The  State  of  a  Body,      -------  266 

CXI.     Ideal  Gases,      . 270 

CXII.     Cyclic  Processes, -         -  272 

CXIII.     Carnot's  and  Clausius'  Theorem, 274 

CXIV.     Application  of  the  Second  Law, 279 


Xll 


CONTENTS. 


SECTION  PAGE 

CXV.  The  Differential  Coefficients, -  280 

CXVI.  Liquids  and  Solids,           -                                                       -  281 

CXVII.  The  Development  of  Heat  by  Change  of  Length,          -  282 

CXVITI.  Van  der  Waal's  Equation  of  State,         -                           -  283 

CXIX.  Saturated  Vapours,                             290 

CXX.  The  Entropy,  -                                                                          -  292 

CXXI.  Dissociation,     ---------  295 


CHAPTER  XIV. 
CONDUCTION  OF  HEAT. 

CXXII.     Fourier's  Equation,  -  298 

CXXIII.     Steady  State,   -  -  300 

CXXIV.     The  Periodic  Flow  of  Heat  in  a  given  Direction,          -  301 

CXXV.     A  Heated  Surface,  -        -  303 

CXXVI.     The  Flow  of  Heat  from  a  Point, 304 

CXXVII.     The  Flow  of  Heat  in  an  Infinitely  Extended  Body,      -  305 

CXXVIII.     The  Formation  of  Ice,     -  -  307 

CXXIX.     The  Flow  of  Heat  in  a  Plate  whose  Surface  is  kept  at 

a  Constant  Temperature, 308 

CXXX.     The  Development  of   Functions  in  Series  of  Sines  and 

Cosines, -  312 

CXXXI.     The  Application  of  Fourier's  Theorem  to  the  Conduction 

of  Heat,      -  -  315 

CXXXII.     The  Cooling  of  a  Sphere, 318 

CXXXIII.     The  Motion  of  Heat  in  an  Infinitely  Long  Cylinder,    -  322 
CXXXIV.     On  the  Conduction  of  Heat  in  Fluids,  -        -        -        -  325 
CXX  XV.     The   Influence   of   the   Conduction  of  Heat  on  the  In- 
tensity and  Velocity  of  Sound  in  Gases,  -  330 


INTRODUCTION. 

IN  the  Science  of  Physics  it  is  assumed  that  all  phenomena  are 
capable  of  ultimate  representation  by  motions,  that  is,  by  changes 
of  place  considered  with  reference  to  the  time  required  for  their 
accomplishment.  We  therefore  begin  with  a  brief  discussion  of  the 
theory  of  pure  motion  (Kinematics).  We  will  treat  first  the  motion 
of  a  point.  The  continuous  line  traced  out  by  the  successive  positions 
which  a  moving  point  occupies  in  space  is  called  its  path.  The 
symbol  s  represents  the  distance  which  the  point  traverses  along  its 
path  in  the  time  t.  In  measuring  these  quantities  the  second  is 
used  as  the  unit  of  time;  the  centimetre,  as  the  unit  of  length. 
The  measures  of  all  the  magnitudes  which  occur  in  the  discussion 
of  motions  may  be  stated  in  terms  of  these  two  units. 

Motions  are  distinguished  by  the  form  of  the  path,  as  rectilinear, 
curvilinear,  or  periodic.  Eectilinear  and  curvilinear  motions  are 
sufficiently  defined  by  their  names.  A  periodic  motion  is  one  in 
which  the  same  condition  of  motion  recurs  after  a  definite  interval 
of  time;  that  is,  one  in  which  the  moving  point  returns  after  a 
definite  time  to  the  same  position  with  the  same  velocity  and  direction 
of  motion. 

Rectilinear  motion  may  be  either  uniform  or  variable.  It  is  uniform 
if  the  moving  point  traverses  equal  distances  in  equal  times.  In  this 
case  the  point  traverses  the  same  distance  in  each  unit  of  time,  and 
the  distance  traversed  in  the  unit  of  time  measures  its  velocity.  If 
the  point  traverses  the  distance  5  in  the  time  t  with  a  uniform  motion, 
the  velocity  c  is  the  ratio  of  s  to  t,  or  (a)  c  =  s/t.  A  velocity  is 
therefore  a  length  divided  by  a  time. 

If  a  point  moves  on  the  circumference  of  a  circle  with  a  constant 
velocity,  the  radius  vector  drawn  to  this  point  sweeps  out  equal 
sectors  in  equal  times.  In  this  case  the  angle  which  is  swept  out  by 
this  radius  vector  in  the  unit  of  time  measures  the  angular  velocity. 


4  INTRODUCTION. 

is  xl  +  x2.  The  component  of  velocity  in  the  direction  of  the  a--axis 
is  ac  =  sc1  +  «2. 

Similar  expressions  hold  for  motions  in  the  directions  of  the  other 
axes.  The  resultant  velocity  is  represented  by  the  diagonal  of  the 
parallelepiped,  whose  edges  are  x,  y,  z,  or  by  s  =  *]d?  +  y-  +  z2. 

Since  an  acceleration  is  the  increment  of  a  velocity,  the  resultant 
acceleration  will  be  determined  in  a  similar  manner.  Let  ajj,  a% 
represent  the  .r-components  of  the  increments  of  velocity  due  to 
the  two  motions.  We  then  have  for  the  total  acceleration  in  the 
direction  of  the  .r-axis,  x  =  xl  +  x.2,  and  for  the  acceleration  of  the 
point,  3  =  ^/(iBj  +  «2)2  +  (yl  +  i/^f  +  (^  +  z2)2,  by  which  s  is  expressed  as 
the  diagonal  of  the  parallelepiped  whose  edges  are  a;,  y,  z. 

If  the  coordinates  of  the  moving  point  are  given  as  functions  of 
the  time,  the  equation  of  the  path  is  obtained  by  determining  the 
values  of  x  and  y  which  hold  for  the  same  time  L  If,  for  example, 
x=fl(t)a,ndy=f2(t),  the  relation  between  x  and  y  is  found  by 
eliminating  the  variable  t  from  the  equations  by  any  appropriate 
method. 

From  this  brief  discussion  of  these  purely  kinematic  questions  we 
turn  to  the  consideration  of  the  causes  of  motion,  taking  as  our 
starting  point  the  researches  of  Galileo  on  freely  falling  bodies. 


CHAPTER  I. 

GENERAL  THEORY  OF  MOTION. 

SECTION  I.     FREELY  FALLING  BODIES. 

THE  investigation  by  Galileo  of  the  motion  of  freely  falling  bodies 
was  the  first  step  in  the  development  of  modern  physics.  It  is 
advantageous  to  start  from  the  same  point  in  our  study  of  the 
subject.  Galileo  concluded  from  his  experiments  that  all  bodies  falling 
freely  in  vacuo  will  fall  at  the  same  rate.  This  is  one  of  the  most 
important  discoveries  in  natural  science,  since  it  shows  that  all  bodies, 
independent  of  their  condition  in  other  respects,  have  one  property 
in  common.  No  parallel  to  this  has  been  found  in  Nature.  It  points 
to  a  unity  in  the  constitution  of  matter,  of  which  we  certainly  do 
not  as  yet  appreciate  the  full  significance. 

Galileo's  conclusions  have  been  confirmed  by  the  careful  experi- 
ments of  Newton,  Bessel,  and  others.  Galileo  concluded  further, 
that  the  distance  s  traversed  by  a  falling  body  in  the  time  t  is  proportional 
to  the  square  of  the  time,  so  that  (a)  s  =  ^gt2,  where  g  is  a  constant. 
The  constant  g  is  called  the  acceleration  of  gravity.  The  falling  body 
has  a  uniformly  accelerated  motion,  since 

=  s  =  t    and 


Its  acceleration  is  therefore  constant.  This  second  law  of  falling 
bodies  is  not  to  be  considered  a  fundamental  law  in  the  sense  in 
which  the  first  is.*  In  the  time  r  immediately  following  the  time  /, 
the  body  traverses  the  distance  <r,  which  is  determined  from  the 
equation  s  +  a-  =  ^(t  +  r)2g.  By  the  use  of  equation  (a)  we  obtain 

*  Later  researches  have  shown  that  the  value  of  the  force  of  gravity  depends 
on  the  distance  of  the  falling  body  from  the  centre  of  the  earth,  and  therefore 
<j  varies  during  the  fall.  However,  the  variation  of  //  is  so  slight  that  it  has 
not  yet  been  detected  by  direct  experiment  on  falling  bodies. 


6  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

(b)  a-  =  gtr  +  ^r2.  During  the  time  r  the  velocity  is  variable,  but  if 
v  is  the  mean  velocity  during  that  time,  we  will  have  v  =  <r/T  =  gt  +  ^gT. 
If  r  is  infinitely  small  and  equal  to  dt,  we  have  <r  =  ds,  and  neglecting 
\gdt  in  comparison  with  gt,  (c)  v  =  ds/dt='s  =  gt.  The  velocity  therefore 
increases  proportionally  to  the  time,  and  g  represents  the  increment  of 
relocity  in  the  unit  of  time. 

The  body  falls  through  the  space  s  in  the  time  t,  which,  from  (a), 
is  determined  by  (d)  t  =  <j2s/g. 

The  velocity  at  the  time  t  is  obtained  by  substituting  this  value 
of  t  in  (c) ;  making  this  substitution,  we  have  (e)  v  =  J2sg.  We 
reach  the  same  result  by  eliminating  t  between  equations  (a)  and  (c). 

From  the  laws  of  falling  bodies  we  deduce  the  law  of  inertia. 

In  order  to  explain  the  fact  that  the  velocity  of  a  falling  body 
increases  uniformly  with  the  time,  we  make  the  assumption,  that 
a  body  retains  a  velocity  once  imparted  to  it  unchanged  in  magnitude  and 
direction;  any  change  of  its  velocity  is  due  to  external  causes.  This  law 
is  called  the  principle  of  inertia. 

At  the  time  (t  +  T)  the  velocity  v'  is  v'  =  gt  +  gr.  The  initial  velocity 
is  here  gt,  to  which,  in  consequence  of  an  external  cause,  namely, 
the  force  of  gravity,  the  velocity  gr  is  added.  Under  the  action  of 
gravity  the  body  traverses  the  space  cr  =  gtT  +  ^gT2  in  the  time  T, 
immediately  following  the  time  t.  The  falling  body  traverses  the 
space  gtr  during  the  time  T,  with  the  velocity  gt  attained  at  the 
end  of  the  time  t ;  the  additional  distance  \grl  traversed  by  the 
falling  body  is  due  to  the  action  of  gravity  during  the  time  T. 

The  principle  of  inertia  holds  not  only  when  the  increment  of 
velocity  is  in  the  same  direction  as  the  original  velocity,  but  also 
when  it  makes  any  angle  whatever  with  the  original  velocity.  This 
principle  justifies  the  application  of  the  methods  of  geometrical 
addition  to  the  motions  and  accelerations  of  bodies. 

The  laws  of  falling  bodies  lead  also  to  an  answer  to  the  question, 
how  forces  are  to  be  measured.  It  is  evident  that  the  gain  in 
velocity  of  a  body,  and  its  pressure  on  a  support,  that  is,  its  weight, 
are  properly  regarded  as  actions  of  one  and  the  same  force.  The 
increasing  velocity  of  a  body  in  its  fall  is  an  evidence  of  that  force, 
and  the  increment  of  velocity  in  the  unit  of  time  gives  a  new  measure 
of  it.  This  definition  shows  what  before  Galileo's  time  was  not 
clearly  understood,  how  the  combined  action  of  several  forces  may  be 
measured.  The  increments  of  velocity  corresponding  to  the  separate 
forces  are  combined  by  the  method  previously  described,  and  the  re- 
sulting increment  gives  a  measure"  of  the  combined  action  of  the  forces. 


SECT.  II.] 


THE  MOTION  OF  PBOJECTILES. 


SECTION  II.    THE  MOTION  OF  PROJECTILES. 

We  will  apply  the  foregoing  principles  to  the  motion  of  projectiles, 
which  is  closely  connected  with  that  of  freely  falling  bodies.  We 
consider 

1.  Vertical  projection,  both  downward  and  upward. 

Galileo,  in  his  study  of  the  motion  of  projectiles,  proceeded  on 
the  assumption  that  a  body  which  is  given  an  initial  motion  in  any 
direction  retains  this  motion,  which  is  combined  with  that  imparted 
to  it  by  gravity  in  accordance  with  the  laws  of  freely  falling  bodies. 
If,  at  the  time  t  =  0,  the  velocity  u  is  given  to  a  body,  directed 
vertically  downward,  its  velocity  r,  after  the  lapse  of  the  time  t, 
is  (a)  r  =  u  +  gt,  and  the  distance  traversed  is  (b)  s  =  ut  +  ^gt2.  If 
the  body  is  given  the  initial  velocity  u,  directed  vertically  upward, 
the  corresponding  formulas  are  (c)  (d)  v  =  u-  gt  and  s  =  ut-  ^gt~. 

2.  Projection  in  a  direction  inclined  to  the  vertical. 

Let  a  body  be  projected  in  the  direction  OA,  making  an  angle  a 
with  the  horizontal.  Let  OA  represent  the  initial  velocity  u  (Fig.  1). 
The  space  which  the  body 
would  traverse  in  the  time 
t  if  gravity  did  not  act  on 
it  is  OB  =  ut.  The  body, 
however,  does  not  reach  B, 
but,  at  the  end  of  the  time 
/,  is  beneath  B  at  the  point 
C,  so  that  BC=\gP.  Let 
the  x-axis  Ox  and  the  #-axis 
Oy  lie  in  the  vertical  plane 
containing  OB  ;  then  the 
coordinates  of  the  point  C 
at  the  time  t  are 
(e)  x=  OD  = 


By  these  equations  the  posi- 
tion of  the  body  at  any 
time  is  determined.  During 


FIG.  1. 

the  time-element  dt  the  coordinates  x  and  y  increase  by 
(f)  dx  =  ucosadt  and  dy  =  u  sin  a  dt  -  gtdt. 

The  distance  ds  traversed  in  the  time  dt  is  determined  by 
ds*  =  dx*  +  df  =  [(u  cos  a)2  +  («  sin  a  -  gt)2]dt*. 


8  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

The  velocity  v  is  given  by 

(g)  (h)  v  =  s  and  v*  =  s2  =  &  +  f  =  u?-  Zugtsin  a  +  gW. 

The  components  of  the  velocity  parallel  to  the  axes  Ox  and  Oy 
are  respectively  v  .  dx/ds  and  v  .  dy/ds,  or,  by  (g),  are  equal  to  x  and  y. 
From  equation  (f  )  we  have  (i)  x  =  u  cos  a  ;  y  =  u  sin  a  -  gt.  Hence 
the  horizontal  velocity  is  constant,  while  the  vertical  velocity 
diminishes  uniformly;  this  follows  because  the  only  force  acting  is 
directed  vertically  downward. 

If  t  is  eliminated  from  the  equations  (e)  we  obtain  for  the  equation 
of  the  path  (k)  y  =  x  tan  a  -  x2/4h  .  (1  +  tan'2a),  where  h  =  %u2/g  is  the 
distance  through  which  the  body  must  fall  under  the  action  of  gravity 
to  attain  the  velocity  u.  Equation  (k)  shows  that  the  path  is  a 
parabola.  The  range,  or  the  distance  reckoned  from  0,  at  which 
the  path  cuts  the  z-axis,  is  given  by  (k)  if  we  set  y  =  0.  We  have 
0  =  tan  «  -\gx(\  +tan2a)/u2.  The  range  W,  or  the  particular  value 
of  x  given  by  this  equation,  is  W=  u2  sin  2a/</  ;  the  maximum  range 
is  attained  when  o.  =  \ir. 

If  the  velocity  u  is  given,  we  may  determine  from  equation  (k) 
the  direction  in  which  a  body  must  be  projected  in  order  to  reach 
a  prescribed  point.  Transposing,  we  obtain 


tan  a  =  (2h  ±  V4A2  -  4hy  -  x2)/x. 

This  equation  shows  that  there  are  in  general  two  directions  in 
which  the  body  may  be  projected  with  the  initial  velocity  u  so  as 
to  reach  a  prescribed  point.  If  the  expression  under  the  radical 
is  zero,  there  is  only  one  possible  direction.  If  the  point  to  be 
reached  by  the  body  is  so  situated  that  4/i2  -  4hy  -  x2<0,  tan  a  will 
be  imaginary,  and  the  body  will  not  reach  the  prescribed  point. 


SECTION  III.    EQUATIONS  OF  MOTION  FOR  A  MATERIAL  POINT. 

In  the  theory  of  motion  we  use  the  word  force  to  designate  the 
causes,  known  or  unknown,  of  a  change  in  the  motion  of  a  body. 
If  a  body  at  rest  is  set  in  motion,  or  if  a  moving  body  comes  to 
rest,  these  changes  are  ascribed  to  the  action  of  a  force.  If  the 
change  is  sudden,  the  force  acting  on  the  body  is  called  an  instantaneous 
force,  or  impulse.  Close  examination  shows,  however,  that  finite 
changes  in  the  motion  of  a  body  are  never  instantaneous,  but  occur 
only  in  a  finite  time.  This  time  may,  in  many  cases,  be  very  small. 
The  motion  of  a  body,  which  is  measured  by  its  velocity,  may  vary 


•SECT,  in.]  EQUATIONS  OF  MOTION.  9 

both  in  amount  and  in  direction.  The  velocity  of  a  freely  falling 
body  varies  only  in  amount ;  the  velocity  of  a  body  revolving  round 
a  centre  varies  in  direction,  and  sometimes  also  in  amount.  Experi- 
ment shows  that  all  changes  in  the  direction,  as  well  as  in  the 
amount  of  velocity,  are  due  to  external  causes,  which  act  during  a 
longer  or  shorter  time,  but  never  instantaneously. 

We  may  set  aside  all  questions  as  to  the  origin  of  force,  and 
measure  the  amount  of  a  force  by  its  action.  We  may  take  as  a 
measure  of  a  force  either  the  space  which  a  body,  starting  from 
rest,  traverses  under  the  action  of  the  force,  or  the  velocity  which 
the  force  imparts  to  the  body  in  a  given  time.  There  is  no  essential 
difference  between  these  two  modes  of  measurement,  but  generally 
the  velocity  produced,  or  better,  the  change  in  velocity,  is  used  for 
the  purpose.  We  measure  the  amount  of  an  impulse  by  the  change  in 
velocity  imparted  to  the  body  by  the  impulse,  and  the  amount  of  a  constant 
force  acting  continually  upon  the  body,  by  the  change  in  velocity  which 
occurs  in  a  second.  Newton  assumed  further,  that  the  force  is  proportional 
to  the  quantity  of  that  which  is  set  in  motion,  that  is,  to  the  mass  m  of 
the  body.  He  therefore  set  F=f.m.  b  where  b  is  the  acceleration  of 
the  body,  and  /  is  a  factor  dependent  on  the  units  of  force,  mass, 
and  acceleration,  or  on  the  units  of  mass,  time,  and  length.  If  we 
set/=l,  then  F=m.b,  and  we  obtain  the  following  definition  for 
the  unit  of  force :  The  unit  of  force  is  that  foi~ce  which  imparts  the  unit 
of  acceleration  to  the  unit  of  mass,  or  which  imparts  to  a  body  in  a  second 
the  unit  of  momentum  (cf.  XVI.).  This  unit  of  force,  called  a  dyne, 
is  therefore  that  force  which,  acting  for  one  second,  imparts  to  a 
mass  of  one  gram  the  velocity  of  one  centimetre  per  second.  Hence 
the  dimensions  of  force  are  MLT~2  (cf.  Introduction). 

The  force  with  which  a  body  is  attracted  by  the  earth  is  called 
its  weight,  and  is  measured  by  the  product  of  its  mass  and  the 
acceleration  which  it  would  have  if  it  were  falling  freely.  If  a  body 
is  prevented  from  falling  by  a  support,  it  exerts  a  pressure  on  the 
support  which  is  equal  to  its  weight.  Conversely,  the  support  exerts 
the  same  pressure  on  the  body,  in  accordance  with  the  law  of  action 
and  reaction.  This  pressure  may  be  determined  by  the  balance,  by 
the  elasticity  of  a  spring,  etc. 

Since  the  velocity  which  is  caused  by  a  force  F  may  be  resolved 
into  components  in  the  directions  of  the  three  axes  of  a  system  of 
rectangular  coordinates,  so,  in  the  same  way,  the  force  F  may  be 
resolved  into  components  along  the  three  coordinate  axes.  If  these 
components  are  represented  by  X,  Y,  and  Z,  we  have  F2  =  X'2  +  Y'2  +  Z2. 


10  GENERAL  THEORY   OF  MOTION.  [CHAP.  i. 

We   may  also  resolve   forces  into  their  components  in  other  ways. 
These  will  be  treated  later. 

If  a  body  moves  with  the  velocity  v  in  the  direction  AB  (Fig.  2)» 
and  if  a  force  acts  on  it  in  the  direction  AC, 
the  path  of  the  body  may  be  determined  by 
the  method  used  by  Galileo  to  obtain  the  law 
of  the  motion  of  projectiles.  Consider  the  motion 
in  the  time  T.  In  that  time  the  body  will  tra- 
verse the  distance  AM  =  VT,  in  consequence  of 
its  initial  velocity ;  in  the  same  time  it  will 
traverse  the  distance  AN=  |yr2  under  the  action 
of  the  force  F,  if  y  represents  the  acceleration 
due  to  the  force  F.  If  the  parallelogram  AMDN 
^  is  constructed,  D  will  be  the  position  of  the  body 

at  the  end  of  the  time  r. 

Let  the  direction  of  motion  make  the  angles  a,  /8,  y  with  the  axes 
OX,  OY,  OZ  of  a  system  of  rectangular  coordinates ;  let  the  direction 
of  the  force  make  the  angles  A,  /j.,  v  with  the  same  axes.  If  the 
coordinates  of  the  point  A  are  a*,  y,  and  z,  the  x  coordinate  of  the 
point  D  is  (b)  .r  +  AM  cos  a  +  DM  cos  A  =  x  +  VT  cos  a  +  £yr2  cos  A.  Since 
the  coordinates  are  functions  of  the  time,  we  may  obtain  an  expression 
for  the  ^-coordinate  of  D  by  the  use  of  Taylor's  theorem.  Applying 
this  theorem,  we  obtain  (c)  x  +  XT  +  |arr2  +  . . .  .  By  comparing  (b)  and 
(c)  it  follows  that  (d)  (e)  v  cos  a  =  x  and  y  cos  A  =  .r.  In  a  similar 
way  we  obtain  v  cos  jK  =  y,  y  cos  n  —  i/ ;  v  cos  y  =  z,  y  cos  v  =  z.  Tb  e 
symbols  x,  y,  z  represent  the  velocities  along  the  coordinate  axes  ; 
this  appeai-s  also  if  we  write  v  =  ds/dt,  and  notice  that  cosa  =  dx/ds, 
etc.,  so  that  v  cos  a  =  dxjds  .  s  =  x.  From  (e)  it  follows  further  that 
ray  cos  \  =  mx.  Since  my  is  the  force  F=JX2+  Y*  +  Z*  and  my  cos  A 
represents  the  x  component  X  of  the  force  F,  we  have  (f)  X=mx; 
similarly  Y=my,  Z  =  mz.  These  equations  (f)  are  the  equations  of 
motion  of  the  particle  m.  If  X,  Y,  and  Z  are  given  functions  of  the 
coordinates,  of  the  time,  and  sometimes  of  the  velocity,  equations 
(f)  will  determine  the  motion  of  the  mass  m,  if  its  position  and 
velocity  are  given  at  the  beginning  of  the  motion.  To  determine 
the  motion,  however,  it  is  necessary  to  integrate  equations  (f),  which 
can  be  done  in  only  a  very  few  cases.  If  the  motion  is  known, 
that  is,  if  .r,  y,  and  z  are  given  as  functions  of  the  time  #,  these 
equations  may  be  more  easily  applied  to  find  the  force  which  causes 
the  motion. 

We  will  now  consider  some  examples. 


SECT.  III.] 


EQUATIONS  OF  MOTION. 


11 


1.  Motion  in  a  Circle. 

Let  a  body  of  mass  in  move  with  constant  velocity  in  the  circle 
ABC,  whose  centre  lies  at  the  origin  of  coordinates,  and  whose 
radius  is  R  (Fig.  3).  Let  T  represent  the  time  of  revolution,  or 
period.  If  w  represents  the  angular  velocity  of  the  body,  and  if  the 
,«-axis  is  drawn  through  the  point  occupied  by  the  body  at  the 
time  t  =  0,  we  have  x  =  Bcos  (at),  y  =  E  sin  (wi).  It  then  follows 
from  (f)  that  X=mx=  -raw2/i?cos  ((at),  Y=my  =  -moPRsm  ((at)  or 
X=  —m<a?x,  Y=  —  rridPy.  The  force  acting  on  the  body  is,  therefore, 
F=  *JX'2  +  1' 2  =  maPll.  The  cosines  of  the  angles  made  by  the  direction 


FIG.  3. 

of  the  force  with  the  x-  and  y-axes  are  respectively  -x/ft&nd  —  y/R. 
The  force  is  therefore  directed  toward  the  centre  of  the  circle.  If 
v  represents  the  velocity  of  the  body  in  the  circle,  we  have 
v  =  /fw  =  27r£/r  and  F=mv*/R  =  4Tr*mfi/T2.  The  acceleration  directed 
toward  the  centre,  the  so-called  centripetal  acceleration,  is  equal  tov2/R  =  Ii(a2. 
F  is  called  the  centripetal  force.  This  result  was  first  obtained  by 
Huygens. 

2.  The  Motion  of  Projectiles. 

Let  a  body  be  projected  from  the  origin  of  coordinates  with  the 
velocity  u  in  a  direction  which  makes  the  angle  a  with  the  horizontal 
#-axis ;  let  the  positive  ?/-axis  be  directed  upward.     Then 
A"=0,     Y=-mg. 


12 


GENERAL  THEORY  OF   MOTION. 


[CHAP. 


The  equations  of  motion  are  mx  =  Q,  mi/  =  -ing.  By  integration 
we  have  x  =  a  +  a^,  y  =  b  +  b1t-  \gfi,  where  a,  av  b,  bl  are  constants. 
Since  the  body  is  at  the  origin  at  the  time  t  =  0,  we  have  a  =  0  and 
6  =  0.  The  components  of  velocity  at  the  time  t  are  x  =  av  y  =  \-gi 
From  the  value  of  the  velocity  at  the  time  t  =  Q  we  have 

al  =  u  cos  a,     l)l  =  u  sin  a. 
We  thus  obtain  again  the  equations  given  in  II.  (e). 


3.  Oscillatory  Motion. 

If  an  elastic  cylindrical  rod,  whose  weight  is  so  small  as  to  be 
negligible,  and  which  carries  on  one  end  a  heavy  sphere,  is  clamped 
firmly  by  the  other  end,  and  if  it  is  then  bent,  it  will  be  urged  back 
toward  its  position  of  equilibrium  by  a  force  which  is  proportional 
to  its  displacement  from  that  position.  If  r  (Fig.  4)  is  the  distance 


FIG.  4. 


from  the  position  of  equilibrium  0  to  the  point  P,  which  the  body 
occupies  at  the  time  t,  the  force  which  urges  it  toward  0  may  be 
set  equal  to  -  m&V,  where  k  is  a  constant.  The  components  of  this 
force  are  X=  -mk'2x,  Y=  -mkzy,  and  the  equations  of  motion  are 


SECT.  III.] 


EQUATIONS  OF  MOTION. 


13 


The  integrals  of  these  equations  are 

x  =  a:  cos  let  +  Jx  sin  kt,     y  =  a2  cos  let  +  b.2  sin  kt, 

where  av  bv  a2,  b2  are  constants.  If  the  coordinates  of  the  point  P 
at  the  time  t  =  0  are  x0,  y0,  and  if  the  components  of  the  velocity  v 
at  the  same  time  are  UQ  and  #0,  we  have 

z0  =  ai'   ^0  =  %;    «o  =  &A   v0  =  bjc. 

With  these  values  of  the  constants,  the  equations  become 
x  =  x0  cos  kt  +  u0/k .  sin  kt,     y  =  yQ  cos  kt  +  vjk .  sin  K 
The  components  of  the  velocity  are 

x=  -kx0sinkt  +  u0coskt,     y  =  -  ky0sin  kt  +  v0cos  kt. 
If,  when  /  =  0,  the  point  P  lies  on  the  axis  Oy,  and  if  its  initial 
velocity  v  is  parallel  to  the  or-axis,  V  =  UQ  and  v0  =  0;   and 

x  =  MO/^  .  sin  kt,    y  =  y0cos  kt. 

For  two  values  of  t  which  differ  by  2ir/k,  the  values  of  x  and  y 
are  the  same.  The  motion  is  therefore  periodic.  The  period  is 
T=2ir/k.  If  we  divide  the  first  equation  by  ujk,  the  second  by  y0, 
and  add  the  squares  of  the  right  and  left  sides  of  both  equations, 
we  eliminate  t,  and  obtain  the  equation  of  an  ellipse  as  the  equation 
of  the  path  of  the  body. 


SECTION  IV.    THE  TANGENTIAL  AND  NORMAL  FORCES. 
Let  MAD  (Fig.  5)  be  a  part  of  the  path  of  a  body  whose  mass 


FIG.  5. 
is  m,  let  AB  be  the  tangent  to  the  path  at  the  point  A,  AC,  the 


14  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

direction  of  the  force  F  acting  on  the  body.  The  directions  of  the 
motion  and  of  the  force  lie  in  the  plane* of  the  path.  We  choose 
this  plane  for  the  ay-plane  of  a  system  of  rectangular  coordinates 
whose  .vaxis  lies  in  the  direction  AB.  The  normal  AH,  drawn  to 
the  same  side  as  the  force  AC,  is  taken  as  the  positive  #-axis.  The 
equations  of  motion  are  mx  =  T,  my  —  N. 

T  and  N  are  the  components  of  the  force  in  the  direction  of  the 
tangent  and  of  the  normal,  and  are  called  in  consequence  tangential 
and  normal  forces.  If  the  small  arc  AD  is  represented  by  s,  if  H 
is  the  centre  of  curvature  of  the  curve  at  the  point  A,  and  if  the 
radius  of  curvature  AH  is  represented  by  It,  the  coordinates  of  D  are 

x  =  E .  sin  (s/R),    y  =  R-Rcos  (s/R). 
Hence  x  =  s.  cos (s/R)  -  sin  (s/R) .  £/R, 

i/  =  s.sin  (s/R)  +  cos  (s/R) .  ^,'R. 

If  s  is  very  small,  we  may  set  cos  (s/R)  =  1  and  sin  (s/R)  =  0.  We 
have  then  x  =  s,  y  =  s2/R  =  v*/R, 

and  therefore  T=ms   and   N=m^jR, 

that  is,  the  tangential  force  is  proportional  to  the  acceleration  in  the  path. 
The  normal  force  is  proportional  directly  to  the  square  of  the  velocity,  and 
inversely  to  the  radius  of  curvature. 


SECTION  V.    WORK  AND  KINETIC  ENERGY.* 

If  a  particle,  under  the  action  of  a  force  S,  moves  along  a  path  ds, 
whose  direction  is  that  of  the  force  S,  the  force  is  said  to  do  work 
equal  to  Sds.  If  the  direction  of  motion  and  the  direction  of  the 
force  make  the  angle  6  with  each  other,  we  must  use  the  component 
of  the  force  in  the  direction  of  motion,  instead  of  the  total  force  S ; 
the  work  done  is  Sds  cos  0.  If  the  body  moves  in  a  given  path  s0s 
under  the  action  of  the  tangential  force  T,  the  work  done  by  motion 
through  the  element  ds  is  Tds,  and  the  work  done  in  the  path  s0s 

is  given  by  the  integral  I  Tds.      If  the  velocity   of  the  particle   is 
represented  by  v,  v  =  ds/dt  and  T=mii  —  mv.     Hence 
(a)  ('Tds  =  fmvvdt  =  \m$  -  \mi-*, 

where  v0  represents  the  velocity  of  the  body  in  its  initial  position  s0. 

The  quantity  £mt'2,  or  the  product  of  one   half  the   mass  and  the 

*  Kinetic  energy  is  also  called  actual  energy  or  vis  viva. 


SECT,  v.]  WORK  AND  KINETIC  ENERGY.  15 

square  of  the  velocity,  is  called  the  kinetic  energy  of  the  body. 
From  equation  (a)  the  gain  in  kinetic  energy  is  equal  to  the  work  done 
by  the  tangential  foi'ce,  or  is  equal  to  the  work  done  by  the  total  force, 
since  in  the  calculation  of  the  work  as  it  has  been  defined  it  is 
necessary  to  consider  only  the  component  of  the  total  force  which 
acts  in  the  direction  of  the  path.  If  a,  (3,  y  are  the  angles  which  ds 
makes  with  the  coordinate  axes,  and  X,  Y,  Z  the  components  of  the 
force  T,  then  the  following  equations  hold : 

T  =  JTcos  a  +  Fcos  ft  +  Zcos  y, 

ds  cos  a  =  dx,   ds  cos  /3  =  dy,   ds  cos  y  =  dz. 

The  work  done  by  the  force  T  in  the  infinitely  small  distance  ds  is 

Xdx  +  Ydy  +  Zdz. 
Equation  (a)  then  takes  the  form 

(b)  \(Xdx  +  Ydy  +  Zdz)  =  %mv2  -  \mv*. 

This  equatipn  may  be  applied  to  advantage  in  many  cases,  especially 
if  the  force  is  a  function  of  the  coordinates  only.  If  the  path  is 
also  given,  we  may  use  this  equation  to  determine  the  velocity  at 
any  point  in  the  path. 

1.  Example. — Let    the    a^-plane   of  a   system    of   rectangular    co- 
ordinates  be   horizontal,  f  and   let   the   y-axis   be   directed   vertically 
upward.     Let  a  body  of  mass  m  be  situated  on  the  «/-axis,  and  let 
the  only  force  acting  on  it  be  gravity.     Its  components  are 

X=Q,    Y=  -mg,   Z=0. 

We  have  therefore  {  (Xdx  +  Ydy  +  Zdz)  =  -  mg(y  -  b),  if  the  body 
begins  to  move  at  the  point  y  =  b.  From  (b)  we  obtain 

(c)  0  =  v<*-2g(y-b). 

Hence  the  velocity  is  determined  by  the  ^-coordinate  alone.  This 
example  is  discussed  in  II. 

2.  Example. — The  force  is  a  function  of  the  distance  of  the  particle 
from  a  fixed  point.     Let  the  force  be  a  repulsion  and  a  central  force, 
that  is,  one  whose  direction  passes  always  through  a  fixed  point  0. 
We  take  this  point  as   the   origin.      The   components   of  the   force 
which  acts  at  the  point  (x,  y,  z)  are 

X=f(r).x/r,    Y=f(r).ylr,   Z=f(r).z/r. 
Using  these  values,  we  have 

I  (Xdx  4-  Ydy  +  Zdz)  -  f^(xdx  +  ydy  +  zdz). 
Since    r2  =  x2  +  f  +  z2,    and    therefore   rdr  —  xdx  +  ydy  +  zdz,   the   work 


16  GENERAL  THEORY   OF  MOTION.  [CHAP,  u 

which  is  done  by  the  force  during  the  movement  of  the  body  from 
the  point  A  to  the  point  B  is    I  f(r)dr,  if  r0  and  r  are  respectively 

the  distances  from  the  point   0  to  the  points  A  and  B.     Let  the 
velocities  at  the  points  A  and  B  be.  respectively  v0  and  r,  then 


The  gain  in  kinetic  energy  depends  only  on  r0  and  r,  and  is  con- 
sequently independent  of  the  form  of  the  path.  The  general  con- 
dition which  must  be  fulfilled  that  the  work  done  may  depend 
only  on  the  initial  and  final  positions  of  the  body,  and  be  inde- 
pendent of  ther  path  traversed,  will  be  examined  in  the  next  section. 


SECTION  VI.    THE  WORK  DONE  ON  A  BODY  DURING  ITS  MOTION 
IN  A  CLOSED  PATH. 

If  a  body  describes  a  closed  path  A  BCD  (Fig.  6)  under  the  action 
of  a  force  whose  components  are  X,  Y,  Z,  the  work  done  upon  it  is 
determined  by  taking  the  integral  (a)  ^(Xdx+  Ydy  +  Zdz)  over  the 
whole  path.  If  the  body,  moving  from  A  with  the  velocity  r0, 


FIG.  6. 


traverses  the  closed  path  in  the  direction  indicated  by   the  arrow, 
and  returns  to  A  with  the  velocity  v,  the  work  done  is  equal  to 


Supposing  v>v0,  kinetic  energy  is  produced  during  the  motion,  and 
will  increase  continuously  if  the  motion  is  continued.  On  the  other 
hand,  supposing  v<v0,  kinetic  energy  will  be  produced  if  the  body 


SECT,  vi.]  MOTION  IN- A  CLOSED  PATH.  17 

traverses  the  path  A  BCD  in  the  Opposite  direction.  Now,  we  know 
by  experience  that  a  body,  under  the  action  of  forces  proceeding 
from  fixed  points,  after  traversing  a  closed  path,  returns  to  the 
starting  point  with  the  same  kinetic  energy  which  it  had  when  it 
started.  It  is  therefore  important  to  investigate  to  what  conditions 
the  components  of  the  force  must  conform  in  order  that  the  integral 
(a),  taken  over  a  closed  path,  shall  be  zero;  that  is,. the  conditions 
which  must  hold  in  order  that  a  body,  moving  through  a  closed 
path,  shall  return  to  its  original  position  with  the  same  kinetic 
energy  with  which  it  started. 

If  the  integral  taken  over  the  closed  path  A  BCD  equals  zero,  that 

{•ABC  rCDA 

is,  if          /      (Xdx  +  Ydy  +  Zdz)  +  J      (Xdx  +  Ydy  +  Zdz)  =  0, 

where  the  letters  connected  with  the  integral  signs  indicate  that  the 
first  integral  is  to  be  taken  over  the  line  ABC,  the  second,  over 
CD  A,  we  have 

/ABC  fADC 

(Xdx  +  Ydy  +  Zdz)  =  J      (Xdx  +  Ydy  +  Zdz). 

If  the  work  done  during  the  passage  of  the  body  from  one  point 
to  another  is  independent  of  the  path  and  dependent  only  on  the 
initial  and  final  points  of  its  path,  the  components  X,  Y,  Z  are 
single  valued  and  continuous  functions  of  the  position  of  the  point. 
Before  we  deduce  the  general  conditions  which  must  hold  in  order 
that  the  work  performed  by  a  force  shall  be  dependent  only  on  the 
initial  and  final  points  of  the  path,  we  will  determine  the  work 
done  in  the  case  in  which  the  area  enclosed  by  the  path  is 
infinitely  small.  Through  the  point  0  (Fig.  7),  whose  coordinates 
are  x,  y,  z,  we  draw  the  lines  Ox,  Oy,  Oz  parallel  to  the  coordinate 
axes,  whose  positive  directions  are  determined  in  the  following  way. 
If  the  right  hand  is  stretched  out  in  the  direction  of  the  positive 
£-axis,  a  line  drawn  from  the  palm  will  give  the  direction  of  the 
positive  y-axis,  and  the  thumb  that  of  the  positive  ?-axis.  A 
positive  rotation  around  the  re-axis  is  that  by  which  the  +«/-axis 
is  brought  by  a  rotation  through  a  right  angle 
into  coincidence  with  the  +2-axis.  This  rule, 
by  cyclic*  interchange  of  the  letters  x,  y,  z, 
gives  the  directions  of  the  positive  rotations 
about  the  y-  and  2-axes.  If  OBDC  is  a  rect-  *~ 
angle  in  the  i/2-plane,  and  if  its  perimeter  is 

*  That  is,  if  y  is  replaced  by  x,  z  will  be  replaced 
by  y,  and  x  by  z. 

B 


18 


GENERAL  THEOEY  OF   MOTION. 


[CHAP. 


traversed  in  the  direction  OBDCO,  the  motion  by  which  it  is  traversed 
is  said  to  be  in  the  positive  direction.  This  convention  as  to  the 
sign  of  the  direction  of  rotation  shall  hold  in  all  our  subsequent 
work.  If  we  set  OB  =  dy,  the  work  done  by  the  transfer  of  the 
body  from  0  to  B  equals  Ydy.  If  the  body  moves  from  B  to  D, 
the  work  done  is  (Z+'dZ/'dy .  dy)dz.  The  work  done  in  the  path 
DC  is  -(Y+dYfd2.dz)dy,  and  that  done  in  the  path  CO  is  -  Zdz. 


FIG.  7. 


Hence  the  total  work  done  is  (c>Zpy>-'dYj'dz)dydz.  In  general,  the 
work  done  by  a  force  during  the  movement  of  a  body  around  a 
surface  element  dS^  which  is  parallel  to  the  y^-plane,  is 

(b)  F.dSx 


In  the  same  way  we  obtain 

G  .  dS,  =  (dX/'dz  - 


SECT.  VI.] 


MOTION  IN  A  CLOSED  PATH. 


19 


F,  G,  and  H  are  the  quantities  of  work  done  during  the  movement 
of  the  body  around  a  unit  area  at  the  point  0,  when  perpendicular 
to  the  x-,  y-,  and  s-axes  respectively. 

If  OABC  (Fig.  8)  is  an  infinitely  small  tetrahedron,  whose  three 
edges  OA,  OB,  OC  are  parallel  to  the  coordinate  axes,  and  if  the 
body  moves  on  the  boundary  of  the  surface  ABC  in  the  direction 
given  by  the  order  of  the  letters,  the  work  done  is  equal  to  that 
which  is  done  by  moving  the  body  in  succession  about  OAB,  OBC, 
and  OCA.  By  this  set  of  motions,  the  distances  AB,  EC,  CA  will 
each  be  traversed  once  in  the  positive  direction,  while  the  distances 


FIG.  8. 


OA,  OB,  OC  will  each  be  traversed  twice  and  in  opposite  directions, 
so  that  the  work  done  in  them  is  zero.  The  work  done  during  the 
movement  of  the  body  about  the  surface  ABC  =  ds  is,  therefore, 
(c)  J  .dS=F  .dS.l  +  G.dS  .m  +  H.dS.n,  where  /,  m,  n  are  the 
cosines  of  the  angles  which  the  normal  to  the  surface  dS  drawn 
outward  from  the  tetrahedron  makes  with  the  coordinate  axes. 
Hence  the  work  J  done  during  the  movement  of  the  body  around 
unit  area  is  (d)  J=Fl  +  Gm  +  Hn,  where  Z,  m,  and  n  determine  the 
position  of  the  unit  area. 


20  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

If  the  work  done  during  the  movement  of  a  body  about  an  infinitely 
small  surface  is  zero,  we  must  have  .7=0  for  all  positions  of  the 
surface,  or  F=G  =  H=Q, 

(  -dZfdy  -  'dY/'dz  =  0,     ax/a?  -  VZ/Vx  =  0, 
\  9F/aB-aX/<ty  =  0. 

When  the  equations  of  condition  (e)  are  satisfied,  the  expression  under 
the  integral  sign  in  (a)  is  the  complete  differential  of  a  function  V 
of  x,  y,  z,  whence  X  =  rdVj'dx,  Y='dF/'dy,  Z='dF/'dz.  The  equations 
of  condition  (e)  are  satisfied  by  this  assumption.  The  function  V  is 
the  potential  of  the  acting  forces ;  we  here  obtain  for  the  first  time 
the  mathematical  definition  of  this  function,  whose  differential  co- 
efficients with  respect  to  x,  y,  z  are  the  components  of  force  X,  Y,  Z. 
If  V  is  a  value  of  the  potential  of  the  acting  forces,  V  is  also  a 
value,  if  V=V+C,  where  C  is  a  constant;  since 

.Y=3F/3z  =  3F7ar,   etc. 

The  value  of  the  potential  therefore  involves  an  unknown  or  arbitrary 
constant.  We  will  return  to  the  consideration  of  this  point  in  the 
next  section. 

If  the  equations  of  condition  (e)  are  everywhere  satisfied,  the 
work  done  during  the  movement  of  the  body  about  a  surface  is  also 

zero    when    the    surface    is    finite. 
//       >^  The  surface   may  be  divided  into 

'- — -t -i L*~*il — /\        surface-elements,  as  shown  in  Fig.  9. 

/       /       //,     I         I     \     If   the    body    moves    about    these 
~7      7      7       7       7       7|    elements  one  after  another  in  the 

v — 4- •+• 4- + -r -f — 7    same  direction,  the  total  work  done 

X^     /       /       /       /       I/     will  equal  zero.     It  is  here  assumed 
"x^,/       7       1^^^          ^Dat  ^e  f°rces  X,   F,  Z  are  con- 
•^~^^_    "      J  tinuous    and    single    valued    func- 

tions  of  the   coordinates.      Every 

line-element  thus  introduced  will  be  traversed  twice  in  opposite 
directions,  with  the  exception  of  those  which  form  the  boundary  of 
the  finite  surface. 

Those  forces  or  systems  of  forces,  which  are  such  that  the  work 
done  by  them  is  independent  of  the  path  in  which  the  body  is  trans- 
ferred from  its  initial  to  its  final  position,  are  called  conservative  forces. 
The  most  important  examples  of  such  forces  are  those  which  act 
from  a  fixed  point,  and  have  values  which  depend  only  on  their 
distance  from  it.  If  the  force  acting  at  the  point  P  depends  only  on 
the  distance  of  that  point  from  the  origin  of  coordinates  0,  that  is, 


SECT,  vi.]  MOTION  IN  A  CLOSED  PATH.  21 

if  it  equals  /(r),  then  X=f(r) .  x/r,  since  x/r  is  the  cosine  of  the  angle 
which  the  line  OP  makes  with  the  #-axis.  We  have  similarly 

X=f(r).x/r,     r=/(r).y/r,     Z=f(r).z/r. 

If  we  set  f(r)/r  =  R,  then  X=Ex,  Y=Ry,  Z=Rz.     We  have  then 
-dZfiy  =  dR/dr .  yz/r,     3F/3*  =  dR/dr .  yz/r. 

The  equation  of  condition  ~dZ/c)y -?>Y/'d2  =  0  is  therefore  satisfied. 
The  same  is  true  of  the  other  equations  of  condition  (e). 

The  work  done  during  the  movement  of  the  body  about  a  surface 
is  given  by  the  integral  \(Xdx  +  Ydy  +  Zdz).  This  work  is  also  done 
if  the  body  moves  in  succession  about  all  the  surface-elements  into 
which  the  finite  surface  is  divided  (Fig.  9).  In  this  process  the 
motion  must  be  uniformly  carried  out  in  the  same  sense.  From  (c) 
this  work  is  equal  to  \(Fl  +  Gm  +  Hn)dS.  If  we  substitute  the  ex- 
pressions for  F,  Gf,  H  formerly  obtained,  we  have,  by  the  use  of 
(a)  and  (b), 

{\(X .  dx/ds  +  Y.  dyjds  +  Z .  dz/ds)ds 
=  I  j  [(dZfdy  -  -dY/3z)l  +  (dX/Vz  -  -dZ/3x)m 
+  (dY/3x-'dXI'dy)n]dS, 

where  s  is  the  perimeter  of  the  surface  S,  and  I,  m,  n  are  the  direction 
cosines  of  the  normal  to  each  surface-element.  Equation  (f)  shows 
that  the  line  integral  along  a  closed  curve  may  be  replaced  by  a 
surface- integral  over  a  surface  bounded  by  this  curve.  The  only 
conditions  which  the  surface  S  must  fulfil  are  that  it  shall  be  bounded 
by  the  curve  and  have  no  singular  points.  The  theorem  contained 
in  (f)  was  discovered  by  Stokes. 


SECTION  VII.     THE  POTENTIAL. 

The  only  applications  of  the  potential  that  we  will  discuss  are 
those  like  the  foregoing,  in  which  the  work  done  during  the  motion 
is  completely  determined  by  the  initial  and  final  points  of  the  path. 
That  this  may  be  the  case,  we  must  have 


We  exclude  from  the  discussion  all  cases  in  which  these  equations 
do  not  hold. 

Let  the  components  of  the  force  in  the  field  be  X,    Y,  Z,     Let 
there  be  a  unit  of  mass  at  the  point  0  (Fig.  10),  whose  rectangular 


22 


GENERAL  THEORY  OF  MOTION. 


[CHAP.  i. 


coordinates  are  a,  b,  c,  and  let  it  move  from  0  to  P  along  the  path  s. 
The  work  V  done  by  the  force  during  this  motion  is 


(a) 


V=  f'(Xdx  +  Ydy  +  Zdz)  =  VP  - 


it  being  assumed  that  A',  Y,  Z  are  the  partial  derivatives  of  a  single 
function  V,  which  itself  is  a  function  only  of  x,  y,  z.  The  work 
required  to  transfer  the  unit  of  mass  from  any  point  0  to  P  is 
equal  to  the  difference  of  the  potentials  VP  and  V0  at  those  points, 
or  is  equal  to  the  difference  of  potential.  Such  differences  of  potential 
are  all  that  can  be  directly  measured.  The  value  of  the  potential  itself 


FIG.  10. 


involves  an  unknown  constant,  and  therefore  cannot  be  completely 
determined.  If  we  assume  that  the  potential  is  zero  at  the  point  0, 
then  VP  is  the  potential  at  P.  Hence  the  potential  at  any  point  is  the 
work  required  to  transfer  the  unit  of  mass  to  that  point  from  a  point 
where  the  potential  is  zero. 

The  potential  V  is  a  function  of  the  coordinates.  The  equation 
(b)  y(x,  y,  z)  =  C,  when  C  is  constant,  represents  a  surface  which  is 
the  locus  of  points,  such  that  the  amount  of  work  required  to 


SECT.  VII.] 


THE  POTENTIAL. 


23 


transfer  the  unit  of  mass  from  a  point  where  the  potential  is  zero 
to  any  one  of  them  is  the  same.  If  different  values  of  C  are  taken, 
we  obtain  a  system  of  surfaces,  which  are  called  level  or  equipotential 
surfaces.  Let  PF  and  QQ'  (Fig.  11)  be  two  infinitely  near  surfaces 
of  this  system ;  let  the  potential  on  PP  be  V,  and  on  QQ'  be  V+dV. 
Let  ds  be  the  element  of  an  arbitrary  curve  crossing  these  surfaces, 
which  is  cut  off  by  them.  If  the  force  acting  in  the  direction  of 
ds  is  T,  the  quantity  of  work  Tds  will  be  done  by  the  transfer 
of  the  unit  of  mass  from  P  to 
Q ;  this  work  is  also  equal  to 
V^-V^dV.  We  have,  there- 
fore, (c)  T  .ds  =  dV  or  T=dVjds. 
Hence  the  force  in  any  direction 
at  a  point  is  determined  from  the 
potential ;  the  relation  between 
force  and  potential  being  given  by 
equation  (c).  Since  the  direction 
of  the  element  ds  is  arbitrary,  we 
may  substitute  for  ds  the  elements 
dx,  dy,  dz,  and  obtain  for  the  com- 
ponents of  the  force  X='dV/'d.r, 
Y=  dF/c)y,  Z^'dF/^z.  From  equa- 
tion (c)  the  force  is  inversely  pro- 
portional to  the  element  ds  drawn 
between  the  two  equipotential  sur-  " 
faces  V  and  V '+  dV.  If  the 
direction  of  ds  is  that  of  the  normal  to  the  surface  PP',  the  force 
has  its  greatest  value.  If  a  series  of  lines  is  drawn  which  cut  the 
equipotential  surfaces  orthogonally,  their  directions  are  the  directions 
of  the  force  at  the  points  of  intersection.  Such  lines  are  conse- 
quently called  lines  of  force.  The  tangent  to  the  line  of  force  at  a 
point  gives  the  direction  of  the  force  at  that  point. 

If  Pl  and  P2  are  two  infinitely  near  points  in  an  equipotential 
surface,  no  work  need  be  done  to  transfer  a  body  from  Pl  to  P2, 
for  VPl  -  VP^  =  §\  the  force  acting  on  the  body  is  perpendicular  to 
the  direction  of  motion. 

1.  Example. — Gravity. — If  at  a  place  near  the  earth's  surface  we 
set  up  a  system  of  rectangular  coordinates,  so  that  the  .r^-plane  is 
horizontal,  and  the  positive  ?y-axis  directed  vertically  upward,  then 
^=0,  Y=  -mg,  Z=0.  Hence  we  have  V=  -mgy,  that  is,  the 
equipotential  surfaces  are  horizontal  planes. 


FIG.  11. 


24 


GENERAL  THEOEY  OF  MOTION. 


[CHAP.  i. 


2.  Example. — In  the  case  discussed  in  V.,  Ex.  2,  the  work 
F=fjf(r)dr 

is  needed  to  move  the  body  from  its  position  at  the  distance  r0 
from  a  fixed  point  to  another  position  at  the  distance  r.  Hence 
we  have  F"=  F(r)  -  jP(r0),  and  the  equipotential  surfaces  are  spheres 
whose  centres  are  at  the  centre  of  attraction  0. 


SECTION  VIII.    CONSTRAINED  MOTION. 

Galileo  investigated  not  only  freely  falling  bodies  and  the  motion 
of  projectiles,  but  also  motion  on  an  inclined  plane  and  the  motion 
of  a  pendulum,  and  thus  made  the  first  step  in  the  investigation  of 
constrained  motion. 

If  a  body  is  compelled  by  any  cause  to  move  in  a  given  path, 
which  is  not  that  which  it  would  follow  if  free  to  yield  to  the 
action  of  the  forces  applied  to  it,  its  motion  is  said  to  be  constrained. 

1.  Example.— The  Inclined  Plane.— Let  the  body  D  (Fig.  12),  acted 
on  by  gravity,  slide  down  an  inclined  plane  AB,  which  makes  the 


FIG.  12. 

angle  a  with  the  horizontal  plane  BC.  We  neglect  any  resistance 
which  may  arise  from  friction.  The  force  or  reaction  exerted  by 
the  inclined  plane  acts  perpendicularly  to  the  plane  AB,  and  does 
not  affect  the  motion  of  the  body.  The  expression  sought  may  be 
best  obtained  by  using  the  relation  between  kinetic  energy  and 
work.  If  in  represents  the  mass  of  the  body,  v  the  velocity  acquired 
at  B,  g  the  acceleration  of  gravity,  and  I  the  length  AB  of  the 
inclined  plane,  we  have  fymv2  =  mg  sin  a  .  /,  assuming  that  the  motion 
begins  at  A,  so  that  the  initial  velocity  is  zero.  If  a  represents  the 
height  AC  of  the  inclined  plane,  we  have  I  sin  a  =  a,  and  the  work 


-SECT    VIII.] 


CONSTEAINED  MOTION. 


25 


<lone  equals  mga.  Hence  the  velocity  of  the  body  at  the  foot  of 
the  inclined  plane  B  is  v  =  .j2ga,  and  is  the  same  as  that  which  it 
would  have  at  C,  if  it  were  to  fall  freely  through  the  distance  AC. 
If  a  body  moves  on  the  curve  AB  (Fig.  13)  under  the  action  of  gravity, 
we  determine  the  velocity  at  B  in  a  similar  way,  from  (the  initial 
velocity  v0  at  A  and  the  distance  of  the  fall  AC.  That  is,  we  have 


FIG.  13. 

%mv2  -  %mv0*  =  mga,  and  therefore  (a)  #2  =  v02  +  2ga.    The  time  t  required 
for  the  movement  of  the  body  from  A  to  B  is 

<">  -jft 

where  ds  is  an  element  of  the  path  AB. 

2.  Example. — The  Pendulum. — If  we  suspend  a  body  A  (Fig:  14) 
at  the  end  of  a  weightless  rod  which  can  swing  freely  about  the 
point  0,  it  is  compelled  to  move  on  the  surface  of  a  sphere  whose 
radius  is  equal  to  the  length  I  of  the  rod.  We  will  treat  only  the 
simple  case  in  which  the  departure  of  the  pendulum  from  its  position 
of  equilibrium  is  small.  If,  at  the  time  t  =  0,  the  body  starts  from 
rest  at  A,  it  will  move  in  the  arc  A  BCD  through  the  point  C  lying 
perpendicularly  under  0.  If  we  set  OA  =  l,  ^AOC  =  a,  i_£OC=0, 
and  if  A  A'  and  BB'  are  drawn  perpendicular  to  OC,  then  the  velocity 
which  the  body  gains  in  moving  from  A  to  B  equals  that  which  would 
be  gained  if  it  were  to  fall  freely  from  A'  to  B'.  The  distance  from 
A  to  B'  is  A'B'  =  l(cos  0-cosa),  and  therefore  the  velocity  at  B  is 

v  =  *j2gl  (cos  0  -  cos  a). 


26 


GENERAL  THEORY  OF  MOTION. 


[CHAP.  i. 


For  v  =  0,  or  for  6=  ±a,  the  pendulum  bob  will  be  at  rest,  and  is 
then  at  A  or  D,  if  i.DOC  =  LAOC.  The  time  t  taken  by  the 
body  to  move  from  A  to  B  is  found  by  substituting  this  value  of  v 
in  (b).  We  thus  obtain 

(c)  t=  -  Ieid6/J2gl(cos  6  -  cos  a), 

•'a 

0 


This  expression  is  easily  integrated  if  a,  and  therefore  0,  are  so 
small  that  we  may  set  cos0=l-£02  and  cosa  =  l-£a2.  The  ex- 
pansion of  the  cosine  in  a  series  is  of  the  form 


and  if  x  is  very  small  we  may  neglect  terms  of  higher  orders  than 
the  second.     Making  this  restriction,  we  have 


and  by  integration  (d)  6  =  acos(t>JgJl).  If  t^/g/l  =  ^Tr,  we  have  0  =  0; 
the  body  moving  from  A  reaches  the  lowest  point  of  its  path  in 
the  time  t  =  £ir  .  *Jl/g.  The  time  T  required  for  the  movement  of 
the  body  from  A  to  D  is  twice  this,  or  (e)  T=irjl/g.  T  is  called  the 
period  of  oscillation.*  The  period  of  oscillation  is  directly  proportional 

*  In  the  case  of  the  pendulum  here  treated,  which  swings  in  a  plane,  it  must 
be  clearly  understood  that  by  the  period  of  oscillation  T  only  one  advancing 
or  returning  beat  is  meant.  In  other  periodic  motions,  the  period  of  oscilla- 
tion is  the  time  between  two  instants,  at  which  the  motion  of  the  body  is 
precisely  similar,  that  is,  at  which  the  body  has  the  same  velocity  and  direction 
of  motion  ;  or,  it  is  the  time  required  for  both  the  advancing  and  returning 
l>eats. 


SECT.  VIIL]  CONSTEAINED  MOTION.  27 

to  the  square  root  of  the  length  of  the  pendulum,  and'  is  inversely  proportional 
to  the  square  root  of  the  acceleration  of  gravity. 

The  equation  (e)  holds  only  for  very  small  arcs.     In  the  case  of 
finite  values  of  a,  we  use  the  formula 

(f) 


It  is  only  for  very  small  arcs  that  the  oscillations  of  the  pendulum 
are  isochronous,  that  is,  independent  of  the  size  of  the  arcs.  If  the 
arc  is  not  infinitesimal,  the  period  will  increase  rather  rapidly  with 
the  length  of  the  arc. 

The  pendulum  may  also  be  studied  in  the  following  way.  Let 
an  oscillating  body  of  mass  m  be  at  the  point  J5,  and  be  acted  on 
by  the  force  mg.  We  may  represent  this  force  by  the  line  OE 
(Fig.  14).  Draw  EF  perpendicular  to  OB;  then  OF  and  FE  are 
components  of  the  force  OE.  The  magnitude  of  the  tangential  force 
is  mgsind.  If  we  set  BC  =  s,  and  reckon  the  tangential  force  posi- 
tive, when  it  tends  to  increase  st  we  have  P=  -mgsin(s/l),  and  if 
we  assume  s  to  be  very  small,  P=  -mgs/L  Hence  the  equation  -of 
motion  is  (g)  ms  =  P  or  »=  -gs/l.  By  integration  we  obtain,  by 
a  suitable  choice  of  constants,  (h)  s  =  a  cos  (t\/g/l).  This  equation 
corresponds  to  (d). 

If  a  body  is  compelled  to  move  on  a  given  surface,  the  deter- 
mination of  its  motion  is  in  general  very  difficult.  We  will  not 
enter  into  the  discussion  of  the  general  case,  but  will  consider  only 
the  motion  of  an  infinitely  small  body  on  a  spherical  surface,  when 
the  body  during  the  motion  always  remains  near  the  lowest  point 
C  of  this  surface,  and  when  gravity  is  the  only  force  acting  on  it. 
We  may  then  assume  that  the  component  of  gravity  which  moves 
the  body  is  directed  toward  the  point  C,  and  that  it  is  equal  to 
mgsjl,  when  I  represents  the  radius  of  the  sphere.  This  assumption 
gives  the  motion  treated  in  III.,  Ex.  3.  The  path  is  an  ellipse  and 
the  time  of  oscillation  is  T=2ir*Jl/g.  The  time  of  oscillation  is 
therefore  independent  of  the  form  and  dimensions  of  the  path. 


SECTION  IX.     KEPLER'S  LAWS. 

In  our  deduction  of  the  principal  theorems  of  the  general  theory 
of  motion,  we  proceeded  from  Galileo's  laws  of  falling  bodies.  We 
turn  now  to  that  force  of  which  gravity  is  a  special  example,  and 


28 


GENERAL  THEORY  OF  MOTION. 


[CHAP.  i. 


from  whose  properties  the  laws  of  planetary  motion  may  be  deduced. 
Starting  with  the  hypothesis  of  Copernicus,  that  the  sun  is  stationary 
and  that  the  earth  rotates  on  its  own  axis  and  also  revolves  round 
the  sun,  Kepler  announced  the  following  laws  : 

1.  A  radius  vector  drawn  from  the  sun   to  a  planet  describes  equal 
sectors  in  equal  times. 

2.  The  orbits  of  the  planets  are  ellipses  with  the  sun  at  one  of  the  foci. 

3.  The  squares  of  the,  periodic,  times  of  two  planets  are  proportional  to 
the  cubes  of  the  semi-major  axes  of  their  orbits. 

These  laws  may  be   expressed"  analytically  in  the  following  way. 
Let  S  be  the  centre  of  the  sun  (Fig.   15)  and  APQ  a  part  of  the 

orbit  of  a  planet.  Let  the 
planet  be  at  A  at  the  time 
=  0,  and  at  P  at  the  time  t. 
In  the  next  time  element  dt 
the  planet  moves  from  P  to 
Q,  and  its  radius  vector  drawn 
from  the  sun  describes  the 
sector  PSQ.  Let  -ASP  =  Q, 
LPSQ  =  dQ,  and  SP  =  r.  The 
surface  PSQ  is  equal  to  \r-dQ. 
Since  by  Kepler's  first  law 
the  surface  described  by  the 
radius  vector  increases  pro- 
portionally to  the  time,  we 
have  r2dQ  =  Mt,  where  k  is 
constant,  or,  writing  the  equa- 
tion in  another  form, 
(a)  7-2.0  =  &. 

Kepler's  first  law  is  a  special 
case  of  a  general  law  which  is 


FIG.  15. 


If  the  force  which  acts  upon  a 
the  surface  described  by  the 


called  the  law  of  areas.     This  law  is 

moving  body  pi'oceeds  from  a  fixt 

radius  vector  drawn  from  that  point  to  the  body  increases  at  a  constant 

rate.     Hence  Kepler's  first  law  holds  for  all  central  forces. 

From  (a)  it  appears  that  the  angular  velocity  0  is  inversely 
proportional  to  the  square  of  the  distance  of  the  planet  from  the 
sun. 

We  represent  the  velocity  of  the  planet  at  P  by  v,  and  the 
perpendicular  from  S  upon  the  tangent  to  the  orbit  at  the  point  P 
by  SN=p. 


SECT,  ix.]  KEPLER'S  LAWS.  29 

If  we  set  PQ  =  ds,  the  area  of  the  sector  PSQ  is  equal  ,to 
^pds  =  fypvdt.  But  it  is  also  equal  to  ^r2dQ  =  \k  .  dt.  Hence  we  have 
pvdt  =  Mt,  or  pv  =  Tc,  that  is,  the  velocities  of  the  planet  at  different 
points  in  its  orbit  are  inversely  proportional  to  the  distances  of  the 
tangents  at  those  points  from  the  sun,  the  centre  of  attraction. 

Let  BPC  be  the  elliptical 
orbit  of  the  planet  (Fig.  16), 
with  the  sun  situated  at  the 
focus  S.  Let  the  major  axis 
be  BC=2a,  and  let  SA  be  a 
fixed  radius  vector  which  makes 
the  angle  a  with  the  major  axis. 
We  set  SP  =  r,  ±_ASP  =  Q.  If 


n          i     ci  if-  v.  FIG.   16. 

F  and  S  are  the  foci;  we  have 

PF+PS=2a,     PF=2a-r, 

and  hence  (2a-?-)2--=4a2e2  +  r2  +  4aercos(6-a), 

if  e  is  the  eccentricity,  and  if,  therefore,  FS=2ae.     From  this  equation 
we  obtain  for  the  equation  of  the  path  in  polar  coordinates 
(b)  1/r  =  [1  +  e  cos  (6  -  o)]/a(l  -  e2). 

From  equation  (a)  we  have  J^r2  .  dO  =  |&  .  T,  if  the  integration  is 
taken  over  the  whole  orbit  and,  if  T  is  the  periodic  time.  The 
integral  is  equal  to  the  area  of  the  ellipse,  or  to  a  .  b  .  TT,  if  b  represents 
the  minor  axis.  Hence  we  have  %Trab  =  k.T.  If  we  notice  that 
a2  =  b2  +  a?e2,  we  obtain  (c)  27r«2V  1  -  e2  =  JcT,  and  squaring, 


By  Kepler's  third  law  T2/a3  is  constant  for  all  the  planets.  We 
must  therefore  have  (d)  /*  =  k2/a(l  -  e2)  =  47r2a3/T2,  a  constant. 

The  velocity  v  may  be  determined  in  the  following  way.  Let 
S  (Fig.  16)  be  the  origin  of  a  system  of  rectangular  coordinates, 
and  let  SA  be  the  a-axis.  We  have  x  =  rcosQ  and  y  =  rsin6,  and 
v2  =  z?  4-  if2.  From  the  equations 

(e)  x  =  r  cos  0  -  r  sin  0  .  0  ;     y  =  r  sin  0  +  r  cos  0  .  9, 

we  obtain  (f)  vz  =  r2  +  r262. 

If  we  substitute  for  r6  its  value  given  by  equations  (a)  and  (b), 
and  for  r  the  value  got  by  differentiating  equation  (b),  we  have 

v2  =  (l  +  2e  cos  (0  -  a)  +  e2)  .  &2/a2(l  -  e2)2. 

Noticing  that  1  +2ecos  (Q  -  a)  +  e2  =  2(l  +  e  cos  (6  -  a)J  -  (1  -e2),  we 
obtain,  by  the  help  of  equation  (b),  «2  =  (2/r-  I/a)  .  &2/a(l  -e2),  or, 
introducing  the  quantity  p.  defined  by  (d),  (g)  v2  =  2p,/r  —  p/a. 


30 


GENERAL  THEORY  OF  MOTION. 


[CHAP.  i. 


SECTION  X.    UNIVERSAL  ATTRACTION. 

We  owe  to  Newton  the  determination  of  the  law  of  the  force  which 
must  act  on  a  planet  in  order  that  its  motions  may  conform  to 
Kepler's  laws.  To  determine  this  force,  we  use  equation  (g)  IX. 
Let  the  centre  of  the  sun  be  the  origin  of  a  system  of  rectangular 
coordinates,  and  let  the  planet  be  situated  at  the  point  (z,  y) 
Represent  the  components  of  the  unknown  force  by  X  and  Y.  From 
the  law  of  kinetic  energy  (V.)  we  have  |#2  -  £t'02  =  {  Xdx  +  Ydy.  If  v0 
is  the  velocity  at  the  distance  r0,  we  obtain,  by  the  help  of  equation 
(g)  IX.,  fJL/r-fjL/rQ  =  \Xdx  +  Ydy.  If  Xdx+Ydy  is  a  complete  differ- 
ential d<f>,  we  will  have 

X  =  'tyfdx**'d(plrydx  and 
or         X  =  -  /x/V2  -  3r/ac  =  -  /iz/r8,     Y 

The  force  R  with  which  the  sun  acts  on  the  planet  is  R=  -  /*/r2, 
that  is,  is  a  force  which  is  inversely  proportional  to  the  square  of  the 
distance  of  the  planet  from  the  sun.  It  is  evident  from  equation  (d)  IX. 
that  the  quantity  p.  has  the  same  value  for  all  the  planets. 


FIG.  17. 

We  may  also  obtain  these  results  from  the  general  equations 
x  =  X  and  y=Y.  The  unknown  components  of  force  X  and  Y  may 
be  represented  by  the  lines  PA  and  PB  (Fig.  17),  and  resolved 
into  a  component  R  in  the  direction  SP  =  r  and  a  component  T 
perpendicular  to  SP.  Setting  -PSX  =  Qt  we  have 

T=  -  XsinQ+  Fcos6. 


SECT,  x.]  UNIVERSAL  ATTRACTION.  31 

By  the  help  of  equation  (e)  IX.,  this  becomes 

(d)  72  =  r-r62  and   T=2fQ  +  re=l/r .  d(r*6)ldt. 

But  since  r26  =  constant  by  Kepler's  first  law,  we  have  T=Q.     The 

attractive  force  is  therefore   directed  toward  the  sun.     Using   equations 

(a)  and  (b)  IX.,  we  obtain  (e)  R=  -£2/a(l -e2>'2  =  -  f*,'r2. 

We  apply  this  result  to  the  motion  of  the  moon.     By  reference 
to  (d)  IX.,  where  the  value  of  p.  is  given,  we  find  that 

R=  -4irV/ZV. 

The  orbit  of  the  moon  is  approximately  a  circle  with  a  radius 
60,27  times  as  great  as  that  of  the  earth.  Setting 

r=a=4.  109.60,27/27rcm, 
we  have  the  acceleration  7  of  the  moon  toward  the  earth, 

7  =  47r2a/r2  =  877 .  60,27  .  109/2  360  6002  cm, 

since  the  period  of  the  moon's  rotation  is  27,322  days  or  2  360  600 
seconds.  Hence  we  have  7  =  0,27183  cm.  If  the  centre  of  the 
moon  were  situated  at  the  distance  of  the  radius  of  the  earth  from 
the  earth's  centre,  it  would  have  an  acceleration  equal  to 


0,27183.60,272cm  =  987cm, 

assuming  that  the  force  is  inversely  proportional  to  the  square  of 
the  distance.  This  value  accords  so  well  with  that  of  the  accelera- 
tion at  the  surface  of  the  earth,  that  we  are  justified  in  assuming 
that  the  motion  of  a  falling  body  is  an  action  of  the  same  force  as 
that  which  keeps  the  moon  and  the  planets  in  their  orbits.  The 
final  proof  of  the  validity  of  Newton's  law  of  mass  attraction  is 
obtained  from  the  complete  agreement  of  the  theoretical  conclusions 
drawn  from  it  with  the  results  of  observations  on  the  heavenly 
bodies. 


SECTION  XI.    UNIVERSAL  ATTRACTION  (continued). 

We  will  now  use  a  method  precisely  the  opposite  of  our  former 
one.  We  will  assume  the  law  of  attraction  known,  and  determine 
the  path  of  a  planet  whose  position  and  velocity  at  the  time  t  =  0  are 
given.  Let  S  be  the  centre  of  the  sun  (Fig.  18),  let  the  attracted 
body  be  situated  at  A  at  the  time  £  =  0,  and  let  AC  represent  the 
velocity  r0,  whose  direction  makes  the  angle  CAD  =  </>  Avith  SA  =  r0 
produced.  If  the  acceleration  which  the  sun  imparts  to  the  planet 
is  set  equal  to  p./r2,  then,  using  a  system  of  polar  coordinates  whose 


32  GENERAL  THEORY  OF  MOTION.  [CHAP.  i_ 

origin  is  at  S,  and  writing  the  force  with  the  minus  sign  because 
it  is  directed  toward  the  sun,  we  obtain 
(a)  (b)  f  -  r62  =  -  fM/r2  and   1/r .  d(r*Q)/dt  =  0. 

It  follows  from  (b)  that  (c)  r29  =  &,  where  k  is  a  constant.  This  for- 
mula was  obtained  from  Kepler's  first  law,  (a)  IX.  Since  ?-6  is  the 
component  of  velocity  perpendicular  to  the  direction  of  r,  we  obtain 
for  t  =  Q,  (d)  k/r0  =  vQsm(f>.  By  the  help  of  equation  (c),  (a)  take& 
the  form  r  -  F/r3  =  -  /*/r2.  If  this  equation  is  multiplied  by  2i-dt, 
we  have  d(f3)+d(k2/r2)  =  2d(p/r),  and  by  integration 
r2  +  F/r2  =  2/i/r  +  Const. 


FIG.  18. 

In  the  initial  point  A,  we  have  v  =  v()  and  r  =  v0cos<f>,  therefore  for 
/  =  0,  we  obtain  ?'02cos2<£  +  F/r02  =  2fi/r0  +  Const.,  from  which  by  the 
help  of  (d)  it  follows  that  ?;02  =  2/*/r0  +  Const.     Hence  we  have 
(e)  r2  =  V-2f0o  +  2/Vr-F/r>. 

Since  the  velocity  v,  from  (f)  IX.,  may  be  expressed  generally  by 
2j  we  obtain  by  the  use  of  (c)  and  (e) 


This  equation  agrees  with  (g)  IX. 

The   same   result   may  be   derived   from   the   theorem    connectin 
kinetic  energy  and  work.     From  (e)  we  have 

(g)  r=±x/V-2ft 


where  the  upper  sign  is  to  be  taken,  if  r  and  t  increase  or  diminish 
together.  It  follows  from  equation  (c)  that  Q  =  k/r2.  Since  r  and  0 
depend  only  on  t,  we  obtain  from  (c)  and  (g) 

ft  .  d(l/r)/dB  =  +  VV-2/ 


SECT,  xi.]  UNIVERSAL  ATTRACTION.  33 

This  is  the  differential  equation  of  the  orbit.     By  adding  and  sub- 
tracting fj.2/k2  under  the  radical  sign,  and  by  noticing  that  p/k  is  a 
constant,   and   that  therefore   its   differential   is   zero,   this   equation 
may  be  written 
(h)  dQ  =  d(k/r  - 


If  we  set  M2  =  v0z  -  2/x/r0  +  p?/k2,  we  get  by  integration 
0  =  arc  cos  (k/ur  -  p/uk)  +  a, 

where  a  is  a  constant. 

Hence  the  equation  of  the  path  is 

(i)  l/r=(l+Jb//tco8(e-o))/<^//*),    . 

In   this  equation   u  may  always  be  considered   positive,    since   a  is 
arbitrary. 

The  polar  equation  of  a  conic  is  (k)  l/r=  (l  +e  cos  (0  -  a))/a(l  -e2), 
which  represents  an  ellipse,  a  parabola,  or  a  branch  of  an  hyperbola 
respectively,  according  as  e<l,  e  =  l  or  e>l.  If  e  =  0  we  have  the 
equation  of  the  circle.  From  the  equation  e  =  ku/p,  by  introducing 
the  value  of  «,  we  obtain  (1)  1  -e2  =  (2/»/r0-i>02)  .  &2//*2.  If  a  body 
approaches  the  sun  from  infinity  to  the  distance  r0,  its  velocity  will 
be  vv  determining  by  the  following  equation 


Therefore  we  have  (m)  e2  =  1  -  (v^  -  v02) .  F//*2.  Hence  the  path  is 
either  an  ellipse,  a  parabola,  or  an  hyperbola,  according  as 

«o<*'i»   vo  =  vi   or  vo>vi'> 

that  is,  the  path  of  the  body  is  an  ellipse,  parabola,  or  hyperbola, 
according  as  the  vis  viva  imparted  to  the  planet  at  the  first  instant 
is  too  small  to  send  it  to  infinity  against  the  attraction  of  the  sun, 
or  exactly  sufficient,  or  more  than  is  sufficient,  to  accomplish  that 
result. 

By  comparison  of  the  formulas  (i)  and  (k),  we  obtain 

(n)  o(l -#)  =  *»//*. 

This  corresponds  to  (d)  IX.  From  (m)  and  (n)  it  follows,  moreover, 
that  (o)  yu.  =  ±  (flj2  -  v02) .  a.  The  upper  sign  is  used  when  v1>v0, 
the  lower  when  v1<v0. 

In  the  first  case,  if  the  value  for  #x2  is  substituted  in  (o),  we 
have  #02  =  2/i/7-0  -  p-ja.  In  conjunction  with  (f)  this  equation  becomes 
v2  =  2/x/r  -  p/a,  which  corresponds  to  (g)  IX. 


34 


GENEEAL  THEOEY  OF  MOTION. 


[CHAP.  i. 


SECTION  XII.    THE  POTENTIAL  OF  A  SYSTEM  OF  MASSES. 

In  the  previous  discussion  Newton's  law  of  gravitation  was  derived 
from  Kepler's  laws  by  the  assumption  that  the  attraction  proceeds 
from  the  centre  of  the  sun,  or,  what  is  the  same  thing,  that  the 
whole  mass  of  the  sun  is  concentrated  at  its  centre.  A  similar 
assumption  was  made  in  the  case  of  the  planets.  These  assumptions 
might  be  made  without  further  demonstration  if  the  radius  of  the 
sun  were  infinitely  small  in  comparison  with  the  orbits  of  the  planets  ; 
since  this  is  not  the  case,  it  is  necessary  to  investigate  with  what 
force  a  mass  distributed  throughout  a  given  space  acts  on  a  body. 
This  problem  in  the  simplest  cases  was  solved  by  Newton.  His 
researches,  and  those  of  other  distinguished  mathematicians,  have 
led  to  results  of  the  greatest  importance,  both  in  physics  and 
mathematics.  The  method  by  which  such  problems  are  treated  is 
due  to  Laplace,  and  the  theory  was  developed  by  Poisson,  Green, 
Gauss,  and  others. 

Let  the  masses  mv  m2,  mB  be  situated  at  the  points  A,  B,  C  (Fig.  19), 
and  let  a  unit  of  mass  be  concentrated  at  the  point  whose  coordinates 
are  x,  y,  z.  The  force  with  which  the  unit  of  mass  is  attracted  by 

ml  is  -fmjr^  where  1\  is  the 
distance  AP,  and  /  a  constant 
dependent  on  the  units  of  mass, 
force,  and  length.  Call  the  co- 
ordinates of  A,  £v  rjv  fr  The 
components  Xv  Yv  Zl  of  the  force 
by  which  P  is  attracted  to  A  are 
evidently  X,  =  -fmjr,2.  (x  -  ^)/rv 
etc.  In  the  same  way  we  calcu- 
late the  components  of  the  forces 
which  originate  at  the  other 
points,  B,  C,  etc.  If  the  sum  of  all  the  X  components  is  repre- 
sented by  X,  we  obtain 

(a)  X=  -/K(z  -  fJW  +  m.(x  -  &)lr*  +...}. 

We  set  (b)  F=  mj^  +  w2/r2  +  m3/r3  +  .  .  .  . 

Now  V  =  (x  -  &)»  +  (y-^Y  +(*-  &)2, 

and  therefore  r^dr^fdx  =(x-  gj,  etc. 

Hence  we  have 


and  (c) 


X=f. 


SECT,  xii.]  POTENTIAL   OF  MASSES.  35 

In  a  similar  way  we  derive  the  equations 

(d)  (e)  F=/.3r/3y  and   Z  =f .  VP'I'dz. 

The  quantity  V  defined  by  equation  (b)  is  (VII.)  the  potential  at 
the  point  P  of  the  given  system  of  masses.  If  the  potential  is  given, 
the  equations  (c)  (d)  and  (e)  determine  the  forces  acting  in  the 
directions  of  the  coordinate  axes.  Since  the  position  of  the  system 
of  coordinates  is  arbitrary,  the  force  acting  in  any  direction  may  be 
derived  from  V.  This  has  been  already  shown  in  VII.  The  force 
acting  in  the  direction  s  is  therefore 

d F/ds  =  3 F/3z .  dx/ds  +  3F/fy .  dy/ds  +  9F/3* .  dz/ds. 

The  work  A  performed  by  the  force  in  moving  a  unit  of  mass 
along  an  arbitrarily  chosen  path  is  given  by 

A  =  l\Xdx  +  Ydy  +  Zdz), 

where  o  and  s  are  respectively  the  initial  and  final  points  of  the 
path.  If  the  values  given  in  the  formulas  (c)  (d)  (e)  are  substituted 
for  X,  Y,  and  Z,  and  the  element  of  the  path  whose  projections  on 
the  coordinate  axes  are  dx,  dy,  dz  is  designated  by  ds,  then 

A  =ff(d  Vj'dx .  dx/ds  +  3  FJ-dy .  dy/ds+  3  F/3z .  dz{ds)ds  =fj'd  V, 

and  hence  we  have  A=f(Tt-F0).  If  the  body  traverses  a  closed 
path,  the  work  done  by  the  forces  equals  zero  (cf.  VI.,  VIL). 
Therefore,  if  we  let  a  body  traverse  a  closed  path  under  the  action 
of  gravity,  the  work  which  gravity  performs  in  moving  the  body 
forward  is  equal  in  absolute  value  to  the  work  which  must  be 
performed  against  gravity  in  order  to  bring  the  body  back  to  the 
starting  point.  There  is  no  surplus  work  performed.  Hence  it  is 
evidently  impossible  to  produce  a  perpetuum  mobile,  that  is,  an 
arrangement  which  continuously  creates  work  out  of  nothing. 

We  have  assumed  that  the  masses  considered  are  concentrated  at 
points;  this,  however,  does  not  occur  in  nature.  Matter  is  more 
or  less  continuously  distributed  throughout  space  or  on  surfaces.  If 
it  is  uniformly  distributed  in  space,  the  mass  p  contained  in  the 
unit  of  volume  is  called  the  density.  If  it  is  not  uniformly 
distributed,  let  a  sphere  of  infinitely  small  radius  be  constructed 
about  the  point  P ;  the  ratio  of  the  mass  contained  in  the  sphere 
to  its  volume  is  the  volume  density  p  at  the  point  P.  If  the 
mass  is  distributed  over  a  surface,  the  surface  density  a-  at  the 
point  P  is  defined  by  the  ratio  of  the  mass  contained  within  a 
circle  of  infinitely  small  radius  drawn  about  the  point  P  as  centre 
to  the  area  of  the  circle. 


36  GENERAL  THEORY   OF  MOTION.  [CHAP.  i. 

If  the   mass  contained   in   the  unit  of  volume  is   p,  the  element 
of  volume  dw  will  contain  the  mass  pe?w.     The  potential  of  a  mass 
which  is  distributed  in  space  is  therefore,  from  equation  (b), 
(g)  F=JJJ^o,/r. 

This  integral  is  extended  over  the  whole  volume  occupied  by  the 
mass,  r  is  the  distance  between  dt»  and  the  point  for  which  the 
potential  is  to  be  determined. 

It  is  sometimes  necessary  to  consider  the  mass  as  distributed  in 
an  infinitely  thin  sheet  over  a  surface.  If  the  mass  on  the  unit  of 
surface  is  a-,  the  surface-element  dS  will  contain  the  mass  crdS. 
The  potential  takes  the  form  (h)  F=\\a-dS/r. 

The  potential  cannot  be  determined  without  some  further  informa- 
tion ;  in  the  next  paragraph  we  will  discuss  some  of  the  simplest  cases. 


SECTION  XIII.    EXAMPLES.    CALCULATION  OF  POTENTIALS. 

The  sun  and  planets  are  approximately  spherical.  On  the  sup- 
position that  they  are  spheres,  their  potential  can  be  easily  calculated, 
if  the  density  p  is  given,  and  if  we  assume  that  it  is  a  function  of 
the  radius,  and  therefore  has  the  same  value  for  all  parts  of  the 
concentric  spherical  layers  which  compose  the  sphere. 


FIG.  20. 
1.  The  Potential  of  an  Infinitely  Thin  Spherical  Shell  of  Constant  Surface 

Density  <r. 

Let  ABD  (Fig.  20)  be  a  sphere,  whose  centre  is  the  point  C  and 
whose  radius  is  R.  The  potential  at  the  point  0  is  to  be  deter- 
mined. If  we  set  OC  =  r,  L.OCB  =  <j>  and  OB  =  u,  we  have 


?=  r-2-n-R  sin  <£  .  Rd<i>  .  <r/u. 


SECT,  xin.]  CALCULATION  OF  POTENTIALS.  37 

Since  u2  =  r'2  +  R2-  2Rr  cos  $  and  udu  =  Rr  sin  <f>d(j>,  the  integral  takes 
the  form  V=  faR/r  .  udu/u  .  <r  =  2TrRa-/r.  \du.  If  0  lies  outside  the 
sphere,  we  have  \du  =  (r  +  R)  -  (r  -  R)  =  2  R  ;  while  if  it  lies  within 
the  sphere,  we  have  \du  =  (R  +  r)-(R-  r)  =  2r.  If  we  designate  the 
potential  outside  the  sphere  by  Va  and  that  within  the  sphere  by 
V-t,  we  have  (a)  Vi  =  IvRa-  •  Va  =  ^irR^a-jr.  The  potential  is  therefore 
constant  inside  the  spherical  shell;  and  at  points  outside  the  spherical  shell 
it  is  inversely  proportional  to  the  distance  from  its  centre.  Hence  the 
potential  for  the  whole  region  is  given  by  the  two  different  ex- 
pressions Va  and  V^  It  is  not  discontinuous  at  the  surface,  since 
for  r  =  R  we  have  Va=Vt  =  iirRo-.  On  the  other  hand,  its  differential 
coefficient  is  discontinuous  at  the  surface.  For  we  have 

dVJdr  =  -±irR?<r/r2  and 
and  therefore 

(b)  [dVJdr]r=Jt  =  -4™  and 

Hence  an  infinitely  thin  spherical  shell  exerts  no  force  at  a  point 
lying  within  it.  The  sphere  acts  only  on'  points  outside  of  it,  as  if 
its  whole  mass  were  concentrated  at  its  centre.  Hence  a  solid 
sphere  made  up  of  homogeneous  concentric  spherical  shells  acts  on 
outside  points  in  a  similar  manner.  If  the  attracted  point  is  situated 
within  the  mass  of  a  spherical  shell,  it  will  be  attracted  to  the  centre 
by  the  portion  of  the  mass  which  lies  within  a  sphere  described  about 
the  centre  with  the  distance  of  the  point  from  the  centre  as  radius. 
The  portion  lying  outside  this  surface  exerts  no  action. 

2.  The  Potential  of  a  Solid  'Sphere. 

We  will  now  calculate  the  potential  of  a  solid  sphere  of  constant 
density  p.     We  have  for  points  outside  the  sphere 


and  for  points  inside  the  sphere 

V,  =  f  47T.R2  .  dR  .  P/r  +  f  iirR  .  dR  .  P  =  ~r*P  +  2irp(R2  -  r2), 

•  0  ~r  O 

or  (d)  F—Zirp^2-^).  In  this  case  also,  the  potentials  within 
and  without  the  sphere  are  represented  by  two  different  expressions 
Vt  and  Va.  Both  values,  however,  coincide  at  the  surface,  since  for 
r  =  R  we  have  for  the  potential  Vi=Va  =  ^irR^p. 

The  function   Vt   which   represents   the   potential   of  a    mass  dis- 
tributed through    space,  is   everywhere   continuous.      Its  differential 
coefficients  with   respect  to  r  are 
(e)  dVt/dr=  -  frrp,  dVJdr=  - 


38  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

and  these  values  at  the  surface,  where  r  =  R,  are  also  equal,  that  is, 


Hence  the  first  derivatives  of  the  potential  of  a  mass  distributed  in 
space  are  nowhere  discontinuous,  but  are  continuous  throughout  all 
space.  On  the  other  hand,  its  second  derivatives  vary  continuously 
in  the  interior  and  the  exterior  regions,  but  on  passage  through  the 
spherical  surface  a  discontinuity  occurs,  That  is,  d2F'/dr2  at  the 
surface  has  two  values,  since 

[cVFJdr*]^^  -$irp  and  [dzFa/dr2]r=J!  =  +fay>. 
From  equation  (e)  the  force  outside  the  sphere  is  inversely  pro- 
portional to  the  square  of  the  distance  of  the  unit  of  mass  from 
the  centre  of  the  sphere.  We  thus  justify  the  assumption  that  the 
planets  and  the  sun  may  be  treated  as  points  in  which  their  re- 
spective masses  are  concentrated.  In  the  interior  of  the  sphere  the 
force  is  proportional  to  the  distance  of  the  attracted  point  from  the 
centre.  If  we  transform  equation  (e)  to  dFt/dr  =  -^m^p.l/r2,  we 
see  that  the  force  proceeds  from  that  portion  of  the  sphere  whose 
distance  from  the  centre  is  less  than  r.  This  only  holds  on  the 
assumption  made  about  p.  The  earth's  density  very  probably  increases 
toward  the  centre;  hence  the  force  of  gravity  will  not  have  its 
greatest  value  at  the  surface,  but  at  some  point  beneath  it.  This 
corresponds  with  the  results  of  experiments  on  the  time  of  vibration 
of  a  pendulum  in  a  deep  mine. 

3.  The  Potential  of  a  Circular  Plate. 

Let  AE  (Fig.  21)  be  a  circular  plate  of  surface  density  a-;  the 
centre  of  the  plate  is  0  and  the  axis  OP. 
The  point  P,  for  which  the  potential  is 
to  be  determined,  lies  on  the  axis  at  the 
distance  x  from  the  plate.  The  potential 

V  is  then 

f* 
F=  I    2irr) .  drj .  o-/<u, 

where  R  is  the  radius  of  the  plate  and 
7;  and  u  are  the  distances  of  a  point  on 
the  plate  from  0  and  P  respectively. 
We  have  v?  =  rf  +  x2,  therefore  ndu»i)d^t 
and  hence 

IG'  V=  fadu  .<r=27ra-(p-  x), 

if  p  is  the  distance  of  the  point  P  from  the  edge  of  the  plate.     If 


SECT.  XIII.] 


CALCULATION  OF  POTENTIALS. 


39 


(f) 


the  x  drawn  from  one  face  of  the   plate  is  considered  positive,  the 
potential  for  negative  values  of  x  is  Pr=2Trcr(p  +  x).     Hence 
Jfora;>0,     Vl  =  2ir(r(p-x) 

If  the  radius  of  the  plate  is  infinitely  great  in  comparison  with  x, 
we  may  set  Prl  =  C-2Tra-x  and  F2—C+2iro-x,  where  C  is  an  infinitely 
great  constant,  since  p  is  infinitely  great  and  o-  remains  finite.  We 
have  for  x>Q,  dVljdx  =  -2ircr, 

and  for  x<0,    d  F2/dx  =  +  27nr. 

By  passage  through  the  surface,  dF/dx,  that  is,   the  force,  changes  dis- 
continuously  by  47rcr. 


4.  The  Potential  of  an  infinitely  long  straight  line. 

Suppose  each  unit  of  length  of  the  line  AB  (Fig.  22)  to  have  the 

.     Let  G  be  a  point  at  the 
distance   a  from   AB,  and   CD  the     "B 
perpendicular  let  fall  from  C  upon 
AB.     The  potential  V,  at  the  point 
(7,  is 


V=  2^  log  (z'/a 

Since  z'  is  infinitely  great  in  com- 
parison with  a,  we  may  neglect  1 
under  the  radical  and  write 

(k)  V=  2p  log  (  2z'/a)  =  C.-p  log  a2,    . 

where  C  is  an  infinitely  great  con- 
stant, if  z'  is  infinitely  great. 
Further,  we  obtain 

dF/da=  -2/*/a, 

that    is,    the  force    is    inversely  pro-  FlG'  22> 

portioned  to  the  distance  of  the  point  from  the  straight  line. 


5.  The  Potential  of  a  Circular 

Represent  the  surface  density  of  a  circular  cylinder  by  a-.     Through 
the  point  P,  for  which  the  potential  is  to  be  determined,  pass  a  plane 


40  GENERAL  THEOEY  OF  MOTION.  [CHAP.  i. 

perpendicular  to  the  axis  of  the  cylinder  (Fig.  23).  Let  R  be  the 
radius  of  the  cross  section  of  the  cylinder  and  r  the  distance  of  the 
point  P  from  its  centre. 


FIG.  23. 


We  then  have  from  (k) 

v=  c-^JJme.a-.  log  a2. 

We  now  find  the  value  of  the  integral 


in  which  a  =  *Jf2  -  2rE  cos  6  +  If2.      First  consider  the  case  in  which 
r>R.     The  integral  may  then  be  written,  if  we  set  a  =  E/r, 

rir  _ 

^  =  2  Jo  Rd6(\og  r  +  log  Vl  -  2a  cos  0  +  a2). 
Since  cos  6  =  \(&9  +  «-**),  we  have  1  -  2a  cos  0  +  a2  =  (1  -  a.ei6}(  1  -  ae~ie), 


and  jf 


Now  developing  the  terms  in  this  integral  in  series,  and  carrying 
out  the  integration,  we  find  that  the  integral  is  equal  to  zero, 
and  hence  that  A  =  Z-n-R  log  r.  Thus  the  mean  value  of  log  a  for 
all  points  of  the  circumference  of  the  circle  is  equal  to  log  r  or  to 
the  logarithm  of  the  mean  distance  from  P  to  the  circumference  of  the 
circle. 

If  now  V<  R,  that  is,  if  P  lies  between  0  and  the  circumference, 
and  if  we  set  a  =  rjR,  we  have 


=  2       Rd6(\0g  R  +  log  N/l-2aCOS0+a2), 

and  hence  A  =  2-rrE  log  R.     In  this  case  also  the  mean  value  of  log  a 


SECT,  xiii.]  CALCULATION  OF  POTENTIALS.  41 

is  equal  to  the  logarithm  of  the  mean  distance  from  P  to  the  circum- 
ference of  the  circle. 

Now  setting  Vn  and  Vi  for  the  potentials   of   points  outside  and 
inside  the  cylinder  respectively,  we  have  from  these  values, 
ra=C-  47r^o-  log  r,     Vi  =  c-  lirEo-  log  R. 

The  potential  is  therefore  constant,  and  the  force  zero  within  the 
cylinder.     Outside  the  cylinder  the  force  is  given  by 
(n)  d7Jdr=  -  4irftr/r, 

that   is,   the  force   is   inversely  proportional  to  the  distance  of  the 
point  from  the  axis  of  the  cylinder. 


SECTION  XIV.     GAUSS'S  THEOREM.     THE  EQUATIONS  or  LAPLACE 
AND  POISSON. 

Let  ABF  (Fig.  24)  be  a  closed  surface,  of  which  AB  =  dS  is  a 
surface-element,  and  at  the  point  0  within  the  surface,  let  the  mass 
m  be  concentrated.  On  the  element  ds  at  C,  construct  the  normal 


CE.  Let  the  length  of  the  line  connecting  0  and  C  be  r,  and  let 
the  normal  CE  make  the  angle  DCE  =  Q  with  00  produced.  If  the 
potential  at  C  due  to  ml  is  Vv  then  Prl  =  ml/r,  and  the  force  ^ 
acting  at  the  point  C  in  the  direction  CE  is  Nl  =  'dVl[dn)  while  the 
total  force  in  the  direction  of  CO  is  mjr2.  We  have 
(a)  JVt  =  wij/r2  .  cos  («•  -  6)  =  -  mjr2 .  cos  0. 


42 


GENEEAL  THEORY  OF  MOTION. 


[CHAP.  i. 


If  a   sphere   of  unit   radius    is    described   about   the   point  0  as 
centre,  the  straight  lines  drawn  to  the  contour  of  (IS  mark  out  on 
this  unit  sphere  a  surface-element  whose  magnitude  is  equal  to 
(b)  du  =  dS.cose/r*. 

From  (a)  and  (b)  we  obtain 

NjdS=  -  mjr2 .  cos  QdS=  -  m^w  and  'dPJ'dn  .  dS=  -  mfa. 
If   there  are   still   other  masses,   m^    my    etc.,    within   the    closed 
surface,  we  obtain  similarly 

3  fy3» .  dS  =  -  m^u,     "d  V^^n, .  dS  =  -  msdo>, .... 
Vv  Vy  Fs  are  the  potentials  at  the  point   C  due  to  the  masses 
mv  m2,  ms  respectively.     For  the  total  potential  at  the  point  C  we 
~  ...,  and  therefore 

.dS=  - 


If  we  designate  the  mass  enclosed  by  the  surface  by  2m,  integration 
over  the  whole  surface  gives  (c)  foV/'dn.  dS=  -  47r2m.  The  force 
acting  in  the  direction  of  the  normal  to  the  surface  S  is  'dF'/'dn ;  we 
call  'dF'/'dn.  dS  the  flux  of  force  which  passes  through  the  element  dS. 
Hence  the  total  flux  of  force  passing  through  a  finite  closed  surface  equals 
the  sum  of  the  acting  masses  contained  within  the  surface  multiplied 
by  -  4:ir.  Hence,  if  the  entire  acting  mass  is  enclosed  by  the  surface, 
and  'dVfdn  is  given  for  all  points  on  the  surface,  the  sum  of  the 
masses  may  be  determined  by  the  help  of  equation  (c). 


SECT,  xiv.]  GAUSS'S  THEOREM.  43 

The  theorem  expressed  in  (c)  also  holds  in  case  the  acting  mass  lies 
outside  the  closed  surface.  Let  the  mass  m'  be  situated  at  the  point 
0'  (Fig.  25)  outside  the  surface  ABB' A.  If  the  surface-element  do> 
is  taken  on  the  surface  of  the  unit  sphere  described  about  0'  as 
centre,  the  straight  lines  drawn  from  0'  through  the  boundary  of  this 
surface-element  mark  out  on  the  closed  surface  the  surface-elements 
AB  =  dS  and  A'B'  =  dS'.  Let  the  normals  to  AB  and  A'ff  directed 
outward  from  the  closed  surface  be  n  and  n'  respectively,  and  let 
the  forces  'dF'/'dn  and  'dF'/'dn'  act  in  the  direction  of  these  normals. 
The  V  in  these  expressions  represents  the  potential  due  to  m'.  If 
the  angles  made  by  the  normals  directed  outward  and  the  straight 
line  drawn  from  0'  are  designated  by  6  and  6'  respectively,  and  if 
we  set  0'A=r,  0'A'  =  r',  we  then  obtain 

-dF'fdn  =  m'/r2 .  cos  (a-  -  9) ;     'dF'/^n'  =  m'/r'2 .  cos  (a-  -  6'), 
dS .  cos  (TT  -  6)  =  r2du ;     dS'  cos  6'  =  r'2 .  du, 

and  therefore  Wpn.dS  +  W/dri  .dS' =  0.  We  therefore  have  (d) 
\?)V'/'dn.dS=Q,  if  the  integral  is  extended  over  the  whole  surface. 
The  flux  of  force  proceeding  from  a  point  outside  a  closed  surface,  and 
passing  through  the  surface,  is  equal  to  zero.  Therefore  the  value  of  the 
integral  is  independent  of  the  mass  outside  the  surface.  We  have 
then,  generally,  (e)  foF'fdn.dS**  -4-jrM,  where  S  is  a  closed  surface, 
V  the  potential,  n  the  normal  directed  outward,  and  M  the  sum  of 
all  the  masses  within  the  surface.  This  theorem  is  due  to  Gauss. 
Equation  (e)  may  be  put  into  another  form.  We  have 

3  Ffdn  =  3  Vfdx .  dx/dn  +  3  F/^ .  dy/dn  +  3  V\"bz .  dz/dn, 

and  dx/dn  =  A,  dyjdn  =  p,  dzldn  =  v,  where  A,  /*,  and  v  are  the  cosines  of 
the  angles  which  the  normal  to  the  surface  makes  with  the  axes. 
We  have  then  3F/3tt  =  \X  +  ^Y+  vZ.  X,  Y,  and  Z  are  the  com- 
ponents of  the  force,  and  /  is  set  equal  to  1.  We  then  obtain  from 
Gauss's  theorem  (f)  J(ZX  +  Yp  +  Zv)dS=  -  1-rrM. 

Let  x,  y,  z  (Fig.  26)  be  the  coordinates  of  the  point  0,  Ox,  Oy,  and  Oz 
be  parallel  to  the  coordinate  axes,  and  00'  be  a  parallelepiped  whose 
edges  are  parallel  to  these  axes.  Let  X,  Y,  Z  be  the  components  of 
the  force  acting  at  0.  The  components  of  the  force  at  the  point  A, 
whose  coordinates  are  x  +  dx,  y,  z,  will  be 

X+VXfdx.dx,  Y+'dY/'dx.dx,  Z+-dZ/Vx.dx. 

We  apply  Gauss's  theorem  to  the  surface  of  this  parallelepiped.  The 
force  acting  normal  to  the  surface  OA'  is  -  X,  that  acting  normal 
to  AO'  is  +X+~dX/c)x.dx.  In  the  same  way  the  force  acting  normal 


44 


GENERAL  THEORY  OF  MOTION. 


[CHAP.  i. 


to  OB1  is    -Y,  and  that  acting  normal  to  BO'  is  Y+'dY/'dy.d 
similar  statement  holds  for  the  z  coordinate.     We  have  therefore 


\-dVfdn,  .  dS  =  Jf  J[  -  Xdydz  +  (X+  VX/Vx  .  dx)dydz] 
+  [  -  Ydxdz  +  (Y+  VY/oy  .  dy)dzdx] 


We  suppose  the  volume-element  00'  to  contain  the  mass  M  of  density 
/>,  so  that  M=pdxdydz.     We  then  have  from  (e) 
(g) 
or  (h) 


On  account  of  the  frequent  use  made  of  this  equation  in  mathe- 
matical physics,  we  use  for  the  sum  of  the  first  derivatives  of  a 
function  /  with  respect  to  the  three  coordinates  the  symbol 


and  for  the  sum  of  the  second  derivatives  with  respect  to  the  same 
variables  the  symbol  V2/=  32//a«2  +  32//c)?/2  +  32//3«2.    With  this  notation 


SECT,  xiv.]  GAUSS'S  THEOREM.  45 

equation  (h)  may  be  written  (i)  V2  F+  4ny>  =  0.  By  the  help  of  this 
equation,  which  was  first  used  by  Poisson,  we  can  determine  the 
density  when  the  potential  is  known.  If  no  matter  is  present  in 
the  region  under  consideration,  that  is,  if  p  =  0,  we  have 

(k)  32r/9a;2  +  32F/3y2  +  32r/o^  =  V2r=0. 

This  equation  was  first  derived  by  Laplace.  It  may  be  obtained 
more  simply  in  the  following  way.  We  start  from 


where  £,  i],  and  £  are  constant,  and  obtain 

3(l/r)/a*  =  -  (x  -  £)/r»,     ^(l/r)/^  =  -  l/r»  +  3(a?  - 

Analogous  expressions  hold  for  38(l/r)/8jy2  and  32(l/r)<322.      Adding 
these  equations,  we  have 

3«2  =  0. 


Since  the  potential  F=2w/r  [(b)  XII.],  this  is  equivalent  to  V2F=0. 

Poisson's  equation  may  also  be  obtained  in  the  following  way. 
Let  the  density  at  the  point  P  be  p.  Describe  a  sphere  of  infinitely 
small  radius  R  so  as  to  contain  the  point  P,  and  suppose  the  density 
in  the  interior  of  the  sphere  to  be  constant.  The  potential  P  at 
the  point  P  consists  of  two  parts,  Fj  and  F~a  ;  Va  is  due  to  the  mass 
outside  the  sphere,  and  Vi  to  the  mass  within  the  sphere.  The 
potential  at  P  is  V=Va+Vi.  If  P  is  at  the  distance  r  from  the 
centre  of  the  sphere,  we  have  from  (d)  XIII. 
(k')  Ft  =  2aX#»  -  £r2)  ;  V=  Va  +  **p(&  -  &*)> 

If  £,  rj,  £  and  x,  y,  z  are  the  coordinates  of  the  centre  of  the  sphere 
and  of  P  respectively,  we  will  have  r2  =  (x  -  £)2  +  (y  -  r))2  +  (z  -  £)2. 
By  differentiation  with  respect  to  x,  we  obtain 

3(r2)/ae  =  2(x  -  £)   and   92(r2)/ae2  =  2, 

therefore  W  =  6,  and  from  equation  (k')  V*V=V*Va-lvp.  Now  Va 
is  the  potential  due  to  the  mass  lying  outside  of  the  sphere,  and 
therefore  V2F"a  =  0  and  V2F+47r/a  =  0.  This  is  Poisson's  equation. 

In  the  parts  of  the  region  where  p  is  infinitely  great,  Poisson's 
equation  loses  its  meaning.  In  this  case  we  return  to  the  fundamental 
equation  (e).  For  example,  let  a  mass  be  distributed  on  a  surface  S 
with  surface  density  a-.  Draw  the  normals  vt  and  va  to  the  element 
dS  on  both  sides  of  the  surface,  and  construct  right  cylinders  on 
both  sides  of  the  surface  on  dS  as  base,  and  with  the  heights  dvi 
and  dva  ;  the  linear  elements  of  these  cylinders  are  lines  of  force.  By 
applying  equation  (e)  to  the  volume  included  in  the  cylinders,  we 


46  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 


obtain  ?>V^vi  .  dS+'dFJ'dva.  dS=  -4WS,  where  Vt  and  Va  represent 
the  values  of  the  potential  on  both  sides  of  the  surface.     Hence 

(1)  3r«/3v,  +  ^Val^va  +  lira-  =  0. 

This  equation  finds  an  application  in  the  theory  of  electricity. 

Comparing  formulas  (e)  and  (h),  and  noticing  that  M=  ^^pdxdydz, 
we  obtain  the  relation  (m)  \\\VzVdxdydz  =  Jf3F/3».  dS.  The  triple 
integral  in  (m)  must  be  extended  over  the  volume  bounded  by  the 
surface  S,  and  the  double  integral  over  the  surface  S.  This  theorem 
may  also  be  proved  by  integration  by  parts. 


SECTION  XV.     EXAMPLES  OF  THE  APPLICATION  OF  LAPLACE'S  AND 
POISSON'S  EQUATIONS. 

The  potential  V  at  the   point  x,  yt  z  is   a  function  of  the  three 
coordinates,  and,  from  the  previous  discussion,  has  the  form 


(a)  V=  Mpd&vdt/J(z  -  &  +  (y-  ^  +  (*  -  02> 

where   the   density   p   at  the   point   (£,  rj,   £)   is   a   function   of  the 

coordinates. 

We  may,  however,  use  the  differential  equation  (b)  V2F+47r/>  =  0 
as  the  starting  point  for  the  determination  of  the  potential;  we 
thus  often  obtain  the  desired  result  by  a  more  convenient  method. 
The  density  p  must  be  given  as  a  function  of  x,  y,  and  z.  The 
integral  of  (b)  is  always  given  by  (a),  but  V  may  often  be  found 
more  conveniently  by  direct  integration  of  Poisson's  equation. 

In  the  solution  of  problems  in  potential,  special  attention  must  be 
paid  to  the  boundary  conditions  which  serve  to  determine  the  functions 
which  are  obtained  by  integration.  We  shall  investigate  the  equations 
of  condition  to  which  the  potential  Vt  within  a  closed  surface  S  and 
the  potential  Va  outside  that  surface  must  conform,  if  the  surface  S 
encloses  all  the  masses  which  are  present  in  the  field,  and  if  no 
mass  is  present  outside  the  surface.  Applying  Poisson's  equation 
to  the  region  enclosed  by  #,  we  have  (c)  V2J^  +  47r/>  =  0.  Outside 
the  surface  S  we  have  (d)  V2F"a  =  0. 

If  0  is  any  point  within  S,  P  a,  point  outside  S,  and  if  we  set 
OP  =  r,  we  have,  when  r  is  very  great,  (e)  Fa  =  M/r.  M  represents 
the  whole  mass  enclosed  by  S.  Hence,  for  r  =  <x>,  the  potential 
Ftt  =  0  and  (f)  lim(rFa)r=00  =M,  that  is,  the  product  rVa  approaches 
the  finite  limit  M  if  the  point  P  moves  off  to  infinity. 


SECT,  xv.]       LAPLACE'S  AND  POISSON'S  EQUATIONS.  47 

If  Pl  and  P2  are  two  points  which  lie  infinitely  near  each  other 
on  different  sides  of  the  surface  S,  the  potentials  at  both  points  are 
equal,  and  we  have  for  all  points  on  the  surface  S,  (g)  Vi  =  Fa.  The 
dash  drawn  over  V  is  used  to  denote  the  value  of  V  at  the  surface. 

From  (1)  XIV.  it  follows  further  that  for  the  points  on  the  surface 
where  o-  =  0,  we  have  (h)  'dFi/'dv  =  'dFa/?)v,  where  the  normal  to  S  is 
designated  by  v  =  -  v;  =  va.  The  potential  is  therefore  everywhere 
finite. 

It  is  here  assumed  that  p  is  everywhere  finite.  For  the  places 
where  p  =  so  we  obtain  other  equations  of  condition,  which  may 
readily  be  derived  from  those  already  given.  For  example,  if  o-  is 
the  surface  density  on  a  surface  S,  in  which,  therefore,  p  is  infinitely 
great,  and  if  p  =  Q  for  all  other  points  in  the  region,  then,  in  our 
former  notation,  we  have  V2F"j  =  0  and  V2F"tt  =  0,  but 
(i)  Vi=V«  Wtl'dvl  +  Wjdva+4:ira-  =  Q 

for  all  points  on  the  surface  S. 

By  these  equations  we  may  determine  the  potential  if  the  density 
p  within  a  sphere  of  radius  R  is  constant.  Outside  the  sphere  p 
is  supposed  to  be  zero.  The  potential  within  the  sphere  is  Ft,  and 
outside  of  it  Fa.  We  have  then  W,  +  4ir/o  =  0,  V2Fft  =  0.  Now 
we  have  'dF/'dx  =  dFfdr  .  x/r, 

&Ffda?  =  d*Fjdr*  .  x2/r2  +  dFjdr  .  1/r  -dF/dr  .  x2/^. 
Similar  equations  hold   for  the  derivatives  of  V  with  respect  to  y 
and  z.     We  thus  obtain  V2F=d?F/dr2  +  2/r  .  dFjdr.     Since,  however, 

d(r  F){dr  =  rd  F/dr  +  V  and  d2(rF)/dr2  =  rd2  F/dr2  +  2d  F/dr, 
we  have  (1)  V2F=  1/r  .  d2(rF)/dr2. 

The  differential  equations  which  Vi  and  Fa  must  satisfy  are  therefore 
(m)  d2(rFt)ldr2  +  l-rpr  =  0,     d2(r  Fa)/dr*  =  0. 

From  these  we  obtain  by  integration 


For  r  =  ao    it  is  assumed   that   Fa  =  0,  so  that    Fa  =  Cl'/r.     Since   Ft 
cannot  become  infinite  for  r  =  0,  we  have  (7  =  0,  and  hence 


Since  the  force  is  a  continuous  function  of  the  coordinates,  arid 
since,  therefore,  for  all  points  on  the  surface,  dFi/dr  =  dFJdr,  we 
will  also  have,  when  r  =  B,  ^TrpR  =  Cl'/R2.  Hence  C^^-n-pB3,  and 
therefore  (n)  Fa  =  £irEsp/r.  Since  F"4=Fa  when  r  =  R,  we  have 
C=27rR2p,  and  therefore  (o)  F^^^R2-^-2).  These  formulas  are 
the  same  as  (c)  and  (d)  XIII. 


48  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

If  the  potential  depends  on  the  distance  of  the  point  under  con- 
sideration from  a  straight  line,  we  choose  this  line  as  the  £-axis  of 
a  system  of  rectangular  coordinates.  Let  the  distance  from  the 
2-axis  of  the  point  for  which  the  potential  is  to  be  determined 
be  r.  We  have  then  r~  =  x2  +  y-,  and  further 


x*/r2  .  dtF/dr*  +  l/r  .  dFfdr  -  x2/^  .  dF/dr,  etc. 
Therefore 
(p)  ^Y^^YI^  +  ijr  .  dp/fir  =  i  IT  t  d(rdF/dr)/dr. 

If  we  are  dealing  with  an  infinitely  long  circular  cylinder  of 
radius  E  and  surface-density  a-,  the  axis  of  which  is  taken  as  the 
z-axis,  we  have  (q)  VsFi  =  0  and  V2F"0  =  0,  while  for  r  =  R  we  have 

(r)  Ft=FM     d  FJdr  -  dFt/dr  =  -  47ro-. 

It  follows  from  equations  (p)  and  (q)  that 

d(rdFt/dr)/dr  =  Q  and  d(rdfra/dr)ldr  =  0. 
Hence  dFt/dr=Cl/r  and  dFa/dr  =  C2/r, 

ri  =  (71logr+(71'  and    Fa=C2\ogr+C2. 

GI  must  be  equal  to  zero,  since  no  force  acts  at  points  in  the  axis 
of  the    cylinder.      Further,    for   r  =  R,    we    have    C^  =  C.2  log  E  +  C2. 
From  equation  (r)  we  have  C2=  -47rKo-,  and  therefore 
V.  =  C2  -  lirRo-  log  R;     Fa  =  C2-  iirRo-  log  r. 
These  equations  are  the  same  as  those  given  in  (m)  XIII. 


SECTION  XVI.     ACTION  AND  REACTION.    ON  THE  MOLECULAR  AND 
ATOMIC  STRUCTURE  OF  BODIES. 

In  our  discussions  up  to  this  point  we  have  considered  the  motion 
of  a  body  under  the  action  of  given  forces ;  but  nothing  has  yet 
been  said  as  to  the  origin  of  these  forces.  A  body  upon  which  no 
forces  act  moves  forward,  by  the  principle  of  inertia,  in  a  straight 
line  with  a  uniform  velocity.  A  change  in  the  motion  can  arise 
only  from  outside  causes.  We  learn  from  experience  that  the  motion 
of  one  body  in  the  presence  of  another  undergoes  a  change,  and  we 
are  therefore  led  to  assume  that  in  the  mutual  action  of  these  bodies 
is  to  be  found  the  reason  for  the  change  of  motion.  We  will  first 
consider  the  mutual  action  of  two  bodies.  We  thus  obtain  the  means 
of  investigating  the  more  general  case  in  which  three  or  more  bodies 


SECT,  xvi.]  ACTION  AND   REACTION.  49 

act  on  one  another.  The  mutual  action  may  be  of  different  kinds. 
If  two  bodies  collide  their  motion  changes.  A  similar  change  occurs 
when  the  bodies  slide  over  each  other.  In  both  cases  the  bodies 
are  at  least  momentarily  in  contact.  Bodies  also  act  on  each  other 
without  contact;  thus,  for  example,  a  magnet  attracts  a  piece  of 
iron,  or  a  piece  of  amber  when  rubbed  attracts  a  feather.  The 
first  serious  eifort  to  explain  these  mutual  actions  or  so-called  actions 
at  a  distance  was  made  by  Descartes.  His  explanation  was  based 
on  the  assumption  that  all  space  is  filled  with  very  small  particles 
in  motion,  and  that  all  observed  motions  of  bodies  are  due  to 
collisions  between  them  and  these  invisible  particles. 

Hence  the  discovery  of  the  laws  of  collision  became  one  of  the 
most  important  tasks  in  the  study  of  physics.  Descartes  investigated 
this  question,  but  without  success.  It  was  not  until  the  close  of 
the  17th  century  that  Huygens,  Wallis,  and  Wren  contemporaneously 
succeeded  in  solving  it.  A  sphere  in  motion  can  set  in  motion  a 
sphere  at  rest;  the  moving  sphere,  therefore,  possesses  energy  of 
itself.  Let  the  collision  be  central,  that  is,  let  the  direction  of  motion 
coincide  with  the  line  joining  the  centres  of  the  spheres.  An  iron 
sphere  produces  a  greater  effect  on  the  sphere  at  rest  than  a  wooden 
sphere  of  equal  size  moving  with  the  same  velocity.  Of  two  equally 
large  spheres  whose  mass  is  the  same,  the  one  produces  the  greater 
effect  which  has  the  greater  velocity.  Hence  the  force  which  the 
moving  sphere  possesses  increases  with  its  mass  and  with  its  velocity 
jointly.  The  product  of  the  mass  and  the  velocity  gives  a  measure 
for  the  force  residing  in  the  body,  and  is  called  its  momentum,  or 
quantity  of  motion. 

The  principal  result  which  Huygens,  Wallis,  and  others  obtained 
was  the  following  :  If  two  bodies  collide,  they  undergo  changes  of 
momentum  which  are  equally  great  and  in  opposite  directions,  or, 
they  act  on  each  other  with  equal  but  oppositely  directed  forces. 
The  action  and  reaction  are  therefore  equal  and  oppositely  directed. 

This  is  one  of  the  most  important  laws  of  natural  philosophy, 
and  we  will  discuss  the  grounds  upon  which  it  is  founded.  It  was 
first  derived  from  observations  on  collision,  without  its  thereby 
becoming  apparent  how  far  it  holds  for  other  interactions  between 
bodies.  Xewton  first  recognized  in  this  law  a  universal  law  of 
nature,  which  always  applies  when  bodies  act  on  one  another.  By 
careful  investigation  of  the  collisions  of  different  bodies  (steel,  glass, 
wool,  cork)  he  found  that  the  action  and  reaction  are  equal,  if 
allowance  is  made  for  the  resistance  of  the  air.  In  order  to  examine 

D 


50  GENEKAL  THEORY   OF  MOTION.  [CHAP.  i. 

whether  the  same  law  holds  for  actions  at  a  distance,  he  mounted 
a  magnet  and  a  piece  of  iron  on  corks  and  floated  them  on  water. 
The  iron  and  the  magnet  approached  each  other  and  remained  at 
rest  after  they  had  come  in  contact,  so  that  the  forces  by  which 
the  iron  and  the  magnet  were  mutually  attracted  were  oppositely 
directed,  and  equal.  He  showed  further  by  the  following  argument 
that  action  and  reaction  are  equal  in  the  case  of  attraction  or 
repulsion :  If  two  bodies  acting  on  each  other  are  rigidly  connected, 
they  should  both  move  in  the  direction  of  the  greater  force  if  action 
and  reaction  were  not  equal ;  this  would  contradict  the  principle  of 
inertia. 

Since  Newton's  time  this  law  has  been  established  in  many  ways, 
and  many  discoveries  in  physics  have  furnished  proofs  of  its  correct- 
ness. It  has  led  in  many  cases  to  new  discoveries,  and  there  is  no 
longer  any  doubt  of  its  universal  applicability. 

The  simplest  conception  of  the  structure  of  bodies  is  that,  according 
to  which  bodies  are  composed  of  discrete  particles,  for  whose  mutual 
action  the  law  of  action  and  reaction  holds.  Starting  from  this 
view,  Newton  calculated  the  action  of  gravity.  Gravity  is  a  function 
of  distance  alone ;  its  value  is  therefore  the  same  so  long  as  the 
distance  is  unchanged.  This  conception  of  the  structure  of  bodies 
has  led  to  important  results  in  other  branches  of  physics.  There 
are,  however,  many  cases  in  which  it  seems  inadequate.  Chemistry 
teaches  that  bodies  are  composed  of  molecules,  which  themselves 
may  be  groups  of  smaller  particles  or  atoms.  These  molecules  have 
certainly  a  very  complex  structure,  and  the  mutual  actions  among 
them,  especially  if  the  distances  between  them  are  great  in  comparison 
with  their  size,  must  therefore  be  of  a  very  complicated  nature.  As 
yet  we  have  little  knowledge  on  this  subject.  In  what  follows 
we  will  confine  ourselves  to  the  treatment  of  the  motions  of  particles 
acting  on  each  other  with  forces  which  are  functions  only  of  the 
distances  between  them. 


SECTION  XVII.     THE  CENTRE  OF  GRAVITY. 

Gravity  acts  on  all  parts  of  a  body;  the  forces  thus  arising  may 
be  considered  parallel  for  all  parts  of  the  same  body.  The  action 
of  gravity  on  all  the  particles  of  a  body  may  be  combined  in  a 
resultant  whose  point  of  application  is  at  the  centre  of  gravity.  If 
the  centre  of  gravity  is  rigidly  connected  with  the  body  and  rests 


SECT.  XVII.] 


THE  CENTRE  OF  GRAVITY. 


51 


on  a  support,  the  body  is  in  equilibrium  in  any  position.  Since 
gravity  is  proportional  to  the  mass,  the  centre  of  gravity  coincides 
with  the  centre  of  mass.  The  resultant  applied  at  the  centre  of  gravity 
is  the  weight  of  the  body.  The  straight  lever  is  in  equilibrium  if  there 
is  applied  to  its  centre  of  gravity  a  force  equal  to  its  weight  and 
acting  in  the  opposite  direction ;  the  particles  on  the  one  side  of 
the  centre  of  gravity  tend  by  their  weight  to  produce  rotation  in 
one  sense  which  is  equal  to  that  produced  in  the  opposite  sense  by 
the  particles  on  the  other  side. 

Let  the  masses  rax  and  m2,  whose  velocities  are  represented  by  A  A' 
and  BB'  (Fig.  27),  be  situated  at  the  points  A  and  B.  Let  the  point 
C  be  so  determined  on  the  line  joining  A  and  B  that  mlAC  =  m.2BC. 
The  point  C  is  then  .called  the  centre  of  gravity  of  the  two  masses 
ml  and  m2.  If  the  point  C'  is  so  determined  on  the  line  joining 
m2B'C',  we  may  consider  CC'  the  velocity  of 


FIG.  28. 

tJie  centre  of  gravity.  If  AD  and  BE  are  equal  and  parallel  to  CO', 
the  velocity  AA  of  the  mass  ml  may  be  resolved  into  the  components 
AD  and  DA',  and  similarly  the  velocity  BB'  may  be  resolved  into 
the  components  BE  and  EB'.  Now,  since  ml/m.2  =  BC/AC=B'C'/A'C', 
the  triangles  A'CfD  and  B'C'E  are  similar,  the  sides  A'D  and  B'E  are 
parallel,  and  hence  ml/m2  =  B'EIA'D^  The  velocities  of  the  masses 
may  be  considered  as  compounded  of  the  velocity  of  the  common 
centre  of  gravity  and  two  velocities  vl  and  v2,  which  are  parallel 
to  each  other  and  inversely  proportional  to  the  masses  ;  so  that 


If,  therefore,  oa  and  ob  (Fig.  28)  represent  the  velocities  of  the 
masses  m-^  and  m2,  and  if  ab  is  divided  by  the  point  c  into  the 
parts  ac  and  be,  which  are  inversely  proportional  to  the  masses,  then 
oc  represents  the  velocity  of  the  centre  of  gravity,  and  ca  and  cb 


52  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

represent  the  velocities  of  the  masses  ml  and  m.2  relative  to  the  centre 
of  gravity.  It  is  convenient  to  resolve  the  velocity  in  this  way, 
because  the  velocity  of  the  centre  of  gravity  is  changed  by  external 
forces  only. 

If  momenta  are  resolved  and  compounded  like  forces,  then  from 
Fig.  28  the  momentum  of  the  centre  of  gravity,  in  which  we  may  consider 
both  masses  united,  equals  the  resultant  of  the  momenta  of  the 
separate  masses  ml  and  m2.  The  momentum  ml.oa  may  be  resolved 
into  m1oc  +  m1c«,  the  momentum  m2ob  into  m2oc  +  m2cb.  Now,  m^ca 
and  m.ycb  are  equal  but  opposite  in  direction.  Hence  we  have  for 
the  resultant  momentum  m^c  +  m.2oc  =  (ml  +  m2)oc. 

The  velocity  of  the  centre  of  gravity  remains  unchanged  if  the 
bodies  m^  and  m2  act  on  each  other  according  to  the  law  of  action 
and  reaction.  In  this  case  both  bodies  receive  momenta  which  are 
equal  but  oppositely  directed,  and  which  annul  each  other.  This 
result  may  be  derived  analytically  in  the  following  way.  If  xl  and  x2 
are  the  coordinates  of  the  particles  m^  and  m.2,  the  line  joining  which  is 
taken  for  the  a-axis,  and  if  the  ,r-components  of  the  forces  with  which 
the  masses  act  on  each  other  are  X^  and  X>,  the  equations  of  motion 
are  (a)  m1xl  =  Xl  and  m^cz  =  X2.  Adding  these  equations,  we  have 

(b)  d*(mlxl+m<p2)/dt-  =  Xl  +  X2.      Since   Xl   and   X2  arise   from   the 
mutual  action  of  the  masses  on  each  other,  they  are  equal  but  opposite 
in  direction,  and  hence  Xl  +  X2  =  Q.     Setting 

(c)  mfa  +  mjK2  =  (TWj  +  ro2)£, 

we  have  £  =  0,  £  =  Const.  Hence  the  point  determined  by  the 
^-coordinate  moves  with  the  constant  velocity  £.  The  .r-coordinate 
of  the  centre  of  gravity  is  £,  because  m1(.r1  -  £)  =  w2(£  -  x2).  Differ- 
entiating equation  (c)  with  respect  to  t,  we  have 


That  is,  the  momentum   of  the  centre  of  gravity  equals  the  sum  of  the 
momenta  of  the  separate  masses. 

By  Newton's  law  of  universal  attraction  two  masses  ml  and  m.2 
act  on  each  other  with  a  force  -fm^m^f2,  where  r  is  the  distance 
between  the  two  masses.  Their  motion  may  be  determined  in  the 
following  way.  The  velocity  of  the  centre  of  gravity  and  the 
velocities  of  the  masses  relative  to  the  centre  of  gravity  are  deter- 
mined from  the  velocities  of  the  masses.  These  act  on  each  other 
with  forces  directed  toward  the  centre  of  gravity,  and  we  can 
therefore  consider  this  as  the  attracting  point.  If  1\  is  the  distance 
of  the  mass  m^  from  the  centre  of  gravity,  then  m1rl  =  m2(r-r1),  and 


SECT,  xvii.]  THE   CENTRE  OF  GRAVITY.  53 

therefore  m.2r  =  (m1  +  m.2)rl.  By  substitution  of  this  value  of  r  we 
obtain  for  the  force  the  expression  fmlm.2Al(ml  +  w2)2ri2-  The  mass 
%  therefore  moves  round  the  centre  of  gravity  as  if  the  force  acting 
on  it  were  due  to  a  mass  Ml  =  m23j(ml  +  m.->)2  situated  at  that  point. 


SECTION  XVIII.    A  MATERIAL  SYSTEM. 

We  will  now  consider  a  system  of  separate  masses  in  vacuo,  which 
act  on  each  other  with  forces  which  are  functions  of  the  distances 
of  the  masses  from  each  other,  and  obey  the  law  of  action  and 
reaction.  The  forces  which  act  in  such  a  way  within  the  system 
are  called  internal  forces.  External  forces,  proceeding  from  bodies 
which  do  not  belong  to  the  system,  may  also  act  on  it.  The  masses 
are  designated  by  %,  m2,  m3,  etc.,  and  the  positions  of  the  masses  are 
determined  by  the  coordinates  x,  y,  z,  with  appropriate  indices.  We 
may  determine  the  position  of  the  system  by  supposing  each  mass 
to  be  made  up  of  different  numbers  of  units  of  mass ;  the  mean 
values  £,  y,  £  of  the  x-,  y-,  ^-coordinates  will  then  be 

(a)  £  =  (mft  +  m^  +  myx3  +  ...)/(ml  +  m2+ ...).  etc. 

£,  -i],  £  are  the  coordinates  of  the  centre  of  gravity  of  the  system  of 
masses.  If  the  equation  (a)  is  differentiated  with  respect  to  the 
time  t,  it  appears  that  the  velocity  of  the  centre  of  gravity  depends 
on  the  velocities  of  the  particles.  That  is, 

(b)  £  =  (m^  +  m^c2  +  m^bs  +  . . . )/(•/»,  +  m2  +  m3+ ... ),  etc. 

The  internal  forces  cannot  change  the  motion 
of  the  centre  of  gravity,  since  by  the  law  of 
action  and  reaction  two  masses  impart  to  each  a, 
other  equal  and  opposite  momenta,  the  sum 
of  whose  projections  on  any  axis  is  equal  to 
zero. 

This  result  may  be  represented  geometrically 
in  the  following  way.  From  any  point  0  (Fig. 
29)  draw  the  lines  Oa,  Ob,  Oc,  etc.,  which  repre- 
sent the  velocities  of  the  masses  mv  m2,  m3,  etc. 
Then  if  the  masses  mv  mv  ms,  etc.,  are  placed 
at  the  points  a,  b,  c,  etc.,  respectively,  and  if 
p  is  the  centre  of  gravity  of  the  masses,  Op  is  the  velocity  of  the 
centre  of  gravity. 


54  GENERAL  THEORY  OF  MOTION.  [CHAP.  i. 

In  order  to  determine  the  motion  of  the  separate  particles  we 
must  know  the  separate  forces  which  act  on  them.  If  the  com- 
ponents of  the  external  forces  acting  on  the  mass  ma  are  designated 
by  XM  Ya,  Za,  if  Fab  is  the  force  with  which  ma  is  attracted  by 
mM  and  if  r,^  is  the  distance  between  rna  and  mb,  the  ,r-component 
of  the  forces  acting  on  ma  is 

Xa  +  Fab.(xa-xb)/rab  +  Fac(x<l-xr)lrae+.... 

We  obtain  similar  expressions  for  mb,  mc,  etc.     Hence  we  have 
(c)  m^a  =  Xa  +  F^ .  (xa  -  xb)/rab  +  FM .  (xa  -  xe)/rM  +  . . . ,  etc. 

Since  by  the  law  of  action  and  reaction  Fn6  =  Fba,  Fac  =  Fca,  we  have 
f 
\ 

If  we  now  introduce  the  coordinates  of  the   centre  of  gravity,  we 
have,  using  equation  (a), 


This  equation  contains  the  law  of  the  motion  of  the  centre  of  gravity, 
which  may  be  thus  stated :  The  centre  of  gravity  of  a  system  of  masses 
moves  like  a  material  point  in  which  all  the  masses  of  the  system  are 
united,  and  at  which  all  the  forces  are  applied. 

The    momentum    of    the    whole    system    is    compounded    of    the 
momenta  of  the  separate  masses.      From  a  point  0  (Fig.  30)  draw 
OA  =  mava  parallel  to  the  direction  of  the  velocity 
va.     In  the  same  way  draw  AB  =  mbrb,  BC=mevc, 
etc.      Taking   account   of  all  particles  we  reach 
2?/  /         a  point  D.      The   line   OD  then   represents  the 

momentum  of  the  system.     The  sums  of  the  pro- 
jections of  the  momenta  on  the  coordinate  axes  are 


By  equation  (b)  these  sums  equal  the  components 
of  the  momentum  of  the  centre  of  gravity.  In 
the  time-element  dt  the  motions  of  the  separate 
masses  are  changed  by  the  forces  which  act  on 
them;  nevertheless,  the  internal  forces  do  not 
change  the  momentum,  since  the  resultant  of  the 
momenta  which  these  forces  occasion  is  zero,  by  the  law  of  action 
and  reaction.  On  the  other  hand,  changes  of  momentum  are  occa- 
sioned by  the  action  of  the  external  forces.  A  force  K  produces 
the  momentum  K  .dt  in  the  time  dt.  If  all  the  momenta  which 
the  external  forces  produce  are  determined  in  this  way,  and  com- 


SECT,  xviii.]  A  MATERIAL  SYSTEM.  55 

bined  with  those  originally  given,  the  actual  momentum  is  obtained. 
This  also  appears  from  equation  (d),  which  may  be  thus  written : 

(f )  d(maxa  +  mbxb  +  m,xe  +...)  =  (Xa  +  Xb  +  Xc  +.  ..)dt. 

Sir  William  Rowan  Hamilton  introduced  the  word  vector  to  repre- 
sent magnitudes  which  have  direction,  and  which  may  be  compounded 
like  motions,  velocities,  forces,  etc.  The  sum  of  vectors  is  called 
their  resultant.  If  we  consider  momentum  and  force  as  vectors,  the 
increase  of  the  momentum  which  a  system  receives  in  the  time  dt 
equals  the  product  of  the  resultant  of  the  external  forces  and  the 
time  dt.  Since  the  momentum  of  the  system  equals  the  momentum 
of  the  centre  of  gravity,  the  law  just  stated  holds  also  for  this  latter. 


SECTION  XIX.    MOMENT  OF  MOMENTUM. 

If  the  mass  m  at  the  point  A  (Fig.  31)  moves  in  the  direction  AB 
with  the  velocity  v,  its  momentum  is  mi:  If  0  is  an  arbitrary 
fixed  point,  and  00 =p  a  line  perpendicular  to  ^ 
AB,  the  product  mvp  is  called  the  moment  of 
momentum  with  respect  to  0.  The  value  of  the 
moment  depends  on  the  position  of  the  point  0. 
If  we  erect  a  perpendicular  on  the  plane  deter- 
mined by  0  and  AB,  and  lay  off  on  it  from  0 
a  length  proportional  to  mvp,  the  vector  deter- 
mined in  this  way  is  called  the  moment  of  ! 

momentum.      This  vector  is  to  be  so  constructed 

that  it  points   in   the   direction   of  the   thumb,  ; 

if  the   right   hand   points   in   the   direction    OC, 

and   the   palm   is   turned   toward   the   direction   of  the  force. 

In  the  same  way  the  vectors  corresponding  to  all  parts  of  the 
system  can  be  determined,  and  compounded  by  the  method  given 
in  Fig  30.  If  neither  external  nor  internal  forces  act  on  the  parts 
of  the  system,  the  moment  of  momentum  of  the  entire  system  is 
invariable,  since  the  separate  moments  remain  invariable.  The 
moment  of  momentum  of  the  system  is  also  not  changed  by  the 
action  of  internal  forces.  If,  for  example,  A  and  B  (Fig.  32)  are 
the  points  occupied  by  two  masses  m^  and  m^  which  repel  each 
other  with  the  force  K,  then  A  receives  in  the  time  dt  the  momentum 
K.dt  in  the  direction  A  A',  and  B  receives  the  same  momentum  in 
the  opposite  direction.  The  moments  of  momentum  of  A  and  B 


56 


GENERAL  THEORY  OF  MOTION. 


[CHAP.  i. 


annul  each  other.  On  the  other  hand,  the  moment  of  momentum 
of  the  system  will  in  general  be  changed  by  external  forces,  but  it 
will  remain  constant  in  case  the  directions  of  all  the  external  forces 
always  pass  through  the  fixed  point  0. 

Hence,  if  the  moments  of  momentum  and  the  moments  of  the 
external  forces  are  considered  as  vectors,  the  increment  of  the  moment 
of  momentum  of  the  system  in  the  time  dt  equals  the  resultant  of 
the  moments  of  the  external  forces  multiplied  by  dt. 


X 


FIG.  32. 


FIG.  33. 


This  may  be  represented  analytically  in  the  following  way.  Let 
AB  —  x  and  AC=y  be  the  velocity  components  of  the  particle  M 
situated  at  A  (Fig.  33).  The  distance  of  the  moving  mass  from  the 
x-axis  is  y,  and  from  the  ?/-axis  is  x.  Hence  the  moments  of  momentum 
with  respect  to  the  £-axis  are  mxy  and  myx.  These  being  oppositely 
directed,  their  difference  mxy  -myx  is  the  moment  of  momentum  of 
m  with  respect  to  the  z-axis.  This  moment  receives  in  the  time  dt 
the  increment  md(xy  -  yx)  =  m(xy  -  yx)dt.  Hence  we  have 

(a)     2m(a#  -  yx)  =  2(zF-  yX),  or  (b)  d2m(xy  -  yd)  =  dfS,(xY-  yX), 


that  is,  the  increment  which  the  moment  of  momentum  about  any  axis 
receives  in  the  time  dt  is  equal  to  the  pi'oduct  of  the  sum  of  the  moments 
of  the  external  forces  about  the  same  axis  and  the  time-element  dt. 


SECTION  XX.    THE  ENERGY  OF  A  SYSTEM  OF  MASSES. 

If  a  particle  m  moves  with  a  velocity  v  =  ds/dt,  its  kinetic  energy 
[V.]  is  £/nfl2  =  £m(<fe/<ft)2  =  £ws2.  Since  ds2  =  dx*  +  df  +  dz2,  this  may 
be  written  %mvz  =  \m(xl  +  y2  +  22).  The  kinetic  energy  of  the  system 


SECT,  xx.]  ENERGY  OF  SYSTEM  OF  MASSES.  57 

is  determined  from  the  velocities  of  the  separate  particles  of  the 
system.  It  is  expressed  by  T  =  % .  *2m(x*  +  yz  +  z2).  If  xt  y,  z  are  the 
coordinates  of  a  particle,  and  £,  rj,  £  the  coordinates  of  the  centre  of 
gravity,  the  coordinates  of  the  particle  with  respect  to  the  centre 
of  gravity  are  x  —  £  =  x',  y  —  ri  =  y',  z  —  £  =  z1.  Using  these  new 
coordinates,  we  obtain 

Swfce2  =  2m(£  +  x)'2  =  £22m  +  Zmx'2  +  2£2m£',  etc. 

From  XVIII.  (a)  we  may  set  ^mx'  =  0,  if  the  centre  of  gravity  is 
chosen  as  the  origin  of  coordinates.  Then 

T=  \  .  (¥+  if  +  C2) .  2m  + 1 .  ^m(x"1  +  y"2  +  z2). 

The  kinetic  energy  of  the  system  is  equal  to  the  sum  of  the  kinetic  energy 
of  the  masses  due  to  the  motion  of  the  centre  of  gravity,  and  the  kinetic 
energy  of  the  masses  due  to  their  motion  relative  to  the  centre  of  gravity. 

The  increment  of  the  kinetic  energy  of  the  system  in  the  time- 
element  dt  equals  the  work  done  by  the  forces  during  that  time. 
This  is  divisible  into  two  parts,  that  of  the  external  and  that  of 
the  internal  forces.  If  we  designate  the  components  of  the  motion 
of  a  particle  parallel  to  the  axes  by  dx,  dy,  dz,  the  work  done  by 
the  external  forces  is  ^(Xdx+Ydy  +  Zdz). 

If  /•  is  the  distance  between  two  particles  and  the  repulsive  force 
F  acts  between  them,  the  work  done  by  the  internal  forces  is  *2.Fdr. 
Hence  we  have  dT=^(Xdx+Ydy  +  Zdz)  +  ^Fdr.  If  the  force  with 
which  the  masses  act  on  each  other  is  a  function  of  the  distance  r 
only,  we  can  set  Fdr  =  d^,  where  ^  is  a  function  of  r  only.  Now 
setting  ^d$  =  dU,  we  have  finally 

(d )  dT  =  2  (Xdx  +  Ydy  +  Zdz)  +  d  U. 

The  function  U  depends  only  on  the  distance  between  the  particles 
or  on  the  configuration  of  the  system.  U  is  the  potential  of  the  system 
on  itself  or  the  internal  potential  energy  of  the  system.  Further,  U  is 
the  work  which  would  be  done  by  the  internal  forces  if  the  particles 
were  to  move  from  their  positions  at  any  instant  into  other  positions 
in  which  their  mutual  actions  are  zero. 

If,  for  example,  the  given  masses  act  on  each  other  according  to 
Newton's  law,  we  have  F=  -fmlm2/r'2,  and  therefore 

Fdr=  -fmlm2dr/r'2  =  +fd(mlm2/r). 

If  several  masses  mv  m.2,  m3,  ...  are  present,  whose  distances  from 
each  other  are  r12,  r13,  r23,  ...  respectively,  we  have 

(e)  d  U =.fd(mlm2lrl2  +  w»1»»8/rI8  +  m2m3/r23  +...). 


58  GENERAL  THEORY  OF  MOTION.  [CHAP.  i.. 

If  the  system  passes  from  one  configuration  to  another,  the  work 
done  by  the  internal  forces  is  determined  only  by  the  initial  and 
final  positions  of  the  particles,  and  does  not  depend  on  the  paths 
traversed  by  them.  If  no  external  forces  act  on  the  system,  we  have 

(f)  clT=  dU  or  T  -  T0  =  U  -  U0. 

Now,  from  the  discussion  in  VII.,  we  may  set  E/"0  =  0,  and  so  obtain 
T=U+T0,  that  is,  the  kinetic  energy  of  the  system  equals  the 
original  kinetic  energy  T0  increased  by  the  work  U  done  by  the 
forces.  In  case  of  a  change  in  the  relative  positions  of  the  particles, 
supposing  no  external  forces  to  act,  a  transformation  of  the  one 
form  of  energy  into  the  other  occurs  without  causing  a  change  in 
the  total  energy  of  the  system,  that  is,  the  sum  of  the  kinetic  and 
potential  energies  of  such  a  system  is  constant. 


SECTION  XXI.    CONDITIONS  OF  EQUILIBRIUM.    RIGID  BODIES. 

We  will  now  consider  the  conditions  of  equilibrium  of  a  system. 
If  the  positions  of  the  separate  masses  at  a  definite  instant,  and 
also  the  internal  and  external  forces  are  given,  the  system  is  in 
equilibrium,  when  the  resultant  of  all  the  forces  acting  on  each 
particle  is  zero.  If  the  internal  forces  are  in  equilibrium,  no  change 
occurs  in  the  motion  of  the  system,  so  long  as  no  external  forces 
act  on  it.  External  forces  will '  as  a  rule  set  the  system  in  motion ; 
but  it  is  also  possible  that  they  will  not  change  the  equilibrium  of 
the  system  as  a  whole  even  if  its  separate  parts  are  set  in  motion. 
If  the  resultant  of  the  external  forces  is  zero,  the  motion  of  the 
centre  of  gravity  remains  unchanged  [XVIII. ] ;  so  that,  for  example, 
if  the  centre  of  gravity,  is  at  rest,  it  remains  at  rest.  But  even 
when  this  is  the  case,  the  external  forces  may  set  the  separate 
masses  in  motion ;  the  relative  positions  of  the  particles  may  be 
changed,  and  changes  of  form  or  rotations  may  occur.  The  con- 
ditions for  such  changes  are  developed  in  the  theory  of  elasticity 
and  in  hydrodynamics.  At  present  we  will  consider  only  the 
behaviour  of  rigid  bodies. 

The  particles  of  such  bodies  are  so  conditioned  that  the  distances 
between  them  are  constant  or  nearly  so.  If  the  positions  of  three 
particles  of  the  body  are  given,  the  positions  of  all  the  other  particles 
are  also  given,  and  the  position  of  the  body  is  determined.  If  the 


SECT.  XXI.] 


CONDITIONS  OF  EQUILIBRIUM. 


body  (Fig.  34)  is  moved  from  its  position  so  that  the  points  A,  B,  C 

are  brought  to  the  points  A',  B',  C'  respectively,  it  can  be  brought 

back  to  its  original  position  by  a  series  of  simple  operations.     The 

body  may  first  be  displaced  parallel  with  itself  through  the  distance 

AA',  so  that  the  point  A'  coincides  with  A,  and  the  points  B'  and 

C'  are  brought  to  b  and  c.     C'c,  B'b,  and  A' A  are  equal  and  parallel. 

The  body  may  then  be  turned  about  an 

axis  passing  through  A,  perpendicular 

to  the   plane  determined   by  BA  and 

bA,  through  the  angle  BA  £>,  so  that  b 

coincides  with  B  and  c  is  brought  to 

c'.      By   a   second   rotation    about  the 

axis  AB,  c'  may  be  made  to  coincide 

with  C.      The   motion  of  the  body  is 

thus  reduced  to  a  translation  and  two 

rotations.      A    rigid    body,    therefore, 

cannot  be  moved,  if  it  can  neither  be 

displaced  nor  rotated. 

In  order  that  a  body  acted  on  by 
external  forces  shall  be  in  equilibrium, 
its  centre  of  gravity  must  remain  at 
rest.  The  necessary  condition  for  this 

is  that  the  resultant  of  the  external  forces  is  zero.  It  must  also 
have  no  rotation  about  any  axis.  If  such  rotations  exist,  its  particles 
receive  a  certain  momentum,  which  has  a  moment  with  respect  to 
the  axis.  Since  each  of  the  elements  in  this  moment  is  positive, 
because  the  parts  of  the  body  all  move  in  the  same  sense,  and, 
therefore,  the  momenta  of  the  separate  particles  have  the  same 
sign,  the  sum  of  the  momenta  can  vanish  only  when  each  one  of 
them  is  separately  zero.  Now  the  sum  of  the  moments  of  momentum 
is  [XIX.]  equal  to  the  product  of  the  moment  of  force  and  the  time 
during  which  the  force  acts.  Hence  it  is  required  for  equilibrium 
that  the  forces  which  act  on  the  body  have  no  moment  with  respect 
to  the  axis.  This  must  hold  for  each  axis  about  which  the  body 
can  turn.  Now,  if  the  moment  of  the  forces  equals  zero,  the  sum 
of  the  moments  of  momentum  equals  zero,  therefore  the  moment  of 
momentum  of  each  particle  equals  zero;  that  is,  each  particle  is  in 
equilibrium.  Furthermore,  since  moments  can  be  compounded  like 
forces,  equilibrium  will  exist  if  the  moments  with  respect  to  three 
arbitrary  axes  are  zero. 


FIG.  34. 


60  GENERAL  THEORY   OF  MOTION.  [CHAP.  i. 

SECTION  XXII.    ROTATION  OF  A  RIGID  BODY.    THE  PENDULUM. 

Let  a  solid  body  revolve  around  an  invariable  axis,  which  is  chosen 
as  the  0-axis  of  a  system  of  rectangular  coordinates.  Let  the  angular 
velocity  of  the  body  be  w.  If  r  represents  the  distance  of  any  particle 
m  from  the  2-axis,  the  velocity  of  this  particle  is  ro>  and  its  kinetic 
energy  ^mr-wr.  Since  tu  has  the  same  value  for  all  particles,  the 
kinetic  energy  T  equals  T=W22mr2.  The  factor  2mr2  is  called  the 
moment  of  inertia  J  of  the  body  with  respect  to  the  3-axis ;  the  moment 
of  inertia  is  equal  to  the  sum  of  the  products  of  the  particles  into 
the  squares  of  their  respective  distances  from  the  0-axis.  Hence  we 
have  T=\(»-J,  that  is,  the  kinetic  energy  of  a  rotating  body  is  equal  to 
its  moment  of  inertia  multiplied  by  half  tJie  square  of  its  angular  velocity. 
A  length  K  may  always  be  found,  such  that  2»w2  =  K'2-m.  This 
length  is  called  the  radius  of  gyration  of  the  body.  It  is  the  distance 
from  the  axis  at  which  a  mass  equal  to  the  mass  of  the  body  would 
have  the  same  moment  of  inertia  with  respect  to  the  axis  as  that 
of  the  body. 

If  the  only  external  forces  which  act  on  the  body  pass  through 
the  axis,  the  work  done  by  them  is  zero,  since  the  axis  does  not 
move.  Since  the  internal  forces  also  do  no  Avork,  the  kinetic  energy 
and  therefore  also  the  angular  velocity  o>  must  remain  constant. 

Since  [XVIII.]  the  centre  of  gravity  moves  as  if  the  resultant  of  all 
the  forces  acted  on  the  mass  of  the  body  concentrated  at  the  centre 
of  gravity,  this  resultant  R  can  be  determined.  If  we  represent 
the  distance  OP  (Fig.  35)  of  the  centre  of  gravity  from  the  r-axis 
by  a,  we  have  [IV.  (b)]  E  =  ^ma-(a-/a  =  ^mao)-.  E  is  the  resultant 
of  the  forces  with  which  the  body  acts 
on  the  axis  of  rotation.  In  general,  the 
forces  applied  to  the  body  so  act  that 
they  have  no  resultant;  they  tend  only 
to  produce  rotation  about  the  axis.  In 
order  to  determine  them,  the  theorem  in 
XIX.  concerning  moments  of  momentum 
must  be  used. 

If  external  forces  act  on  the  body,  its 
angular  velocity  changes.  The  amount 
of  this  change  is  determined  from  XIX. 
The  momentum  of  a  particle  m  is  represented  by  mro>,  and  its 
moment  of  momentum  by  mrtar.  Hence  the  moment  for  all  particles 
of  the  body  is  <o2wr2  =  wJ.  If  the  moment  of  the  forces  with  respect 


SKCT.  xxii.]  ROTATION  OF  A  RIGID  BODY.  .  61 

to  the  i-axis  is  represented  by  M,  we  have,  from  XIX.,  d(<oj)  =  Melt 
or  (c)  Jdwfdt  =  M. 

If,  for  example,  the  moment  is  constant,  the  angular  velocity 
increases  in  direct  ratio  with  the  time. 

If  the  moment  is  due  to  gravity,  the  body  under  certain  conditions 
performs  oscillations.  We  suppose  the  a-axis  taken  parallel  to  the 
direction  of  gravity  and  represent  the  force  of  gravity  by  g.  The 
position  of  the  centre  of  gravity  P  (Fig.  35)  is  determined  by  the 
angle  POX=Q  and  the  angular  velocity  o>  by  dS  =  <adt.  The  moment  of 
force  with  respect  to  the  s-axis  is  -  (m^  +  rn.^  +...)#=  -?/.  g?m, 
where  -/;  is  the  y-coordinate  of  the  centre  of  gravity.  2m  denotes  the 
sum  of  all  the  masses.  Since  t]  =  a  sin  0,  we  have  from  (c) 

/.  0  =  -  a  sin  0  .  g~2m. 

If  6  is  very  small,  so  that  we  may  set  sin  9  =  9,  this  becomes 
(d)  9=  -aeg'Sm/J. 

Comparing  this  equation  with  that  given  in  VIII.  (g),  we  find  that 
they  are  identical  when  we  set  l/7  =  «2m//.  The  period  of  oscillation 
of  the  physical  pendulum  is  therefore  (e)  t  =  vxjlfg  =  irJjjgaSm.  Since 
J=2m(x-  +  y2  +  z2),  we  get  by  transferring  the  origin  of  the  system 
of  coordinates  to  the  centre  of  gravity  (£,  77,  £),  if  #',  y',  z'  are  the 
coordinates  with  respect  to  the  new  origin, 

(f  )       J  =  2m{(z'  +  £)2  +  (y'  +  ,)2  +  (z'  +  O2}  =  2ma*  +  ?m(yf*  +  y">  +  ^\ 

since  the  terms  £2w.?',  rj^my'  and  £2mtf  vanish. 

Now,  if  we  set  /=  a22w  +  &22m,  where  k  is  the  radius  of  gyration, 


we  obtain  from  (e)  (g)  t  =  irj(a2  +  k*)/ga.  We  call  l  =  —  the 
reduced  length  of  the  pendulum  or  the  length  of  the  equivalent  simple 
pendulum.  The  point  S  which  is  at  the  extremity  of  the  line  OS  =  l 
(Fig.  35),  drawn  through  0  and  P,  is  called  the  centre  of  oscillation. 
If  an  axis  is  passed  through  S  parallel  to  the  £-axis,  and  the  body 
oscillates  about  it,  the  reduced  length  of  the  pendulum  I'  is 


Since,  however,  l-a  =  k2/a,  we  have  I'  =  (a2  +  k2)/a  =  I.  The  reduced 
length  of  the  pendulum  and  therefore  the  time  of  oscillation  are 
the  same  for  this  new  axis  as  for  the  former  one. 


CHAPTER   II. 

THE  THEOEY  OF  ELASTICITY. 

SECTION  XXIII.     INTERNAL  FORCES. 

IF  all  parts  of  a  body  are  in  equilibrium  and  if  no  tensions  or  pressures 
act  on  them,  yet  internal  forces  must  be  present  acting  between  the 
separate  parts  of  the  body.  Every  action  produces  changes  of  form 
in  the  body,  and  thus  develops  forces  in  its  interior,  which  act  in  a 
sense  opposite  to  the  external  forces.  These  internal  forces  con- 
dition the  nature  of  the  body,  determining,  for  example,  the  difference 
between  solids  and  fluids.  No  sharp  distinction  can  be  drawn,  how- 
ever, between  these  two  classes  of  bodies.  Viscous  fluids  and  jelly- 
like  solids  are  bodies  which  seem  to  be  transition  forms  between 
true  solids  and  fluids. 

If  a  pressure  acts  on  the  surface  of  a  fluid,  it  must  be  equally 
great  on  equal  areas  of  the  surface  at  all  points,  and  it  must  be 
perpendicular  to  the  surface,  if  the  fluid  is  to  be  in  equilibrium. 
This  pressure  is  exerted  throughout  the  whole  mass ;  all  equal 
surface-elements  at  a  point  are  subjected  to  equal  pressures,  which 
are  always  perpendicular  to  the  surface-elements.  We  call  such  a 
pressure  hydrostatic  pressure.  A  similar  pressure  may  also  be  present 
in  solids.  If  a  solid,  a  piece  of  glass,  for  example,  which  fills  the 
volume  enclosed  by  its  external  surface,  is  immersed  in  a  fluid  on 
which  a  pressure  is  exerted,  the  same  pressure  exists  at  every  point 
in  the  surface  of  the  glass  as  in  the  fluid.  The  pressure  is  everywhere 
the  same,  and  perpendicular  to  the  surface-elements.  We  may  there- 
fore speak  of  hydrostatic  pressure  in  solids  also. 

Yet,  in  general,  internal  forces  in  solids  are  very  different  from 
those  in  fluids.  Let  a  cylindrical  rod  be  fastened  at  one  end,  and 
let  the  force  V  be  applied  at  the  other  end  so  as  to  lengthen  the 

62 


CH.  II.  SECT.  XXIII.] 


INTERNAL   FORCES. 


63 


rod.  In  a  cross  section  perpendicular  to  the  axis  of  the  cylinder 
the  internal  forces  are  everywhere  equal.  Let  the  area  of  the  cross 
section  be  A  (Fig.  36),  then  the  force  V\A  acts  on  unit  of  area  in  A. 
This  quotient  represents  the  stress  S  in  the  rod.  If 
another  plane  cross  section  B  is  taken  in  the  rod, 
which  makes  the  angle  <£  with  A,  the  force  &  acts 
on  each  unit  of  area  of  B,  so  that 

&  .  AJ  ^  *b .  ^4  =  o  .  Jj  cos  <PJ 

and  hence  (a)  S'  =  Scos<J>.  The  stress  S'  is  no  longer 
perpendicular  to  the  surface  B  on  which  it  acts ;  its 
magnitude  decreases  with  cos  <f>  and  vanishes  for 
<£  =  |TT.  A  surface-element  within  the  cylinder  and 
.  parallel  to  its  axis  is  therefore  subjected  neither  to 

pressure  nor  to  tension ;  this  conclusion  holds  for 
an  element  of  the  surface  of  the  cylinder.  We  may 
resolve  S'  into  two  components,  one  of  which,  T, 

is  tangent,  and  the  other,  N,  normal  to  B,  and  have 

(b)  N=Scos-<f),      T  =  S  cos  <f>  sin  <J>. 

If  internal  forces  of  this  type  exist  within  a  body,  we  call  the  stresses 

axial.     In  the  direction  of  the  axis  the  stress  is  S;  a  unit  of  surface 

whose  normal   makes  the  angle  <£  with  the  axis  is  acted  on  by  a 

force  S  cos  <£  in  the  direction  of  the  axis. 

We  will  consider  a  rectangular  parallelepiped  (Fig.  37),  of  which 

the  lines  OA,  OB,  and  OC  are  adjacent  edges.     The  stresses  which 

act   on   each   unit  of  area  of  the  faces 

which  are  perpendicular  to  OA,  OB,  and 

OC  are  Sa,  Sb,  Sc  respectively.      If  the 

normal   to   an   arbitrarily   situated   unit 

of  surface  /  makes  the  angles  a,  (3,  y, 

with  the  edges  OA,  OB,  OC  respectively, 

the   force   acting   on  /  is   the  resultant 

of   the   forces  Sacos  a,   Sbcos  /3,   Sccos  y, 

which   are   parallel   to  OA,  OB,  OC  re-    / 

spectively.      If  the    stresses    Sa,   Sb,   Sc 

have  the  same  value  S,  this  resultant  is  &/  cos2a  +  cosL'/3  +  cos'-'y  =  S. 

Hence   three  equal   stresses  which  are  perpendicular  to   each  other 

cause  a  hydrostatic  stress,  since  their  resultant  has  the  same  value 

whatever  may  be  the  position  of  the  surface.     Since  the  components 

of  this  stress  are  S  cos  a,  S  cos  ft,  and  S  cos  y,  it  is  perpendicular  to 

the  unit  area  /. 


y 

B 

,/ 

' 

A     x 

o 

/ 

/ 

J3' 

FIG.  37. 

64 


THE   THEOEY   OF  ELASTICITY. 


[CHAP  n. 


On  the  other  hand,  if  Sc  =  0  and  Sa  =  Sb  =  S,  that  is,  if  two  stresses 
act  at  right  angles  to  each  other,  while  the  stress  perpendicular  to 
them  both  is  zero,  the  components  in  the  directions  OA,  OB,  OC 
respectively  are  S  cos  a,  S  cos  /3,  0.  Hence  the  force  acting  on  /  is 


Sj  cos-a  +  cos2/?  =  &Jl  -  cos-y  =  S  sin  y, 

and  is  perpendicular  to  OC.  Such  a  state  of  stress  in  a  body  may 
be  called  equatorial.  The  plane  which  contains  OA  and  OB,  or  rather, 
every  plane  parallel  to  both  these  lines,  may  be  called  an  equatorial 
plane.  The  same  stress  S  acts  on  each  unit  area  perpendicular  to 
the  equatorial  plane.  If  the  normal  to  the  surface  /  makes  the  angle 
(f)  with  the  equatorial  plane,  the  stress  on  it  is  proportional  to  cos  <£. 


SECTION  XXIV.     COMPONENTS  OF  STRESS. 

Let  the  surface  F  (Fig.  38)  divide  a  body  into  two  parts,  A  and 
B.  If  the  portion  of  A  which  touches  the  element  dF  of  the  surface  F 
is  removed,  a  force  must  act  on  dF  to  keep  B  in  equilibrium.  This 
force  SdF  is  not,  as  a  rule,  perpendicular  to  the  element  dF.  The 

forces  acting  at  the  various  points 
of  F  are,  in  general,  different. 
If  the  force  tends  to  move  the 
element  dF  into  the  space  occu- 
pied by  B,  it  is  called  a  pressure 
on  the  surface  dF;  if  it  tends  to 
move  the  element  dF  into  the 
space  occupied  by  A,  it  is  called 
a  tension.  In  all  cases  we  call  the 
force  S  a  stress;  if  this  acts  as  a 
tension,  it  is  a  positive  stress,  if  as 
a  pressure,  it  is  a  negative  stress. 
If  the  part  of  B  which  touches 
dF  is  removed,  then  to  maintain 
equilibrium  in  A  a  force  SdF 
must  act  on  dF,  since  action  and 
reaction  are  equal  Hence  both  forces  which  act  on  an  element  of 
surface  within  a  body  are  equal,  but  oppositely  directed.  It  is 
characteristic  of  a  stress  that  it  may  be  looked  on  as  made  up  of 
two  equal  and  opposite  forces. 


FIG.  38. 


SECT,  xxiv.]  COMPONENTS  OF  STRESS.  65 

If  the  surface-element  dF  remains  in  its  original  place  in  the  body, 
but  is  turned  about  one  of  its  points,  a  particular  value  of  the  stress 
corresponds  to  every  one  of  its  positions;  for  special  positions  the 
stress  may  be  zero.  When  the  body  in  which  the  surface  is  drawn 
is  a  fluid,  the  stress  is  independent  of  the  position  of  the  surface. 
We  assume  in  the  body  a  system  of  rectangular  coordinates.  The 
stresses  in  the  surface-elements,  which  are  perpendicular  to  the 
directions  of  the  axes,  are  determined  by  their  components. 

Let  the  surface-element  dF  be  perpendicular  to  the  z-axis,  and  let 
dF  =  dy .  dz.  If  that  part  of  the  body  is  removed  which  lies  on 
the  positive  side  of  the  surface-element  dydz,  the  positive  side  being 
determined  by  the  positive  direction  of  the  z-axis,  then,  to  maintain 
equilibrium,  a  force  Sdydz  must  act  on  the  surface  dydz.  The 
force  AS'  is  resolved  into  the  components  Xa  Ya  Z^  which  are 
respectively  parallel  to  the  coordinate  axes.  The  index  indicates 
that  the  forces  act  on  an  element  which  is  perpendicular  to  the 
a--axis.  Xx  is  perpendicular  to  the  surface-element;  it  is  therefore 
called  the  normal  force ;  Yx  and  Zx  are  tangential  forces.  Xow,  let 
the  element  dF  remain  in  the  same  place,  but  be  turned  so  that 
it  is  perpendicular  to  the  y-axis.  We  may  then  set  dF=dzdx. 
As  before,  there  are  three  components  of  force  X^  Yy,  Zy  acting  on 
the  surface-element  dzdx,  of  which  Yy  is  the  normal  force,  Xy  and  Zy 
are  the  tangential  forces.  If  the  surface-element  dF  is  turned  so  as 
to  be  perpendicular  to  the  z-axis,  we  have  as  components  Xa  Ya  Z,, 
of  which  Zt  is  the  normal  force  and  Xz  and  Yz  are  the  tangential 
forces.  There  are  therefore,  in  all,  nine  components, 
Xa  Y^Z*;  X0Y,,Z,;  Xa  Yz,  Zz. 

By  these  components  the  stress  on  any  surface  is  determined.     Let 
OA,  OB,  00  (Fig.   39)  represent  line-elements,  parallel   respectively 
to  the  x-,  y-,  2-axes.     Let  a  plane  be  passed 
through  A,  B,  and  (7,  so  as  to   form   the 
tetrahedron  OABC.     Let  P,  Q,  and  R  be 
the  components  of  the  stress  in  the  direc- 
tions  of  the   coordinate   axes   at   a    point 
in  the  base  ABC  of  the  tetrahedron.     We 
now  form  the  equation  of  condition,  which 
must  hold  that  the  tetrahedron  shall  not    '  FIG. 

move  in  the  direction  of  the  z-axis.     The 

forces  which  tend  to  move  the  tetrahedron  in  that  direction  are 
P.  ABC  acting  on  its  base,  and  -XX.OBC,  -Xy.OAC,  -XZ.OAB 
acting  on  its  faces.  Hence  the  force  which  urges  the  tetrahedron  in 


66  THE   THEORY   OF   ELASTICITY.  [CHAP.  n. 

the  direction  of  the  o-axis  is  P.  ABC  -  Xx  .  OBC-Xy  .OAC-  X:  .  OAB. 
Designate  by  a,  /3,  y  the  angles  made  with  the  axes  by  the  normal 
to  the  surface  ABC  drawn  outward  from  the  tetrahedron;  then  the 
expression  for  the  force  in  the  direction  of  the  ,r-axis  becomes 

(P  -  Xx  cos  a  -  Xy  cos  (3  -  Xz  cos  y)  .  ABC. 

Now,  if  no  external  attractions  or  repulsions  act  on  any  part  of  the 
body,  the  conditions  of  equilibrium,  obtained  by  setting  this,  and 
the  two  similar  expressions  which  hold  for  the  other  axes,  equal  to 
zero,  are 

f  P  =  Xx  cos  a  +  Xy  cos  ft  +  Xz  cos  y, 
(a)  I  Q=Yxcosa+  Fycosp+Yscosy, 

[^=^cosa+  Zy  cos  /3  +  Zt  cos  y. 

If  other  forces  besides  the  stresses  act  on  the  parts  of  the  body, 
these  must  be  taken  into  account  in  equations  (a).  If  the  force  X 
acts  on  the  unit  of  mass  in  the  direction  of  the  a;-axis,  the  force 
acting  in  that  direction  on  the  tetrahedron  is  Xpdv,  if  dv  represents 
its  volume,  and  p  its  density.  The  condition  of  equilibrium  in  the 
direction  of  the  z-axis  then  becomes 

(P  -  Xx  cos  a  -  X,  cos  j3  -  X,  cos  y)  .  ABC  +  Xpdv  =  0. 

Now,  since  dv  =  %h  .  ABC,  where  h  is  the  height  of  the  tetrahedron, 
this  equation  is  equivalent  to 

P  -  Xx  cos  a  -  Xy  cos  (3  -  Xz  cos  y  +  $hpX  =  0. 

Since  the  height  h  of  the  tetrahedron  is  infinitely  small,  we  may 
neglect  the  term  containing  it,  and  again  obtain  the  first  of  equations 
(a),  which  hold  generally. 

In  order  to  exhibit  the  meaning  of  equations  (a),  we  will  consider 
the  following  case.  Suppose  a  tension  S  to  act  in  the  direction  of  the 
re-axis,  and  a  pressure  of  the  same  value  to  act  in  the  direction  of  the 
y-axis.  Then  XX  =  S,  Yy  =  -  S,  and  all  other  components  of  stress  are 
equal  to  zero.  Hence  P  =  S  cos  a,  Q  =  -Scosfl,  E  =  0. 

The  resultant  A  of  these  components  is  ^=$siny.      If  A,  //,,  v 

are    the    angles    between   A    and    the    axes,    we   have   cos  A 


B 

cos  M  =  —  -.  —  —  .  cos  v  =  0.      The  angle  e  between  A  and  the  normal 

smT  o  .      ,    •>/} 

to  the  surface-element  considered  is  determined  by  cos  e  =  —  '  —.  -  -. 

smy 

If   the    surface  -  element    is    parallel    to    the    ^axis,   y  =  a    A=S, 
COS  e  =  COS  2a,   €  =  ±  2a. 


SECT.  XXIV.] 


COMPONENTS   OF   STEESS. 


67 


If  a=       then  e  =  ^,  the  resultant  is  a  tangential  force.     Thus  the 
surface  of  a  prism  whose  axis  is  parallel  to  the  s-axis,  and  whose 


o 


1 


FIG.  39  a. 


make  angles  of  45°  with  the  xz-  and  y^-planes,  is  acted  on  only 
by  tangential  forces,  each  equal  to  S. 


SECTION  XXV.    RELATIONS  AMONG  THE  COMPONENTS  OF  STRESS. 

The  force  which  acts  on  the  volume-element  dxdydz  (Fig.  40)  is 
determined  from  the  components  of  stress.  Let  the  components 
acting  at  the  point  0  be  given,  and  let  the  force  which  acts  on  OA' 
in  the  direction  of  the  x-axis  be  equal 
to  -  Xjlydz.  By  development  by  Mac- 
laurin's  theorem  we  obtain  for  the  force 
acting  on  AO'  the  expression 
(Xx  +  'dXx/'dx.dx)dydz. 
The  resultant  of  these  two  forces  is 
'dXJ'dx  .  dxdydz.  The  forces  -  Xydxdz 
and  ( Xu  +  'dXJ'dy .  dy}dxdz,  whose  re- 
sultant is  'dXj'dy .  dxdydz,  act  on  the 
surfaces  OB'  and  O'B  respectively  in 
the  direction  of  the  z-axis.  The  resultant  of  the  forces  acting  in 
the  same  direction  on  the  surfaces  O'C  and  OC'  is  'dXJ'dz .  dxdydz. 
Hence  the  total  force  acting  on  the  parallelepiped  dxdydz  in  the 
direction  of  the  £-axis  is 


68  THE  THEORY   OF   ELASTICITY.  [CHAP.  n. 

If  (X),  (Y),  and  (Z)  represent  the  components   of  the   force  with 
which  the  stresses  act  on  unit  of  volume,  we  have 


(X)  =  'dXJ 

(Y)  =  VYJ-dx  +  3  YJ-dy  +  'd  YJ-dz, 

(Z)  =  c)ZJ3x  +  'dZyfdy  +  'dZJ'dz. 

If  the  body  is  acted  on  only  by  stresses,  equilibrium  will  exist 
if  the  three  components  (X),  (Y),  (Z)  are  each  equal  to  zero.  The 
equations  (b)  in  this  case  are  three  differential  equations  which  the 
components  of  stress  must  satisfy.  If  a  force  whose  components 
are  X,  Y,  Z  acts  on  each  unit  of  mass,  and  if  the  density  of  the 
body  is  p,  we  obtain  the  conditions  of  equilibrium, 

[  VXJ'dx  +  VXJdy  +  'dXJ^z  +  PX=  0, 

(c)  I  'dYJ'dx  +  'dYJ-dy  +  'dYJ-dz  +  pY^O, 

(  -d  Zx/*dx  +  3  ZJdy  +  ~d  ZJ-dz  +  PZ=0. 


Internal  forces  produce  both  translations  and  rotations  in  the 
The  tangential  components  tend  to  rotate  the  parallelepiped  00' 
about  the  £-axis.  The  tangential  force  Xy  acts  on  the  surface  OB' 
in  the  .negative  direction,  while  the  tangential  force  Xy  +  3  XJ'dy  .  dy 
acts  on  the  opposite  surface  O'B  in  the  positive  direction.  These 
two  forces  form  a  couple  acting  on  the  parallelepiped  with  a  moment 
Xy  .  dxdz  .  dy,  if  terms  of  an  order  higher  than  the  third  are  neglected. 
This  moment  tends  to  turn  the  parallelepiped  about  the  2-axis  in 
the  negative  direction.  The  tangential  forces  acting  on  the  surfaces 
OA'  and  O'A  have  the  moment  Yz  .  dydz  .  dx,  which  tends  to  turn 
the  parallelepiped  in  the  positive  direction.  The  total  moment  which 
tends  to  rotate  the  parallelepiped  about  the  z-axis  is  (Yx-  Xy)dxdydz. 
If  the  body  is  in  equilibrium  under  the  action  of  the  stresses  con- 
sidered, this  moment  must  be  zero,  that  is,  (d)  Yx  =  Xy,  and  similarly 
Zy  =  Ya  X,=  Z^  The  last  two  equations  are  derived  in  the  same 
way  as  the  first.  If  attractive  forces,  such  as  gravity,  or  in  general, 
if  any  forces  acting  at  a  distance  act  on  the  body,  equations  (d) 
will  still  be  applicable.  The  point  of  application  of  such  forces,  in 
infinitely  small  bodies,  coincides  with  the  centre  of  gravity  ;  such 
forces,  therefore,  cannot  produce  rotations,  and,  therefore,  cannot 
make  equilibrium  with  the  forces  which  tend  to  rotate  the  body. 

It  appears  from  equations  (d)  that  six  quantities  are  sufficient  to 
determine  the  stress  at  a  point  in  a  body,  namely,  Xx,  I7,,,  Zz  ; 
Zy=Ya  X2  =  ZX,  Yx  =  Xy.  The  first  three  are  normal  forces,  the  other 
three  tangential  farces.  It  is  possible  to  express  these  forces  by  a 


SECT,  xxv.]  COMPONENTS   OF   STEESS.  69 

simpler  notation,  but  we  will  retain  the  above,  which  has  the  advantage 
that  it  exhibits  more  clearly  than  any  other  the  true  significance  of 
the  quantities  involved.  It  must  be  borne  in  mind  that  the  value 
of  a  component  of  stress  remains  unchanged  if  the  direction  of  the 
force  and  the  direction  of  the  normal  to  the  surface-element,  on 
which  the  stress  acts,  are  interchanged. 


SECTION  XXVI.     THE  PRINCIPAL  STRESSES. 

In  order  to  obtain  a  better  understanding  of  the  nature  of  internal 
forces,  we  will  examine  if  it  is  possible  to  pass  a  surface  through  a 
given  point  in  a  body  in  such  a  position  that  no  tangential  force 
acts  on  it.  We  may  anticipate  our  conclusion  by  the  statement  that 
three  such  surfaces  may  be  drawn  through  any  point  and  that  they 
are  perpendicular  to  each  other.  To  show  this,  we  proceed  from 
the  equations  [XXIV.  (a)] 

(  P  =  Xx  cos  a  +  Xy  cos  (3  +  Xz  cos  y, 

(a)  \  Q=  Yxcosa+Yycos/3  +  F.cosy, 

[  R  =  Zx  cos  a  +  Zy  cos  (3  +  Zt  cos  7, 


in  which  a,  /3,  y  are  the  angles  between  the  normal  to  the  surface 
and  the  axes,  and  determine  the  position  of  the  surface  on  which 
the  components  of  stress  P.  Q,  R  act.  It  is  to  be  shown  that  this 
surface  may  have  such  a  position  in  the  body  that  the  stress  acts 
perpendicularly  to  it  ;  we  will  call  the  stress  in  this  case  the  principal 
stress  S.  The  angles  which  the  direction  of  S  makes  with  the  axes 
are  as  before,  a,  /3,  y,  and 

(b)  P  =  Scosa,     Q  =  Scos(3, 

Introducing  these  values  in  (a),  we  have 


f  (Xx  -  S)  cos  a  +  Xy  cos  ft  +  X,  cos  y  =  0, 
J 


(c) 

If  cos  a,  cos  (3,  cosy  are  eliminated  from  these  equations,  we  obtain 

(S*-(Xt+Y,,  +  Zt)S*  +  (XxY,+  YJt  +  ZtX,-Z*-X*-Y1?)S 
\      -  (XxYJt  +  2ZyXz  Yx  -  XJ?  -  YyX*  -  ZZYX2)  =  0. 

This  equation  has  always  one  real  root  A,  and  we  can  find  the  cor- 
responding values   of  a,  /?,   y  from   equations   (c)   and   the  relation 


70  THE  THEOEY  OF   ELASTICITY.  [CHAP.  n. 

cos2a  +  cos2/?  +  cos2y  =  1.  Therefore,  through  any  point  in  the  body 
there  may  be  passed  at  least  one  plane  having  the  property  that  no 
tangential  forces  act  on  it.  We  call  such  a  plane  a  principal  plant. 
Let  the  system  of  coordinates  be  so  rotated  that  this  principal 
plane  is  parallel  to  the  7/2-plane.  On  this  supposition,  we  have 
XX  =  A,  Yx  =  0,  ZX^Q.  The  equations  (c)  then  become 

(A  -  S)  cos  a  =  0  ;     (Y,,-  S)  cos  p+Yzcosy  =  Q; 


These  equations  are  satisfied  when  we  set 

S=A,     cosa  =  l,     cos  [3  =  cos  y  =  0. 

We  thus  return  to  the  principal  plane  already  found,  with  its  appro- 
priate normal  stress  A.  The  same  equations  are  also  satisfied  if  we 
set  cos  a  =  0  ;  cos  /3/cos  y  =  -  Yj(Yy  -  S)  =  -  (Zt  -  S)/Z,r 

Since  cos  a  =  0,  and  a  =  |TT,  the  new  principal  planes  are  perpen- 
dicular to  the  first  one.     We  have  further, 


and  cos  /3/cos  y=\(Y!l-Z2±  J(  Yy  -  Z 


These  equations  present  two  values  of  S  and  two  values  each  of 
(3  and  y.  If  we  represent  the  values  of  P  and  y  by  /5'  and  /3",  y' 
and  y"  respectively,  we  have 

cos  (3'  cos  P"  /  cos  y'  cos  y"  =  -1, 
and  hence  cos  ft'  cos  /?"  +  cos  y'  cos  y"  =  0. 

Since  the  corresponding  values  of  a  are  equal  to  JTT,  it  follows  that 
the  two  new  principal  planes  are  perpendicular  to  each  other. 

It  is  thus  proved  that,  in  general,  through  any  point  in  a  body, 
there  may  be  drawn  three  surface-elements,  and  only  three,  on 
which  only  normal  forces  act,  and  that  they  are  perpendicular  to 
one  another.  The  normal  stresses  corresponding  to  the  three  planes 
may  be  designated  by  A,  B,  and  C.  From  (d)  the  following  relations 
hold  among  these  normal  stresses  and  the  components  of  stress, 


(e) 

(     ABC=  XJ^i  +  1ZVX.YX  -  XtZ*  -  YJt?  -  ZY?. 

The  first  of  these  equations  should  be  especially  noticed;  it  shows, 
that  the  sum  of  the  normal  forces  for  three  planes  perpendicular  to  each 
other  is  constant. 


SECT.  XXVI.] 


THE   PRINCIPAL   STRESSES. 


71 


If  the  axes  of  the  system  of  coordinates  are  parallel  to  the  directions 
of  the  principal  stresses  A,  B,  and  C,  equations  (a)  become 

P  =  Aco$a,     Q  =  Bcos[3,     E=Ccosj. 

IfA>E>C,  and  we  set  A  =  B+SV  C=B-S2,  the  principal  stresses 
can  be  replaced  by  a  hydrostatic  stress  B  and  two  axial  stresses  Sl 
and  S.2,  the  first  of  which  is  a  tension,  the  second  a  pressure. 

This  investigation  shows  that  through  any  point  in  a  body  three 
planes  can  always  be  passed  which  are  acted  on  only  by  normal 
stresses,  equal  to  the  principal  stresses  A,  B,  and  C.  A,  B,  and  C 
are  the  three  roots  of  equation  (d) ;  their  directions  may  be  determined 
by  the  help  of  equations  (c).  A  makes  the  angles  a,  (3,  y,  with  the 
coordinate  axes.  We  write  cos  ax  =  /1}  cos  f3l  =  mv  and  cos  yl  =  nv  The 
corresponding  notation  for  B  and  C  is  exhibited  in  the  following 
table  : 


(g) 


From    equations    (c)    the    following    relations    hold    among    these 
quantities  : 


Am1  =  YJ1  + 


C13 


Em.  =  Yxlz  +  Yym.,  +  Yzn2  : 
Bnz  =  ZJZ  +  Zum.2  +  Zpz 
A>3, 


J3  +  Zym3  +  Ztna. 

These  equations  can  be  solved  for  the  components  of  stress  Xx,  Yy,  etc. 
These  quantities  may,  however,  be  determined  more  easily  in  the 
following  way.  Through  a  point  P  draw  the  lines  PA',  PB',  and  PC' 
parallel  to  the  directions  of  the  principal  stresses  A,  B,  and  C.  These 
three  lines,  together  with  a  plane  F  parallel  to  the  p-plane,  deter- 
mine a  tetrahedron.  The  plane  F  is  so  placed  that  the  tetrahedron 
is  infinitely  small ;  its  base  is  dF.  The  areas  of  the  faces  which  meet 
at  P  are  I^dF,  l2dF,  and  lsdF.  The  force  acting  on  unit  area  in 
1-^dF  in  the  direction  of  the  o>axis  is  Al^ ;  the  forces  acting  on  unit 
area  in  the  two  other  faces  are  B12,  Cls,  respectively,  and  the  force 


72  THE   THEORY  OF   ELASTICITY.  [CHAP.  n. 

acting  on  unit  area  in  dF  is  Xx.     That  the  tetrahedron  shall  not  move 
in  the  direction  of  the  ar-axis  we  must  have 


2B  .  l2dF+  13C  .  l3dF=  XxdF  or  X,  =  Al*  +  Bl.22  +  C732. 

By  a  similar  process  we  obtain  for  the  other  components  the  following 
equations  : 


2  +  Cl3n5, 
Cl3m3. 

It  may  easily  be  seen  that  these  values  of  the  components  of  stress 
satisfy  equations  (h),  if  the  known  relations  among  the  quantities 
given  in  (g)  are  taken  into  account. 


SECTION  XXVII.     FARADAY'S  VIEWS  ON  THE  NATURE  OF  FORCES 

ACTING   AT   A   DISTANCE. 

Newton  considered  the  action  between  two  masses  as  an  action  at 
a  distance  which  is  not  propagated  from  particle  to  particle  of  the 
medium  surrounding  the  masses.  Faraday,  on  the  other  hand,  in 
discussing  electrical  action,  held  that  the  intervening  medium  is  the 
seat  of  the  action  between  two  charged  bodies,  and  that  the  action 
is  transferred  from  particle  to  particle.  In  each  of  these  particles 
electricity  is  displaced  in  the  direction  of  a  line  of  force,  one  end 
of  which  becomes  positively  and  the  other  negatively  electrified. 

In  a  body  thus  polarized  the  particles  are  so  arranged  that  poles 
of  opposite  name  are  contiguous.  Hence  the  lines  of  force  tend 
to  contract,  and  a  state  of  stress  arises  in  the  medium.  This  stress 
is  similar  to  the  elastic  stress,  and  was  called  by  Maxwell  electrical 
elasticity.  In  Chapter  V.  of  his  Treatise  on  Electricity,  Maxwell, 
using  Faraday's  hypothesis,  developed  a  theory  which  we  will  now 
proceed  to  discuss.  Since  electrical  and  magnetic  forces  conform  to 
the  same  law  as  that  of  universal  attraction,  the  discussion  may  be 
made  perfectly  general,  and  applicable  to  all  forces  between  bodies 
which  are  inversely  proportional  to  the  squares  of  the  distances 
separating  the  bodies. 

Let  the  potential  $  be  given  for  all  points  of  the  region.  The 
density  p  is  determined  from  the  potential  by  Poisson's  equation 

(a)  32^/?.r2  +  3-Y/3y2  +  ^/dz2  +  ±TTP  =  0. 


SECT,  xxvii.]         FORCES  ACTING  AT  A  DISTANCE.  73 

The  mass  pdv  contained  in  the  volume-element  d-v  is  acted  on  by 
a  force  whose  components  are 


The  upper  sign  holds  for  magnetic  or  electrical  attractions,  the  lower 
for  mass  attractions.  Introducing  the  value  of  p  given  in  (a)  the 
component  acting  in  the  direction  of  the  .r-axis  becomes 


This  quantity  must  be  capable  of  representation  as  the  sum  of  three 
differential  coefficients  with  respect  to  x,  y,  and  z.     We  have 


Hence   the   force  which   acts  in  the   direction  of  the  a-axis  on  the 
volume-element  dv  is 


If  we  designate  the  components  of  force  which  act  on  the  unit 
of  volume  by  (X),  (Y),  and  (Z)  [XXV.],  and  if,  for  brevity,  we  set 

X  =  -  -d^px,     Y=- 
we  obtain 

f  (X)  =  ± 

(b)  |  (F) 

(  (Z)  =  ±  I/STT  .  [Zd(XZ)/-dx  +  Zd(YZ)l*dy  +  3(Z2  -X2- 
Since  these  equations  are  perfectly  analogous  to  those  which  deter- 
mine the  force  with  which  stresses  act  on  the  unit  of  volume,  we 
may  consider  forces  acting  at  a  distance  as  arising  from  stresses  in 
the  medium.  If  we  are  dealing  with  universal  mass  attraction,  the 
ether  may  be  assumed  to  be  the  intervening  medium;  if  we  are 
discussing  electrical  actions,  the  dependence  of  the  stress  in  the 
ether  on  the  matter  which  fills  the  region,  air,  water,  etc.,  must  be 
taken  into  account.  It  is  not  necessary  to  enter  upon  this  question 
in  our  treatment  of  the  subject. 

A  comparison  of  equation  (b)  with  equation  XXV.  (b)  shows  that 
Xx  =  ±  (X2  -  Y2  -  Z2)/Sir,     Y 

(c)  -I  Yy=±(Y2-X2-Z2)/87r,     Zx 

Zz=±  (Z2  -X2-  FO/STT,    X, 


74  THE   THEORY   OF   ELASTICITY.  [CHAP.  11. 

To  determine  the  principal  stresses  in  the  medium,  we  use  equations 
XXVI.  (e),  Avhich  give 

A  +  B  +  C  =  +  (X-  +  F-  +  Z2)/8ir, 

BC+AC+AB=  -  ((A'2+  F2  +  Z2)/87r)2, 

ABC  =  ±  ((X'2  +  Y1  +  Z-')/8ir)8, 

If  we  set  (d)  (X'2+Y*  +  Z'2)/8ir  =  S,  A,  B,  and  C  are  the  roots  of 
the  equation  D3  ±  SD'2  -  SW  +  Ss  =  0  or  (D  +  S)  (D  ±  S)'2  =  0.  We 
have  therefore  either  (e)  A  =  +  S,  B=C  =  -S  or  A  =-  S,  B=C  =  +  S. 
Hence  two  principal  stresses  are  always  equal.  In  order  to  deter- 
mine their  directions,  a,  (3,  and  y  must  be  calculated  from  XXVI.  (c). 
It  is  easiest  to  determine  the  directions  of  the  equal  stresses  B  and 
C.  If  the  values  of  ±S,  given  in  (d),  are  substituted  for  S  in  the 
equations  referred  to  [XXVI.  (c)],  using  the  negative  value  of  S  in 
combination  with  the  positive  value  of  Xx,  etc.,  and  vice  versa,  we 
obtain  (f)  A"  cos  a  +  Ycos  (3  +  Zcosy  =  Q.  Hence  both  of  the  equal 
principal  stresses  are  perpendicular  to  the  direction  of  the  force ;  the 
third  principal  stress  is  in  the  direction  of  the  force,  and  is  equal 
to  the  square  of  the  force  divided  by  STT. 

It  has  thus  been  shown  that  all  forces  acting  at  a  distance  may 
be  explained  by  a  state  of  stress  in  an  intervening  medium.  From 
this  point  of  view  universal  mass  attraction  is  replaced  by  a  negative 
stress,  that  is,  a  pressure,  in  the  direction  of  the  lines  of  force,  and 
a  positive  stress,  that  is,  a  tension  in  all  directions  perpendicular  to 
the  force.  A  surface-element  which  lies  perpendicular  to  the  direction 
of  the  force  is  acted  on  by  a  tension  which  is  equal  to  the  force. 
In  the  case  of  magnetic  and  electrical  attractions  the  opposite  holds 
true.  There  is  no  independent  evidence  for  the  existence  of  such 
stresses  in  the  case  of  gravity ;  but  several  phenomena  in  electricity 
indicate  that  the  medium  between  two  electrified  bodies  is  in  a  state 
of  stress,  and  no  facts  are  known  that  are  inconsistent  with  the 
assumption  that  this  stress  is  the  cause  of  the  forces  acting  on  the 
bodies. 


SECTION  XXVIII.    DEFORMATION. 

If  a  body  changes  its  shape  or  its  position  in  space,  one  of  its 
points,  whose  coordinates  are  originally  x,  y,  .:,  may  be  so  displaced 
that  its  coordinates  become  x  +  £,  y  +  ??,  z+  £  £,  ->;,  £  are  the  pro- 
jections of  the  path  which  P  has  traversed  or  the  components  of 


SECT.  XXVIII.] 


DEFORMATION. 


the  displacement.  If  £,  77,  £  are  given  as  functions  of  the  time,  the 
position  of  the  point  P  at  any  instant  is  determined.  The  motions 
of  the  separate  points  of  the  body  are  in  general  different,  that  is, 
£,  ?/,  £  are  functions  of  x,  ?/,  z.  We  will  first  consider  some  simple 
motions  of  the  body. 

If  £,  77,  £  are  equal  for  all  points  of  the  body,  the  points  all 
move  through  equal  distances  and  in  the  same  direction  ;  the  motion 
is  a  translation.  In  this  motion  all  parts  of  the  body  remain  at  fixed 
distances  from  each  other,  and  there  are  no  internal  forces  developed. 
This  holds  also  in  the  case  of  a  rotation  of  the  body  about  an  axis. 
Let  the  axis  of  rotation  be  parallel  to  the  o;-axis,  and  pass  through 
the  point  P  (Fig.  41),  whose  coordinates  are  x,  ?/,  s.  Let  a  point  Q, 


whose  coordinates  are  x',  y  ,  z,  traverse  the  path  Qli  =  lix  .  r,  where 
r=QS  is  the  distance  of  the  point  Q  from  the  axis,  aud  hx  is  the 
angle  of  rotation.  By  this  rotation  the  ^/-coordinate  is  diminished 
by  BB'  =  QR(z'  -z}/r  =  hx(z'-  z),  and  the  .^-coordinate  is  increased  by 
CO'  =  QR(y'  -y}lr  =  hx(ij  -y).  If  the  body  rotates  at  the  same  time 
about  two  other  axes,  which  are  parallel  to  the  y-  and  z-axes,  and 
if  the  angles  of  rotation  are  designated  by  liy  and  hz  respectively, 
the  coordinates  of  Q  are  increased  by  £,  17,  £,  which  have  the  following 
values  : 


*  W  --9)*.-  &-*)*,- 

We  may  now  proceed  to  the  discussion  of  the  general  case,  in 
which  the  points  of  the  body  change  their  relative  positions.  Let 
the  point  P,  whose  coordinates  are  .T,  y,  z,  pass  during  this  motion 
to  the  point  P',  whose  coordinates  are  x  +  g,  y  +  r,,  z  +  £;  let  another 


76  THE  THEORY  OF  ELASTICITY.  [CHAP.  n. 

point  Q,  whose  coordinates  are  originally  x\  y',  z1,  pass  to  the  point 
Q,  whose  coordinates  are  x'  +  £',  y'  +  i/,  z'  +  £  '.  If  £  is  a  known  func- 
tion of  x,  y,  z,  we  will  have 

£'  =  £  +  (x'  -  x^j-dx  +  (y  -  y)3£/3//  +  (z'  -  *)3£/3*  +  .  .*  . 
We  may  assume  that  P  and  Q  are  infinitely  near,  so  that 

x'  —  x  =  dx,     y'  —  y  =  dy,     z'  —  z  =  dz. 

Neglecting  terms  of  the  second  order  we  obtain  the  following  relations, 

£'  =  £  +  3£/d.r  .  dx  +  ^I'dy  .  dy  +  V£/3z  .  dz, 

77'  =  ?;  +  'drf/'dx  .  dx  +  'drj/'dy  .  dy  +  'dq/'dz  .  dz, 

('  =  £  +  3£/3.r  .  dx  +  'dtfdy  .  dy  +  *dtfdz  .  dz. 

By  introducing  the  following  notation, 


yx  =  xy 

we  obtain 

(£'  =  £  4-  x,dx  +  xvdy  +  x.dz  -  hzdy  +  hjlz, 
rj'  =  r]  +  yjlx  +  yydy  +  yjlz  -  hjz  +  hzdx, 
T  =  C  +  M*  +  z^y  +  zzdz  -  hydx  +  h,dy. 

These  equations  determine  the  motion  of  a  point  in  the  neighbour- 
hood of  P.  This  motion  is  compounded  of  a  translation,  whose 
components  are  £,  »;,  £  a  rotation,  whose  components  are  h^  ky,  h:, 
and  two  motions,  determined  by  x^  y^  zz  and  zf  xa  yf  If  we  confine 
our  attention  to  the  way  in  which  the  form  of  the  body  changes, 
we  need  only  consider  the  motion  whose  components  e?£,  d^,  d£  are 
determined  by  the  following  equations  : 


dr)  =  yrd 

d£  =  ztdz  +  z,dx  +  zydy. 

To  interpret  the  coefficients  xz,  y^  sz  and  z^  a;,,  y^  we  assume  that 
all  except  xx  are  equal  to  zero.  Then  dg  =  xx.dx  and  dr)  =  d£=Q.  The 
change  of  form  corresponding  to  this  is  a  dilatation  of  the  body  in 
the  direction  of  the  ar-axis,  by  which  dx  increases  by  d£.  The  co- 
efficient xx  therefore  represents  the  dilatation  of  a  unit  of  length 
parallel  to  the  x-axis,  or  is  the  dilatation  in  the  direction  of  the 
;e-axis.  Hence  yv  and  zt  are  the  dilatations  in  the  directions  of  the 
y-  and  2-axis  respectively. 


SECT,  xxvin.]  DEFORMATION.  77 

If,  on  the  other  hand,  all  the  coefficients  vanish  with  the  exception 
of  zy,  we  have  dg  =  Q,  d^=zy.dz,  d£=zy.dy.  The  particles  are  dis- 
placed in  a  plane  parallel  to  the  yz-plane,  and  their  distances  from 
the  yz-plane  remain  unchanged.  Let  the  original  coordinates  of  the 
point  P  (Fig.  42)  be  x,  y,  z;  let  ABCD  be  a  square,  the  length  of 
whose  sides  is  2a.  The  point  A,  whose  original  coordinates  were 
x,  y  +  a,  z  +  a,  referred  to  the  axes  PY  and  PZ,  is  displaced  to  A\ 
whose  coordinates  are  a  +  zya,  a  +  zya. 
A  therefore  lies  on  PA  produced.  The 
points  B  and  D  are  displaced  to  B'  and 
Z>',  which  lie  on  BD ;  C  is  displaced  along 
AC  produced  to  C".  The  square  ABCD 
becomes  the  rhombus  A'B'C'D.  This 
change  of  form  is  called  a  shear;  the 
quantities  zy,  xa  yx  are  called  components 


of  shear. 

In  the  theory  of  elasticity  we  consider    ^  FIG  49 

only  very  small  deformations  of  the  body  ; 

the  components  xx,  yy,  etc.,  are  consequently  small  quantities,  whose 
second  and  higher  powers  may  be  neglected.  The  volume  of  the 
body  is  not  changed  by  a  shear;  the  square  whose  area  is  4a2  will 
become  a  rhombus  A'B'C'D'  whose  area  is 

2PA' .  PB'  =  2(a  4  zya)j2(a  -  zya)j2  =  4a2(l  -  z?). 

If  we  neglect  zy-,  the  area  of  the  square  is  equal  to  that  of  the 
rhombus;  hence  the  volume  will  not  be  changed  by  the  shear. 

From  Fig.  42  it  is  evident  that  the  infinitely  small  angle  between 
AB  and  A'B'  is  equal  to  azy/a  =  zy;  hence  the  right  angle  DAB  is 
diminished  by  the  shear  by  2zy,  so  that 


As  the  result  of  a  dilatation  determined  by  xa  yy,  zz,  the  volume 
of  the  parallelepiped  dxdydz  becomes  dxdydz(\  +  xx)(l  +yy)(l+zz).  If 
the  components  of  dilatation  are  supposed  infinitely  small,  we  may 
neglect  their  second  and  higher  powers.  Hence  the  increase  in 
unit  volume  is  0  =  xx  +  yy  +  zz.  0  is  called  the  volume  dilatation.  Sub- 
stituting the  values  of  x^  y^  za  we  have  also 

(e)  Q  =  'd£fdx  +  'fr)l'dy  +  'dtrdz. 

Let  dr  be  an  element  of  a  straight  line  which  makes  the  angles 
a,  /3,  y,  with  the  coordinate  axes ;  then 

dx  =  dr  cos  a,     dy  =  dr  cos  (3,     dz  =  dr  cos  y. 


78  THE   THEORY  OF  ELASTICITY.  [CHAP.  n. 

By  the  deformation  dr  becomes  dr',  and  makes  the  angles  a,  (3',  y' 
with  the  axis,  so  that 

dx  +  dg  =  dr'cosa  ;     dy  +  djj  =  dr' cos (B' ;     dz  +  d£  =dr' cosy', 

from  which  dg,  dt],  d£  may  be  determined  by  equations  (d).  If  the 
direction  of  the  line  dr  remains  unchanged,  we  have  a  =  a,  /3  =  /3', 
and  7  =  7',  and  hence  dg  =  dpcosa,  drj  =  dp  cos  (3,  d£=dpcosy,  where 
dp  =  d(r'-r).  The  length  dp  is  the  elongation  of  dr,  and  dp/dr  is 
the  dilatation  s  in  the  direction  of  the  line  dr.  Hence  we  have 

.9  =  dp/dr. 
Equations  (d)  then  assume  the  following  form: 

{(xx  -  s)cos  a  +  xycos  (3  +  xzcos  7  =  0, 
yxcos  a-  +  (yy-  s)  cos  (3  +  yzcos  y  =  0, 
2^ cos  a  +  zy  cos  (3  +  (zz-  s)  cos  7  =  0. 

A  comparison  of  these  relations  with  those  of  XXVI.  (c)  shows  that 
they  both  may  be  interpreted  in  a  similar  way. 

There  are  therefore  three  directions  perpendicular  to  each  other, 
called  the  principal  axes  of  dilatation,  in  which  only  dilatations  occur ; 
every  line-element  which  is  parallel  to  one  of  these  three  directions 
contains  after  deformation  the  particles  which  were  in  it  before  the 
deformation.  This  conclusion  holds  only  on  the  supposition  that 
the  body  does  not  rotate,  a  supposition  which  has  been  made  in 
deducing  equations  (d).  If  the  principal  dilatations  thus  deter- 
mined are  called  a,  b,  c,  we  have,  as  in  XXVI.  (e), 

f  a  +  b  +  c  =  xx  +  yy  +  zz, 

(g)  j  be  +  ac  +  ab  =  z$,  +  x^  +  ypx  -  z*  -  x*  -  y*, 

{  abc  =  xji^t  +  2ZfCjff  - x^f - yp* - zjy*. 

The  first  of  these  equations  shows  that  the  volume  dilatation  does  not 
depend  on  the  position  of  the  system  of  coordinates. 

In  the  same  way  as  that  in  which  the  components  of  stress  are  ex- 
pressed in  terms  of  the  principal  stresses  [XXVI.  (i)]  xa  xy,...  may  be 
expressed  in  terms  of  the  principal  dilatations  a,  b,  and  c.  Denoting 
the  cosines  of  the  angles  which  the  direction  of  a  makes  with  the  axes 
by  lv  mv  nv  and  the  cosines  of  the  angles  which  b  and  c  make  with 
the  axes  by  12,  m.2,  n2;  13,  m3,  %,  we  obtain 

f  xx  =  all2  +bl.22  +c/32;     2,«om1n 

(h)  |  yy  =  a™,l2  +  bm22  +  cffl./ ;   xt-=alln1 

I  zt  =  an^  +  bn.2-  +  en./ ;    yx  =  al-pi^ 


SECT.  XXIX. 


STRESSES   AND   DEFORMATIONS. 


SECTION   XXIX. 


RELATIONS  BETWEEN  STRESSES  AND 
DEFORMATIONS. 


The  study  of  the  deformations  of  an  elastic  body  has  shown  that 
a  parallelepiped  which  is  stretched  by  forces  applied  to  its  ends,  is 
increased  in  length  and  diminished  in  cross  section.  If  we  only 
consider  forces  which  are  so  small  that  the  limits  of  elasticity  are 
not  exceeded,  the  elongation  s  per  unit  of  length  is  s  =  S/E,  where 
E  is  the  coefficient  of  elasticity  and  S  the  force  acting  on  the  unit  of 
surface.  The  contraction  s'  per  unit  of  length  parallel  to  the  end 
surfaces,  is  given  by  s'  =  k  .  S/E,  where  k  is  a  constant.  It  is  assumed 
that  the  body  is  isotropic,  that  is,  equally  elastic  in  all  directions 
and  at  all  points. 

We  will  first  consider  a  rectangular  parallelepiped,  whose  edges 
are  parallel  with  the  coordinate  axes.  The  normal  forces  are  denoted 
by  Xx,  Yy,  Za  and  a  unit  of  length  which  is  parallel  to  the  a-axis 
increases  by  xx;  the  units  of  length  which  are  parallel  to  the  y-  and 
2-axes  respectively  increase  by  yy  and  zz.  We  then  have 


-k(Xx+Yy)/E. 
From  XXVIII.  (e)  the  volume  dilatation  0  is 


From  these  equations  we  deduce 

Xx  =  kEQ/(l  +  k)(l-  2k)  +  Exx/(l  +  k). 
Setting  A.  =  kE/(  I  +  k)  (1  -  2k),    p.  =  i£/(l  +  k), 

we  obtain  (a)  X,  =  XO  +  2/tr,  ;    Yy  =  XQ  +  2p.yy; 
and  by  addition 
(b)  Xx  +  Yy  +  Zz  =  (3X  +  2^)9. 

To  investigate  the  relation  between  the 
shears  and  the  tangential  forces  we  may  use 
the  following  method,  due  to  V.  v.  Lang.* 
If  the  prism  ABCD  (Fig.  43)  is  stretched  by 
the  tension  S  applied  to  each  unit  of  surface 
of  its  ends  AB  and  CD,  it  takes  the  form 
AB'C'D'.  Four  plane  sections  EF,  FG,  GH, 
and  HE  are  passed  through  the  prism,  which 
mark  out  the  rectangle  EFGH  on  a  plane 
parallel  to  the  axis;  the  rectangle  EFGH  becomes  by  deformation 
*  V.  v.  Lang,  Theoretisc.he  Physik;  §  411. 


80  THE   THEOEY  OF   ELASTICITY.  [CHAP.  n. 

the  parallelogram  E'F'G'H'.  The  angle  AFE  is  represented  by  </>. 
The  tangential  stress  T,  which  acts  on  the  surface  EF  in  the 
direction  EF,  is  given  [XXIII.  (b)]  by  T  =  S  sin  <f>  cos  <£. 

Since  <  BFG  =  |?r  -  <£,  the  same  tangential  stress  T  acts  on  GF 
in  the  direction  GF.  On  deformation  the  angle  AFE  becomes 
AF'E'  =  <f>  +  d<f>,  and  we  have 


Now  since  s=--SjE  and  s'  =  kS/E  are  infinitely  small,  we  have 

(  1  +  s)/(l  -  s')  =  1  +  s  +  s'  =  1  +  (1  +  k)S/E. 
Further,  we  have 


and  tg  (<£  +  dfy  =  tg  <f>  +  d(f>/co$2<j>, 

so  that  d<l>  =  (l+k)S  sin  <f>  cos  <£/  E  =  (  1 

Hence  the  change  of  the  angle  <£  is  proportional  to  the  tangential 
stress  T.  Since  the  same  tangential  stress  acts  on  GF  as  on  EF,  the 
angle  BFG  increases  by  d<f>,  the  angle  EFG  diminishes  by  2d(f>,  and 
the  angle  FGH  increases  by  2d<f).  The  shear  is  thus  equal  to  2e?<£, 
and  we  have  2d<f>=2(l+k)T/E.  But  2d<f>  is  the  quantity  2zy  pre- 
viously introduced,  when  the  rectangle  EFGH  is  parallel  to  the 
2/3-plane;  and  hence  T=Zy  and  zy  =  (l+k)ZJE.  If  we  set 


we  have  (c)  ^  =  2^,     A>2/^z,     Yx  =  2^yx. 

The  equations  (a)  and  (c)  are  the  solution  of  the  problem,  to  find 
the  components  of  stress,  when  the  deformations  are  given,  and 
conversely.  They  contain  only  two  constants,  A  and  /*,  which  involve 
the  deformations  caused  by  simple  dilatation  in  the  following  way  : 

f   X  =  *£/(l  +*)(!-  2*);     ,*  = 

t£-<8V+4i*V(*+p);  * 

Since  A  and  /A  are  positive,  &  must  be  less  than  \. 

The  relations  between  the  elastic  forces  and  the  deformations  may 
also  be  derived  by  another  method.  Let  the  principal  stresses  A,  B, 
and  C,  at  the  point  P,  be  known  in  magnitude  and  direction  [cf.  XXVI. 
(g)].  An  infinitely  small  parallelepiped,  whose  edges  are  parallel 
to  the  directions  of  the  stresses  A,  B,  and  C,  is  extended  in  those 
three  directions.  The  increments  a,  b,  and  c  of  the  unit  of  length 
are  parallel  to  A,  B,  and  (7,  and  as  in  (a),  we  have 
(e)  A  =  \e  +  -2iM,  B  =  XQ  +  2pb, 

when  0  =  a  +  b  +  c,  or  [XXVIII.  (g)],  Q  = 


SECT,  xxix.]  STRESSES   AND   DEFORMATIONS.  81 

By  applying  the  formula  XXVI.  (i),  we  obtain  the  equation 

Xx  =  X0  +  2/x(a/12  +  bl/  +  c/32), 

which,  from  XXVIIL  (h),  becomes  Xx  =  X9  +  2/xa^     The  expressions  for 
Yy  and  Zt  are  obtained  in  a  similar  way. 

From  XXVI.  (i),  we  have 


and  hence  [XXVIIL  (h)]  Zy  =  2fj.zy.  We  obtain  the  expressions  for  A", 
and  Yx  in  a  similar  way. 

The  coefficients  E  and  k  depend  on  the  nature  of  the  body.  It 
was  at  one  time  believed  that  k  had  the  same  value  for  all  bodies. 
This  opinion  was  first  expressed  by  Navier.  He  assumed  that  bodies 
are  made  up  of  material  points  which  repel  one  another,  and  on  this 
assumption  concluded  that  k  =  £.  Poisson  also  had  the  same  opinion. 

While  k  is  a  mere  number,  the  coefficient  of  elasticity  E  is  deter- 
mined by  E  =  S/s;  the  fraction  \\E  is  called  the  modulus  of  elasticity. 
Sis  the  force  which  acts  on  the  unit  of  surface,  and  [III.]  its  dimensions 
are  LT~2M/L2  =  L~1T~'2M.  Since  s  is  the  ratio  between  the  elongation 
and  the  original  length,  it  is  also  a  mere  number.  Hence  the  dimen- 
sions of  E  are  L~1T-'2M. 

In  practical  units  E  denotes  the  number  of  kilograms  which  would 
produce  an  elongation  in  a  rod  of  one  square  millimetre  cross  section, 
such  that  its  length  is  doubled.  In  order  to  transform  it  into  absolute 
measure,  we  notice  that  the  weight  of  one  gram  is  about  equal  to 
981  dynes,  and  that  therefore  the  weight  of  one  kilogram  is  equal 
to  981  000  dynes.  The  cross  section  must  be  taken  equal  to  1  sq.  cm., 
and  the  number  must  therefore  be  multiplied  by  100,  so  that  the 
factor  of  transformation  becomes  98,100,000.  According  to  Wertheim, 
E  equals  17278  in  practical  units  for  English  steel  ;  therefore,  in 
absolute  units  it  equals  17278.  981  .  105=  1,695  .  1012. 


In  the  case  of  fluids,  the  discussion  is  simplified  by  the  condition 
that  a  fluid  always  yields  to  tangential  forces,  so  that,  when  it  is  in 
equilibrium,  there  are  no  tangential  forces  acting  in  it.  This  condition, 
from  (c),  enables  us  to  set  /^  =  0.  If  the  fluid  is  subjected  to  the 
pressure  p,  and  if  its  volume  v  is  thereby  diminished  by  dv,  we 
have  from  (b) 


(f)  -{  or,  since  /^  =  0, 

dv  =pv/X. 


82  THE  THEORY  OF   ELASTICITY.  [CHAP.  n. 

If,  for  example,  the  unit  volume  of  water  is  diminished  by  0,000  046, 
when  the  pressure  is  increased  by  1  atmosphere,  we  have 
X=pv/dv  =  7&  .  13,596  .  981/0,000,046  =  2,204  .  1010. 
In  the  case  of  gases,  if  we  represent  the  original  pressure  by  P, 
and  its  increase  by  p,  Mariotte's  law  gives  the  equation 

Pv  =  (P+p)(v-dv). 
Assuming  that  p  is  very  small  in  comparison  with  P,  we  obtain 

dv=pv/P, 
and  therefore,  for  gases,  we  have  from  (f),  (g)  P=X. 


SECTION  XXX.     CONDITIONS  OF  EQUILIBRIUM  OF  AN  ELASTIC  BODY. 

If  a   force   whose   components   are  X,   Y,  Z  acts  on   the   unit  of 
mass  of  a  body,  we  have  from  XXV.  (c) 

(a)  VX,J'dx  +  'dX,l'dy  +  'dXJ'dz  +  pX=Ot  etc. 
Further  [XXIX.  (a)  and  (c)]  we  have 

(b)  Xx  =  XQ  +  2fj..^px,     Zy=Yi  =  p.(dil'dy  +  'd-ql^z\  etc. 

If  the  values  for  XM  etc.,  are  substituted  in  (a),  it  follows  that 

c  (X  +  p)  .  ae/a-e  +  /*V2£  +  pX  =  0, 
(c) 


I 

By  the  use  of  our  former  symbols  for  the  components  of  rotation,  viz., 

|  2A 

\ 
equations  (c)  become 


?  - 

<e)  (X  +  2/t) 

I  (x  +  2/*) 

If  the  first  equation  is  differentiated  with  respect  to  .r,  the  second 
with  respect  to  y,  and  the  third  with  respect  to  ,r,  we  have  by 
addition  when  p  is  constant, 

(X  +  2/x)y26  +  P(dXfix  +  VY^ij  +  -dZI-dz)  =  0. 

If  X,  Y,  and  Z  are  the  derivatives  of  a  potential  ^,  and  if  y'-^  =  0 
everywhere  within  the  body,  then  (f)  y20  =  0. 

This  result  must  be  supplemented  by  the  conditions  of  equilibrium 
of  the  surface  of  the  body.  The  force  acting  on  the  surface-element 
dS,  whose  components  are  FQ'R,  is  in  equilibrium  with  the  elastic 


SECT,  xxx.]         STRESSES   IN   A   SPHERICAL  SHELL.  83 

forces  which  act  on  the  parts  of  the  body  contiguous  to  dS.  If 
.Xx,  Yy'}  etc.,  are  the  components  of  the  elastic  forces,  we  have 

(g)  F  =  XX  COS  a  +  Xy  COS  ft  +  Xi  COS  7. 

There  are  similar  values  for  Q'  and  R'.  The  symbols  a,  (3,  7 
represent  the  angles  which  the  normal  to  the  surface  directed  out- 
ward makes  with  the  coordinate  axes. 

We  assume  (h)  £  =  ax,  r)  =  by,  {=cz,  where  a,  b,  and  c  are  constants. 
.£,  ij,  and  £  are  therefore  linear  functions  of  x,  y,  and  z,  and  £  depends 
only  on  x,  77  only  on  y,  and  £  only  on  z.  On  this  supposition 
[XXVIII.  (b)]  the  deformation  of  the  body  is  made  up  of  dilatations 
•only.  The  volume  dilatation  is  0  =  a  +  i  +  c;  hence  the  values 
assumed  for  £,  r/,  £  satisfy  equations  (c),  if  we  neglect  the  action  of 
external  forces.  We  have  further 


If  we  set  XX  =  S,    Yy  =  0, 

we  have  b  =  c  and 

S=X(a  +  2b)  +  '2pa;     0  = 
The  last  equation  gives 

b/a=  -£A/(X  +  /*)=  -k, 

and  the  first  #  =  a(3A/*  +  2/*2)/(A  +  p)  =  Ea. 

The  equations  thus  obtained  give  the  law  of  the  expansion   of  an 
elastic  prism. 


SECTION  XXXI.     STRESSES  IN  A  SPHERICAL  SHELL.  ' 

Suppose  a  spherical  shell,  bounded  by  two  concentric  spheres, 
whose  radii  are  rl  and  r2.  Of  these  we  assume  r2  >  rr  Suppose  a 
constant  hydrostatic  pressure  pl  applied  to  the  inner  surface,  and 
a  similar  pressure  p.2  applied  to  the  outer  surface.  The  pressures  p^ 
and  p2  are  perpendicular  to  the  surfaces.  Let  the  centre  0  of  the 
sphere  be  the  origin  of  coordinates,  and  let  the  distance  from  0  of 
any  point  in  the  shell  be  r.  On  the  hypothesis  that  has  been  made 
with  respect  to  the  pressures,  all  points  lying  in  the  same  spherical 
surface  having  the  centre  0,  receive  equal  displacements  from  the 
centre.  Let  the  displacement  of  the  point  considered  be  er,  where  e 
is  a  very  small  quantity.  We  then  have 


84  THE  THEORY  OF   ELASTICITY.  [CHAP.  n. 

Since  e  is  a  function  of  r  only,  we  may  set 

£  =  fr .  x/r  =  d<j>/dr .  "drfdx  =  'd^j'dx, 

where  <f>  is  a  new  function  of  r.  We  may  represent  ij  and  £  in  a 
similar  way,  so  that 

(b)  £  =  3.£/3.£,     ^30/fy,     t=3</>/32. 

Hence  we  have  (c)  6  =  V2<£. 

The  equations  XXX.  (c),  if  the  action  of  gravity  is  neglected, 
become 

(A  +  2fi).V23<£/9.c  =  0,     (A.  +  2/z).  V23</>/3y  =  0,     (A  +  2/*) .  V^fdz  =  0, 
so  that  (d)  0  =  V-<£  =  a,  where  a  is  a  constant. 
From  XXX.  (b)  the  components  of  stress  are 

c  Xx  =  Aa  +  2p . 
(e)  J   Yy  =  Xa  +  2 p. 

I  Z,  =  Xa  +  2/i.32^/?38;    F,  =  2/t .  32</>/3o%. 
The   stress   in   a   surface-element  perpendicular   to  r  is   given   by 
XXIV.  (a),  if  we  set 

cos  a  =  z/'r,     cos  (3  =  y/r,     cos  y  =  z/r. 
If  the  components  of  stress  are  P,  Q,  and  R,  we  have 

P  =  Xa .  a-/r  +  2/*(.r/r .  32<^>/3a;2  +  y/r .  ^fdxdy  +  zlr .  &<t>/'dzdz). 
Using  the  equations 

Pt/dr*  -  x*/i* .  d<f>/dr  +l/r.  dj/dr, 
xy/f* .  d*<j>/dr*  -  xy/r* .  d<j>/dr, 
xz/r* .  d^/dr*  -  xzji*  .  d<f>/dr, 
we  have  P  =  (\a  +  2p. .  d^jdr*) .  xfr. 

Similar  expressions  may  be  obtained  for  Q  and  It.  Hence  a  prin- 
cipal stress  (f)  A  =  Aa  +  2/* .  d*<j>/dr2  acts  on  the  surface-element 
considered. 

For  a  surface-element  which  contains  r,  the  components  are  obtained 
in  the  same  way.  If  a,  /?,  y  are  the  angles  which  the  normal  to  the 
surface-element  makes  with  the  axes,  we  have 

P  =  Aa  cos  a  +  2/*(92<£/3;e2 .  cos  a  +  32<£/3.c3y .  cos  ft  +  32</>/3x3^ .  cos  y). 
If  we  notice  that  in  this  case 

cos  a .  x/r  +  cos  (3 .  y/r  +  cos  7 .  z/r  =  0, 

and  use  the  expressions  given  above  for  the  differential  coefficients, 
we  have  P  =  (\a  +  2/*/r .  d<J>/dr)  cos  a. 

We  may  obtain   Q  and  R  by  replacing  a  by  ft  and  y  respectively. 


SECT,  xxxi.]  TOESION.  85 

Hence  the  principal  stress  B  acting  on  the  element  is 

From  (d)  and  XV.  (1)  we  have 

and  therefore 

(h)  d<j>/dr  =  *ar  +  b/r* ;     d2<j>/dr2  =  ^a-  26/r3. 

From  (f)  and  (g)  it  follows  that 

A  =  (X  +  |/i)a  -  4/xJ/r3  ;     B  =  ( A.  +  |/z)a  +  fyb/r3. 
For  r  =  rv  A=  -pv  and  for  r  =  rz,  A=  -p2,  therefore 
a  =  3/(3X  +  2/x) .  (pfi 


and       A- 
B  = 


SECTION  XXXII.    TORSION. 

Let  us  consider  a  circular  cylinder  whose  axis  coincides  with  the 
£-axis;  and  let  the  circle  in  which  the  xy -plane  cuts  the  cylinder 
be  the  end  of  the  cylinder  and  be  fixed  in  position.  If  torsion  is 
applied  to  the  cylinder,  a  point  at  the  distance  r  from  the  axis  describes 
an  arc  r<f>,  parallel  to  the  xy-plane,  whose  centre  lies  on  the  z-axis. 
This  angle,  in  the  case  of  pure  torsion,  is  proportional  to  the  distance 
of  the  point  from  the  zy-plane,  so  that  <f>  =  kz,  where  k  is  a  constant. 
The  displacement  of  this  point  is  krz,  and  its  components  £,  77,  £  are 
(a)  £=-kyz,  r,  =  kxz,  f=0. 

Using  these  values,  we  find  that  the  volume-dilatation  0  is  zero,  that 
is,  pure  torsion  dots  not  cause  a  change  of  volume.  We  have  further 
[XXX.  (b)],  Xx  =  0,  Yy  =  0,  ZZ  =  Q, 

and  hence  no  normal  forces  act  on  the  surfaces  which  are  parallel 
to  the  coordinate  planes.     On  the  other  hand,  we  have 
Z^fjJcx,     Xt=-fdy,     Yx  =  0. 

A  surface-element  perpendicular  to  the  £-axis  is  acted  on  by  the 
tangential  forces  Y,=  +  fj.kx  and  Jfz=  - pJcy,  whose  resultant  pkr  is 
perpendicular  to  the  radius  r  and  to  the  z-axis. 

By  XXIV.  (a)  we  reach  the  same  result.     That  is,  we  get 
f  P  =  -  ftky  cos  y,     Q  =  pkx  cos  7, 
I  R  =  -nkycosa 


86  THE   THEORY  OF   ELASTICITY.  [CHAP.  n. 

For  the  stress  on  the  surface  of  the  cylinder  we  must  set 

cos  a  =  x/r,     cos  f3  =  y/r,     cos  y  =  0. 
We  will  then  have          P  =  0,     Q  =  0,    ^  =  0. 

Hence  a  surface-element  perpendicular  to  the  radius,  or  which  is  part  of 
the  surface  of  a  circular  cylinder  whose  axis  is  the  z-axis,  is  not  acted  on- 
by  a  force. 

To  find  the  surface-elements  on  which  the  only  forces  which  act 
are  normal  forces,  we  use  equation  XXVI.  (d),  which,  in  the  case 
before  us,  becomes 


If  A,  B,  and  C  are  the  roots  of  this  equation,  we  can  set 
A  =  0,     B  =  pkr,     C=-ph\ 

If  the  angles  between  the  axes  and  the  normal  to  one  of  these  surface- 
elements  are  represented  by  a,  /3,  y, 

S  cos  a  =  —  p.ky  cos  y,    S  cos  f3  =  pJcx  cosy,    Scosy  =  —  //.£?/  cos  a  +  //Xvr  cos  /£. 

If  we  substitute  in  these  equations  the  particular  values  of  S  given 
by  A,  B,  and  (7,  the  values  of  a,  J3,  y  thus  obtained  show  that  the 
stress  A  =  0  acts  on  a  surface-element  perpendicular  to  r  ;  and  that 
B  and  C  act  in  directions  perpendicular  to  the  radius  r,  and  making 
angles  of  45  degrees  with  the  2-axis.  B  acts  in  the  same  direction 
as  the  torsion,  C  in  the  opposite  direction. 

For  example,  considering  a  point  which  lies  in  the  surface  of  the 
cylinder  and  in  the  .s-plane,  and  setting  therefore  Y=0   and  A"  =  r, 
we  have       S  cos  a  =  0,     S  cos  /?  =  pkr  cos  y,     S  cos  y  =  pier  cos  p. 
When   S  =  Q   we  have   y  =  /3  =  ^7r;    when  S—  ±pkr  we  have   a  =  ^7r, 
cos  (3  =  cosy.     Since  cos2a  +  cos2/?  +  cos2y  =  1  ,  we  have  eos/?  =  ±s/|. 

The  moment  of  force  M  to  which  the  torsion  of  the  cylinder  is 
due  is 

M=  (f&r.  Zirrdr.r. 

The  upper  limit  R  of  the  integral  is  the  radius  of  the  cylinder, 
Integrating,  we  have  M=^TrfjJcR*  =  ir<f>ii.R4j2l,  where  I  is  the  length 
of  the  cylinder  and  </>  the  angle  of  torsion. 

The  factor  T  =  Trp.R*/2l  is  called  the  moment  of  torsion  of  the  cylinder. 
It  depends  only  on  the  dimensions  of  the  cylinder  and  the  constant 
of  elasticity  /x.  For  this  reason  //,  is  called  the  coefficient  of  torsion. 


SECT.  XXXIII. 


FLEXURE. 


87 


0. 


a 


SECTION  XXXIII.    FLEXURE. 

It  is  not  possible  to  give  a  rigorous  discussion  of  the  flexure  of 
a  prism.  We  will,  therefore,  confine  ourselves  to  an  approximate 
calculation  in  one  very  simple  case*. 

Let  ABCD  (Fig.  44)  be  the  prism  considered.     Its  length  is  supposed 
horizontal  and  coincident  with  the  axis  Ox.     The  axis  Oz  is  directed 
perpendicularly   upward,   and    the   axis    Oy  is    therefore    horizontal. 
After   flexure,  the   cross   section  AB  is 
displaced  to  A'£',  which  may  lie  in  the 
same  plane  as  AB.     Another  plane  cross 
section   FG,   also   perpendicular    to    the 
axis,  is  displaced  by  the  flexure  to  F'G' ; 
we  assume  that  the  section  F'G'  is  also 
plane,  and  that  the  plane  F'G'  cuts  the 
plane  A'B'  in  a  horizontal  line  passing     y 
through  P.     This  line  of  intersection  is 
supposed  to  be  common   to   the   planes 
of  all  sections  perpendicular  to  the  axis. 
The  parts  of  the  prism  which  originally 
lay  in  OQ  lie  after  the  flexure  in  OQ',  FlG  44 

which  we  will  consider  as  the  arc  of  a 

circle  whose  centre  is  P.  Such  a  flexure  is  called  circular.  All  the 
lines  in  the  prism  which  were  originally  parallel  to  the  x-axis  become 
circles,  whose  centres  lie  on  the  straight  line  passing  through  P. 

Represent  the  original  coordinates  of  a  point  M  in  the  section 
AB  by  0,  y,  z,  and  its  coordinates  after  flexure  by  0,  y  +  ^0,  2+f0* 
The  same  changes  occur  in  the  other  cross  sections,  for  example  in 
FG.  If  the  coordinates  of  a  point  M'  in  FG  are  originally  x,  y,  z, 
they  will  become  by  flexure  x  +  g,  y  +  t],  z  +  £  We  set  L.  OPQ'  =  <f>, 
OP  =  pa,nd  OQ  =  OQ'.  This  last  assumption  is  admissible,  since  there 
is  always  one  line  whose  length  does  not  change  by  flexure,  and 
since  we  have  as  yet  made  no  assumption  as  to  the  position  of  the 
z-axis.  We  therefore  obtain 


*  + 1  =  *  + to -(/>  +  *  + &)(!  -  cos  <£). 
If  p  is  very  great  in  comparison  with  x,  z,  and  £0,  we  may  set 

sin  <£  =  x/p ;     1  —  cos  <f>  —  a^/2p8 
and  obtain  (a)          £  =  xz/p,     n  =  ^     t=to-x2/2P- 

*  Barre  de  Saint- Venant,  Mem.  pres.  par  div.  Savants.     T.  14.    Paris,  1856. 


88  THE   THEORY   OF   ELASTICITY.  [CHAP.  n. 

We  may  so  determine  %  and  £0  that  all  the  components  of  stress 
except  Xx  vanish  ;  hence  we  may  write 
(1  )  X,  =  A6  +  2^/p  =  S,  (4)  Z,  = 

(2)  y,=Ae+2/u3v^=0>        (5)  xz 

(3)  Zt  =  \Q  +  2iidf0fdz  =  Q,  (6)  Y, 

Further,  we  have  Q  =  z/p  +  'dr] 

From  (2)  and  (3)  it  follows  that  (b) 

Comparing  (b)  with  XXIX.  (d)  it  appears  that  the  contraction  of  the 

cross  section  is  to  the  increase  in  length  in  the  ratio  of  |A.  to  A  +  /z, 

or  of  k  to  1.      Further,  since  T?O  and  £0  do  not  involve  x,  we  have 

from  (b)  rj0=  -  kyz/p  +/(*),     £0  =  -  kz*/2p  +  g(y), 

when  /  and  g  designate  two  unknown  functions.     From  (4)  we  have 

-ty/p+f'(z)  +  g'(y)  =  0,  and   hence  f\z)  =  c,  where  c  is   an   unknown 

constant.     It  follows  that 

f(z)  =  cz  +  c',     g(y)  =  kf/  2p-cy  +  c" 

and  T,O  =  -  kyz/p  +  cz  +  c',     £0  =  k(V2  ~  Z<2)/%P  ~W  +  c". 

At  the  point  0,  where  y  =  z  =  Q,  we  have  Vo  =  ^>  &  =  0>  and  hence 
c'  =  0  and  c"  =  0.  Since  the  prism  does  not  turn  about  the  ;r-axis 
during  flexure,  it  follows  that  for  y  =  Q,  ^  =  0  also,  and  consequently 
c  =  0.  We  obtain  therefore 


and  further,  from  (a), 

(d)  t  =  xzl*     ri=-  kyz/p,     t=k(y*- 

These  values  for  ^,  rj,  and  ^  satisfy  the  equations  XXX.  (c),  since 
by  hypothesis  X=Y=Z=0.  The  equations  1-6  show  that  the 
conditions  of  equilibrium  are  fulfilled.  From  (1)  and  (b)  we  get 
X,  --=S  =  (3X/A  +  2/*2)/(A.  +  p.)  .  zfp.  If  we  introduce  the  general  coefficient 
of  elasticity  E  [XXIX.  (d)],  we  have  (e)  S  =  Ez/P. 

The  resultant  E  of  the  forces  S  is  (f  )  R  =  Ejp  .  \z  .  dydz,  and  is  equal 
to  zero,  if  the  a;-axis  passes  through  the  centre  of  gravity  of  the 
prism.  If  we  assume  this  and  then  determine  the  moment  M  of 
the  forces  S  with  respect  to  a  horizontal  line  passing  through  the 
centre  of  gravity,  we  will  have  M  =  \Szdydz  =  Ejp  .  \zzdydz  =  EJjp, 
where  J  is  the  moment  of  inertia  of  the  cross  section.  In  order  to 
bend  the  prism  so  that  an  axis  passing  through  the  centre  of  gravity 
of  the  prism  becomes  a  circle  of  radius  p,  a  rotating  force  of  moment 
M  must  act  on  each  end  surface  ;  the  axes  of  the  rotating  forces 
are  perpendicular  to  the  plane  of  the  circle,  and  are  oppositely 
directed. 


SECT,  xxxin.]         MOTION  OF  AN   ELASTIC   BODY.  89 

The  cross  section  of  the  prism  is  noticeably  altered  by  the  flexure. 
Since  the  parts  on  the  convex  side  of  the  prism  are  extended,  and 
the  parts  on  the  concave  side  compressed,  the  former  tend  to  contract 
in  the  directions  of  the  y  and  .s-axes,  the 
latter  to  expand.  If,  for  example,  the 
cross  section  is  a  rectangle  ABCD,  as  in 
Fig.  45,  ABCD  takes  the  form  A'FC'D'. 
The  two  plane  surfaces  whose  projections 
are  represented  in  the  figure  by  AB  and  '  r"~ 

CD  are  transformed  into  surfaces  of  double 
curvature.  We  may  consider  A'B'  and 
C'D'  as  arcs  with  the  centre  E,  while  A'D' 
and  B'C'  are  straight  lines  which  intersect 
at  E.  The  lines  A' I?  and  B'C'  are  not 
changed  in  length,  AB  is  shortened  and  FlG-  45- 

CD  lengthened.      If  z  =  %BC,  it  follows  from  the  definition  of  k  [cf. 
XXIX.]  that 

A'B'  =  AB(\  -  kz/p),     C'D'  =  CD(l  +  kz/p). 
UOE  =  p,  then 

A'B' /C'D'  =  (p'  -  z)I(p  +z)  =  (l-  kz/P)/(l  +  kz/p), 

from  which  it  follows  that  p  =  kp.     This  relation  has  been  applied 
to  the  determination  of  k  for  glass  prisms. 


SECTION  XXXIV.    EQUATIONS  OF  MOTION  OF  AN  ELASTIC  BODY. 

The  resultant  with  which  the  elastic  forces  act  on  an  infinitely 
small  volume-element  dv  of  an  elastic  body  in  the  direction  of  the 
3-axis  is  [XXV.],  (dXx/^x  +  'dXy!dy  +  'dXipz)dv.  If  the  body  is  acted 
on  besides  by  attractions  or  repulsions,  whose  component  in  the 
direction  of  the  z-axis  is  X,  the  element  dv  is  also  acted  on  by  the 
component  of  force  X.dv.p,  where  p  is  the  density  of  the  body. 
Hence  the  .r-component  of  the  acting  forces  is 

(dXjVx  +  VXJVy  +  VXfiz  +  pX)dv. 

If  this  resultant  is  not  equal  to  zero,  motion  occurs  in  the  direction 
of  the  .T-axis,  and  the  momentum  imparted  to  the  part  of  the  body 
under  consideration  in  unit  time  is  pdvd-(x  +  g)/dt?  —  pdvd'2g/dt2,  where  t 
denotes  the  time.  Hence  we  have 

=  VXJdx  +  VXJciy  +  "dXJdz  +  PX. 


90  THE   THEOEY   OF   ELASTICITY.  [CHAP.  ir. 

If  the   components   of  stress   are   expressed   by   £,    ry,  and    £  as   in 
XXX.  (b),  we  obtain  the  equation 

(a)  p£  =  (*  +  /*)•  39/3*  +  )"V2£  +  pX 

The  equations  for  V)  and  £  are  similar.  « 

As  in  XXX.  (e)  the  equations  (a)  take  the  form 

(b)  p£=  (A  +  2/*) .  36/3x  +  2/i(3A,/3«  -  3/*2/3y)  +  /oA'. 

If  the  force  whose  components  are  AT,  F,  and  Z  has  a  potential,  and 
if  therefore      X=  -3*/3a:,     F=  -3¥/3y,     2T  =  -3¥/3*, 
by  differentiation  of  equation  (b)  with  respect  to  #,  y,  2  respectively, 
and  by  addition,  we  have 

(c)  Pe  =  (A  +  2/*)V2e-/>V2¥. 

In  what  follows  we  assume  that  no  external  forces  act,  so  that 
the  components  X,  Y,  Z  are  zero.  Therefore  V2^  drops  out  of 
equation  (c). 


SECTION  XXXV.    PLANE  WAVES  IN  AN  INFINITELY 
EXTENDED  BODY. 

Lame*  treated  this  form  of  motion  in  the  following  way.  Suppose 
a  plane  wave  propagated  in  a  direction  which  makes  the  angles  a,  J3,  y 
with  the  axes;  let  the  velocity  of  propagation  be  V,  and  let  the 
direction  of  vibration  make  with  the  axes  the  angles  a,  b,  c.  If  u 
represents  the  distance  of  a  point  from  its  position  of  equilibrium, 
U  the  amplitude,  and  T  the  period  of  vibration,  the  vibration  at  the 
origin  may  be  expressed  by  u=  Ucos(2trt/T).  At  any  other  point, 
whose  coordinates  are  x,  y,  z,  we  have 


(a)  «=  tfcosW. 


We  have  further 

(b)  £  =  u  cos  a,     rj  =  u  cos  b,     £  =  ti  cos  c. 

If  the  angle  between  the  direction  of  propagation  and  the  direction 
of  vibration  is  represented  by  <£,  we  have 

cos  <f>  =  cos  a  cos  a  +  cos  bcos/3  +  cos  c  cos  y. 
For  brevity  we  set 


-  tfsin  jar/ 


**,    t  - 


*  Lame,  Theorie  de  1'elasticit^,  p.   138.     Paris,  1866. 


SECT,  xxxv.]  PLANE   WAVES.  91 

and  obtain 

9  =  2KS/T  V.  cos  </>,     36/ar  -  -  ^2u/Tz  V-  .cos  a.  cos  & 

V2£  =  -  47r2M/r2  V1  .  cos  a,     £  =  -  47r2M/r2  .  cos  a. 
By  the  help  of  these  relations  and  corresponding  ones  for  t\  and 
£  we  obtain  from  XXXIV.  (a) 

r  (A  +  p.)  cos  a  cos  <£  +  (p.  -  pV2)  cos  a  =  0, 


(c) 

I  (A  +  /*)  cos  y  cos  <£  +  (/x  -  pV2)  cos  c  =  0. 
If  these  equations  are  multiplied  by  cos  a,  cos/2,  cosy  respectively 
and  then  added,  we  have  (X  +  2/X-/3/72)  cos  <£  =  0.  We  therefore  have 
either  (d)  (e)  pV'2  =  \  +  2fj.  or  cos<£  =  0.  In  the  first  case,  equations  (c) 
become 

cos  a  =  cos  a  cos  <£,  cos  b  =  cos  (3  cos  <£,  cos  c  =  cos  7  cos  <£. 
If  the  right  and  left  sides  of  these  equations  are  squared  and  added, 
we  obtain  (f)  cos2(/>=l,  so  that  either  <£  =  0  or  <£  =  TT.  The  vibrations 
therefore  occur  in  the  direction  of  propagation  ;  they  are  called 
longitudinal  vibrations.  In  the  second  case  <£  =  JTT,  that  is,  the  vibra- 
tions are  perpendicular  to  the  direction  of  propagation  ;  they  are 
called  transverse  vibrations. 

Longitudinal  Vibrations.  —  The  velocity  of  propagation  12  of  these 
vibrations  is  determined  by  (d),  (g)  tt  =  <J(\  +  '2/j.)lp.  Hence  con- 
densations and  rarefactions  occur,  since 


To  determine  the  stresses  we  assume  that  the  waves  are  propagated 
in  the  direction  of  one  of  the  coordinate  axes,  say  the  0-axis.  In  this 
case  we  have 

£  =  0,     17  =  0,     C=Ucos{27r/T.(t-z/^)}. 

From  XXX.  (b)  the  tangential  forces  are  zero  ;  the  normal  forces  are 
(h)  Xx  =  Y,  =  27rX/7'I2  .  U  sin  {  2^/T  .  (t  -  g/Q)}. 

(i)  Zt  =  2v(\  +  2p)/TQ.  Usin{2Tr/T.(t-z/tt)}. 

Transverse  Vibrations.  —  The  velocity  of  propagation  w,  from  (c),  equals 
(k)  w  =  x//Vp.  Since,  for  these  vibrations,  we  have  cos<£  =  0,  we 
also  have  9  =  0,  that  is,  neither  condensations  nor  rarefactions  occur. 
If  the  wave  is  propagated  in  the  direction  of  the  2-axis,  and  if  the 
vibrations  are  parallel  to  the  .T-axis, 

£=  Ucos{2ir/T.(t-z/^)},     77  =  0,     £=0. 

All  components  of  stress  vanish  with  the  exception  of  the  tangential 
force  Za  (1)  Zx  =  2^/T^  .Usin{  2v/T.  (t  -  z/w)  }. 


•92  THE   THEORY   OF  ELASTICITY.  [CHAP.  n. 

In  a  solid,  therefore,  two  different  wave  motions  may  exist,  which 
^re  propagated  with  different  velocities  12  and  u>.  From  formulas 
(g)  and  (k)  the  velocity  12  of  the  longitudinal  vibrations  is  always 
greater  than  the  velocity  w  of  the  transverse  vibrations.  In  liquids 
and  gases  the  only  vibrations  which  can  occur  are  longitudinal,  since 
for  these  bodies  p.  =  0. 

For  gases  we  have  A.  =JPJXXIX.  (g)],  and  hence  the  velocity  of 
sound  in  air  is  (m)  12  =  *JP/p.  P  must  here  be  expressed  in  absolute 
units.  According  to  Regnault  the  density  of  atmospheric  air  at 
Paris  equals  0,0012932  under  a  pressure  of  76  cm.  of  mercury,  and 
At  a  temperature  of  0°C.  Since  the  acceleration  of  gravity  at  Paris 
is  980,94,  the  pressure  of  the  air  on  a  square  centimetre  equals 
76.13,596.980,94  in  absolute  units.  Hence  the  density  p  of  the 
air  under  a  pressure  P  in  absolute  units  is 
<n)  P  =  0,001  293  P/76  .  13,596.  980,94  =  P.  1,2759  .-lO'9. 

Using  this  value  of  P/p  we  obtain  12  =  27  996  cm,  or  approximately 
280  metres  per  second  at  0°C.  Since  the  density  of  the  air  at  f  C 
is  p  =  P.  1,2759.  10~9/(l+af)  the  velocity  of  sound  at  f  is 


where  a  is  the  coefficient  of  expansion  of  air  0,00  366.  The  result 
obtained  from  this  form  of  the  theory  does  not  agree  with  that  found 
by  observation.  Observation  shows  that  12  is  about  330.  The 
reason  why  theory  and  observation  are  not  in  accord  will  be  discussed 
later  in  the  theory  of  heat. 

The  velocity  of  sound  in  water  is  obtained  in  a  similiar  manner. 
For  water  at  15°  C.  we  have  A  =  2,22.  1010.  At  the  same  tempera- 
ture we  have  P  =  0,999  173,  whence  12  =149  060  cm. 

In  a  research  carried  out  on  the  Lake  of  Geneva,  Colladon  and 
Sturm  found  that  the  velocity  of  sound  in  water  at  8,1°C.  is  12  =  143  500 
centimetres;  the  difference  between  the  observed  and  calculated 
values  is  explained  by  the  difference  in  temperature,  since  A.  increases 
rapidly  for  water  as  the  temperature  rises.* 

No  observations  have  been  made  on  wave  motions  in  large  masses 
of  metal  ;  but  the  velocity  of  sound  has  been  determined  in  a  metallic 
wire.  In  such  a  body,  however,  sound  is  propagated  with  a  different 
velocity  from  that  which  it  would  have  in  an  extended  body.  If 
the  wire  is  parallel  to  the  z-axis,  and  if  we  consider  only  the  motion 
of  the  particles  in  the  direction  of  this  axis,  the  stress  Z,  at  the 
•distance  z  from  the  .r//-plane  is,  in  our  usual  notation,  Zt  = 
*  Fogliani  und  Vicentini,  Wied.  Beibl.  Bd.  8.  S.  794. 


SECT,  xxxv.]  WAVE  MOTIONS.  93 

At  the  distance  (z  +  dz)  the  stress  is 

Zt  +  dZJdz .  dz  =  E(dffd3  +  32£/322 .  dz). 

Hence  a  portion  of  the  wire  whose  length  is  dz  and  whose  cross 
section  is  A,  is  acted  on  by  a  force  given  by  AE&ydz2 .  dz.  The 
equation  of  motion  is  pA  .  dz.  ^=AE^i/dz2 .  dz  or  (o)  £=  F232t/9.s2, 
where  V=jE/p.  The  integral  of  the  differential  equation  (o)  is 
(p)  £=cos  {2ir/T.  (t-zlV)} ;  the  velocity  of  propagation  Vis  obtained 
from  equation  (p). 

According  to  the  researches  of  Wertheim  the  velocity  calculated 
from  (p)  agrees  fairly  well  with  the  results  of  observation. 


SECTION  XXXVI.     OTHER  WAVE  MOTIONS. 

Spherical  Waves. — We  will  investigate  the  circumstances  of  the 
propagation  of  spherical  waves  in  an  infinitely  extended  elastic  body, 
when  the  direction  of  vibration  of  every  particle  passes  through  the 
same  point.  We  take  this  point  as  the  origin  of  coordinates.  As 
in  XXXI.  (b)  we  set 

(a)  £  =  30/3a,     i7  =  30/3y,     £=30/3*, 

where  e£  is  an  unknown  function  of  t  and  of  the  distance  r  from  the 
origin.     The  equations  of  motion  [XXXIV.  (b)]  give  (b)  0  =  Q2V2<£. 

In  this  case  [XV.  (1)]  we  may  set  V2<£  =  !/?• .  32(r0)/3r2,  and  hence 

(c)  32(r<£)/3<2  =  fi232(r<£)/3r2.     This  equation  is  satisfied  by 

(d)  0  =  a/r.cos{27r/r.(*-r/fl)}, 
when  a  is  a  constant  and  T  the  period  of  vibration. 

The  distance  u  of  a  point  from  its  position  of  equilibrium  is 
u  =  30/3r  =  -  fl/r2 .  cos  { ^JT .  (t  -  r/fl)}  +  ^irafBr .  sin  {2-n-fT.  (t  -  r/Q)}, 
where  B  =  £lT.     If  r  is  very  much  greater  than  the  wave  length  we 
can  neglect  the  first  term  on  the  right,  and  have 

u  =  A/r.sin{-27r/T.(t-rltt)}. 

The  wave  motion  is  therefore  one  in  which  the  wave  surfaces  are 
spheres  propagated  with  the  velocit}^  12. 

Since  the  expressions  (a)  satisfy  the  equations  of  motion,  if  <f>  has 
the  value  given  in  (d),  these  equations  are  also  satisfied  if  </>  is  replaced 
by  30/3x,  or  by  another  differential  coefficient  taken  with  respect 
to  one  or  more  coordinates. 

Vibrations  Due  to  Torsion.— Let  the  axis  of  a  circular  cylinder 
coincide  with  the  2-axis,  and  its  separate  parts  oscillate  in  arcs 


94  THE   THEORY  OF   ELASTICITY.  [CHAP.  n. 

about  the  same  axis.  The  components  of  displacement  of  a  particle 
from  its  position  of  equilibrium  may  be  expressed  [XXXII.  (a)]  by 
{e)  £=  -<^y,  fi  =  <t>x,  £=0,  where  <£  is  a  function  of  z. 

From  XXVIII.  (e)  we  have  0  =  0;  therefore  condensation  and 
rarefaction  do  not  occur.  The  equations  of  motion  are  [XXXIV.  (a) 

and  XXXV.  (k)],  £  =  w2V2£,      77=0,^,, 

whence  we  again  obtain  (f)  <j>  =  w?'d2<t>/'dz'2.  This  equation  is  satisfied 
by  (g)  <£  =  asin  {27T/T.  (t-zjw)}.  Hence  w  =  v//x//j  is  the  velocity 
with  which  a  wave  motion  is  propagated  in  the  direction  of  the 
axis  of  the  cylinder.  From  XXX.  (b)  the  components  of  stress  are 

Zt  =  -  Apx,     Xz  =  +  AM  where  A  =  Zira/Tta  .  cos  {2ir/T.  (t  -  z/u)}. 

The  other  components  of  stress  are  zero. 

If  the  cylinder  is  of  finite  length,  stationary  waves  can  exist  in 
it,  that  is,  waves  such  that  certain  definite  points  of  the  cylinder 
called  nodal  points  are  at  rest,  while  on  both  sides  of  a  nodal  point 
the  vibrations  are  in  opposite  phase.  The  amplitude  of  the  vibration 
is  greatest  half  way  between  two  nodal  points,  at  the  ventral  segments. 
Stationary  waves  are  formed  when  waves  which  have  passed  over  a 
certain  point  return  to  that  point  again  in  the  opposite  direction. 
To  find  the  period  T  of  these  vibrations,  we  notice  that  equation  (f) 
will  be  satisfied  not  only  by  (g),  but  also  by  <}>  =  b  sin{2?r/T.  (t  +  z  /«>)}, 
and  in  general  by 

<£  =  B  sin  27rt/T.  cos  (ZirzjTu)  +  Ccos  (2irt/T)  .  sin  (2vz/T<o), 

where  B,  C,  and  T  are  constants.  If  the  points  for  which  ^  =  0  are 
fixed,  the  constant  B  will  be  zero  and  (i)  <j>=Ccos  (2vtjT)  .  sin  (2irz/Tw). 
If  I  represents  the  length  of  the  cylinder,  and  if  the  points  for  which 
.z  =  l  are  also  fixed,  we  will  have  <£  =  0  when  z  =  l,  and  therefore 
pTr,  where  p  is  a  whole  number.  Hence 


If,  on  the  other  hand,  one  end  of  the  rod  is  free,  Yz  =  X2  =  0  when  z  =  I. 
.Since  Xt  =  -  w  .  3<£/a?,     YZ=+IM.  3<f>/'dz, 

we  have  9<£/3.z  =  0  when  z  =  l. 

In  this  case  we  obtain  from  equation  (i)  2irl/T<a  =  ^(2p+  1).  TT,  where 

j?   is   a   whole   number;    and    hence    T=M/(2p+  1).  v//>//x.      If  both 

ends  of  the  rod  are  free,  T=2l/ 


SECT,  xxxvii.]  VIBRATING  STRINGS.  95 

SECTION  XXXVII.    VIBRATING  STRINGS. 

Although  the  problem  of  the  motion  of  vibrating  strings  is  only 
slightly  connected  with  the  theory  of  elasticity,  a  simple  example 
of  this  form  of  motion  will  be  considered  here.  We  suppose  a 
perfectly  flexible  string  stretched  between  two  fixed  points  A  and 
B.  If  P  is  the  stress  in  the  string,  /0  the  length  of  the  string  before 
the  application  of  the  stress,  and  /  its  length  while  the  stress  is 
applied,  p  the  cross  section  of  the  string,  E  the  coefficient  of  elasticity, 
we  have  1-10  =  P10/FE.  Let  the  string  be  slightly  moved  from  its 
position  of  equilibrium,  that  is,  the  straight  line  which  joins  A  and 
B,  and  let  the  new  form  of  the  string  be  designated  by  ACDB.  By 
this  deformation  the  length  of  the  string  is  increased  by  dl  =  dP  .  10'FE. 
It  is  here  assumed  that  dP  is  infinitely  small  in  comparison  with  P, 
so  that  we  may  set  the  stress  in  the  string  everywhere  equal  to  P. 

For  the  sake  of  simplicity  we  suppose  that  the  motion  of  the 
string  is  always  in  one  plane,  say  the  z#-plane.  Let  A  be  the  origin 
of  coordinates,  and  let  B  lie  on  the  a>axis  at  the  distance  I  from  A. 
The  distance  of  any  point  C  of  the  string  from  A  may  be  represented 
by  5,  and  that  of  the  infinitely  near  point  D  by  s  +  ds.  The  com- 
ponents of  stress  at  C  in  the  directions  of  the  x-  and  y-axis  respectively 
are  Pdx/'ds  and  P'dy/'ds.  For  the  point  D  the  similar  components  are 
PCdx/'ds  +  'd-x/^s2  .  ds)  and  P(byl^s  +  32y/3s2  .  ds). 

The  infinitely  short  portion  CD  of  the  string  is  therefore  acted 
on  by  the  force  P^x/cts-  .  ds  in  the  direction  of  the  z-axis,  and  by 
the  force  P&y/'ds2  .  ds  in  the  direction  of  the  y-axis.  If  the  string 
is  displaced  only  very  slightly  from  its  position  of  equilibrium,  we 
can  set  s  =  x;  the  z-component  then  vanishes  and  the  particles  of 
the  string  oscillate  perpendicularly  to  the  z-axis.  If  m  represents  the 
mass  of  unit  length  of  the  string,  the  equation  of  motion  is 


or,  if  we  set  ma?  =  P,  (a)  ij  =  a2d*y/3a:2.  The  integral  of  this  differential 
equation  is  (b)  y  =  AH  cos  (mrat/l)  .  sin  (mrx/l),  where  n  is  a  whole 
number.  When  2  =  0  and  x  =  l  we  have  y  =  0,  and  when  /  =  0, 

y  =  Ansin(mrx/l)  ; 
this  is  the  equation  of  a  sinusoid. 

If,  in  the  general  case,  the  form  of  the  string  when  t  =  0  is  given 
by  the  equation  y=f(x),  then 

f(x)  =  Al  sin  (TTX/I)  +  AZ  sin  (2irx/l)  +  A3  sin  (3vx/l)  +  ...  . 


96  THE   THEORY   OF  ELASTICITY.  [CHAP.  H. 

The  coefficients  Av  A2,  A3  ...  are  determined  in  the  following  way. 
Let  the  general  term  of  the  series  be  Ansinn<f>,  where  <t>  =  -xjl.  If 
both  sides  of  the  last  equation  are  multiplied  by  sin  «<£,  we  will  have 
/(/^>/7r)sinn^)  =  ^1  sin  <£sin?t<£  +  ^2sin2(£sin72<£  +  ...  +  Ansin-n<f>  +  .... 
If  this  equation  is  multiplied  by  dfa  and  integrated  between  the 
limits  0  and  IT,  we  will  have 


ff(l<f>/Tr)  sin  n<f>  .  d<f>  =  An  /  s 
For  if  in  and  n  are  different  numbers,  we  have 

~w.  »» 

/  sin  m<f>  sin  n<f>d(f>  =  \  I  (cos  (m  -  w)<£  -  cos  (m  -f  n)$)e?<£  =  0. 
But  when  they  are  equal,  /  sin%<£d<£  =  ITT.     Hence 
(c)  A  „  =  2/7T  .    /W«-)  sin  n<f>  d<f>  =  2  //  .    f(x)  sin  (n«c//)  .  rfz. 


If,  for  example,  the  string  is  so  displaced  from  its  position  of 
equilibrium  that  a  point  in  it  at  the  distance  p  from  the  end  A  is 
moved  through  the  distance  h  in  the  direction  of  the  y-axis,  we  have 
f(x)  =  hxjp  for  0  <  x  <p,  but  f(x)  =  h(l  -  x)j(l  -p)  for  p  <  x  <  /.  Hence 


..  .  . 

p  I  JP     l-p  I 

and  therefore 

An  =  2hF/p(l  -p)  .  (sin  ^\  /»V2. 
We  obtain  for  y, 

y  =  2a%/(a  -  l);r2  .  fl/12  .  sin  -  .  sin  ^  .  cos 

[_  a  /  / 

+  1/22.  sin  27r.  sin  ^-.cos^+ 

at  I 

where  a  =  Z/p.     If  the  string  is  struck  in  the  middle,  we  have  a  =  2  and 


SECTION  XXXVIII.     POTENTIAL  ENERGY  OF  AN  ELASTIC  BODY. 

When  an  elastic  body  changes  its  form  work  will  be  done.  This  is 
stored  up  in  the  body  as  potential  energy  if  the  body  is  perfectly 
elastic,  which  we  will  assume  to  be  the  case.  The  work  necessary 
to  bring  about  a  particular  change  is  equal  to  the  potential  energy 
gained  by  the  body,  and  may  be  determined  in  the  following  way. 


SECT,  xxxvni.]  VIBRATING  STRINGS.  97 

Let  A',  B1  and  C"  represent  the  principal  stresses  at  a  point  in  the 
body ;  about  this  point  construct  the  infinitely  small  parallelepiped, 
whose  edges  u,  v  and  ic  are  parallel  to  the  principal  stresses.  When 
the  stresses  are  applied,  the  edges  of  the  parallelepiped  are  extended, 
u  becoming  u(l+a'),  v  becoming  v(l+b'),  and  iv  becoming  iv(l+c'). 
From  XXIX.  (e)  we  have 

A'  =  A9'  +  2/iw',     B"  =  A6'  +  2/aJ',     C'  =  A9'  +  2/xc'. 

If  a',  b'  and  c'  change  by  the  increments  da',  db'  and  dc  respectively, 
the  edges  of  the  parallelepiped  are  increased  by  uda,  vdb'  and  wdc', 
and  the  parallelepiped  undergoes  an  infinitely  small  change  of  form. 
The  forces  which  act  in  the  directions  of  the  edges  are  vwA',  uwB' 
and  uvC".  Hence  the  work  done  by  the  stresses  during  the  change 
of  form  is 

(A'da  +  B'dV  +  C'dc')uvw  =  (  A6W  +  ^(a'da  +  b'db'  +  c'dc'))uvw, 
since  0'  =  a'  +  b'  +  c'. 

To  change  the  form  of  the  parallelepiped  by  an  amount  which  is 
determined  by  the  elongations  a,  b,  c,  the  work 

4(  A62  +  2/n(a2  +  b2  +  c2))uvw 

must  be  done.  If  we  set  dv  for  the  volume  uvw  of  the  parallelepiped, 
the  potential  energy  Ep  of  the  whole  body  is  given  by 

(a)  ^  =  |J(Ae2  +  2/>i(a2  +  &2  +  c2))cfo. 

If  we  introduce  the  principal  stresses  A,  B,  C  [XXIX.  (e)],  we  have 

(b)  Ep  =  ^{(A  +  £  +  C)2/E-(AS  +  £C+CA)lfi}dv. 

By  this  equation  (b)  the  potential  energy  is  determined,  the  com- 
ponents of  stress  and  of  elongation  are  known.  We  confine  ourselves 
to  the  statement  of  the  following  relation, 

(c)  Er  =  i  Jfe  +  Yjy,  +  Z:zz  +  *Zjs,  +  2A>,  +  2YjJdr, 

from  which  the  others  can  easily  be  deduced. 

Galileo  was  the  first  to  study  the  properties  of  elastic  bodies;  he 
failed,  however,  to  reach  correct  results.  The  physical  basis  for  the 
theory  of  elasticity  was  given  by  Eobert  Hooke,  Avho  in  1678  published 
a  treatise,  De  potential  restitutira,  in  which  he  showed  by  experiment 
that  the  changes  of  form  of  an  elastic  body  are  proportional  to  the 
forces  applied  to  it.  Among  earlier  investigations  those  of  Mariotte 
and  Coulomb  deserve  especial  mention.  More  recently  the  theory 
of  elasticity  has  been  developed  principally  by  the  French  mathe- 
maticians, Cauchy,  Poisson,  Lame,  Barr6  de  Saint-Venant,  and  others. 


98  THE  THEORY  OF  ELASTICITY.  [CH.  u.  SECT,  xxxvin. 

We  owe  to  Cauchy  the  theory  of  the  components  of  stress  in  the 
form  here  given.  For  more  extended  accounts  of  the  theory  of 
elasticity  we  may  mention  Lame,  Theorie  Mathe'matique  de  V  Elasticity 
des  Carps  Bolides.  Paris,  1866.  Clebsch,  Theorie  der  Elasticitdt  fester 
Korper.  Leipzig,  1862.  Among  the  more  important  recent  treatises 
on  the  theory  of  elasticity  we  mention :  Boussinesq,  Application  des 
Potentiel  a  I'Etude  de  I'Equilibre  et  du  mouvement  des  Solides  Elastiques. 
Paris,  1885.  Barr6  de  Saint- Venant,  Me'moire  sur  la  Torsion  des  Prismes. 
M6m.  d.  sav.  Mr.  T.  XIV.  Paris,  1856  ;  Mtmoire  sur  la  Flexion  des 
Prismes.  Liouville  I.,  1856.  William  Thomson,  Elements  of  a 
Mathem.  Theory  of  Elasticity.  Phil.  Tr.  London,  1856;  Dynamical 
Problems  on  Elastic  Spheroids.  Phil.  Tr.  London,  1864. 

Further  researches  on  the  theory  of  elasticity  have  been  carried 
out  in  recent  years  by  W.  Voigt. 


CHAPTER   III. 

EQUILIBRIUM   OF   FLUIDS. 

SECTION  XXXIX.    CONDITIONS  OF  EQUILIBRIUM. 

THE  principal  difference  between  solids  on  the  one  hand  and  liquids 
and  gases  on  the  other  consists  in  the  fact  that  the  latter  do  not, 
like  the  former,  offer  a  great  resistance  to  change  of  form.  A  force 
is  always  needed  to  change  the  form  of  a  fluid  mass,  but  the  resistance 
offered  by  the  fluid  is  determined  by  the  rate  at  which  the  change 
of  form  proceeds,  and  will  be  infinitely  small  if  it  proceeds  very 
slowly.  We  assume  that  the  motion  by  which  the  condition  of 
equilibrium  is  attained  proceeds  very  slowly,  and  we  may  therefore 
assume,  in  hydrostatics,  that  a  fluid  offers  no  resistance  to  change  of 
form,  so  long  as  this  does  not  involve  change  of  volume. 

Each  infinitely  small  change  of  form  of  an  infinitely  small  part  of  the 
body  may  [XXVIII.]  be  treated  as  if  it  were  produced  by  the  dilatations 
«,  b,  c  in  three  directions  perpendicular  to  each  other.  The  lengths 
u,  v,  w  drawn  in  these  three  directions  become  u(\  +  a),  v(l  +b),  w(\  +c). 
If  A,  B,  C  are  the  corresponding  normal  forces  per  unit  of  surface 
which  act  on  the  surfaces  vw,  uw,  iw  respectively,  the  work  done 
by  the  normal  forces  in  this  change  of  form  is 

Avwua 4- Buwvb  +  Cuvwc  or  (Aa  +  Bb  +  Cc)u .v.w. 

The  change  of  form  considered  will,  in  general,  involve  an  increase 
of  volume,  given  by  uvw(\  +  a}(\  +b)(l  +c)  -  uvw.  Since  a,  b,  c  are 
infinitely  small,  the  increment  of  the  volume  equals  (a  +  b  +  c)u.  v.  w, 
if  we  neglect  infinitely  small  quantities  of  a  higher  order. 

If  we  start  from  the  assumption  that  the  work  done  by  the  forces 
equals  zero  if  the  volume  is  not  changed,  we  have  at  the  same 
time  Aa  +  Bb  +  C'c  =  Q  and  a  +  b  +  c  =  0.  These  equations  can  both  be 
true  only  if  A  =  B  =  C. 


100 


EQUILIBRIUM  OF   FLUIDS. 


[CHAP.  in. 


The  equations  for  the  components  of  stress  [XXVI.  (i)]  give 

Xt=Y,  =  Z,  and  Z,  =  0,  X,  =  0,  Yx  =  0. 
There  are,  therefore,  no  tangential  forces  in  a  fluid  in  equilibrium. 

If  we  start  from  the  condition  that  the  only  forces  which  act  on 
fluids  in  equilibrium  are  perpendicular  to  their  surfaces,  we  reach  the 
same   result,  namely,  that   the   normal   stresses   are   all   equal.      To 
show  this  we  set        Zy  =  Q,     A'2  =  0,     1^  =  0, 
and  have,  from  XXVI.  (a), 

P=Xxcosa,     Q=Y,cos/3,     fi  =  Z;Cosy. 

P,  Q  and  R  are  the  components  of  stress  for  a  surface  whose  normal 
makes  the  angles  a,  (3,  y  with  the  axes.  The  stress  acting  on  this 
surface  is  <JP-  +  Q-  +  R\  the  normal  force  N  is  determined  by 

N=  Pcosa  +  Qcosfi  +  R  cos  y. 
The  tangential  force  T  is 

T-  =  (P2  +  Q*  +  R2}  -  (P  cos  a  +  Q  cos  P  +  R  cos  y)2. 
Introducing  the  values  of  P,  Q  and  R,  we  have 
(X,  -  F,)2  cos2a  cos2j3  +  (Yy-  Zy-  cos2£  cos'y  +  (Z2  -  Xx)*  cos2«  cos2y  =  0, 
and  hence  Xx=Yy  =  Zz. 

From  the  expression  given  above  for  N,  it  follows  that  N=XX, 
that  is,  the  normal  force  acting  on  a  surface-element  in  the  interior  of 
the  fluid  is  independent  of  the  position  of  the  element. 

If  we  neglect  the  force  of  cohesion  of  the  fluid,  which  will  be 
treated  later,  the  normal  force  will  be  a  pressure  ;  if  this  is  designated 
by  p,  we  have 

(a)  X,=  Y,  =  Z,=  -p;    Zv  =  0,     X,  =  Q,     Yx  =  0. 

If  p  is  the  density  of  the  fluid,  we  obtain  from  XXV.  (c)  the  condi- 
tions of  equilibrium 

(b)  "dp/fa  =  pX,  dp/'dy  =  pY,  oppz  =  PZ. 
The  components  of  the  force  acting  on 
the  unit  of  mass  are  A",  Y,  Z;  we  may 
consider  p  as  constant  in  liquids ;  in  gases 
p  is  a  function  of  the  pressure. 

Equations  (b)  may  also  be  developed  in 
/C ~B'  the  following  way.     Represent  the  sides 

of  the  parallelepiped  00'  (Fig.  46)  by 

OA=dx,     OB  =  dy,     OC  =  dz. 

The  pressure  on  OA'  ispdydz,  the  pressure  on  O'A  is  (p  +  "dp/dx .  dx)dydz. 
The  resultant  of  the  pressures  is  the  pressure  -  "dp/*dx .dx.dy.dz  in 


SECT,  xxxix.]  EQUILIBRIUM  OF  FLUIDS.  101 

the  direction  of  the  a-axis.  The  force  pXdxdydz  also  acts  on  the 
parallelepiped  in  the  same  direction.  The  condition  of  equilibrium 
is  therefore  (  -  'dp/'dx  +  pX)dxdydz  =  0,  from  which  we  obtain  the  first 
of  equations  (b). 

If  equilibrium  obtains,  p  must  satisfy  equations  (b)  ;  the  conditions 
for  this  are 
( 
1 

These  equations  will  hold  if  a  function  4>  exists,  such  that 
(d)  -d3>px  =  pX,     -d3>py  =  pY,     ?&fiz  =  f>Z. 

Equations  (c)  are  the  essential  conditions  of  equilibrium  ;   if  they 
are  satisfied,  p  may  be  determined  from  the  equation 

dp  =  p(Xdx  +  Ydy  +  Zdz). 
If  the  forces  have  a  potential  i/s,  so  that 


we  will  have  (e)  dp=  -  pd^. 

In  gases  p  is  a  function  of  p  •  in  liquids  p  may  be  considered 
constant.  In  the  latter  case  we  obtain  (f)  p  =  c  -  pif>,  where  c  is 
constant. 


SECTION  XL.    EXAMPLES  OF  THE  EQUILIBRIUM  OF  FLUIDS. 

The  conditions  of  equilibrium  of  a  liquid  mass  contained  in  a 
vessel,  and  acted  on  by  gravity  only,  may  be  determined  in  the 
following  way  :  Suppose  the  position  of  the  particles  of  the  liquid 
referred  to  a  system  of  rectangular  coordinates,  whose  2-axis  is  directed 
perpendicularly  upward ;  we  then  have 

X  =  0,     F-0,     Z=-g, 

and  therefore  $  =  gz.  Since  the  density  p  is  considered  constant, 
equilibrium  can  obtain  under  the  action  of  gravity.  From  XXXIX.  (f) 
we  have  p  =  c-gpz.  Hence  the  pressure  at  the  same  level  is  every- 
where the  same. 

We  now  determine  the  pressure  in  a  liquid  contained  in  a  vessel,  which 
rotates  about  a  perpendicular  axis  A  with  constant  angular  velocity  w. 
The  fluid  will  turn,  like  a  solid,  about  the  axis  A  with  the  same 
angular  velocity  as  that  of  the  vessel. 

A  particle  at  the  distance  r  from  the  axis  A  is  acted  on  both  by 
gravity  and  by  a  centrifugal  force  whose  acceleration  is  w2r.  We 


102  EQUILIBEIUM   OF  FLUIDS.     [CHAP.  HI.  SECT.  XL. 

refer  it  to  a  system  of  rectangular  coordinates  whose  2-axis  is  directed 
perpendicularly  upward,  and  coincides  with  the  axis  of  rotation. 
We  then  have 

X=tfx,     Y=^y,     Z=-g, 

and  the  potential  ^  is  ^=  -%<azr-  +  gz.  From  XXXIX.  (f)  the  pres- 
sure is  p  =  c  +  p( Jw'V2  -  gz).  The  surfaces  of  constant  pressure  are 
paraboloids  of  revolution  with  the  common  axis  A. 

We  will  make  a  third  application  of  the  conditions  of  equilibrium 
to  the  determination  of  the  pressure  in  the  atmosphere.  We  suppose 
gravity  directed  toward  the  centre  of  the  earth,  its  acceleration  y 
may  then  be  expressed  by  7=  -  ga?/f2,  where  g  is  the  acceleration 
at  the  surface  of  the  earth,  a  the  earth's  radius,  and  r  the  distance 
of  the  point  considered  from  the  centre  of  the  earth.  We  then  have 
\}/=  -ga2/r.  If  the  temperature  is  constant  p  =  k.p,  where  k  is  con- 
stant. We  have  then  [XXXIX.  (e)] 

dp=  -k.pdif*  or  log  p  =  c  -  k^. 

If,  at  the  earth's  surface,  the  pressure  is  p0,  and  the  potential  «/-0,  we 
have  logpQ  =  c-k$0  and  Iog(pjp)  =  k(\f'-^0)  =  kg(ar-a2)/r.  If  the 
difference  r-a  is  very  small  in  comparison  with  a,  we  can  set 
Iog(p0/p)  =  kgh,  where  h  =  r  —  a  is  the  height  of  the  point  considered 
above  the  earth's  surface,  and  k  is  equal  to  1,2759  .  10~9  for  dry  air 
at  0°C. 


CHAPTER   IV. 

MOTION   OF   FLUIDS. 

SECTION  XLI.    EULER'S  EQUATIONS  or  MOTION. 

IN  the  study  of  the  motion  of  fluids  very  serious  difficulties  are 
encountered,  and  thus  far  only  a  few  problems  have  been  completely 
solved.  In  the  following  chapter  we  will  consider  only  the  so-called 
ideal  fluids,  and  therefore  neglect  the  friction  between  their  moving 
particles  and  the  forces  of  adhesion  and  cohesion,  which  will  be 
treated  later.  We  further  assume  that  the  fluids  considered  are  in- 
compressible, and  thus  limit  our  discussion  to  liquids,  which  are  only 
slightly  compressible.  On  these  assumptions  several  characteristics 
of  the  motion  of  liquids  may  be  derived.  As  the  study  of  the  motion 
of  gases  is  extremely  difficult,  and  as  little  success  has  so  far  been 
obtained  in  it,  we  will  not  enter  upon  it  here. 

For  a  complete  determination  of  the  motion  of  a  fluid  the  path  of 
each  separate  particle,  as  well  as  the  position  of  the  particle  in  the 
path  at  any  instant,  may  be  given.  The  coordinates  x,  y,  z  of  the 
particle  M  may  be  given  as  functions  of  the  time  t. 

An   easier   method   is   one   in   which   the   motion  is  expressed  in 
terms  of  the  components  of  velocity,  which  according  to  circumstances 
shall  be  designated  by   U,   V,   W,  or  u,  v,  w.     Suppose   U,  F,  W  to 
be  the  components  of  velocity  of  a  particular  particle  of  the  fluid  ; 
if  the  particle  at  the  time  t  is  situated  at  P  and  at  the  time  /  +  dt  at  P', 
U,  V,  W  will  be  the  components  of  velocity  at  the  point  P,  and 
TT    dUlt         r    dVlt      ,„    d,W,. 
U+lltdt>      V+lTtdt>     W+-^ 

the  components  of  velocity  at  P.  The  quantities  dU,  dF,  dW  repre- 
sent the  increments  which  U,  V,  W  respectively  receive  during  the 
time  dt,  if  our  attention  is  confined  to  the  motion  of  a  particular 
particle. 

103 


104  MOTION   OF   FLUIDS.  [CHAP.  iv. 

The  other  symbols,  u,  v,  iv,  represent  the  components  of  velocity 
at  a  definite  point  in  space,  where  one  particle  replaces  another  in 
the  course  of  the  motion.  If  x,  y,  z  are  the  coordinates  of  the  point 
considered,  u,  v,  w  are  the  components  of  velocity  of  a  particle  situated 
at  that  point  at  the  time  t.  After  the  lapse  of  the  time  dt  the  same 
point  is  occupied  by  another  particle,  whose  components  of  velocity 
are  u  +  du/'dt.dt,  v  +  'dv/'dt.dt,  w  +  'dw/^t.dt. 

A  particle  situated,  at  the  time  t,  at  a  point  whose  coordinates  are 
x  +  dz,  y  +  dy,  z  +  dz,  has  a  velocity  whose  projection  on  the  ,r-axis  is 

u  +  'du/'dx .  dx  +  "dufdy .  dy  +  "dufdz .  dz. 

The  velocities  u,  v,  w  are  everywhere  functions  of  x,  y,  z  and  /.     If 
u,  v,  w  are  the  components  of  velocity,  at  the  time  /,  at  the  point  P, 
whose  coordinates  are  x,  y,  z,  the  components  at  the  time  t  +  dt  at 
another  point  P,  whose  coordinates  are  x  +  dx,  y  +  dy,  z  +  dz,  will  be 
u  +  'du/'dt .  dt  +  'du/'dx .  dx  +  "dufdy .  dy  +  'du/'dz .  dz,  etc. 

If  the  fluid  particle  is  situated  at  the  time  t  at  P  and  at  the  time 
t  +  dt  at  P',  then  we  have  U=u  and 

U+  d  U/dt  .dt  =  u+  oupt .  dt  +  'duj'dx .  dx  +  ^fdy .  dy  +  'du/'dz .  dz, 
or  (a)  d  U/dt  =  Vu/?)t  +  Vu/'dx .  dx/dt  +  'du/'dy .  dyfdt  +  'dufdz .  dz/dt. 

The  particle  considered  traverses  the  distance  PP'  in  the  time  dt, 
hence  its  velocity  is  PF/dt,  the  projections  of  which  on  the  coordinate 
axis  are  evidently  dxfdt  =  u,  dy/dt  =  v,  dz/dt  =  w.  Thus  we  obtain 

d  U/dt  =  'du/'dt  +  u .  du/'dx  +  v .  'du/'dy  +  w .  'du/'dz. 
The  equations  for  dVjdt  and  dlVjdt  are  similar. 

To  find  the  equations  of  motion  of  a  fluid  let  us  cut  from  it  a 
parallelepiped  dw,  whose  edges  are  dx,  dy,  dz,  and  on  which  a  force 
acts  whose  components  are  X,  Y,  Z.  In  the  time  dt  the  parallelepiped 
receives  an  increase  of  momentum,  whose  components  are 

pXdudt,     pYdwdt,     pZdwdt, 
when  p  denotes  the  density  of  the  fluid. 

The  pressure  p  acting  at  the  point  .r,  y,  z,  as  has  been  shown  in 
a  former  chapter,  imparts  to  dto  the  components  of  momentum 

-  'dp/'dx'.  d<adt,     -  'dp/'dy .  dwdt,      -  ty/'dz .  dwdt. 

Under  the  action  of  these  forces  the  body  receives,  in  unit  time,  an 
increment  of  velocity  whose  components  are  dU/dt,  dVjdt,  dWjdt, 
and  hence  we  have 

(b) 


SECT.  XLL]  MOTION  OF   FLUIDS.  105 

By  the  help  of  equations  (a)  we  then  obtain 

f  "duj'dt  +  udul'dx  +  v'dufiy  +  wdttfiz  =  X-  l/p.'dp/'dx, 

(c)  i.'dv  fit  +  u*v  fix  +  vdv  fiy  +  wdv  fiz=Y-l/p.  Vpfiy, 

\  "divfit  +  tidwfiz  +  vdwfiy  +  wdwj'dz  =  Z  -\jp.  *dpfiz. 

These  equations  are  due  to  Euler,  and  are  known  as  Euler's  Equations 
of  Motion.  To  them  must  be  added  the  so-called  equation  of  continuity, 
which  is  found  in  the  following  way :  The  parallelepiped  receives  in 
the  time  dt,  through  the  face  dydz,  the  quantity  of  fluid  pudydzdt ;  and 
loses,  through  the  opposite  face,  the  quantity  (pu  +  'd(pu)j'dx.dx)dydzdt. 
The  difference  between  the  quantities  flowing  through  the  two  surfaces, 
which  indicates  a  loss  of  fluid,  if  ~d(pu)fix .  dx  is  positive,  will  be 
3(ptt)/3a; .  do> .  dt. 

By  a  similar  argument  applied  to  the  two  other  pairs  of  faces  it 
appears  that  the  total  difference  between  the  quantities  of  fluid  which 
leave  and  enter  the  parallelepiped  is 

(d(pu)fix  +  ^(pv)fiy  +  -d(pw)fiz)d<» .  dt. 

The  parallelepiped  at  first  contained  the  quantity  p .  do> ;  after  the 
lapse  of  the  time  dt  it  contains  the  quantity  (p  +  ~dpfit .  d£)d<a ;  the 
difference  of  these  two  quantities  is  -~dpfit.dw.dt.  By  equating 
these  two  expressions  for  the  same  quantity  we  get  the  equation  of 
continuity  (d)  "dp fit  +  ^pujfix  +  ?>(pv)fiy  +  ^(pw^fiz  =  0.  If  the  density 
p  of  the  fluid  is  constant,  the  equation  of  continuity  becomes 

(e)  'dU'/'dx  +  "do  fiy  +  Vwfiz  =  0. 

Euler's  equations  are  specially  suited  to  investigations  of  the  motion 
in  fluid  masses  with  fixed  boundaries.  If  the  surface  of  the  fluid 
changes  there  will  be  points  which  will  lie  sometimes  within  and 
sometimes  without  the  fluid ;  the  velocity  at  such  a  point  cannot  be 
determined  by  the  method  here  given.  Lagrange's  method  is  the  one 
then  employed.  To  this  we  will  return  later. 

In  equations  (c)  and  (e)  there  are  contained  four  unknown  quan- 
tities u,  v,  10  and  p,  for  whose  determination  we  have  four  equations 
given.  To  determine  the  constants  of  integration  the  conditions  of 
the  motion  of  the  fluid  must  be  given  at  a  definite  time.  If  the 
fluid  is  bounded  by  a  fixed  surface,  the  components  of  velocity  in 
the  direction  of  the  normals  to  the  bounding  surface  are  zero.  If 
u,  v,  w  are  the  components  of  velocity  of  a  particle  at  the  boundary  of 
the  fluid,  and  if  the  normal  to  the  bounding  surface  makes  the  angles 
a,  ft,  7  with  the  axes,  we  have 

(f )  u  cos  a  +  v  cos  ft  +  w  cos  7  =  0. 


106  MOTION   OF   FLUIDS.  [CHAP.  iv. 

SECTION  XLIL    TRANSFORMATION  OF  EULER'S  EQUATIONS. 

In  k  fluid  in  motion  an  elementary  parallelepiped,  whose  edges 
are  originally  dx,  dy,  dz,  not  only  changes  its  position  in  space  but 
may  also  rotate  and  change  its  form  at  the  same  time.  Its  motion 
at  any  instant  is  determined  by  the  components  of  velocity  u,  r,  v:  ; 
the  rotations  and  changes  of  form  may  be  determined  in  the  following 
way  :  In  the  theory  of  elasticity  the  component  of  rotation  hx  of  such 
an  element  is  expressed  by  hx  =  ^(dC'/'dy  -  ty'/'dz),  if  £'  and  rf  are  the 
infinitely  small  changes  of  the  coordinates  z  and  y  introduced  by  the 
motion.  We  may  set  £'  =  iv.dt,  rj'  =  v.dt,  and  obtain 


If  £  is  the  corresponding  angular  velocity,  we  will  have  hx  =  %.  dt 
and  hence 


(a)  £  =  £(3tp/By-a0/az);  77  =  $ 
The  equations  for  t]  and  £  may  be  derived  in  the  same  way  as  the 
first;  £,  »/,  £  are  the  components  of  angular  velocity  in  a  rotation 
about  the  three  coordinate  axes. 

If  no  rotation  exists  in  the  fluid,  we  have  £  =  y  =  {=Q  or 


These  equations  are  the  conditions  for  the  existence  of  a  function  4> 
of  x,  y,  z  and  t  which  has  the  property  that 

it  =  -  3^/Da;,     v  =  -  3<£/9y,     tr  =  -  "d^fdz. 

This  function  </>  is  called  by  v.  Helmholtz  the  velocity  potential,  since 
the  components  of  velocity  U,  v,  w,  are  related  to  each  other  in  the 
same  way  as  the  components  of  a  force  if  it  has  a  potential. 

The  equation  of  continuity  [XLI.  (e)],  on  the  assumption  that  a 
velocity  potential  exists  and  that  the  fluid  is  incompressible,  becomes 
*  =  V2<t>  =  Q.     The  velocity  h  of  a  particle  is 


From  equations  a  it  follows  that  'du['dy  =  'dv/'dx  -  2£  'du/'dz  =  'dtv/'dx  +  2r). 
The  first  of  equations  XLI.  (c)  becomes 

dw/3*  +  2(ui)  -  vQ  +  u  .  'du/'dx  +  v  .  dr/da:  4-  w  .  'dwj'dx  =  X-l/p.  'dpjox. 
This  equation  may  be  written 

f  3tt  fit  +  2(v»)  -  rfl  =  X  -  1  Ip  . 

\  We  have  similarly 

I  -dv  fit  +  2«  -w£)=Y-\lp. 


SECT.  XLII.]  VORTEX   MOTIONS.  107 

where  h  is  the  velocity  of  a  particle.  "We  may  eliminate  p  from 
equations  (b)  by  differentiating  the  second  of  those  equations  with 
respect  to  z,  and  the  third  with  respect  to  y  and  subtracting.  We 
thus  obtain 


If  we  use  the  equation  of  continuity  'du/'dx  +  'dv/'dy  +  'dw/'dz  =  0,  and 
the  relation  following  from  (a)  3£/3z  +  cfy/3?/  +  3£/?z  =  0,  we  obtain 
3£/3<  +  u  .  3£/3z  +  v  .  9£/3y  +  w  .  m'dz  -  £  .  'du/ox  -  rj  .  'duf'dy  - 


If  &  7/'  £  represent  the  components  of  rotation  at  a  point  in  the 
region  containing  the  fluid  at  the  time  /,  we  may  use  &,  H,  Z  to 
represent  the  components  of  rotation  of  a  particle  at  the  time  t  +  dt, 
whose  components  at  the  time  t  were  £,  -rj,  £. 

The  connection  between  the  components  £,  77,  £  and  &,  H,  Z  may 
be  established  in  the  same  way  as  that  previously  used  to  find  the 
relation  between  the  velocity  at  a  point  in  space  and  the  velocity 
of  a  particle  of  the  fluid.  We  have 

£  =  S,  dS/dt  =  3£/3rf  + 
Using  this  equation  we  obtain 
(c)  dS/dt  =  £  .  'du/'dx  +  rj  .  -duj-dy  +  £.  'du/'dz  +  tfdZ/'dy  - 

If  at  any  time  no  rotation  exists  in  the  fluid,  and  if  therefore  £  =  y  =  £=0 
at  any  point  in  the  fluid,  a  rotation  may  still  be  set  up  if  Z  and  Y  have 
no  potential.  If,  on  the  other  hand,  Z  and  Y  have  a  potential  so 
that  Z=  -W/oz  and  F=-3¥/3y,  we  will  have  dSjdt  =  0.  If 
besides  X=  -o¥/'dx,  we  have  dU/dt  =  Q  and  dZ/dt  =  Q.  Hence  no 
rotation  can  be  set  up  in  an  ideal  fluid  if  the  fwces  have  a  potential  In 
this  case,  the  particles  which  rotate  already  continue  to  rotate,  but  the  particles 
which  do  not  rotate  from  the  beginning  will  never  rotate.  This  theorem 
was  first  given  by  v.  Helmholtz. 


SECTION  XLIII.    VORTEX  MOTIONS  AND  CURRENTS  IN  A  FLUID. 

In  researches  on  the  motion  of  fluids  it  is  important  to  observe 
whether  the  particles  rotate  or  not.  If  there  is  rotation  it  is  called 
cortex  motion.  We  then  have 


(a)    £  =  $ 

From  this  it  follows  at  once  that  (b)  3£/3z  +  3^/3y  +  3f/3^  =  0.     The 

equation  of  continuity  is  (c)  'du/'dx  +  'dv/'dy  +  'dw/'dz  =  0. 


108  MOTION   OF  FLUIDS.  [CHAP.  iv. 

If  the  forces  have  a  potential  it  follows  from  equations  XLIL  (c)  that 
C  dS/dt  =  £  .  3n  /  cte  +  17  .  'dw  I'dy  +  £  . 


(d) 

{  dZ  fdt  =  £  .  Sw/dx  +  77  .  dw?/3#  +  £  .  'dwj'dz. 


In  these  equations  £,  77,  £  are  the  components  of  rotation  at  the 
point  x,  y,  z;  S,  H,  Z  are  the  same  components  for  a  particle  which 
at  the  .time  t  is  situated  at  the  point  x,  y,  z,  but  which  at  the  time 
t  +  dt  is  situated  at  the  point  x  +  dx,  y  +  dy,  z  +  dz. 

On  the  other  hand,  if  the  components  £,  77,  £  are  zero  at  every  point 
in  the  fluid  at  a  definite  instant  they  are  equal  to  zero  at  any  time, 
from  equations  (d).  In  this  case  we  call  the  motion  a  flow.  It  is 
characterized  by  the  equations 

(e)  'dw/'dy  =  Zv/'dz,     ?m/dz  =  'dw/'dx,     Vvj'dx  =  'du/'dy. 

From  XLII.  u,  v  and  w  then  have  a  velocity  potential  <f>,  which  depends 
in  general  on  x,  y,  z  and  t.  The  equation  of  continuity  is 

(g)  32<£/ar2  +  a^/a^8  +  a^/a?2  =  v-<£  =  o. 

Euler's  equations  XLI.  (c)  take  the  form 


r  X  =  - 
J  Y  =  - 


(h)  Y  =  -  ^pfdy  +  ^W/Vy  +  1  fp  . 


Hence  such  a  motion  can  exist  only  when  the  forces  have  a  poten- 
tial ¥.  In  case  this  condition  holds,  we  obtain  from  (h)  by  integration 
(i)  ¥  +  T=  'd^/'dt  -  %h?  -p/p,  where  T  is  a  function  of  the  time  only. 

In  order  to  have  a  simple  example  of  the  two  classes  of  motion 
just  described,  we  consider  the  case  of  an  infinite  fluid  mass,  all 
particles  of  which  move  in  circles  parallel  to  the  zy-plane  whose 
centres  lie  on  the  z-axis.  All  particles  at  the  same  distance  from  the 
.z-axis  move  with  the  same  velocity  and  in  the  same  sense.  We 
have  then  from  XXXVI.  (e)  u=  -wy,  v=  +ux,  w  =  0.  w  depends 
only  on  the  distance  r  of  the  particle  from  the  z-axis.  Since  we  have 
'du/'dx  =  -  xy/r  .  dw/dr  and  'dvj'dy  =  +  xy/r  .  dw/dr.  the  equation  of  con- 
tinuity is  satisfied,  because  'du/'dx  +  "dv/^y  —  0. 

In  general  there  is  rotation  of  the  separate  particles,  since 
'duj'dy  =  -  (a  -  y2/r  .  dw/dr  ;     'dv/'dx  =  w  +  x2/r  .  dw/dr, 

and  therefore  £=<»  +  £?•  .do>/dr.  Since  £  =  0,  77  =  0  and  u,  r,  w  are 
independent  of  z,  the  equations  of  motion  (d)  are  satisfied. 

We  assume  f  =  £0  for  r  <r0  and  £=0  for  r  >  r0,  where  £0  is  a  constant. 
In  the  first  case,  we  have  w  =  f0  +  Cjr2,  where  C  is  a  new  constant. 


SECT.  XLIII.]  STEADY  MOTION.  109 

C  must  vanish,  because  otherwise  the  particles  at  the  axis  would  have 
an  infinitely  great  velocity.  Hence  o>  =  £0  for  the  part  of  the  fluid 
lying  within  a  circular  cylinder  whose  radius  is  ?-(),  and  whose  axis 
coincides  with  the  2-axis.  These  fluid  particles  therefore  rotate  about 
the  £-axis,  just  as  if  they  formed  a  solid  body.  If,  on  the  other  hand, 
r>r0,  and  hence  {=0,  the  angular  velocity  w'  will  be  ot'=C'/rz.  The 
linear  velocity  is  rw'  or  C'/r,  and  therefore  inversely  proportional  to  the 
distance  of  the  particle  from  the  axis.  On  the  condition  that  there  is 
no  discontinuity  in  the  motion  of  the  fluid,  we  have  for  r  =  r0,  £0  =  C'/r02. 
Hence  for  r  >  r0,  we  have  rw'  =  r02f0/r.  If  r0  is  infinitely  small  and  £0 
infinitely  great,  we  obtain  a  so-called  vortex  filament. 

The  action  of  the  vortex  filament  on  the  surrounding  fluid  depends 
on  its  cross-section  and  its  angular  velocity.  If  we  set  m  =  irr0^0, 
the  velocity  h  of  a  fluid  particle  which  does  not  belong  to  the  vortex 
is  h  =  rw  =  m/Trr. 

Vortex  filaments  may  have  other  forms ;  they  were  first  investigated 
by  v.  Helmholtz,*  and  afterwards  by  William  Thomson,  and  several 
others.  We  see  from  this  example,  that  the  separate  parts  of  the 
fluid  do  not  need  to  turn  about  themselves  as  their  centres  of  gravity 
describe  circles :  although  the  fluid  surrounding  the  vortex  filament 
revolves  about  the  2-axis,  the  separate  drops,  into  which  the  mass 
may  be  divided,  do  not  rotate  about  themselves. 


SECTION  XLIV.    STEADY  MOTION  WITH  VELOCITY-POTENTIAL. 

If  the  components  of  velocity  are  independent  of  the  time,  or  if 
the  condition  of  motion  at  any  definite  point  in  the  fluid  does  not 
change,  the  motion  is  called  steady.  If  a  velocity-potential  <£  exists, 
we  have  (a)  u=  —  'd<f>/'dz,  v—  —  c)(£/3y,  w=  -'dtfij'dz,  where  <f>  is  a 
function  of  x,  y,  z  only.  The  same  holds  for  the  potential  M*1  of  the 
forces ;  hence  the  function  T  in  XLIII.  (i)  must  be  constant.  If  we 
set  T=  -C,  we  have  (b)  V+p/p  +  $h2  =  C.  If  the  only  forces  which 
act  are  pressures  within  the  fluid,  we  may  set  ^  =  0  and  conclude 
that  the  velocity  of  the  particles  increases  as  they  pass  from  places 
of  higher  pressure  to  places  of  lower  pressure,  and  inversely. 

For  a  motion  for  which  there  is  a  velocity-potential,  the  equation 
of  continuity  is  (c)  V2<£  =  0. 

As  an  example  of  such  a  motion,  we  will  consider  a  sphere  at 
rest  in  an  infinitely  extended  fluid.  The  particles  of  the  fluid  which 
*  Helmholtz,  Crelle's  Journal,  Bel.  55,  S.  25,  1858. 


110  MOTION   OF   FLUIDS.  [CHAP.  iv. 

are  at  a  great   distance  from  the   centre  of  the   sphere  move  with 
equal  velocities  in  the  same  direction. 

Let  the  sphere  be  placed  so  that  its  centre  is  at  the  origin  of 
coordinates  0,  and  let  the  radius  of  the  sphere  be  R.  The  particles 
whose  distance  r  from  0  is  infinite  are  supposed  to  move  in  a 
direction  parallel  to  the  positive  2-axis  with  the  velocity  w0.  We 
set  the  velocity-potential  (d)  $=V-  u-^z,  in  Avhich  V  =  0  when 
r  =  oo.  Then  (e)  u  =  -'dF/'dx,  v=  -'dF/'dy,  10  =  -'dPrjdz  +  tc0.  Using 
equation  (c)  we  have  (f)  V2F=0.  If  we  set  P=l/r,  or  equal  to  a 
differential  coefficient  of  1/r  taken  with  respect  to  x,  y,  or  z,  the 
equation  (f)  will  be  satisfied.  Since  the  arrangement  around  the 
.r-axis  is  symmetrical,  we  will  consider  if  the  assumption 


will  satisfy  the  given  conditions. 

The  particles  of  the  fluid  move  over  the  surface  of  the  sphere, 
and  hence  the  component  of  velocity  in  the  direction  of  the  radius 
is  equal  to  zero,  that  is  (3</>/<3r)r=Jj  =  0.  If  we  set  z/r  =  cosy,  we 
have 

</>=-(?  cos  y/r2  -  ra'0  cos  y  and  d<f>/dr  =  2C  cos  y/r3  -  w?0  cos  y. 
From  this  it  follows  that  (h)  C=±wQR*. 

From  equations  (d),  (g),  and  (h),  it  follows  that 


Using  equations  (e)  we  obtain 

u  =  -  %wQR3zx/r>,  v  =  -  f  w0R?zyli*t    w=  -  ^R^Sz^/r5  -  1/r3)  +  zr0. 

If  we  set  tt,-2  +  t>2  =  s2  and  x2  +  y2  =  q-,  we  will  have  s=  -^w^qz/r5. 
If  q  and  z  are  the  coordinates  of  the  path  of  a  particle,  the  equation 
of  the  path  will  be  (k)  dq/dz  =  s/w. 

Remembering  that  r2  =  q2  +  z2,  we  may  integrate  equation  (k)  and 
obtain  q~(\  -R?/r3)  =  c.  If  c  is  constant,  this  is  the  equation  of  a 
stream  line.  If  c  =  0,  we  will  have  either  r  =  R  or  <?  =  0;  in  the  first 
case,  we  get  the  equation  of  a  great  circle,  in  the  second,  the  equa- 
tion of  the  z-axis. 

The  pressure  p  may  be  determined  by  the  help  of  equation  (b). 
Since  ^P  =  0,  we  have  p  =  p(C-  ^h'2).  Now  h2  =  u2  +  v*  +  w2,  and  hence 
for  a  point  on  the  surface  of  the  sphere  we  have  h  =  %w(>q/E. 

Hence  the  pressure  p  on  the  part  of  the  sphere  which  lies  toward 
the  positive  side  of  the  0-axis,  is  as  great  as  that  on  that  part  of 
the  sphere  which  lies  on  the  negative  side  of  the  2-axis  ;  the  moving 
mass  of  fluid  will  therefore  impart  no  motion  to  the  sphere.  And 


SECT.  XLIV.]  LAGRANGE'S   EQUATIONS.  Ill 

further,  a  sphere  which  moves  with  constant  velocity  in  any  direc- 
tion in  an  infinite  mass  of  fluid  experiences  no  resistance  during  its 
motion.  This  result,  which  is  at  first  sight  so  startling,  is  explained 
by  the  fact  that  the  resistance  offered  by  friction  is  not  taken  into 
account. 


SECTION  XLV.     LAGRANGE'S  EQUATIONS  OF  MOTION. 

Suppose  a  particle  P  of  a  fluid  to  be  originally  situated  at  the 
point  whose  coordinates  are  a,  b,  c,  and  after  the  lapse  of  the  time 
dt,  to  have  reached  the  point  x,  y,  z.  The  general  coordinates  x,  y,  z, 
are  functions  of  t,  a,  b,  c;  if  t  alone  varies  in  these  functions,  we 
obtain  the  path  of  a  particular  particle.  If,  on  the  other  hand,  we 
give  to  the  coordinates  a,  b,  c,  all  possible  values,  and  keep  t  constant, 
we  have  the  positions  of  all  the  particles  of  the  fluid  at  the  same 
time.  If  the  pressure  is  designated  by  p,  and  the  density  of  the 
fluid  by  p,  and  if  we  set  U=x,  V=y,  JF=z,  we  obtain  from 
XLI.  (b) 

(a)       x  =  X-l/P.dp/dx,     ij=Y -Ijp.dpldy,     z  =  Z-l/p.dp/dz. 
In  order  to  eliminate  the  differential  coefficients  with  respect  to 
x,  y,  z,  we  multiply  these   equations   respectively  by   dx/da,  dy/da, 
dz/da,  by  dx/db,  dy/db,  dz/db,  and  finally  by  dx/dc,  dy/dc,  dz/dc. 
By  addition  we  then  get  the  following  equations : 

f  (x  -  X) .  dx/da  +  (y-Y).  dy/da  +  (z-Z).  dz/'da  +l/p.  dp/da  =  0, 
(b)    -   (x-X).  dx/db  +  (y-Y).  dy/'db  +  (z-Z).  dzjdb  +  l/p .  dp/db  =  0, 
(  (x  -  X] .  dx/dc  +  0/-Y).  dy/dc  +  (z-Z).  dz/'dc  +  l(p.  dp/dc  =  0, 
These  equations  are  due  to  Lagrange. 

To  these  equations  there  must  be  added  a  relation  which  expresses 
the  fact  that  the  volume  of  the  fluid  does  not  change.  The  particles 
originally  situated  in  a  rectangular  parallelepiped  with  the  edges 
da,  db,  dc,  are  at  the  time  t  contained  in  a  parallelepiped,  the  pro- 
jections of  whose  edges  are 

dx/da .  da,  dy/da .  da,  dz/da .  da ; 
dx/db  .  db,  dy/db .  db,  dz/db  .  db  ; 
dx/dc .  dc,  dy/dc  .  dc,  dz/dc  .  dc. 

The  volume  of  the  parallelepiped  at  the  time  t  will  therefore  be 
dx/da,     dy/da,     dz/da 


dx/db,     dy/db,     dz/db 
dx/dc,     dy/dc,     dz/dc 


.dadbdc. 


112  MOTION   OF   FLUIDS.  [CHAP.  iv. 

Since   the   fluid   is    assumed    incompressible,    the    equation    of  con- 
tinuity is 

'dxj'da,     'dy/da,     'dz/'da 
(c)  -dxj-db,     -dy/cib,     -dz/Zb    =1. 

'dx/'dc,     'dy/'dc,     'dzj'dc 

To   apply   Lagrange's    equations,    we   will   consider   a   fluid   mass, 
which   turns  with   a   constant   angular   velocity  <u  about   the   z-axis, 
directed   vertically  downward.     We  then   have   X=0,    F=0,    Z  =  g, 
and  we  set  z  =  c,    a  =  r  cos  (f>,    b  =  r  sin  </>,    and  further, 
x  =  r  cos  (<f>  +  (at)  =  a  cos  ut  -b  sin  W, 
y  =  r  sin  (<£  +  W)  =  b  cos  tat  +  a  sin  wt. 
From  these  relations  it  follows  that 

dx/'da  =  cos  (at,    ~dx/db  =  -  sin  (at,    'dx/'dc  =  0, 
~dy/da  =  sin  (at,    ~dy/db  =  cos  (at,        'dy/'dc  =  0  ; 


i1  =  —  o>2a;,     y——  (a2y,    z  =  0. 

The  equation  of  continuity  (c)  is  satisfied  and  the  equations  of 
motion  (b)  are  'dp/'da  =  /ooj%,  'dp/'db  =  p(a2b,  'dp/'dc  =  gp.  Hence  we 
obtain  by  integration  p=  C  +  p(iw2(a-  +  J2)+^c).  This  solution  agrees 
with  that  given  in  XL. 


SECTION  XLVI.    WAVE  MOTIONS. 

Lagrange's  equations  may  be  used  to  advantage  in  investigations 
on  wave  motions  in  a  fluid  acted  on  by  gravity.  All  the  particles 
of  the  fluid  may  be  assumed  to  move  in  plane  curves  parallel  to 
the  x^-plane ;  let  the  a;-axis  be  horizontal,  and  the  £-axis  be  directed 
perpendicularly  downward.  Then,  if  we  set  y  =  b,  we  obtain 

'dx/'db  =  0,   9y/3a  =  0,    dy/3&  =  l,   3y/3c  =  0,   "dz[db  =  0   and   y  =  0. 

If  we  further  set  p  =  pP,  the  equations  of  motion  given  in  XLV. 
(b)  become 

,.  (  x.  'dx/da  +  (z-g).  'dz/'da  +  ?>P/da  =  0, 

t  x .  dx/'dc  +  (z-g).  dz/dc  +  'dPj'dc  =  0, 
and  the  equation  of  continuity  XLV.  (c)  takes  the  form 
(b)  'dx/'da .  'dz/dc  -  'dzj'da .  "dx/dc  =  1 . 

Suppose  the  particle  B  (Fig.  47),  having,  while  in  its  position  of 
equilibrium,  the  coordinates  OA=a  and  AB  =  c,  to  move  in  a  circle 


SECT.  XLVI.] 


WAVE   MOTIONS. 


113 


DFE,  whose  centre  is  at  C.  Let  D  be  the  position  of  the  particle, 
represent  the  angle  between  CD  and  the  perpendicular  CE  by  0, 
and  set  BC=s,  CD  =  r.  We  then  have 

(c)  x 


Here  m  and  n  are  constants  and  r  and  s  are  functions  of  c.     We 
therefore  obtain 

"dzfda  =  1  +  nr  cos  0,    'dz/'da  =  -  nr  sin  0,   'dxj'dc  =  'dr/'dc .  sin  0, 

"dz/dc  =  1  +  'dsj'dc  +  9r/3c .  cos  0. 
By  these  relations,  equation  (b)  takes  the  form 

3s/3e  +  nr .  'dr/'dc  +  {nr(l  +  3s/3c)  +  3r/3c}  cos  0  =  0. 
Since  this  equation  holds  for  all  values  of  t  or  0,  we  have 
(d)  3s/9e  +  nr.3r/dc  =  0  and  3r/3c  +  »r(l+a*/3c)  =  0. 

We  obtain  further  from  equations  (a)  the  relations 
(e) 

-mh:'drl'dc- 
the  last  of  which  is  transformed  by  the  help  of  equations  (d)  into 

(f )  -  (w2  -  gnfrCdrfdc  +  ( 1  +  ^s/'dc)  cos  0  }  -  g  +  'dP/'dc  =  0. 

If  the  pressure  depends  on  c  only,  it  follows  from  (e)  and  (f)  that 

(g)  (h)  m2  =  gn,   P  =  gc,  if  the  constant  is  set  equal  to  zero;  the  pres- 
sure therefore  disappears  for  c  =  0.     This  condition  must  hold  at  the 
free  surface  of  the  fluid. 

The  paths  of  the  particles  are  circles.  If  the  time  required  by  the 
particle  to  traverse  its  path  is  T,  that  is,  if  T  is  the  period  of  oscilla- 
tion, we  have  m  =  27r/77  and  6  =  2ir/T. 


114  MOTION  OF   FLUIDS.         [CHAP.  iv.  SECT.  XLVI. 

If  X  is  the  wave  length  and  h  the  velocity  of  the  wave,  we  will 
have  h  =  Tg/2Tr  and  h  =  \/T,  from  which  it  follows  that 
(i)  h  =  *Jg\j'2ir  and  n  =  2ir/X. 

The  motion  of  the  particle  is  such  that  it  describes  a  circle  whose 
centre  lies  a  little  above  the  position  of  equilibrium  of  the  particle. 
From  the  first  of  equations  (d)  we  have  s=  -^nr2,  where  the  con- 
stant disappears,  since  s  and  r  vanish  simultaneously.  Hence  also 
(k)  s=  -7rr2/A.  From  the  second  of  equations  (d)  it  follows  that 
d  log  r  +  nd(c  +  s)  =  0,  and  hence,  by  integration,  log  r  +  n(c  +  s)  =  k,  when 
k  is  a  constant.  For  a  particle  on  the  surface  we  have  c  =  0  ;  if  the 
values  of  r  and  s  for  this  particle  are  designated  by  E  and  S,  we 
have  log B  +  nS=k.  We  have  further  log (r/E)  +  n(c  +  s-S)  =  Q.  The 
factor  c  +  s-S=H  is  the  perpendicular  distance  between  the  centre 
of  the  path  of  the  particle  considered  and  the  centre  of  the  path  of 
a  particle  in  the  surface.  We  have  therefore  (1)  r  =  Ee~Zirai'x. 

If  ds/dc  is  eliminated  from  equations  (d)  we  obtain 

dr  +  nr(dc  -  nrdr)  =  0, 

and  by  integration  1  ,'n .  log  r  +  c  -  |nr2  =  Tc'.     For  the  particles  on  the 
surface  we  have  1/n.  log  E  -  ^nE2  =  k' .     Hence 
(m)  c  =  A/27T .  log  (E/r)  -  ir/A  .  (E2  -  r2). 

The  free  surface  can  be  thought  of  as  formed  by  the  rolling  of  a 
circular  cylinder  on  the  under  side  of  a  horizontal  surface  AB  (Fig.  48), 


FIG.  48. 


which  lies  at  the  height  OA  =  X/2ir  over  the  centres  of  the  paths 
which  the  particles  in  the  surface  describe.  The  free  surface  is  then 
represented  by  a  straight  line  whose  distance  from  the  axis  of  the 
cylinder  is  E, 


CHAPTER   V. 


INTERNAL   FRICTION. 

SECTION  XLVII.     INTERNAL  FORCES. 

IN  the  discussion  of  the  motion  of  fluids,  the  friction  among  the 
fluid  particles  has  not  been  considered.  Friction  is  excited  in  different 
degrees  between  the  particles  of  the  fluid  when  they  move  among 
themselves  at  different  rates.  In  consequence  of  friction  the  viscosity 
of  the  fluid  is  more  or  less  great.  We  will  try  to  determine  the 
friction  caused  by  the  motion  of  the  fluid. 

We  will  suppose  that  the  particles  of  a  fluid  mass  are  moving  in 
a  direction  parallel  to  the  ar-axis,  and  that  those  situated  at  the  same 
distance  from  the  ar^-plane  have  the  same  velocity.  The  velocity 
increases  in  proportion  to  the  distance  from  the  xy-pl&ne.  One  sheet 
of  the  fluid  glides  over  another  and  thereby  gives  rise  to  a  definite 
frictional  resistance,  which,  according  to  Newton,  may  be  assumed 
proportional  to  the  rate  of  change  of  velocity  with  respect  to  the 
distance  from  the  ^-plane,  so  that  du/dy  =  e.  The  friction  between 
two  contiguous  sheets  is  then  propor- 
tional to  the  difference  of  their  velocities, 
and  inversely  proportional  to  the  distance 
between  them.  We  therefore  set  the 
velocity  u  =  <uQ  +  ey.  Let  00'  (Fig.  49) 
be  a  part  of  the  fluid  mass ;  on  each 
unit  of  surface  of  O'B  there  acts,  in  the 
direction  O.r,  a  tangential  force 
(a)  T=  p. .  du/dy  =  /*«, 

where  //,  is  the  coefficient  of  friction.      A  FlG-  49- 

force  -  T  acts  on  .OB'  in  the  direction  Ox.  Further,  the  tangential 
forces  T  [XXV.  (d)]  must  act  on  O'A  and  OA',  of  which  the  one 
acting  on  O'A  is  in  the  direction  Oy,  and  the  one  acting  on  OA'  is 
in  the  direction  yO. 

115 


116 


INTERNAL   FRICTION. 


[CHAP.  v. 


If  the  fluid  mass  moves  in  the  direction  of  the  ^/-axis  with  a  velocity 
v  =  v0  +  e'x,  the  tangential  force  necessary  to  produce  this  motion 
is  T'  =  p  .  dv/dx.  If  both  motions  exist  simultaneously,  a  tangential 
force  Xy  acts  on  the  fluid,  such  that  we  have 

X,  =  T+T'  =  iJ.(dujdy  +  dv/dx). 

In  order  to  examine  the  physical  meaning  of  this  expression,  we 
will  consider  a  fluid  particle,  originally  situated  at  the  point  x,  y,  z, 
which  has  moved  through  an  infinitely  small  distance,  whose  pro- 
jections on  the  axes  are  £,  7;,  £  We  then  have  u  =  3£/9/,  fl=cty/^, 
and  X=p.'dl'dt('dgfiy  +  'dii/'dx)  =  2p.'dz,rdt.  This  expression  gives 
the  tangential  force  which  arises  from  motion  in  a  fluid  in  terms  of 
the  friction.  "We  may  therefore  set 


Zy  =  2[j.  .  "dzj'dt 


2/t  .  'dyj'dt 


v  fis), 
+  'div/'dx), 
/*(9t>  fix  +  *dufiy). 


"We  know  by  experiment  that  ft  is  independent  of  the  pressure.  The 
meaning  of  the  other  quantities  in  (b)  is  clear  without  explanation. 
By  the  help  of  the  formulas  given  in  (b),  we  can  determine  the 
tangential  forces  which  must  act  in  the  fluid  to  overcome  the  frictional 
resistances.  We  will  now  determine  the  magnitudes  of  the  normal 

forces  which  are  necessary  for 
G  d'  D  J)'  the  extension  of  a  viscous  fluid 
in  a  given  direction.  Let  the 
fluid  move  in  a  direction  parallel 
to  AB  (Fig.  50),  and  let  the  velo- 
city of  a  particle  situated  at  the 
distance  i/  from  this  line  be  equal 
to  u.  As  before,  we  may  set 

u  =  ?/o  +  *y- 

After  the  lapse  of  the  time  dty 

A  has  traversed  the  distance  u0dt, 

and  G  the  distance  (u0  +  fAC)dt. 

CC'  represents  the  motion  of  the 

point  C  relative  to  A,  and  we  have  CC'  =  e  .  AC  .  dt.     If  we  designate 
the  angle  GAG'  by  d<f>,  we  have  (c)  d<f>  =  f.dt. 

The  rectangle  EFGH,  described  in  the  rectangle  ABCD,  transforms 
into  the  parallelogram  EF'G'H',  and  we  determine  the  increments 
which  the  sides  EH  and  EF  receive  by  this  transformation.  Eepre- 
senting  the  angle  HEB  by  ^,  and  noticing  that  HH'  and  FF  are 


SECT.  XLVII.]  INTERNAL   FORCES.  H7 

parallel  to  AB,  we  have  EH'  =  EH  '+  HH1  'cos  &  EF'  =  EF-  FF'sin  +. 
We  further  have 


HH'  =  BH.  d<j>  =  EHsint.d<t>,    FF'  =  AF.  d<j>  =  EFcos  $  .  d<t>, 
and  hence  (EH'  -  EH)/EH=  sin  ^  cos  $d<j>  ; 

(£P"  -  EF)/EF  =  -  sin  ^  cos  ^ty. 

If  the  increment  of  length  per  unit-length  of  EH  is  designated  by  ds, 
we  have  (d)  ds  =  sin  ^  cos  \W<£  ;  ds  is  also  the  diminution  of  length 
per  unit-length  of  EF. 

To  bring  about  the  deformation  considered,  a  tangential  force  T 
must  act  on  ABCD,  which  is,  from  (a),  (e)  T=p.e.  This  force  acts 
on  the  surface  corresponding  to  CD  in  the  direction  CD,  and  on  that 
corresponding  to  AB  in  the  direction  BA  ;  on  the  other  two  surfaces 
the  forces  act  in  the  directions  CA  and  BD.  To  determine  the 
normal  force  N  acting  on  the  surface  EF,  we  'set  in  XXIV.  (a)  a  =  ^, 
/3  =  |TT  -  ^,  y  =  £TT,  and  Xy=Yx  =  T.  Since  all  the  other  components 
of  stress  are  equal  to  zero,  we  obtain 

(f)  N=Pcost  +  Qsinf  =  2TsiniI>cost. 

From  (c)  and  (d)  we  have  ds  =  sin  ^  cos  ^  .  e  .  dt,  and  from  (e)  and  (f  ) 
JV=2/xesin  ^cos  \j/.  Hence,  we  have  (g)  N=2p..ds/dL  The  stress 
acting  on  the  surface  EH  is  -N.  It  has  been  shown  that  a  unit  of 
length,  in  the  direction  EF,  is  increased  by  -  ds.  If  a  normal  stress 
were  to  act  on  the  surface  of  ABCD,  it  would  have  no  influence  on 
the  deformation;  but  a  normal  force  S  +  N  would  act  on  EF,  and  a 
normal  force  S-N  on  EH. 

If  the  normal  stresses  XM  Y^  Zt  act  on  a  rectangular  parallelepiped 
whose  edges  are  parallel  to  the  coordinate  axes,  when  the  fluid  is  in 
motion,  they  cause  deformations  and  a  change  of  volume.  If,  as  in 
the  theory  of  elasticity,  we  set  the  volume  dilatation  Q  =  xx  +  y,  +  za 
then  xx  -  JO  is  the  part  of  the  increase  in  the  direction  of  the  ic-axis 
which  is  here  considered.  Similarly,  we  set  3S  =  Xx  +  Yy  -f  Za  and 
Xx  -  S  is  the  part  of  the  normal  force  in  the  direction  of  the  a;-axis 
which  causes  the  deformation.  By  the  help  of  (g),  we  obtain 


If,  finally,  we  set  for  -  S  a  quantity  p,  which  may  be  considered  a 
pressure,  on  account  of  its  analogy  with  the  pressure  in  ideal  fluids 
and  gases,  and  remember  that  xx  =  'd£l'dx,  'dxt/'dt  =  'du/'dx,  etc.,  we  will 
have  (h)  Xx  =  -p  +  2/j.  .  'duf'dx  -  ^('du/'dx  +  dv/3y  +  "dwfdz).  Analogous 
expressions  hold  for  Yy  and  Zf 


118  INTERNAL   FRICTION.  [CHAP.  v. 

From  equation  (h)  the  dimensions  of  ju,  are  ML~lT~l.  The  co- 
efficient p.  has  been  determined  for  many  fluids  and  gases.  It  changes 
very  much  with  the  temperature.  The  following  values  hold  for  0°  C. : 

Water,  0,01775;   Alcohol,  0,01838;   Air,  0,000182. 


SECTION  XLVIII.    EQUATIONS  OF  MOTION  OF  A  Viscous  FLUID. 

We  will  now  present  the  equations  of  motion  of  a  fluid  exhibiting 
internal  friction.  From  XXV.  the  components  of  stress  act  on  the 
unit  of  volume  in  the  direction  of  the  x-axis  with  the  force 

(X)  =  -dXJ-dx  +  aX,/3y  +  *dXJVz. 

If  U  is  the  velocity  of  a  single  particle  of  the  fluid  in  the  direction 
of  the  z-axis,  then  pU=(X)  +pX,  where,  in  the  usual  notation,  A" 
denotes  the  component  of  force  in  the  direction  of  the  x-axis.  From 
XL  VII.  (b)  and  (h),  we  have 

(a)  pU=pX-  dpfdx  +  p-V2u  +  £/x  .  3(3tt/3z  +  dv/dy  +  'dw/'dz)/'dx. 

This  equation,  and  those  analogous  to  it,  which  hold  for  the  velocities 
V  and  W  in  the  directions  of  the  y-  and  £-axes,  are  due  to  Stokes.* 
They  hold  in  connection  with  the  equation  of  continuity 

(b)  3/0/3*  +  3{pw)/3a;  +  'd(pv)/Vy  +  3(pw')/^  =  0. 
We  assume  that  the  fluid  is  incompressible,  and  have 

(c)  'du/'dx  +  'dv/'dy  +  'dw/'dz  =  0. 

(  p(u+u.  'du/'dx  +  v  .  'duj'dy  +  w  .  'du/'dz)  =  pV2u  +  pX  -  "dpfdx, 

(d)  J  p(v  +  u  .  ?>v  fdx  +  v  .  'dv  fdy  +  w.d>v  fdz)  =  /*V2?  +pY-  3p/3y, 
{  p(w  +  u  .  'dw/'dx  +  v  .  'dw/'dy  +  w  .  'dw/'dz)  =  p.V2w  +  pZ  -  'dp/'dz. 

The  equations  are  simplified  if  the  motion  is  steady,  that  is,  if 
ft  =  0,  v  =  0,  w  =  0.  If  the  velocity  is  very  small,  the  terms  u  .  'du/'dx, 
v  .  'du/'dy,  etc.,  may  be  neglected  ;  we  then  have 

(e)  nV2u  +  PX-'dpl'dz  =  0,  fj.V2v  +  pY-'dp/'dy  =  0,  p&w  +  pZ  -  'dp/'dz  =  0. 
If  the  forces  have  a  potential  ¥,  (f)  V2p  +  pV2^  =  0.     If  we  introduce 
the  components  of  rotation 

£  =  ±(dw/dy  -  'dv/dz),     y  =  $(dufdz  -  dw/dx), 
and  if  the  forces  have  a  potential,  we  have  from  (e) 
(g)  V2£  =  0,     V2»/  =  0,     V2C=0. 

Further  we  have  (h) 


*  Stokes,  Cambridge  Phil.   Tr.,  Vol.  vin.,  p.  297,  1845. 


SECT.  XLVIII.]          MOTION   OF  A   VISCOUS   FLUID.  119 

With  reference  to  the  boundary  conditions,  it  is  assumed  that  the 
particles  of  the  fluid  which  are  in  contact  with  solid  boundaries  have 
no  relative  motion  with  respect  to  them ;  at  the  boundary  of  the  fluid 
we  have  therefore  w  =  0,  v  =  0,  w  =  0,  if  u,  v,  and  w  represent  the 
components  of  velocity  at  the  bounding  surface.  If  solids  are  present 
in  the  moving  fluid,  we  may  generally  assume  that  each  particle  in 
the  surface  of  the  solid  has  the  same  velocity  as  the  particle  of  the 
fluid  which  is  in  contact  with  it. 


SECTION  XLIX.    FLOW  THROUGH  A  TUBE  OF  CIRCULAR 
CROSS  SECTION. 

We  consider  a  viscous  fluid  moving  slowly  through  a  narrow  tube, 
which  is  set  horizontal,  so  that  gravity  does  not  influence  the 
motion.  Let  the  axis  of  the  tube  be  taken  as  the  z-axis,  and  suppose 
that  the  particles  of  the  fluid  move  parallel  to  it.  We  then  have 


Equations  XLVIII.  (e),  (c)  and  (f)  then  become 
(a)    9p/a»  =  0,    3p/3y  =  0,   ^w^'dp/'dz;     (b),  (c)  dw/'dz  =  0  ;   \72p  =  0. 

From  (c)  it  follows  that  (d)  d2p/dzz  =  Q  and  p=fa+p(),  where  /  and 
pQ  are  constant. 

It  follows  further  from  (a)  that  ju,y%=/.  Since  w  depends  on 
the  distance  r  of  the  particle  from  the  axis  of  the  tube,  we  have, 
since  r2  =  x1  +  y2, 

V2w  =  d2w/dr2  +  l/r  .  dw/dr. 

Hence  we  obtain  d^/dr^+l/r.dw/dr^f/fjL.     By  integration 
w  =  c  log  r  +fr2/4:fj.  +  WQ. 

Since  w  has  a  finite  value  for  r  =  0,  the  constant  c  must,  equal  0. 
Therefore  (e)  w  =  wl)+fr2/4/JL,  where  WQ  is  the  velocity  in  the  axis 
of  the  tube.  If  the  pressure  is  equal  to  p0  when  z  =  0  and  to  pl 
when  z  =  l,  we  have  from  (d)  f=(pi-p0)/l-  If  we  substitute  this 
value  of/  in  equation  (e),  we  have  w  =  w0-rz.  (p0-pi)/4fd-  For 
all  particles  of  the  fluid  which  are  in  contact  with  the  wall  of  the 
tube,  we  have  w  =  0.  Representing  by  R  the  radius  of  the  tube,  we 
will  therefore  have  0=»W0-^.(p0-j^)/4/tZ.  We  obtain  finally 


120  INTERNAL   FRICTION.         [CHAP.  v.  SECT.  XLIX. 

The  volume  m  of  the  fluid  which  flows  in  one  second  through  a 
cross-section  of  the  tube  is  given  by 

(f)  m  =  I  2irrdr .  w  =  7r(pQ  -p^W/Spl, 

that  is,  the  volume  of  the  fluid  is  directly  proportional  to  the  fourth  power 
of  the  radius  of  the  tube,  inversely  proportional  to  its  length,  and  inversely 
proportional  to  the  constant  /*. 

Poiseuille  was  the  first  who  investigated  the  flow  of  a  fluid  through 
narrow  tubes ;  he  was  led  to  results  which  agree  with  the  above 
formulas.* 

*  Among  recent  works  on  hydrodynamics  are  to  be  mentioned  :  Lamb,  Treatise 
on  the  Motion  of  Fluids.  Cambridge,  1879.  Auerbach,  Die  Theoretische  Hydro- 
dynamik.  Braunschweig,  1881. 


CHAPTER  VI. 


CAPILLARITY. 

SECTION  L.— SURFACE  ENERGY. 

THE  form  of  a  fluid  mass  on  which  no  external  forces  act  is  determined 
by  the  forces  with  which  its  particles  act  on  one  another.  If  the 
mass  is  very  great,  it  will  take  the  spherical  form,  in  consequence  of 
the  gravitational  attraction  of  its  parts ;  if,  on  the  other  hand,  the 
mass  is  small,  the  force  of  gravitation  between  the  particles  will 
have  no  perceptible  influence.  If  the  force  of  gravitation  can  be 
neglected,  the  force  of  cohesion,  which  acts  in  every  fluid  mass,  tends 
to  bring  it  into  the  same  spherical  form.  From  researches  which  have 
been  made  on  the  mode  of  action  of  this  force,  it  appears  that  it 
acts  only  between  particles  which  are  at  very  small  distances  from 
one  another.  The  law  of  its  dependence  on  the  distance  between  the 
particles  is  not  yet  known.  We  may  nevertheless  develop  the  laws 
of  capillarity  by  the  use  of  a  method  which  does  not  require  a 
knowledge  of  that  law. 

If  the  form  of  a  fluid  mass  is  originally  a  sphere,  work  must  be 
done  to  change  it  into  any  other  form.  If  the  fluid  offers  no  frictional 
resistance,  this  work  can  be  due  only  to  the  fluid  particles  situated 
in  or  near  the  surface ;  since  the  only  particles  which  can  act  on  a 
particle  at  a  greater  distance  from  the  surface  are  those  which 
immediately  surround  it;  these  either  remain  in  their  positions  or 
are  replaced  by  others  which  act  in  the  same  way  as  those  replaced. 
The  work  done  is  therefore  expended  in  adding  new  particles  to 
those  already  present  in  the  surface,  or,  what  is  the  same  thing,  in 
enlarging  the  surface.  To  increase  the  surface  S  of  the  fluid  by  the 
infinitely  small  quantity  dS,  the  work  CdS  is  necessary ;  C  is  constant 
and  may  be  called  the  capillary  constant. 

Two  bodies  in  general  meet  in  a  surface,  and  C  depends  on  the 
character  of  these  two  bodies.  In  the  case  of  a  falling  raindrop  the 

121 


122  CAPILLARITY.  [CHAP.  vi. 

two  bodies  in  contact  are  water  and  air.  At  the  surface  of  a  drop 
of  oil  which  floats  in  a  mixture  of  water  and  alcohol,  as  in  the 
well-known  experiment  of  Plateau,  two  liquids  are  in  contact.  Even 
when  a  fluid  is  in  contact  with  a  solid,  or  when  two  solids  are  in 
contact,  the  common  surface  possesses  a  definite  surface  energy. 

Let  the  capillary  constant  of  two  bodies  a  and  b  be  Cab,  and  let 
S  be  the  surface  in  which  the  two  bodies  meet.  The  potential  energy 
Ep  of  the  surface  S  is  (a)  Ep  =  Cab .  S.  Since  Cab  =  EP/S,  and  since 
the  dimensions  of  Ep  and  S  are  L2T~2M  and  L-  respectively,  the 
dimensions  of  the  capillary  constant  are  T~'2M. 

The  surface  of  the  fluid  is  under  a  definite  tension  somewhat 
analogous  to  that  of  an  elastic  membrane.  If  a  rectangle  DEFG 
is  described  in  a  plane  surface  of  a  fluid,  and  if  three  sides  of  it 
retain  their  positions  unchanged  while  the  fourth  side  FG,  along  with 
the  particles  of  the  fluid  present  in  it,  is  moved  through  the  distance 
FH  in  the  direction  EF,  the  surface  is  enlarged  by  the  area  FG .  FH, 
and  the  surface  energy  is  increased  by  C .  FG .  FH,  where  C  is  the 
capillary  constant.  To  produce  the  motion  considered,  a  force  K  must 
act  on  FG ;  the  work  done  is  therefore  equal  to  K .  FG .  FH.  It 
follows  that  (b)  K=C,  or  the  tension  per  unit-length  of  the  fluid  surface 
is  numerically  equal  to  the  capillary  constant. 

This  tension  existing  in  the  surface  exerts  a  pressure  in  the  fluid. 
Let  P  (Fig.  51)  be  a  point  in  the  surface  which  is  supposed  convex 
in  the  neighbourhood  of  P.  Suppose  two  plane 
sections  erected  at  P,  which  contain  the  normal 
to  the  surface  at  that  point.  One  of  these 
planes  cuts  the  surface  in  the  curve  PA,  the 
other  in  the  curve  PR  PA  and  PB  shall 
intersect  at  right  angles  and  their  radii  of  curva- 
ture shall  be  the  principal  radii  of  curvature  of 
the  surface  at  the  point  P.  A  third  plane 
containing  the  normal  to  P  cuts  the  surface 
in  the  curve  PF,  whose  radius  of  curvature  R  is  determined  from 
Euler's  theorem  by  the  equation  (c)  l/E  =  cos2<f>/fil  +  sin2(f)/R2,  where  (f> 
is  the  angle  between  PA  and  PF.  About  the  point  P  as  a  centre  we 
suppose  described  a  sphere  of  infinitely  small  radius  which  cuts  the 
surface  in  the  curve  AFBDE.  The  element  FG  of  this  curve  is  acted 
on  by  the  tension  C .  FG,  proceeding  from  the  adjacent  parts  of  the 
surface.  FG  may  be  set  equal  to  rd<f>  and  the  tension  to  Crd<f).  Its 
direction  makes  an  angle  with  the  normal  to  the  surface,  whose 
cosine  is  r/R;  hence  the  force  acting  in  the  direction  of  the  normal 


SECT.  L.]  SURFACE  ENERGY.  123 

is  Cr2/E  .  d<f>.     The  surface  tension  therefore  draws  the  surface-element 
ABDE  toward  the  interior  of  the  fluid  with  a  force 

Cr2  [*'d<f>/lt. 

Jo 

Therefore  if  P  represents  the  pressure  on  unit  of  surface  we  have 

PTTT-  =  Cr2  l^d^jR    and    P  =  C/ir  .  f^'d^/B. 
If  for  R  we  introduce  the  value  given  in  (c),  we  obtain 


(d)  P 

It  is  probable  that  in  addition  to  the  pressure  here  found,  which  arises 
from  the  curvature  of  the  surface,  there  also  exists  a  constant  pressure 
M  which  acts  in  the  fluid  when  its  surface  is  plane.  The  total 
pressure  due  to  capillary  forces  is  therefore  M+C(l/Rl  +  l/R2),  where 
M  and  C  depend  on  the  character  of  the  two  bodies  which  are  in 
contact  in  the  surface.  Since  the  phenomena  of  capillarity  do  not 
permit  of  the  measurement  of  this  quantity  M,  it  need  not  be  further 
considered. 

At  20°  C.  the  value  of  C  for  the  surface  of  contact  between  water 
and  air  is  81,  between  mercury  and  air  540,  and  between  mercury 
and  water  418. 


SECTION  LI.     CONDITIONS  OF  EQUILIBRIUM. 

Equilibrium  exists  in  a  fluid  mass  if  its  potential  energy  remains 
unchanged  when  the  position  and  form  of  the  mass  are  changed  by 
an  infinitely  small  amount.  Since  the  energy  depends  on  the  extent 
of  surface  we  must  obtain  an  expression  for  the  increment  SS  of  the 
surface.  Suppose  a  fluid  A  surrounded  by  another  fluid  B,  the  two 
fluids  being  such  that  they  do  not  mix.  If  no  external  forces  act  on 
them,  A  will  assume  the  spherical  form. 

Suppose  the  surface  S  to  be  concave  toward  A  and  to  move  toward 
B  so  that  it  undergoes  an  infinitely  small  change  of  form.  Let  s 
be  the  contour  of  the  surface  S,  and  S'  represent  that  surface  after 
the  change  of  form  has  occurred.  The  contour  of  S'  may  be  repre- 
sented by  s'.  Erect  at  all  points  of  s  normals  to  S  which  cut  the 
surface  S'  in  a  new  curve  o-,  which  may  be  supposed  to  lie  within  s'. 
If  we  designate  the  infinitely  small  distance  between  a-  and  s'  by  81, 
the  part  of  S'  which  lies  between  o-  and  s'  will  be  given  by  (b)  j£/ .  ds. 


124  CAPILLARITY.  [CHAP.  vi. 

We  now  erect  at  a  point  P  in  S  the  normal  PP  which  cuts  «S" 
at  F ;  set  PP'  =  8t>,  and  draw  through  P  on  the  surface  S  two  curves 
PE  and  PF,  one  of  which  corresponds  to  the  maximum  curvature 
of  the  surface  at  P,  the  other  to  the  minimum.  These  principal  curves 
and  two  others  infinitely  near  them  will  bound  a  rectangle  PEQF, 
whose  sides  PE  =  a  and  PF=b  are  infinitely  small.  If  P^  and  722 
are  the  principal  radii  of  curvature,  there  will  always  be  two  angles 
a  and  (3  such  that  a  =  Rla,  b  =  R.2(3,  and  therefore  dS  =  a.b  =  filR2a/3. 
The  normals  to  S  erected  at  E,  Q  and  F,  intersect  S'  at  E',  Q\  F'. 
We  set  FE'  =  a',  P'F  =  b'  and  obtain  a'  =  (R1  +  8v)a,  b'  =  (Rz  +  8v)/3 
If  Si  is  the  part  of  S'  bounded  by  <r,  we  will  have 

dSi  =  a'b'  =  (R^  +  (Rl  +  R2)8v)a{3  and  d(Sl'  -S)  =  (l JP^  +  I JR2)8vdS. 
We  have  therefore  (c)  £/  -S=\(l/Rl  +  \/R2)8vdS.  The  total  incre- 
ment 8S  which  S  receives  in  consequence  of  the  change  of  form  is 
therefore  (d)  8S=  f (l/^  +  l/R2)8vdS  +  \8l .  ds.  This  expression  remains 
valid  even  if  81  and  8v  at  particular  points  or  at  all  points  of  the 
surface  S  are  negative.  If  the  fluid  mass  is  bounded  by  a  single 
surface,  the  contour  s  will  be  zero;  if  the  contour  is  fixed  we  have 
$  =  0.  In  both  cases  the  condition  of  equilibrium  is 

(e)  \(I/E1  +  l/R2)8vdS=0. 

Since  the  space  occupied  by  the  fluid  mass  is  supposed  constant,  we 

have  (f)  ftvdS  =  0,  since  ftvdS  represents  the  increment  of  volume. 

From  (e)  and  (f)  it  follows  that  (g)  l/Rl  +  ljRz  =  ct  where  c  is  a 
constant.  The  same  result  is  also  given  by  L.  (d)  if  we  notice  that 
the  pressure  in  the  fluid  mass  must  be  constant. 

If  three  fluids  which  do  not  mix  meet  in  a  line,  the  three  angles 
which  the  surfaces  of  the  fluids  make  with  one  another  may  be 
determined.  Such  relations  occur  if  a  drop  of  oil  lies  on  the  surface 
of  water.  In  this  case  the  three  fluids  which  meet  are  water,  oil, 
and  air.  We  shall  designate  these  fluids,  for  greater  generality,  by 
a,  b,  and  c;  let  the  energy  of  a  unit  area  of  the  surface  separating 
a  and  b  be  Cab ;  let  €„.  and  (76c  have  similar  meanings.  It  is  sufficient 
to  examine  the  case  in  which  the  edge  is  a  straight  line.  The 
directions  of  the  three  surface  tensions  Gab,  C^,  and  (7^,  determine 
the  inclination  of  the  surfaces  to  each  other ;  equilibrium  exists  when 
the  three  forces  Cab,  C^.,  C^  are  in  equilibrium.  Let  a,  /?,  y  be  the 
three  angles  sought,  belonging  respectively  to  the  three  fluids  a.  b, 
and  c.  We  then  have  as  the  condition  of  equilibrium, 
(h)  CJ  sin  a  =  CJ  sin  (3  =  CJ  sin  y. 


SECT.  LI.] 


CONDITIONS   OF   EQUILIBRIUM. 


125 


From  these  equations  we  may  in  general  determine  a,  (3,  y.  If, 
however,  one  of  the  tensions,  say  C^,  is  greater  than  the  sum  of 
the  other  two,  equilibrium  cannot  exist.  In  this  case,  the  fluid 
spreads  out  into  a  very  thin  sheet  which  separates  the  fluids  b  and 
c  ;  we  have  as  an  example  the  behaviour  of  a  drop  of  oil  of  turpen- 
tine on  water. 

The  theorem  given  in  (h)  may  be  also  obtained  by  the  following 
method  :  If  the  edge  is  displaced  by  an  infinitely  small  distance 
from  its  original  position,  the  sum  of  the  surface  energies 


increased   by  Cbc8S1  +  Coc8S2  +  Cab8S3.     This   increment  must  be  equal 
to  zero,  if  equilibrium  exists;  we  thus  obtain  equations  (h). 

If  a  solid  c  (Fig.  51)  is  in  contact  with  two  fluids  a  and  b,  the 
edges  may  be  displaced  infinitely  little  along  the  surface  of  the 
solid.  In  this  case  we  have  8S:  =  -  8S2,  8S3  =  -  S^j  .  cos  a,  and  further, 
@ic  -  Cac  =  CM  -  cos  a-  Hence  we  have  (i)  cos  a  =  (C^  -  C^)ICaM  where 
a  is  the  so-called  contact  angle. 


SECTION  LII.    CAPILLARY  TUBES. 

To  make  an  application  of  the  foregoing  principles,  we  will  consider 
a  cylindrical  tube  c  (Fig.  51  A)  placed  perpendicularly,  the  lower  end 
of  which  is  immersed  in  a  fluid  b ;  the  upper  part  of  the  tube  is 
surrounded  by  air,  which  may  be  repre- 
sented by  a.  The  bounding  surfaces  may 
be  represented  as  before  by  Sv  S2,  and 
S3.  We  may  call  the  surface  of  contact 
between  the  fluid  and  the  tube  Sv  that 
between  the  air  and  the  tube  S2,  and 
that  between  the  air  and  the  fluid  S3.  ~M  N 
The  fluid  surface  MM  outside  the  tube 
may  be  infinitely  great;  it  may  also  be 
considered  as  at  rest,  even  when  the  surface  in  the  tube  is  in  motion. 
Take  the  surface  MM  for  the  xy-plsuue,  and  let  the  z-axis  be  directed 
perpendicularly  upward.  If  g  represents  the  acceleration  of  gravity, 
and  p  the  density  of  the  fluid,  the  potential  energy  of  a  particle  of 
the  fluid  pdv,  will  be  g .  pdv .  z.  Hence  the  potential  energy  of  the 
fluid  mass  lying  above  the  xy-plane  is 


i  b 


FIG.  51  A. 


M 


126  CAPILLARITY.  [CHAP.  vi.  SECT.  LIT. 

if  x,  y,  z  are  from  now  on  considered  as  belonging  to  S3.  The  part 
of  the  potential  energy  Ep  whose  variations  are  to  be  considered  is 

EP  =  &P\  \Mxdy  +  CabS3  +  CJ3,  +  CJ3V 
In  the  case  of  equilibrium  we  have  8EP  =  Q,  or 
<a)  0  -  gp\  \z8zdxdy  +  Cab8S3  +  CM8S2  +  CJS^ 

If  s  is  the  length  of  the  line  of  section  of  the  surface  S3  with 
the  inner  surface  of  the  tube,  if  </>  represents  the  angle  between  ds  and 
the  xy-pla.ne,  and  if  all  points  of  the  surface  S3  are  elevated  by  the 
same  infinitely  small  amount  Sz,  where  8z  is  constant,  we  will  have 
8S3  =  Q,  &§!=  -8S2  =  \cos<j>8zds.  The  equation  (a)  then  becomes 
(b)  gp\  \zdxdy  =  (CM-  C7fc) J cos  <frds. 

Hence  the  fluid  will  be  displaced  by  the  difference  of  the  tensions 
in  the  surfaces  S.2  and  Sr 

If,  on  the  other  hand,  the  line  of  section  s  retains  its  position, 
and  if  the  only  change  is  the  change  in  the  shape  of  the  surface  S3, 
we  will  have  from  LI.  (d)  8S3  =  ^(l/El  +  l/Sa)8vdSs  and  8Sl  =  8S2  =  Q, 
in  which  8v  is  an  element  of  a  normal  lying  between  the  surfaces 
S  and  -ST. 

We   may   replace   Szdxdy  by  8vdS3,  and   obtain  from  equation  (a) 

f  {gpz+  C^l/E,  +  l/R2)}8vdS3  =  0. 

Since  8v  is  arbitrary,  we  must  have  (c)  gpz  +  Cab(l/Rl  +  1/A>2)  =  0.  If 
the  curvature  of  the  surface  is' expressed  by  the  differential  coefficients 
of  z  with  respect  to  x  and  y,  we  will  obtain  from  (c)  a  differential 
equation  for  the  determination  of  the  form  of  the  surface.  If  the 
contact  angle  is  also  given,  the  surface  is  completely  determined. 

If  the  cross-section  of  the  tube  is  circular  and  very  narrow,  we 
may  assume  that  approximately  B^  =  R.2  =  r/  cos  a,  where  r  is  the 
radius  of  the  tube  and  a  the  contact  angle.  The  height  z  to  which 
the  fluid  rises  is  then  z  =  -  2  cos  a/gpr .  C^.  We  obtain  the  same 
result  from  equation  (b)  if  we  set  ^zdxdy  =  Trr2z,  Jc08<kfca*2rr,  and 
use  equation  LI.  (i). 

The  theory  of  capillarity  was  discussed  by  Laplace  in  a  supple- 
ment to  the  tenth  book  of  the  Mtcanique  Celeste.  Poisson  wrote  a 
larger  work  on  the  subject,  called  Nmivelle  Thdorie  de  I' Action 
Capillaire :  Paris,  1831.  Finally  Gauss  made  an  epoch-making  in- 
vestigation on  the  theory  of  capillarity,  published  in  the  Commentationes 
Soc.  Scient.  Gottingensis.  Vol.  VII.  1830.  (Works.,  Vol.  V.,  p.  29.) 
The  most  elaborate  recent  publication  on  the  subject  is  that  of  Mathieu, 
Tktorie  de  la  Capillarity :  Paris,  1883. 


CHAPTER   VII. 

ELECTEOSTATICS. 

SECTION  LIII.    FUNDAMENTAL  PHENOMENA  OF  ELECTRICITY. 

THE  theory  of  electricity  is  founded  upon  the  observation  that 
amber  and  other  bodies  obtain  by  friction  the  property  of  attracting 
light  bodies.  Gray  showed  that  this  property  may  be  transferred 
from  one  body  to  another.  The  conception  was  thus  suggested  that 
this  property  depends  upon  the  presence  of  a  fluid  which  is  formed  or 
set  free  by  friction  in  the  body,  and  which,  under  certain  conditions, 
can  pass  from  body  to  body.  Dufay  first  showed  that  there  are  two 
so-called  electrical  conditions,  or  according  to  the  conception  just 
stated,  two  fluids,  which  Franklin  named  positive  and  negative,  since 
they  can  completely  neutralize  each  other. 

The  hypothesis  of  two  fluids  has  had  an  extraordinary  influence 
on  the  development  of  the  theory  of  electricity.  Poisson  proceeded 
from  this  conception  in  his  researches  on  electrical  distribution, 
and  W.  Weber  founded  on  it  his  theory  of  electrical  currents. 

In  opposition  to  this  theory,  which,  in  its  mathematical  treatment, 
proceeds  from  the  conception  that  electrical  action  is,  like  gravity, 
a  force  acting  at  a  distance,  Faraday  adopted  the  view  that  the 
electrical  forces  propagate  themselves  from  particle  to  particle,  not 
immediately,  therefore,  but  by  the  action  of  an  intervening  medium. 
On  this  view  it  is  not  possible  to  explain  the  phenomena  of  electricity 
completely  without  introducing  a  hypothetical  medium,  the  ether, 
and  we  thus  meet  with  peculiar  difficulties,  which  have  not  yet  been 
entirely  overcome.  Fruitful  as  Faraday's  conceptions  have  been, 
we  still  cannot  explain  many  phenomena,  even  by  their  help,  and  a 
complete  systematic  discussion  of  electricity  cannot  yet  be  given. 
In  the  following  presentation  it  has  not  been  possible  to  consider  the 
whole  subject  from  one  point  of  view  ;  we  have  only  endeavoured 
to  present  the  most  important  results  which  have  been  obtained. 

127 


128  ELECTROSTATICS.  [CHAP.  vn. 

The  starting  point  of  our  study  is  Coulomb's  investigation  of  the 
mechanical  force  which  two  electrified  bodies  exert  on  each  other. 
Two  bodies  which  carry  the  charges  el  and  e2,  measured  in  any  manner 
and  separated  from  each  other  by  the  distance  r,  will  then  act  on 
each  other,  according  to  Coulomb's  law,  with  a  force  ^  =  c .  e^/r2, 
where  c  is  a  constant.  According  as  the  two  charges  are  similar  or 
dissimilar  the  bodies  repel  or  attract  each  other.  If  the  distribution 
of  electricity  on  extended  bodies  is  known,  the  mechanical  force 
with  which  the  bodies  act  on  each  other  may  be  calculated.  As  a 
rule,  however,  the  distribution  of  electricity  on  a  body  cannot  be 
considered  given.  If  electricity  is  developed  by  friction  on  a  glass 
rod  or  a  stick  of  sealing  wax,  that  is,  on  relatively  poor  conductors, 
a  slow  discharge  of  the  same  occurs  with  lapse  of  time.  The  distribution 
of  electricity  on  good  conductors  depends  on  their  form,  on  the  char- 
acter of  the  body  surrounding  them,  and  on  the  charge.  In  deter- 
mining the  distribution,  we  start  from  the  assumption  that  the  same 
force  acts  between  two  quantities  of  electricity  as  between  two  bodies 
which  are  charged  with  these  quantities  of  electricity.  If  a  definite 
charge  is  imparted  to  an  insulated  conductor  its  separate  parts  will 
act  on  each  other,  and  there  will  be  a  definite  distribution. 

Charged  bodies  excite  an  electrical  distribution  in  neighbouring 
bodies.  The  attraction  which  a  charged  body  exerts  on  an  uncharged 
body  is  explained  by  the  assumption  that  positive  and  negative 
electricities  in  equal  quantities  are  present  in  the  latter  body,  and 
that  under  the  influence  of  the  charged  body  a  separation  of  the 
opposite  electricities  occurs,  the  force  which  proceeds  from  the  charged 
body  acting  as  an  electromotive  force.  The  distribution  thus  pro- 
duced acts  against  the  external  electromotive  force,  and  a  condition 
of  equilibrium  is  brought  about  if  the  electromotive  force,  which  arises 
partly  from  the  force  acting  from  without  which  causes  the  distri- 
bution, and  partly  from  the  electricity  separated  in  the  body  itself 
and  therefore  free,  is  everywhere  zero  within  the  conductor.  In  poor 
conductors  also  as,  for  example,  the  air,  an  electromotive  force  must 
arise  under  similar  circumstances,  whose  action  we  will  at  present 
not  consider. 


SECTION  LIV.    ELECTRICAL  POTENTIAL. 

Suppose  that  electricity  of  density  p  is  contained  within  the  body 
L  (Fig.  52),  and  that  electricity  of  density  a-  is  present  on  its  surface. 
A  volume-element  dr  then  contains  the  quantity  of  electricity  p .  dr, 


SECT.  LIV.]  ELECTRICAL  POTENTIAL.  129 

and  a  surface-element  dS  contains  the  quantity  o-  .  dS.     Suppose  that 

unit  quantity  of  electricity  is  pre- 

sent   at    the    point    P,    whose    co- 

ordinates are  x,  y,  z,  and  that  the 

charges  within  L  and  on  its  surface     (          L  ]       "jT      Q 

act  upon  it.      Let  the  coordinates 

of  any  point  in  the  body  L  be  £, 

»7,  £  and  let  X,  Y,  Z  be  the  com-  FlG  52 

ponents  of  the  force  which  acts  at 

the  point  P.     We  then  have  from  Coulomb's  law  (cf.  XII.), 

A'  =  \(x  -  £)/»«  .  pdr  +  \(x  -  £)/r»  .  a-dS, 
y=  (y  -  1*  • 


where  r*  =  (x  -  £)2  +  (y-  r,f  +  (z-  ()2. 

If  we  set  (b)  ¥  =  J/o  .  dr/r  +  Jo-  .  dS/r,  it  follows  that 

(c)  X=-Wrdx,     Y=  -'d^py,     Z=  -W/Vz. 

*¥  is  the  electrical  potential.  If  the  unit  charge  moves  from  P  to  Q, 
and  if  r'  is  the  distance  of  the  point  Q  from  the  point  (£,  ?/,  £)  of 
the  body  L,  we  will  have  W  =  ^p.  dr/r'  +  Jo-  .  dS/rr  The  work  done 
by  the  electrical  forces  during  the  motion  is 

\(Xdx  +  Ydy  +  Zdz)  =  ¥-¥'. 

If  the  point  Q  is  so  far  distant  from  the  body  L  that  ¥'  =  0,  the 
work  done  will  equal  M*.  We  can  therefore  say  :  the  electrical  potential 
at  a  point  is  equal  to  the  work  done  by  the  electrical  forces  when  a  unit 
of  electricity  moves  from  the  point  considered  to  a  point  at  an  infinite  dis- 
tance from  the  charged  body. 

If  the  point  P  is  within  the  body  L,  we  describe  a  sphere  K  of 
radius  E  about  the  point  P  as  centre,  so  small  that  the  density  p 
within  it  may  be  considered  constant.  The  force  due  to  the  charge 
in  K  is  then  zero  (cf.  XIII.)  ;  the  potential  within  L  due  to  the 
electricity  present  in  it  cannot,  therefore,  be  infinite. 

The  potential  has  the  same  value  on  loth  sides  of  a  surface  charged 
with  electricity.  If  the  surface  density  is  o-,  the  potential  ^  which 
arises  from  the  surface  distribution  is  ^F  =  Jo-  .  dS/r.  We  will  consider 
the  values  "^  and  W.2  of  the  potential  at  two  points  Pl  and  P2 
(Fig.  53)  which  lie  on  each  side  of  the  surface  AB,  and  are  separated 
by  an  infinitely  small  distance.  Let  the  line  P^PZ  be  a  normal  to 
the  surface.  We  suppose  an  infinitely  small  portion  of  the  surface 

i 


130  ELECTROSTATICS.  [CHAP.  VH. 

cut  off  by  a  circular  cylinder,  whose  axis  is  the  line  /^P.,  and  whose 

radius  is  R.     The  potential  ^  is  made  up  of  two  parts,  one  of  which 

*Py  arises  from  the  part  of  the  surface  which  is  cut  off  by 

„         the  cylinder,  and  the  other  ^rl  -  ^Py  from  the  remaining 

7  part  of  the  surface.  The  latter  value,  ^  -  "*?/,  is  neither 
discontinuous  nor  infinite  in  the  distance  from  Pl  to  P0. 
The  value  ¥2  is  likewise  made  up  of  two  parts.  The 
radius  of  the  infinitely  small  circle  cut  out  of  the  surface 
is  R.  Let  n  be  the  distance  of  Pl  from  the  surface ;  we 
will  then  have 


It  therefore  appears   that  >Py  vanishes   if  R  and  n  are 
infinitely  small. 

It  has  been  shown  [XIV.]  that 

a2(i/r)/az2 + a2(i/r)/a#2 + a2(i/r)/a^2 = o. 

FIG  53        ^e  ^ave  tneref°re>  from  (b),  for  a  point  outside  of  L, 

(d)         a^pyaa;2 + a^/a^2 + a^/a^2 = v'2^ = o. 

On  the  other  hand,  if  P  lies  within  the  body  we  have  [XIV.  (h)] 
(e)  V21?  +  4irp  =  0.  If  we  designate  the  normals  to  the  surface  drawn 
inward  and  outward  by  vs  and  va,  we  have  for  the  surface  density 
o-  [XIV.  (1)],  (f)  W,/3v,  +  a^0/ava  +  47TO-  =  0.  The  horizontal  lines  over 
the  differential  coefficients  indicate  that  their  values  are  to  be  taken 
at  the  surface.  Hence  if  v  represents  the  electrical  density  on  a  surface, 
the  sum  of  the  forces  acting  in  the  direction  of  the  normals  drawn  outward 
from  both  faces  equals  lira-. 

These  properties  of  the  potential  hold  for  every  system  of  bodies 
charged  in  any  way  with  electricity.  If  the  distribution  is  given,  the 
potential  can  be  determined  either  from  (b)  or  from  (e)  and  (f).  If 
the  potential  is  given,  the  densities  p  and  o-  are  determined  from 
(e)  and  (f),  while  the  components  of  the  electrical  force  are  given 
from  (c).  The  force  F,  acting  in  any  direction  ds,  is  F=  - 


SECTION  LV.    THE  DISTRIBUTION  OF  ELECTRICITY  ON  A 
GOOD  CONDUCTOR. 

If  a  charge  e  is  communicated  to  a  good  conductor,  it  distributes 
itself  over  the  conductor.     \Ve  will  determine  its  volume  density  p 


SECT.  LV.]  DISTRIBUTION  ON  A  CONDUCTOR.  131 

and  the  surface  density  <r.     If  M^  is  the  potential  inside,  and  *Pa  the 
potential  outside  the  conductor,  we  have  from  LIV.  (d)  and  (e) 
(a)  V2^,.  +  47rp  =  0  and  V*P,,  =  0. 

After  equilibrium  is  attained  there  is  no  electrical  separation  in  the 
interior,  that  is,  we  have  (b)  3¥,/ae  =  0,  3¥y3z/  =  0,  3¥,/a?  =  0  for 
all  points  in  the  interior  of  the  conductor.  Hence  (c)  V2^i==0,  and 
therefore  from  (a)  />  =  0.  Hence  the  electricity  is  distributed  only  on 
the  surface  of  the  conductor. 

From  (b)  ^  is  constant  in  the  interior  of  the  conductor  and  equal 
to  ¥",  where  ¥  is  the  value  of  the  potential  on  the  surface.  We  may 
determine  ¥rt  from  the  equations  y2^fa  =  0  and  '¥„  =  *¥  for  all  points 
of  the  surface.  The  surface  density  is  given  by 

3*V9i'«  +  'd¥m[dvil  +  47TO-  =  0. 

Since  ¥,  is  constant  we  obtain  (d)  4iro-  =  -^J?>va. 

If  we  represent  the  force  acting  outward  at  the  surface  of  the 
conductor  by  F,  we  have  (e)  F—  -  'dj¥a/dva  =  47ro-,  that  is,  the  force 
acting  at  a  point  on  the  surface  of  the  conductor  in  the  direction  of  the 
normal  is  equal  to  the  surface  density  a-  at  that  point  multiplied  by  47r. 

If  ds  is  an  element  of  a  curve  drawn  on  the  surface  of  the  con- 
ductor, we  have  3^/3$  =  0,  since  "*"„  is  equal  to  ^f  everywhere  on 
the  surface,  as  has  just  been  shown.  Hence  F  has  no  components 
in  the  surface.  The  surface  of  the  conductor  is  a  surface  of  constant 
potential,  and  the  direction  of  the  force  F  is  everywhere  perpendicular 
to  it. 

To  determine  the  potential  of  the  conductor  its  charge  e  must  be 
calculated  ;  we  have  e  =  J Jo- .  dS,  and  therefore 

<f)  'e  =  1/47T .  JJ F.  dS=  -  1  ,/47T .  J j3*V3v0 .  dS. 

Since  there  is  no  charge  in  the  interior  of  the  conductor,  we  have 
*Pa  =  \\<r.dSlr.  If  the  electricity  on  the  conductor,  whose  density 
is  <r,  is  in  equilibrium,  it  will  remain  so  if  the  density  becomes  wr, 
where  n  is  a  number.  If  one  distribution  which  is  in  equilibrium  is 
superposed  on  another,  the  new  distribution  is  still  in  equilibrium. 
If  the  density  is  everywhere  rw,  the  potential  has  the  value  ri*Fa. 
The  potential  is  therefore  proportional  to  the  charge.  If  the  charge  C  will 
bring  the  conductor  to  the  potential  1,  the  charge  (g)  Q  =  C^r  must 
be  imparted  to  the  conductor  in  order  to  bring  the  potential  from 
0  to  "V.  We  call  C  the  capacity  of  the  conductor.  The  capacity  C 
is  the  ratio  of  the  charge  Q  of  a  conductor  to  its  potential  ¥.  In 
order  to  give  a  means  of  representing  the  magnitude  and  direction 


132 


ELECTROSTATICS. 


[CHAP.  vn. 


of  the  electrical  force  in  the  region  around  the  conductor,  we  determine 
the  position  of  the  surfaces  of  constant  potential  which  surround 
it.  Their  equation  is  ^  =  c,  where  c  is  a  constant.  The  first  equi- 
potential  surface  is  the  surface  of  the  conductor,  for  which  ¥  =  XK 
At  a  distance  which  is  very  great  in  comparison  with  the  dimensions 
of  the  conductor,  the  equi-potential  surfaces  will  be  spheres,  since  in 
that  case  the  expression  reduces  to  ^  =  ejr. 

From  VII.  the  surfaces  of  constant  potential  are  perpendicular  to 
the  direction  of  the  force.  If  two  such  surfaces  are  considered  which 
are  infinitely  near  each  other,  it  appears  from  VII.  that  the  force 
at  every  point  in  one  of  them  is  inversely  proportional  to  the  distance 
between  the  surfaces.  If  a  line  of  force  PP1P2PB  (Fig.  54)  is  drawn 
from  a  point  on  the  surface,  and  if  the  points  Pv  P2,  P3,  etc.,  are 
so  chosen  that  F.  PPl  =  Fl.  P1P2  =  F.2.  P2P3,  etc.,  where  Flt  F^  are 
the  electrical  forces  at  the  points  Pv  P2,  we 
have  a  relation  which  may  be  carried  out  to 
any  distance.  If  we  draw  the  equi-potential 
surfaces  in  such  a  way  that  the  potential,  as 
we  pass  from  one  to  the  next,  increases  or 
diminishes  regularly  by  the  same  amount,  the 
product  Fn.(PJPn+1)  will  be  constant.  The 
further  we  pass  from  the  acting  quantity  of 
electricity  the  greater  will  be  the  distance 
between  the  successive  equi-potential  surfaces. 
If  the  magnitude  of  the  force  at  one  of  the 
equi-potential  surfaces  is  given,  its  magnitude 
at  another  point  of  the  figure  may  be  calculated  from  the  distance 
between  the  successive  equi-potential  surfaces,  and  if  the  magnitude 
of  the  electrical  force  F  at  the  surface  of  the  body  is  given,  its 
magnitude  at  every  other  point  of  the  figure  may  be  determined. 


FIG.  54. 


SECTION  LVI.    THE  DISTRIBUTION  OF  ELECTRICITY  ON  A 
SPHERE  AND  ON  AN  ELLIPSOID. 

1.  The  Sphere,  Suppose  the  charge  Q  given  to  an  insulated  sphere 
of  radius  R.  We  are  to  determine  the  potential  "*P0  at  a  point  outside 
the  sphere.  Let  the  centre  of  the  sphere  be  taken  as  the  origin  of 
coordinates.  We  have  V2^a  =  0.  Since  ¥„  is  a  function  of  the  dis- 
tance from  the  centre  0,  we  have  [XV.  (1)]  1/r  .d'2(rVn)/dr*  =  0,  and 
hence  "(Pa  =  cl  +  c<Jr,  where  cx  and  c2  are  constants. 


SECT.  LVL]  DISTEIBUTION  ON  AN  ELLIPSOID.  133 

We  assume  that  no  charged  bodies  are  present  besides  the  sphere, 
so  that  ¥„  =  0  when  r  =  <x>  ,  and  hence  cx  =  0.  The  electrical  force 
at  the  distance  r  from  the  centre  is  F=  -d^rjdr  =  c.2/r2. 

From  LV.  (e)  we  have  further 


The  potential  ¥„  and  the  capacity  C  are  therefore 

¥a=0/r,    C=QI^  =  R. 
The  dimensions  of  capacity  are  therefore  those  of  a  length. 

2.  The  Ellipsoid.  Eepresent  the  semi-axes  of  the  ellipsoid  by  a,  b,  c, 
and  its  charge  by  Q.  It  is  most  natural  to  assume  that  the  surfaces 
of  constant  potential  are  confocal  ellipsoids.  The  equation  of  a 
system  of  such  surfaces  is 

(a)  E  =  3?  I  (a?  +  A)  +  f/(b2  +  A)  +  *2/(c2  +  A)  =  1  . 

On  our  assumption  the  potential  must  be  a  function  of  A,   so  that 
we  will  write  ¥  =/(A).     To  find  this  function  /  we  proceed  from 


and  the  analogous  expressions  for  y  and  z.     These  give 

<b)     V2*  =  d'WJdX*  .  (  (3A/3.C)2  +  (3A/ay)2  +  (3A/3z)2)  +  dV/dX  .  V2A. 

We  will  set  for  brevity, 

A  =  .«2/(o2  +  A)2  +  f/(b*  +  A)2  +  ^/(c2  +  A)2 

B  =  ./:2/(«2  +  A)3 
We  then  have 
(c) 
(d) 

Analogous  expressions  hold  for  the  differential  coeflBcients  \vith  respect 
to  y  and  z.     Hence  we  have 
c>2£/a*;2  =  2/(tt2  +  A)  -  ^  .  92  A/3.7:2  -  2ar/(a2  +  A)2  . 
or,  using  equation  (c), 

(e)  32£/ac2  =  2/(a2  +  A)  -  A  .  32A/3x2  +  '2B. 
From  equations  (c)  and  (a)  it  follows  that 

(f)  A2((d\px)*  +  (3A/3y)2  +  (9A/o?)2)  =  4A, 
and  from  (e)  and  (f)  that 

(g)  A  .  V2  A  =  2/(a2  +  A)  +  2/(ft2  +  A)  +  2/(c2  +  A). 

By  the  help  of  equations  (f)  and  (g)  it  follows  from  (b)  that 
<h)   A  .  V2*"  =  4d-^/dX2  +  2d¥jdX  .  (l/(fl2  +  A)  +  l/(62  +  A)  +  l/(c2  +  A)). 


134  ELECTEOSTATICS.  [CHAP.  vn. 

Outside  the  ellipsoid   we   have   y2^  =  0.     If  C  is  a  constant,   we 
have  from  (h),  (i)  dV/d\  =  -  Cjjfa*  +  A)(62  +  A)(c2  +  A)  and 


A)(c2  +  A)  +  <72. 

At  an  infinitely  distant  point,  for  which  x  =  y  =  z  =  QC  ,  ¥"  is  assumed 
equal  to  zero  ;  and  for  such  a  point  equation  (a)  shows  that  A.  =  x>  . 
The  potential  ¥  at  any  point  (x,  y,  z)  is  therefore  given  by 


+  A)(62  +  A)(c2  +  A), 

where  A  is  known  from  the  equation 
z2/(a2  +  A)  +  ?/2/(&2  +  A)  +  s2/(c2  +  A)  =  1  . 

If  the  charge  on  the  ellipsoid  is  Q,  the  potential  at  a  point  at  a 
great  distance  from  the  ellipsoid  is  Qj^JX  ;  by  comparison  with  (k) 
we  then  obtain 


(1)  ¥  =  Q/2  .      d\lJ(a*  +  A)  (ft2  +  A)(c2  +  A). 

The  electrical  force  F  and  its  components  X,  Y,  Z  are  determined 
from  the  equations 

X  =  -  dVdX  .  3Aar      Y=  -  dVdX  .  9A3       Z  =  - 


F  =  - 

From  (f)  and  (1)  we  have 
F=  - 


If  we  represent  the  perpendicular  let  fall  from  the  origin  on  the 
plane  tangent  to  the  ellipsoid  at  the  point  x,  y,  z  of  its  surface  by  Nt 
we  have  JA-\IN>  and  hence  F=Q.  JV/>/(^+A)(6*  +  A)(c2  +  A). 

The  surface  density  o-  on  the  ellipsoid  itself  is  determined  by  the 
equation  47ro-  =  F,  and  A  =  0  for  this  ellipsoid,  so  that  (m)  a-  =  N.  Q/lirabc. 
Hence  the  electrical  density  at  a  point  on  the  ellipsoid  is  proportional  io 
the  perpendicular  let  fall  from  the  centre  on  the  plane  tangent  to  the 
ellipsoid  at  that  point. 

We  will  now  consider  several  special  cases.  In  the  case  of  an 
ellipsoid  of  rotation  a  =  b,  and  therefore  from  (1),  if  ^  is  the  potential 
of  the  ellipsoid,  we  have 

%  =  Q/2  .  J°°fU/(a2  +  AjVc^+A: 

Hence  for  a  >  c,  (n)  ¥0  =  Qf-Ja*  -  c2  .  (\TT  -  arctg  c/V«2  -  c2)  ;    for   «  =  c, 
(o)^P0  =  Q/a;  and  for  a<c, 

(P)  ^o  =  £/2v/c^2  .  log  [(c  +  N/^^)/(C  -  Vc2^^)]. 


SECT.  LVI.] 


DISTEIBUTION  ON  AN  ELLIPSOID. 


135 


If  in  (n)  we  set  c  =  0  the  ellipsoid  becomes  a  circular  plate,  and  its 
capacity  is  C=  Q/yir0  =  a/(^Tr).  For  an  ellipsoid  of  rotation  whose  length 
is  great  in  comparison  with  the  equatorial  diameter,  we  have  from  (p) 
¥0  =  Q/c .  log  (2c/o)  and  C'=c/log(2c/o). 

The  surface  density  a-,  from  equation  (m),  is 


a-  =  Q/l-rrabc .  l/Na:2/ 

If  z  is  eliminated  by  the  help  of  the  equation  of  the  ellipsoid,  and 
if  c  is  infinitely  small,  we  have  for  the  density  on  an  elliptical 
plate  whose  semi-axes  are  a  and  b,  a-^Q/^Trab.  !/>/!  -x2/a2-y'2/b'2. 
If  the  plate  is  circular,  that  is,  if  a  =  b,  and  if  we  set  x2  +  y'2  =  r2, 
we  have  <r  =  Qjlira .  l/>Ja?  -  r*.  At  a  point  whose  distance  u  from 
the  edge  is  very  small,  we  have  o-  =  Q/4:Tra.  l/*/2au.  In  this  case, 
therefore,  the  density  is  inversely  proportional  to  the  square  root 
of  the  distance  of  the  point  from  the 


SECTION  LVII.    ELECTRICAL  DISTRIBUTION. 

If  several  charged  conductors  are  present  in  a  region,  the  distribu- 
tion of  electricity  on  the  conductors  is  determined  not  only  by  their 
form  and  magnitude,  but  also  by  their  mutual  action.  The  deter- 
mination of  the  conditions  of  electrical  equilibrium  is,  as  a  rule,  very 
difficult.  The  most  important  work  on  this  subject  has  been  done 
by  Poisson  and  William  Thomson.  We  will  here  make  use  of  the 
method  of  electrical  images  given  by  Thomson. 

(a)  Distribution  on  a  Plane  Surface. — Suppose  the  quantity  of 
electricity  e  present  at  the  point  0  (Fig.  55) ;  let  AS  be  the  plane 
surface  of  a  very  large  conductor  L, 
which  is  in  conducting  contact  with 
the  earth.  The  potential  ^  of  L  is 
therefore  zero,  since  we  assume  the 
potential  of  the  earth  equal  to  zero 
(cf.  VII.).  We  are  to  determine  the 
surface  density  o-  of  the  distribution 
on  the  surface.  Let  the  potential  at 
an  arbitrary  point  in  space  due  to  the 
conductor  L  be  "*",.,  so  that  ^  is  the 
work  done  by  the  electrical  forces  of  FlG-  55- 

the  conductor  if  unit  quantity  of  electricity,  which  is  supposed  to  be 
merely  a  test  charge  and  to  have  no  effect  on  the  electrical  distribution 


136 


ELECTROSTATICS. 


[CHAP.  vu. 


on  L,  is  transferred  from  the  point  P  to   infinity.     If  OP  =  r,  the 
potential  ^F  at  P  will  be  ^  =  e/r  +  V,. 

We  now  suppose  a  quantity  of  electricity  -  e  situated  at  the  point 
0'  (Fig.  55),  which  is  the  image  of  the  point  0  with  respect  to  the 
plane  AB.  This  imaginary  quantity  at  0'  would  act  on  all  points 
lying  on  the  same  side  of  the  plane  AB  as  0,  in  the  same  way  as 
the  quantity  of  electricity  which  is  distributed  on  AB ;  for  the 
potential  which  arises  from  the  quantities  at  0  and  0'  satisfies 
Laplace's  equation  at  all  points  which  lie  on  the  same  side  of  the 


FIG.  56. 

plane  AB  as  0,  except  at  the  point  0  itself.  Further,  the  potential 
vanishes  at  all  points  of  the  plane  AB,  since  all  points  of  that  plane, 
which  passes  perpendicularly  through  the  middle  point  of  the  line 
00',  are  equally  distant  from  the  points  0  and  0',  and  hence  for 
all  points  of  the  plane  AB  we  have  e/r  -  e/r'  =  0,  where  r  and  r' 
represent  the  distances  of  a  point  in  the  plane  from  0  and  0'.  Now 
if  a  function  satisfies  Laplace's  equation  and  assumes  assigned  values 
over  a  given  surface,  and  if  the  function  itself  and  its  differential 
coefficients  are  continuous,  it  is  single-valued  and  determinate.  This 
theorem  is  known  as  Dirichlet's  Principle. 


SECT.  LVII.]        DISTRIBUTION.     ELECTRICAL  IMAGES.  137 

It  should  be  noticed  that  "9~e  =  e/r',  and  that  the  potential  ¥  at  the 
point  P  is  V  =  e/r -e/r'. 

A  unit  quantity  of  positive  electricity  lying  at  a  point  P  (Fig.  56) 
in  the  plane  AB  is  acted  on  by  two  forces,  K=e/r2  and  K  =  e/r''2, 
whose  directions  coincide  with  OP  and  PO'  respectively.  Hence  the 
direction  of  the  resultant  force  F  is  parallel  to  00'  and  equal  to 
F=  -  2eOI»/OP3,  if  we  consider  the  force  positive  when  it  is  directed 
toward  the  region  in  which  0  lies.  Now  since  4:Tra-  =  F  we  obtain 
<T=  -  e/2ir  .  OB/OPS.  The  surface  density  at  the  point  P  (Fig.  56) 
is  therefore  inversely  proportional  to  the  cube  of  the  distance  of  the  point  P 
from  the  point  0,  at  which  the  quantity  of  electricity  +e  is  situated. 
The  potential  and  the  surface  density  are  calculated  in  the  same 
way  when  several  points  carrying  charges  are  present  in  the  region. 

(b)  The  Sphere. — Suppose  that  the  quantities  of  electricity  e  and  e' 
are  situated  at  the  points  0  and  0'  (Fig.  57).  The  equi-potential 


FIG.  57. 

surface  for  which  the  potential  vanishes  is  given  by  the  equation 
4/r  +  e'/r'  =  Q  if  r  and  r  represent  the  distances  of  a  point  on  the 
equi-potential  surface  from  0  and  0'  respectively.  If  e  and  e'  have 
the  same  signs,  this  equation  represents  a  surface  lying  at  infinity ; 
if  e  and  e'  have  opposite  signs  it  represents  a  sphere  and,  in  the 
limiting  case,  a  plane. 

The  centre  C  of  the  sphere  lies  on  the  line  00'  and  we  have 
00 :  0'C=  e2 :  e"2  and  CO'/CB  =  CBJCO  =  e'/e.  The  triangles  CO'B  and 
CBO  are  similar,  and  the  radius  CB  of  the  sphere  is  the  mean 
proportional  between  the  distances  of  the  points  0  and  0'  from  the 
centre  of  the  sphere. 

If  a  hollow  sphere  of  very  thin  sheet  metal,  in  conducting  connection 
with  the  earth,  is  brought  into  the  place  occupied  by  this  spherical 
surface  of  zero-potential,  the  potential  of  points  in  the  region  will 
not  be  changed,  either  within  or  without  the  sphere  ;  the  electrical 
action  depends  only  on  the  quantities  of  electricity  e  and  e'. 


138  ELECTROSTATICS.  [CHAP.  vn. 

If  the  sphere  remains  in  conducting  contact  with  the  earth,  and  if 
we  remove  the  quantity  of  electricity  e'  from  the  interior  of  the 
sphere,  the  potential  within  the  sphere  will  become  zero,  while 
outside  the  sphere  it  retains  its  former  value,  since  the  quantity  of 
electricity  e  does  not  change  its  position  and  the  potential  of  the 
sphere  still  remains  zero.  Hence,  the  quantity  e  lying  outside  the 
sphere,  kept  at  potential  zero,  together  with  the  electricity  induced 
on  the  sphere,  exerts  on  points  outside  the  sphere  the  same  action 
as  if  the  induced  electricity  were  replaced  by  the  mass  e'  lying  inside 
the  sphere.  We  call  the  point  0',  where  the  mass  e'  is  situated,  the 
electrical  image  of  the  point  0.  The  quantity  of  electricity  e'  at  that 
point  will  exert  the  same  action  as  the  quantity  of  electricity  actually 
present  on  the  sphere.  In  optics  a  point  which  appears  to  emit 
light  from  behind  a  mirror  or  lens  which,  if  it  were  self-luminous, 
would  emit  rays  in  the  same  direction  as  those  which  proceed  from 
the  mirror  or  lens,  is  called  a  virtual  image.  Hence,  we  may  consider 
0'  the  electrical  image. 

If  we  set  C0'=f,  (70  =  a,  and  CB  =  R,  we  have 

e'/e  =  0'B/OB=f/E  =  R/a; 

and  hence  e'  =  Be/a.  We  set  OB  =  r  and  0'B  =  r'.  The  force  e/r2 
acts  at  B  in  the  direction  OB,  and  the  force  e'/r'2  acts  at  the  same 
point  in  the  direction  BO'.  The  former  of  these  may  be  resolved 
into  the  components  e/r'2.  a/r  along  OC  and  e/r2.  Rjr  along  CB,  the 
latter  into  -e'/r'2.f/Y  along  OC  and  -e'/r'2.R/r  along  CB.  The 
two  components  in  the  direction  OC  are  equal  but  oppositely  directed, 
and  therefore  annul  each  other.  The  other  two  combine  to  give  the 
force  efi/r3  -  e'R/r'3  =  -  (a2  -  R?)/E  .  e/r5,  which  acts  in  the  direction 
CB.  The  sphere  is  there/we  an  equi-potential  surface,  since  the  direction  of 
the  force,  at  any  point  of  its  surface,  coincides  tvith  the  direction  of  the 
normal  to  that  surface. 

The  density  <r,  as  obtained  from  F=  4™,  is  cr=  -  (a-  -  R^/^irR.e/r3, 
and  hence  is  inversely  p-oportional  to  the  third  power  of  the  distance 
from  the  charged  point  0.  The  quantity  of  electricity  on  the  sphere 
is  -  e'  =  -  Re  I  'a,  since  this  charge  produces  the  same  potential  in  the 
region  as  the  actual  charge  on  the  sphere,  and  therefore  [LV.  (f)]  must 
be  equal  to  it. 

The  sphere  is  attracted  by  the  point  0  with  the  force 


If  the  distance  (70  =  a  is  very  great,  this  force  becomes  Re^/a3. 


SECT.  LVII.]  ELECTRICAL  IMAGES.  139 

(c)  If  the  sphere  is  originally  insulated  and  uncharged,  we  can 
find  the  electrical  distribution  on  it  by  assuming  that,  besides  carrying 
a  charge,  distributed  as  above  described,  it  also  carries  a  uniformly 
distributed  charge  whose  surface-density  is  e'/4:-!rR2  =  e/4:TrRa.  The 
sum  of  these  two  charges  or  the  charge  of  the  sphere  is  equal  to 
zero.  The  surface-density  is  then  o-  =  e/4;rj?.(l/a-  (a2  -  R2)/^).  The 
surface-density  cr  is  zero  on  a  circle  whose  periphery  is  distant 
r  =  af/1  -  ft?  /a2  from  the  point  0.  The  plane  of  this  circle  lies  nearer 
to  0  than  the  centre  of  the  sphere.  In  order  to  find  the  potential^ 
of  the  sphere,  we  determine  it  for  the  centre.  Since  the  charge  on 
the  sphere  is  zero,  the  potential  due  to  that  charge  is  also  zero  ;  the 
potential  at  the  centre  is  therefore  ^  =  eja.  This  follows  from  the 
remark  that  the  induced  charge  -  e'  and  the  charge  e  at  0  together 
have  no  effect  on  the  potential  of  the  sphere  ;  the  potential  is  due  to 
the  additional  charge  +  e',  which  makes  the  potential  e  /R  =  e/a. 

The  force  with  which  the  sphere  is  attracted  by  0  is  in  this  case 
very  much  smaller  than  if  the  sphere  were  in  conducting  connection 
with  the  earth.  It  is 

ee'/(a  -/)2  -  ee'/a2  =  Re2/as.  R2('2a2  -  R2)/(a2  -  R2)2. 

When  R  is  very  small  in  comparison  with  a,  the  force  is  approximately 
Re2/a3.  2R2fa2.  In  this  case  we  have  a  simpler  expression  for  the 
surface-density  a-.  Designating  the  angle  BCO  by  6,  we  have 


If  a  is  so  great  that  the  higher  powers  of  R/a  can  be  neglected,  we 
have  r~3  =  a~3(l  +  3R/a  .  cos  6).  If  we  designate  the  inducing  force 
e/a?  which  proceeds  from  0  by  X,  we  have  a-  =  -  3  cos  6/47r  .  X,  if, 
in  the  formula  for  o-,  we  consider  the  radius  R  as  infinitely  small  in 
comparison  with  a  and  substitute  the  value  just  given  for  r~3. 


SECTION  LVIII.     COMPLETE  DISTRIBUTION. 

If  a  charged  body  A  (Fig.  58)  is  situated  in  the  interior  of  a  metallic 
shell  BC,  there  will  be  a  distribution  of  electricity  on  the  shell.  If 
A  is  charged  positively,  the  inner  surface  B  will  be  negatively,  and 
the  outer  surface  C  positively  electrified.  Let  the  charge  on  A  be 
e,  that  on  B  -  e',  and  that  on  C  +  e'.  We  will  show  that  the 
quantity  of  induced  electricity  e'  is  equal  to  the  quantity  of  induc- 
ing electricity  e.  Let  us  suppose  a  closed  surface  D  drawn  in  the 
interior  of  BC.  If  ¥  is  the  potential  in  the  shell  BC,  and  v  the 


140 


ELECTROSTATICS. 


[CHAP.  vii. 


normal  to  the  element  dS  of  the  surface  D,  we  have  from  XIV. 
(c),  4ir(e  -e')=-  \(dy/Vv)dS.  Since  the  potential  "*~  in  the  shell  is 
constant  the  integral  vanishes,  and  hence  e  =  e'.  The  charge  e  on  C 
can  be  conducted  off,  or  another  charge  can  be  conducted  to  it 
without  causing  any  change  in  the  charges  A  and  B.  If  the  inducing 
body  is  entirely  surrounded  by  the  body  in  which  electricity  is 
induced,  we  may  say  that  the  induction  is  complete ;  the  inducing  and 
induced  quantities  of  electricity  are  equal. 

C 


FIG.  58. 


FIG.  59. 


This  theorem  may  be  used  in  the  comparison  of  the  charges  of 
different  conductors.  If  the  outer  surface  of  BC  is  connected  with 
an  instrument  which  will  measure  potentials,  and  if  the  charged 
bodies  to  be  tested  are  brought  successively  into  the  hollow  within 
BC,  the  potentials  indicated  by  the  instrument  are  proportional  to 
the  magnitudes  of  the  charges. 

Suppose  the  sphere  A,  whose  radius  is  R,  to  be  charged  with  the 
quantity  e  (Fig.  59).  Let  BC  be  a  spherical  shell  concentric  with  A, 
whose  radii  are  R2  and  Ry  The  inner  surface  of  BC  is  then 
charged  with  the  quantity  -  e.  If  there  is  no  electricity  on  the 
outer  surface  (7,  the  potential  ¥  at  A  is  *  =  ejRl  -  e/R2.  The  potential 
at  a  point  within  the  shell  BC  or  outside  the  surface  C  is  zero,  since 
both  charges  act  with  respect  to  an  external  point,  as  if  they  were 
concentrated  at  the  common  centre  of  the  spheres.  From  the 
definition  of  the  capacity  (7,  we  have  [cf.  LV.  (g)] 

(a)  C=e/V  =  RiRJ(Rz  -  A). 

The  induction  can,  in  many  cases,  be  almost  complete,  even  when 
the  one  conductor  is  not  completely  surrounded  by  the  other.  Let 
ABC  and  DEF  (Fig.  60)  be  two  conductors  whose  surfaces  BC  and 


SECT.  LVIII.] 


COMPLETE  DISTRIBUTION. 


141 


DF  lie  very  near  each  other.     On  the  surface  BC  describe  a  closed 

curve  GH,  and  from  all  points  of  it  draw  lines  of  force,  which  cut 

out  on  DF  the  curve  KJ.      Now  draw  a  closed 

surface  G'GJJ'K'KHH'  in  such  a  manner  that  the 

two  curves  GH  and  JK  lie  in  it.      The  surfaces 

bounded  by  the  curves  G'H'  and  J'K'  lie  inside 

the  conductors  and  are  congruent  to  the  surfaces 

GH  and  JK  respectively.     Let  dS  be  a  surface- 

element  of  the  closed  surface  G'H'  J'K',  and  let  e 

and  e'  be  the  charges  of  the  surfaces  GH  and  JK. 

The  potential  is  represented  by  ¥  and  the  normal 

by  v.     We  then  have  from  LV. 

(b)  ±Tr(e  +  e')=  -  \\Wfiv  .  dS. 

The  integral  vanishes,  since  ^  is  constant  within  the  conductors,  and 
since  between  the  conductors  the  force  is  parallel  to  the  closed  surface. 
Hence  we  have  e  =  -e'.  If  the  surfaces  BC  and  DF  lie  very  near  each 
other,  we  also  have  <r=  -a-',  that  is,  the  densities  on  the  two  surfaces 
are  equal  but  of  opposite  sign. 

If  a  is  the  distance  between  the  surfaces  BC  and  DF,  and  if  ^ 
is  the  potential  of  ABC,  and  ^2  the  potential  of  DEF,  the  electrical 
force  F  in  the  intervening  space  is  [VII.  (e)] 

¥j  =  ¥2  +  Fa,  F=(¥1-  ¥2)/a 


-  ¥2)/47 


The 


The  surface-density  o-  is  [cf.  LV.  (e)]  a-=  -  <r'  = 
charge  on  the  surface  S  is  e  =  (^  -  ¥2)/47ra  .  S. 

If  the  conductor  DF  is  connected  with  the  earth,  that  is,  if  ^  =  0, 
the  capacity  C  will  be  C*  =  $/47ra,  that  is,  the  capacity  is  inversely 
proportional  to  the  distance  between  the  conductors, 

This  formula  is  used  in  air-condensers,  when  a  is  very  small. 


SECTION  LIX.     MECHANICAL  FORCE  ACTING  ON  A  CHARGED  BODY. 

If  an  element  of  volume  contains  the  quantity  of  electricity  pdv, 
it  is  acted  on  by  a  force  whose  components  are  Xpdv,  Ypdv,  and  Zpdv. 
From  LV.  (a)  the  ,r-component  can  be  expressed  by 

1/47T .  Wrdx .  ^dv  =  +  1/47T .  X(dX/*dx  +  VY/dy  +  dZfdz)dr. 

In  the  interior  of  a  good  conductor  the  force  and  electrical  density 
are  zero ;  but  a  force  acts  on  each  element  of  its  surface,  where  the 
density  is  not  zero.  This  force  is  determined  in  the  following  way  : 


p+- 

1     E 


142  ELECTROSTATICS.  [CHAP.  vn. 

Let  AD  (Fig.  61)  be  the  electrified  surface,  and  BC  =  dS  the  surface- 
element,   which  is  cut  out  by  an  infinitely   small   sphere   described 
-  about  the  point  P  as  centre, 'with  the  radius  PB  =  PC. 

Let  P2  lie  on  the  normal  PP2  to  BC,  and  let  P.2P  be 
infinitely  small  in  comparison  with  PB.  A  unit  of 
electricity  at  the  point  P2  is  acted  on  by  the  force 
27ro-  arising  from  the  distribution  on  the  surface  BC 
[cf.  XIII.  (3)].  At  the  corresponding  point  Pl  on 
the  opposite  side  of  BC  the  force  acting  is  -  27nr.  If 
/,  m,  n  are  the  cosines  of  the  angles  which  PjP.,  makes 
with  the  axes,  and  X,  Y,  Z  are  the  components  of 
force  which  arise  from  all  the  electricity  present  except 
FIG.  61.  that  on  BC,  we  have 

X.,  =  -  3¥2/ae  =  X  +  2Tro-l,    Xl  =  -  3¥,/az  =  X  -  2Tro-l. 

X2,  Xl  and  ¥2,  ^x  are  the  components  of  force  and  the  potentials 
at  the  points  P2  and  Pl  respectively.  We  have  (c)  X=^(X<,  +  Xl). 
Analogous  expressions  hold  for  the  other  components  of  force.  X 
represents  the  force  Avhich  acts  on  a  unit  of  electricity  on  dS  in  the 
direction  of  the  a:-axis.  The  element  dS  is  therefore  moved  in  the 
direction  of  the  #-axis  by  the  force  (d)  Xa-dS  •=  %(X.2  +  XJo-dS. 

If  the  element  dS  is  part  of  the  surface  of  a  good  conductor,  and 
if  Pl  lies  within  the  conductor,  the  force  at  Pl  is  equal  to  zero.  If 
we  represent  the  force  acting  at  P.2  by  F,  the  force  which  acts  on 
(IS  is,  from  (c)  and  (d),  (e)  ^Fa-dS.  Since,  from  LV.  (e),  we  have 
F=  +47ro-,  the  force  sought  will  be  (f)  2Tro-'2dS=l/8ir.F2.dS. 

W.  Thomson  has  made  an  interesting  application  of  this  equation 
in   the   construction  of  his  absolute  electrometer.     This  consists  of  an 
insulated  metal  plate  EF  and  a  smaller  circular  plate  CD,  which  is 
parallel  with  EF.     CD  forms  a  part  of  the  base  of  a  metallic  cylinder 
AB  (Fig.  62).  '  If  the  potential  of  EF 
is  ¥,  and  that  of  CD  and  AB  is  zero, 
CD  will  be  attracted  to  EF  by  a  force 
which  is  determined  in  the  following 
way :    Since  AB  and  CD  are  almost 
like  a  single  continuous  body,   there 
is  no  perceptible  surface  distribution 
on  the  inner  surface  of  A  BCD.     Re- 
FIG.  62.  present  the   density  on  the  external 

surface  of  CD  by  o-,  the  distance  between  CD  and  EF  by  a,  and  let 
the  surface  CD  =  S.  The  force  K  which  attracts  CD  toward  EF  is, 


C  D 


SECT.  L1X.] 


FORCE  ON  CHARGED  BODY. 


143 


from  (e),  equal  to  K=$F<rS.  We  have  Fa  =  ^  [cf.  VII.],  ^<r  =  F, 
and  therefore  K=S^'2l%iraz.  We  determine  the  weight  which  is 
necessary  to  counterbalance  the  electrical  attraction.  If  M  grams  are 
necessary  for  this  purpose,  we  have  ^  =  a»j8TrMg/St  where  g  is  the 
acceleration  of  gravity. 


A      B 


FIG.  63. 


FIG.  64. 


SECTION  LX.    LINES  OF  ELECTRICAL  FORCE. 

All  actions  by  which  electricity  is  produced,  such  as  friction,  induc- 
tion, etc.,  produce  equal  quantities  of  positive  and  negative  electricity  ; 
for  this  reason  we  are  led  to  assume  that  in  every  unelectrified  body 
equal  quantities  of  positive  and  negative  electricity  are  present.  Let 
A  and  B  (Fig.  63)  be  two  bodies  electrified 
by  friction  which  are  gradually  separated 
further  and  further  from  each  other,  as  in 
Fig.  64 ;  during  this  separation  they  retain 
equal  but  opposite  charges.  Suppose  A  and 
B  to  be  good  conductors  and  to  be  insulated.  " 
Now,  construct  lines  of  force  from  all  points 
of  the  contour  of  a  surface-element  dS  on  A. 
They  will  determine  a  surface-element  dS'  on 
B.  The  region  bounded  by  the  lines  of  force 
and  the  two  elements  dS  and  dS'  may  be 
called  a  sphondyloid.  If  the  density  is  o-  on  dS  and  a-'  on  dS',  we  may 
prove,  as  in  the  discussion  following  LVIII.  (b),  that  a-dS  -  a-'dS'  =  0. 
The  surface-densities  on  the  two  charged  conductors  A  and  B  are 
therefore  inversely  proportional  to  the  surfaces  limited  by  the 
sphondyloid. 

If  the  conductor  A  is  charged  with  a  quantity  Q  of  positive  elec- 
tricity, and  if  the  surface  of  A  is  cut  into  Q  parts,  each  one  of  which 
is  charged  with  unit  quantity,  the  lines  of  force  drawn  from  the 
contours  of  the  Q  parts  on  A  cut  the  surface  B  also  into  Q  parts, 
each  one  of  which  is  charged  with  unit  quantity  of  negative  elec- 
tricity. 

If  a  conductor  A  BCD  is  charged  in  any  manner  to  the  potential 
""P,  the  equi-potential  surfaces  about  it  are  spheres,  at  all  distances 
which  are  great  in  comparison  with  its  dimensions.  If  we  construct 
the  equi-potential  surface,  whose  potential  is  (¥  -  1),  it  lies  nearest 
the  conductor  at  the  points  A,  B,  and  D  (Fig.  65).  At  these  points 
[LV.]  the  electrical  force  and  the  surface-density  are  greatest.  The 


144 


ELECTROSTATICS. 


[CHAP.  vn. 


lines  of  force  also  lie  in  closest  proximity  to  one  another  at  these 
points.  If  the  conductor  has  edges  or  points  projecting  outward, 
the  density  on  them  is  very  great;  it  will  be  infinitely  great  on  a 
perfectly  sharp  edge.  On  this  depends  the  so-called  action  of  points  ; 
the  density  of  the  electricity  is  greatest  at  these  points,  and  therefore 
the  electricity  flows  out  from  them  with  especial  ease. 


FIG.  66. 


FIG.  65. 


If  A  (Fig.  66)  is  a  conductor  charged  with  positive  electricity, 
and  B  an  insulated  conductor  without  charge,  negative  electricity 
will  be  present  at  all  parts  of  B  which  are  met  by  the  lines  of  force 
proceeding  from  A.  Lines  of  force  also  proceed  from  the  other  points 
of  B ;  the  number  of  the  lines  which  fall  upon  B  is  equal  to  the 
number  of  those  proceeding  from  B.  Hence,  the  surface  of  B  is 
divided  into  two  parts  with  opposite  charges.  The  parts  are  separated 
by  a  curve  encircling  the  body  B,  along  which  the  surface-density 
a-  and  therefore  also  the  electrical  force  are  zero.  This  curve  is  the 
line  of  intersection  of  B  and  an  equi-potential  surface  around  A. 

Let  ABC  (Fig.  67)  be  a  conductor  on  whose  surface  the  potential 
is  constant  and  equal  to  ¥,  and  let  A'B'C"  be  an  equi-potential 
surface  at  which  the  potential  is  *Pr  We 
suppose  that  the  charge  of  each  surface- 
element,  for  example  of  AB  =  dS,  is  moved 
outward  in  the  direction  of  the  lines  of 
force  and  transferred  to  the  surface  A'B'C'. 
If  this  surface  is  a  conductor,  the  electricity 
transferred  to  it  is  in  equilibrium.  If  the 
potential  within  the  surface  A'B'C'  is  ¥j, 
and  if  it  retains  its  former  values  outside 
of  that  surface,  the  condition  V^a  =  0  is 
fulfilled  for  all  external  points.  If  the  electrical  forces  at  AB  and 
A'B'  are  F  and  F'  respectively,  we  have  from  LV.  (f),  since  AB  and 


SECT.  LX.]  FORCE  ON  CHARGED  BODY.  145 

A'B'  are  bounded  by  lines  of  force,  AB .  F=  A'B' .  F'.  If  o-  and  a-' 
represent  the  densities  at  AB  and  A'B'  respectively,  we  have  further 
AB.a-  =  A'B'.a-'.  We  therefore  obtain  F/<r  =  F'/(r'.  We  have,  how- 
ever, F=±irv  and,  therefore,  also  F'  =  4Tra-'. 

Hence,  if  we  divide  the  surface  of  the  conductor  into  elements,  each 
of  which  contains  unit  quantity  of  electricity,  and  if  we  draw  lines 
of  force  outward  from  the  boundary  of  the  element,  these  lines  bound 
a  tube.  The  tube  cuts  the  equi-potential  surfaces  surrounding  the 
conductor  in  such  a  way  that  for  all  of  them  we  have 

F'/<r'  =  F"/<r"  =  ...F/v. 

From  the  form  of  the  tubes  of  force  we  obtain  a  representation  of 
the  distribution  of  electrical  force  in  the  region.  And  also  from  the 
distribution  of  the  lines  of  force  we  may  distinguish  between  the 
attractive  and  repulsive  forces.  Two  lines  of  force  proceeding  in  the 
same  sense  repel  each  other,  so  that  the  repulsion  maintains  equili- 
brium with  the  tension  which  acts  along  the  lines  of  force  [cf.  XXVII]. 


SECTION  LXI.    ELECTRICAL  ENERGY. 

A  conductor  charged  with  the  quantity  of  electricity  e  can  do 
work  in  consequence  of  that  charge  ;  it  possesses  electrical  energy.  If 
the  charged  surface  ABC  (Fig.  67)  is  extended  so  that  it  assumes  in 
succession  the  form  of  its  equi-potential  surfaces,  the  forces  acting 
on  the  electrified  surface  do  an  amount  of  work  which  can  be  deter- 
mined. If  the  surface  of  the  conductor  has  the  potential  ^,  and 
if  it  is  extended  until  it  coincides  with  the  equi-potential  surface 
whose  potential  is  ^r  +  d^r,  the  work  done  is  calculated  in  the  follow- 
ing way  :  A  surface-element  dS  carrying  the  charge  crdS  is  acted 
on  by  the  force  i  .  Fa-dS.  We  represent  by  dv  the  distance  of  the 
equi-potential  surface  V  +  cW  from  the  conductor.  The  work  done 
on  the  element  dS  during  its  motion  is  ^Fa-dSdv.  The  total  work 
done  is  therefore  ^^Fo-dSdv. 

From  the  definition  of  the  equi-potential  surface  [cf.  LV.  (e)]  we 
have  Fdv  =  -d^f  •  we  therefore  have  for  the  total  work  done 


If  the  surface  of  the  body  is  extended  until  it  coincides  with  the 

equi-potential  surface  at  which   the  potential  is  zero,  the  work   W 
done  is 

(a)  W= 


146  ELECTEOSTATICS.  [CHAP.  vn. 

All  the  work  which  can  be  done  on  the  given  conditions  is  represented 
"by  W,  and  hence  it  is  called  the  potential  energy  of  the  conductor. 

The  electrical  energy  of  the  conductor  can  be  geometrically  repre- 
sented in  the  following  way :  About  the  conductor  L  (Fig.  68), 
whose  potential  is  M*",  we  construct  the 
equi-potential  surfaces  whose  potentials  are 
successively  IP-I,  ¥-2,  ¥-3,  etc.,  and 
we  divide  the  surface  of  the  body  L  in 
such  a  way  that  each  part  carries  unit 
charge.  The  space  surrounding  the  body 
L  is  divided  into  e^f  parts  by  the  lines 
of  force  starting  from  the  boundaries  of 
the  separate  parts  and  by  the  equi-potential 
surfaces ;  the  number  of  these  parts  is 
double  the  value  of  the  electrical  energy. 

If  we  designate  the  electrical  energy  by  W  and  the  capacity  by 
C,  we  have  e^CY,  and  thus  (b)  JF-^IVjC^-l^/a  We  have 
seen  before  that  the  energy  is  also  given  by 

W=  \\  \FvdSdv  =  I / Sir .  J  \F\lSclv. 

If  X,  Y,  Z  are  the  components  of  the  electrical  force,  and  if  dxdydz 
is  a  volume-element,  we  will  have 

(c)  W  =  1  /Sir .  J  J  \(X2  +  T-  +  Z*)dxdydz. 

If  the  two  conductors  A  EG  and  A'B'C'  (Fig.  67)  are  charged  with  the 
quantities  of  electricity  +  e  and  -  e  respectively,  and  have  the  potentials 
M^  and  *Fy  we  obtain  their  potential  energy  in  the  same  way  by 
supposing  the  body  ABC  carrying  the  charge  +e  to  be  gradually 
extended  so  as  to  coincide  with  the  equi-potential  surfaces  which 
surround  it.  In  this  way  the  charge  e  is  finally  transferred  to  A'B'C'. 
The  integral  in  (a)  then  becomes 

(d)  W=  -  \  l**edV  =  \e(^  -  ¥2). 

This  method  of  treatment  may  be  applied  in  all  cases.  A  system 
of  conductors  whose  charges  are  ev  e2,  e3,  ...,  and  whose  potentials  are 
"*?!,  ^2,  ^Pg,  ...,  have,  with  respect  to  a  conductor  whose  potential  is 
^PO,  the  potential  energy 

<e)  W=  K(*i  -  %)  +  &(*«  -  *0)  +  -  - 

If  the  sum  of  all  the  charges  is  zero,  that  is,  if  el  +  e.2  +  . . .  =  0 
we  have  (f)  ir=%el¥l  +  ie2¥2+ ... .  Therefore,  if  all  charged  con- 
ductors are  brought  to  the  same  potential,  the  electrical  energy  is  inde- 
pendent of  the  value  of  the  common  potential. 


SECT.  LXI.]  ELECTRICAL  ENERGY.  147 

The  expression  for  the  electrical  energy  may  be  derived  in  still 
another  way.  In  a  system  of  conductors  the  electrical  distribution 
is  determined  if  the  density  p  is  given  at  all  points  in  terms  of  the 
coordinates  x,  y,  z.  The  potential  at  any  point  is  then,  in  the  usual 
notation,  *  =  \pdr/r.  The  potential  increases  and  diminishes  pro- 
portionally to  p.  If  the  density  of  the  electricity  is  doubled  at  all 
points,  the  value  of  the  potential  also  is  doubled. 

If  the  charge  l/n.pdr  is  removed  from  every  volume-element,  the 
potential  becomes  (n  —  1  )/n  .  "*F.  In  order  to  transfer  the  quantity  of 
electricity  l/n.  \pdr  to  a  distant  and  very  large  body  whose  potential 
is  •*•„,  for  instance  to  the  earth,  the  work  l/n.  f/orfrC'P  -  ¥•„)  must  be 
done,  if  n  is  a  very  large  number.  If  the  quantity  l/n  .  \pdr  is  again 
removed,  the  work  required  is 


If  the  whole  charge   is  at  last   transferred  to  the  earth,  the   work 
W  which  is  done  is 


W=  l/n .  \pdr( I  +  (n  -  I )/n  +  (n  -  2)/n  +  ...  +  l/rt)*  -  *0f pdr. 
Now,    we    have     1  +(n-  l)/n  +  (n-  2)/n+  ...  +  l/n  =  n(n+  l)/2n,    and 
if  n  is  very  great   (g)   W=  ^^pdr-^Q  .  \pdr.      If  the   sum   of  the 
quantities  of  electricity  present  is  zero,  we  will  have  (h)  W=  ^^pdr. 
Since  V2*  =  32*/3x2  +  VW/'dy2  +  32*/os2  =  -4^,  we  have 

Now,  I  HV.'&VI'da?.  dxdydz 

=  f  J(*  .  3*/ae) .  dydz  -  J  { f (3*/9z)2 .  dxdydz. 

Hence,  by  partial  integration  extended  over  the  whole  volume,  we 
obtain 

<i)      W=  1  /STT  .  {((3*/3z)2  +  (d^j'dyf  +  (3*/^)2)^r  =  I /Sir .  \F*dr. 
This   result   has    already    been  derived    for   the   energy   in   a    good 
conductor. 

SECTION  LXII.    A  SYSTEM  OF  CONDUCTORS. 

If  several  insulated  conductors  Av  Ay  A3  are  given,  and  if  a  unit 
of  electricity  is  imparted  to  one  of  them,  say  to  Av  while  the  others 
have  no  charge,  then  the  potential  of  A^  becomes  plv  while  the  potentials 
of  A 2  and  Ay  etc.,  become  pn  and  pl3  respectively.  If  A.2  were  to 
become  charged  with  unit  quantity  while  the  other  conductors  were 
to  remain  uncharged,  the  potential  of  Az  would  equal  p22,  and  the 


148  ELECTEOSTATICS.  [CHAP.  vn. 

potentials  of  Av  As,  A4,  .  .  .  would  be  p.2v  p.2.A,  p.2v  .  .  .  respectively.  Now, 
if  the  conductor  Al  is  charged  with  the  quantity  e^,  the  conductor 
A  2  with  the  quantity  e2,  etc.,  the  potentials  M-^,  ^2,  ...  of  the  con- 
ductors Av  A2,  A5,  ...  respectively  will  be  expressed  by 


The  total  energy  of  the  electrical  system  is 

(b)  W= 

Hence, 
(c) 


If  an  infinitely  small  quantity  of  electricity  Sel  is  communicated  to 
one  of  the  conductors,  for  example  to  Av  the  energy  of  the  system 
will  be  increased  by  8/iT=  ^Er18e1.  This  increment  of  the  energy  may 
also  be  expressed  by  the  help  of  (c),  since  we  have 

(d)  8fT=  (p^  +  |(p12  +p^e2  +  $(pl3  +p3l)e3  +  ...  )&r 

Now,  since  §W  is  in  this  case  equal  to  ^f18e1,  it  follows  from  the 
first  of  equations  (a)  that  (e)  S?F=  (^>11e1  +^21e2  +^)31e3  +  ...)Ser  Com- 
paring the  formulas  (d)  and  (e),  we  obtain 


Therefore,  (f)  p.2l  =plv    psl  =^13,  and  in  general  pmn  =pnm. 

The  electrical  energy  W,  of  the  system  may  therefore  be  expressed 
as  a  homogeneous  quadratic  function  of  the  charges  by 

(g)     W*  =  &nei2  +  \Pve*  +  fce32  +  •  •  •  +Pi2eie2  +Pi 
where  the  coefficients  pmn  are  called  coefficients  of  potential. 
If  equations  (a)  are  solved  for  ev  e.2,  es,  we  obtain 


(h) 


If  the  charge  8el  is  communicated  to  the  conductor  Al  and  the  charge 
8e2  to  the  conductor  A^  etc.,  so  that  the  potential  ^  is  increased 
by  8¥v  while  the  other  potentials  retain  their  original  values,  we  have 


The  increment  of  the  energy  is  therefore 
(i)  8/ 


SECT.  LXII.]  A  SYSTEM  OF  CONDUCTORS.  149 

From  equation  (h)  the  energy  W  is  given  by 

(k) 

If  ^  is  increased  by  8WV  we  have 

Comparing  equations  (i)  and  (1),  it  follows  that  (m)  qmn  =  qnm.     The 
energy  JF^,,  expressed  in  terms  of  the  potentials,  is  therefore  given  by 


•"  +028*^8+  •••• 


The  coefficients  <?„„,  in  which  the  indices  are  the  same,  are  the  capacities 
of  the  different  conductors  ;  the  coefficients  qmia  in  which  the  indices 
differ  from  each  other,  are  the  coefficients  of  induction.  The  energy 
can  therefore  be  expressed  by  the  charges  as  well  as  by  the  potentials  ; 
in  the  former  case  it  is  represented  by  We,  in  the  latter  by  W^,. 

The  significance  of  the  coefficients  qlv  q^,  ...  is  shown  as  follows: 
If  Al  (Fig.  69)  is  an  insulated  conductor  having  the  charge  ev  and 
if  A^  Ay  etc.,  are  connected  with  the  earth,  we  have 

ei  =  fti*!.     e2  =  qlz^rv     e3  =  ql3¥v  ..., 

since  "*P2,  ^F3,  etc..  are  equal  to  zero.  The  coefficient  ^n  is  the 
capacity  of  the  conductor  Al  under  these  conditions.  Hence,  the  capacity 
of  a  conductor  is  the  quantity  of  electricity 
which  it  must  contain  in  order  that  its  potential 
shall  be  unity,  while  the  potential  of  all  other 
conductors  is  zero.  The  quantity  of  elec- 
tricity induced  on  the  conductors  A^  A3 
when  connected  with  the  earth  is  given 
by  <?12,  <?13,  ...,  if  A^  is  electrified  to  unit 
potential.  The  coefficients  <?12,  ql3  are 
negative.  The  lines  of  force  proceeding 
from  Al  may  either  pass  to  the  earth  or  terminate  on  the  conductors 
Ac,,  Ay  etc.  Since  a  positive  charge  is  present  at  the  points  at 
which  they  leave  Av  the  points  at  which  they  fall  upon  the  other 
conductors  must  have  a  negative  charge. 

On  the  other  hand,  the  coefficients  pl2,  pl3  ...pmn  are  positive.  If  the 
charges  of  the  conductors  A2,  A3,  .  .  .  are  e2  =  e3  =  e±  =  .  .  .  =  0,  we  have 
^i  ~P\\e\i  ^2  =Pi2ev  •••  •  -^s  many  lines  of  force  enter  the  uncharged 
conductors  A2,  A3  as  pass  out  from  them.  Since  lines  of  force  pass 
from  points  of  higher  to  points  of  lower  potential,  the  potential  of 
an  uncharged  conductor  in  an  electrical  field  cannot  be  a  maximum  ; 


150  ELECTROSTATICS.  [CHAP.  VH. 

it  lies  between  the  greatest  and  least  values  of  the  potential  in  the 
field.  If  the  conductor  Al  is  charged  with  the  unit  of  electricity, 
its  potential  is  pn.  At  a  point  infinitely  distant  the  potential  is  zero. 
Hence,  we  have  pu  >J912,  in  general  pnn  >pmn  and  pmm  >pnm.  Further, 
the  value  of  pnm  lies  between  pnn  and  zero,  and  since  pnn  is  positive, 
pnm  is  also  positive.  The  potentials  of  the  two  conductors  are  equal 
only  when  the  charged  conductor  encloses  the  uncharged  conductor. 
If  one  conductor  does  not  enclose  the  other,  we  will  always  have 

Pnn>Pmn     an(i    Pmm>Pmn- 


SECTION  LXI1I.    MECHANICAL  FORCES. 

Let  us  suppose  a  set  of  insulated  conductors ;  their  charges  will 
remain  unchanged  in  quantity  when  the  conductors  are  displaced. 
Their  potentials  depend  on  the  charges  in  the  manner  given  in  LXII. 
The  forces  acting  on  the  charged  surfaces  tend  to  set  the  conductors 
in  motion.  We  assume  that  all  the  conductors  except  Al  retain 
their  relative  positions;  that  Al  can  move  in  the  direction  of  the 
z-axis ;  and  we  then  determine  the  force  which  tends  to  move  Al  in 
this  direction.  Let  the  displacement  of  A^  be  8.1:  The  energy  We 
of  the  system  will  be  diminished  in  consequence  of  this  displacement 
by  X8x.  At  the  end  of  the  motion  the  energy  is  We  +  §Wn  and 
hence  we  have  W.-X.  8x=  W.  +  8JT.,  and  (a)  X=  -8J7f/8.v.  Now, 
from  LXII.  (g)  we  have 

(b)  X  =  |e128/>11/&c  +  %e2~8p22/8x  +  . . .  +  «1e28p12/Sj:  +  . . . , 

because  the  charges  do  not  change  during  the  motion,  and  are  inde- 
pendent of  the  displacement  8x.  This  method  may  be  always  applied 

if  the  motion  of  the  conductor  is. 
one  for  which  the  mechanical  work 
done  by  the  system  may  be  repre- 
sented in  the  form  X8x. 

We  now  determine  the  force 
with  which  one  of  the  conductors 
will  move  in  the  direction  of  the 
.r-axis  if  the  potentials  remain  con- 
stant. Let  Av  A2,  A3,  ...  be  the 
FIG.  70.  given  conductors  (Fig.  70).  They 

are  supposed  to  be  connected  by  very  thin  wires  with  the  very  large 
conductors  Bv  B2,  B3,  ...,  whose  potentials  are  ^f1,  ^o,  ^3,  ...  respec- 
tively, and  which  are  so  remote  from  the  system  of  conductors  A 


SECT.  Lxiii.j  MECHANICAL  FORCES.  151 

that  they  have  no  influence  upon  it  by  induction.  If  the  conductor 
Al  is  displaced  by  Sx,  the  charges  ev  e2,  es,  ...  increase  by  8elt  8e2,  8e3,  ..., 
and  we  have  from  LXII.  (h) 


The  electrical  energy  of  the  system  thus  increases  by 
SJr=  ^tej  +  ¥2&2  +  ¥3Se3  +  .  .  .  , 


,  +  .  .  .  . 

But  from  LXII.  (n)  the  energy  is  equal  to  W^,  +  Sf^  in  the  new 
configuration  of  the  system,  where 


The  work  done  is  JTSa;.     The  sum  of  the  energy  7F^,  originally  present 
and  the  energy  8fF  supplied  is  equal  to  the  sum  of  the  energy  in 
the  new  position  and  the  work  done.     We  therefore  have 
tf*  +  877=17^,  +  SIT*  +  X  .  8x,     X.8x  =  8fT  -  8!^. 
If  we  substitute  the  expressions  found  for  8JF  and  8JF^,,  we  have 


v  +  ^2^3^23  +  «.  . 

Hence,    we   obtain   (d)    X.8x  =  8JF^,    or    X=8fTyjSx,    and   further 
8JF=  '2X  .  8x. 

The  electrical  energy  supplied  to  the  system  is  therefore  twice  as  great 
as  the  mechanical  work  done.  Now,  if  during  the  displacement  of 
the  conductors  their  potentials  do  not  change,  energy  must  flow 
from  Blt  By  B3,  etc.,  to  the  conductors  Av  A%,  A3,  etc.  One-half 
of  the  energy  SJF  supplied  is  expended  in  doing  the  mechanical 
work,  the  other  half  in  increasing  the  electrical  energy. 


SECTION  LXIY.    THE  CONDENSER  AND  ELECTROMETER. 
1.  Parallel  Plates. 

If  two  bodies  at  different  potentials  are  placed  near  each  other, 
a  relatively  great  quantity  of  electricity  can  be  collected  on  the 
surfaces  which  face  each  other.  If  A  and  B  are  two  such  bodies, 


152  ELECTROSTATICS.  [CHAP.  vn. 

whose  potentials  are  ^  and  Mfg  respectively,  and  if  the  opposing 
surfaces  of  the  bodies  are  planes,  the  electrical  force  in  the  inter- 
vening space  is  everywhere  constant,  except  near  the  edges  of  the 
plane  surfaces.  If  a  represents  the  distance  between  the  planes, 
and  if  ^fl  is  greater  than  ^fy  so  that  this  force  is  directed  from 
A  to  B,  we  have  from  VII.  (c),  (a)  ¥1  =  ¥2  +  JFa  and  F=(¥1-¥2)la. 
The  surface-density  on  Al  is  determined  from 

471-0-  =  ^  or  <r  =  (*1-'4r2)/4ira. 

If  S  represents  the  surface  of  the  conductor  A  which  faces  B,  we 
have  for  the  charge  el  on  S,  (b)  e1  =  S<r  =  (>¥l-V2)  .  S/faa.  The 
charge  e2  on  B  is  equal  to  -  er  The  electrical  energy  W^  of  the 
system  is 


or  (c)  JT^  =  I/Sir.  (^-^^.S/a,  where  the  energy  is  expressed 
in  terms  of  the  potentials.  From  (b)  and  (c)  it  follows  that 

(d)  JFe  =  2irael2/S. 

If  the  z-axis  is  perpendicular  to  the  plane  surfaces  of  the  conductors, 
and  if  it  is  directed  from  A  to  B,  we  have,  representing  the  x-co- 
ordinate  of  the  plane  face  of  A  by  xv  and  that  of  the  plane  face 
of  B  by  x2,  a  =  x2-xl,  and 

W*  =  1/87T  .  ppj  -  ¥/  .  S/(x2  -  xj  ;   W.  =  2we^(x2  -  xJ/S. 
The  mechanical  force  which  acts  on  A  is  [LXIIL] 

Xl  =  -  SWJSa^  =  2^/S  ;  Xl  =  2™^  =  $Fev 
This  corresponds  to  LIX.  (e).     We  further  have  [LXIIL  (d)] 

X1  =  Sfr+ISXi  =  1/87T  .  (^  -  ¥2)2  .  S/(X2  -  Xrf-  =  1  1  Sir  .  F*  .  S. 

This  agrees  with  LIX.  (f).  From  the  expressions  which  have  been 
given  for  W^,,  we  also  have  W^,=  l/8-n-.  F2  .  S(x<,  -  xj,  which  agrees 
with  LXI.  (i). 

The  capacity  C  is  C=e1/(^rl-^r2)-)  if  ^2  =  0.  we  have  C=S/4ira. 

2.  Concentric  Spherical  Surfaces. 

If  a  sphere  Av  whose  radius  is  R,  is  enclosed  by  the  concentric 
spherical  shell,  whose  internal  and  external  radii  are  7?2  and  Rz, 
and  if  Al  is  given  the  charge  ev  and  A2  the  charge  e2,  the  inner 
surface  of 
the  charge 


SECT.  LXIV.]          CONDENSER  AND  ELECTROMETER.  153 

The  potentials  within  the  sphere  Al  and  within  the  spherical 
shell  As  are  therefore  "¥l  =  eJP^  -  el/R2  +  (el  +  e2)/E3  ;  ¥2  =  (e2  +  el)jR3  ; 
and  hence  ex  =  R^(RZ  -  RJ  .  ^  -  JR2.  Rlf(Ra  -  RJ  .  ¥/  and 


These  equations  agree  with  those  given  in  LXII.  (h).     The  potential 
^  in  the  space  between  the  two  spheres  is 


where  /•  is  the  distance  of  the  point  considered  from  the  common 
centre  of  the  spheres.  The  potential  outside  the  spherical  shell  is 
^P0  =  (e1  +  e2)/r.  The  capacity  C  of  the  inner  sphere  is  determined 
by  el  =  C^¥v  when  ¥2  is  set  equal  to  zero  ;  we  have  therefore  C  =  R^RJat 
if  we  represent  by  a  the  distance  between  the  surface  of  the  inner 
sphere  and  the  inner  surface  of  the  spherical  shell. 

3.  Coaxial  Cylinders. 

Suppose  two  coaxial  cylindrical  surfaces,  Al  and  A^  confronting 
each  other.  Let  their  potentials  be  ^  and  ^2,  and  their  radii  Rl 
and  R0  respectively.  Let  a  point  in  the  space  between  Al  and  A2 
be  at  the  distance  r  from  the  common  axis  of  the  cylinders.  The 
potential  ^f  must  satisfy  the  equation  V2xF  =  0  for  this  space.  Since 
the  equi-potential  surfaces  in  the  space  considered  are  cylindrical 
surfaces  coaxial  with  Alt  the  equation  V2^"  =  0  may  be  given  the 
form  [cf.  XV.]  d*P/dr2  +  2lr.dV/dr  =  (),  and  we  obtain  by  integration 
*  =  c  log  r  +  cr  For  r  =  R^  we  have  ¥  =  ¥lf  and  for  r  =  R^  ¥  =  ¥„, 
therefore 

*.  =  *!•  (log  R2  -  log  r)/(log  E2  -  log  RJ 
+  V,  .  (log  r  -  log  ^)/(log  R,  -  log  Rj). 
For  a  point  outside  the  outer  cylinder,  the  potential  is 


where  r  is  the  distance  of  the  point  considered  from  the  axis  of 
the  cylinder.  The  constant  c  cannot  be  determined  from  the 
potentials  alone.  The  electrical  force  F  in  the  intervening  space  is 
F  =  -  <t9tjdr  =  -  OP2  -  "¥"!)/  (log  R2  -  log  RJ  .  1/r.  The  surface-densities 
o-j  and  o-2  on  Al  and  A2  are  [LV.  (e)] 

2  -  log  RJ  . 

2  -  log  .Bj)  . 

The  charge  on  a  portion  of  the  cylinder  Av  whose  length  is  /,  is 
or   e1  =  //2.(Y1-'»P2)/(log^2-log^1).      The   charge   on 


154 


ELECTEOSTATICS. 


[CHAP.  vn. 


the  inner  surface  of  A.2  is  equal  and  opposite  to  this.     The  capacity 
C  of  a  portion  of  the  inner  cylinder,  whose  length  is  /,  is 


4.  Tlie  Quadrant  Electrometer. 

Suppose  AlAl  to  be  two  metal  plates  whose  potential  is  1Ir1,  and 
A2A2  to  be  two  similar  plates  whose  potential  is  ^2  (Fig.  71). 
In  the  middle  between  the  two  pairs  of  plates  there  is  placed  a 
plate  A3  whose  potential  is  ¥3.  We  assume  that  ^1<->P2<'»Ir3.  If 
A3  is  displaced  by  the  distance  8x  in  the  direction  from  Az  to  Av 
and  therefore  in  the  direction  of  its  length,  the  area  b8x  is 
displaced  from  right  to  left,  if  b  represents  the  width  of  the 


I 

1 

Jt 

Az 

\ 

plate  Ay  If  the  distance  of  As  from  Al  and  A2  is  a,  the  force 
acting  between  Al  and  A3  is  ^  =  ("*F3  —  "ty^/a,  and  that  acting  between 
A.2  and  ^43  is  ^  =  (¥3  -  ^2)/a,  on  the  assumption  that  the  points 
considered  do  not  lie  near  the  edges  of  As,  or  in  the  space  between 
Al  and  Ay  From  LXI.  (i)  the  electrical  energy  is  7F=l/87r.  \F-d-r. 
During  the  motion  considered,  the  energy  on  the  left  side  of  the  pair 
of  plates  is  increased  by 


that  on  the   right   side  is  diminished   by    l/8ir.  (M/3-'*J/2)2/a2.  2ab8x. 
The  gain  of  energy  is  therefore 


This  expression  does  not  fully  represent  the  gain  of  energy,  since 
no  account  is  taken  of  the  relations  at  the  edges.     We  therefore  set 


SECT,  LX.IV.]          CONDENSES  AND  ELECTEOMETEE. 


155 


The  force  X  which  tends  to  move  A3  from  right  to  left  is  then 
[LXIIL  (d)]  X=k(Vt-Vl).(Vi-Mp1  +  VJ).  We  may  apply  this 
result  to  the  quadrant  electrometer.  If,  in  this  instrument,  the 
movable  aluminium  plate  turns  through  the  angle  6,  we  may  set 
approximately 

6  =  aC*,  -  Yj) .  (^3  -  £C*-j  +  *2))» 

where  ^  and  ^  are  the  potentials  of  the  quadrants,  ¥3  is  the 
potential  of  the  aluminium  plate,  and  a  is  a  constant  whose  value 
depends  on  the  form  and  dimensions  of  the  apparatus. 


SECTION  LXV.    THE  DIELECTRIC. 

We  have  assumed  until  now  that  the  bodies  considered  were 
either  good  conductors  or  perfect  insulators,  on  which  the  charge 
was  immovable.  Experiment  shows,  however,  that  there  are  no 
perfect  insulators. 

Electricity  on  insulators  is  often  lost  by  conduction,  which  for 
the  most  part  is  due  to  the  film  of  fluid  deposited  on  them  by 
the  air.  But  even  if  this  film  is  removed  by  careful  drying,  con- 
duction still  persists.  If  a  charge  of  electricity  is  communicated 
to  one  part  of  an  insulator,  it  is  distributed  after  a  considerable 
time  in  the  insulator  in  the  same  way  as  it  would  be  in  a  good 
conductor.  Besides  this,  another  action  also  exists  which  is  instan- 
taneous. When  a  movable  insulator  is  brought  into  the  neighbour- 
hood of  a  charged  conductor,  the  insulator  sets  itself 
in  the  same  way  as  a  good  conductor,  from  which  it 
follows  that  an  instantaneous  distribution  of  electricity 
takes  place  in  it.  According  to  Faraday,  insulators  con- 
sist of  very  small  conductors  which  are  separated  by 
an  insulating  medium.  The  capacity  of  a  condenser 
is  increased  by  replacing  the  air  which  serves  as  the 
insulator  between  its  surfaces  by  other  insulators,  such 
as  glass,  shellac,  calc  spar,  etc. 

Let  A  and  B  (Fig.  72)  be  two  conducting  plates  which    (^  j 

are  separated  by  the  insulator  CD.  Let  A  be  brought 
to  the  potential  ^fv  and  B  to  the  potential  ^2;  let  the 
surface-density  on  A  be  a-,  that  on  B  is  then  —  <r. 

In  order  to  explain  this,  Faraday  assumed  that  a  peculiar  electri- 
fication exists  in  the  insulator  CD,  by  which  each  of  the  conducting 
particles  contained  in  it  acquires  negative  electricity  on  its  right 


C 


D 

FIG.  72. 


156  ELECTROSTATICS.  [CHAP.  vn. 

and  positive  electricity  on  its  left.  Just  as  a  mechanical  force  can 
give  rise  to  an  elastic  displacement,  the  forces  proceeding  from 
the  plates  of  the  condenser  produce  an  electromotive  action  by  setting 
up  a  current  of  electricity  in  the  particles  of  the  dielectric.  The 
positive  electricity  flows  in  the  particles  toward  the  left,  the  negative 
toward  the  right. 

By  this  process,  which  we  may  call  dielectric  displacement,  there 
arises  a  polarization  of  all  the  particles. 

The  condition  in  the  dielectric  can  be  compared  with  the  polarity 
of  the  particles  of  a  permanent  magnet. 

The  quantity  $)  which  flows  through  a  unit  of  area  parallel  to 
A  and  B  must  be  equal  to  a-.  A  unit  of  area  of  the  surface  of  the 
insulator  at  A  receives  the  charge  -a-,  and  that  at  B  the  charge  +  0-. 
Let  ft  be  the  distance  between  A  and  B,  whose  difference  of  potential 
is  ¥x  -  ¥2.  The  force  in  the  intervening  space  is  (a)  F=  (M^  -  ^.2)/a. 
If  the  quantity  of  electricity  £)  which  flows  through  the  unit  of  area 
is  proportional  to  the  force  acting  in  the  insulator,  we  can  set 
(b)  £)  =  #/47r..F, 

where  K  is  a  constant.  Hence,  the  surface  of  A,  whose  area  is  S, 
will  have  the  charge  (c)  S3)  =  K/±ir .  (¥\  -  ¥2)/a .  S.  A  comparison 
of  this  equation  with  LXIV.  (b)  shows  how  many  times  greater  the 
capacity  of  the  condenser  becomes  if  another  insulator  is  used  in 
place  of  air.  We  call  K  the  dielectric  constant ;  for  air,  which  is  chosen 
as  the  standard  medium,  we  set  K=  1. 

The  dielectric  constant  of  an  insulating  medium  is  the  ratio  of  the 
capacity  of  a  condenser  having  that  medium  as  an  insulator  to  the  capacity 
of  the  same  condenser  when  air  is  the  insulator.  It  is,  however,  more 
correct  to  set  K=  1  for  a  vacuum ;  it  has  been  shown  that  K  for 
gases  is  a  little  greater  than  1.  As  examples  of  the  values  of  the 
dielectric  constant,  we  have  for 

Glass,  -  #=5,83—6,34, 

Paraffin,     -        -         -  #=2—2,32, 
Sulphur,  -         -  #=3,84, 

Shellac,      -        -        -  #=3—3,7, 
Bi-sulphide  of  carbon,  #=2,6, 
Oil  of  turpentine,       -  #=2,2. 

On  the  whole,  the  results  obtained  by  different  observers  are  not 
very  consistent. 

Let  us,  as  in  LVIIL,  suppose  that  the  body  A  is  brought 
into  the  space  enclosed  by  the  metallic  shell  BC.  Suppose  that 


SECT.  LXV.]  THE  DIELECTEIC.  157 

B  receives  the  charge  -  Q,  and  C  the  charge  Q.  The  quantity 
Q  of  positive  electricity  flows  outward  through  the  closed  surface  7> 
taken  within  BC.  Hence,  when  the  quantity  Q  is  introduced  through 
the  closed  surface  D,  the  same  quantity  flows  out  through  the  same 
surface.  We  are  therefore  justified  in  assuming  that  the  quantity 
enclosed  by  the  surface  D  is  always  zero.  This  holds  for  the  closed 
surface  E,  which  may  be  drawn  in  the  insulator  surrounding  BC, 
and  thus  we  obtain  the  general  theorem  that  the  total  quantity  of 
electricity  contained  within  a  closed  surface  is  equal  to  zero. 

If  the  quantity  of  electricity  £>,  which  flows  through  a  unit  of  area 
perpendicular  to  the  direction  of  electrical  force,  is  proportional  to 
that  force  at  every  point,  we  will  have  (d)  £)  =  Kjkir .  F.  If  /,  g, 
and  h  are  the  quantities  which  pass  through  three  units  of  area  in 
an  isotropic  body  taken  perpendicular  to  the  three  coordinate  axes, 
that  is,  if  they  are  the  rectangular  components  of  the  displacement  £),  and 
if  X,  Y,  Z  are  the  components  of  the  electromotive  force  F,  we  have 

(e)  f=K/4:Tr. X,    g=K/4ir.Y,    h  =  K/4:Tr.Z.      These    expressions   are 
consistent  with  the  relations 

(f)  Z=-3*/3ar,     Y=  -3*"/3y,     Z= - 


SECTION  LXVL     CONDITIONS  OF  EQUILIBRIUM. 

Suppose  the  closed  surface  AS'  to  enclose  a  portion  of  the  electrical 
system,  passing  partly  through  the  dielectric,  partly  through  the 
conductors.  We  will  make  use  of  our  previous  conclusion,  that  the 
total  quantity  of  electricity  within  S  is  zero.  If  we  represent  by 
e  the  total  quantity  enclosed  by  S,  and  by  £)'  the  quantity  which 
flows  out  through  a  unit  of  area  in  consequence  of  the  displacement 
in  the  dielectric,  we  have  e—fo'.dS,  The  normal  to  the  surface 
S  directed  outward  makes  angles  with  the  axes  whose  cosines  are 
I,  TO,  7i.  We  have  ^)'  =  K/^ir.  Fcose,  if  e  is  the  angle  between  the 
normal  to  dS  and  the  direction  of  the  electromotive  force  F.  Since 
F  cos  e  =  XI  +  Ym  +  Zn,  we  obtain 

(b)  (c)         e  =  1/47T .  \K(Xl  +  Ym  +  Zn)dS  =  \(fl  +  gm  +  hn}dS. 

If  we  apply  equation  (c)  to  an  infinitely  small  parallelepiped,  whose 
edges  are  dx,  dy,  and  dz,  we  find  that  the  quantity  which  enters, 
it  is  fdydz  +  gdxdz  +  hdxdy,  and  that  which  leaves  it 

'/+     dydz  +   g  +     dydxdz  +   h  +     dz 


158  ELECTROSTATICS.  [CHAP.  vn. 

if  p  is  the  volume-density,  the  charge  contained  in  the  parallelepiped 
is  pdxdydz;  now  the  total  quantity  of  electricity  within  it  equals 
zero,  and  therefore 

fdydz  +  gdxdz  +  hdxdy  +  pdxdydz 

=  (f  +  ?fdx}dydz  +  (g  +  ^dy}dxdz  +  (h  +  fdz}dxdy. 
\      ox    J  \      oy    J  \       oz    J 

From  this  it  follows  that  the  volume-density  p  of  the  free  electricity 
within  the  body  [cf.  LXVIIL]  is  given  by  (d)  p  =  'd/J'd 
-and  from  LXV.  (e)  we  obtain 


In  terms  of  the  potential,  this  becomes 

<f  )     -o(K  .  3¥/az)/ae  +  -d(K.  3¥,%)/3y  +  3(K.  3¥/3*)/3,?  +  4rp  =  0. 

If  we  consider  a  surface  on  which  the  surface-density  is  o-,  we 
obtain  by  the  same  method  as  that  used  in  LIV.  (g)  a-  =  2)1  +  $)2,  if 
<£)l  and  £)2  are  the  polarizations  in  the  directions  of  the  normals 
to  the  surface  drawn  outward.  If  the  forces  along  the  same  normals 
are  NI  and  JV2,  we  have  (h)  a-  =  KJiir  .  ^  +  K^v  .  N.2,  where  K^ 
And  K.2  are  the  dielectric  constants  on  the  opposite  sides  of  the 
surface. 

By  means  of  this  equation  questions  on  electrical  distribution  and 
•on  the  relations  between  density  and  potential  may  be  solved.  If 
K  is  constant,  it  follows  from  (f  )  that  K  V2^  +  kirp  =  0.  In  a  region 
where  the  dielectric  constant  is  K,  the  potential  which  arises  from 
.a  given  charge  p  is  equal  to  only  the  Ktb  part  of  that  which  arises 
from  the  same  charge  in  a  region  where  K=  1.  In  the  latter  case 
ithe  potential  Hf'  is  determined  by  V2¥  '  +  4ny)  =  0,  so  that  ¥  =  V/K. 

The  electrical  force  is  also  diminished  in  the  same  ratio.     If  the 
•charges  el  and  e.2  are  placed  at  the  points  A  and  B  respectively,  and 
if  the  distance  AB  =  r,  the  charges  repel  each  other  with  the  force  E, 
B-l/K.qiJA 


SECTION  LXVII.    MECHANICAL  FORCE  AND  ELECTRICAL  ENERGY  IN 
THE  DIELECTRIC. 

Suppose  that  the  dielectric  constant  is  constant  in  the  region  con- 
sidered. The  forces  acting  in  the  directions  of  the  axes  on  the 
parallelepiped  dxdydz  =  dr,  in  which  the  density  is  p,  are 


SECT.  LXVII.]      MECHANICAL  FORCE  IN  DIELECTRIC.  159 

Using  LXVI.  (e),  we  have 

(X)  =  l/4ir(X .  -d(KX)l^x  +  X .  V(KY)rdy  +  X .  d(KZ)pz). 
Since    [LXV.    (f)]   the    forces    have    a    potential,   we    obtain    from 
XXVII.  (b) 

(X)  =  K/8ir .  (d(X*  -Y2-  Z2)/cx  +  2^(XY)py  +  23(A^)/3.?). 
The  force  which  acts  on  the  volume-element  dr  may  be  considered 
as  due  to  the  stresses  Xx,  Yy,  etc.  [cf.  XXVII.  (c)],  where 

f  Xx  =  K/STT  .  (X2  -  Y 2  -  Z2),     YZ=ZU=  A747T .  YZ, 

[  Zt  =  K/8ir  .  (Z2  -  A'2  -  72),      Xy  =  Yx  =  A747T  .  YX. 
If  the  direction  of  the  a;-axis  is  that  of  the  electrical  force,  we  have 

and  the  tangential  components  vanish.  In  the  direction  of  the  electro- 
motive force  F,  there  is  a  tension  S,  and  in  all  directions  perpen- 
dicular to  the  force  F,  a  pressure  T,  such  that  (d)  S=T=KF'2/8Tr. 

If  A  and  B  (Fig.  73)  are  two  conducting  surfaces  which  are 
separated  by  an  insulator  whose  dielectric  constant  is  K,  and  if  AB 
is  a  line  of  force,  a  tension  S  acts  along  that  line, 
and  a  pressure  T  =  S  acts  perpendicularly  to  it. 
A  surface-element  at  A  is  under  a  tension 

1/Sir.KF2, 

acting  in  the  direction  of  the  normal  drawn 
outward.  When  K=  I  this  reduces  to  the  result 
reached  in  LIX. 

For  example,  let  A  B  be  a  hollow  sphere  of  glass  whose  inner  and 
outer  radii  are  rl  and  r2  respectively.  Let  the  potential  of  the  surface 
A  be  ^v  and  that  of  B  be  0.  The  potential  in  the  interior  of  the 
spherical  shell  is  determined  from  LXVI.  (f ) ;  when  p  =  0,  we  have 
V2*P"  =  0.  Since  the  potential  depends  only  on  the  distance  r  from 
the  centre,  we  have  from  XV.  d*¥/dr*  +  2/r  .d^/dr=0,  and  hence 
M*1  =  A  +  Bfr.  Having  regard  to  the  boundary  conditions,  we  obtain 
W  =  rl(r2-r)/(r2-rl).yi/r.  The  forces  F1  and  F2  which  act  at  the 
inner  and  outer  surfaces  are 

Representing  the  stresses  on  these  surfaces  by  pl  and  pv  we  obtain 
These  stresses  may  be  regarded  as  pressures  which  act  on  the  surfaces. 


160  ELECTROSTATICS.  [CHAP,  vn 

If  d<f>/dr  represents  the  increment  of  r  which  results  from  the  pressure 
on  the  surface,  we  have  from  XXXI.  (h)  d<f>/dr  =  ^ar+b/r2,  where 
a  =  3/(3  A  +  2/0  .  (prf  -pff)/(r*  -  r*)  ;  b  =  1  /4/»  .  (Pl  -pjr*  .  r2*/(rs»  -  rf). 
From  this  it  follows  that 


If  we  set  K^^/Sir .  fjf j/[(*j  —  **i)(^23  ~  ri3)] =  -^r»  we  w^  have 

The  volume  contained  by  the  hollow  sphere  will  be  increased  by 
the  action  of  the  electrical  force. 

If  we  represent  by  60  the  increment  of  the  unit  of  volume,  we  have 
47T/3.(r  +  d<£/e?r)3  =  47r/3.r3(l+60),  so  that  00  =  3d(^/rdr.  For  the 
volume  of  the  sphere,  for  which  r  =  rv  we  obtain 

60  =  3(l/(3A  +  2/0  +  (r2* - r-*)i(r,  - rj .  l/iprflN. 

If  we  set  r2-  ^==8,  and  if  5  is  very  small  in  comparison  with  rv 
we  have  00  =  9JV.  (A  +  /*)//*(3A  +  2^),  where  H=K'¥l*/8ir.  1/332.  It 
thus  follows,  using  XXIX.  (d),  that  (e)  60  =  3/E8* .  K^/Sv.  The 
increase  of  volume  here  considered  has  been  observed  for  various 
condensers. 

If  a  region,  in  which  the  dielectric  constant  K  is  a  function  of 
the  coordinates,  contains  electrical  charges,  whose  density  is  p,  and 
which  give  rise  to  the  potential  ¥,  the  energy  W  is  determined 
as  in  LXI.  by  (f)  W=\\f&dr. 

If  p  is  expressed  in  terms  of  the  potential  by 

we  have  for  the  energy  W> 

JF=-1/87T. 


By  integration  by  parts,  we  obtain 

W=  1  1  Sir  .  J  J  f  tf((3¥/aB)2  +  Cd¥  pyy-  + 
where  the  integration  is  extended  over  the  entire  region,  and  it  is 
assumed  that  the  force  and  the  potential  vanish  at  infinity.  If  F 
represents  electrical  force,  we  have  (g)  JF=lj8ir. 


Electrical  Double  Sheets.  We  have  seen  in  LVIII.  that  two  conductors 
which  are  very  near  each  other,  and  are  kept  at  the  two  different 
potentials  ^ifl  and  W^  are  oppositely  charged.  The  surface-densities 
<r  and  -a-  of  their  charges  are  given  by  o-  =  (^r1  —  ¥2)/47r«,  where  a 


SECT.  LXVII.]     MECHANICAL   FORCE   IN   DIELECTRIC.  161 

is  the  distance  between  the  surfaces.  If  a  is  taken  infinitely  small, 
there  is  still  a  finite  difference  of  potential  between  the  surfaces  ; 
calling  this  difference  V,  we  have  (a)  a-  =  F/4?ra. 

Now,  there  are  several  ways  by  which  such  finite  potential  differ- 
ences may  be  established  across  a  surface  ;  for  example,  it  may  be 
done  by  friction  or,  what  amounts  to  the  same  thing,  by  contact. 
This  being  so,  we  must  necessarily  assume,  as  was  first  remarked  by 
v.  Helmholtz,  that  a  double  sheet  of  electricity  is  formed  on  the  two 
surfaces  which  are  near  each  other,  in  which,  if  the  distance  a  is 
extremely  small,  the  density  o-  must  be  extremely  great  if  the  potential 
difference  V  is  to  be  finite.  When  two  such  bodies  are  separated 
from  each  other,  provided  they  still  retain  the  electricity  thus  disposed 
on  them,  they  will  both  be  very  strongly  charged.  It  is  usually  not 
possible  to  separate  them  without  discharging  them,  but  if  one  or 
both  of  them  are  insulators  a  very  considerable  charge  remains. 
V.  Helmholtz  thus  explains  the  action  of  the  rubber  of  the  frictional 
electrical  machine. 

In  the  same  way  v.  Helmholtz  explained  several  remarkable  pheno- 
mena, for  example  the  phenomena  of  electrical  convection,  studied 
by  Quincke  and  Wiedemann.  Let  us  consider  a  capillary  tube, 
of  circular  cross-section,  whose  inner  radius  is  R  and  whose 
length  is  /.  A  liquid  is  supposed  to  be  flowing  through  this  tube. 
If  there  is  a  difference  of  potential  V  between  the  liquid  and  the 
wall  of  the  tube,  a  layer  of  electricity  forms  around  the  liquid,  whose 
density  a-  is  determined  by  equation  (a).  Setting  the  radius  of  this 
cylindrical  electrical  layer  equal  to  r,  so  that  a  =  E-r,  we  have  (b) 
o-  =  VI±ir(R  -  r).  Let  the  velocity  of  the  flow  at  the  place  at  which 
this  layer  is  present  be  w.  The  quantity  of  electricity  which  is  carried 
on  by  the  current  in  unit  time  through  any  cross-section  of  the  tube  is 


Now,  we  have  found  [XLIX.]  that 

')t  and  hence 


We  may  in  this  equation  set  E  =  r,  and  obtain 


(0 

where  S  is  the  area  of  the  cross-section  of  the  tube. 

If  the  pressure  at  the  two  ends  of  the  tube  is  the  same,  while  a 
difference   of  potential   ^  -  M*^   exists   between   them,   an   electrical 


162  ELECTROSTATICS.  [CHAP.  vn. 


_ 

force  F=  —  ~  —  2  will  act  within  the  tube,  and  therefore  a  force  Fa- 

will  act  on  every  unit  of  area  of  the  electrical  layer.  The  liquid  will 
thus  be  set  in  motion.  The  velocity  increases  from  the  wall  of  the 
tube,  where  it  is  0,  to  the  layer  a-,  where  it  may  be  called  u.  By 

the  definition  of  internal  friction  [XLVIL],  we  then  have  F<r  =  p.~. 

If  or  is  expressed  in  terms  of  the  potential  difference,  we  have 

VF 
(d)  u  =  .  — 


CHAPTER  VIII. 

MAGNETISM. 
SECTION  LXVIII.    GENERAL  PROPERTIES  OF  MAGNETS. 

IT  was  very  early  known  to  the  Greeks  that  in  the  city  of  Magnesia, 
in  Asia  Minor,  stones  were  to  be  found  which  had  the  power  of  attract- 
ing iron.  If  such  a  stone,  which  consisted  mostly  of  magnetic  oxide 
of  iron,  were  thrown  into  iron  filings,  they  would  adhere  with  special 
strength  to  it  at  certain  points.  A  long  magnet  is  especially  active 
in  the  vicinity  of  its  ends,  which  are  called  poles.  If  a  bar  magnet  is 
suspended  horizontally,  so  that  it  can  turn  about  a  vertical  axis  passing 
through  its  centre,  it  assumes  of  itself  a  definite  direction  which 
approximately  coincides  with  the  meridian  of  the  place.  The  end 
of  the  bar  which  points  toward  the  north  is  called  the  north  pole,  the 
other  end  the  south  pole.  Following  the  indication  of  this  experiment, 
we  assume  the  existence  of  two  magnetic  fluids,  which  are  separated 
in  each  particle  of  iron  by  the  influence  of  a  magnetic  force.  If  a 
bar  magnet  is  floated  on  a  fluid  at  rest,  it  assumes  a  definite  direction 
when  exposed  to  the  influence  of  a  magnetic  force,  but  the  force 
does  not  set  the  floating  magnet  in  motion  if  the  dimensions  of  the 
magnet  are  very  small  in  comparison  with  its  distance  from  the  seat 
of  the  magnetic  force.  We  conclude  from  this  that  the  north  and 
south  magnetic  fluids  are  present  in  a  magnet  in  equal  quantities. 
Let  us  represent  the  quantity  of  one  fluid  by  +  m,  that  of  the  other 
by  —  m,  the  north  magnetism  being  conventionally  taken  as  positive. 
Coulomb  proved  that  the  poles  of  two  magnets  repel  each  other  with 
a  force  F,  which  is  given  by  (a)  F^m^^r^  where  m1  and  m2  are 
the  quantities  of  magnetism  at  the  poles  and  r  is  the  distance 
between  the  poles. 

If  a  magnet  is  broken  into  many  small  pieces,  each  part  is  still  a 
163 


164 


MAGNETISM. 


[CHAP.  vni. 


magnet.  Because  of  this  fact,  we  conclude  that  every  magnet  is 
made  up  of  a  very  great  number  of  very  small  magnets. 

If  a  magnet  is  broken,  positive  magnetism  appears  on  one,  and 
negative  magnetism  on  the  other,  of  the  surfaces  formed  by  the 
fracture,  and  their  quantities  are  equal  unless  the  magnetization  is 
changed  by  the  jar  given  to  the  magnet  when  it  is  broken.  We  will 
represent  by  +  0-  the  quantity  of  magnetism  that  is  present  on  a  unit 
of  area  of  one  of  the  newly  formed  faces.  For  each  point  on  this 
face  a-  has  a  definite  value,  dependent  on  the  position  of  the  point 
on  the  face  and  of  the  face  in  the  original  magnet.  Construct  a  normal 
to  this  face  drawn  outward ;  <r  then  depends  on  the  coordinates  x,  y,  z 
of  the  point  and  on  the  direction  of  the  normal,  which  makes  angles 
with  the  axes  whose  cosines  are  /,  m,  n.  Let  ABC  =  dS  (Fig.  74) 
be  an  element  of  the  positive  face 
and  0  a  point  in  the  magnet  infinitely 
near  it.  A,  B,  C  are  points  in  which 
the  surface  dS  is  met  by  the  lines 
Ox,  Oy,  Oz  parallel  to  the  axes  drawn 
through  0.  The  surface  OBC,  one  of 
the  surfaces  of  the  tetrahedron  OABC, 
may  be  considered  as  a  negative  face. 
Suppose  that  the  magnet  is  mag- 
netized with  the  components  A,  B, 
C  in  the  directions  OA,  OB,  and 
OC,  the  surface  ABC  of  the  tetrahedron  exhibits  positive  magnetism, 
the  surfaces  OBC,  etc.,  negative  magnetism.  Represent  the  surface- 
density  perpendicular  to  the  z-axis  by  A ;  each  unit  of  surface  of 
OBC  then  exhibits  the  quantity  of  magnetism  -  A.  The  unit  of 
surface  of  OA  C  and  OB  A  will,  in  a  similar  notation,  exhibit  the 
quantities  of  magnetism  -  B  and  -  C  respectively.  The  position  of 
the  surface  ABC  is  determined  by  the  cosines  /,  m,  n  of  the  angles 
which  the  normal  to  the  surface  makes  with  the  axes ;  we  have 
OBC  =  l.dS,  OAC=m.  dS,  OB  A  =n.dS.  If  h  represents  the  perpen- 
dicular let  fall  from  0  to  the  surface  ABC,  and  if  the  quantity  of 
magnetism  in  the  unit  of  volume  is  represented  by  p,  the  total  quantity 
of  magnetism  contained  in  the  tetrahedron  is 

(<r-lA-mB-  nC)dS+  ^hpdS. 

Now,  since  the  total  quantity  of  magnetism  in  a  magnet  is  zero,  and 
the  altitude  h  of  the  tetrahedron  is  assumed  to  be  infinitely  small,  we 
have  (b)  o-  =  Al  +  Bm  +  Cn.  Hence,  if  the  surface-density  on  three 
perpendicular  surface-elements  passed  through  a  point  is  known,  the 


FIG.  74. 


SECT.  LXVIII.] 


PROPERTIES  OF  MAGNETS. 


165 


density  on  any  other  surface  passed  through  the  same  point  is  deter- 
mined from  (b).  If  we  set 

(c)  J*  =  A*  +  B*  +  C*  and   A=JX,    B  =  Jp.,    C  =  Jv, 

where  A2  +  /*2  +  v-=  1,  it  follows  from  (b)  that  a-  =  J(l\  +  mp,  +  nv).  J 
is  the  intensity  or  strength  of  magnetization.  The  direction  of  the  inten- 
sity makes  angles  with  the  coordinate  axes,  whose  cosines  are  A.,  /*,  v. 

If  we  set  l\  +  mp.  +  nv  =  cos  e,  where  e  is  the  angle  between  the 
intensity  of  magnetization  and  the  normal  to  the  surface-element, 
we  have  0-  =  J  cose,  that  is,  a-  is  the  component  of  the  intensity  of 
magnetization  along  the  normal  to  the  surface-element.  The  greatest 
value  of  o-  is  reached  when  «  =  0,  that  is,  when  the  direction  of  the 
intensity  of  magnetization  coincides  with  the  normal  to  the  surface- 
element  on  which  the  density  is  o-.  A  surface  may  be  passed  through 
any  point  in  a  magnet  for  which  the  surface-density  is  a  maximum. 
The  direction  of  the  intensity  of  magnetization  /  lies  in  the  normal 
to  this  surface.  For  such  a  surface,  whose  normal  coincides  with  the 
direction  of  magnetization,  cr  =  t/;  that  is,  the  quantity  of  magnetism 
on  the  unit  of  area  of  this  surface  is  /.  We  construct  a  parallelepiped, 
one  of  whose  ends,  dS,  lies  in  the  surface  considered,  and  whose  edges 
perpendicular  to  the  surface  are  ds ;  J.dS.ds  is  called  the  magnetic 
moment  of  this  parallelepiped.  The  intensity  of  magnetization  J  is 
equal  to  the  ratio  of  the  magnetic  moment  of  the  magnet  to  its  volume. 

The  magnetic  condition  of  a  magnet  is  defined  by  the  components 
of  magnetization  A,  B,  C.  The  density  o-  on  the  surface  of  the  magnet 
is  determined  by  (b)  from  these 
components.  Free  magnetism 
may  be  present  in  the  interior 
of  the  magnet  also.  If  00'  (Fig. 
75)  is  a  rectangular  parallel- 
epiped, whose  edges  are  dx,  dy, 
dz,  taken  within  the  magnet,  and 
if  A,  B,  C  are  the  components 
of  the  intensity  of  magnetization 
at  the  point  0,  OA'  will  con- 
tain the  quantity  of  magnetism 
-Adydz,  and  O'A  the  quantity 

(A  +  'dA/'dx .  dx)dydz.  Analogous  expressions  hold  for  the  other 
surfaces.  Representing  by  p  the  quantity  of  magnetism  contained 
in  the  unit  of  volume,  we  have 

('dA/'dx  +  "dBfdy  +  'dC/'dz  +  p)dxdydz  =  0, 


FIG.  75. 


166  MAGNETISM.  [CHAP.  vin. 

since  the  total  magnetism  in  the  parallelepiped  is  zero.  Hence  (e) 
P=  -(dA/'dx  +  'dBfdy  +  'dCfdz).  While  we  may  consider  A,  B,  C  as 
the  natural  and  direct  expressions  for  the  condition  of  a  magnet,  we 
determine  its  free  magnetism  from  the  derived  magnitudes  p  and  or. 


SECTION  LXIX.    THE  MAGNETIC  POTENTIAL. 

The  force  with  which  a  magnet  acts  on  a  pole  at  which  unit 
quantity  of  magnetism  is  concentrated  is  called  the  magnetic  force. 
Its  components  are  represented  by  a,  ft,  y.  These  are  determined 
from  the  potential  in  the  same  way  as  the  components  of  the  electrical 
force.  Let  the  quantities  of  magnetism  m,  m',  m",  ...  be  present  at 
given  points;  let  the  pole  P,  for  which  the  potential  is  to  be  deter- 
mined, be  distant  r,  r',  r",  ...  respectively  from  those  points.  The 
potential  is  then  given  by 

F=m/r  +  m'/r'+m"/r"+...,  and  m  +  m'  +  m"+  ...  =  0. 
If  N  is  a  north  pole  with  magnetism  +m  (Fig.  76)  and  S  a  south 
pole  with  magnetism  -  m,  and  if  P  is  the  point  for  which  the  potential 
^p  is   to  be  determined,  and  whose  distances 
from  the  north  and  south  poles  are  r  and 
fS/  r   respectively,  we  have 

V=  m/r  -  m/r'  =  m(r  -  r)/rr'. 
If  the  length  /  of  the  magnet  is  very  small 
in  comparison  with  r  and  /,  and  if  we  set 
2,  we  have  (a)  r= 


90}  =  lm  is  called  the  moment  of  the  magnet  ; 

SNM  is  called  the  magnetic  axis.     Its  positive  direction  is  that  from 
the  south  to  the  north  pole. 

We  now  determine  the  potential  of  a  magnet  whose  components  of 
magnetization  A,  B,  and  C  are  given.  Let  the  coordinates  with  respect 
to  any  point  taken  as  origin  be  £,  ?/,  £;  A,  B,  C  are  then  functions 
of  these  three  coordinates.  A  parallelepiped  whose  edges  are  d£, 
drj,  d£  will  have  the  magnetic  moment  Adijd{.d£,  if,  for  the  present, 
we  consider  only  the  magnetization  determined  by  A.  If  the  co- 
ordinates of  the  point  P,  for  which  the  potential  is  to  be  determined, 
are  x,  y,  z,  and  if  its  distance  from  the  point  £,  77,  £  is  r,  we  have 

r2  =  (x  -  £)2  +  (y  -  rj)2  +  (z-  C)2,    cos  9  =  (x  -  £)/>-. 

The  potential  due  to  the  element  d%  .  d^  .  d£  arising  from  the  com- 
ponent of  magnetization  A  is,  from  (a),  Ad£dr]d£/r2  .  (x  -  £)>.    B  and  C 


SECT.  LXIX.]  THE    MAGNETIC   POTENTIAL.  167 

give  rise  to  the  potentials  Bdgdrjdflr2 .  (y  -  rf)jr,  Cdgdrjdflr2 .  (z  -  £)/r. 
The  sum  of  these  three  potentials,  integrated  over  the  whole  magnet, 
gives  the  total  potential  V, 

(b)  V=  \\\\A(x  -  £)  +  B(y  -  r,)  +  G(z  -  fl^W- 
Since  r'dr/'dx  =  x-g,   rdrj'dy  =  y—%   rdrfdz  =  z-£,  we  have 

(c)  V=  -  \\\(A  .3(l/r)/3a;  +  £.3(l 
If  we  set 

^  =  Jff^/r.d£eMt  *,=  \\lBfr. 
we  have  (d)  F"=  -(3</'1/3a;  +  3i/'2/3y  +  3^3/32),  since  the  components  of 
magnetization  ^4,  -6,  (7  are  independent  of  the  coordinates  x,  y,  z  of 
the  point  P. 

Another  more  general  transformation  may  be  made  by  help  of  the 
equations,  r .  3r/3£  =  -  (x  -  £),  r .  3r/3?/  =  -  (y  - 17),  r .  3r/3£=  -  («  -  0, 
by  the  use  of  which  we  obtain  from  (b), 

(e)  V=  \\\[A  .  3(l/r)/3£  +  B .  3(l/r)/3,  +  (7.  3(l/r)/3f]^^df. 

Let  the  normal  to  the  surface-element  make  angles  with  the  axes 
whose  cosines  are  /,  m,  n.  By  integration  by  parts,  it  follows  that 

(f)  V=  \\(Al  +  Bm  +  Cn)/r .  dS  -  J  J J(3^/3^  +  3£/3ry  +  3(7/30 .  dgdr)d{/r, 
and  using  LXVIII.  (b)  and  (e),  (f)  becomes 

(g)  r=JforfS/r+Jffpd^<*0r. 

The  correctness  of  the  last  equation  is  immediately  evident  from 
the  meaning  of  <r  and  p. 

The  components  a,  /3,  y  of  the  magnetic  force  are  expressed,  as 
in  the  theory  of  electricity,  by 

(h)  a=-3F/3z,   /3=-3F/3z/,    y=-3F"/3z. 

The  force  N,  which  acts  in  the  direction  of  the  element  of  length  dv, 
is  JV=  -3F/3^.  If  the  potential  is  F]  inside  the  magnet  and  Va 
outside  the  magnet,  we  have  at  the  surface  Vi=Va.  If  vt  is  the 
normal  to  a  surface-element  on  which  the  density  is  cr  drawn  into 
the  magnet,  and  va  the  normal  to  the  same  element  drawn  outward, 
we  have  from  the  general  laws  of  the  potential  [cf.  XIV.  (1)],  which 
are  applicable  here,  (i)  3Fi/3vi  +  3F0/3v(l+  47r<r  =  0.  For  every  point 
within  the  magnet,  we  have  (k)  V2F|-  +  top  =  0,  or  introducing  the  value 
for  P  given  in  LXVIII.  (e),  (1)  V2Fi  =  ^(dA/'dx  +  dB/'dy+aC/^z),  if 
p  and  A,  J5,  C  are  functions  of  x,  y,  and  z.  Outside  the  magnet  we 
have,  on  the  other  hand,  (m)  V2F"ra  =  0.  If  S  is  a  closed  surface,  v 
its  normal  drawn  outward,  and  M  the  total  magnetism  enclosed  by 


168  MAGNETISM.  [CHAP.  vm. 

the  surface,  we  have  from  XIV.  (c)  4irM  =  -  JJdF/3v .  dS,  or  designating 
by  ^)n,  the  magnetic  force  in  the  direction  of  the  normal  to  the  surface, 

(n)  4irM=l\$ndS. 

The  equations  (i)  and  (k)  are  special  cases  of  this  equation. 

SECTION  LXX.    THE  POTENTIAL  OF  A  MAGNETIZED  SPHERE. 

If  the  components  of  magnetization  are  given  functions  of  the 
coordinates,  and  £,  77,  £  are  the  coordinates  of  a  point  within  the 
magnet,  we  obtain  the  potential  most  easily  by  using  the  formula 
LXIX.  (d).  If  A,  B,  C  are  constant,  the  problem  is  to  determine  the 
potential  of  a  body  of  constant  volume-density.  Hence  we  set 

(a)  t 

and  obtain  (b)  V=  -  (A  .  -d^fdx  +  B .  3</y  cty  +  C .  tyfiz).  The  potential 
of  a  sphere  whose  components  of  magnetization  are  A,  B,  C  is  to  be 
determined  by  means  of  this  equation.  Take  the  origin  of  the  system 
of  coordinates  at  the  centre  of  the  sphere.  The  potential  $  has 
different  values,  according  as  the  point  for  which  the  potential  is  to 
be  determined  lies  inside  or  outside  the  sphere.  In  the  usual  notation 
we  have,  from  XIII.  (c)  and  (d),  ^  =  47r^/3r;  i}>t  =  2ir(It2-r2/3), 
where  R  is  the  radius  of  the  sphere. 

If  we  represent  the  magnetic  potential  for  points  outside  the  sphere 
by  ^a>  an(l  for  points  inside  it  by  Viy  we  have 
(c),  (d)      Va  =  47T/3 .  ^/r3 .  (Ax  +  By  +  Cz)  •  Vi  =  47r/3 .  (Ax  +  By  +  Cz). 

Let  /  be  the  intensity  of  magnetization,  and  let  its  direction  make 
angles  with  the  axes  whose  cosines  are  A,  /*,  v.  Let  6  be  the  angle 
between  the  direction  of  J  and  the  line  r.  We  then  have 

Xx/r  +  fjiy/r  +  vz/'r  =  cos  0 

and  (e)  Va  =  47r/3 .  1PJ cos 0/r2 ;    F;  =  47r/3./rcose. 

Hence,  the  potential  outside  the  magnet  is  the  same  as  that  which 
is  set  up  by  an  infinitely  small  magnet,  whose  magnetic  moment  is 
3K  =  4ir/3.W  [LXIX  (a)]. 

If  the  a-axis  lies  in  the  direction  of  magnetization,  the  potential  in 
the  interior  of  the  magnet  is  Vi  =  4?r/3 .  Jx.  The  magnetic  force  ^) 
inside  the  sphere  is  therefore  constant,  and  is  expressed  by 

(f)  $=- -Air/3.  J. 

Outside  the  sphere  we  divide  the  force  into  two  components,  one  of 


SECT.  LXX.]   POTENTIAL  OF  A   MAGNETIZED  SPHERE.  169 

which,  P,  acts  in  the  direction  of  the  line  r,  the  other,  Q,  perpendicularly 
to  that  line.  We  then  have  P=  -Vrjdr,  Q=  -l/r  .'dTJ'dQ,  or 

P  =  87T/3 .  Wcos  0/7-3  =  2%R/r3 .  cos  0 ; 
Q  =  47T/3 .  E3Jsin  0/r3  =  Wfr3 .  sin  0. 

From  LXVIII.  the  surface-density  is  determined  by  cr  =  /cos9.  The 
resultant  forced  is  F^^ft/r3.^/!  +  3cos26.  We  have  further  tgQ  =  2Q/P. 
If  </>  is  the  angle  between  the  direction  of  the  force  F  and  the  direc- 
tion of  r,  we  have  tg<f)  =  Q/P  =  \tgQ. 

SECTION  LXXI.    THE  FORCES  WHICH  ACT  ON  A  MAGNET. 

Let  us  suppose  that  the  magnetic  forces  of  a  magnet,  whose  com- 
ponents we  may  represent  by  a,  /?,  y,  are  functions  of  the  coordinates. 
Let  us  suppose  also  another  magnet, 
whose  components  of  magnetization  are 
A,  B,  C.  its  action  on  the  first  magnet 
is  to  be  determined.  Consider  the  in- 
finitely small  parallelepiped  00'  within 
the  second  magnet  (Fig.  77) ;  on  the 
face  OA'  there  is  a  quantity  of  magnet- 
ism present  equal  to  -  Adydz,  which  is 
acted  on  by  the  force  -  Adydza  in  the  '  Flo 

direction  of  the  positive  z-axis.     The  face 

AO',  on  which  is  the  quantity  of  magnetism  Adydz,  is  acted  on  by 
the  force  (a  +  'da/'dx.dx)Adydz  in  the  same  direction.  The  resultant 
of  these  two  forces  is  A .  cta/dx .  dxdydz.  On  the  surface-element 
OB'  there  is  the  quantity  of  magnetism  -  Bdxdz,  and  on  BO'  the 
quantity  Bdxdz.  The  former  is  acted  on  by  the  force  —Bdxdz.a  in 
the  direction  of  the  positive  .r-axis,  and  the  latter  by  the  force 
+  B(a  +  'da/'dy .  dy)dxdz  in  the  same  direction.  The  resultant  of  these 
two  forces  is  B .  'da/'dy .  dxdydz.  For  the  surface-elements  00'  and  O'C, 
we  obtain  the  resultant  C.  'daj'dz.  dxdydz.  We  form  the  sum  of  these 
three  resultants,  integrate  over  the  whole  volume  occupied  by  the 
magnet,  and  obtain  for  the  force  X,  which  tends  to  move  the  magnet 
in  the  direction  of  the  z-axis, 

(a)  X=  \H(A  .  -dafdx  +  B .  3a/3y  +  C .  Va/c)z)dxdydz. 

Analogous  expressions  hold  for  the  forces  Y  and  Z. 

If  the  magnetic   force   whose   components   are   a,  /?,   y  is  due  to 
a  system  of  magnets  which  give  rise  to  the  potential  Fat  the  point 


170  MAGNETISM.  [CHAP.  vm. 


z,  y,  z,  we  have  a  =  -3F/3x,  /3=  -3F/3y,  7=  -^Vfdz.  We  then 
have  also  9a/3y  =  'dfi/'dx,  "dafdz  =  7tyfdx,  and  hence 

(b)  X=  \\\(A  .  -dafdx  +  B  .  Vpfix  +  C  .  Vy/-dx)dxd>/dz. 

We  now  determine  the  moment  of  the  forces  which  tend  to  turn 
the  magnet  about  one  of  the  coordinate  axis,  say  the  z-axis,  on  the 
assumption  that  the  magnetic  forces  are  constant.  Let  the  coordinates 
of  the  point  0  (Fig.  77)  be  x,  y,  z.  The  force  which  acts  on  the  surface 
BO'  in  the  direction  of  the  z-axis  has  a  moment  with  respect  to  the 
x-axis  equal  to  Bdxdz  .  y.  (y  +  dy).  The  force  acting  on  OB'  has 
a  moment  with  respect  to  the  same  axis  equal  to  -  Bdxdz  .y.y. 
Neglecting  small  terms  of  higher  order,  the  resultant  moment  is 
Bydxdz  ,  dy.  The  forces  acting  on  the  surfaces  O'C  and  OC'  give 
rise  to  the  moment  -  Cftdxdy  .  dz.  The  moment  L,  which  tends  to 
turn  the  magnet  about  the  z-axis,  is  therefore 

(c)  L  =  \\\(By-C($)dxdydz. 

The  moments  of  rotation  M  and  N,  with  respect  to  the  two  other 
axes,  are  determined  from  analogous  expressions. 

If  the  magnet  is  subjected  to  the  action  of  the  earth's  magnetism 
only,  the  magnetic  force  may  be  considered  as  constant  both  in 
magnitude  and  direction.  The  components  a,  (3,  y  are  then  inde- 
pendent of  x,  y,  z,  and  therefore  X=F=Z=Q.  The  centre  of  gravity 
of  a  magnet  does  not  move  under  the  action  of  the  earth's  magnetism. 
The  magnet  is,  however,  acted  on  by  a  moment  of  rotation,  which 
may  be  determined  in  the  following  way  : 

Let  the  magnetic  moment  of  the  magnet  be  9)?,  and  suppose  its 
direction  with  respect  to  the  coordinate  axes  to  be  determined  by 
the  angles  whose  cosines  are  I,  m,  n.  We  then  have 

Wl  =  \\\A.dT,    mm  =  \\\B.dr,    Wn  =  \\lC.dr, 

and  hence  L  =  Wl(ym  -  /3n)  ;  M=$tt(an  -  yl)  ;  N=  Wl(/3l  -  am).  From 
these  equations,  it  follows  that  La  +  Mtf  +  Ny  =  0  and  LI  +  Mm  +  Nn  =  0, 
that  is,  the  resultant  moment  is  perpendicular  to  the  magnetic  force  and 
also  to  the  magnetic  axis  of  the  magnet.  If  the  direction  of  the  force 
is  parallel  to  the  or-axis,  and  if  the  magnetic  axis  lies  in  the  xy-  plane 
and  forms  with  the  ,T-axis  the  angle  6,  we  will  have 

(d)  Z  =  0,    M=Q,    N=  -2tta.sine. 

If  a  magnet  can  turn  about  a  vertical  axis,  the  moment  which  tends 
to  increase  the  angle  0  between  the  magnetic  axis  and  the  magnetic 
meridian  is  -~9RlT.dll  6,  where  H  denotes  the  horizontal  component 


SECT.  LXXI.] 


FORCES   ON   A   MAGNET. 


171 


of  the  earth's  magnetism.      If  w  is  the  angular  velocity  of  the  magnet 
and  /  its  moment  of  inertia,  we  have,  from  XXII.  (c), 

d(Ja>)  =  -  WH .s'mQ.dt,    or  since    w  =  dQ/dt  =  0, 
(e)  je=  -WlH.sinQ. 

If  the  angle  6  is  very  small,  the  period  of  oscillation  of  the  magnet 
is  given  by  XXII.  (e),  (f)  T 


SECTION  LXXII.    POTENTIAL  ENERGY  OF  A  MAGNET. 

By  the  potential  energy  of  a  magnet  is  meant  the  work  which  is 
needed  to  transfer  the  magnet  from  a  position  in  which  no  magnetic 
forces  act  on  it  to  the  position 
in  which  the  magnetic  potential 
is  V.  We  will  first  consider  an 
infinitely  small  parallelepiped 
(Fig.  78),  whose  components  of 
magnetization  are  A,  B,  C.  In 
order  to  bring  the  magnetic 
surface  OA'  to  the  position  in 
which  the  potential  is  V,  the 
work  -  A  .  dydz .  V  must  be 
done.  The  opposite  surface  O'A 
is  brought  to  the  position  in 

which  the  potential  is  F'+'dV/'dx.dx,  and  the  work  done  on  it  is 
A .  dydz.  (P'+'dT/'dx.dx).  The  work  done  on  these  two  surfaces 
therefore  amounts  to  A .  'dVj'dx..  dxdydz.  If  we  obtain  in  like 
manner  the  work  done  on  the  two  other  pairs  of  surfaces,  we 
find  that  the  whole  work  W  done  in  transporting  the  magnet  is 

(a)  W=  \\\(A.^  Vfdx  +B.3  Vj'dy  +  C .  9  Vfdz)dxdydz, 

or  since  a,  /?,  y,  the  components  of  the  magnetic  force,  are 


FlG  78 


we  have  (b)  JF=  -\\\(Aa  +  B(3  +  Cy)dxdydz.  We  will  apply  this 
equation  to  the  case  of  a  magnet  subjected  to  the  action  of  the  earth's 
magnetism  only.  Let  its  magnetic  moment  be  3ft,  and  let  the 
direction  of  its  magnetic  axis  make  angles  with  the  coordinate  axes 
whose  cosines  are  /,  m,  n.  We  then  have 

.  dxdydz  =  m,    \\\B  .dxdydz  =  nm,    \\\C  .dxdydz  =  nW, 


172  MAGNETISM.  [CHAP.  vui. 

Representing  the  magnetic  force  by  §,  and  supposing  its  direction 
to  make  angles  with  the  coordinate  axes  whose  cosines  are  A,  p,  i>, 
we  obtain  W=  -  Wl^lX  +  mp  +  nv).  Letting  6  represent  the  angle 
between  the  magnetic  axis  of  the  magnet  and  the  magnetic  force, 
we  have  (c)  W=  -  3J?^) .  cos0.  If  the  direction  of  the  force  is 
parallel  to  the  z-axis,  as  in  LXXf.,  and  if  the  magnetic  axis  lies 
in  the  ay-plane,  the  work  done  in  turning  the  magnet  through  the 
angle  dQ  is  dW=  +  9tf£  .  sin  6 .  dQ.  This  agrees  with  LXXI.  (d). 

We  will  now  consider  a  very  small  magnet  situated  near  a  very 
strong  magnet.  If  the  small  magnet  has  sufficient  freedom  of  motion, 
it  will  turn  so  that  its  magnetic  axis  is  parallel  to  the  direction  of 
the  magnetic  force.  In  this  case  we  have  6  =  0  and  its  potential 
energy  W  is  W=  -  3ft$. 

Since  the  motion  of  the  small  magnet  involves  the  loss  of  potential 
energy,  it  moves  in  such  a  way  that  W  diminishes.  This  occurs  by 
the  last  equation,  when  ^)  increases ;  the  magnet  therefore  moves  in 
the  direction  in  which  the  magnetic  force  increases.  A  particle  of 
a  paramagnetic  substance  therefore  tends  to  move  towards  the  place 
where  the  magnetic  force  is  greatest.  On  the  other  hand,  diamagnetic 
bodies  move  toward  the  place  where  the  magnetic  force  is  a 
minimum. 

In  order  to  find  the  magnetic  energy  residing  in  a  system  of  magnets, 
we  proceed  in  the  following  way :  The  potential  at  every  point  within 
the  system  varies  proportionally  with  the  values  of  the  components 
of  magnetization.  We  assume  that  the  components  of  magnetization 
change  only  in  such  a  way  that,  in  successive  instants,  they  always 
increase  by  the  same  fraction  of  their  final  values.  On  these  con- 
ditions the  potential  increases  in  the  same  proportion.  If  the  com- 
ponents of  magnetization  arc  originally  zero,  the  potential  is  also 
originally  equal  to  zero.  Let  the  final  values  of  the  components  of 
magnetization  be  A,  B,  C.  At  a  particular  instant  during  the  increase 
of  the  components  of  magnetization,  let  these  be  represented  by  nA, 
nB,  nC,  where  n  is  a  proper  fraction.  At  the  same  time  the  potential 
at  any  point  is  equal  to  nV.  If  the  components  of  magnetization 
increase  by  A.dn,  B.dn,  G.dn  respectively,  the  potential  at  the 
point  considered  increases  by  Fdn.  If  A,  B,  C  increase  by  A  .  dn, 
B.dn,  G.dn  respectively,  the  work  needed  to  accomplish  this  is, 

\>y  (a), 

\\\(A .  dn .  ridF/dx  +  B.dn.  ndVj'dy  +C.dn.  n 
=  n.dn\\\(A  . 


SECT.  LXXII.]     POTENTIAL  ENERGY  OF  A   MAGNET.  173 

Now,   if  n  increases  from  0  to  1,  we  have    /  ndn  =  ^,  and  the  work 
done  is 

(d)  W=  \\\\(A  .  -dVftx  +  B  .  -dFj-dy  +  C  .  VFj?)z)dxdydz, 
or,  by  introducing  the  components  of  the  magnetic  force, 

(e)  W=  -  \\\\(A*  +  Bfl  +  Cy)dxdydz. 

The  energy  of  a  magnetic  system  may  be  expressed  in  another 
way.  The  same  method  by  which  we  before  obtained  an  expression 
for  the  energy,  shows  that  (f)  W=*\\<r.  V.  dS+\\\\p.  F.dxdydz, 
where  cr  and  p  represent  the  surface  and  volume-densities  respectively. 

If  V  and  its  first  differential  coefficient  vary  continuously,  we  have 
o-  =  0  and  W=\\\\p.V.dxdydz.  Now,  V2F+4«/>  =  0,  and  hence 

W  =  -  1  /STT  .  J  f  J  V(  32  Vfdy?  +  92  Vfif  +  32  Vfiz^dxdydz. 
By  integration  by  parts  over  the  whole  infinite  region,  we  obtain 


or  (g)  W=-\l&ir.  ^\(az  +  /3'2  +  y2)dxdydz.     Similar  expressions  hold  for 
dielectric  polarization  [cf.  LXI.]. 


SECTION  LXXIII.    MAGNETIC  DISTRIBUTION. 

A  piece  of  soft  iron  brought  into  a  magnetic  field  becomes  magnetized 
by  induction.  "We  assume  that  the  intensity  of  magnetization  at 
any  point  is  a  function  of  the  total  magnetic  force  acting  at  that 
point.  We  assume  that  the  intensity  of  magnetization  is  proportional 
to  the  magnetic  force,  or  that  (a)  A=ka,  B  =  kf3,  C=ky,  where  k  is 
a  constant.  The  magnetizing  force  proceeds  partly  from  the  per- 
manent magnets  present  in  the  field,  and  partly  from  the  quantities 
of  magnetism  induced  in  the  soft  iron.  The  potential  due  to  the 
former  may  be  designated  by  V,  that  due  to  the  latter  by  U, 
so  that 

A  =  -  k  .  3(  F+  U)fix,   B  =  -  k  .  3(  F+  U)fdy,    C  =  -  k  .  3(  F+  U)fiz. 
Now,  in  the  space  not  occupied  by  permanent  magnets,  we  have  V2F"=  0, 
and  therefore  'dA/'dx  +  'dB/'dy  +  'dC/'dz^  -kV2U,  or  since,  from  LXVIII. 
(e),  p=  -(dAj'dx+oBj'dy  +  'dCI'dz},  we  have  finally  V2U-p/k  =  0. 

Since  the  potential  W  is  due  to  the  components  of  magnetization 
A,  B,  C,  the  equation  that  holds  within  the  soft  iron  is  V2  U  +  4-rp  =  0. 
From  the  last  two  equations  we  obtain  (b)  (1  +  ±irk)p  =  Q,  and  hence 
p  =  0  ;  that  is,  there  is  no  free  magnetism  present  within  the  soft  iron.  The 


174  MAGNETISM.  [CHAP.  vm. 

magnetism  present  is  therefore  situated  on  the  surface  of  the  iron. 
We  will  now  determine  the  surface-density  cr  of  this  distribution. 
For  this  purpose  we  use  the  equation 


where  va  and  vt  are  the  normals  drawn  from  any  point  on  the  surface 
of  the  iron  inward  and  outward  respectively.  Ut  and  Ua  are  the 
values  of  the  potential  due  to  the  induced  magnetism  inside  and 
outside  the  iron  mass.  Now  we  have 

3  VfiVl  =  -  3  F/3vrt,  and  hence  4™  +  3  £7,/3v,  +  3  UJdva  =  0. 
The  magnetizing  force  just   outside   the  surface  of  the  soft  iron  is 
-3(P"+  £7,)/3v4   in  the  direction  of  vt.     The  free  magnetism  on  the 
corresponding  surface-element  is  therefore  cr  .dS  =  k.'d(V+  Ui)/'dvi.dS. 
Hence  we  have 

(c)  4vk  .  3  F/'dvt  +  (  1  +  47r£)3  U^v,  +  3  Ujova  =  0 

The  relation  (c)  in  connection  with  the  equations  (d)  V2Ut  =  Q,  V2f/0  =  0 
serves  to  determine  the  potentials  Ut  and  Ua. 

As  an  example  of  the  theory  here  presented,  we  will  consider 
the  magnetization  of  a  sphere  subjected  to  the  action  of  a  constant 
magnetizing  force  ^)  which  acts  in  the  direction  of  the  £-axis.  Let 
the  intensity  of  magnetization  of  the  sphere  in  the  direction  of  the 
a-axis  be  A  ;  the  force  due  to  the  magnetization  and  acting  in  the 
direction  of  the  z-axis  is  [LXX.  (f)]  equal  to  -4TT/3.A.  From 
equation  (a)  we  have,  therefore,  A  =  k(5g>  -  4ir/3  .  A),  and  hence 
^=&<p/(l+47r/3.&).  We  have  in  this  case  [LXX.  (e)] 

Ut  =  47T/3  .  Ax  ;    Ua  =  4;r/3  .  R3A  .  cos  0/r2. 
These  values  satisfy  equation  (c),  since  V=  -$gx. 

SECTION  LXXIV.    LINES  OF  MAGNETIC  FORCE. 

If  M  represents  the  free  magnetism  within  a  closed  surface,  and 
<^)n  the  component  of  magnetic  force  in  the  direction  of  the  normal 
to  the  surface,  we  have,  from  LXIX.,  (a)  4jr.3/=f  ]"£)„  .dS. 

If  we  lay  a  sheet  of  paper  on  a  magnet  which  lies  horizontally, 
and  scatter  iron  filings  over  it,  they  arrange  themselves  in  curves 
which  are  called  lines  of  magnetic  force.  Let  DE  (Fig.  79)  be  a  small 
surface  of  area  dS;  lines  of  force  proceed  from  its  perimeter  which 
bound  a  tube  of  force.  If  D'E'  represents  another  section  cut  through 
the  tube  of  force,  equation  (a)  can  be  applied  to  the  part  of  the 
tube  of  force  thus  bounded.  Since  the  direction  of  the  magnetic 


N 


SECT.  LXXIV.]  LINES   OF   MAGNETIC    FOKCE.  175 

force  coincides  with  the  direction  of  the  lines  of  force,  the  normal 

force  over  the  surface  of  the  tube  is  everywhere  zero  except  at  the 

ends  DE  and  D'E'.     Let  <p  be  the  force  acting  at  DE  and  £)'  that 

acting    at    D'E';    let    0    and    6'   be 

the  angles  between   the  direction  of 

the  force  and  the  directions  of  the 

normals  to  DE  and  D'E'  respectively.       ™ 

Since   there    is    no    free    magnetism 

present  in  the  interior  of  the  tube, 

we  have 

-  £)  .  (IS  .  cos  6  +  £)'  .  dS'  .  cos  9'  =  0. 

If  the  sections  dS  and  dS'  are  perpendicular  to  the  lines  of  force, 
we  have  ^)/^)'  =  dS'/dS.  The  force  is  therefore  inversely  proportional  to 
the  cross-section  of  the  tube  of  force.  The  lines  of  force  may  be  so 
drawn  that  their  distances  from  one  another  furnish  a  representation 
of  the  magnitude  of  the  magnetic  force  in  the  field. 

A  tube  of  magnetic  force  cannot  return  into  itself  or  form  a  hollow  ring. 

If  this  were  not  so,  the  work  which  is  done  by  the  magnetic 
forces  during  the  transfer  of  a  unit  quantity  of  magnetism  from  any 
point  over  a  closed  path  back  to  the  same  point  again  would  not 
be  zero.  If  ds  is  an  element  of  the  tube  of  force,  the  work  done 
would  be  J^).cfc>>0,  if  the  direction  of  motion  coincides  with  the 
direction  of  the  force.  If  the  magnetic  potential  is  V,  we  have, 
however,  ^)=  -dF/ds,  and  therefore,  for  a  closed  path, 


since  the  potential  is  a  single-valued  function  of  the  position  of  the 
point. 

Any  tube  of  magnetic  force  must  begin  and  end  on  the  surface  of  a 
magnet.      If  the  tube  ends  with  the  cross-section  PQ  (Fig.   80),   so 
that  a  magnetic  force  is  present  in  the  tube  TUQP,  while 
it  is  zero  outside  the  tube  at  R  and  S,  we  may  apply      Jt,---,S 
equation  (a)  to   the  region  TUQSRP.     Since  a  magnetic 
force    acts    at    the    surface    TU,    but   not   in   the   region 
PQSB,  the  surface  integral  taken  over  TUQSRP  cannot 
be  zero.      Magnetism  must   therefore   be   present   within 
the    closed    surface,    which    contradicts    our    assumption. 
Therefore,  any  tube  of  magnetic  force  ends  at  the  surface       FlG 
of  a  magnet. 

In   order   to   represent   the   magnitude  and  direction   of  magnetic 
forces,  Faraday  used  lines  of  magnetic  force  ;  he  assumed  that  the  lines 


176 


MAGNETISM. 


[CHAP.  vm. 


of  force  are  continued  in  the  body  of  the  magnet.  His  mode  of  repre- 
sentation has  become  of  very  great  importance.  If  a  magnet  is  broken, 
and  the  surfaces  exposed  by  the  fracture  are  placed  so  as  to  face 
each  other  and  separated  by  only  a  small  distance,  a  strong  magnetic 


T 


FIG.  81. 

force  acts  in  the  region  PQEU  (Fig.  81).  This  force  is  due  partly 
to  the  free  magnetism  in  the  interior  of  the  magnet  and  on  its  original 
surface,  and  partly  to  the  free  magnetism  on  the  newly-formed  sur- 
faces. The  force  due  to  the  former  cause  is  directed  from  the  north 
pole  n  to  the  south  pole  s,  that  due  to  the  latter  from  s  to  n.  The 
latter  force  is  in  practice  the  stronger,  so  that  we  may  say  with  a 
certain  propriety  that  the  magnetic  tubes  of  force  are  produced  through 
the  interior  of  the  magnet  along  the  path  D'F'FD  (Fig.  81). 


FIG.  82. 


If  S  (Fig.  82)  is  a  closed  surface  lying  outside  all  the  magnets  in 
the  field,  and  therefore  containing  no  magnetism,  we  have  from  XIV., 


It  is  customary  to  express  this  result  in  the  following  way  :  The 
integral  J^)n  .  dS  which  is  extended  over  a  part  of  the  surface  may  be 


SECT.  LXXIV.]  LINES  OF  MAGNETIC  FORCE.  177 

divided  into  the  parts  $£>nl .  dSlt  ^)n2 .  dS2,  etc.  Let  them  be  so  taken 
that  they  are  all  equal,  and  let  their  common  value  be  taken  as  unity. 
Since  the  product  §£)n.dS  is  constant  for  the  same  tube  or  line  of 
force,  the  integral  J.£)n .  dS  gives  the  number  of  lines  of  force  which  traverse 
the  surface.  If  this  integral  is  zero,  as  many  lines  of  force  entei'  the  sur- 
face as  leave  it. 

This  holds  for  a  surface  which  contains  one  or  more  magnets,  for 
the  sum  of  the  magnetism  in  every  magnet  is  zero.  On  the  other 
hand,  the  theorem  does  not  hold  if  the  surface  cuts  through  a  magnet. 
Nevertheless,  if  the  magnet  is  divided  into  two  parts,  MNQP  (Fig.  81) 
and  RSTU,  and  if  they  are  situated  infinitely  near  each  other,  the 
theorem  holds  for  either  of  them  if  the  surface  considered  contains 
one  part,  but  excludes  the  other.  Now,  if  this  mode  of  division  of 
the  magnet  produces  no  disturbance  in  its  magnetization,  the  theorem 
can  be  expressed  in  the  following  way : 

We  represent  the  components  of  the  magnetic  force  by  a,  (3,  y. 
In  the  part  of  the  surface  lying  outside  the  cleft,  no  other  magnetic 
force  is  acting;  on  the  other  hand,  the  free  magnetism  +0-  on  the 
element  dS  of  PQ  (Fig.  81),  and  —  o-  on  the  corresponding  element 
of  RU,  produce  a  force  which  can  be  determined  in  the  following 
way :  From  XIII.  a  surface  on  which  the  surface-density  is  o-  exerts 
an  attractive  force  27ro-  on  a  unit  of  mass  lying  very  near  it.  In 
the  case  of  magnetism,  the  force  lira-  is  a  repulsion.  If  there  are 
two  parallel  surfaces,  on  one  of  which  the  density  of  the  magnetic 
distribution  is  o-,  while  on  the  other  it  is  -  o-,  the  magnetic  force 
acting  between  the  surfaces  is  47r<r.  If  the  normal  to  the  surface- 
element  dS  directed  outward  makes  angles  with  the  axes  whose  cosines 
are  /,  m,  n,  we  have,  if  A,  B,  C  are  the  components  of  magnetization, 
a-  =  lA  +  mB  +  nC.  The  magnetic  force  in  the  direction  of  the  normal 
is  la  +  m/3  +  ny.  Since,  in  the  case  considered,  the  surface-integral  must 
be  zero,  we  have  {](Ai+Jit0+ny+4xv)d9«0,  or,  from  LXVIII.  (b), 

\[l(a  +  lirA)  +  m(p  +  ±irB)  +  n(y  +  4irC)]dS=  0. 
We  set  (b)  a  =  a  +  4vA,   b  =  P  +  ±irB,   c  =  y  +  lirC,  and  obtain 
(c)  f  (al  +  bm  +  cn)dS  =  0. 

The  quantities  a,  b,  c  are  the  components  of  magnetic  induction.  If, 
therefore,  the  directions  of  the  lines  of  force  are  determined  by  the 
directions  of  the  resultants  of  the  magnetic  induction,  it  follows  that 
the  lines  of  force  may  be  considered  as  continued  through  the  magnet  itself, 
and  that  they  therefore  return  into  themselves.  Now,  equation  (c)  shows 

M 


178 


MAGNETISM. 


[CHAP.  vin. 


that  as  many  lines  enter  a  surface  as  leave  it.  If  we  consider  an 
arbitrary  surface  S  so  drawn  as  to  pass  through  every  point  of  a  closed 
curve  5,  and  determine  the  magnetic  induction  whose  components 
are  a,  b,  c,  the  magnitude  N=  \(al  +  bm  +  cn)dS  is  determined  by  the 
boundary  s  of  the  surface  S.  We  may  say,  the  curve  s  encloses  N 
lines  of  magnetic  force. 

From  equations  (b)  it  follows  that 

'da/'dx  +  'db/'dy  +  'dc/'dz 

=  Vapz  +  Vppy  +  V7rdz  +  tor(dAfdx  +  'dBj'dy  +  Wj'dz)  =  -  V2  V-  4ir/o. 
Hence,  we  have 


SECTION  LXXV.    THE  EQUATION  OF  LINES  OF  FORCE. 

We   will   develop   the  equation    of  the   lines  of  force  of  a   small 
straight  magnet  NS  (Fig.  83)  which  is  magnetized  in  the  direction 

9 

fl 


S 


FIG.  83. 

of  its  length  with  the  intensity  of  magnetization  J.  Let  N  and  S 
represent  the  north  and  south  poles  respectively.  Free  magnetism 
is  present  on  the  end  surfaces  S  and  N,  supposed  to  be  plane  and  to 
have  the  area  dA ;  the  quantity  at  the  north  end  is  /.  dA,  that  at 
the  south  end  -  .7 .  dA.  Let  the  centre  of  the  magnet  be  the  origin 
of  coordinates,  and  the  x-axis  coincide  with  the  length  of  the  magnet. 
We  are  to  determine  the  components  a  and  (3  of  the  magnetic  force 
which  acts  at  the  point  P,  whose  coordinates  are  x  and  y.  If  21  is 
the  length  of  the  magnet,  and  if  PS  =  rv  PN=r2,  J.dA  =  q,  we  have 
a  =  q.(z-  Z)/r./  -  q  .  (x  +  J)/V,  ft  =  q .  y/r*  -  q  .  y/rf. 


SECT.  LXXV.]  EQUATION  OF  LINES  OF  FOECE.  179 

If  dx  and  dy  are  the  projections  of  an  element  of  the  line  of  force, 
we  have  dy/dx  =  f3/a,  or 

(a)  (x  -  l)tr/ .  dy  -  (x  +  l)/r^ .  dy  =  y/r/ .  dx  -  y/r^ .  dx. 

If  we  set  ^PSx  =  Qv  ^PNx  =  Q.2,  we  have 

cosei  =  (x  +  l)/rl  and  cos  0,  =  (x  -  T)lrr 

If  x  and  y  increase  by  dx  and  dy  respectively,  cos0  will  increase  by 
or 


In  the  same  way  d  cos  02  =  y2/rzs .  dx  -  (x  -  l)y/r/ .  dy.  From  equation 
(a)  we  obtain  d(cosSl  -  cos62)  =  0,  or,  if  c  is  a  constant,  cos61  -  cos02  =  c. 
This  is  the  equation  of  the  lines  of  force. 


SECTION  LXXVI.    MAGNETIC  INDUCTION. 
The  components  of  magnetic  induction  are 


If  we  consider  a,  b,  c  as  components  of  flux,  they  have  a  property 
similar  to  that  of  the  components  of  flow  of  an  incompressible  fluid 
[cf.  XLI.  (e)],  for  'dal'dx  +  'db/'dy  +  'dcj'dz  =  0,  or,  what  amounts  to  the 
same  thing,  ^(al  +  bm  +  cn)dS  =  0.  If  23  represents  the  resultant  of 
a,  b,  c,  and  €  the  angle  between  23  and  the  normal  to  the  surface-element 
dS  of  the  closed  surface  S,  we  have  (53.  cose.  ^=0. 

Let  EF=dS  (Fig.  84)  be  an  element  of  the  surface  of  the  magnet, 
and  let  the  quantity  of  magnetism  ar.dS  be  present  on  dS.  Let  the 
induction  outside  the  surface  EF,  gr  p' 

in   the   direction   EE',   be   230,   and  __  E  '          ~4^L_ 

inside    that    surface,    in    the    same    -^""^  '  ------  -  £ 

f"          f" 
direction,   be  23;.      Suppose   perpen- 

diculars erected  on  the  perimeter  of 

the    element  EF  and   the   surfaces    E'F'   and   E"F"  drawn   parallel 

to  EF.     We  then  have 

(S3i-S3re).f^=0,  or  23,  =  23rt. 

Outside  the  surface,  93a  is  the  same  as  the  magnetic  force  in  the  direc- 
tion EE'  ;  it  will  be  23rt=  -3F0/3va,  if  the  potential  outside  the  sur- 
face is  Va  and  the  normal  EE'  is  equal  to  dvn.  Let  V.  be  the  potential 
within  the  surface.  The  induction  23;  is  then  [LXXIV.] 


and  we  have,  therefore,  3Fi/3v<  +  3F"a/3va  +  4jrcr  =  0.     This  is  the  same 


180 


MAGNETISM. 


[CHAP.  vui. 


equation  as  LXIX.   (i).      If  the  body  considered  is  a  mass  of  iron 
whose  coefficient  of  magnetization  is  k,  we  have 

A=ka,    JB  =  kP,    C  =  ky,    fJ.=  l+±irk,    a  =  pa,    b  =  pP,    C  =  /xy. 
It  follows  that  'da/'dx  +  'dbl'dy  +  30/^  =  1*.  V2Vi  =  0  within  the  mass  of 
iron.      Within  that  mass,    therefore,    no   free   magnetism  is  present. 
The  magnetic  induction  perpendicular  to  the  surface  has   the  same 
value  on  both  sides  of  it,  that  is  [cf.  LXXIII.  (c)], 


The  magnitude  /*,  which  is  the  ratio  of  the  magnetic  induction  to 
the  magnetic  force,  may  be  called  the  magnetic  inductive  capacity  (mag- 
netic permeability}.  The  coefficient  of  induction  or  the  permeability  p  is 
equal  to  unity  in  vacuo,  where  k  =  Q  ;  in  paramagnetic  bodies  /x  >  1,  in 
diamagnetic  bodies  /*<  1. 


SECTION  LXXVII.    MAGNETIC  SHELLS. 

Suppose  a  thin  steel  plate  to  be  magnetized  so  that  one  face  is 
covered  with  north  magnetism  and  the  other  with  south  magnetism. 
At  any  point  A  in  the  face  N  (Fig.  85)  draw  a  normal  to  the  plate 
which  cuts  the  surface  S  at  B.  Let  the  plate  be  so  magnetized 


FIG.  85. 

that  -o-  represents  the  magnetic  surface-density  at  B,  and  +0-  that 
at  A.  We  set  AB  =  e,  and  call  o-e  =  4>  the  strength  of  the  shell*  at 
the  point  under  consideration.  If  the  plate  is  infinitely  thin  and  the 
surface-density  infinitely  great,  <f>  has  a  finite  value.  Such  a  plate  is 
called  a  magnetic  "shell" 

The  potential  of  such  a  shell  may  be  expressed  in  the  following 
way:    Let   LM  (Fig.   86)  be  the  shell,  dS  a  surface-element  on  its 
positive  face,  BC  the .  normal  to  this  surface-element,  and  P  the  point 
*  In  the  original,  the  moment  of  the  surface. — TB. 


SECT.  LXXVII.]  MAGNETIC  SHELLS.  181 

at  which  the  potential  is  to  be  determined.  We  represent  the  angle 
between  BC  and  BP  by  e.  The  potential  at  the  point  P,  due  to 
that  part  of  the  shell  whose  end-surface  is  dS,  is  [LXIX.] 

dF=(r.dS.e.cosf/rz. 

Hence,  if  the  strength  of  the  shell  is  constant,  (a)  V=  <£ .  J{  cos  e .  dS/r2, 
where  the  integral  is  to  be  extended  over  the  whole  surface.  If 
the  solid  angle  subtended  by  dS  at  the  point  P  is  called  dw,  and  if 
we  set  BP  =  r,  we  have  dS .  cos  e  =  r^dw. 

Therefore  dpr=or.  e.  e?w  =  <J?.  dw,  and  hence  (b)  F=<&.<a,  where  w 
is  the  solid  angle  subtended  by  the  shell  at  the  point  P.  We  may 
call  W  the  apparent  magnitude  of  the  shell  seen  from  the  point  P. 

If  the  point  for  which  the  potential  is  to  be  determined  lies  on 
the  opposite  side  of  the  surface,  say  at  P',  and  if  w'  is  the  solid 
angle  subtended  by  the  shell  at  the  point  P',  we  have  V  =  -  * .  w'. 
If  the  points  P  and  P'  approach  each  other  until  they  are  infinitely 
near,  but  on  opposite  sides  of  the  shell,  we  have  (c)  V  =  —  & .  (47r-  w), 
since  4?r  is  the  total  solid  angle  about  a  point.  Hence,  finally, 

V-V'  =  \TT.<$>. 

If  PQP'  (Fig.   87)   is  a  curve  which  does  not  cut  the  shell,  and 
whose  ends  lie  infinitely  near  each  other  on  opposite  sides  of  it,  the 
work  done  by  the  magnetic  forces  in  moving  a  unit         „---,. 
magnet   pole  over  the    path   PQP'  is  equal   to    4ir«i». 
This  theorem  holds  even  if  other  magnets  are  present 
in  the  field.     They  act  on  the  pole  with  forces  which 
have  a  single-valued  potential,  and  the  work  done  by 
them  during  the  motion  of  the  unit  pole  in  the  curve 
PQP'  is  equal  to  zero ;   for  this  curve  may  be  con- 
sidered as  a  closed  curve,  since  P  and  P'  are  infinitely 
near  each  other. 

After  obtaining  an  expression  for  the  potential  of  a 
magnetic  shell,  we  determine  the  force  with  which 
the  shell  acts  on  a  magnet  pole  of  unit  strength.  The  normal  to 
the  shell  makes  angles  with  the  axes  whose  cosines  are  I,  m,  n; 
let  £  */>  £  ke  the  coordinates  of  a  point  in  the  shell,  and  x,  y,  z 
the  coordinates  of  the  point  outside  the  shell  for  which  the  potential 
is  to  be  determined.  We  then  have 

cos  e  =  I.  (x -  £)/r  +  m.(y-  ^)/r  +  n .  (z -  f)/r, 

where  r'2  =  (x  -  £)2  +  (y-  rjf  +  (z-  £)2.  From  equation  (a)  the  potential 
is 

r=  & .  J  J [(,-  _  g)  .1  +  (y  _  rj)m  +  (z  -  f  )«]/r» .  dS. 


182  MAGNETISM.  [CHAP.  vm. 

Since  3r~1/c3£  =  (as-£)/rsi  we  obtain 

V=  *f  J(/  .  Br-1/^  +  wi  .  Br"1/^  +  n  .  'dr~lj'dt)dS. 

Represent  the  components  of  the  magnetic  force  in  the  direction  of 
the  .r-axis  by  a,  we  then  have  (d)  a=  -'dP'/'dx,  and 


because   3r~1/3«=  -Br"1/^.     If  the  shell  does  not  pass  through  the 
point  .T,  y,  2,  7-  will  never  become  zero,  and  we  have 

3V-i/3£2  +  3V-  W  +  BV-'/Bf2  =  0, 
(e)  a  =  +  *.  J  J[TO.32i-V3i/3£  +  «.  32r-Y3£3£  -  l(Wr~ll'dr 
From  the  theorem  of  VI.  (f),  we  have 

f  \(X.  d£/ds  +  Y  .  drj/ds  +  Z  .  dC/ds)ds 

+  n 


WesetJf=0,    Y=  +*.3r~1/9t  2=  -^.'dr~1/'dri,  by  which  the  right 
sides  of  equations  (e)  and  (f)  become  identical,  and  then  obtain 
(g)  a  =  *  .  {Br-VBf  -  *?/<**  -  Sr-1/^  .  dC/ds)ds. 

Analogous  expressions  hold  for  (3,  •/.     By  carrying  out  the  differen- 
tiation, we  obtain  (h)  a  =  $  .  J[(z  -  fl/r8  .  diy/rfe  -  (y  -  1?)/?-3  .  d{/ds]ds. 

The  force  is  therefore  determined  by  the  contour  and  the  strength  of  the 
magnetic  shell.  This  result  follows  from  the  fact  that  the  potential 
is  determined  by  the  solid  angle  and  the  strength  of  the  shell. 

In  order  to  find  the  geometrical  meaning  of  equation  (h),  we  use 
the  following  method  :  Let  £,  77,  £  be  the  coordinates  of  the  point  0 
(Fig.  88)  ;  Oy  and  Oz  represent  the  directions  of 
the  y-  and  s-axes  respectively.  Let  the  element 
ds  be  parallel  to  the  2-axis,  and  represented  by 
OA=d£;  we  then  have  drj  =  Q.  Let  the  point  P, 
for  which  the  potential  is  to  be  determined,  lie 
in  the  y^-plane,  and  let  OP  =  r.  We  set 


and  have  y  -rj  =  r .  sin  6.  The  magnetic  force 
due  to  ds  =  OA  is,  (i)  a  =  -  * .  ds .  sin  6/r3 ;  it 
is  perpendicular  to  the  y.^-plane.  Its  direction  may  be  determined 
in  the  following  way  :  If  the  right  hand  is  held  so  that  the  fingers 
point  in  the  direction  of  ds,  and  the  palm  is  turned  toward  the  pole  P, 
the  thumb  gives  the  direction  of  the  force. 

Finally,  we  determine  the  work  which  must  be  done  to  bring  a 
magnetic  shell  from  an  infinite  distance  to  a  place  where  the  magnetic 
potential  is  equal  to  V.  Let  the  shell  be  divided  into  elements  dS. 
In  order  to  bring  the  surface-element  which  carries  the  quantity 


SECT.  LXXVIL]  MAGNETIC  SHELLS.  183 

o- .  dS  of  south  magnetism  to  its  final  position,  work  equal  to  -  o- .  dS .  V 
must  be  done.  In  order  to  bring  the  corresponding  surface-element 
carrying  the  same  quantity  of  north  magnetism  to  its  place,  work 
equal  to  (F'+dF'/dv.  e)ar .  dS  must  be  done,  if  v  represents  the  normal 
to  the  surface-element  dS.  Hence  the  total  work  done  is 

A  =  f \dVldv .&r.dS=3>.  \\dVldv .  dS. 
Now,  since  dVjdv  =  -  (la  +  mfi  +  ny),  we  obtain  for  the  work  done 

A=  -3>.l\(loi  +  mp  +  ny)dS. 

If  JV  represents  the  number  of  lines  of  force  contained  by  the  contour 
of  the  shell,  we  have  A  =  -  & .  N. 


CHAPTER   IX. 

ELECTRO-MAGNETISM. 
SECTION  LXXVIII.    BIOT  AND  SAVART'S  LAW. 

OERSTED  discovered  that  the  electrical  current  exerts  an  action  on 
magnets  ;  the  law  of  the  magnetic  force  which  is  due  to  an  electrical 
current  was  discovered  by  Biot  and  Savart.  Let  AB  (Fig.  89)  be 
a  conductor  traversed  by  a  current  which  is 
S*P  measured  by  the  quantity  of  electricity  flowing 
/*  in  unit  time  through  any  cross-section.  Let 


D 


0,-' 


the  quantity  of  magnetism  p.  be  situated  at  the 


point  P,  and  let  the  conductor  AB  be  divided 
into  infinitely  small  parts  ds.  If  CD  =  ds  is  an 
infinitely  small  part  of  the  conductor  CP  =  r,  and 
6  the  angle  between  r  and  the  direction  of  the 
current  in  CD,  the  direction  of  the  force  will  be 
perpendicular  to  the  plane  determined  by  r  and  ds. 
If  the  right  hand  points  in  the  direction  of  the  current  and  the  palm  is 
turned  toward  the  magnet  pole,  the  direction  of  the  force  exerted  by  the 
current-element  on  the  pole  is  given  by  the  direction  of  the  thumb.  The 
magnitude  of  the  force  K  is  (a)  K=fi .  i.ds/r2.  sin0. 

The  magnetic  force  which  is  due  to  any  system  of  electrical  currents, 
whose  direction,  strength,  and  position  in  space  are  known,  may 
be  calculated  from  (a). 

If  the  current  forms  a  closed  circuit,  and  if  the  intensity  of  the 
current  is  the  same  at  all  points  in  the  conductor,  we  may  determine 
the  force  due  to  the  current  and  also  the  potential  which  the  current 
produces.  The  force  due  to  any  current-element  is  equal  to  that 
exerted  by  a  line-element  of  the  same  length,  which  forms  part  of 
the  contour  of  a  magnetic  shell,  whose  strength  is  equal  to  the  current- 

184 


CH.  ix.  SECT.  Lxxviii.JBIOT  AND  SAV  ART'S  LAW.  185 

strength.       This    follows   from    a    comparison   of   equation    (a)   with 
LXXVII.  (i). 

From  LXXVII.  (b),  the  potential  V  of  a  closed  circuit  of  strength  i, 
at  the  point  P,  is  (b)  V=iw,  where  w  is  the  solid  angle  subtended 
by  the  circuit  at  P.  If  a,  /?,  y  are  the  components  of  the  magnetic 
force  acting  at  P,  we  have,  from  LXXVII.  (h), 

(  a  =  i  .  f  ((*  -  {)/rS  .  d-nlds  -(y-  n)li*  .  dtfds)ds, 
(c)  J8  =  i  .  \((x  -  £)/r3  .  dflds  -(z-  f)/» 


7  =  *  • 

If  ABC  (Fig.  90)  is  a  conductor  through  which  a  current  i  flows 
in  the  direction  indicated  by  the  arrow,  and  if  a  unit  magnet  pole 
moves  around  the  current  in  the  direction 
found  by  using  the  right  hand  in  the  manner 
before  described,  the  work  done  by  the  mag- 
netic forces  during  the  movement  over  the  /j 
path  DFED  is,  from  (b),  equal  to  4iri.  If  the 
path  of  the  pole  encircles  several  currents  i,  i', 
i"  etc.,  the  magnetic  forces  due  to  these 
currents  do  upon  it,  during  its  motion,  the 
work  A,  given  by  (d)  A  =  4ir(i  +  i'  +  i"  +  ...), 
in  which  the  currents  which  flow  in  one  FlG 

direction  are  to  be  reckoned  positive,  and 
those  which  flow  in  the  opposite  direction,  negative.  Hence,  the 
potential  which  an  electrical  current  produces  at  the  point  F  is 
not  determined  only  by  the  position  of  that  point.  If  we  bring  a 
unit  pole  (Fig.  90)  to  F  over  the  path  GF  from  an  infinite  distance, 
the  work  which  is  done  will  be  equal  to  V,  the  potential  at  the 
point  F.  If  the  pole  then  passes  around  the  current  over  the  path 
FEDF,  the  work  k-n-i  will  be  done,  and  the  potential  at  F  becomes 
V+±iri.  If  the  pole  passes  n  times  around  the  current  in  the 
same  way,  the  potential  at  F  becomes  V+  47rni.  Hence,  the  potential 
at  the  point  F  has  an  infinite  number  of  values.  The  differential 
coefficients  of  the  potential  with  respect  to  x,  y,  z  are  nevertheless 
completely  determined. 

If  a  pole  of  strength  ^  passes  once  around  the  current,  the  work 
done  on  it  is  iTrip;  if  a  magnet  passes  once  around  the  current  and 
returns  to  its  original  position,  the  work  done  is  47ri2/*,  when  2/* 
represents  the  sum  of  the  quantities  of  magnetism  in  the  magnet. 
But  since  for  any  magnet  2/*  =  0,  the  work  done  is  in  this  case  equal 
to  zero. 


186  ELECTRO-MAGNETISM.  [CHAP.  ix. 

Since  an  electrical  current  may  be  replaced  by  a  magnetic  shell, 
we  can  obtain  the  magnetic  moment  of  an  infinitely  small  closed 
current.  If  i  is  the  current-strength,  the  strength  of  the  equivalent 
magnetic  shell  is  a-e=--i.  If  dS  is  the  surface  of  the  shell,  we  have 
i.dS  =  <re.dS.  The  quantity  of  magnetism  on  one  side  of  the  shell 
is  <rdS,  and  the  thickness  of  the  shell  is  e.  Hence,  the  magnetic 
moment  of  the  current  equals  the  product  of  the  current-strength 
and  the  area  enclosed  by  the  circuit. 


SECTION  LXXIX.     SYSTEMS  OF  CURRENTS. 

Let  a  conductor  be  wound  around  a  cylinder,  so  that  the  distances 
of  the  separate  turns  from  each  other  are  equal.  We  may  approxi- 
mately determine  the  magnetic  action  of  the  system  in  the  following 
way :  If  L  is  the  length  of  the  cylinder,  N  the  number  of  turns, 
and  i  the  current-strength,  the  current  flowing  in  unit  length  of 
the  cylinder  is  Ni/L,  and  that  flowing  in  the  length  dx  is  Ni.dxjL. 
A  portion  of  the  cylinder  whose  length  is  dx  may  be  replaced  by 
a  magnetic  shell,  whose  thickness  is  dx  and  whose  surface-density 
is  a-,  if  (a)  <r.dx  =  Ni.dx/L  and  o-  =  Ni/L.  If  this  substitution  is 
carried  out  for  the  whole  length  of  the  cylinder,  the  actions  of  the 
positive  and  negative  faces  of  the  substituted  magnetic  shells  annul 
each  other  everywhere  except  on  the  ends  of  the  cylinder.  If  the 
current  flows  in  the  way  shown  in  Fig.  91,  A  exhibits  negative  and 
B  positive  magnetism.  Such  a  system  of  currents  is  called  a  solenoid. 
Outside  the  cylinder,  the  only  magnetic  forces  which  act  are  those 
which  proceed  from  the  poles  A  and  B.  If  the  length  of  the  solenoid 
is  great  in  comparison-  with  its  diameter,  the  magnetic  force  outside 
of  it  vanishes  near  its  middle  point.  The  force  in  the  interior  of 
the  solenoid  may  be  determined  in  the  following  way :  Let  the 
line  CD  be  parallel  to  the  axis  of  the  solenoid  (Fig.  91),  let  CF  and 

DE  be  perpendicular  to  its  sur- 
face,  and  let  the  line  FE  be 
parallel  to  CD.  We  assume  that 
the  magnetic  force  y  is  parallel 
FIG  91  to  the  axis  of  the  solenoid  in  its 

interior,    and   that   no   magnetic 

force  acts  outside  of  it.  Suppose  a  unit  pole  to  traverse  the  closed 
path  CDEF.  The  work  done  by  the  magnetic  force  is  y .  d:c,  if  we 
set  CD  =  dx.  From  LXXVIII.  (d),  we  have  (b)  y.dx=4ir.Ni.dxlL, 


SECT.  LXXIX.]  SYSTEMS  OF  CURRENTS.  187 

y  =  ^irNijL.  If  we  represent  the  number  of  turns  in  unit  length  by 
N,  we  have  y  =  ±irNli,  that  is,  the  magnetic  force  (number  of  lines 
of  force  per  square  centimetre)  in  the  interior  of  the  solenoid  and 
\ecir  its  middle  point  is  given  by  the  product  of  the  current-strength  i 
into  the  number  N  of  turns  per  unit  length  of  the  solenoid. 

We  will  now  determine  the  magnetic  force  of  a  sphere  on  whose 
surface  a  conductor  is  wound.     Suppose  that  a  conductor  is  wound 
on  a  sphere  of  radius  It,   so  that  the  planes   of 
the  turns  are  parallel    and   separated   from   each 
other   by  the  distance   a.     Suppose   that  ABCD 
and  EFGH  (Fig.  92)  are  two   of  the   turns.     If 
the  current-strength   is   i,   a   single  turn  may  be 
replaced  by  a  magnetic  shell  whose  surface-density 
is   s,   if  as  =  i.      For    points   outside   the   sphere, 
the    action    of   the    positive    magnetism    on    the 
surface  EG  will  be  nearly  annulled  by  the  action 
of  the   negative   magnetism   on  the  surface  EH; 
the  only  effective  part  of  the  two  surfaces  is  the     "         FlG  92 
circular  ring,   whose  width  is  BE.     Suppose  the 
magnetism  on  this  ring  to  be  distributed  with  the  density  <r,  over 
the  zone  BFGC.     We  have  then 

BE .s  =  BF.v,  or  <r/s  =  BE/BF=  cos  9, 

if  6  is  the  angle  between  the  radius  OB,  and  the  line  OP  perpendicular 
to  the  plane  of  the  coils.  Using  the  relation  as  =  i,  we  obtain 
a-  =  i/a.  cos  6.  From  LXX.  (e),  (f)  the  magnetic  potential  for  points 
outside  the  sphere  is  given  by  Va  =  47r/3 .  RHja .  cos  6/r2,  since  i/a  is 
equivalent  to  the  intensity  of  magnetization  /. 

In  determining  the  magnetic  force  in  the  interior  of  the  sphere, 
we  must  remember  that  the  magnetic  lines  of  force  due  to  the  currents 
are  continuous,  and  that  the  magnetic  force  in  this  case  is  the  same  as 
the  magnetic  induction  of  the  equivalent  magnetized  sphere  [LXXIV.]. 
To  find  it,  we  suppose,  as  in  that  section,  that  the  magnetized  sphere 
is  divided  into  two  parts  by  cutting  out  an  infinitely  thin  section 
perpendicular  to  the  lines  of  magnetic  force,  and  that  the  unit  pole 
is  placed  in  the  opening  between  these  two  parts.  It  will  then  be 
subjected  to  the  force  '2-n-s  directed  toward  the  north  end  of  the 
magnetized  sphere  due  to  the  repulsion  of  the  north  magnetism 
exposed  on  one  face  of  the  cut,  to  the  force  2-n-s  due  to  the  attraction 
of  the  south  face  of  the  cut  and  in  the  same  direction,  and  to  the 
force  -  f Tri/a,  [LXX.  (f)]  due  to  the  distribution  on  the  outside  of  the 


188  ELECTRO-MAGNETISM.  [CHAP.  ix. 

sphere  and  in  the  opposite  direction.  The  total  force  F  acting  on 
the  pole  is  therefore  F  =  4iri/a  -  £iri/a  =  ^ni/a. 

If  the  whole  number  of  turns  is  N,  we  have  F=§ir.Ni/R,  that  is, 
the  force  in  the  interior  of  the  sphere  is  proportional  directly  to  the  current- 
strength  i  and  to  the  number  of  turns  A7,  and  inversely  to  the  radius  It 
of  the  sphere.  We  get  the  same  result  if  we  consider  that  this  sphere 
behaves  like  a  magnetized  iron  sphere  whose  intensity  of  magnet- 
ization is  J=i/a. 

In  this  way  we  can  set  up  an  almost  constant  magnetic  field,  which 
may  be  applied  in  the  construction  of  instruments  used  to  determine 
current-strength. 

If  the  solenoid  forms  a  closed  ring,  we  obtain  a  system  of  currents 
which  has  many  applications.  Let  AB  (Fig.  93)  be  a  circle  whose 

centre  is  0  and  whose  radius  is  r,  and 
let  R  be  the  distance  between  the 
centre  0  and  a  straight  line  CD, 
which  lies  in  the  plane  of  the  circle. 
If  the  circle  rotates  about  the  axis 
CD,  it  describes  a  circular  ring.  Sup- 
~j)  pose  that  on  this  ring  there  are  JV 
turns  of  wire,  through  which  the 

current  i  flows.  We  replace  the  separate  turns  by  magnetic  shells, 
and  determine  the  magnetic  force  in  the  interior  of  the  ring.  We 
reach  this  result  most  simply  if  we  suppose  a  unit  pole  to  move 
on  a  circle  of  radius  R  about  the  axis  CD.  The  work  done  by  the 
magnetic  force  ^),  which  acts  in  the  interior  of  the  ring,  when  the 
unit  pole  has  completed  one  revolution,  is  2-n-IiQ.  This  work  is  also 
equal  to  ItrNi  if  the  path  of  the  pole  is  in  the  interior  of  the  ring ; 
it  is  equal  to  zero  if  the  path  is  outside  the  ring. 

Hence  we  have,  in  the  former  case,  2irB§£>  =  faNi  and  ^)  =  INifR  ; 
in  the  latter  case,  ^p  =  0. 


SECTION  LXXX.    THE  FUNDAMENTAL  EQUATIONS  OF 
ELECTRO-MAGNETISM. 

Up  to  this  point  we  have  considered  the  path  of  the  electrical  current 
as  a  geometrical  line.  In  reality  the  current  always  occupies  space, 
and  is  determined  by  its  components  along  the  coordinate  axes.  For 
example,  if  dy .  dz  is  a  surface-element  perpendicular  to  the  z-axis, 
and  if  the  quantity  of  electricity  u.dy.dz.  dt  passes  through  it  in 


SECT.  LXXX.]     EQUATIONS  OF  ELECTRO-MAGNETISM.  189 

the  positive  direction  in  the  time  dt,  u  is  the  component  of  current 
in  the  direction  of  the  re-axis.  The  components  of  current  in  the 
directions  of  the  two  other  axes  are  represented  by  v  and  w.  If  Oy 
and  Oz  are  drawn  through  the  point  0  (Fig.  94), 
whose  coordinates  are  x,  y,  z,  parallel  to  the  cor- 
responding coordinate  axes,  and  if  the  rectangle 
OBDC  is  constructed  with  the  sides  dy  and  dz, 
the  current  u.dy . dz  flows  through  the  element 
OBDC.  If  the  components  of  the  magnetic  force 
are  represented  by  a,  p,  y,  and  if  a  unit  pole  o 


moves  about  the  rectangle  in  the  direction  OBDCO, 
the  work  done  by  the  magnetic  forces  will  be 

(3.dy  +  (y  +  'dy/'dy  .  dy)  .dz-(p  +  3/3/3z  ,dz).dy-y.dz 


This  is  [LXXVIII.  (d)]  equal  to  4?r  .  u  .  dy  .  dz.  Hence,  we  obtain 
the  equations 

(a)  4iru  =  (dy[dy-'dpfdz),  4m  =  (da[dz-'dy[dx)i  4irw  =  (d(J/'dx-'da/'dy). 

These  equations  express  the  current  in  terms  of  the  magnetic  force.  In  a 
region  where  there  is  no  current  we  have  u  =  0,  v  =  0,  w  —  0,  and 
therefore  oy/3y  =  3/3/32,  da/32  =  3y/3z,  'dfij'dx  =  3a/3y, 
or  a.  dx  +  p.dy  +  y.dz=  -dV.  Therefore,  in  a  region  where  there  is  no 
current  the  magnetic  forces  have  a  potential.  In  this  case  the  forces  arise 
from  magnets. 

From  equations  (a)  the  magnetic  force  is  not  determined  only  by 
the  components  of  current.  If  u,  v,  w  are  given  and  a,  p,  y  so  deter- 
mined that  equations  (a)  are  satisfied,  these  equations  will  also  be 
satisfied  if  we  replace  a,  p,  y  by 


where  V  is  an  arbitrary  function.  The  potential  due  to  the  magnets 
present  in  the  region  is  V. 

We  will  now  consider  a  few  simple  examples  : 

(a)  Suppose  the  direction  of  the  magnetic  force  to  be  parallel  to 
the  0-axis,  and  its  magnitude  to  be  a  function  of  the  distance  r  from. 
this  axis  (Fig.  95).  We  then  obtain  from  equations  (a) 

47TM  =  +dyjdr.y/r,    4irv  =  -dy/dr.x/r,   w  =  0. 

The  current  is  parallel  to  the  zy-plane  and  perpendicular  to  r.  The 
current-strength  J  is 

J=  u  cos  (uJ)  +  v  cos  (vJ),   J—  -u.  y/r  +  v  .  x/r  =  —  l/4?r  .  dy/dr. 


190 


ELECTRO-MAGNETISM. 


[CHAP.  ix. 


If  y  is  constant  in  the  interior  of  a  cylinder  whose  radius  is  OA=rl 
(Fig.  95),  and  equal  to  zero  outside  a  cylinder  of  radius  r.2,  the  current 
in  the  unit  length  of  the  cylinder  is 


j  .  dr  =  - 


.  dr 


FIG.  95. 


which  agrees  with  LXXIX.  (b). 

(b)  If  the  current-strength  is  given, 
we  can  find  the  magnetic  force  by 
integrating  equations  (a).  Let  u  and 
v  be  zero,  and  w  be  a  function  of  the 
distance  r  from  the  ^-axis.  We  then 
have  from  (a) 


These  equations  will  be  satisfied  if  we  assume  that  y  =  0  and  that 
a  and  (3  are  functions  of  x  and  y  only.  Suppose  that  the  magnetic 
force  is  resolved  into  two  components,  one  of  which,  R,  acts  in  the 
direction  of  the  prolongation  of  r,  and  the  other,  S,  is  perpendicular 
to  r.  We  then  obtain  a  =  R  .  x/r  -  S  .  y/r,  fl  --=  R  .  yjr  +  S  .  .r/r,  and 
therefore  47rw  =  dS/dr  +  S/r  =l/r.  d(Sr)/dr. 

If  the  conductor  is  a  tube  bounded  by  two  coaxial  cylinders  whose 
radii  are  Rl  and  R2,  and  if  w  is  constant  in  the  conductor  we  have, 
if  Cv  C2,  C3  are  constants,  Slr=Cv  2Trwr2  +  C.2  =  S.2r,  S3r=C3.  The 
first  of  these  equations  holds  for  the  interior,  the  second  for  the  con- 
ductor, the  third  for  the  space  outside  the  conductor.  From  the  nature 
of  the  problem,  Sl  must  have  a  finite  value  in  the  axis  ;  we  therefore 
have  C\  =  0.  Since  the  magnetic  force  changes  continuously,  we  have 
£,  =  0  when  r  =  Rly  and  therefore  C.2  =  -  2-n-wR^,  S.2  =  Airier  - 


Since  irw(R.22  -  R^)  is  equal  to  the  current-strength  i  in  the  conductor, 
we  have  S3  =  2i/r. 

Therefore,  an  infinitely  long  straight  linear  current  exerts  a  magnetic 
force  at  a  given  point,  which  is  inversely  pi-oportional  to  the  distance  of  that 
point  from  the  current. 


SECTION  LXXXI.     SYSTEMS  OF  CURRENTS  IN  GENERAL. 

The  components  of  current  and  the  components  of  magnetic  force 
are  connected  by  the  equations  [LXXX.  (a)] 
(a)    4?rM  =  'dyj'dy -  'dfi'fdz,    ITTV  =  'da/'dz  -  'dy/'dx,    lirw  =  'dft/'dx  -  'da/ay. 


SECT.  LXXXI.]  SYSTEMS  OF  CUREENTS.  191 


From    these   equations   it   follows   that  (b) 

This  equation  corresponds  with  the  equation  of  continuity  in  mechanics, 
and  asserts  that  the  total  quantity  of  electricity  contained  in  a  closed  region  is 
constant.  It  thus  appears  that  the  current,  whose  components  are  u,  v,  w, 
moves  like  an  incompressible  fluid.  There  is  never  any  accumulation 
of  electricity,  but  only  a  displacement  of  it.  This  apparently  con- 
tradicts experience  ;  in  order  to  be  consistent  with  our  method  of 
treatment  we  assume  with  Faraday  that  an  electrical  polarization  or 
an  electrical  displacement  occurs.  We  represent  the  components  of 
this  displacement  by  /,  g,  h.  If  one  of  the  components,  say  /,  increases 
by  the  increment  df  in  the  time  dt,  df/dt=f  represents  the  quantity 
of  electricity  which  passes  in  unit  time  through  a  unit  of  area  per- 
pendicular to  the  z-axis,  in  consequence  of  the  change  of  polarization. 
If  p,  q,  r  represent  the  components  of  the  electrical  current  which  is 
due  to  the  flow  of  electricity  through  the  body,  we  have 
(c)  u=p  +  df/dt,  v  =  q  +  dg/dt,  w  =  r  +  dhjdt. 

These  quantities,  u,  v,  iv,  are  the  components  of  the  actual  current, 
which  is  made  up  of  the  current  conducted  by  the  body  and  the  current 
arising  from  the  change  of  polarization  or  the  electrical  displacement. 

If  the  components  of  the  current  are  finite,  the  components  of 
magnetic  force  vary  continuously  when  no  magnets  are  present  in  the 
region.  The  components  of  force  perpendicular  to  the  surfaces  of  the 
magnets,  if  any  are  present,  is  in  general  discontinuous.  We  assume 
that  currents  of  infinite  strength  do  not  occur  in  practice  ;  however 
we  sometimes  consider  the  flow  in  a  surface,  in  which  case  we  must 
assume  that  the  components  of  current  in  the  surface  are  infinite. 
In  this  case  the  components  of  force  parallel  to  the  surface  vary  dis- 
continuously  on  passage  from  one  side  of  the  surface  to  the  other. 
If  a:  and  a2  represent  these  components  of  force,  and  J  the  quantity 
of  electricity  which  flows  through  a  unit  of  length  perpendicular  to 
the  components,  we  have  from  LXXVIII.  (d)  47r/=a2-a1.  We  may 
obtain  the  same  result  from  (a)  as  follows  :  We  consider  two  surfaces 
whose  equations  are  z  =  cl  and  z  =  cy  We  obtain  from  the  first  two 
equations  (a) 

477  .  Fu  .  dz  =      33  .dz-      +  fi       ±-  .  /"*»  .  dz  =  a  -<L-  f*33fc  .  dz. 


Pi  and  p.2  are  the  components  of  the  magnetic  force  in  the  direction 
of  the  y-axis  on  both  sides  of  the  plane  surface  ;  ^  and  a.2  have  similar 
meanings.  If  c.2  -  cl  =  c  is  infinitely  small,  and  u  and  v  are  infinitely 

great,  the  integrals   /  udz  and   /   vdz  are  the  quantities  of  electricity 


192  ELECTRO-MAGNETISM.  [CHAP.  ix. 

which  flow  in  a  surface.  The  integrals  on  the  right  side  vanish 
simultaneously,  and  we  obtain  the  value  given  in  (d)  for  the  difference 
between  the  components  of  magnetic  force  on  the  opposite  sides  of 
the  surface. 


SECTION  LXXXII.    THE  ACTION  OF  ELECTRICAL  CURRENTS  ON 
EACH  OTHER. 

The  work  which  must   be   done   upon  two   conductors  A  and  E 
(Fig.  96),  carrying  the  currents  i  and  i',  in  order  to  bring  them  nearer 
each  other,   can   be  determined  in  the 
following  way :  Let  the  position  of  the 
current  A  be  fixed,  while  B  is  brought 
toward    it    from    an    infinite    distance. 
B  may  be  replaced  by  a  magnetic  shell 
whose    surface-density  is  o-  and   whose 
thickness   is   e.      Suppose   ODE  to  be 
FIG.  96.  a    straight    line,   normal   to    the    shell, 

which   cuts   it   on  its  negative   face   at 

C  and  on  its  positive  face  at  D.  A  surface  element  dS'  at  C  carries 
with  it  the  quantity  of  magnetism  -  <rdS'.  If  V  represents  the 
potential  which  A  produces  at  C,  the  work  done  upon  the  element 
dS'  at  C  is  -  Fa- .  dS'.  Setting  CD  =  dv,  the  work  done  upon  the 
element  dS'  at  D  is  (F'+'dF'/'dv  .dv).a- .dS'.  Hence  the  work  arising 
from  this  portion  of  the  shell  is  'dF'j'dv  .dv.a-.  dS'.  But  dv .  a-  =  i',  and 
hence  the  work  done  upon  the  whole  shell  is  (a)  W=i  .  \\dVfov.  dS'. 

The  force  acting  in  the  direction   CE  is   -'dF'/'dv.     If  CE  makes 
angles  with  the  axes  whose  cosines  are  /',  m',  ri,  we  have 

(b)  W=  -  i' .  Jf  (I'a  +  m'p  +  riy)dS'. 

The  quantities  a,  (3,  y  are  determined  in  LXXVIII.  (c),  and  may 
be  put  in  the  form 

a  =  i\(dr~l/dy .  d£/ds  -  'dr^/dz .  drj/ds)ds, 
ft  =  i \(dr-lldz .  dg/ds  -  dr~l/dx .  d£/ds)ds, 
y  =  tj(3r- 1  fa .  dri/ds  -  Vr~lfdy .  dg/ds)ds. 
From  the  theorem  of  VI.  (f)  we  have 

\(X .  dx/ds'  +  Y.  dy/ds'  +  Z .  dz/ds')ds' 

z)  +  m'(dX/dz  -  dZ/dx)  +  n'(dY/dx  -  dX/dy)]dS'. 


SECT.  LXXXII.]  MUTUAL  ACTION   OF  CURRENTS.  193 

We  now  set 

X  =  Jii'/r .  dg/ds  .ds,    Y=  J«'/r .  dij/ds  .ds,   Z=  Jii'/r .  dQds .  ds, 
and  obtain 

ii'\\(d£/ds .  dx/ds  +  drj/ds .  dy/ds'  +  dflds .  dz/ds')dsds'/r 
=  f  J  jM'PX^-Vfy  •  dt/ds  -  Vr-ifdz  .  drj/ds) 
+  m'@r-lfdz .  dg/ds  -  dr^fdz .  d£/ds) 
+  n'(dr-i/-dx .  drj/ds  -  3r~V3y .  d£/ds)]dsdS', 

or  ii'\\(dg/ds .  dx/ds'  +  drj/ds .  dy/ds'  +  d£(ds .  dz/ds')dsds'/r 

=  \\i'(l'a  +  m'p  +  riy)dS'  =  -  W. 
Hence  from  (b) 
(c)       W=  -  ti\\(d£lds .  dx/ds'  +  drjjds .  dy/ds'  +  dflds .  dzlds')dsds' /r, 

where  ds  is  an  element  of  one  conductor  and  ds'  of  the  other. 

If  we  represent  the  angle  between  two  elements  ds  and  ds'  by  c, 
we  obtain  F.  E.  Neumann's  expression  for  the  potential  energy  of  two 
electrical  currents,  (d)  W=  -w'JJcose/r.  dsds'. 

If  we  represent  the  magnetic  force  which  is  perpendicular  to  an 
arbitrary  surface  containing  the  circuit  B  by  ^),  we  have  from  (a), 
(e)  W=  -  i' .  J^)' .  dS'  =  -i'N,  if  N  represents,  in  Faraday's  nomen- 
clature, the  number  of  lines  of  force  enclosed  by  the  conductor. 
Therefore  the  potential  energy  of  a  current  equals  the  negative  pi'oduct  of 
the  current-strength  and  the  number  of  lines  of  force  enclosed  by  the  con- 
ductor. Hence  it  follows  that  a  current  always  tends  to  move  so 
that  the  number  of  lines  of  force  enclosed  by  it  shall  become  as  great 
as  possible.  The  positive  direction  of  the  lines  of  force  is  the  direc- 
tion in  which  a  north  pole  moves  under  the  action  of  the  current 
[cf.  LXXIV.]  The  above  theorem  has  only  been  proved  on  the 
assumption  that  the  magnetic  force  ^p  is  due  to  another  current. 
But  because  currents  and  magnets  are  equiv- 
alent, the  law  is  true  generally.  Since  the 
surface  containing  the  circuit  B  may  be 
arbitrarily  chosen,  the  energy  W  depends 
only  on  the  contour  of  the  circuit. 

The  force  which  acts  on  an  element 
AB  =  ds  of  the  current  ABO  (Fig.  97) 
may  be  determined  in  the  following  way : 
Suppose  that  the  conductor  ABC  moves  so 
that  AB  is  displaced  to  A'B'  in  such  a 

manner  that  A  A'  and  BB'  are  perpendicular  to  AB.     If  AA'  =  dp, 

N 


194  ELECTEO-MAGNETISM.  [CHAP.  ix. 

the  area  contained  by  the  conductor  increases  by  ds.dp.  If  the 
magnetic  force  at  A  is  equal  to  £),  and  if  its  direction  makes  the  angle 
a  with  the  normal"  to  the  surface  ABB' A',  the  component  K  of  the 
force  normal  to  the  surface  is  K=^>cosa.  Hence  the  increment  of 
the  potential  energy  of  the  conductor  is,  from  (e),  dW '=  -iK. ds.dp, 
if  i  is  the  current-strength.  In  order  to  cause  the  motion  here  described, 
a  force  X  must  act  on  ds  in  the  direction  A  A',  which  is  determined 
by  X.dp  =  -iK.  ds.dp,  X=  -iK.ds.  Hence  the  force  -IK.ds  acts 
on  the  current-element  in  the  aforesaid  direction,  and  if  the  current- 
element  is  free  to  move,  the  direction  of  its  motion  is  perpendicular 
to  the  direction  of  the  magnetic  force  as  well  as  to  its  own  direction. 
The  direction  of  the  motion  is  determined  by  laying  the  right  hand 
on  the  current  [cf.  LXXVIII.]. 

It  follows  further  that  the  force  which  acts  on  an  element  ds  of 
the  current  i  is  perpendicular  to  the  plane  determined  by  the  current 
and  the  direction  of  the  magnetic  force  Q.  If  we  represent  the  angle 
between  the  direction  of  the  force  and  the  direction  of  the  current 
by  <£,  this  force  will  equal  $£>i .  ds .  sin  <£. 


SECTION  LXXXIII.    THE  MEASUREMENT  or  CURRENT-STRENGTH  OR 
THE  QUANTITY  or  ELECTRICITY. 

(a)  Constant  Currents. 

To  measure  constant  currents  we  generally  use  a  galvanometer  con- 
sisting of  parallel  circular  conductors  carrying  the  current  whose 
strength  is  to  be  determined.  A  magnet  whose  dimensions  are  small 
in  comparison  with  the  radius  of  the  coils,  is  suspended  in  the  centre 
of  the  apparatus,  which  is  so  placed  that  the  coils  are  parallel  to 
the  magnetic  meridian.  The  current  sets  up  a  magnetic  force  whose 
value  is  Gi  perpendicular  to  the  direction  of  the  earth's  magnetic  force, 
whose  horizontal  component  is  called  H.  G  depends  on  the  con- 
struction of  the  galvanometer.  If  G  is  constant  in  the  region  in  which 
the  magnet  moves,  the  angle  <£  by  which  the  magnet  is  turned  from 
its  position  of  rest  by  the  current  is  determined  by 

(a)  tg<t>  =  GilH,    i  =  H/G.tg<j>, 

that  is,  the  current-strength  is,  in  this  case,  proportional  to  the  tangent  of 
the  angle  of  deflection. 


SECT.  LXXXIII.]  MEASUREMENT  OF  CURRENT-STRENGTH.  195 

(b)   Variable  Currents. 

It  is  very  difficult  to  determine  the  strength  of  currents  of  short 
duration  at  any  instant.  We  may,  however,  easily  measure  the  total 
quantity  of  electricity  Q  which  flows  through  the  conductor.  From' 
LXXI.  (d)  the  moment  which  tends  to  turn  the  magnet  about  a 
perpendicular  axis  is  -  SCfta  sin  0,  if  9JJ  is  the  magnetic  moment  of 
the  magnet,  a  the  magnetic  force,  and  0  the  angle  between  the  direc- 
tions of  9Ji  and  a.  Setting  a  =  Gi,  where  G  is  the  galvanometer 
constant,  and  assuming  Q  =  ^TT,  the  directive  J wee  exerted  by  the  current 
on  the  magnet  is  equal  to  £0J6ri.  The  total  moment  caused  by  the 
current  is  therefore  \^llGi.dt  =  ^lG.Q,  if  we  write  Q  =  \i.dt.  Q  is  the 
quantity  of  electricity  which  passes  through  the  conductor  during 
the  discharge. 

If  .7  is  the  moment  of  inertia  of  the  magnet,  and  <o  its  angular 
velocity,  Ju>  will  be  its  moment  of  momentum.  We  thus  obtain  the 
equation  (b)  *$lGQ  =  Jo>.  If  the  period  of  oscillation  of  the  magnet 
is  called  T,  we  have,  by  LXXI.  (f), 

(c),  (d)  T  =  Tr.Jj/WH  and  therefore  Q  =  Hr^/Gir2. 

The  kinetic  energy  which  the  magnet  receives  from  the  impulse 
given  to  it  by  the  current  is  |/w2,  in  consequence  of  which  it  turns 
through  the  angle  6.  Its  potential  energy  thereby  increases  from 
-WlH  to  -WlHcosQ;  the  work  done  on  it  is  9Ji#(l -cos0).  We 
therefore  have  iJo)2  =  29JlH"sin2(0/2),  or,  if  6  is  very  small,  (e)  G-TW/TT 
and  Q  =  Hr/TrG .  0,  that  is,  if  there  is  no  damping  action  on  the  magnet, 
and  if  its  angular  displacement  is  small,  the  quantity  of  electricity  flowing 
through  a  section  of  the  conductor  is  proportional  to  the  angular  displace- 
ment of  the  magnet. 

(c)  Damping  Action. 

The  oscillations  of  the  magnet  generally  diminish  rather  rapidly 
in  consequence  of  what  is  called  damping  or  damping  action.  Damping 
arises  from  resistance  of  the  air  and  the  action  of  currents  induced 
by  the  motion  of  the  magnet  in  neighbouring  conductors.  If  there 
is  no  damping,  we  have  from  LXXI.  (e)  and  (f),  when  the  oscillations 
are  small,  d'2Q/dt2  =  -  7r2/r2 . 0,  T  is  therefore  the  period  of  oscillation 
of  the  undamped  magnet.  We  may  assume  that  the  damping  action 
is  proportional  to  the  angular  velocity  dQ/dt.  Taking  the  damping 
into  account,  we  have,  to  determine  the  deflection  0,  the  differential 
equation  (f)  0  +  2?w0  +  7r2/r2 .0  =  0.  The  factor  m  depends  on  the  size 
and  character  of  the  oscillating  magnet,  on  the  density  of  the  air, 


196  ELECTRO-MAGNETISM.  [CHAP.  ix. 

and  on  the  size,  character,  and  position  of  the  masses  of  metal  in 
which  currents  are  induced.  If  we  set  TT/T  =  n  and  6  =  eat,  we  have 

a2  +  2ma  +  n-  =  0   and   a  =  -  m  ±  \/n2  -  m?*J  -  1 

in  which  it  is  assumed  that  n  >  m.  Setting  (g)  n2  -m2  —  n-2/Ti2>  we 
have  6  =  (A  sin  (rtfa)  +  B  COS^/T^)  .  e~mt.  If  0  =  0  at  the  time  /  =  0, 
we  obtain  Q  —  A  .  e~mt  .  sin  (irf/Tj).  dQ/dt  =  <*>  at  the  time  /  =  0,  and 
therefore  9»r1«*/«>.tf^"*..nn(vl/T1),  To  find  the  magnitude  of  the 
deflection  we  set  dQ/dt  =  Q,  and  obtain  (h)  tg^/Tj)  =  7r/wrr  If  TO  is 
the  smallest  root  of  this  equation,  the  successive  roots  are  TO  +  TJ, 
TQ  +  2rv  ....  The  oscillations  are  therefore  isochronous.  If  we  repre- 
sent the  deflections  by  0U  62,  03,  ...,  we  have 

.e-mr<>.  sin  (^TO/TJ), 
.  e  -  m(To+Ti>  .  sin  (TT^/TJ  ), 
~  «<To+2Ti)  .  sin  (*-TO/TJ  ). 


If  the  position  of  equilibrium  is  designated  by  A0,  the  first  point 
of  reversal  by  Av  the  second  by  A.2,  etc.,  the  ratio  between  the 
oscillations  AA*  and  A.A  is 


(ej-e^^Og-e,)  and  (e1-e2y/(e3-e2)=e"-r'. 

We  set  w-r^A,  and  obtain  (i)  A  =  log  nat  [(ex  -  02)/  '(03  -  62)].  A  is 
the  logarithmic  decrement,  which  can  be  very  exactly  determined  from 
a  series  of  oscillations.  From  (g)  the  period  of  oscillation  TI}  is 
(k)  TJ  =  T  .  Vl  +  A'2  --.  Therefore  the  period  of  oscillation  is  increased 
by  the  damping  action.  If  we  set  T  =  TO  in  equation  (h),  we  have 

tg(7rro/Ti)  =  vlmri  and  7rTo/Ti  =  arctg(ir/X),    mr0  =  A/V  .  arctg(7r/  A), 

sm(irr0/T1)  =  l/Vl+A2/ir2. 
Hence   we   have   further  0:  =  T^/TT  .  g-V*-.arctg(ir/A)  ,  i/^/j  +  xay^-'    and 

(1)  fc)  =  TrOj/T!  .  Vl  +  A*/JT*  .  «W»"  •  arctg  (x/X), 

We  obtain  from  (d)  and  (k)  Q  =  Hr^jGir-  .  w/(l  +  A2/jr2),  and  using 
equation  (1),  (m)  #  =  ei.fir1/GV.«x/r-1TO»*0r/x>.  l/Vr+A^. 

In  order  to  determine  the  quantity  of  electricity  sent  through  a 
conductor  by  an  electrical  current,  whose  duration  is  small  in  com- 
parison with  the  period  of  oscillation  of  the  magnet,  we  must  determine 
the  logarithmic  decrement  and  the  period  of  oscillation  of  the  magnet. 
Q  is  then  determined  from  these  quantities,  if  we  know  in  addition 
the  intensity  of  the  earth's  magnetism  and  the  constant  of  the 
galvanometer. 


SECT.  LXXXIII.]  MEASUREMENT  OF  CURRENT-STRENGTH.  197 

Setting  arctg  7r/X  =  ^7r-x  we  have  tg.r  =  A/7r.  If  A  is  very  small, 
we  have  x  =  X.jir  and  arctg  (jr/A)  =  \TT  —  A/TT.  If  the  damping  action 
is  insignificant,  we  can  neglect  higher  powers  in  the  series  in  which 
the  exponential  may  be  developed,  and  obtain 

2  and    Q  = 


SECTION  LXXXIV.    OHM'S  LAW  AND  JOULE'S  LAW. 

We  have  up  to  this  point  assumed  the  existence  of  the  electrical 
current  and  have  not  discussed  the  question  of  the  way  in  which  it 
is  started  and  maintained.  This  mode  of  treatment  in  many  respects 
lacks  clearness.  We  will  therefore  state  such  facts  as  are  well  established 
by  observation.  The  so-called  galvanic  elements  can  establish  and 
maintain  an  almost  constant  current.  In  order  to  maintain  a  constant 
current  in  a  conductor,  an  electromotive  force  must  act  in  the  direction 
of  the  current.  If  u  is  the  quantity  of  electricity  which  flows  in  unit 
time  through  a  unit  of  surface  of  the  #>/-plane  in  the  direction  of 
the  .r-axis,  we  can  set  w=C.X,  if  C  is  the  conductivity  and  X  the 
component  of  the  electromotive  force  in  the  direction  of  the  o-axis. 
C  depends  on  the  nature  of  the  conductor,  and  may  be  supposed  to 
have  the  same  value  in  the  conductor  in  all  directions.  If  the  com- 
ponents of  current  and  of  force  in  the  other  two  directions  are  v,  w, 
and  Y,  Z  respectively,  we  have  (a)  u  =  CX,  v=CY,  w=CZ.  Hence 
'fafdx  +  'dordy  +  'du;l'dz=C('dX[dx  +  'dY[dy  +  'dZ[dz).  If  the  steady  state 
of  the  electrical  current  has  been  reached,  the  left  side  of  the  equa- 
tion equals  zero,  and  hence  the  right  side  is  also  equal  to  zero.  If 
the  electromotive  forces  have  a  potential  V,  we  have  (b)  V2F~=0. 
This  equation  states  that  no  free  electricity  is  present  within  the  conductor- 
as  soon  as  the  current  becomes  steady.  The  electromotive  forces  must 
therefore  arise  from  the  free  electricity  on  the  surface  of  the  conductor. 

Suppose  ABC  (Fig.  98)  to  be  an  electrical  conductor.  We  will 
consider  a  portion  of  it  which  is  bounded  by  the  infinitely  small 
cross-sections  A  A '  =  S  and  BB'  =  S,  separated 

from  each  other  by  the  distance  I,  which  is     ______ — ^r- 1 £— __ 

also  infinitely  small.     If  AB  is  parallel  to     ^ ]b — -7- 

the  ./--axis,  the  component  of  current  u  equals 

CX,  and  therefore  the  quantity  of  electricity  FlG-  98' 

i  =  uS=CX.S  flows  through  the  cross-section  S.      If  V  and  V"  are 

the  potentials  at  A  and  B  respectively,  we  have  i  =  C .  S .  ( V- 


198  ELECTRO-MAGNETISM..  [CHAP.  ix.  SECT.  LXXXIV. 

and  further  (c)  t  =  (V-  P)/(//(7S)  =  (F-  V")/K  The  resistance  I!  is 
directly  proportional  to  the  length  of  the  conductor  and  inversely  propor- 
tional to  its  cross-section,  and  to  the  conductivity  of  the  substance  constituting 
the  conductor.  The  difference  of  potential  between  A  and  B  is  V  -  V. 
Equation  (c)  contains  Ohm's  law,  according  to  which  the  current- 
strength  is  directly  proportional  to  the  difference  of  potential  and  inversely 
proportional  to  the  resistance. 

The  quantity  of  electricity  i  .  dt  flows  through  the  cross-section  A  A' 
in  the  time  dt  and  passes  from  A  to  B  under  the  influence  of  the 
electromotive  force  X.  The  work  done  is  therefore 


The  work  done  in  this  part  of  the  conductor  by  the  electromotive 
forces  in  unit  time  is  (d)  A  =  i(V-  V)  =  i'2R  This  work  is  trans- 
formed into  heat  in  the  conductor.  Therefore  the  quantity  of  heat 
developed  in  a  conductor  is  proportional  to  the  square  of  the  current-strength 
and  the  resistance  of  the  conductor.  This  theorem  was  proved  experi- 
mentally by  Joule  and  deduced  theoretically  by  Clausius. 


CHAPTER   X. 

INDUCTION. 
SECTION  LXXXV.    INDUCTION. 

FARADAY  was  the  first  to  demonstrate  that  a  current  is  set  up  in 

a  conductor  if  a  magnet  or  a  conductor  carrying  a  current  is  moved 

in  its  neighbourhood.     F.  E.  Neumann  discovered  the  laws  of  these 

induced  currents.      Faraday  himself  afterwards  described  a  method 

of  determining  the   strength   and   direction  of  the  induced   current 

which   possesses   great   advantages,    because  it  makes   it   possible  to 

visualise    the    process.      Suppose    ABO   to    be    a    closed    conductor 

(Fig.  99),  and  DE,  D'E',  etc.,  the  lines  of  force 

enclosed  by  it.      Let  us  designate  an  element 

of  a    surface    bounded   by   the    conductor  by      *"*  — 

dS,    the    components    of    the    magnetic    force 

(cf.  LXXI.)  by  a,  /?,  y,  and  the  angles  which 

the  normal  to  dS  makes  with  the  axes  by  I, 

m,  n.     An    electromotive   force   arises   in   the 

conductor  if  the  interal 


(a) 

changes  its  value.  If  magnetizable  bodies  are  enclosed  by  the  circuit 
which  have  a  greater  permeability  for  lines  of  force  than  air,  the 
components  of  force  must  be  replaced  by  the  components  of  magnetic 
induction.  An  electromotive  force  then  arises  in  the  conductor  if 
the  integral  (cf.  LXXIV.)  N=  \(al  +  bm  +  cn)dS  changes  its  value.  An 
induced  current  arises  if  the  number  of  lines  of  f  wee  enclosed  by  the  con- 
ductor is  changed.  If  the  change  in  the  number  of  lines  of  force 
is  an  increase,  the  induced  current  tends  to  diminish  the  number  of 
the  enclosed  lines  of  force,  for  the  direction  of  the  induced  current 
is  such  that  its  own  lines  of  force  are  opposite  to  the  lines  of  force 
formerly  existing.  If  the  direction  indicated  in  the  figure  by  the 


200  INDUCTION.  [CHAP.  x. 

arrow  is  taken  as  positive,  the  induced  electromotive  force  acts  in 
a  negative  direction. 

According  to  Lenz's  law,  the  current  induced  by  the  motion  of  a  circuit 
tends,  by  its  electrodynamic  action,  to  oppose  the  motion,  by  which  it  is 
induced. 

In  order  to  determine  the  magnitude  of  the  induced  electromotive 
force,  we  suppose  that  the  current-strength  at  a  given  instant  is 
equal  to  i.  If  we  move  the  conductor  in  the  magnetic  field,  we 
must,  by  LXXXIL,  do  the  work  -i.dN,  and,  at  the  same  time,  the 
quantity  of  energy  Ei2 .  dt  is  transformed  into  heat.  "We  have  at 
once  -  i .  dN=  EP .  dt,  and  therefore,  because  Ei  equals  the  electro- 
motive force  e,  (b)  e  =  -  dN/dt,  that  is,  the  induced  electromotive  force 
is  equal  to  the  decrease  in  unit  time  of  the  number  of  lines  of  foi'ce  enclosed 
by  the  circuit. 

The  induced  electromotive  force  depends  on  the  value  of  the 
magnetic  induction,  whose  components  are  a,  b,  c,  not  on  the  magnetic 
force,  whose  components  are  a,  /3,  y.  If  there  is  no  magnet  near, 
and  if  the  coefficient  of  magnetization  of  the  region  is  &  =  0  (cf. 
LXXVL),  the  induction  and  the  magnetic  force  have  the  same  value, 
and  a,  b,  c  may  be  replaced  by  a,  /?,  y. 

For  example,  if  the  circuit  ABC  at  the  time  t  carries  a  current  of 
strength  i,  the  number  N  of  lines  of  force  passing  through  the  circuit 
in  the  positive  direction  is  N=Li.  L  is  the  number  of  lines  of  force 
if  the  current-strength  is  unity.  It  is  called  the  coefficient  of  self- 
induction.  If  the  current  diminishes  there  arises  an  electromotive 
force  (c)  e  =  —  d(Li)/dt.  According  to  Ohm's  law  we  have  e  =  7iV.  if  we 
represent  the  resistance  by  E,  and  therefore 

(d)  Ei  =  -  d(Li)/dt  =  -  L .  flijdt, 

provided  that  the  coefficient  of  self-induction  L  is  constant.  This 
coefficient  depends  on  the  permeability  of  the  region  and  also  on 
the  form  of  the  conductor. 

If  i0  is  the  current-strength  at  the  time  /  =  0,  we  have  (e)  i  =  i0 .  e~R!L-'. 
The  current-strength  therefore  diminishes  the  more  rapidly  the  greater  the 
resistance  and  the  smaller  the  coefficient  of  self-induction. 

We  obtain  from  (c) 

J"eidt  =  -Zp.di  =  Ui02. 

From  LXXXIV.  (d),  the  left  side  of  this  equation  is  an  expression  for 
the  work  done,  which  appears  as  heat  in  the  conductor.  Hence  we 
obtain  for  the  electro-kinetic  energy  T  of  a  conductor  whose  self-induction 


SECT.  LXXXV.]  INDUCTION.  201 

is  L  and  which  carries  a  current  of  strength  i,  (f)  T=^Li2.  The 
electro-kinetic  energy  of  the  circuit  is  therefore  equal  to  half  the  product  of 
the  coefficient  of  self-induction  L  and  the  square  of  the  current-strength  i. 

If  ABC  and  A'B'C'  (Fig.  100)  are  two 
conductors  carrying  currents  whose  respec- 
tive strengths  are  ^  and  i2,  the  current 
^  sets  up  a  number  Llil  of  lines  of  force 
which  pass  through  the  conductor  ABC. 
The  current  i<,  also  sets  up  lines  of  force, 
and  the  number  of  them  which  pass 
through  ABC  may  be  represented  by 
M2li2.  The  total  number  of  lines  of  force 
enclosed  by  ABC  is  therefore 
(g)  N^L^  +  Mrf* 

The  number  of  lines  of  force  enclosed  by  A'B'C'  is  (h)  N2  = 
From  LXXXII.  and  the  discussion  at  the  beginning  of  this  section, 
we  have,  in  the  usual  notation, 

=  *     COS  er  '  dS^ 


nlc2)dS1  =  i2\  cos  e/r  . 

The  integral  with  respect  to  dS2  equals  i^f^,  that  with  respect  to 
dS1  equals  i2M2l.     From  LXXXII.  we  have 

M12  =  M2l  =  |  cos  e/r  .  ds^dsy. 
M12  =  M2l  is  the  coefficient  of  mutual  induction  of  the  two  circuits. 

If  R1  and  R0  are  the  resistances  of  the  conductors  ABC  and  A'B'C' 
respectively,  we  have 

Rfr  =  6l=  -  dNJdt  =  -  Ll  .  dijdt  -  M2l  .  di2/dt, 
Rt,i2  =  «2  =  -  dN2/dt  =  -  L.2  .  dijdt  -  M12  .  dijdt. 

Hence  we  have,  for  the  electro-kinetic  energy  T  of  a  system  of  two  con- 
ductors which  carry  the  currents  zt  and  i2, 

T  =  J^i,  +  e2i2)dt  =  SLj*  +  Ml2ij2  +  \L^. 

For  the  electro-kinetic  energy   T  of  any  system  of  conductors,    we 
find  in  the  same  way, 

(i)  T=i(Llil2  +  L2i2^  +  Lsi32+  ...  +2M,2ili2  +  2Ml3ili3  +  2M23i2i,+  ...). 
If  Nv  N2,  N3  ...  denote  the  number  of  lines  of  force  enclosed 
respectively  by  the  conductors  1,  2,  3...  so  that,  for  example, 
N1  =  Lli1  +  M2li2  +  M3li3  +  ...,  the  expression  for  the  electro-kinetic 
energy  T  becomes  (k)  r=£(JV1*'1  +  A~2»'2  +  JV3«3+  ...)  =  pM,  that  is,  the 


202 


INDUCTION. 


[CHAP.  x. 


electro-kinetic  energy  of  a  system  of  currents  is  equal  to  the  sum  of  the  pro- 
,  ducts  of  the  number  of  lines  of  force  enclosed  by  each  conductor'  and  the 
strength  of  the  current  present  in  the  conductor. 

Electrical  currents  arise  not  only  if  the  neighbouring  currents  change 
in  strength,  but  also  if  they  change  their  position,  so  that  7l/12,  M13  . . . 
vary,  and  also  if  the  conductor  itself  changes  its  form.  In  all  cases 
the  induced  current  is  determined  by  the  change  in  the  number  of 
lines  of  force  enclosed  by  the  conductor. 


SECTION  LXXXVI.     COEFFICIENTS  OF  INDUCTION. 

In  the  investigation  of  variable  electrical  currents  flowing  in  wire 
coils,  the  coefficients  of  induction  between  different  turns  in  any  one 
coil  and  between  separate  coils  are  of  great 
importance.  The  calculation  of  these  co- 
efficients is  in  most  cases  very  difficult  ;  we 
will  consider  only  one  simple  case.  We 
suppose  two  circular  conductors  whose  radii 
are  r:  and  r.2  (Fig.  101).  They  have  a 
common  axis,  and  are  separated  by  the 
distance  b.  Suppose  r2>r1.  We  have  to 
calculate  the  integral 


FIG.  101. 


.  cos  e. 
We  first  evaluate  the  integral  m, 

m  =  ^dsjr  .  cos  e. 
We   have  r-  =  b2  +  r22  -  2r1?'2  .  cos  c  +  1\2.      If 


p  is  the  shortest  and  q  the  longest   distance   between  points  of  the 
two  conductors,  we  have 


and  m  =  2  I  ra  .  cos  e  .  dtj*ffP  +  (q*  -  p'z)  sm2£«. 

If  a  is  a  small  angle,  so  chosen  that  qa  is  very  great  in  comparison 
with  p,  we  can  set 

m  =  2        .  deJP  +     .J/i  +  2  f.  *,  .  df  .  (1  -  2  sin2|e)/?  sin  it. 

-a 

f»«/2  sin  £e  -  f  sin  ^«  .  efeT 

.  [Iog(2r1a/p)  -  log(a/4)  -  2], 

.  [log(8r1//>)  -  2]  =  2[log(8r1/J9)  -  2]. 


SECT.  LXXXVI.]         COEFFICIENTS   OF   INDUCTION. 


203 


With  this  value  for  m,  we  obtain  Ml.2  =  47rr2(\og(8rl/p)  -  2).  On  the 
assumptions  which  have  been  made  we  may  set  rl  =  r.2  =  R,  so  that 

(a)  Mn  =  ^E(\og(SE/p)-2). 

The  coefficient  of  induction  between  two  coils,  the  number  of  whose 
turns  is  n^  and  ?i.>  respectively,  and  for  which  the  mean  value  of 
logp  is  expressed  by  logP,  is  given  by 

(b)  ^12  =  4rJl%7rJ??(log(8^/P)-2). 

On  the  same  assumptions  the  coefficient  of  self-induction  L  of  a 
single  coil,  if  n  is  the  number  of  turns,  is  given  by 

(c)  L  =  4irn2E(\og(8B/P)  -  2). 

We  will  not  go  further  into  the  calculation  of  coefficients  of  induction  ; 
in  most  cases  they  are  determined  experimentally  by  one  of  the 
following  methods  : 


Methods  of  Determining  the  Coefficients  of  Induction. 

(a)  If  the  coefficient  of  mutual  induction  M  of  two  coils  Zx  and  L.2 
is  known,  we  may  determine  in  the  following  way  the  coefficient 
of  mutual  induction  M'  for  two  other  coils  Z/  and  L.2'. 

Let  Zj  and  L.2  (Fig.  102)  be  the  coils  whose  coefficient  is  known, 
and  L±  and  L.2  those  which  are  to  be  investigated.  A  current  is 


FIG.  102. 


passed  through  the  coils  Z/  and  L.2  from  the  voltaic  cell  E.  The 
coils  Z2  and  L.2  are  joined  by  conductors,  and  conductors  are  joined 
from  the  points  a  and  b  to  the  galvanometer  G.  If  the  current  / 
which  passes  through  L^  and  Z/  is  suddenly  broken,  electromotive 


204  INDUCTION.  [CHAP.  x. 

forces  e  and  e'  arise  in  L.2  and  L.2'.  If  /2  and  J.2'  are  the  strengths 
of  the  currents  induced  in  L.2  and  L.2\  we  have 

e  =  -  d(L^J2  +  MJ)/dt,   e  =  -  d(L.2'J.2  +  M'J)/dt, 

where  L.2  and  Z-2'  represent  the  coefficients  of  self-induction.  Applying 
KirchhofTs  laws  to  the  circuit  L.2G  and  L2G,  it  follows  that 

-  d(L.2  /2  +  M  J)/dt  =  R2  J2  +  G(J2  -  J2'), 

-  d(L2J2  +  M'J)/dt  =  R2'J.2r  -  G(J2  -  J2'), 

if  the  resistance  of  the  galvanometer  is  designated  by  G,  and  the 
resistances  of  the  coils  L2  and  L2  by  E.2  and  R2  respectively.  We 
multiply  these  equations  by  dt  and  integrate  from  t  =  0  to  t  =  T,  where 
T  is  a  very  small  time-interval.  If  the  current  /  is  broken  at  the 
instant  t  =  0,  a  current  is  induced  in  the  circuit  L.y'GL2  which,  in 
the  time  T,  sets  in  motion  in  the  circuit  L.2  the  quantity  of  electricity 
C2,  in  L2  the  quantity  C2,  and  therefore  in  the  galvanometer  the 
quantity  C2-C2.  At  the  time  t  =  Q,  J=J  and  J.2  =  J.2'  =  C2  =  C2'  =  0, 
and  at  the  time  t  =  r,  J=0  and  the  induced  current  has  also  vanished, 
so  that  J2  =  /2'  =  0.  Hence  we  have 

MJ=  E2C,  +  G(C2 -  G,'),    M'J=  R2'C2  -  G(C2 -  C2). 
C2  -  C2  =  J(M/E2  -  M'/R2)/(1  +  G/R2  +  G/R2). 

C2  -  C2  is  the  induced  quantity  of  electricity  which  flows  through 
the  galvanometer,  that  is,  the  total  current.  The  galvanometer  shows 
no  deflection  if  the  resistances  satisfy  the  equation  M'/M=R2'/R2. 

(b)  The  comparison  between  two  coefficients  of  self-induction  can 
be  carried  out  in  the  following  way  :    Let  A  BCD  (Fig.    103)  be  a 


FIG.  103. 

Wheatstone's  bridge- with  a  galvanometer  inserted  in  the  arm  BD. 
L^  and  L2  are  two  coils  inserted  in  the  arms  AB  and  BC,  whose 
coefficients  of  self-induction  are  to  be  compared.  The  current  entering 
at  A  and  passing  out  at  C  distributes  itself  in  the  conductors,  and 
causes  a  deflection  of  the  galvanometer  needle.  Let  the  resistances 


SECT.  LXXXVI.]         COEFFICIENTS  OF  INDUCTION.  205 

&L,  Ry  Sj_  and  S2  in  AB,  EG,  AD  and  DC  respectively,  be  so  adjusted 
that  no  current  passes  through  the  galvanometer  ;  we  then  have 


When  the  circuit  is  broken  at  E,  an  electromotive  force  arises 
by  induction  in  £:  and  L2,  in  consequence  of  which  a  current  of 
strength  g  flows  through  the  galvanometer.  Let  ilt  «'„  yv  y.2  respec- 
tively be  the  current-strengths  in  the  conductors  AB,  BC,  AD,  DC. 
They  are  connected  by  the  relations  il  =  yv  «2  =  7o-  We  then  have 

-  d(L&)ldt  =  (Rl  +  Sfo  +  Gg,    -  d(L.2i2)/dt  =  (E2  +  S2)i2  -  Gg. 

At  the  instant  £  =  0,  when  the  circuit  is  broken  at  E,  the  same 
current  i0  was  in  AB  and  BC  ;  whence 

LJQ  =  (Rl  +  SJ  f\  .dt+GTg.dt;    L,i0  =  (Rz  +  S2)  l\  .dl-oTg.  dt, 

Jo  Jo  Jo  >>o 

where  r  denotes  a  very  short  time.     No  deflection  is  caused  in  the 
galvanometer  by  the  current  g  if   I  g.dt  =  Q. 

Since  ^  =  i.2  +  g  we  have  i:  =  i2,  if  no   current   flows   through   the 
galvanometer,  and  hence  in  that  case  LJL2  =  (El  +  <S'1)/(-K2 
By  the  use  of  the  relation  Rl:fi.2  =  Sl:  S2,  it  follows  that 


that  is,  the  coefficients  of  self-induction  of  the  coils  are  in  the  same  ratio 
as  the  resistances  of  the  two  arms  in  which  the  coils  are  introduced. 

If  therefore  the  needle  of  the  galvanometer  is  not  deflected  either 
by  a  constant  or  a  variable  current  passing  through  the  circuits  Z: 
and  L2,  we  have  a  means  of  determining  the  ratio  between  the 
coefficients  of  self-induction  L  and  £.,. 


SECTION  LXXXVII.     MEASUREMENT  OF  RESISTANCE. 

The  strength  of  the  electrical  current  in  a  conductor  is  determined 
by  its  magnetic  action ;  the  electromotive  force  is  determined  by 
the  change  in  the  number  of  lines  of  force  contained  by  the  con- 
ductor. The  resistance  in  the  conductor  may  then  be  determined 
by  the  help  of  Ohm's  law.  Many  methods  of  measuring  resistance 
are  in  use.  We  will  confine  ourselves  to  the  description  of  some  of 
the  simplest. 

In  one  of  these,  used  by  W.  Weber,  the  essential  part  of  the  apparatus 
is  a  wire  coil  which  can  rotate  about  a  vertical  axis.  This  coil  is 


206 


INDUCTION. 


[CHAP.  x. 


set  perpendicular  to  the  magnetic  meridian  and  is  then  turned  through 
180°.  The  coil  is  in  connection  with  a  galvanometer;  the  total 
resistance  of  the  circuit  is  R.  If  the  plane  of  the  coil  makes  an 
angle  <£  with  the  magnetic  meridian  at  a  particular  instant,  the 
number  of  lines  of  force  passing  through  the  coil  is  SH  sin  <£.  if  <S' 
represents  the  area  enclosed  by  the  coils  and  H  the  horizontal  inten- 
sity of  the  earth's  magnetism.  If  L  denotes  the  coefficient  of  self- 
induction  of  the  coil  and  galvanometer  and  i  the  current-strength, 
we  have  (a)  -d(SHsin  tfi/dt  -d(Li)  dt  =  Ri. 

Since  </>  changes,  during  the  rotation  of  the  coil,  from  +  i~  to  —  ^TT, 
and  since  i  is  zero  both  at  the  beginning  and  at  the  end  of  the 
motion,  we  have  (b)  2SH=  RQ,  where  Q  denotes  the  total  quantity 
of  electricity  which  flows  through  the  conductor.  If  Q  is  measured 
by  the  method  described  in  LXXXIIL,  we  have 


The  absence  of  H  from  this  expression  shows  that  it  is  not  necessary 
to  know  the  intensity  of  the  earth's  magnetism  in  order  to  determine 
the  resistance. 


Sir  William  Thomson's  (Lord  Kelvin's)  Method. 

If  the  coil  above  described  turns  with  a  constant  angular  velocity  w, 
we  have  by  (a)  -  SHw  cos  <£  -  L  .  di/dt  =  Ri.  The  integral  of  this  equa- 
tion is  i  =  i0.  e~KiL  -  A  .  cos(^-a).  If  the  rotation  is  continued  for 
a  considerable  time,  the  exponential  term  vanishes  and  need  be  no 
longer  considered.  To  determine  A  and  a,  we  have 

(c)  A=SHt»/(Rcosa  +  Lwsma);    tga^Lu/R, 

and  therefore      A  =  SHu/Rjl  +  LW/R*  =  SHu/R  .  cos  a. 

It  thus  appears  that  the  self-induction  appar- 
ently increases  the  resistance. 

If  ON  (Fig.  104)  is  the  magnetic  meridian 
and  if  the  line  OM  is  perpendicular  to  it, 
the  coil  acts  on  a  magnetic  needle  at  its 
centre  with  the  force  Gi,  whose  direction  is 
that  of  the  line  OP  perpendicular  to  the 
plane  of  the  coil.  The  components  of  this 
FIG.  104.  force  are 


OM=  a  =  Gi  .  cos  <£   and 


SECT.  LXXXVII.]      MEASUREMENT  OF  RESISTANCE. 


207 


Let  aa  and  bl  denote  the  mean  values  of  these  forces.     We  then  have 

a,  =  1  /2ir .  /  Gi .  cos <f> .  d<k  ;    ft,  =  1 /2-n- .  I  Gi.sin<f>.  d<f>. 
Now  *  Jo 

I  cos  (<f>  —  a) .  cos  <£ .  d(f>  =  TT  .  cos  a,     /  cos  (<f>  —  a) .  sin  </> .  d<f>  =  TT  .  sin  a, 
and  hence  flx  =  -  \GA  .  cos  a,    1^=- \GA.sm  a. 

The  magnet  at  the  centre  of  the  coil  turns  from  the  meridian  in 
the  same  sense  as  that  in  which  the  coil  rotates.  If  its  angular 
displacement  is  represented  by  B,  we  have 

tg  6  =  -  «!/(#  +  bj)  =  GA  cos  a/(2H-  GA  sin  a), 

or,  introducing  the  value  of  A,  tg  6  =  GS<»  cos2a/(27i  -  GSw  sin  a  cos  a). 
This  equation,  in  connection  with  (c),  serves  to  determine  the  resistance 
R.  If  a  is  very  small  we  have  B=GSot/'2tgQ. 


L.  Lorenz's  Method. 

Suppose  that  a  metallic  disk  ABC  (Fig.  105),  whose  radius  is  a, 
turns  with  constant  velocity  about  an  axis  passing  perpendicularly 
through  its  centre.  Around  the  rim  of  the  disk,  and  concentric  with 


FIG.  105. 

it,  let  there  be  placed  a  coil  EF,  through  which  flows  an  electrical 
current  of  strength  i,  arising  from  the  voltaic  battery  H.  This  current 
sets  up  a  magnetic  force,  whose  component  perpendicular  to  the 
plane  of  the  disk  may  be  set  equal  to  mi,  where  m  is  a  function 
of  the  distance  from  the  centre  of  the  disk  0.  If  the  disk  turns 
from  B  to  A,  and  if  the  current  flows  in  the  same  direction,  the 
electromotive  force  induced  in  the  disk  is  directed  from  the  centre 
to  the  periphery.  A  spring  is  placed  at  the  point  B,  and  is  connected 


208  INDUCTION.  [CHAP.  x. 

by  the  conductor  BGDEO  with  a  rod  which  touches  the  disk  at  its 
centre.  DE  is  the  conductor  whose  resistance  E  is  to  be  determined. 
If  the  current  from  the  battery  also  flows  through  the  conductor 
DE,  we  may  so  adjust  the  angular  velocity  of  the  disk  that  no 
current  flows  through  the  conductor  which  connects  E  with  the 
disk.  When  this  condition  is  attained,  the  galvanometer  needle 
shows  no  deflection.  If  the  electromotive  force  induced  in  the 
disk  is  represented  by  e,  we  then  have  e  =  Ei,  where  i  denotes  the 
current  flowing  through  the  resistance.  To  determine  e,  we  consider 
the  disk  replaced  by  a  ring  BAC  and  a  straight  conductor  OA. 
The  circuit  OAGDEO  is  divided  into  the  two  circuits  OAB  and 
BGDEOB.  The  number  of  lines  of  force  passing  through  the  latter 
circuit  is  not  changed  during  the  motion.  On  the  other  hand,  the 
number  passing  through  the  circuit  OAB  increases  in  unit  time  by 

/  mino .  dr  =  iw  I  mr .  dr. 
Jo  JQ 

This  change  in  the  number  of  lines  of  force  gives  the  induced 
electromotive  force.  We  therefore  have 

,-a  -a 

Ei  =  i<ol  mr.dr,     E=wl  mr.dr. 

Jo  Jo 

If  n  is   the    number   of  revolutions    made    by   the    disk    in    one 
second,  we  have  o>  =  'lirn  and 

n .  2irr  .  dr. 


The  integral  gives  the  coefficient  of  mutual  induction  M  between 
the  coil  EF  and  the  disk  ABC,  or  the  number  of  lines  of  force 
which  pass  through  the  disk  if  a  unit  current  is  flowing  in  the 
coil.  Hence  we  have  E  =  nM.  Therefore  to  measure  the  resistance 
we  determine  the  number  of  revolutions  per  second  which  must  be 
given  to  the  plate  in  order  that  no  current  shall  flow  through  G. 


SECTION  LXXXVIII.    FUNDAMENTAL  EQUATIONS  OF  INDUCTION. 

We  have  hitherto  determined  the  electromotive  force  induced  in 
the  conductor  s  by  the  change  in  the  number  N  of  lines  of  force 
which  are  enclosed  by  s.  N  represents  the  number  of  lines  of  force 
which  pass  through  an  arbitrary  surface  containing  the  conductor  S ; 
it  is  therefore  determined  by  the  conductor  alone.  Hence  three 
quantities,  F,  G,  H,  can  be  so  determined  that  the  line  integral 
l(F .  dx/ds  +  G .  dy/ds  +  H.  dz/ds)ds 


SECT.  LXX  xvm.]         EQUATIONS  OF   INDUCTION.  209 

is  equal  to  the  surface  integral  N=\\(al  +  bm  +  cn}dS.     It  is  necessary 
for  this,  by  VI.  (f),  that 
(a)        a  =  a#/3y-3fiya?,  b  =  'dF/'dz--dH/'dx,  c 
We  may  also  obtain  these  equations 
of  condition  by  the  assumption  that, 
for  example,  dS  is  equal  to  the  surface- 
element  di/dz,  which  is  represented  by 
OB  DC  (Fig.  106).     The  line  integral 
then  becomes 


,  dz)dy  -  Hdz 


=  (dH/dt/-'dG!'dz)dydz.  FIG.  106. 

Since  the  surface  integral  in  this  case  is  a  .  dydz,  we  obtain  the  first 
of  equations  (a). 

If  these  equations  are  supposed  solved  for  F,  G,  H,  we  have 

(b)  N=  ^(F.  dx/ds  +  G  .  dy/ds  +  H.  dz/ds)ds. 

If  the  part  of  the  region  considered  is  at  rest,  the  induced  electro- 
motive force  e  is  determined  by 

(c)  e  =  -  dNjdt  =  -  \(dF/dt  .  dx/ds  +  dG/dt  .  dy/ds  +  dH/dt  .  dz/ds)ds. 

If  we  set  (d)  P=  -dF/dt,  Q  =  -dG/dt,  R=  -dH/dt,  we  may  consider 
P,  Q,  E,  as  the  components  of  the  electromotive  force  (£,  and  if 
the  integration  is  extended  along  the  whole  conductor,  we  obtain 
as  an  expression  for  the  total  induced  electromotive  force 

(e)  e  =  $(P.dx+Q.dy  +  R.dz). 

These  components  may  also  be  determined  directly  by  variation  of 
the  value  of  the  magnetic  induction,  whose  components  are  a,  b,  c. 
We  thus  obtain  from  (a)  and  (d) 

-  da/dt  =  ?>R[dy  -  VQfdz, 

(f)  -  db/dt 


Suppose  that  the  electrical  current,  as  has  been  remarked  in 
LXXXI.,  is  made  up  of  two  parts,  namely,  the  current  of  conduction, 
which  is  proportional  to  the  electromotive  force,  whose  components 
are  p,  q,  r,  and  the  current  due  to  the  changes  in  the  electrical 
polarization,  whose  components  are  /,  g,  h.  Then  for  the  components 
of  the  total  electrical  current,  we  obtain 

u  =p  +  df/dt,   v  =  q  +  dg/dt,   w  =  r  +  dh/dt. 
o 


210  INDUCTION.  [CHAP.  x. 

If  C  represents  the  conductivity,  we  have  (g)  p  =  CP,  q=CQ,  r=CR 
and  (h) '  f=kP/4v,  g  =  kQ''±ir,  h  =  kR/4Tr,  where  k  is  the  dielectric 
constant  measured  in  electromagnetic  units.  Hence  we  have 
(i)  u=CP  +  k/4ir.dP/dt,  v=CQ  +  k/±ir.dQ/dt,  w=CR  +  kjlir  .dR/dt. 
From  LXXVI.  the  components  of  the  magnetic  induction  33  and  the 
components  of  the  magnetic  force  ^p  are  connected  by  the  equations 
(k)  a  =  p.a,  &  =  ju/3,  e  =  /ry,  where  p.  is  the  magnetic  permeability  of 
the  substances. 

We   have   already  found   the   following   equations   connecting   the 
magnetic  force  and  the  components  of  current  (cf.  LXXX.) : 
(1)      47Ttt  =  3y/3y-30/3s,    4™  =  'dafiz  -  "dy/'dx,   4-n-w  =  3/3/3a;  -  "daffy. 


SECTION  LXXXIX.    ELECTRO-KINETIC  ENERGY. 

By  LXXXV.  (k),  the  electro-kinetic  energy  of  any  system  of  con- 
ductors is  expressed  by  T=^Ni.  By  using  LXXXVIII.  (b)  we 
obtain  T=^\(Fi.dx/ds  +  Gi  .dy/ds  +  Hi.  dz/ds)ds.  If  u,  v,  w,  are  the 
components  of  current,  a  current  i,  in  a  conductor  whose  cross-section 
is  A,  may  be  expressed  by  i  =  A.  -Ju2  +  v*  +  w2,  and  we  have  also 
i.dx/ds  =  uA,  etc.  If  we  set  dxdydz  =  A.ds,  we  obtain 

(b)  T=  $  J  J JCFw  +  Gv  +  Hw}dxdydz. 

If  u,  v,  w  are  here  expressed  by  the  components  of  the  magnetic  force 
[LXXXVIII.  (1)],  we  have 

T=  1/87T  .  J  Jf  |7(3y/3y  -  3/3/3«)  +  GCdafdz  -  Vy/dx) 

+  HCdfi/'dx  -  'da/'dy)] .  dxdydz. 

If  the  separate  terms  are  integrated  by  parts,  and  the  integration 
is  extended  over  the  whole  infinite  region,  it  follows,  since  at  the 
boundary  of  this  region  a,  (3,  y  are  infinitely  small  of  the  third 
order,  that 

f  \\H .  3a/3?/ .  dxdydz  =  -  f  J  fa .  ^Hffy .  dxdydz 
and  Jf  f  G .  Vafdz  .  dxdydz  =  -  J  J  J  a  .  VGfdz .  dxdydz. 

Analogous  expressions  hold  for  the  other  integrals. 
By  reference  to  LXXXVIII.  we  therefore  obtain 

(c)  T=  I/Sir .  f  f  f  (oo  +  /3b  +  yc)dxdydz. 

If  no  magnets  or  no  bodies  which  can  acquire  an  appreciable 
magnetization  are  present  in  the  region,  we  have  a  =  a,  b  =  f3,  c  =  y, 
and  (d)  T=  I/Sir.  Hl 


SECT,  xc.]  ABSOLUTE   UNITS.  211 

SECTION  XC.    ABSOLUTE  UNITS. 

In  Physics  we  generally  take  the  centimetre,  (/mm  and  second  as 
units  of  length,  mass  and  time  respectively.  These  are  the  units  which 
are  used  in  the  theory  of  electricity.  We  will  now  proceed  to 
express  in  terms  of  them  the  most  important  electrical  and  magnetic 
quantities.  They  are  designated  by  the  symbols  L,  M,  T  respectively 
(cf.  Introduction). 

(a)  The  Electrostatic  System  of  Units. 

In  electrostatics  the  force  £y,  with  which  two  quantities  of  electricity 
^  and  e.2  act  on  each  other,  is  expressed  by  (cf.  LIII.)  ^  =  «1«2/r2» 
where  r  is  the  distance  between  the  quantities.     If  «1  =  e2  =  «,  we  have 
e  =  rJ$,  and  hence  the  dimensions  of  a  quantity  of  electricity  e  are 
[e]  =  [LL*M*T-*]  =  [tfM*T-*]. 

The  electrical  force  F,  which  acts  on  a  unit  quantity  of  electricity, 
has  the  dimensions  of  the  quantity  e/r2,  and  therefore 


TJie  electrostatic  potential  ¥  (cf.  LIV.)  has  the  dimensions  of  the 
quantity  e/r,  and  therefore  [¥]  =  [L^M^T^/L]  =  [L*M*T~1]. 

The  capacity  C  [cf.  LV.  (g)]  has  the  dimensions  of  the  quantity  «/¥", 
and  therefore  [£]  =  [£].  The  capacity,  therefore,  has  the  dimensions 
of  a  length. 

The  surface-density  cr  (cf.  LIV.)  is  the  quantity  of  electricity  present 
on  the  unit  of  surface,  and  therefore  [<r]  =  [Liflf*  T~l/Lz]  =  [L-*M*T-*]. 
The  dimensions  of  surface-density  are  the  same  as  those  of  electrical 
force  [cf.  LV.  (e)].  Since  the  electrical  displacement  or  polarization 
(cf.  LXV.)  is  the  quantity  of  electricity  which  passes  through  unit 
of  surface  in  the  dielectric,  its  dimensions  are  also  [lT^M^T~1^,  The 
ratio  between  the  electrical  displacement  ®  and  the  electrical  force 
is  expressed  by  KJ^TT,  where  K  is  the  dielectric  constant.  K  is  there- 
fore a  mere  number. 

The  electrical  energy  W  [cf.  LXI.  (a)]  is  measured  by  the  product 
of  the  difference  of  potential  and  the  quantity  of  electricity.  Hence 
[JF]  =  [/AMT~2].  These  are  also  the  dimensions  of  all  other  forms 
of  energy. 

(b)  The  Electromagnetic  System  of  Units. 

Two  magnet  poles  which  contain  the  quantities  of  magnetism  /^ 
and  f*2,  repel  each  other  with  the  force  F  [cf.  LXVIII.  (a)],  /'=/*1/*2/r2. 


212  INDUCTION.  [CHAP.  x. 

Hence  quantity  of  magnetism  has  the  same  dimensions  in  the  electro- 
magnetic system  as  quantity  of  electricity  in  the  electrostatic  system. 
This  relation  holds  throughout  between  the  two  systems  for  corre- 
sponding quantities  in  electrostatics  and  magnetism  ;  we  will  therefore 
not  consider  each  case  separately. 

The  dimensions  of  the  strength  of  the  electrical  current  are  deter- 
mined from  Biot  and  Savart's  Law  (cf.  LXXVIIL).  According  to 
this  law  the  force  K,  with  which  a  current  -element  ds  acts  on  a  magnet 
pole  containing  the  quantity  of  magnetism  p.,  is  K=p.  ids.  sin0/V2. 
Since  K  is  a  mechanical  force,  the  dimensions  of  the  current-strength 
»  are 


Since  the  quantity  of  electricity  q  may  be  considered  as  a  product 
of  current  strength  and  time,  we  have  [q]  =  [ZA/lf*]. 

The  electromotive  force  e  which  arises  in  a  closed  conductor  has  been 
defined  [cf.  LXXXV.  (b)]  by  e=  -dNjdt,  where  N=\\(al  +  bm  +  cm)dS. 

Since  magnetic  induction,  by  definition,  has  the  same  dimensions 
as  magnetic  force,  the  dimensions  of  electromotive  force  are 


In  this  system  the  electromotive  force  per  unit  of  length  of  the 
conductor  may  be  considered  as  the  measure  of  the  electrical  force, 
whose  components  are  P,  Q,  R.  Hence  the  dimensions  of  electrical 


According  to  Ohm's  law  the  resistance  is  equal  to  the  ratio 
between  the  electromotive  force  in  the  conductor  and  the  current- 
strength  ;  the  dimensions  of  resistance  are  therefore 


that  is,  the  dimensions  of  resistance  are  the  same  as  those  of  velocity. 

Surface-density  of  electricity  and  electrical  polarization  have  the  dimen- 
sions of  a  quantity  of  electricity  divided  by  an  area,  and  are  therefore 


By  LXXXVIII.  (h)  the  dielectric  constant  k  in  this  system  has 
he  same  dimensions  as  the  ratio  between  the  dielectric  polarization 
and  the  electrical  force,  or  [L-*M*/L*M*T-*]  =  [L^-T2]. 


SECT,  xc.]  ABSOLUTE  UNITS.  213 

(c)  Comparison  of  the  Two  Systems. 

If  we  measure  a  quantity  of  electricity  electrostatically  by  Coulomb's 
torsion  balance  and  electromagnetically  by  a  galvanometer,  we  have 
two  values  for  the  same  quantity.  In  the  first  system  this  quantity 
is  expressed  by  e  .  [I$M*T~l],  in  the  second  system  by  q  .  [L^M?]  ;  the 
ratio  V  between  these  expressions  is  V=  e/q  .  [LT~1]. 

This  ratio  is  therefore  a  velocity.  It  was  first  measured  by  Weber 
and  Kohlrausch.  Its  value,  as  found  by  them,  is  F=3,1.1010. 
which  is  very  closely  the  velocity  3,0.  1010  of  light  in  air.  Subsequent 
experiments  have  made  it  probable  that  V  is  actually  the  same  as 
the  velocity  of  light.  It  is  thus  shown  that  an  electromagnetic  unit 
of  electricity  is  equal  to  V  electrostatic  units. 

If  a  certain  quantity  of  electricity  flows  through  a  portion  AB  of 
a  conductor  it  produces  heat  in  the  conductor,  which,  considered  as 
energy,  must  be  independent  of  the  system  of  measurement  employed. 
The  energy  in  electrostatic  units  is  e  .  ^,,  in  electromagnetic  units 
q*Pmi  where  ""P,  represents  the  difference  of  potential  between  A  and  B 
in  electrostatic  units,  "*Pm  the  same  difference  of  potential  in  electro- 
magnetic units.  Hence  we  have  e^f,  =  cflf  m.  We  have  shown  that 
e=  Vq,  and  therefore  ^rm  =  ^ft.  V;  that  is,  an  electrostatic  unit  of  potential 
is  equal  to  V  electromagnetic  units  of  potential. 

If  Ave  designate  electrical  force  in  the  electrostatic  system  by  Fn 
in  the  electromagnetic  system  by  Fm,  the  difference  of  potential 
between  two  points,  which  are  distant  from  each  other  by  dx,  is  in 
the  first  system  *Pt  =  Ft.  dx,  in  the  second  system  *Pm  —  Fm  .  dx.  Since 
"*Pm  =  ty,  .  V  we  have  Fm  =  V  .  Ft.  Hence  one  unit  of  electrostatic  force 
is  equal  to  V  units  of  electromagnetic  force. 

The  dielectric  polarization  !£)  is  connected  with  the  force  F.  by 
the  equation  LXV.  (d)  (£)=K/'±ir.  Fn  if  the  electrostatic  system  is 
used.  In  the  electromagnetic  system  this  equation  takes  the  form 
/=  k/4ir  .  Fm.  Since  ©  and  /  are  quantities  of  electricity  divided  by 
areas,  and  since  e  =  qpr,  we  have  f£)  =  Vf.  Since  Fm  is  an  electrical 
force  measured  in  electromagnetic  units,  we  have 

k  =  tirf>Fm  =  47r£)/  V*F.  =  K!  V2. 

Hence  in  the  electromagnetic  system  the  equations  connecting  the 
components  of  dielectric  polarization  with  the  electrical  force  are 


We  thus  obtain  ground  for  the  assumption  that  V\<J]L  is  the  velocity 
of  light  in  a  medium  whose  dielectric  constant  is  K. 


214  INDUCTION.  [CHAP.  x.  SECT.  xc. 

(d)  Practical  Units. 

In  practical  work  these  absolute  units  are  often  discarded  in  favour 
of  others,  called  practical  units.  The  unit  of  current-strength  in  this 
practical  system  is  the  ampere,  equal  to  10"1  electromagnetic  units 
of  current.  The  unit  of  resistance  is  the  ohm,  equal  to  109  absolute 
units  of  resistance.  The  ohm  is  nearly  equal  to  the  resistance  of  a 
column  of  mercury,  whose  cross-section  is  1  sq.  mm.  and  whose  length 
is  106,3  cm.  The  unit  of  electromotive  force  then  follows  from  Ohm's 
law;  it  is  called  the  wit,  and  is  equal  to  10s  absolute  units  of  electro- 
motive force. 

The  unit  of  quantity  of  electricity  is  the  quantity  which  flows 
in  one  second  through  any  cross-section  of  a  conductor  in  which  the 
current-strength  is  an  ampere.  This  unit  is  called  a  coulomb.  The 
capacity  of  a  condenser,  one  of  whose  coatings  is  charged  with  one 
coulomb  when  the  difference  of  potential  between  its  coatings  is  one 
volt,  is  called  a.  farad;  it  is  equal  to  IQ~1/IQ8=  10~9  absolute  units 
of  capacity.  A  body  whose  capacity  is  unity  in  the  absolute  electro- 
magnetic system  must  be  charged  with  unit  quantity  of  electricity 
in  order  to  reach  unit  potential.  These  quantities  are  the  same  as 
the  quantity  of  electricity  V  and  the  potential  \/V  in  the  electro- 
static system.  The  electrostatic  capacity  of  the  body  is  therefore 
V-.  It  follows  from  this  that  a  farad  is  equal  to  F^/IO9  electro- 
static units  of  capacity.  Since  this  capacity  is  very  great,  the 
millionth  of  a  farad  or  a  microfarad,  is  generally  used  as  the  practical 
unit  of  capacity. 


CHAPTER  XL 

ELECTRICAL  OSCILLATIONS. 
SECTION  XCI.    OSCILLATIONS  IN  A  CONDUCTOR. 

IF  a  conductor  is  traversed  by  alternating  currents,  that  is,  by  currents 
which  reverse  their  directions  at  regular  intervals,  we  say  that  electrical 
oscillations  exist  in  the  conductor.  Such  alternating  currents  may  be 
produced  by  induction,  as  in  the  well-knoAvn  experiments  of  Feddersen. 
Let  AB  (Fig.  107)  be  a  condenser  whose  plates  are  joined  by  con- 
ducting wires  to  two  small  metallic  spheres  C  and 
D.  If  the  condenser  is  so  charged  that  A  has  the 
potential  ¥  and  B  the  potential  zero,  and  if  the 
distance  CD  is  sufficiently  diminished,  a  spark  will 
pass  from  C  to  D.  Careful  investigation  has  proved 
that  this  spark  consists  of  a  series  of  sparks,  which 
correspond  to  currents  in  opposite  directions.  Hence 
electrical  oscillations  will  be  set  up  by  the  discharge. 
But  if  the  current-strength  i  varies  in  the  conducting 
wires  AC  and  DB,  an  electromotive  force  will  be 
induced  in  them,  which,  from  LXXXV.,  is  equal  to 
—  L .  di/dt,  where  L  is  the  coefficient  of  self-induction. 
If  we  represent  the  electrical  resistance  in  the  conducting  wires  and 
in  the  spark-gap  by  r,  the  current-strength  i  is  given  by 

(a)  V-L.di/dt  =  r.i. 

If  c  is  the  capacity  of  the  condenser  in  electromagnetic  units,  the 
charge  q  of  the  condenser  at  the  time  t  is  c*P;  at  the  time  t  +  dt,  it 
is  c*P  +  c .  d^/dt .  dt.  Hence  we  have 

(b)  c¥  =  c¥  +  c .  d^jdt  .dt  +  idt;  i  =  -  c .  dV/dt. 

From  (a)  and  (b)  it  follows  that  Lc .  d-i/dt2  +  cr .  di/dt +  i  =  0.  An 
integral  of  this  equation  is  i  =  A.  emit  +  B .  e"^\  where  ml  and  m2  are 

215 


216  ELECTRICAL  OSCILLATIONS.  [CHAP.  xi. 

the  roots  of  the  equation  Lcm2  +  crm  +1  =  0.  The  roots  of  this 
equation  are  m=  -  r/2L  ±  */r2/4:L-  -  l/Lc.  If  r  is  sufficiently  small,  the 
roots  m  will  be  imaginary  and  we  will  have,  neglecting  r2/4Z2, 


Using  the  real  part  and  the  real  coefficient  of  the  imaginary  part 
as  particular  solutions  we  can  then  express  i  by 

i  =  e-*»L  .  [A1  .  sin(//VZ<J)  +  V  •  cos(*/VZc)]. 

Hence  the  current-strength  changes  periodically  and  diminishes  with  the 
time.  This  equation  shows  that  the  period  T  of  the  oscillation  is 
given  by  T/jLc  =  27r)  and  hence  T=2irsjLc.  Therefore  as  the  capacity 
of  the  condenser  is  diminished,  the  period  T  diminishes  also. 

The  amplitude  of  the  oscillations  is  proportional  to  e~rt/2i  ;  it  there- 
fore diminishes  continually  as  the  time  increases.  The  ratio  between 
the  amplitudes  of  two  successive  oscillations,  the  so-called  damping 
coefficient,  is  e~rt/2Z  :  g  -•*+*)/«  =  erT/'2L.  The  damping  is  therefore  increased 
as  the  resistance  of  the  connections  is  increased,  and  with  a  given 
period  it  is  diminished  as  the  self-induction  L  is  increased.  If  we 
substitute  for  T  its  value  T=  27rx/Lc,  the  damping  coefficient  becomes 
eirrVw..  Hence  the  effect  of  damping  on  the  oscillations  is  diminished 
when  the  capacity  of  the  discharging  conductor  is  diminished.  Instead 
of  making  observations  on  the  damping  coefficient  itself,  we  generally 
deal  with  its  natural  logarithm,  the  so-called  logarithmic  decrement  8. 
We  have  8  =  rT/2L  =  irr  .  Jc/L.  H.  Hertz  obtained  very  rapid  oscilla- 
tions by  the  use  of  the  apparatus  represented  in  Fig.  108.  It  con- 

/^~x.4         C  T)        B  s^sts  °^  two  ^ar§e  sPneres  °f  equal  size  at 

(  QO  ^     j    A  and  B,  which  are  fastened  on  the  ends 

^—^     of  the  copper  rods  AC  and  BD.      The 
other  ends  of  the  rods  AC  and  DB  carry 
small  spheres,  separated  by  a  distance  of 
about    1    centimetre.      The    spheres   are 
FlG-  m  charged  by  the  induction  coil  EF  '•    the 

discharge  occurs  in  the  gap  between  the  spheres  C  and  D.  If  at 
a  definite  instant  a  current  of  strength  i  passes  from  A  to  B,  and 
if  the  potentials  of  A  and  B  have  respectively  the  values  '*P1  and 
¥"2,  then  ^rl  -  ^2  -  L  .  di/dt  =  ri.  If  the  capacity  of  each  of  the  large 
spheres  is  represented  by  c,  we  have  cl  .  d^PJdt  =  -i,  cl  .  d^2/dt  =  i, 
or  by  using  c  =  \cv  i=  -c.  d(tirl  -  ^2)/dt.  The  current-strength  is 
therefore  given  by  the  same  differential  equation  as  before,  and  we 
obtain  from  it  the  same  expression  for  the  period  T. 


SECT,  xcn.]  CALCULATION   OF  THE   PERIOD.  217 


SECTION  XCIL     CALCULATION  OF  THE  PERIOD. 

In  order  to  determine  the  period,  we  must  first  determine  the 
coefficient  of  self-induction.  The  method  of  determining  coefficients  of 
self-induction  for  closed  conducting  circuits  has  already  been  given. 
In  the  present  case  we  have  to  deal  partly  with  actual  currents  in 
the  cylindrical  conductors,  partly  with  polarization  or  displacement 
currents  in  the  surrounding  dielectric.  The  principal  effect  must 
be  due  to  the  induction  in  the  conductor  itself,  since  in  it  the  distance 
between  the  inducing  and  the  induced  currents  is  least.  The  full 
treatment  of  the  question  would  be  very  difficult,  because  the  current- 
strengths  in  the  different  parts  of  the  cross-section  are  not  equal. 
We  will  therefore  neglect  these  differences  in  the  subsequent  dis- 
cussion, and  calculate  the  coefficient  of  self-induction  L  in  a  cylinder 
on  the  assumption  that  the  current-strength  in  all  parts  of  the  cross- 
section  is  the  same.  According  to  F.  Neumann,  the  electromotive 
force  E  induced  in  a  conductor  s'  by  the  action  of  a  variable  current 
i  flowing  through  another  conductor  s,  is  determined  by  the  variation 
of  the  integral  Li,  where  L  =  ^cose/r.dsds'.  In  this  integral,  which 
is  to  be  taken  over  all  the  elements  of  both  conductors,  e  denotes  the 
angle  between  ds  and  ds',  and  r  the  distance  between  ds  and  ds. 

Let  AB  and  CD  be  two  parallel  lines  (Fig.  109)  which  together 
with  AC  and  BD  form  a  rectangle.  We  set  AB=CD  =  l,  and  CF=s'. 


f- 

fJs' 

1 
1 

>;>';"     i*    ""G 

i 

A 

fc        b,       G              12 

FIG.  109. 

In  the  present  case  the  integral  which  is  to  be  evaluated  becomes 

/  |  ds.ds'/r,  because  (Fig.  109)  cose  =  l.     In  order  to  find  first  the 

-'o  ^o 

value  of  the  integral  P=  I  ds/r  we  draw  FG  perpendicular  to  CD 
and  AB,  and  write  FG  =  a.  If  AG  =  bv  BG  =  b9  FA  =  r1,  FB  =  r.2, 
we  have  P  =  \  ds/r  +  I  ds/r,  where  s  is  the  distance  of  any  point  on 
AB  from  G.  Now  since 

\dsfr  =  Ids/Jat  +  s*  =  log  nat(s/a  +  Jl+sz/az) 
we  obtain  P  =  \ogna,t[(rl  +  bl)(r2  +  b2)/a2'] 

=  log  nsA[(AF  +  AG)(BF+  BG)/a?]. 


218  ELECTRICAL  OSCILLATIONS.  [CHAP.  xi. 

If  a  is  very  small  in  comparison  with  I,  we  may  write 
AF=AG=*'  and   FB  =  GB  =  l-s', 

and  have  />=lognat[4s'(^  -s')/«2].  We  then  calculate  the  value  of 
the  integral  \P.ds  and  obtain 

P.ds'  =  2l.  [log  nat(2//a)  -  1]. 

We  will  now  calculate  the  coefficient  of  self-induction  of  a  wire  with 
circular  cross-section.  Let  AB  (Fig.  110)  be  the  cross-section  of  a 
cylindrical  conductor,  of  length  /  and  radius 
R.  Let  the  current-density  u  be  constant 
throughout  the  cross-section.  The  current- 
strength  i  is  then  given  by  i  =  r-iru.  We 
will  first  consider  the  inductive  action  due 
to  a  filament  D  whose  cross-section  is  ds; 
this  filament  is  supposed  to  act  on  a  line 
which  is  parallel  to  the  axis  of  the  cylinder, 
FIG.  110.  an(j  passes  through  the  point  c.  If  DC=a, 

the  inductive  action  is  obtained  by  the  variation  of  the  integral 

u.dS.  2/(log  nat(2//a)  -l)  =  u.dS.  (M+Nlog  nat  a), 

where,  for  the  sake  of  conciseness,  we  use  M  and  N  as  symbols 
for  the  quantities  Jl/=2£(lognat2i-l)  and  N=-'2l.  We  must 
distinguish  between  two  cases  :  OD  =  r  may  be  either  greater  or  less 
than  OC=rv  First  let  r>rr  The  elements  dS  may  be  taken  so 
as  to  form  the  surface  of  a  ring  whose  area  is  2irr  .  dr.  From  the 
demonstration  of  XIII.  log  a  equals  the  logarithm  of  half  the  sum  of 
the  greatest  and  least  values  which  a  can  take.  These  values  are 
respectively  r  +  rl  and  r  -  rr  The  mean  value  required  is  therefore 
logr.  Hence  we  have  the  integral 

*'2irr  .  dr  .  (M+  N\og  nat  r)  =  mt{M(Sr  -  rtf 

+  N[R2  log  nat  R  -  r*  log  nat  rx  -  $(R2  -  r-f)]  }. 

For  that  part  of  the  cylinder  whose  distance  from  the  axis  is  less 
than  1\  the  mean  value  of  the  greatest  and  least  values  of  a  will 
equal  rr  Hence  the  integral  for  this  part  is 

u  I  Jfcw  .  dr  .  (M+  N\og  nat  r:)  =  Tru(Mi\2  +  Ni\z  .  log  nat  rx). 
The  sum  of  both  integrals  is 


[ 

-' 


SECT,  xcn.]  CALCULATION  OF  THE  PERIOD.  219 

In  order  to  obtain  the  mean  value  of  this  quantity  for  all  the  filaments 
composing  the  cylinder,  we  need  only  find  the  mean  value  of  rx2, 
since  all  the  other  quantities  are  constant.  But  since 


it  follows  that  the  mean  value  sought  is  iruR2(M  +  .ZV(log.E  -  J)}.    If 

we  introduce  into  this  equation  the  values 

M=  2J(log  21-1)  and  N=  -  21 

and  set  iruR2  =  i,  we  obtain  for  the  quantity,  the  variation  of  which 
gives  the  self-induction,  2/i(log(2?/^)  -  f  ).  We  therefore  obtain  for  the 
quantity  L,  L=  '2l(\og(2l/E)  -  f).  In  Hertz's  investigation,  1=150, 
.#  =  0,25  and  therefore  Z=1902,  where  all  lengths  are  expressed  in 
centimetres. 

In  order  to  calculate  the  period  of  oscillation,  we  will  next  determine 
the  capacity  of  a  sphere  with  a  radius  of  15  cm.,  such  as  Hertz  used. 
If  Q  is  its  charge  and  >F  its  potential,  the  capacity  C  in  electrostatic 
units  is  C=Q/y.  Eepresenting  the  charge  and  the  potential  in 
electromagnetic  units  by  Q'  and  ¥"  respectively,  and  using  V=3.  1010, 
the  velocity  of  light  in  vacuo,  we  have  Q=7(^  and  ¥  =  ¥7^".  The 
capacity  c  in  electromagnetic  units  is  therefore  c=Q'lW  =  C/V2. 
The  period  of  oscillation  is  then  given  by  T=2irjLC/V. 

Using  the  symbol  introduced  at  the  end  of  XCL,  we  have  c  =  ^cv 
where  Cj  is  the  capacity  in  electromagnetic  units  of  each  of  the  two 
large  spheres.     Hence  we  must  set  £=15/2,  and  obtain  r=2,5/108 
seconds.     The  corresponding  wave  length  in  air  is 
2,5.10-8.3.1010 


SECTION  XCIII.    THE  FUNDAMENTAL  EQUATIONS  FOR  ELECTRICAL 
INSULATORS  OR  DIELECTRICS. 

Maxwell  shows  that  it  follows,  as  a  consequence  of  his  theoretical 
views  of  the  nature  of  electricity,  that  a  change  in  the  electrical 
polarization  of  the  dielectric  can  set  up  electrical  oscillations.  The 
results  which  he  obtained  are  so  important  that  we  will  consider 
some  of  them  here.  For  this  purpose  we  will  follow  Hertz  in  sub- 
stituting in  the  fundamental  equations  of  LXXXVIII.  electrostatic 
units  for  the  electrical  quantities,  while  the  magnetic  quantities  shall 
be  measured  in  electromagnetic  units.  The  quantity  of  electricity 
which  is  displaced  by  the  electrical  force  at  a  point  in  the  dielectric 


220  ELECTRICAL  OSCILLATIONS.  [CHAP.  xi. 

through  a  surface-element  which  stands  perpendicular  to  the  direction 
of  the  force,  is,  according  to  LXV.,  equal  to  K!±TT  multiplied  by 
the  magnitude  of  the  force  F.  Representing  the  components  of  the 
electrical  displacement  by  /,  g,  h,  and  the  components  of  the  electrical 
force  by  X,  Y,  Z,  we  have  f=KXJ^  g  =  KY/4ir,  li  =  KZ^.  If  the 
component  X  increases  by  dx  in  the  time  dt,  the  quantity  of  electricity 
df  flows  through  unit  area  in  the  direction  of  the  z-axis  ;  the  com- 
ponent u  of  the  current-strength  in  the  dielectric  is  eqiial  to  dfjdt, 
and  we  have 

(a)  u  =  K/±ir.'dXI-dt,    v=K/4ir.c)YI'dt,    w  =  K/4ir  .  -dZfdt. 

The  equations  LXXX.  (a)  express  the  fact  that  the  work  done  by 
the  magnetic  forces  in  consequence  of  the  movement  of  a  unit  pole 
about  the  current  is  equal  to  the  current-strength  multiplied  by  4?r. 
If  the  current-strength  is  measured  in  electrostatic  units,  we  have 

(b)  4irM/F=3y/3y-a)8/as,  4ir»/F=3a/3«-3y/aB,  4irw/F=3/3/az-'9a/3y, 
since   the   electromagnetic   unit   of  quantity   of  electricity  equals   V 
electrostatic  units. 

The  electromotive  force  induced  equals  -  dNjdt,  if  N  represents  the 
number  of  lines  of  force  enclosed  by  the  circuit.  Since  the  electro- 
motive force  in  electromagnetic  units  equals  the  electromotive  force 
in  electrostatic  units  multiplied  by  F,  we  have  from  LXXXVIII. 
(f)  and  (k) 


f  -  >*/  F.  da/3/  =  ?)Z  /-dy  -  -dY/-dz  ; 
-  * 


(c)  -  ,*/  F  .  3(3^  =  -dXfdz  -  'dZ/'dx  • 

(  -  /*/  V. 


From  (a)  and  (b)  we  obtain 

KIV.  'dX/'dt 

Jf/F.  -dYpt  =  'da/'dz  - 


If  we  now  set  J='dX/'dx  +  ?>Y/'dy  +  ~dZfdz,  we  obtain  from  (c)  and  (d) 
fj.K/F'2.d2X/'dl2  =  V2X-'dJ/ox.  If  we  are  dealing  with  a  region  in 
which  K  is  constant,  K/4ir  .  /=  ?//3.c  +  'bgj'dy  +  "dhfdz.  If  there  is  no 
electrical  distribution  in  the  region,  we  have,  from  LXVI.  (d),  J=0; 
and  hence 

(e)  pKIF*.&XIW=V*X',  pK/r*.&Y/df2  =  V*Y;  nK/r2.VZ/-dP  =  VZ. 
These  equations,  in  connection  with  equations  (c)  and  (d),  give 

(f)  fiK/F^.^a/'df  =  V2a-  (JiK/F2.-d2pfdt*  =  r-p;  nKj  V*-  .  32y/^2  =  V2y. 
at  the   same  time   3o/daf-f  dj9/3y+dy/d*««0  from   LXXVI.,   if  u  is 
constant. 


SECT,  xciv.]       PLANE   WAVES   IN   THE   DIELECTRIC.  221 

From  LXVII.  (g),  the  electrical  energy  W  is  expressed  by 
(g)  W=  l/8ir .  J Jf K(X*  +  Y*  +  Z*)dxdydz. 

The  electrokinetic  energy  T,  according  to  LXXXIX.  (c),  is 
(h)  T=l/8ir.  J{jM(a2  +  p2  +  y2)dxdydz. 


SECTION  XCIV.     PLANE  WAVES  IN  THE  DIELECTRIC. 

We  will  now  investigate  the  movement  of  plane  waves  in  a  dielectric. 
Let  the  plane  waves  be  parallel  to  the  ys-p\&ne.  The  components 
of  the  electrical  force  are  then  functions  of  .r  only,  and  from  the 
equations  XCIII.  (e),  we  have 


At  the  same  time  also  'dXj'dx  +  VYfdy  +  'dZ/'dz  =  0.  Since  Y  and  Z 
are  independent  of  y  and  zt  we  have  'dX/'dx  =  0,  and  since,  in  this 
case,  the  only  forces  which  occur  are  periodic,  X=0.  The  direction 
of  the  electrical  force  is  therefore  parallel  to  the  plane  of  the  wave.  By 
a  rotation  of  the  coordinate  axes  we  can  make  the  y-axis  coincide 
with  the  resultant  of  the  components  Y  and  Z.  We  therefore  need 
to  discuss  only  the  equation  /JT/F2.  327/9^  =  32F/3a;2.  The  integral 
of  this  equation  is  (a)  Y=bsin[2Tr/T  .  (t  -  a;/o>)],  where  T  is  the  period 
of  oscillation  and  w  the  velocity  of  propagation.  The  differential  equation 
is  satisfied  if  w=  F'/Jp-K.  For  vacuum  /*=  1,  K=  1  ;  hence  V  is  the 
velocity  of  propagation  of  plane  electrical  oscillations  in  vacuo.  For 
ordinary  transparent  bodies,  /*=!.  The  velocity  of  propagation  in 
such  bodies  is  therefore  V/\/K.  Maxwell  assumed  that  electrical 
oscillations  are  identical  with  light  waves.  It  has  been  shown  by 
experiment  that  o>  =  V/N,  where  N  represents  the  index  of  refraction 
of  the  dielectric.  The  electromagnetic  theory  of  light  gives  w  =  Vj^K; 
hence  we  have  K=N2,  that  is,  the  specific  inductive  capacity  of  a  medium 
is  equal  to  the  square  of  its  index  of  refraction.  The  fact  that  this 
theorem  holds  for  a  large  number  of  bodies  is  a  strong  confirmation 
of  Maxwell's  hypothesis.  From  this  hypothesis  almost  all  of  the  pro- 
perties of  light  can  be  deduced. 

According   to   XCIII.    (c)    we   have,    under   the   above   conditions, 
a  =  0,  13  =  0,  and 

(b)  /*y  =  7b/u>  .  sin  [27T/J  .  (t  -  ar/w)]  =  Nb  sin  [lirjl  .  (t  -  x/u)]. 

The  direction  of  the  magnetic  force  is  therefore  parallel  to  the  plane  of 
the  wave  and  perpendicular  to  the  direction  of  the  electrical  force. 


222  ELECTRICAL  OSCILLATIONS.  [CHAP.  xi. 

We  will  supplement  this  discussion  by  the  following  examination 
of  the  relation  between  the  electrical  and  the  magnetic  forces.     Let 
an  electrical  force  act  in  the  y^-plane,  parallel  to  the  axis  Oy  (Fig.  Ill), 
and  suppose  it  to  increase  uniformly  from 
& ,£7    zero  to  YQ  in  one  second.     In  consequence 


of  this  an  electrical  current  v  is  set  up  in 
the  same  direction,  and  because  'dY[dt=  Y0, 
we  have  from  XCIII.  (a), 

This  electrical  current  will  set  up  magnetic 
forces,  which  are  parallel  to  the  ^-axis. 

r  IG.   111.  ___  .__ 

We  Avill  assume  that  the  magnetic  force 

increases  uniformly  from  zero  to  y0.  In  consequence  of  this  an 
electromotive  force  will  be  induced  in  the  surrounding  region.  We 
will  assume  that  the  electrical  and  magnetic  actions  advance  in  one 
second  over  the  distance  O.r  =  w  (Fig.  111). 

The  magnetic  force  decreases  uniformly  from  ;c  =  0  to  x  =  u;  the 
same  statement  holds  for  the  electrical  force  OD=Y0.  The  electrical 
current,  on  the  other  hand,  has  the  same  strength  everywhere  between 
0  and  x.  This  is  explained  by  remarking  that  the  electrical  force 
at  a  point  F,  whose  distance  from  x  equals  l/n.  Ox,  has  acted  only 
during  Ijn  seconds,  and  has,  during  this  time  interval,  increased  from 
0  to  l/n.  Yn;  its  increase  in  one  second  is  therefore  equal  to  F0. 

Let  a  unit  pole  move  in  the  rectangular  path  OzBxO  (Fig.  111). 
The  magnetic  force  y0  acts  only  in  the  path  Oz  and  acts  in  the 
direction  of  motion ;  hence  the  work  done  by  the  magnetic  forces  is 
equal  to  y0.  Oz.  The  quantity  of  current,  measured  in  electromagnetic 
units,  which  the  unit  pole  has  encircled,  is  v/F~.  Oz.  Ox.  From  LXXX. 
we  have  therefore  y0 .  OZ  =  ±TTV.  Oz .  Ox/F~,  or  because  OX  =  (D,  we 
obtain  (d)  Fy0  =  47rz;w. 

The  electromotive  force,  measured  in  electromagnetic  units,  which 
is  induced  by  the  motion  about  a  closed  path,  is  e  =  -  dNjdt,  if  A7" 
represents  the  number  of  lines  of  force  enclosed  by  the  path.  We 
have  therefore  JV=  —  ^e .  dt.  The  mean  value  of  the  electromotive 
force  in  the  direction  Oy  is  | .  F0 .  V,  in  electromagnetic  units.  The 
value  of  the^  before-mentioned  integral,  extended  over  the  rectangular 
path  0>/Cx,  is  |F0F.  Oy.  The  mean  value  of  the  magnetic  force  per- 
pendicular to  the  surface  OyCz  is  Jy0;  hence  the  mean  value  of  the 
magnetic  induction  is  \ .  fj.y0.  We  therefore  have  the  equation 


SECT,  xciv.]      PLANE  WAVES  IN  THE  DIELECTRIC.  223 

and  hence  (e)  FT0  =  /ry0w,  NYQ  =  p.y(),  if  the  index  of  refraction  N  is 
substituted  for  F'/w.  In  this  connection  it  must  be  noticed  that  the 
wave  is  propagated  in  the  direction  of  the  .r-axis,  that  on  our  assump- 
tion the  electrical  force  acts  in  the  direction  of  the  y-axis,  and  that 
then  the  magnetic  force  acts  in  the  direction  of  the  £-axis.  Hence  if 
the  right  hand  is  held  so  as  to  point  in  the  direction  in  which  the  wave 
is  propagated,  with  the  palm  turned  toward  the  direction  of  the  electrical 
force,  the  thumb  will  point  in  the  direction  of  the  magnetic  force.  If  we 
represent  the  magnetic  force  by  M  and  the  electrical  force  by  F, 
NF=  pM.  From  (c)  and  (d)  it  follows  that  Fy0  =  -KT0w.  Substituting 
in  (e)  the  value  of  y0,  we  obtain  V1  =  /JTu>2.  Hence  the  velocity  of 
propagation  is  w  =  VjJp.K.  From  (a)  and  (b)  it  follows  that  the 
relations  (e)  between  the  electrical  and  magnetic  forces  hold  also 
in  the  case  of  plane  waves.  In  vacuo  both  forces  have  the  same 
numerical  value. 


SECTION  XCV.     THE  HERTZIAN  OSCILLATIONS. 

H.  Hertz  succeeded  in  producing  very  rapid  oscillations  in  a  straight 
conductor,  which  also  caused  oscillations  in  the  surrounding  dielectric. 
We  can  form  some  idea  of  the  nature  of  these  oscillations  in  the 
following  manner,  due  to  Hertz : 

Let  the  middle  point  of  a  conductor  coincide  with  the  origin  of 
coordinates,  and  let  the  oscillations  take  place  along  the  2-axis.  The 
magnetic  lines  of  force  are  then  circles,  whose  centres  lie  on  the 
£-axis.  The  electrical  lines  of  force  have  a  more  complicated  form. 
We  start  with  the  differential  equation  XCIII.  (f)  for  the  magnetic 
forces.  For  the  sake  of  conciseness  we  set  VjKp-  =  w. 

We  first  investigate  an  integral  of  the  differential  equation 

(a)  l/o>2.c>2w/3/2  =  V%, 

on  the  hypothesis  that  u  is  a  function  of  t  and  of  r  =  >Jx2  +  y2  -j-  z*. 
We  have  then,  from  XV.  (1),  V%  =  1/r .  32(™)/3r2,  and  therefore 
l/o>2.32(rM)/3/2  =  92(?-tt)/3r2.  If  we  set  k=2Tr/T  and  J  =  2jr/7w,  where 
I7  is  a  constant,  then  (b)  u  =  a/r .  sin  (kt  -  lr)  is  a  particular  integral 
of  the  differential  equation.  The  function  u,  as  well  as  its  differential 
coefficients  taken  with  respect  to  x,  y  and  z,  therefore  satisfies  the 
differential  equations  XCIII.  (f)  for  the  components  of  magnetic 
force  a,  /?,  y.  In  the  case  under  consideration  y  =  0,  and  hence  we 
have  'da/'dx  +  'dft/'dy  =  0.  As  the  simplest  solution  of  the  differential 


224  ELECTRICAL  OSCILLATIONS.  [CHAP.  xi. 

equation  we  obtain  (c)  a  =  -  fPu/'dfdy,  /3  =  ^ufdfdx,  where  the  differ- 
entiation with  respect  to  t  is  introduced  for  use  in  the  subsequent 
calculation.  From  (c)  we  obtain  a  =  -  'd-uj'dfdr  .yjr;  (3  =  'd-u/'dfdr  .  x/r. 
The  resultant  magnetic  force  is  therefore 

M=  a*M/a/3r  .  Jx2  +  f/r  =  Wupfdr  .  sin  6, 

where  0  is  the  angle  between  the  radius  vector  from  the  origin  and 
the  s-axis.  The  force  M  is  perpendicular  to  the  plane  which  contains 
the  point  considered  and  the  .e-axis. 

If  we  set  kt-lr  =  <f>,  we  have  M  =  ka(l.  sin  <£/r  -  cos  <£/r2)  .sin  6.     If 
r  is  very  small  in  comparison  with  l/7  =  o>r/27r,  we  have 


that  is,  the  force  is  determined  by  Biot  and  Savart's  law,  the  oscillations 
in  the  conductor  acting  like  a  current  element. 

For  greater  distances  the  magnetic  force  is 
(d)  M=  4*-%/rW  .  sin  [2-rr/T  .  (t  -  r/o,)]  .  sin  6. 

Hence  the  magnetic  waves  proceed  forward  in  space  with  the  velocity 
of  light. 

We  will  now  calculate  the  electrical  forces.     From  (c)  and  XCIII. 
(d)  we  obtain 


Since  u  depends  only  on  r  and  f,  the  carrying  out  of  the  differentiations 

,  xz/r2 ; 


(e) 

KZI  F=  32t*/3r2 .  (r2  -  z2)/7-2  +  cto/'dr .  (r2  +  z2 
The  electrical  force  E  in  the  direction  r  is  (Xx  +Yy  +  Zz)jr,  and  hence 
from  (e),  KRj  V=  2/r .  'duf'dr .  cos  6  =  -  2a(l  cos  <£/r2  +  sin  ^/r3)  cos  G.  If 
#  =  0,  the  electrical  force  is  tangent  to  a  sphere  whose  radius  is 
determined  from  the  equation  tg(kt  —  Ir)  —  —  Ir.  Let  the  next  spherical 
wave  have  the  radius  r' ;  then  tg(kt  -  Ir)  =  -  Ir'.  From  this  it  follows 
that  tgl(r'  -r)  =  l(r'  -  r)/(l  +  I2rr').  If  the  radius  of  the  waves  is  very 
large  we  may  set  l(r'  -  r)  =  ir.  But  since  /  =  277/X,  where  A.  =  T<a,  we 
have  r'  -  r  =  |X.  We  thus  obtain  at  last  equidistant  spherical  waves. 


SECTION  XCVI.    POYNTING'S  THEOREM. 

Let  there  be  an  electrical  current  i  flowing  from  A  to  B  through 
a  long  cylindrical  conductor  (Fig.  112)  of  circular  cross-section.  There 
is  then  a  magnetic  force  M  acting  at  every  point  in  the  region  around 


SECT,  xcvi.]  POYNTING'S  THEOREM.  225 

the  conductor,  given  by  the  equation  2?rr .  M—  47ri,  where  r  =  OC, 
the  distance  of  the  point  from  the  axis  of  the  cylinder.  We  therefore 
have  M=2i/r.  The  equipotential  surfaces  of  the  electrostatic  field 
of  force  within  the  conductor  are  planes  perpendicular  to  the  axis 
of  the  conductor.  Outside  the  conductor  they  s 

are  likewise  perpendicular  to  the  axis,  at  least 
in  the  vicinity  of  the  conductor.  The  equi- 
potential surfaces  of  the  magnetic  field  of  force 
are  planes  which,  like  the  plane  OF,  contain 
both  the  direction  of  the  electrical  force  and 
the  axis  of  the  conductor.  Let  the  electrical 
force  in  the  surface  of  the  conductor  and  in  its 
vicinity  be  F'.  If  we  designate  by  S  the 
cross-section  of  the  conductor,  and  by  C  its 
conductivity,  we  have  from  Ohm's  law  i/S=CF'. 
The  heat  produced  in  the  conductor  during  one  A 

second  is  determined  as  follows  :  If  the  quantity  ^IG-  1^. 

of  electricity  i  flows  through  the  conductor,  whose  length  we  may 
call  /,  the  electrical  force  does  work  equal  to  F'il.  If  /  represents 
the  mechanical  equivalent  of  heat,  the  quantity  of  heat  thus  developed 
equals  F'iljJ.  The  work  done  by  the  electromotive  force  is  therefore 
F'il  =  \.MrFl, 

Now  Poynting  assumes  that  this  quantity  of  energy  enters  the 
conductor  through  its  surface.  That  portion  of  the  surface  which 
is  to  be  considered  is  equal  to  1-xrl.  The  quantity  of  energy  which 
enters  the  conductor  through  unit  area  on  its  surface  is  therefore 

IjlTT.F'M. 

This  quantity  of  energy  moves  in  the  direction  CO,  which  is  deter- 
mined by  the  intersection  of  the  electrical  and  magnetic  equipotential 
surfaces.  The  relation  of  the  direction  in  which  the  energy  is  pro- 
pagated to  the  directions  of  the  electrical  and  magnetic  forces  is 
determined  in  the  same  way  as  that  given  for  the  propagation  of 
waves  in  XCIV. 

The  electrical  force  is  here  measured  in  electromagnetic  units;  if 
we  express  it  in  electrostatic  units,  and  represent  it  by  F,  then 
F'  =  VF.  The  quantity  of  energy  which  enters  in  one  second  through 
a  unit  area  of  the  surface,  which  is  parallel  to  the  directions  both 
of  the  electrical  and  of  the  magnetic  forces,  equals  V/lir.F.M. 

We  will  now  treat  a  more  general  case.  Eepresent  the  magnetic 
force  at  a  point  in  the  region  by  M,  the  electrical  force  at  the  same 
point  by  F,  and  the  angle  between  the  two  forces  by  (M,  F).  We 


226  ELECTRICAL  OSCILLATIONS.  [CHAP.  xi. 

assume  that  the  quantity  of  energy  passing  in  one  second  through 
unit  area,  which  is  parallel  to  the  directions  of  M  and  F,  is 


The  direction  in  which  the  energy  flows  makes  angles  with  the 
coordinate  axes  whose  cosines  are  /,  m,  n.  We  have  then 

I  =  (yy-  pZ)/MFsm  (MF)  ;    m  =  (aZ-  7X)/MF  sin  (MF)  ; 
n  =  ((3X  -  aY)/MFsin  (MF), 

if  a,  p,  y,  are  the  components  of  the  magnetic  force  M,  and  X,  Y,  Z, 
the  components  of  the  electrical  force  F.  From  these  equations 
it  follows  that  la  +  m/3  +  ny  =  ();  lX  +  mY+nZ=0;  l'2  +  m2  +  n2=l. 
Hence  the  energy  flows  in  a  direction  which  is  perpendicular  to  the 
directions  both  of  the  magnetic  and  of  the  electrical  forces. 

If  we   represent  by   Ex,  E^  Ea  the   components   of  the   flow   of 
energy  in  the  directions  of  the  coordinate  axes,  we  have 

Ex  =  F/47T  .  MFsin  (MF)  .  I 
Hence  we  obtain  the  equations 
Ex=T^7r.(yY-^Z};  Ey  =  F/47T  .  (aZ  -  yX)  ;   E._  =  V\hr  .  ((BX-aY). 

The  energy  present  in  a  parallelepiped,  whose  edges  are  dx,  dy,  dz, 
increases,  in  the  time  dt,  by  an  amount  equal  to 

-  (dEJVx  +  VEJ-dy  +  VEJ-dz)dxdydzdt. 

Hence  the  increase  of  energy  A  which  a  unit  of  volume  receives 
in  the  unit  of  time,  is  A=  -(dEJ'dx  +  3EJ'dy  +  'dEt/'dz).  If  we  sub- 
stitute in  these  equations  the  values  previously  given  for  E&  E^  E,, 
we  have 

A  =  T/47T  .  [X(dyfdy  -  3/3/3*)  +  Y(dafdz  -  3y/3a;) 
o/3y)]  -  F/47T  .  [a(dZfdy  -  'dY/'dz) 
-  VZ/'dx)  +  yCdY/Vx  -  aX/3y)]. 
By  the  help  of  equations  (c)  and  (d)  of  XCIIL,  we  obtain  from  this 
A  =  KjSir  .  d(X2  +Y2  + 


On  comparing  this  expression  with  those  given  in  (g)  and  (h)  XCIIL 
for  the  electrostatic  and  electrokinetic  energies,  we  see  that  A  repre- 
sents the  total  increase  of  energy  which  the  unit  of  volume  receives 
in  the  unit  of  time.  Poynting's  Theorem  is  thus  proved,  provided 
the  dielectric  is  not  in  motion.  The  demonstration  can  easily  be 
extended  to  the  conductor  if  we  use  the  developments  of  LXXXVIII. 


SECT,  xcvi.]  POYNTING'S  THEOREM.  227 

and  remember  that  a  part  of  the  energy  absorbed  by  the  conductor 
is  transformed  into  heat. 

We  will  now  apply  Poynting's  theorem  to  a  simple  case.  According 
to  XCIV.,  we  may,  in  the  case  of  vibrations  in  a  plane,  express  the 
electrical  force  F,  there  designated  by  Y,  by  F=  b .  sin  [2ir/jP.  (t  -  z/w)], 
and  the  magnetic  force  M,  there  designated  by  y,  by 

M=  Fb/nw .  sin  [2ir/r .  (t  -  */»)]. 

During  any  complete  vibration  there  passes  through  unit  area  the 
quantity  of  energy 

F^/^TT/tw .  rsin2[2;r/r.  (t  -  x/o>)]dt  =  F^r/fywrw. 

-0 

We  may  set  F=o>  and  /*  =  !,  and  obtain  for  the  quantity  of  energy 
which  passes  in  one  second  through  a  unit  area  perpendicular  to 
the  plane  of  the  wave,  the  quantity  F62/87r. 

The  quantity  of  heat  which  a  square  centimetre  receives  in  one 
minute  from  the  light  of  the  sun  is  equal  to  about  three  gram-calories. 
This  quantity  of  heat  corresponds  to  the  energy  3 .  4,2 . 10"/60  de- 
veloped in  one  second.  If  we  now  set  F=3.  1010,  we  have  6  =  0,04. 
Since  the  unit  of  electrical  force  in  the  electrostatic  system  equals 
300  volts,  we  obtain  for  the  maximum  electrical  force  of  sunlight 
12  volts  per  centimetre.  The  maximum  magnetic  force  is  0,04,  and 
amounts  therefore  to  a  fifth  of  the  horizontal  intensity  of  the  earth's 
magnetism  in  middle  latitudes. 


The  experimental  basis  for  the  mathematical  treatment  of  electro- 
statics was  given  by  Coulomb.  Poisson  handled  a  number  of  problems 
in  electrostatics  and  gave  the  general  method  for  their  solution.  Sir 
William  Thomson  (Lord  Kelvin)  also  treated  the  same  problems  in 
part  by  a  new  and  very  ingenious  method ;  his  papers  are  specially 
recommended  to  the  student  (Reprint  of  Papers,  2nd  Ed.,  1884). 
Faraday  (1837)  developed  new  views  of  electrical  polarization  or  dis- 
placement. On  the  foundation  of  these  concepts,  Maxwell  constructed 
his  development  of  the  theory  of  electricity  (Treatise  on  Electricity 
and  Magnetism,  1873).  Helmholtz  treated  electrostatics  in  a  different 
way  and  solved  new  problems.  His  papers  may  be  found  in 
Wiedemann's  Annalen. 

The  theory  of  magnetism  advanced  parallel  with  the  theory  of 
electrostatics.  The  same  authors,  and  sometimes  even  the  same 
works,  deal  with  both  subjects. 


228  ELECTEICAL  OSCILLATIONS.  [CHAP.  xi.  SECT.  xcvi. 

Ampere  in  his  Thforie  MatMmatique  des  Phe"nomenes  Electrodynamiques, 
Paris,  1825,  discussed  the  theory  of  electrical  currents.  This  work 
forms  the  foundation  for  all  the  recent  development  of  that  theory. 
The  new  concepts  of  the  magnetic  and  inductive  actions  of  electrical 
currents,  developed  by  Faraday,  were  given  a  mathematical  form 
by  Maxwell  in  his  work :  Treatise  on  Electricity  and  Magnetism,  1873. 
We  have  followed  Maxwell's  methods  in  dealing  with  these  topics. 
On  the  other  hand  Gauss,  W.  Weber,  F.  E.  Neumann,  Kirchhoff,  and 
Lorenz,  proceed  from  Ampere's  theory. 

We  are  indebted  to  William  Thomson  and  G.  Kirchhoff  (Poggendorff's 
Annalen,  121)  for  the  theory  of  electrical  oscillations  in  conductors. 
Maxwell  and  Lorenz  showed  that  electrical  oscillations  may  also  exist 
in  the  dielectric.  By  the  investigations  of  H.  Hertz,  the  theory  of 
electrical  oscillations  has  been  given  such  extension  and  significance, 
that  no  one  can  predict  the  consequences  to  which  it  may  lead. 


CHAPTER    XII. 

REFRACTION  OF  LIGHT  IN  ISOTROPIC  AND 
TRANSPARENT  BODIES. 

SECTION  XCVII.    INTRODUCTION. 

As  the  number  of  facts  discovered  by  the  study  of  light  increases, 
and  as  additional  relations  are  found  between  light  and  other  natural 
phenomena,  it  becomes  increasingly  difficult  to  construct  a  theory 
of  light.  According  to  the  emission  theory,  which  in  its  main  features 
may  be  attributed  to  Newton,  and  which  was  handled  mathematically 
by  him,  energy  is  transferred  by  minute  bodies,  called  light  corpuscles, 
which  pass  from  the  luminous  to  the  illuminated  body.  It  was 
supposed  that  these  light  corpuscles  carried  with  them  not  only 
their  kinetic  energy  but  also  another  kind  of  energy,  to  which  the 
luminous  effects  were  due.  In  the  last  century  the  emission  theory 
was  sufficient  to  explain  the  phenomena  then  known.  But  its  develop- 
ment could  not  keep  pace  with  the  advances  of  experimental  knowledge; 
this  became  evident  early  in  this  century  in  connection  with  the 
great  discoveries  in  optics  which  were  made  by  Young,  Fresnel  and 
Malus.  In  opposition  to  this  theory,  Fresnel  developed  his  first  form 
of  the  wave  theory,  which  originated  with  Huygens ;  in  this  form  of 
the  theory,  the  light  waves  were  supposed  to  be  longitudinal. 
According  to  the  wave  theory,  the  space  between  the  luminous  and 
illuminated  bodies  is  filled  with  a  material  medium.  By  the  action 
of  the  particles  of  this  medium  on  each  other,  the  energy  emanating 
from  the  luminous  body  is  propagated  from  particle  to  particle 
through  this  medium  to  the  illuminated  body.  Hence  energy  resides 
in  the  medium  during  the  transfer  of  light  from  one  body  to  the 
other.  The  wave  theory  has  many  advantages  over  the  emission 
theory.  Notably  the  phenomena  of  interference  are  explained  by  it 
in  a  perfectly  natural  way.  It  succeeds  also  in  explaining  some  of 
the  phenomena  of  double  refraction.  But  the  explanation  of  the 

229 


230  REFRACTION  OF  LIGHT.  [CHAP.  xn. 

polarization  of  light  by  this  theory  offered  difficulties  which  could  be 
overcome  only  by  the  assumption  that  the  direction  of  the  light  vibra- 
tions are  perpendicular  to  the  direction  of  the  rays.  Since  Fresnel 
retained  the  idea  that  the  medium  in  which  the  light  vibrations 
are  propagated,  the  ether,  is  a  fluid,  he  encountered  an  obstinate 
resistance  to  his  new  form  of  the  wave  theory ;  Poisson  rightly 
maintained  that  transverse  vibrations  can  never  be  propagated  in  a 
fluid.  Although  the  wave  theory,  in  its  original  form,  was  some- 
what open  to  criticism,  and  in  many  respects  was  insufficient,  in 
that,  among  other  matters,  it  could  not  explain  the  dispersion  of 
light,  yet  it  was  a  decided  advance  on  the  emission  theory. 

Since  the  phenomena  of  optics  cannot  be  explained  on  the 
assumption  that  light  is  due  to  vibrations  in  an  elastic  medium, 
not  even  when  this  medium  is  supposed  to  be  a  solid,  we  must 
endeavour  to  explain  them  in  another  way.  Among  recent  efforts 
in  this  direction,  the  electromagnetic  theory  of  light,  developed  by 
Maxwell,  has  special  advantages.  In  Maxwell's  view,  light  is  also 
a  wave  motion,  but  it  consists  of  periodic  electrical  currents  or  dis- 
placements, which  take  the  place  of  the  etherial  vibrations  of 
Fresnel's  theory.  Maxwell  determined  on  this  assumption  the 
velocity  of  light  in  vacuo  and  in  transparent  bodies,  and  reached 
conclusions  which  agree  very  well  with  the  facts.  Polarization 
and  double  refraction  can  also  be  readily  explained  by  Maxwell's 
theory,  and  it  has  even  been  applied  successfully  to  the  study  of 
dispersion. 

Since  Fresnel's  formulas  are  of  great  importance  for  our  subsequent 
study,  we  will  develop  them  at  the  outset.  We  may  here  recall 
briefly  the  principal  laws  of  light  which  hold  for  isotropic  and 
perfectly  transparent  bodies.  The  knowledge  of  these  laws  is  neces- 
sary for  the  deduction  of  Fresnel's  formulas,  but  is  not  sufficient. 

I.  Light  is  propagated  in  any  one  medium  with  a  velocity  which 
depends  on  the  wave  length  of  the  light,  but  not  on  its  intensity. 
The  velocity  of  light  has  different  values  in  different  media. 

II.  If  a   ray   of    light   falls   on   a   plane   surface,    separating   two 
different  media,  both  refraction  and  reflection  occur   at  this    surface. 
All  three  rays — that  is,  the  incident,  the  refracted,  and  the  reflected 
— lie  in   the   same   plane,  which   is  perpendicular  to  the   refracting 
surface.      If    a  represents   the   angle   of    incidence,   ft  the    angle   of 
refraction,  and  y  the  angle  of  reflection,  we  have 

7  =  0.  and  sin  a/ sin  (3  =  N. 
The  index  of  refraction  N  is  constant  for  homogeneous  light. 


SECT,  xcvn.]  INTRODUCTION.  231 

III.  If  co   represents   the   velocity   of    light   in    the   medium   con- 
taining the  reflected  ray,  and  w'  its  velocity  in  that  containing  the 
refracted  ray,  we  have  JV=  to/to',  and  therefore  sin  a/sin  /2  =  w/w'. 

IV.  Light  can  be  considered  a  wave  motion  in  a  medium,  called 
the  ether.     It  is  a  matter  of  indifference  whether  we  here  consider 
the  bodies  themselves,  or  an  unknown  substance,  or  perhaps  changes 
in   the   electrical   or   magnetic   condition  of  the   bodies.      We  wish 
only  to  indicate  that  the  luminous  motion  may  be  expressed  by  one 
or  more  terms  of  the  form  a  cos  (2irt/T +<(>),  where  a  is  the  ampli- 
tude,  T  the  period  of  vibration,  <f>  the  phase,  and  i  the   variable  time. 
The  intensity  of  light  is  then  expressed  by  a2. 

V.  The  motion  of  the  ether  is   perpendicular  to  the  direction  of 
the  ray  of  light;  that  is,  the  vibrations  are  transverse.     Either  the 
motion  takes  place  always  in  the  same  direction,  in  which  case  the 
ray  is  rectilinearly  polarised,  or  two  or  more  simultaneous  rectilinear 
motions  may  give  the  ether  particles  a  motion  in  a  curve,  which  is 
in   general   an   ellipse.     Rays   of  light   of  this   sort  are   said    to   be 
elliptically  polarized.     If  the  path   of  the   ether  particle  is  a   circle, 
the  light  is  circularly  polarised.     Fresnel's  conception  of  natural  light 
was  that  its  vibrations  were  also   perpendicular  to  the  direction  of 
the    ray   and   rectilinear,   but    that    they    changed    their  directions 
many  times,  and   on    no   regular  plan,  in   a  very  short  interval  of 
time. 


SECTION  XCVIII.     FRESNEL'S  FORMULAS. 

Suppose  the  plane  surface  OP  (Fig.  113)  to  be  the  surface  of 
separation  of  two  transparent  isotropic  media.  We  represent  the 
velocity  of  light  in  the  medium  above  the  surface  of  separation 
OP  by  w,  and  that  in  the  medium  below  the  surface  by  w'.  If  N 
represents  the  index  of  refraction  of  the  ray  of  light  in  its  passage 
from  the  first  to  the  second  medium,  we  have  w  =  JVw'.  We  select 
the  point  0  in  the  bounding  plane  as  the  origin  of  a  system  of 
rectangular  co  ordinates,  and  draw  the  z-axis  perpendicularly  upward 
and  the  y-axis  in  the  plane  of  incidence,  that  is,  the  plane  passed 
through  the  normal  to  the  surface  at  the  point  of  incidence  and 
the  incident  ray  SO.  The  2-axis  is  therefore  perpendicular  to  the 
plane  of  incidence.  Further,  let  SO  be  the  incident,  OT  the  reflected, 
and  OB  the  refracted  ray.  We  designate  the  angle  of  incidence 
by  a,  the  angle  of  refraction  by  ft.  The  amplitude  of  the  vibrations 


232 


KEFRACTION  OF  LIGHT. 


[CHAP.  xn. 


of  the  incident  ray  may  be  called  wa,  and  that  of  the  vibrations  of 
the  refracted  and  reflected  rays  u2  and  u3  respectively.  The  planes 
of  vibration  of  these  rays  make  angles  with  the  plane  of  incidence, 
which  are  represented  by  <f>v  <£2,  <f>3  respectively.  The  components 
of  motion  along  the  coordinate  axes  are  £1?  tjv  {v  for  the  incident 
ray ;  £2,  rjy  £2,  for  the  refracted  ray ;  £3,  ?;3,  £3,  for  the  reflected  ray. 


It  is  further  advantageous  to  introduce  symbols  for  the  components 
of  motion  which  lie  in  the  plane  of  incidence  and  are  perpendicular 
to  the  direction  of  the  rays.  These  components  of  motion  for  the 
three  rays  are  designated  by  sv  sy  ss,  respectively.  We  then  obtain 
the  following  equations : 


(a) 


=  -s3sma,   773  =  s3cosa,    VV  +  Cii   tg'fcj  =  Wsv 
In  order  to  express  s^  s3,  and   £2,  £3,  in  terms  of  sl  and   ^,   we 
must  make  certain  assumptions  with  regard  to  the  behaviour  of  the 
light  at  its  passage  from  one  medium  to  another. 

I.  Fresnel  assumed  first,  that  no  light  is  lost  by  reflection  and  refraction, 
or,  the  sum  of  the  intensities  of  the  reflected  and  refracted  light  is  equal 
to  that  of  the  incident.  This  law  is  only  a  statement  of  the  law  of  the 
conservation  of  energy  in  a  particular  case,  it  being  merely  the  assertion 
that  the  kinetic  energy  of  the  incident  ray  is  equal  to  that  of  the 
reflected  and  refracted  rays.  Let  OPSS'  (Fig.  113)  be  a  cylinder, 
the  area  of  whose  base  OP  is  A,  and  whose  slant  height  SO  is  equal 
to  w,  the  velocity  of  light.  If  we  represent  the  density  of  the  vibrating 


SECT,  xcvm.]  FEESNEL'S  FOEMULAS.  233 

medium  by  p,  the  kinetic  energy  Lv  of  the  light  contained  in  the 
cylinder  considered  is  L^  =  J .  p .  o>  cos  a .  A  .  u^. 

After  the  lapse  of  a  second  this  kinetic  energy  is  divided  between 
the  reflected  and  refracted  rays.  The  kinetic  energy  L3  in  the 
reflected  ray  is  L3  =  \ .  p  .  w  cos  a .  A .  u32,  and  the  kinetic  energy  L2 
in  the  refracted  ray  is  L2  =  \p  .  o>  cos  ft .  A  .  u22,  when  p  represents 
the  density  of  the  vibrating  medium  below  the  bounding  surface. 
Therefore,  on  the  assumption  made  by  Fresnel,  we  have 
Ll  =  L2  +  L3  or  /3w(w12  —  w32)  cos  a  =  pV .  u22 .  cos  (3. 

Taking  into  account  the  relations  <a  =  N.to   and   sin  a  =  N.  sin  /?,  we 
may  give  this  equation  the  form 
{b)  p .  (u-^  —  u32) .  sin  a .  cos  a  =  p'u22 .  sin  /3 .  cos  f3. 

If  the  vibrations  of  the  ray  lie  in  the  plane  of  incidence,  ul  =  sv 
and  we  obtain 

{c)  p .  (s^  -  s32) .  sin  a .  cos  a  =  p  .  s22 .  sin  (3 .  cos  /?. 

But  if  the  vibrations  of  the  ray  are  perpendicular  to  the  plane  of 
incidence,  we  have  u^  =  £v  and  hence 

(d)  p(£i2  -  £32)  sin  a .  cos  a  =  p  .  £22 .  sin  (3 .  cos  p. 

II.  Fresnel    assumed,    secondly,    that   the   components   of  the   vibra- 
tions, which  are  parallel  to  the  bounding  surface,  and  on  either  side  of  it, 
are  equal.     If  the  vibrations  lie  in  the  plane  of  incidence,  we  have, 
on  this  assumption,  ^  +  ^3  =  ^2'  or 

(e)  (sl  +  s3)  cos  a  =  $2 .  cos  [I. 

But  if  the  vibrations  are   perpendicular  to  the   plane  of  incidence, 
we  obtain  (f)  Ci  +  Cs  =  &•     Ife  follows  from  (c)  and  (e)  that 
s2  =  sl .  2p .  sin  a .  cos  a/(p .  sin  a .  cos  (3  +  pr .  cos  a .  sin  [3), 
s3  =  s^p.  sin  a.  cos  (3  -  p. cos  a.  sin  /3)/(p.  sin  a.  cos  (3  +  p'.  cosa.  sin  f3), 
and  from  (d)  and  (f)  that 

i  =  d  •  -p  •  sin  a .  cos  a/(p .  sin  a .  cos  a  +  p' .  sin  (3 .  cos  (3), 

.  =  d .  (p-  sin  a.  cos  a-p'.  sin/3,  cos  (3) /(p.  sin  a.  cosa  +  p.  sin  (3.  cos  (3). 

III.  Since  the  relation  between  p  and  p  is  entirely  unknown,  Fresnel 
was  compelled  to  make  a  third  assumption,  so  he  assumed  that  the 
elasticity  of  the  ether  is  everywhere  the  same,  but  that  its  density  differs 
in  different  media.     On  the  other  hand,  F.  E.  Neumann  assumed  that 
the  density  of  the  ether  is  the  same  in  all  media,  but  that  its  elasticity  is 
different  in  different  media.     Fresnel's  assumption  was  natural,  because 
he  considered  the  ether  as  a  gaseous  body,  but  this  is  not  justified, 
as  has  already  been  remarked.     He  further  assumed  that  w  and  w' 


234  REFRACTION  OF  LIGHT.  [CHAP.  xn. 

can  be  expressed  in  the  same  way  as  in  the  theory  of  elasticity 
[cf.  XXXV.  (k)],  and  therefore  set  w  =  V/Vp,  w/  =  vV'//3'-  Now,  on 
Fresnel's  assumption,  p.  =  p,  and  hence,  (i)  p'/p  =  <a2/<a"2  =  N2.  By  the 
use  of  the  third  assumption  the  equations  (g)  and  (h)  can  be  given 
the  form 

s.2  =  sl  .  2  cos  a  .  sin  /3/(sin(a  +  (3)  cos  (a  - 
I  &  =  Ci  •  2  cos  «  .  sin  /3/  sin  (a 


These  formulas  are  due  to  Fresnel. 

Experiment  alone  can  decide  as  to  the  value  of  these  formulas. 
It  follows,  from  the  expression  for  s3,  that  $3  =  0,  if  a  +  ft  =  ^ir  or  if 
tga.  =  N.  Brewster  showed  that  the  light  which  is  polarized  per- 
pendicularly to  the  plane  of  incidence,  according  to  the  definition 
of  Malus,  is  not  reflected  when  tga  =  N.  This  value  of  the  angle  a 
is  called  the  angle  of  polarization.  It  is  that  at  ivhich  the  reflected  and 
refracted  rays  are  at  right  angles  to  each  other.  This  conclusion  agrees 
with  experiment,  if  we  make  the  assumption  that  the  vibrations  of 
polarized  light  are  perpendicular  to  the  plane  of  polarization.  On  the 
whole,  Fresnel's  formulas  agree  very  well  with  the  results  of  experi- 
ments on  the  intensity  of  the  reflected  light. 

In  the  notation  already  employed,  the  plane  of  vibration  of  the 
incident  ray  makes  the  angle  <^  with  the  plane  of  incidence,  the 
corresponding  angle  for  the  reflected  ray  is  (f>3.  Brewster  found  that 
tg<£3  =  tg<£:  .  cos(o,  -  /i?)/cos  (a  +  (3).  This  follows  from  Fresnel's  formulas, 

since  tg<£3  =  £3/ss  =  fi  cos(a  ~  P)/s\  cos(a  +  /*)  =  fcg^i  •  cos  (a  "  /*)/cos  (a  +  /*)• 
This  agreement  argues  for  the  correctness  of  Fresnel's  formulas. 

Fresnel  assumed  that  the  elasticity  of  the  vibrating  medium  is 
the  same  on  both  sides  of  the  refracting  surface.  We  have  seen 
that  this  assumption  is  to  some  extent  arbitrary.  On  the  other 
hand  F.  E.  Neumann  assumed  that  p  =  p.  We  obtain  on  this  latter 
assumption  from  (g)  and  (h) 

f  s9  =  2  sin  a  .  cos  a  .  Sj/sin(a  +  /3), 
(1)  J  £2  =  2  sin  a.  cos  a.  £/  sin  (a  +  /3)  cos  (a  -J3), 

[53  =  sin(a-/3).5l/sin(a  +  ^,  &  =  tg  (a-/3).  ^/tg  (a  +  /3). 

These  equations  agree  with  the  results  of  experiment,  on  the 
assumption  that  the  vibrations  occur  in  the  plane  of  polarization,  as  well 
as  those  of  Fresnel. 

We  will  now  consider  the  components  of  motion  along  the  normal 
during  the  passage  of  light  from  one  medium  to  the  other.  This 


SECT,  xcviu.]  FEESNEL'S  FORMULAS.  235 

component  above  the  bounding  surface  is  £x  +  £3,  that  below  it  is  gy 
It  follows  from  (a)  and  (g)  that 

£j  4-  £3  =  2/>'.  sin  a .  cos  a .  sin  ft .  s^(p .  sin  a .  cos  (3  +  p  .  cos  a .  sin  /3), 
£2  =  2p  .  sin  a .  cos  a .  sin  )6 .  s-^Kp .  sin  a .  cos  (3  + p.  cos  a.  sin  (3). 

From  these  equations  it  follows  that  £i  +  gs  =  i2- P/P-  If  we 
assume  with  Neumann  that  p  =  p,  we  have  £x  +  £3  =  £2  i  ^a^  ig>  ^e 
components  of  vibration  perpendicular  to  the  bounding  surface  are  equal 
above  and  below  it.  On  the  other  hand  we  obtain  from  Fresnel's 
assumption  (m)  £x  +  £3  =  N2 .  £,. 

Fresnel's  equations  agree  fully  with  experiment  only  when  the 
index  of  refraction  is  about  1,5;  it  is  in  this  case  only  that  s3  =  0 
at  the  angle  of  polarization.  In  other  cases  S3  is  a  minimum,  but 
does  not  vanish.  Several  attempts  have  been  made  to  explain  this 
fact.  Thus,  for  example,  Lorenz  assumed  that  the  passage  of  the 
light  from  one  medium  to  another  occurs  through  an  extremely  thin 
intervening  layer  of  varying  density,  and  that  therefore  the  change 
of  density  is  not  discontinuous. 


SECTION  XCIX.    THE  ELECTROMAGNETIC  THEORY  OF  LIGHT. 

In  XC1V.  it  was  proved  that  electrical  vibrations  are  propagated 
in  vacuo  and  in  a  great  number  of  insulators  with  the  velocity  of 
light.  This  fact  suggests  the  assumption  that  light  consists  of 
electrical  vibrations.  It  was  also  shown  that,  when  the  wave  is 
plane,  the  electrical  and  magnetic  forces  lie  in  the  wave  front.  In 
the  most  simple  case  the  electrical  force  F  is  perpendicular  to  the 
magnetic  force  M.  If  the  permeability  //,  of  the  medium  is  equal 
to  1,  which  is  approximately  the  case  for  most  dielectrics,  we  have 
(XCIV.),  (a)  M=NF,  where  N  is  the  index  of  refraction. 

We  will  now  develop  the  usual  expressions  for  the  reflected  and 
transmitted  light,  considering  first  the  boundary  conditions.  Since 
no  free  magnetism  is  present,  and  since  the  electrical  current 
strength  is  everywhere  finite,  the  magnetic  force  varies  continuously 
during  the  passage  from  one  medium  to  another.  "\Ve  take  the 
refracting  surface  as  the  y^-plane,  and  the  z-axis  a$  the  normal  to  it. 
Hence  if  a,  /3,  y  are  the  components  of  the  magnetic  force  on  one  side 
of  the  refracting  surface  and  a',  /?',  y  those  on  the  other  side,  we  have 

(b)  a  =  a',P  =  (3',y  =  y, 


236 


REFRACTION  OF  LIGHT. 


[CHAP.  xn. 


The  electrical  force  arises  partly  from  induction  and  partly  from 
the  free  electricity  on  the  refracting  surface.  Let  a-  denote  the 
density  on  this  surface,  X,  Y,  Z  the  components  of  the  electrical 
force  immediately  above,  and  X',  V,  Z  those  immediately  below 
this  surface.  We  then  have,  from  LXVL, 

4ir<r  =  X-KX',  Y=Y',  Z=Z. 

The  components  of  the  electrical  force  which  are  parallel  to  the 
refracting  surface  vary  continuously  during  the  passage  from  one  side 
of  the  surface  to  the  other. 

Let  SO,  OB,  and  OT  (Fig.  114)  be  the  directions  of  the  incident, 
refracted,  and  reflected  rays  respectively.  Suppose  the  direction  of 
the  magnetic  force  to  lie  in  the  plane  of  incidence,  and  therefore  the 
electrical  force  to  be  perpendicular  to  that  plane.  The  direction  of  the 


FIG.  114.  FIG.  115. 

electrical  force  may  be  found  by  the  rule  given  in  XCIV.  Since 
the  magnetic  force  varies  continuously,  we  obtain,  by  referring  to  the 
directions  indicated  in  Fig.  114. 

(d)  Ml .  cos  a  -  Ma .  cos  a  =  Jl/2 .  cos  (3, 

(e)  Ml .  sin  a  +  Ms .  sin  a  =  M2  sin  /3. 

Since  the  electrical  force  also  varies  continuously,  we  have 

(f)  Fl  +  F3  =  F2. 

But  from  equation  (a)  M^  =  Fv  M3  =  F3,  and  M2  =  NF2,  if  N  =  sin  a/ sin  ft 
is  the  ratio  between  the  velocities  in  the  first  and  second  media. 


SECT,  xcix.]    ELECTROMAGNETIC  THEORY  OF  LIGHT.  237 

Hence  equations  (e)  and  (f)  are  identical.     We  obtain  from  (f)  and  (d) 
/  F2  =  Fl .  2  cos  a .  sin  /3/sin  (a  +  (3)  • 
\JP3=-JF1sin(a-/3)/sin(a  +  /?). 

Now  let  the  electrical  force  be  parallel  to  the  plane  of  incidence.  If  the 
electrical  forces  are  positive  in  the  directions  indicated  in  Fig.  115, 
the  positive  directions  of  the  magnetic  forces  may  be  obtained  by 
the  rule  given  in  XCIV. 

The  boundary  conditions  are 
(h)  Fl .  cos  a  +  F3 .  cos  a  =  F2 .  cos  /?, 

(i)  Ml-Ms  =  M2. 

In  this  case  also  Ml  =  Fl,  M3  =  F3,  MZ  =  NF^,  so  that  (k) 
JFo  =  ^1.cosa.sin/3/sin(a  +  /3).cos(a-/3)  and  F3= -Frtg(a-p)/tg(a  +  (3). 

Equations  (g)  and  (k)  correspond  to  Fresnel's  equations  XCVIII.  (k). 
Hence,  the  electromagnetic  theory  of  light  leads  to  the  same  results  as  those 
which  are  contained  in  Fresnel's  formulas,  provided  that  the  electrical  force 
is  parallel  to  the  direction  of  the  vibrations  assumed  by  Fresnel. 

We  can  further  show,  by  the  help  of  Poynting's  theorem  (XCVL), 
that  the  energy  which  in  a  given  time  is  transported  to  the  refracting 
surface  by  the  incident  ray  is  equal  to  that  which  is  carried  away 
from  it  in  the  reflected  and  refracted  rays.  Since  in  this  case  the 
electrical  and  magnetic  forces  are  perpendicular,  the  energy  passing 
through  unit  area  equals  VMF\±Tr.  The  bounding  surface  S  receives 
the  quantity  of  energy  1/47T.  PrMlFl .  S  cos  a.  in  unit  time.  We  may 
write  similar  expressions  for  the  energies  of  the  reflected  and  trans- 
mitted rays,  and  have  the  relation 
1 ,  4vr .  VM^ .  S .  cos  a  =  l/4;r .  VM.2F^ .S.cos(B  +  l/47r .  VM3FS  .S.cos  a, 

or  (M^  -  M3FS) .  cos  a  =  M.2F2 .  cos  /3.     By  reference  to  the  relations 
between  the  electrical  and  magnetic  forces,  we  obtain 
(^2  _  ^CQS  a  =  NF* .  cos  /3. 

It  appears  from  equations  (g)  and  (k)  that  this  equation  is  satisfied 
if  the  electrical  force  is  either  perpendicular  or  parallel  to  the  plane 
of  incidence. 


SECTION  C.    EQUATIONS  OF  THE  ELECTROMAGNETIC  THEORY  OF 
LIGHT. 

If  we  at  first  consider  only  bodies  in  which  there  is  no  absorption 
of  light,  and  in  which  the  velocity  of  light  is  the  same  in  all  directions, 


238 


EEFRACTION  OF   LIGHT. 


[CHAP.  xii. 


we  have,  from  XCIII.  (e),  the  differential  equations  of  the  electrical 
force, 

t  l/o>2 .  WXfdt*  =  VIST,    1/w2 . 32F/9£2  =  V-Y, 
\  l/w2 .  wzrdP  =  V2Z,    -dXfdx  +  ^Yj-dy  +  VZ/Vz  =  0. 
The  boundary  conditions  are  obtained  by  remarking  that  the  com- 
ponents of  the  electrical  and  magnetic  forces  parallel  to  the  bounding 
surface  are  equal  on  both  sides  of  it.      Therefore,  if  the   z-axis   is 
perpendicular  to  the  refracting  surface,  we  have 
<b)  Y=Y',  Z=Z';   /?  =  /?',  y  =  y'. 

The  last  two  conditions  (b)  may  be  put  in  the  form  [XCIII.  (e)], 
/  -dXj-dz  -  -dZfdx  =  VX'fdz  - 


FIG.  116. 

Let  us  suppose  that  a  plane  wave  moves  in  a  direction  which 
makes  angles  with  the  axes  whose  cosines  are  I,  m,  n.  Let  the  elec- 
trical force  /  at  the  origin  be  expressed  by  f=F.cos(2irt/T).  The 
electrical  force  at  a  point  whose  coordinates  are  x,  y,  z  is  then 

/=  F.  cos[27r/r(^  -  (Ix  +  my  +  n«)/o»)]. 

If  the  direction  of  the  electrical  force  makes  angles  with  the  axes 
whose  cosines  are  A.,  p.,  v,  we  have  X=\f,  Y=pf,  Z=vf.  These 
expressions  satisfy  the  equations  (b) ;  that  they  may  also  satisfy  the 
last  of  equations  (a)  we  must  have  l\  +  mp.  +  nv  =  0,  that  is,  the  direc- 
tion of  the  electrical  force  is  perpendicular  to  the  direction  of  propagation. 

Let  OP  (Fig.  116)  represent  the  refracting  surface,  and  SO,  OB 
and  OT  the  incident,  refracted,  and  reflected  rays  respectively.  The 


SECT,  c.]     EQUATIONS  OF  ELECTROMAGNETIC  THEORY.  239 

system  of  coordinates  is  drawn  in  the  same  way  as  in  XCVIII.  For 
the  incident  wave,  in  which  the  direction  of  the  electrical  force  lies  in 
the  plane  of  incidence,  we  can  set 

I  =  -  cos  a,    m  =  sin  a,    n  =  0  ; 
X=      sin  a,     ju,  =  cosa,     v  =  0. 
We  have,  for  the  reflected  rays, 

/  =      cos  a,    m  =  sin  a,    n  =  0  ; 
A=-sina,     fji  =  cos  a,     i/  =  0, 
and  for  the  refracted  rays, 

I  =  -  cos  (3,    m  =  sin  /?,    n  =  0  ; 
A  =     sin  /?,    fi  =  cos  f3,    v  =  0. 

If  the  direction  of  the  electrical  force  is  perpendicular  to  the  plane  of  incidence 
we  have  A  =  0,  M  =  0,  v  =  1 . 

If  Fv  F2,  F3  are  the  electrical  forces  in  the  incident,  refracted,  and 
reflected  rays,  when  they  lie  in  the  plane  of  incidence,  and  if  Zv  Zz,  Zz 
are  the  electrical  forces  in  the  same  rays  when  they  are  perpendicular 
to  the  plane  of  incidence,  we  have 

/  X=Fl.sina.cosTl,    Y=  F1 .  cos  a.  cos  Fv   Z  =  Zl.cosF'l, 
\  F!  =  [27T/T  .(t-(-xcosa  +  ysin  a)/a>)]. 

We  obtain  for  the  refracted  ray,  if  w'  is  the  velocity  of  light  in  the 
second  medium, 

=  ^2.sin/3.cosF2,    F=  F2 .  cos /3 .  cos  T2,    Z=Z.cosT 


(d)       l  rr,  =  [27r/r.(^-(-a;cos^  +  ysm/3)/o>')]. 

For  the  reflected  ray  we  have 

r=^3.cosr3, 


1  ^ j  =  [2-ir/T .  (t  -  (x  cosa  +  y  sin  a)/o>)]. 

To  simplify  the  calculation  we  replace  the  trigonometrical  form 
(f)  cos£(^-  (  -  z  cos  a  +  y  sin  a)/o>) 

by  the  expression  (g)  ^(M-scosa+ysineO/a^  where  i  =  \/-  1  and  k=%Tr/T' 
In  the  final  result  we  use  only  the  real  part  of  (g),  namely  (f).  Both 
expressions  satisfy  the  same  differential  equation,  and  therefore  in 
calculations  one  of  them  may  be  replaced  by  the  other. 

If  the  refraction  occurs   at  a   plane  surface,  we  may  replace  the 
expressions  (c),  (d)  and  (e)  by  the  following : 

X=Fl.$ina. 
(b) 

*,< 


240  EEFEACTION  OF  LIGHT.  [CHAP.  xn. 

,  X=Fa.s\n(3.eki(t-<--*coal3+ys 
(i) 


{X=  —  F3 .  sin  a .  «**{*-(* 
Y=F3.cosa.eki(t-<-XC(iS 
£ —  £      gki(t-(xcosa+ysin 
3  * 


a+y  sin 


These  equations  express  the  components  of  the  electrical   force   for 
the  incident,  refracted,  and  reflected  rays. 

In  order  to  satisfy  the  conditions  (b)  and  (c)  it  is  necessary  that 
sin  a/a>  =  sin  /3/w'.  Since  the  velocities  of  propagation  w  and  w'  are 
constant,  we  can  set  N=  sin  a/sin  (3,  where  N  is  the  index  of  refraction. 

From  equations  (b)  we  have 

(1)  (Fl  +  F3)cosa  =  F.2.cos(3;  Z1  +  Zs  =  Zy 

From  (c)  we  obtain 

(m)  (Fl-F3)siu/3  =  F2.sina-  (Zl  -  Z3)cosa.  sin/?  =  ^2.  sin  a.  cos  (3. 
From  (1)  and  (m)  we  obtain  Fresnel's  equations  [XCVIII.  (k)]  for 
the  reflected  and  refracted  waves.  The  problem  is  solved  when  (3 
is  not  imaginary.  /3  becomes  imaginary  when  sin  /?>  1,  and  therefore 
when  sin  a  >  N.  In  this  case  we  must  use  the  complete  expressions 
(i)  and  (k). 

If  the  electrical  force  is  perpendicular  to  the  plane  of  incidence,  the 
reflected  wave  is  determined  by  the  real  part  of  the  expression 

(n)  -  Zl .  Sin(a  -  /3)/sin(a  +  /?)  .  ei-i(<-(xcosa+ysina)/W)> 

We  get  this  expression  by  the  use  of  the  last  of  equations  XCVIII. 
(k).     But  since  cos(3  =  Jl  -  siri2a/N'\  and  therefore 
(o)  Ni .  cos  /3  =  Jsirfa-N12 

we  have 

-  sin  (a  -  (3) /sin  (a+(3)  =  (cos  a  +  i*Jsin?a  -  N2)/(cos  a  -  K/sin-a  -  ,/V2). 
If  we  now  set 

(p)  cos  a  =  C .  cos  Jy,    J  sin2a  -  N'2  =  C .  sin  |y, 

we  obtain 

(q)  tg  |  y  =  N/  sin'2a  -  N2/cos  a,    C?=l-N2; 

(r)  -sin(a-/3)/sin(a  +  £)  =  ^. 

Hence  the  real  part  of  the  expression  (n)  is 
(s)  Z^ .  cos  [k(t  -(xcosa  +  y  sin  a)/w)  +  y]. 

In  this  case  the  reflection  is  total,  since  the  component  Zl  appears 
in  the  expression  for  the  incident  as  well  as  in  that  for  the  reflected 


SECT,  c.]     EQUATIONS  OF  ELECTROMAGNETIC  THEORY.  241 

wave.  But  while,  in  the  case  of  ordinary  reflection,  no  difference 
of  phase  arises  between  the  two  waves,  we  have  in  this  case  a  difference 
of  phase  7,  which  may  be  determined  from  (q). 

If  the  electrical  forces  far  the  incident  wave  are  parallel  to  the  plane  of 
incidence,  we  determine  the  real  part  of  the  expression 

(t)  -  Fl  .  tg(a  -  /3)/tg(a  +  /2)  .  ett(«-Cscosa+3/sina)/w). 

Using  equation  (o),  we  have 
tg(a  -  /2)/tg(a  +  /3)  =  (TV2.  cosa 
If  we  set 
(u)  JV~2.cosa 

so   that   (v)   tg  £8  =  Vsin'2a  -  N2/N2  cos  a, 

we  obtain  (x)  tg(a-  /3)/tg(a  +  p)  =  eiS.     Hence  the   real   part  of  the 

expression  (t)  is  (y)    -F1cos[k(t-  (zcosa  +  ysina)/w)  +  S]. 

The  reflection  is  therefore  total.  To  determine  the  difference  of  phase 
8  between  the  reflected  and  incident  waves,  we  may  use  equation  (v). 
We  obtain  from  (q)  and  (v)  tg  |(8  -  y  )  =  V  sin2a  -  N2/  sin  a  tg  a.  If  a0 
is  the  limiting  angle  of  total  reflection  or  critical  angle,  we  have  N  =  sin  <x0, 
and  hence  (z)  tg|(8-  y)  =  N/sin(a  +  a0).  sin(a-a0)/smatga. 

Since  8  and  y  are  not  equal,  a  linearly  polarized  ray  of  light,  in 
which  the  vibrations  make  any  angle  with  the  plane  of  incidence, 
is  elliptically  polarized  after  reflection. 

If  the  electrical  force  is  perpendicular  to  the  plane  of  incidence,  the 
transmitted  light  is  determined  by  the  real  part  of  the  expression 
(a)  Z±.  2cosa 

Referring  to  (p),  we  have 

2  cos  a  sin  /?/  sin  (a  +  /?)  =  2  cos  a/  C  . 


and  (  -  x  cos  /?  +  y  sin  fB)/<o'  =  (ix-Jsin^a  -N2  +  y  sin  a)/o>. 

Hence  the  real  part  of  (a)  is 

(/3)  2  cos  a/(7  .  etov/sini!a-^/w  .  zi  .  cos  [k(t  -  y  sin  o/u)  +  ly]. 

Since   C2  =  1  -  JV2,  we  obtain  4  cos2a/(l  -  JV2)  .  Z*  .  (**#/***>•-  W\t  for 

the  intensity  of  the  transmitted  light,  where  X  denotes  the  wave  length. 

The  expression  shows  that,  in  this  case  also,  a  motion  exists  which 
corresponds  to  the  refracted  ray  in  the  case  of  ordinary  reflection; 
it  is,  however,  appreciable  only  within  a  very  small  distance  from 
the  refracting  surface. 

Similar  results  are  obtained  in  the  investigation  of  the  refracted 
ray  if  the  electrical  force  of  the  incident  light  is  parallel  to  the  plane  of 

Q 


242  REFRACTION   OF   LIGHT.  [CHAP.  xn. 

Remark:  In  order  to  obtain  the  real  part  of  an  expression  of  the 
form  (n)  we  may  use  the  following  method.  The  expression  (n)  is 
thrown  into  the  form 

(A  +  Si) .  0*  =  (A  +  Bi)(cosV  +  i  sin  ¥). 

The   real   part   of  this   is   R--=A .  cos^-  B.  sin^.     Now   if  we   set 
A  =  C.cosy,    B  =  C.siuy,    we   have    (y)  li=  C.  cos  fF  +  y),  where   0 
and  y  are  determined  by 
f 

|tgy= 
The  expression  (n)  then  takes  the  form 

Z^ .  (cos  a  +  iv/sin2a  -  JV2)/(cos  a  -  ijsin'2a  -  N'2). 
In  the  case  considered,  therefore, 

A+Bi  =  Zl.  (cos  a  +  i\/sin2a  _  JV2)/(cos  a  -  z'v/sin-a  -  N'2). 
If  +  i  and  —  i  are  interchanged,  AVC  obtain 

A  -  Bi  =  ^(cosa  -  iv/sin-a  —  N'2)/(cos  a  +  iv/sin2a- 
By  multiplication  of  the  two  expressions  we  obtain  C~ 
From  (8)  it  further  follows  that 

tg  y  =  2  cos  o\/sin2a  —  N2/(co& 2a  -  sin 2a  +  N'2) . 
This  equation  may  be  also  obtained  from  (q). 


SECTION  CI.    REFRACTION  IN  A  PLATE. 

We  will  consider  the  case  of  a  plane  wave  of  light  falling  on  a 
plane  parallel  glass  plate,  whose  thickness  is  a  and  whose  index  of 
refraction  is  N.  We  can  determine  the  intensities  of  the  reflected 
and  transmitted  light  in  the  following  way.  We  choose  one  surface 
A  of  the  plate  as  the  y^-plane,  and  draw  the  positive  a-axis  outward 
from  this  surface.  Let  a  represent  the  angle  of  incidence,  /5  the  angle 
of  refraction,  w  and  a/  the  velocities  of  the  light  inside  and  outside 
the  plate.  A  part  of  the  refracted  ray  is  reflected  toward  the  surface 
A  at  a  point  E  of  the  surface  B.  This  part  is  again  divided  at  the 
surface  A,  a  part  of  it  passing  through  that  surface  in  the  direction 
FG,  while  the  other  part  is  again  reflected  toward  B.  The  light 
is  thus  reflected  within  the  plate  repeatedly.  Since  the  plate  is 
bounded  on  both  sides  by  the  same  medium,  the  angle  of  exit  is 
equal  to  the  angle  of  incidence  a.  Now  plane  waves  which  move 
in  the  same  direction  may  be  compounded  into  a  single  plane  wave. 


SECT.  CI.] 


REFRACTION   IN   A  PLATE. 


243 


Besides  the  incident  wave  we  have  to  consider  four  others,  namely, 
the  wave  reflected  from  A,  the  wave  passing  through  B,  and  two 
waves  in  the  plate  itself. 

I.  We  will  first  consider  the  case  in  which  the  electrical  force  of  the, 
incident  wave  is  perpendicular  to  the  plane  of  incidence.  The  component 
of  the  electrical  force  outside  A  is  expressed  as  in  the  former  para- 
graph by 


FIG.  117. 

Similar  expressions  hold  for  the  component  Z'  of  the  electrical 
force  in  the  plate,  which  are  obtained  by  replacing  a  by  /?,  u>  by  to', 
and  introducing  the  new  constants  Z^  and  Z±.  Thus  we  obtain 

Z'  =  Z2  .  gW(«-(-*«M|8+y8ii»j8)/w')  +  Z^  .  gKft-fccos/S+if  sinjSyW). 

We  have  for  the  component  Z"  of  the  transmitted  ray 

Z"  =  Z-3  .  el'l(l  -  (  -  x  cos  °-+y  sin  a>/w). 

The  boundary  condition  Z=Z'  when  x  =  Q,  gives  (a)  Zl  +  Z3  =  Z2  +  Z4. 
Similarly  Z'  =  Z"  when  x=  -a,  or 

Z2.e~  kia  •  cos  jS/w  +  Z±  .  ekia  •  cos  /3/w/  =  Z6.e~  kia  •  cos  a/w. 
Now  if  we  set  ka  .  cos  /5/w'  =  ?/,  ka  .  cos  a/w  =  v,  we  can  write  the  last 


condition  in  the  form  (b) 


+  Z±  .  eui  =  Z5  .  e~vi. 


We  have,  further,  when  x  =  0,  ?>Z'/'dx  =  'dZ/'dx,  and  similarly,  when 
x  =  -a,  'dZ'/'dx  =  'dZ"  '  fdx.     These  equations  of  condition  give 
(c)  (Zl  -  Z3)  cos  a  .  sin  /?  =  (Z2  -  Z4)  sin  a  .  cos  /3 

and  (d)  (Z2  .  e~ui  -  Z±  .  eui)  sin  a  .  cos  ft  =  Z5  .  e'".  cos  a  .  sin  /?.     It  follows 
from  (b)  and  (d)  that 

ZJZ.2  =  e-2i".  sin  (a  -  /3)/sin(a  +  /5), 


244  EEFEACTION  OF  LIGHT.  [CHAP.  xn. 


or  if  sin(a-/3)/sin(a  +  /3)  =  €,  ZJZ^  =  t  .  e~*ui.  From  (a)  and  (c)  we 
have  also  Z3/Zl  =  (-  eZ2  +  Z4)/(Z2-  fZj,  and  therefore 

(e)  Z3/Z,  =  -  («*  -  O/(V«  •  *ui  -  e  •  O- 

If  N  is  greater  than  1,  u  is  always  real,  and  we  may  set 
ZJZi  =  -  2w  .  sin  «/[(!  -  €2)cos  M  +  (1  +  e2)f  .  sin  «]. 

Designating  the  intensity  of  the  reflected  light  by  C2,  we  obtain, 
by  the  method  indicated  at  the  end  of  C., 

C*  =  Zl*.  4e2.  sin2«/[(l  -  e2)2  +  4e2.  sin2tt]. 

But  because  k  =  2-!r/T  =  -2ino/\  and  u  =  2-nr  .  N.  cos  /3  .  a/A,  it  follows  that 
/f)  (CZ  =  Z*.  4e2.  sin2(27rJV.  Cos  /3  .  a/A)/[(l  -  £2)2 

\  +  4«2.  sin2(27rAr.  cos  0  .  a/A)]. 

Hence  no  light  is  reflected  if  2-n-N  .  cos  /3  .  a/A  =pr,  where  p  is  a 
whole  number.  This  result  is  of  special  importance  in  the  study 
of  Newton's  rings. 

On  the  other  hand,  if  N<1  and  at  the  same  time  sma>N,  [3  will 
be  imaginary.  In  this  case  we  can  no  longer  use  equation  (f).  We 
then  have  [C.  (o)]  Ni  .  cos  (3  =  >/sin2a  -  N'2,  and  hence 

ui  =  2ira/X.  .  *Jsiri2a-N2. 
If  we  set  m  =  ui  and  e=  -e+{7  [C.  (r)],  equation  (e)  takes  the  form 

Z3/Z,  -  (e--  e-™)/(e>"-v  -  «-*+*); 

Designating  by  C2  the  intensity  of  the  reflected  light,  we  obtain  in 
the  same  wajr  as  before, 
(g)  C2  =  Z*.  1/[1  +  4  sin2y/(em  -  e-™)2], 

where  tg  |y  =  >/sin2a  -  7V2/cos  a  and  m  =  27ra/A.  .  Vsin2a  -  JV2. 

The  relations  which  we  have  here  considered  occur  in  the  case 
of  two  transparent  bodies  which  are  separated  by  a  layer  of  air.  If 
the  thickness  of  the  layer  of  air  is  very  much  greater  than  the 
wave  length  of  the  light,  total  reflection  will  occur.  This  is  in  accord 
with  equation  (g),  which  in  this  case  gives  C2  =  Zl2.  On  the  other 
hand,  if  a  is  small  in  comparison  with  the  wave  length,  all  the  light 
passes  through  the  layer  of  air.  In  consequence  of  this  a  black  spot 
is  seen  if  the  hypotenuse  of  a  right-angled  glass  prism  is  placed  on 
the  surface  of  a  convex  lens  of  long  focus.  If  the  angle  of  incidence 
a  in  the  glass  prism  is  less  than  the  critical  angle,  a  dark  spot  appears 
surrounded  by  coloured  rings  ;  but  if  the  angle  of  incidence  is  greater 
than  the  critical  angle,  the  rings  disappear  while  the  spot  remains. 


SECT,  ci.]  KEFRACTION  IN  A  PLATE.  245 

The  spot  is  larger  for  red  than  for  blue  light.  This  result  is  contained 
also  in  the  expression  for  the  intensity  of  the  reflected  light.  The 
transmitted  light  is  complementary  to  the  reflected  light. 

II.  If  the  direction  of  the  electrical  force  of  the  incident  light  is  parallel 
to  the  plane  of  incidence,  the  disturbance  outside  the  surface  A  is 
determined  by 

X=Fl  .  Sin  a  .  ei-i(«-(-*«>sa+y8ina)/W) 

-  Fs  .  Sin  a  .  gK(t-(*coea+y  sinetf/w), 
Y=  Fl  .  COS  a  .  eki(t  -  (  -  *  cos  a+y  sin  o)/w) 

4  7^3  .  cos  a  .  eki(l  ~  (x  cos  a+y  sln  a)M. 
The  disturbance  inside  the  plate  is  given  by 

X'  =  FZ.  Sin  ft  .  ««(«-(-*  cos  p+y  sin  0)/w) 

-  F4  .  Sin  ft  .  6«('-(*  cos/3+y  sin/3)/W)) 
F  =  1*2  .  COS  0  .  €**(«  -<-*  c°s/3+y  sin/3)/u>') 

+  /^  .  COS  /3  .  ««(*-(*  cosp+y  sin  £)/&>')  . 

and  outside  the  surface  .5  by 

X"  =  FK  .  Sin  a  .  gti{«-(-arcoso+y  sinoyw^ 
Y"  =  F-0  .  COS  a  .  ^('-(-^cosa+J/sinaVw). 

We   must   now   determine   the  constants   ^  J^,  jF4,  Fb.      When 
x  =  0  the  boundary  conditions  give  F=  F,  or 
(h)  (^  +  .F3)  cos  a  =  (F2  +  F4)  cos  /8. 

Similarly,  when  z=  -a, 

,F2  .  cos  ft  .  e~kia  •  co80/w  +  ^4  .  cos  ft  .  ekia  •  ™spl"'  =  F5.cosa.  e~kia  •  cos  a/w. 
Using  the  same  notation  as  before,  we  have 
(i)  F2  .  cos  ft  .  e~ui  +  F±  .  cos  ft  .  eui  =  f5  .  cos  a  .  e-"'. 

We  have,  further,  when  #  =  0, 


or  (k)  (^  -  F3)  sin  ft  =  (F2-  Fj  sin  a. 

The  same  condition  holds  when  x  =  -  a,  or 

(1)  f^  .  sin  a  .  e-"1  -  .P4  .  sin  a  .  e'"  =  ^  .  sin  ft  .  r«. 

We  obtain  from  equations  (i)  and  (1)  FJF2  =  e~-ui  .  tg(o-j8)/tg(a  +  )8). 

But  if  we  set  tg(a-/3)/tg(a  +  /3)  =  e',   we   will  have   FJF2  =  ^.e~M. 

It  follows  from  (h)  and  (k)  that  ^/^  =  (  -  e'  +  e'  .  e-a"«)/(l  -£'2.e-2'»), 

or  (m)  ^3^=  -(e!(i-e-ui)/(  1  /«'•«"*  -e'.e~Mi)-      We   thus   obtain   the 

intensity  Z>2  of  the  reflected  light  in  the  same  way  as  we  obtained 

the  expression  (f)  from  (e), 


™     „„  .. 

1   '  (  I  -  e'  2)2  +  4e  a  .  sina(2^V  cos  /3a/  A)' 


246  REFRACTION  OF  LIGHT.  [CHAP.  xn. 

If  sin  a  >  N  and  if  (3  is  therefore  imaginary,  we  obtain  the  in- 
tensity of  the  reflected  light  in  the  following  way  :  We  have 
€  =  tg(a-  /?)/tg(a  +  f3)  =  e8i,  if,  as  in  C.  (u),  we  set 

JV2  .  cos  a  =  Z)  .  cos  JS,   >/sin2a  -N'2  =  D.  sin  |3. 
If  we  further  set  ui  =  m,  it  follows  that 


The  intensity  D2  of  the  reflected  light  is  then 
(o)  Z>2  =  FS  .  l/(  1  +  4  sin28/(«w  -  e~mf-), 

in  which  expression 

tg  |<5  =  x/sin2u  -  N'2/N2  cos  a,    m  =  '2-n-a/X  . 

The  expressions  (n)  and  (o)  for  the  intensity  of  the  reflected  light 
when  the  direction  of  the  electrical  force  is  parallel  to  the  plane  of 
incidence,  lead  to  essentially  the  same  results  as  equations  (f)  and 
(g),  which  hold  when  the  direction  of  the  electrical  force  is  perpen- 
dicular to  the  plane  of  incidence.  We  only  remark  that,  from 
equation  (n),  D-  vanishes  if  e'  =  0  or  (a  +  /2)  =  |TT.  In  this  case  the 
angle  of  incidence  is  equal  to  the  angle  of  polarization. 


SECTION  CII.    DOUBLE  REFRACTION. 

Up  to  this  point  we  have  assumed  that  the  value  of  the  dielectric 
constant  K  is  independent  of  the  direction  of  the  electrical  force. 
Boltzmann,  however,  has  shown  that  the  dielectric  constant  of  crystals 
has  different  values  in  different  directions,  and  depends  on  the 
direction  of  the  electrical  force.  Let  Kv  K2,  JT3  represent  the  value 
of  the  dielectric  constant  in  three  perpendicular  directions  which 
are  those  of  the  coordinates  a;,  y,  z.  Then  in  place  of  equations 
XCIII.  (a),  we  use 

(a)          u  =  Kl/4v.'dXI'dt,   v  =  K2/4ir .'dY/dt,  w  =  K3/4ir . 'dZ/'di. 
Equations  XCIII.  (d)  and  (c)  become 

KJ  V.  'dX/ot  =  3y/3y  -  3/2/3,2,    A'2/  V.  3  Yfdt  =  3a/3^  - 

K3/  V.  'dZj'dt  =  3£/3*  -  3a/3y, 

and  if  we  set  the  magnetic  permeability  /*  =  1,  we  have 
-  I/  V.  3a/3*  -  *dZfdy  -  'dY/'dz, 
-\\V.  3/?/3*  =  3JST/32  - 
-\\V.  3y/3*  =  3F/3.C  - 


SECT,  cn.]  DOUBLE  REFRACTION.  247 

Further,  if  we  set 


we  obtain 

f  I/a2  .  - 
(b)  J  1/42.  327/9*2 


We  will  consider  a  plane  wave  moving  through  a  body  to  which  these 
equations  apply.  Its  direction  of  propagation  is  determined  by  the 
angle  whose  cosines  are  I,  m,  n;  the  direction  of  the  electrical  force 
/  is  determined  by  the  angle  whose  cosines  are  A,  /*,  v.  We  then 
have 

(c)  X=\f,    Y=rf,   Z=vf,  f=F.cos[2ir/T.(t-(lx  +  my  +  nz)/u>)']. 

F  is  constant,  and  the  velocity  of  propagation  w  depends  only  on 
the  direction  in  which  the  wave  is  propagated.  It  follows  from  (c) 
that  V2A"=  -  47T2A//T2<o2,  and  if  cos  8  =  /A  +  m/*  +  nv,  we  obtain 

(d)  /=  L>7r/ro)  .  F  .  cos  8  .  sin  [2ir/r  .  (t  -  (Ix  +  my  +  ««)/«)]. 

We  obtain  from  the  first  of  equations  (b),  (d')  A  -  I.  cos  8  —  w2  .  A/a2. 
This  equation  and  the  two  similar  to  it  take  the  forms 

(e)  (a2  -  <o2)A  =  a2/  .  cos  8,  (b2  —  or)/*  =  b2m  .  cos  8,  (c2  —  «2)v  =  c2n  .  cos  8. 
We    use    these    equations    to   obtain    the    physical    meaning  of  the 
magnitudes  a,  b,  c.     If  w  =  a  we  have  either  1  =  0  or  cos  8  =  0.     In 
the  latter  case  //,  =  v  =  0  and  A=  ±1  and  therefore  also  1  =  0.     Hence 
a  plane  wave  parallel  to  the  x-axis  is  propagated  with  the  velocity 
a  when  the  electrical  force  is  parallel  to  the  same  axis.     The  meaning 
of  the  magnitudes  b  and  c  is  obtained  in  a  similar  way.     By  the 
optical  axes  of  elasticity  we    mean    the    three   directions   in   a   body 
which  have  the  property  that  a  plane  wave,  in  which  the  electrical 
force  or  the  direction  of  vibration   is   parallel  to  one  of  the  axes, 
for  example  a,   is   propagated  with  the  velocity  a  in  all  directions 
perpendicular  to  the  axis  a. 

We  can  find  the  velocity  of  propagation  and  the  direction  of  the 
force  from  equations  (e)  and  (d),  in  connection  with  the  relation 
A2  +  p.2  +  v2  =  1.  If  equations  (e)  are  multiplied  by  /,  m,  n,  respectively, 
and  added,  it  follows  from  (d)  that 

a2Z2/(a2  -  w2)  +  62m2/(Z>2  -  w2)  +  c%2/(c2  -  w2)  =  1. 

For  brevity  we  will  write  for  this  equation  2a2£2/(a2  -  w2)  =  1  .  But 
because  a2  =  a2  -  w2  +  w2,  we  also  have 

-  w2)  =  2/2  +  2a>2/2/(a2  -  a>2)  =  1. 


248 


REFRACTION   OF   LIGHT. 


[CHAP.  xn. 


Since  2Z2  =  1,  it  follows  that 

(f )  I2 /(a2  —  (o2)  +  m2/(i2  -  o>2)  +  %2/(c2  -  or)  =  0. 

This  equation  is  of  the  fourth  degree  in  w.     Since  two  of  its  roots 
are  numerically  equal   to   the   other  two   but   of  opposite   sign,  the 
electrical  wave  has  two  velocities  of  propagation,  u^  and  0)3. 
We  may  give  equation  (f)  the  form 

(g)  <»*-(l2(b2  +  c2)  +  m2(a2  +  c2)  +  n2(a2  +  b2))u2  +  I2b2c2  +  mW  +  n2a2b2  =  0. 
If  1  =  0,  that  is,  if  the  plane  wave  is  parallel  to  the  z-axis,  we  have 

The  roots  of  this  equation  are  w1  =  a,    o>2  =  *Jm2c2  +  n2b2. 

This  result  can  be  represented  by  drawing  lines  in  the  yz-plane 
from  the  point  0  (Fig.  118),  which  are 
proportional  to  the  velocities  of  propaga- 
tion. The  ends  of  these  lines  then  lie 
on  two  curves,  one  of  which  is  given  by 
w1  =  a,  and  is  a  circle;  the  other  is  given 
by  w2,  and  is  an  oval.  If  a  >  b  >  c,  the 
minor  semi-axis  c  of  the  curve  given  by 
o>2  lies  in  the  y-axis,  and  its  major  semi- 
axis  b  in  the  £-axis.  The  relations  of  the 
"  plane  waves  which  are  parallel  to  the  y- 
and  £-axes  respectively,  are  given  in  Figs. 
119  and  120.  The  relation  in  the  axz-plane  is  especially  peculiar 
(Fig.  119).  In  that  case,  we  have,  for  m  =  0,  <al  =  b,  o>2  =  v7/-c-  +  7i2a2 


c 

FIG.  118. 


O  tax. 

FIG.  119. 

and    l2  +  n?=l.      The    direction    of    propagation   in   which    the   two 
velocities  Wj  and  o>2  are  equal  is  given  by 

/=  ±J(a2-b2)/(a2-c2) ;   n  =--  ±  J(b2  -  c'2)/(a2  -  c2). 


SECT,  cm.]  VELOCITIES  OF  PROPAGATION.  249 


SECTION  CIII.    DISCUSSION  OF  THE  VELOCITIES  or  PROPAGATION. 

If  Wj  and  o>2  represent   the  two  velocities  of  propagation  of  the 
same  plane  wave,  we  have  [CIL  (g)] 


(a)  2 

{  Wl2  .  w22  =  We2  +  m2a?c2  +  n2a2b2, 

and       (Wl2  -  o>22)2  =  (I2(b2  +  c2)  +  m2(a2  +  c2)  +  n2(a2  +  b2))2 

-  4(/W  +  m2a2c2  +  n-a2b2). 

If  we  multiply  the   last  term  on  the  right  side  of  this  equation  by 
I2  +  m2  +  ri2  =  1  we  obtain 


1  -  c2)2  +  m4(c2  -  a2)2  +  7i4(a2  -  b2)2  +  2m2n2(a2  -  b2)(a2  -  c2) 

'  -  c2)(b2  -  a2)  +  2l2m2(c2  -  a2)(c2  -  b2). 
If  a  >  b  >  c,  it  follows  that 
(M^  -  o>22)2  =  l*(b2  -  c2)2  +  m*(a2  -  c2)2  +  n\a2  -  b2)2  +  2m2n2(a2  -  b2)(a2  -  c2) 

-  2l2n2(b2  -  c2)(a2  -  b2)  +  2l2m2(a2  -  c2)(b2  -  c2), 
or 

(Wl2  -  o>22)2  =  (I2(b2  -  c2)  +  m2(a2  -  c2)  +  n2(a2  -  b2))2 

(b)^  -^v-^xs2-^2), 


Hence  the  two  velocities  ^  and  w2  are  equal  for  certain  directions 

of  the  wave  normals.     This  equality  exists  when  m  =  0  and  either 

Ijb2  -  c2  +  nja?^b'2  =  0,  or  ijb2^2  -  W«2  -  6s  =  0. 


These  conditions  are  satisfied  by  m  =  Q  and  Z/»=  ±J(a'2-b-)/(b2  -  c2). 
These  equations  represent  four  directions,  which  are  parallel  to  the 
rc^-plane  and  perpendicular  to  the  axis  of  mean  elasticity  b.     If  we 
represent  the  cosines  of  the  angles  made  by  these  directions   with 
the  coordinate  axes  by  10,  m0,  w0,  we  have 
(c)         m0  =  0,    I0=±j(a2-b2)/(a2-c2),    n0  =  ±  J(b2  -  c2)/(a2  -I2). 
We  call  the  directions  in  the  crystal,  defined  by  equations  (c),  the 
optic  axes.     There  are  two  such  axes,  since  each  of  these  equations 
represents  two  opposite  directions. 

If  Oa  and  Oc  (Fig.  121)  represent  the  axes  of  greatest  and  least 
elasticity  a  and  c,  and  if  OAl  is  one  of  the  directions  in  which  wx 
and  w2  are  equal,  they  are  equal  not  only  in  the  opposite  direction 
OB  but  also  in  the  directions  OA2  and  OB2,  if  OA2  makes  the  same 
angle  with  Oa  as  that  made  by  OAr 


250 


REFRACTION  OF   LIGHT. 


We  will  now  express  the  velocity  of  propagation  in  any  direction  in 
terms  of  the  angles  made  by  this  direction  with  the  optic  axes  OAl 
and  OA 2.  The  cosines  of  the  angles  which  the  direction  of  propagation 
of  the  plane  wave  makes  with  the  axes  are  Z,  m,  ?<.  We  then  have 

f  cos  El  =  / .  V(a2  -  6*)/(tt2  -  c2)  +  n  .  V(&2  -  c2)/(a2  -  c2), 


(c') 


From  this  it  follows  that 

(21  =  (cos  El  +  cos  Ez) .  >/(a2  -  c2)/(a2  - 
1  2«  =  (cos  El  -  cos  E2) .  J(tf^ 


and 


2  -  c2). 


FIG.  121. 

If  we  eliminate  m  from  equation  (a)  by   the  help  of  the  equation 

P  +  m2  +  n'2=l,w6  obtain  o^2  +  o>22  =  «2  +  c-  -  l'2(a2  -  b-)  +  n-(b-  -  c2),  from 

which  we  obtain  by  use  of  equation  (d), 

(e)  <a*  +  o>22  =  a2  +  c2  -  (a2  -  c2)  .  cos  E1  .  cos  ^2. 

From  equations  (c')  we  obtain 

(a2  -  c2)  .  sin2^  =  a2  -  c2  -  Z2(a2  -  ft2)  -  w2(62  -  c2)  - 

(a2  -  c2)  .  sin2£2  =  a2  -  c2  -  /2(a2  -  62)  -  n2(&2  -  c2)  +  2/n  . 
Further,  since  Z2  +  m2  +  ri2  =  1  ,  we  also  have 

Z2(62  -  c2)  +  m2(a2  -  c2)  +  n2(rt2  -  i2)  =  a2  -  c2  -  ^(a2  -  ft2)  -  rc2(62  -  c2). 
By  help  of  these  relations  we  obtain  from  the  first  of  equations  (b) 
(f  )  w/2  -  w22  =  ±  (a-  -  c2)  .  sin  El  .  sin  E2,  and  from  (e)  and  (f  ), 
/  2Wl2  =  a2  +  c2  -  (a2  -  c2)  .  cos(^1  -  E2), 
\  2o>22  =  a2  +  c2  -  (a2  -  c2)  .  cos  (E^  +  E.2). 
The  greatest  value  of  the  velocity  of  propagation  is  a  and  the  least  c. 
This  follows  if  we  set  El  =  E2  and  El  +  E2  =  7r.      If  the  normal  to 
the  waves  is  parallel  with  one  of  the  optical  axes,  for  example  OAlt 
we  have  El  =  0  and  cos  \E.2  =  /0,  and  hence  Wl  =  o>2  =  b.     The  velocity 
of  propagation  is  then  equal  to  the  axis  of  mean  elasticity. 


SECT,  civ.]  THE   WAVE   SURFACE.  251 


SECTION  CIV.    THE  WAVE  SURFACE. 

Suppose  a  plane  wave  to  start  from  the  origin  of  the  system  of 
coordinates,  in  the  direction  in  which  its  normal  makes  angles  with 
the  axes  whose  cosines  are  I,  m,  n.  After  the  lapse  of  a  unit  of  time, 
the  distance  of  the  wave  from  the  origin  is  co.  If  about  each  of  the 
points  of  the  plane  wave  we  construct  a  wave  surface  as  it  would 
appear  after  the  lapse  of  unit  time,  the  plane  wave  thus  propagated 
is  the  envelope  of  all  the  wave  surfaces.  If  x,  y,  z  are  the  coordinates 
of  a  point  of  the  plane  wave  in  its  new  position,  we  have 

(a)  Ix  +  my  +  nz  =  (o, 
and  co  is  determined  by 

(b)  Z2/(a2  -  co2)  +  m2/(b2  -  co2)  +  n2  (c2  -  co2)  =  0. 

AVe  have,  further,  (c)  l2  +  m2  +  n2=l.  If  I,  m,  n,  and  co  vary,  the 
planes  (a)  envelope  a  surface,  which  is  called  the  wave  surface.  Hence 
if  we  consider  all  possible  plane  waves  passed  through  a  point,  and  if  we 
determine  the  position  of  the  same  waves  after  unit  time,  the  wave  surface 
is  the  envelope  of  all  the  plane  waves  thus  determined.  We  will  now 
investigate  the  equation  of  this  wave  surface.  AVe  obtain  from 
(a),  (c),  and  (b), 

(d)  x .  dl  +  y  .  dm  +  z .  dn  =  rfto, 

(e)  I .  dl  +  m .  dm  +  n  .  dn  =  0, 

(f )  I .  dl/(a2  -  co2)  +  m  .  dm/(b2  -  co2)  +  w .  dn/(c2  -  to2)  +  Fu  .  da>  =  0, 
where  (g)        F=  Pj(az  -  co2)2  +  m2/(&2  -  to2)2  +  7i2/(c2  -  co2)2. 

If  we  eliminate  e?co  by  means  of  (d)  from  equation  (f),  we  have 
[Ftax  +  //(a2  -  a>2)]d/  +  [F<ay  +  m/(b2  -  u2)]dm  +  [Fva  +  n/(c2  -  ta2)]dn  =  0. 

We  add,  to  the  left  side  of  this  equation,  equation  (e)  multiplied  by 
a  factor  A.  Since  dl,  dm,  dn  may  be  considered  as  arbitrary  quantities, 
we  have 

00 

If  these  equations  are  multiplied  in  order  by  /,  m,  n  respectively  and 
added,  we  obtain,  by  reference  to  (a)  and  (b),  A  =  -  Fur.     Therefore 
ra2  _  W2j  =  Fu(lu  -  x),   m/(b'2  -  to2)  =  Fn>(mo)  -  y), 

n/(c2  -  to2)  =  Fwtyw  -  z). 
If  we  square  both  sides  of  these  equations,  add  them,  and  use  equation 
(g),  we  have  1  =  /V(co2  -  2co(£c  +  my  +  nz)  +  r2),  in  which  r-  = : 


252  REFRACTION  OF  LIGHT.  [CHAP.  xn. 

Further,  by  reference  to  (a)  we  obtain  (k)  F<^(r2-  w2)  =  l.      F  may 
now  be  eliminated  from  equations  (i)  by  means  of  k,  and  we  have 
a1  -  o>2)  =  Z<o(a2  -  r2),   y(V-  -  to2)  =  mo>(Z>2  -  r2), 


These  equations  enable  us  to  determine  the  point  of  contact  between 
the  wave  surface  and  the  plane  wave,  and  therefore  the  direction  of 
propagation  of  the  ray.  The  plane  wave  moves  in  the  direction 
determined  by  Z,  m,  n. 

If  we  multiply  both  sides  of  equations  (1)  by  x,  y,  z  respectively, 
and  add,  we  have 

z2(a2  -  w2)/(a2  -  r2)  +  yz(b*  -  o>2)/(62  -  r2)  +  z2(c2  -  w2)/(c2  -  r2)  =  o>2, 
since  by  (a)  Ix  +  my  +  nz  =  w.     This  equation  may  be  written  in  the 
abbreviated  form  2z2(a2  -  w2)/(a2  -  r2)  =  w2,  or  in  the  form 
Sr2(a2  -  a>2)/(a2  -  r2)  =  2*2(a2  -  r2  +  r2  -  a>2)/(a2  -  r2) 


But  since  in  this  notation  2z2  =  x2  +  y2  +  z2  =  r2,  we  have  finally 

(r2-w2)(l+2z2/(a2-r2))  =  0. 
?%e  equation  of  the  wave  surface  is  therefore 


-  r-)  +  1-0. 
But  because 

Sa^/r2  =  1    and    2(x2/(a2  -  r2)  +  z2/r2)  =  2a2.c2/(a2  -  r2)  =  0 
we  may  also  write  the  equation  of  the  wave  surface  in  the  form 
(m)  aV/(a2  -  r2)  +  &y/(62  -  r2)  +  c%2/(c2  -  r2)  =  0. 

We  can  easily  transform  this  equation  into 

(n)  (aV  +  jy  +  c2*2)r2-(62  +  c>V-  (a2  +  c2)&V-(a2  +  62)c252  +  «2i2c2  =  0. 
7%g  equation  of  the  wave  surface  is  therefoi'e  of  the  fourth  degree.  In 
order  to  investigate  this  equation  we  set  x=fr,  y  =  gr,  z  =  hr.  By 
substitution  of  these  values  in  the  equation  of  the  wave  surface, 
it  becomes 
2,-2(a2/2  +  bY  +  c%2)  -  [(fr2  +  c>2/2  +  (a2  +  c2)62^2  +  (a2  +  52)c2A2]  =  ±  R, 


From  which  we  get 

B*  =  [(a2  -  c2)&V2  +  (afJW^c*  + 
x  [(a2  -  c2)6V  +  (a/v/^^c2  - 
Hence   a   straight   line   drawn   from   the   origin   of  coordinates   cuts 
the  surface  in  two  points,  which  coincide  when  R  =  Q  or  when 
(o)  .9  =  0  and  f/h  =  ±c/a.  J(a'-b'2)/(b'2  -c2). 


SECT.  CIV.] 


THE   WAVE   SURFACE. 


253 


In  this  case 

f=±e/b. V(«2 - &2)/(tt2 -~c~2),   h  =  ± ajb . *J(b2 - c2)/(a2 - c2). 
There  are   therefore   four  such   points   in   the   wave   surface,  all  of 
which  lie  in  the  o^-plane.      Hence  the  wave  surface  is  a  surface  of 
the   fourth   degree  with  two   nappes.      The   four  points   which   the 
two  nappes  have  in  common  are  called  umbilical  points. 

To  exhibit  the  form  of  this  surface  we  will  determine  the  curves 
formed  by  the  intersection  of  the  wave  surface,  and  the  coordinate 
planes  yz,  xz,  yx.  If,  for  this  purpose,  we  set  in  equation  (n)  z  =  0, 
y  =  0,  2  =  0  successively,  we  obtain 

-  Z>2c2)  =  0, 

2c2)  ^  Q, 

W)  =  0, 

Hence  the  curves  formed  by  the  intersection  of  the  wave  surface 
with  the  coordinate  axes  are  circles  and  ellipses,  as  represented  in 
figures  122,  123,  and  124.  The  curves  in  the  o^-plane  are  of 


9 


FIG.  122. 


b 

FIG.  123. 


c   b 
FIG.  124. 


special  interest.  The  equation  zz  +  x2  =  b2  represents  a  circle  of 
radius  b.  The  equation  c2£2  +  a?x2  -  a?c2  =  0  represents  an  ellipse 
whose  semi-axes  are  a  and  c.  On  the  assumption  that  a>b>c,  the 
circle  and  the  ellipse  intersect  at  a  point  P,  and  this  point  is  one 
of  the  umbilical  points. 

Equations  (1)  and  (h)  serve  to  determine  the  coordinates  of  the 
point  of  contact  between  the  wave  surface  and  a  plane  wave  which 
moves  in  a  direction  determined  by  I,  m,  n. 

The  case  in  which  the  wave  is  propagated  in  the  direction  of  one  of 
the  optic  axes  is  of  special  interest.  In  this  case  [CIIL],  the  velocity 
equals  b,  and  the  direction  of  propagation  is  given  by  the  equations 


since  we  here  consider  only  that  optic  axis  which  lies  between  the 
positive  directions  of  the  z-  and  a:-axes.      Equations  (1)  then  become 

z(a2  -  b2)  =  lb(a2  -  r2),   z(b*  -  c~)  =  nb(r2  -  c2). 


254  REFRACTION  OF  LIGHT.  [CHAP.  xn. 

If  we  introduce  in  these  equations  the  values  for  /  and  n  given 
above,  we  have 


(p)       xj(a2  -  62)(a2  -  c2)  =  b(a2  -  r2),   W(&2  -  c'2)(a*  -  c-)  =  b(r*  -  c2). 
These  equations  represent  two  spheres,  in  whose  lines  of  intersection 
lie  the  points  of  contact  of  the  wave  plane  and  the  wave  surface, 
therefore  in  this  case  the  plane  of  the  wave  touches  the  wave  surface  in 
a  circle. 

We  may  also  obtain   this   result   in   the  following  way.     By  the 
use  of  equations  GUI.  (c),  we  give  (p)  the  form 
(q)     x  =  b(a*-x2-f-z*)/l0(a2-c>-),   z  =  b(x*  +  y2  +  z'2-c2)/n0(a2-c2). 
The   curve   represented   by   these   two   equations    is    a   plane    curve 
because  (r)  xl0  +  zn0  =  b. 

We   now   introduce  a  new  system  of  coordinates  with  the   same 
origin  ;   suppose   the   ?/-axis   to   coincide   with    the   y-axis,  while  the 
£axis  coincides  with  the  optic  axis.     To  effect  this,  we  set 
(s)  z  =  £ra0  +  $o>   y  =  >?,   z=  -S/o  +  fV 

The  equation  (r),  which  represents  a  plane,  then  becomes  (t)  £=b, 
that  is,  the  plane  of  the  curve  of  intersection  is  perpendicular  to  the 
direction  of  the  optic  axis  and  passes  through  its  end  point.  The  first 
of  equations  (q),  by  the  use  of  (s)  and  (t),  takes  the  form 

(u)  f  +  ^Vo(rt2-<;2)/Z'  +  '?2  =  0• 

This  represents  a  circle,  which  passes  through  the  point  £  =  0, 
*7  =  0,  and  £=&,  or  through  the  end  point  of  the  optic  axis.  The 
radius  r  of  the  circle  is  r  =  >J(b2  -  c2)(a2  -  b'2)j  2b,  and  the  coordinates 
of  its  centre  are  £=—?',  t]  =•  0,  £=  b.  Thus  the  circle  is  determined 
in  which  the  plane  perpendicular  to  one  of  the  optic  axes  at  its 
end  point  touches  the  wave  surface. 


SECTION  CV.     THE  WAVE  SURFACE  (continued). 

Let  ON  (Fig.  125)  be  the  normal  to  a  plane  wave;  the  direction 
of  the  normal  is  determined  by  the  cosines  /,  m,  n.  Let  OPl  and 
OP2  be  the  two  velocities  of  propagation  of  the  wave  considered. 
Let  Ql  and  Q2  represent  the  points  of  contact  between  the  plane 
wave  and  the  wave  surface.  We  then  have  OQl  =  rl  and  0$0  =  ?V 
We  represent  the  coordinates  of  the  points  Ql  and  Q2  by  xv  yv  z^ 
and  x2,  y2,  z2  respectively.  If  QlPi=pl  and  Q.f^p^  are  the  per- 
pendiculars let  fall  from  the  points  of  contact  on  the  directions  of 
propagation,  we  have 


SECT.  CV.] 


THE   WAVE   SURFACE. 


255 


A 


The  connection  between  the  direction  of  the  normal  and  the  points 
of  contact  is  given  by  equations  (1)  CIV.     We  will  investigate  more 
particularly  the  directions  of  the  lines  p1  and  p2.     The  projection 
of    PQ   on    the    a-axis    is    ul-x.      If  we 
represent  the  cosines  of  the  angles  which 
p  makes  with  the  axes  by  A.',  /*',   v',  we 
will  have 

A'  =  (wZ  -  x)/p,    p!  =  (<am  -  y)/p, 

v'  =  (wn-z)/p. 

Introducing  in  this  equation  the  values  of 
x,  y,  z,  given  in  CIV.  (1),  we  obtain 
(a)     A'  =  lo>p/(a?  -  w2),    fjf  =  mupf(b2  -  to2), 

v'  =  n<apj(c2  -  w2). 

In   order  to  find  the  angle  between  PlQl 
and  P2Q.2,   we  determine  its  cosine 

= 


FIG.  125. 


But  because  [CIL  (f)] 

2/2/(a2  -  wj2)  =  0  and  2/2/(a2  -  o>22)  =  0, 
we  also  have          (wx2  -  w22) .  2/2/(a2  -  m^)(az  -  u>22)  =  0. 
Hence,  if  the  values  of  co3  and  to2  are  different,  we  have  cos(PlQlP2Q2) 
equal  to  zero,  and  the  angle  between  P^  and   P2Q2  a  right  angle. 
But  if  w1  equals  o>2,  the  points  Pl  and  P2  coincide,  as  we  saw  in  CIV. 
In  this  case,  there  is  an  infinite  number  of  points  of  contact  which 
lie  on  a  circle  passing  through  the  wave  normals. 

If  the  lines  P1Tl  and  P2r2  are  drawn  from  Pl  and  P2  perpendicular 
to  OQl  and  OQ2  respectively,  and  if  we  set  PlT1  =  ql  and  P27T2  =  j2, 
we  have  q:p  =  a)-.r,  and  therefore  q=p(ojr.  Further,  OT^aPjr.  If 
A,  /j.,  v  are  the  cosines  of  the  angles  which  q  makes  with  the  coordinate 
axes,  we  have  A  =  (coZ  -  OT .  x/r )/q,  etc.,  and  hence  [CIV.  (1)], 
(b)  A(a2  -  co2)  =  Ia2pjr,  p.(b2  —  co2)  =  mb2p/r,  v(c2  -  co2)  =  nc2pjr. 

If  we  compare  this  result  with  the  expressions  in  CIL  (e),  which 
determine  the  direction  of  the  electrical  force  F,  whose  components 
are  X,  Y,  Z,  we  see  that  the  electrical  force  is  parallel  to  q.  If 
we  introduce  in  the  equation  CIL  (d)  the  values  for  A,  /*,  v  given 
above,  and  notice  that  Ix  +  my  +  nz  =  to,  we  have  cos  8  —p/r.  Since 
there  are  two  directions  of  q,  namely  ql  and  q2,  there  are  two  direc- 
tions, ql  and  q2,  of  the  force  in  any  plane  wave.  These  lie  in  two 


256  EE FRACTION   OF   LIGHT.  [CHAP.  xn. 

planes  perpendicular  to  the  plane  wave.  There  are  two  values  for  8, 
namely,  ^OQ1P1  and  L.OQ.,PZ;  these  angles  are  equal  to  Ll^P-fl  and 
t-T^P^O  respectively. 

The  electrical  forces  X,  Y,  Z  cause  an  electrical  polarization,  whose 
variation  may  be  looked  on  as  an  electrical  current.  The  components 
of  current  [OIL  (a)  and  (c)]  are 

«  =  KJtir .  VX^i  =  Kj_ A/47T  .  9  Upt,  etc. 

If  A.0,  /x0,  v0  are  the  cosines  of  the  angles  made  by  the  axes  with 
the  direction  of  the  current,  we  have  A0 :  /*0  :  v0  =  A/a2 :  /z/52 :  v/c2. 
But  we  obtain,  by  the  help  of  OIL  (e), 

A0  :  pQ  :  v0  =  //(a2  -  w2) :  m/(b*  -  a>2)  :  n/(c2  -  w2). 

Hence  the  current  has  two  directions,  corresponding  to  the  two  values 
of  w.  From  equation  (a)  the  same  ratio  holds  between  the  cosines  of 
the  angles  which  p  makes  with  the  axes  as  between  the  cosines 
determining  the  directions  of  the  current.  Hence  the  two  directions 
of  the  current  are  parallel  to  p{  and  p.2  respectively. 

In  order  to  determine  the  direction  of  the  electrical  force  and  the 
current,  we  proceed  in  the  following  way.  If  a  plane  wave  moves 
in  the  direction  determined  by  the  normal  ON,  we  construct  two 
planes  which  touch  the  wave  surface  and  are  parallel  to  the  plane 
wave.  These  planes  are  those  constructed  at  Ql  and  Q2.  We  then 
draw  QlPl  and  Q2P2  perpendicular  to  the  wave  normal.  The  electrical 
currents,  which  are  in  the  wave  planes,  are  parallel  to  $iA  and  Q2P2. 
The  corresponding  velocities  of  propagation  are  OPl  and  OP2.  There 
are  two  directions  of  current  in  every  plane  wave,  which  are  perpen- 
dicular to  each  other.  The  electrical  forces,  which  are  connected  with 
these  directions  of  current,  are  parallel  to  P^T^  and  P2T2, 


SECTION  CVI.    THE  DIRECTION  OF  THE  RAYS. 

When  a  plane  wave  is  propagated  in  an  isotropic  medium,  the 
direction  of  the  normal  to  the  wave  coincides  with  the  direction  of 
the  ray.  In  doubly  refracting  media,  the  direction  of  the  ray  is  in 
general  different  from  the  direction  of  the  wave-normals.  We  will 
now  determine  the  direction  of  the  ray.  Let  MN  (Fig.  126)  be  the 
surface  of  a  doubly  refracting  body  on  which  the  cylinder  of  rays 
KOPL  falls  perpendicularly.  By  Huygen's  principle,  the  separate 
points  in  the  bounding  surface  OP  may  be  considered  as  centres  of 
luminous  disturbance.  The  luminous  disturbance  is  propagated 


THE   DIRECTION   OF  THE   RAYS. 


257 


through  the  body  in  such  a  way  that,  after  unit  time,  it  reaches  the 
wave  surfaces  which  are  constructed  about  the  separate  points  of 
the  bounding  surface  OP.  Therefore,  if  the  wave  surfaces  RA,  SO, 
etc.,  are  constructed  about  0,  P,  and  the  intervening  points,  we  obtain 
a  plane  ES  which  touches  every  wave  surface  and  is  congruent  to 
and  similarly  situated  with  OP.  The  direction  OR  or  PS  is  then  the 
direction  of  the  rays.  If  from  the  point  0  we  let  fall  a  perpendicular 
OB  on  the  plane  US  tangent  to  the  wave  surface  RA,  OB  =  u  is  the 
velocity  of  propagation  of  the  wave.  If  /,  m,  n  represent  the  direction 
cosines  of  the  normal  to  the  wave  surface,  o>  is  determined  by 
equation  GIL  (f) 

(a)  /2/(a2  -  "2)  +  m2l(W  -  w2)  +  n2/(c2  -  co2)  -  0. 


-fl 
M                 0 

Li 

P              JT 

J, 

•</ 

/ 

R       jE 

f         5 

FIG.  126. 

The  position  of  the  point  of  contact  of  the  plane  US  and  the  wave 
surface  RA  is  given  [CIV.  (1)]  from  the  equations 


(b)   x(a2  - 


-  r2),  y(bz  - 


where  x,  y,  z  are  the  coordinates  of  the  point  desired,  and  OR  =  r  is 
its  distance  from  the  origin  of  coordinates.  OB  represents  the  velocity 
of  propagation  of  the  wave,  OR  the  velocity  of  propagation  of  the  ray. 

Instead  of  the  wave  surface  itself  we  may  sometimes  use  to 
advantage  another  surface,  called  the  reciprocal  wave  surface.  Let  0 
be  the  centre  of  the  wave  surface,  AR  a  part  of  the  surface  itself,  and 
BR  a  plane  which  touches  the  wave  surface  at  the  point  R.  From 
the  point  0  we  let  fall  the  perpendicular  OB  on  the  tangent  plane. 
The  point  B',  in  the  perpendicular  OB  produced,  is  determined  in 
such  a  way  that  (c)  OB'  =  r'  =  s2/«>,  where  <i>  =  OB,  and  s  is  constant. 
The  reciprocal  wave  surface  is  then  the  locus  of  the  points  determined 
by  (c).  This  surface,  like  the  wave  surface,  is  a  surface  of  two 
nappes.  Its  equation  is  obtained  in  the  following  way.  If  I,  m,  n 


258 


EEFRACTION   OF  LIGHT. 


[CHAP.  xii. 


represent  the  direction  cosines  of  OB  =  to,  and  x',  y',  zf  the  coordinates 
of  the  point  B1,  we  have  (d)  x'  =  lr\  y'  =  mr',  d  =  nr'.  But  because 

l2/(a?  -  w2)  +  w2/(62  -  w2)  +  n2/(c2  -  w2)  =  0, 
it  follows  by  (c)  and  (d)  that 

x'2/(a2r'2  -  s4)  +  2rY(6V2  -  s4)  +  *'2/(cV2  -  s4)  =  0. 

We  set  (e)  a'  =  s2/a,  b'  =  s2jb,  c'  =  s2/c,  and  obtain  the  equation  of  the 
reciprocal  wave  surface  in  the  form 

(f )  a'  2x'2/(a'2  -  r"2)  +  b'2y'2j(b'2  -  r'2)  +  c'-Y2/(c'2  -  r'2)  =  0. 

This  surface  differs  from  the  ordinary  wave  surface  [cf.  CIV.  (m)] 
only  in  that  its  constants  a',  I',  c'  are  the  reciprocals  of  the  constants 
a,  b,  c  of  the  wave  surface. 

If  we   draw    through   the    point   B'   (Fig.    127)   a   plane    tangent 
to  the   reciprocal   wave  surface  FA',  we  can   show  that  the   plane 


B'R  is  perpendicular  to  the  prolongation  of  OR,  and  that  therefore 
OR  is  perpendicular  to  B'R.  Further,  if  OR  =  u>,  Ofi  =  r,  we  have 
(g)  w'  =  s2/r.  This  follows  by  the  same  method  by  which  we  have 
passed  from  one  surface  to  the  other.  We  can  also  prove  it  directly. 
If  the  direction  OR  is  determined  by  the  cosines  I',  m',  n',  we  have 
[CIV.  (1)]  (h)  z'(a'2-to'2)  =  rto'(a'2-'/2)  etc. 

The  equations  (h)  determine  w',  I,'  m',  n'.     Setting  u/  =  s'2/r  and 

l'  =  xfr,   m'  =  y/r,  ri  =  z/r, 

and  using  equations  (c),  (d),  (e),  (g),  equation  (h)  takes  the  form 
x(a?  —  w2)  =  Zo(a2  -  r2),  etc.  Since  these  equations  are  identical  with 
those  in  CIV.  (1),  it  follows  that  the  point  of  intersection  of  OR 


SECT,  cvi.]  THE  DIRECTION  OF  THE  RAYS.  259 

and  the  wave  surface  is  the  point  at  which  the  tangent  plane 
touches  the  wave  surface.  It  follows  further  from  (e)  and  (g)  that 
(i)  r'to  =  ra>  or  OB  .OB' =  OR  .OR.  In  order  to  determine  the  direction 
of  a  ray  from  the  reciprocal  wave  surface,  we  produce  the  wave  normal 
until  it  cuts  that  surface.  The  direction  of  the  ray  is  then  perpendicular 
to  the  tangent  plane  at  the  point  of  intersection. 


SECTION  CVII.    UNIAXIAL  CRYSTALS. 

If  two  of  the  constants  a,  b,  c  are  equal,  for  example  if  b  =  c,  the 
equations  become  much  simplified.  The  bodies  for  which  this  relation 
holds  are  called  uniaxial  crystals.  In  order  to  find  the  velocities  of 
propagation  w1  and  o>2,  we  apply  equation  CII.  (g)  which  is  trans- 
formed into  (a)  to4  -  [b2  +  r-b2  +  (l-  /2)a2>2  +  b2[l2b2  +  (1  -  I2)a2]  =  0. 
From  this  equation  we  obtain  (b)  o^2  =  b2,  o>22  =  I2b2  +  (1  -  l'2)a2.  Hence 
the  velocity  wx  is  constant ;  the  velocity  o>2  depends  on  the  direction 
of  the  wave  normal,  or  on  the  angle  which  the  wave  normal  makes 
with  the  axis  of  elasticity  a.  This  axis  is  called  the  optic  axis;  it 
coincides  with  the  principal  axis  of  the  crystal.  In  the  direction 
of  this  optic  axis  there  is  only  one  wave  velocity,  arid  therefore  also 
only  one  ray  velocity.  If  we  designate  the  angle  between  the  wave 
normal  and  the  optic  axis  by  e,  we  have  (c)  o>22  =  «2sin2€  +  62cos2e. 

Hence,  a  plane  wave,  on  its  passage  from  an  isotropic  to  an 
uniaxial  medium,  is  divided  into  two  waves,  one  of  which  is  propagated 
with  a  velocity  o^,  which  is  independent  of  the  direction  of  the  wave 
normal.  This  wave  is  called  the  ordinary  wave.  The  velocity  of  the 
other  or  extraordinary  wave  changes  with  the  direction  of  the  wave 
normal. 

AVe   obtain    the    equation    of   the 
wave    surface    for    uniaxial    crystals 
from  CIV.  (n),  by  setting  b  =  c.    We 
thus  obtain 
<d)  (r2  -  b2)(a2x2  +  b2(y2+z2)  -  aW)  -  0. 

Hence,  the  wave  surface  consists  of 
a  sphere  whose  radius  is  b,   and  an 
ellipsoid  of  revolution  whose  polar  and 
equatorial  axes  are  26  and  2a  respec- 
tively;  the  sphere  and  the  ellipsoid  FlG- 
touch   at  the    extremities    of  the    polar   axis.      In   Fig.    128, 
represents  the   polar  or  optic  axis,  AE1A1  a  plane  section  through 


260  EEFRACTTON  OF  LIGHT.  [CHAP.  xn. 

the  sphere  and  AR0A1  a  plane  section  through  the  ellipsoid.  Let 
OB^  be  the  normal  to  the  plane  wave  POQ,  B^D  and  B^R.2,  two 
planes  tangent  to  the  wave  surface,  which  are  both  perpendicular 
to  the  wave  normal.  OR^  and  OB.2  are  the  velocities  of  propagation 
in  the  direction  of  the  wave  normal.  We  call  such  a  plane  section, 
which  contains  the  optic  axis  as  well  as  the  wave  normal,  the 
principal  section.  The  direction  of  the  rays  of  the  extraordinary 
wave  is  represented  by  OR.2,  if  the  plane  B2R.2  touches  the  ellipsoid 
at  the  point  R.2.  The  direction  of  the  electrical  force  is  given  by 
B2U.»  which  is  perpendicular  to  OR.2.  The  direction  of  the  rays  and 
the  wave  normal  of  the  ordinary  wave  coincide,  and  the  direction 
of  the  electrical  force  is  perpendicular  to  the  plane  of  the  figure. 

The  polar  axis  AAl  =  2b  (Fig.  128)  is  greater  than  the  equatorial 
axis  2a  ;  the  crystals  for  which  this  occurs  are  called  positive  crystals. 
If  a>b,  the  crystal  is  called  negative.  The  sphere  can  enclose  the 
ellipsoid  or  inversely  ;  crystals  of  the  first  kind  are  called  positive,  those 
of  the  second  kind  negative.  Iceland  spar  is  a  negative  crystal, 
quartz  is  a  positive  crystal. 

If  we  set  o>  =  5  in  OIL  (e),  we  obtain  (e)  \1  =  Q  and  SI  =  ^TT,  that 
is,  the  direction  of  the  electrical  force  in  the  ordinary  wave  is  perpendicular 
to  the  optic  axis  as  well  as  to  the  wave  normal  :  it  is  therefore  perpendicular 
to  the  principal  section. 

In  order  to  obtain  the  direction  of  the  electrical  force  in  the 
extraordinary  waves  from  CIL  (e),  we  introduce  in  it  the  value 
for  <o2  given  in  (b),  and  notice  that  I  =  cos  e,  m  =  sin  e,  n  =  0.  We 
then  obtain 

^  _      a2cos62  _          &2cos8.2        v  _Q 

/2~(a2_^)COS£»    **2-     (a*  -b'2)  sine     V~~ 

Hence  the  direction  of  the  electrical  force  in  the  extraordinary  wave  is 
parallel  to  the  principal  section.  It  follows  from  the  last  equation 
that  l/cos282  =  (a4sin2e  +  &4cos2e)/(a2-&2)2sin2ecos2€,  and  hence 


(f  )  tg  82  =  ±  (a2sin2e  +  62cos2e)/(a2  -  &2)sin  e  cos 

We  thus  obtain  the  equations 


±a2sine/Va 

(g) 

,  =  0. 


SECT.  CVIH.]     DOUBLE   REFRACTION  OF  A  CRYSTAL.  261 

SECTION  CVIII.    DOUBLE  KEFRACTION  AT  THE  SURFACE  OF  A 
CRYSTAL. 

When  a  ray  of  polarized  light  falls  on  the  plane  surface  of  a 
doubly  refracting  medium  both  reflection  and  refraction  occur.  Let 
the  x-axis  of  the  system  of  coordinates  be  parallel  to  the  normal 
to  the  surface  drawn  outwards,  and  the  £-axis  perpendicular  to  the 
plane  of  incidence  ;  the  ?/-axis  is  then  parallel  to  the  line  of  intersec- 
tion between  the  plane  of  incidence  and  the  refracting  surface.  For 
the  components  of  the  electrical  force  of  the  incident  ray  we  have, 
as  in  C., 

(a)  X,  =  A/;,  Yt  =  pji,  Zt  =  v/i,  ft  =  fjCospir/T.  (t  -  (  -  xcos  a  +  y  sin  a)/0)], 
or,  using  only  the  real  part,  (b)/=JP«^-(-*wa+»atoBJWl.  In  these 
equations  a  is  the  angle  of  incidence  and  fi  the  velocity  of  the  light 
outside  the  crystal.  In  addition  to  these  equations  we  have  the 
condition  that  the  electrical  force  is  perpendicular  to  the  direction 
of  the  ray.  Hence,  since  the  direction  of  the  incident  ray  makes  the 
angles  TT  -  a,  |-TT  -  a  and  |TT  with  the  axes,  we  have 

(c)  -  AjCOS  a  +  /AjSin  a  =  0. 

In  the  corresponding  notation  we  have  for  the  reflected  ray 

(d)  A;  =  Xrfr,    Yr  =  /zr/r,    Zr  =  i>rfr,  fr  =  Fr.  ew  -  (irx+™ry+»r*)/n]. 

That  the  electrical  force  shall  be  perpendicular  to  the  direction  of 
the  ray,  we  must  have  (e)  Xrlr  +  p,rmr  +  vrnr  =  0.  Finally,  for  the  re- 
fracted ray,  we  have 

(f)     A>A6/6,    Yb  =  nJb,    Zb=vJM      (g)ft  =  F^-«**+™»+«»W. 

w  depends  on  lb,  mb,  nb,  or  on  the  direction  of  the  propagation  of  the 
refracted  wave.  The  boundary  conditions  are  the  same  as  those  of 
isotropic  bodies.  We  have  everywhere  in  the  bounding  surface,  for 
which  x  =  0,  Yt+Yr=YM  or 


Since  this  equation  must  hold  for  all  values  of  y  and  z,  we  have 
(h)  sin  a/0  =  mr/0  =  mb/o>  and  (i)  0  =  nr/£l  =  «6/w.  By  the  last  equation 
0  =  7ir  =  7iM  that  is,  the  wave  noi-mals  of  the  reflected  and  refracted  waves 
lie  in  the  plane  of  incidence.  It  follows  from  (h)  that  wr  =  sina,  that 
is,  the  angle  of  reflection  is  equal  to  the  angle  of  incidence.  Therefore 
the  direction  of  the  reflected  ray  is  determined  in  the  same  way  as 
in  the  case  of  reflection  by  an  isotropic  body. 


262 


REFRACTION   OF   LIGHT. 


[CHAP.  xn. 


If  (3  represents  the  angle  of  refraction,  we  have 

lb  =  -  cos  ft,   mb  =  sin  j3,    nb  =  0, 

therefore  from  (h)  sin  a/12  =  sin  /3/w.      If  we  determine  the  direction 
of  the  wave  normal  by  the  cosines  of  the  angles  which  it  makes  with 
the  axes  of  elasticity,  we  have,  to  determine  w,  the  equation 
(1)  /2/(a2  -  a)2)  +  TO2/(62  -  w2)  +  rc2/(c2  -  w2)  =  0. 

If  (xa),  (ya\  etc.,  denote  the  angles  between  the  axes  of  elasticity 
and  the  coordinate  axes,  we  have  l  =  lt>cos(xa)  +  mbcos(ya)  +  nl>cos(za). 
Introducing  here  the  values  for  lb,  mt,  etc.,  given  above,  we  obtain 

c    I  =  -  cos  (3 .  cos(za)  +  sin  /3 .  cos(ya) 

(m)  -j  m  =  -  cos  (3 .  cos(a$)  +  sin  /3 .  cos(yb) 

\.  n=  -cos/?,  cos(zc)  +  sin  ft.  cos  (yc). 

By  the  help  of  equations  (m)  and  (1),  w  can  be  expressed  in  terms 
of  (3.  The  equation  thus  obtained  in  connection  Avith  (k)  determines 
the  angle  of  refraction.  In  general  we  obtain  two  values  for  f3,  one 
or  both  of  which  may  be  imaginary ;  if  this  is  the  case  the  reflection 
is  total. 

We  can  find  the  direction  of  the  wave  normal  and  that  of  the  ray 
by  a  construction  given  by  Huygens.     About  the  point  0  (Fig.  129) 

as  centre  construct  the  sphere  PD, 
whose  radius  is  OD  =  Q,  where  12 
denotes  the  velocity  of  light  in  air. 
If  the  incident  ray  is  produced,  it 
meets  the  sphere  at  the  point  D. 
The  plane  which  the  sphere  touches 
at  D  cuts  the  refracting  surface  in 
a  straight  line,  whose  projection 
on  the  plane  of  the  figure  is  Q. 
The  plane  QR  containing  this  line 
is  drawn  tangent  to  the  wave  sur- 
129.  face  FR,  whose  centre  is  at  0. 

The  perpendicular  OB  =  i»  is  let  fall  from  0  on  the  tangent  plane 
QR.  The  normal  to  the  refracted  wave  is  then  OB  and  L'OB  =  (3, 
if  LOL'  is  the  normal  to  the  surface.  Now  OB=<a  is  the  velocity 
of  propagation  in  the  direction  OB,  and  also  OQ  =  OD/sina  =  OI>lsin(3, 
or  fl/sina  =  w/sin/:J,  so  that  equation  (k)  is  also  satisfied.  OB  is  the 
direction  of  the  wave  normal  of  the  refracted  wave,  and  OR  the 
direction  of  the  corresponding  ray.  Since  the  wave  surface  in  general 
has  two  nappes,  two  planes  tangent  to  the  wave  surface  can  be 


SECT,  cvni.]     DOUBLE   EEFEACTION   OF  A  CEYSTAL.  263 

drawn  through  Q.  The  construction  therefore  determines  two  wave 
normals  and  two  ray  directions. 

This  construction  really  serves  only  as  a  representation  of  the 
refraction;  it  cannot  be  used  for  the  determination  of  the  direction 
of  propagation  so  long  as  the  construction  is  confined  to  the  plane, 
because  the  point  of  contact  R  does  not  lie  in  the  plane  of  incidence ; 
we  can,  however,  obtain  the  direction  of  the  wave  normal  by  a 
construction  in  the  plane  of  incidence  given  by  MacCullagh. 

If  we  draw  through  D  (Fig.  129)  the  line  DE  perpendicular  to 
the  refracting  surface,  the  point  of  intersection  B'.  of  DE  and  the 
wave  normal  OB  is  so  situated  that  OB .  OB  =  OD2,  for  we  have 
OB  =  OQ.sinf3,  OB'  =  OE/sin(3,  and  further,  as  may  easily  be  seen 
from  Fig.  129,  OQ.  OE=OD2.  From  this  follows  the  relation 

OB.OB'  =  OD\ 

But  we  have  OB  =  u,  OD  =  Q,  and  if  we  set  OB'  =  r',  it  follows  that 
(n)  r'  =  12'2/w.  Therefore  the  point  B'  lies  on  the  reciprocal  wave 
surface  whose  equation  is  [CVI.  (f)] 

a'2x2/(a'2  -  r2)  +  b'2f/(b'2  -  r2)  +  c'2z2/(c'2  -  r2)  =  0 

if  the  coordinate  axes  are  parallel  to  the  axes  of  elasticity.     In  this 
equation  a'  =  fl2/a,   b'  =  ^2/b,   c'  =  02/c.     If  we  set 
JV^fi/a,    ^2  =  ^/6,    JV3  =  0/c, 

and  choose  as  the  unit  of  length  the  velocity  of  light  fi  in  the 
surrounding  medium,  it  follows  that 

( o)  N^KN^  -  r2)  +  N2y*l(N2  -  r2)  +  N32z2/(Ns2  -  r2)  =  0. 

This  is  the  equation  of  the  reciprocal  wave  surface.  It  follows 
further  from  the  discussion  of  CVI.  that  the  direction  of  the  rays 
OR  is  perpendicular  to  the  plane  tangent  to  the  reciprocal  wave 
surface  at  the  point  B. 

We  can  therefore  construct  the  wave  normal  in  the  following  way. 
About  the  point  0  as  centre  with  unit  radius  we  construct  the  circle 
PD ;  about  the  same  point  we  draw  the  curve  of  intersection  between 
the  plane  of  incidence  and  the  surface  (o).  This  curve  is  represented 
in  (Fig.  29)  by  B'F'.  W e  then  produce  the  incident  ray  to  the 
point  D  lying  on  the  circle,  and  draw  the  straight  line  DB'  perpen- 
dicular to  the  refracting  surface  and  cutting  F'B'  at  B'.  OB  is  then 
the  direction  of  the  wave  normal,  while  the  direction  of  the  ray 
OE  is  perpendicular  to  the  plane  tangent  to  the  surface  F'B  at 
the  point  B'.  We  can  easily  derive  the  condition  for  total  reflection 
from  this  construction.  It  can  also  be  applied  to  the  reflection  of 
light  within  the  crystal  itself. 


264  REFRACTION   OF   LIGHT.  [CHAP.  xn. 


SECTION  CIX.    DOUBLE  REFRACTION  IN  UNIAXIAL  CRYSTALS. 

Using  the  same  notation  as  in  CVIIL,  we  have,  to  determine  the 
angle  of  refraction  of  the  wave  normal,  the  equation 
(a)  sin  a/0  =  sin  /?/a>. 

In  the  case  of  uniaxial  crystals  w  has  the  values  w:  and  w2  which  are 
[CVII.  (b),  (c)]  o)12  =  i'2,  w92  =  a'2sin2e  +  £>2cos2e.  In  the  first  case  the 
angle  of  refraction  /31  is  obtained  from  the  equation 

sin  a  =  JV0  sin  /^  where  NQ  =  &/(OV 

the  so-called  ordinary  index  of  refraction.  The  corresponding  direc- 
tion of  the  electrical  force  is  perpendicular  to  the  principal  section. 
If  the  second  wave  normal  makes  the  angle  /32  with  the  normal  to 
the  surface,  we  have  sin  a/fl  =  sin  /?2/w.,.  If  we  represent  the  angles 
made  by  the  axis  of  the  crystal  with  the  coordinate  axes  by  (xa), 
(ya),  (za)  and  notice  that  TT  -  (3y  |z-  -  jS^  |TT,  are  the  angles  made 
by  the  refracted  ray  with  the  coordinate  axes,  we  have 

cosc=  -cos(««)cos/?2  +  cos(?/a)sin/32. 
Hence,  for  the  calculation  of  /32,  we  have  the  equation 

fl2sin2/32/sin2a  =  a2  -  (a2  -  Z>2)(cos(.ra)cos  /32  -  cos  (ya)sin  /32)2. 
The  corresponding  direction  of  the  electrical  force  is  parallel  to  the 
principal  section.     If  the  optic  axis  lies  in  the  plane  of  incidence  we 
set  cos(:ra)  =  cos  \f/,  cos  (ya)  =  sin  ^,  and  then  obtain 

J22sin2£2/sin2a  =  a2  -  (a2  -  62) .  cos2(^  +  /32). 

If  ^  =  a2sin2r/'+62cos2^,  JB  =  a2cos2^  +  62sin2^,  (7=(«2-62)sin  ^cos  f, 
we  have  A£-C2  =  aW2  and  I22/sin2a  =  A  cotg2/?2  +  2(7  cotg  /3.2  +  H. 
From  this  follows 


(b)  A  cotg  /32  =  -  C  +  v 

If  the  axis  of  the  crystal  is  perpendicular  to  the  plane  of  incidence,  we 
have  (xa)  =  (ya)  =  ^TT,  from  which  sin  a  =  JV^sin  /?2,  where  Ne  =  Q.ja,  is 
the  extraordinary  index  of  refraction.  If  a  and  b  are  expressed  in 
terms  of  N,  and  N0,  we  have  from  (b) 

r  (A^sin2^  +  JV^cosV)  cotg  /32  =  -  (N02  -  N*)  sin  ^  cos  $ 
I  +  JV0JVeX/sin--'a( JV^sin2 ^  +  ^Ve-'cos2 ^)  -  1 . 

In  order  to  obtain  the  equation  of  the  reciprocal  wave  surface,  we 
set  fl  =  l,  and  substitute  N,  for  a,  N0  for  b,  in  the  equation  for  the 
wave  surface.     Thus  we  obtain  [CVII.  (d)], 
(d)  (r>  -  N*%N&  +  NQ  %2  +  *2)  -  NJN*]  =  0, 


SECT,  cix.]  DOUBLE  REFRACTION  IN  UNIAXIAL  CRYSTALS.     265 


FIG.  130. 


as  the  equation  for  the  reciprocal  wave  surface  referred  to  the  axes  of 
elasticity  as  coordinate  axes.  We  obtain  the  same  result  from 
CVIIL  (o),  if  we  set  N^N.  and  N2  =  N3  =  N0. 

In  Fig.  130,  OP  is  the  refracting  surface,  and  OA  the  optic  axis, 
supposed  to  lie  in  the  plane  of  incidence ;  AMl  and  AM.2  are  the 
curves  in  which  the  plane  of  incidence  cuts  the  reciprocal  wave 
surface.  AM^  is  a  circle  with 
radius  NQ,  AM2  an  ellipse  whose 
semi-major  axis  OA  equals  N0, 
and  whose  semi-minor  axis  OM2 
equals  Ne.  We  draw  a  circle  of 
radius  OD  =  1 ,  which  cuts  at  D 
the  prolongation  of  the  incident 
ray.  The  line  ED,  perpendicular 
to  the  refracting  surface,  cuts  the 
reciprocal  wave  surface  at  the 
points  Bl  and  B.2.  The  normals  to  the  refracted  waves  are  then 
OBl  and  OB%.  For  the  ordinary  wave  the  direction  of  the  ray 
coincides  with  the  wave  normal  OBl ;  for  the  extraordinary  Avave  it 
is  perpendicular  to  the  plane  tangent  to  the  ellipsoid  at  the  point  B.2. 

If  the  crystal  is  immersed  in  a  fluid  whose  index  of  refraction  is 
greater  than  that  of  the  crystal,  the  circle  PD  is  replaced  by  another 
circle  of  greater  radius,  for  example  PD'.  If  this  circle  cuts  the 
prolongation  of  the  incident  ray  at  D',  the  directions  of  the  wave 
normals  are  determined  by  the  point  of  intersection  between  the 
reciprocal  wave  surface  and  the  line  UE',  perpendicular  to  the 
refracting  surface.  In  this  case  total  reflection  can  occur.  If  HE' 
does  not  cut  the  reciprocal  wave  surface  there  will  be  no  refraction; 
if  UE'  cuts  only  one  curve,  there  is  only  one  refracted  ray.  If,  as 
in  Fig.  130,  D'E'  touches  the  ellipse  at  a  point  C,  refraction  will 
occur ;  the  direction  of  the  ray  is  parallel  to  the  bounding  surface  OP. 


Our  presentation  of  optics  is  based  on  Maxwell's  conception  of  light 
as  an  electrical  vibration.  A  more  extended  discussion  on  this  same 
basis  has  been  given  by  H.  A.  Lorenz.  Glazebrook  published  a 
discussion  of  the  most  important  optical  theories  in  the  Report  for 
1885  of  the  British  Association  for  the  Advancement  of  Science, 
von  Helmholtz  has  lately  given  a  theory  of  the  dispersion  of  light  in 
which  he  employs  the  electromagnetic  theory  of  light. 


CHAPTER   XIII. 

THERMODYNAMICS. 
SECTION  CX.    THE  STATE  OF  A  BODY. 

IF  the  particles  of  a  system  are  in  motion  and  exert  force  on  one 
another,  the  system  possesses  a  certain  energy  U.  The  energy  of 
a  system  of  discrete  particles  is  made  up  of  their  kinetic  and  potential 
energies.  The  former  depends  on  the  velocities  of  the  particles  at 
any  instant,  the  latter  on  their  distances  apart,  or  on  the  configuration 
of  the  system ;  together  they  determine  the  state  of  the  body.  Thus 
the  energy  at  any  instant  depends  only  on  the  state  of  the  system 
at  that  instant,  and  is  independent  of  its  previous  states.  The 
principle  of  energy  has  been  proved  only  for  a  system  of  discrete 
particles ;  we  make  the  assumption  in  the  mechanical  theory  of 
heat,  that  the  same  principle  or  a  corresponding  one  holds  for  all 
systems  of  particles. 

A  certain  amount  of  energy  is  inherent  in  every  body.  This  we 
call  its  internal  energy,  since  we  take  no  account  of  that  part  of  its 
energy  which  arises  from  its  mutual  actions  with  other  bodies.  By 
the  possession  of  this  internal  energy  the  body  is  in  a  condition 
to  do  work ;  thus  variations  occur  in  its  form,  volume,  temperature 
etc.  The  energy  is  determined  solely  by  the  state  of  the  body ; 
if  the  body  in  a  certain  state  possesses  the  energy  U,  and  if  it  is 
subjected  to  any  variations  of  form,  magnitude,  etc.,  and  finally 
returns  to  its  original  state,  the  internal  energy  will  be  again 
equal  to  U. 

To  determine  the  internal  energy  of  a  body  it  is  necessary  to 
know  the  quantities  which  determine  its  state.  From  Boyle's  and 
Gay-Lussac's  laws  the  state  of  an  ideal  gas  is  completely  determined 
by  its  pressure  and  volume.  The  temperature  is  given  if  these  two 
quantities  are  known.  Boyle's  and  Gay-Lussac's  laws  furnish  an 
equation  which  expresses  the  relations  between  pressure,  temperature, 


CHAP.  xni.  SECT,  ex.]    THE   STATE   OF  A   BODY. 


267 


aud  volume ;  we  call  it  the  equation  of  state  of  a  gas,  because  it 
enables  us  to  determine  the  state  of  an  ideal  gas  under  any  con- 
ditions, if  it  is  known  under  definite  conditions,  for  instance,  at 
0°  C.  and  760  mm.  pressure.  The  behaviour  of  real  gases  cannot 
be  accurately  represented  by  an  equation  embodying  Boyle's  and 
Gay-Lussac's  laws,  but  conforms  to  other  equations  which  include 
these  laws  as  a  limiting  case.  The  state  of  a  fluid  is  in  general 
determined  by  the  same  quantities ;  it  depends  to  some  extent  on 
the  form  of  the  surface  and  the  nature  of  the  bodies  in  contact 
with  it.  The  actions  of  electrical  and  magnetic  forces  may  come 
into  play  in  both  gases  and  fluids.  As  a  rule  the  knowledge  of  a 
great  number  of  quantities  is  required  to  express  the  state  of  a 
solid,  especially  if  it  is  subjected  to  the  action  of  forces.  The 
equation  which  unites  all  quantities  which  determine  the  state  of 
a  body  is  called  the  equation  of  state. 

Since  the  state  of  a  gas  only  depends  on  the  pressure  p  and  the 
volume  v,  it  may  be  represented  by  a  point  in  a  plane  with  the 
coordinates  p  and  v ;  a  series  of  such  points,  or  a  curve,  represents 
a  series  of  successive  states.  The  t'-axis  of  this  system  (Fig.  131) 
is  drawn  horizontal ;  and  the 
^?-axis  vertical.  We  represent 
the  volume  and  pressure  of  the 
gas  in  its  original  state  by  r0 
and  p0 ;  its  state  is  then  given 
by  the  point  A.  If  the  gas  ex- 
pands under  constant  pressure, 
its  state  is  represented  by  a 
horizontal  line  AB,  parallel  to 
the  v-axis.  This  is  called  the 
curve  of  constant  pressure.  The 
curves  of  constant  volume  are 
vertical  straight  lines.  If  heat 
is  communicated  to  the  gas  whose  original  state  is  given  by  A,  at 
constant  volume  r0,  the  variation  of  the  state  of  the  gas  is  represented 
by  the  straight  line  AC,  and  its  pressure  increases.  If  the  tem- 
perature of  a  gas  remains  constant  during  its  successive  states,  we 
have,  from  Boyle's  law,  v.p  =  const.  Hence  the  curves  of  constant 
temperature  or  the  isothermal  lines  are  rectangular  hyperbolas  whose 
asymptotes  are  the  coordinate  axes.  In  order  to  change  the  state 
of  a  gas  in  such  a  way  that  its  temperature  remains  constant, 
there  is  required  either  compression  with  abstraction  of  heat  or  expansion 


J) 


B 


FIG.  131. 


268  THERMODYNAMICS.  [CHAP.  xin. 

with  communication  of  heat.  If  a  gas  whose  original  state  is  represented 
by  A  is  subjected  to  compression  with  abstraction  of  heat,  or  to 
expansion  with  communication  of  heat,  in  such  a  way  that  its 
temperature  remains  constant,  its  successive  states  will  be  represented 
by  the  hyperbola  DAE.  We  may  suppose  the  gas  enclosed  in  a 
receptacle  put  in  connection  with  an  infinitely  great  source  of  heat, 
whose  temperature  is  equal  to  that  of  the  gas  at  the  point  A.  If 
we  change  the  volume  of  the  gas,  the  source  of  heat  sometimes 
takes  up  heat  and  sometimes  gives  it  out,  but  the  gas  retains  the 
temperature  of  the  source.  If  the  gas  is  enclosed  in  an  envelope 
through  Avhich  heat  cannot  pass,  it  is  heated  by  compression  so 
that  its  temperature  rises,  or  cooled  by  expansion  so  that  its  tem- 
perature falls.  In  this  case  the  changes  of  state  are  called  adiabatic 
and  the  curve  which  represents  them  is  called  an  adiabatic  or  isentropic 
curve. 

The  state  of  a  solid  cannot  in  general  be  represented  in  a  plane, 
since  it  depends  on  more  than  two  variables. 

A  series  of  changes  by  which  the  state  of  a  body  is  altered  in  any 
manner,  and  which  is  such  that  the  body  finally  returns  to  its  original 
state,  is  called  a  cyclic  process.  If  a  body  goes  through  a  cyclic  process 
the  energy  which  it  receives  from  surrounding  bodies  is  equal  to  that 
which  it  gives  up  to  them.  The  steam-engine  is  a  system  of  bodies 
which  periodically  returns  to  the  same  state.  It  appears  from  the 
action  of  the  steam-engine,  that  heat  and  work  are  similar  or  equivalent 
quantities,  which  can  be  transformed  into  one  another,  and  are  both, 
therefore,  forms  of  energy.  This  conclusion  has  been  established  by 
accurate  experiment.  The  quantity  of  energy  produced  in  the  one 
form  is  always  proportional  to  that  applied  in  the  other  form.  This 
law  of  the  equivalence  of  heat  and  energy  was  first  formulated  by  K. 
Mayer  (1842).  The  later  observations  of  Joule  and  others  have  shown 
that  the  quantity  of  work  which  is  equivalent  to  a  unit  of  heat, 
or  to  the  quantity  of  heat  which  will  raise  the  temperature  of  a  gram 
of  water  by  1°  C.  is  equal  to  4.2.107  absolute  units  of  work  (C.G.S.). 
This  result  is  called  the  first  law  of  thermodynamics.  It  may  be  thus 
stated :  Heat  and  work  are  equivalent ;  work  can  be  obtained  from 
heat  and  heat  from  work.  The  work  equivalent  or  the  mechanical 
equivalent  of  the  unit  of  heat  is  designated  by  J. 

If  the  quantity  of  heat  dQ  is  communicated  to  a  body  it  receives  the 
energy   / .  dQ.      This  goes  partly  to  increase  the  internal  energy  U 
of  the  body,  partly  to  do  the  work  dW.     We  then  have 
(a)  J.d 


SECT.  CX.] 


THE   STATE   OF  A   BODY. 


269 


This  equation  is  called  the  first  fundamental  equation.  We  will  apply 
it  to  the  case  in  which  the  work  dW  is  done  by  expansion  against 
external  pressure. 

We  consider  the  body  ABC  (Fig.  132),  which  is  subjected  to  the 
hydrostatic  pressure  p  at  every  point  on  its  surface.  When  the  body 
expands  its  volume  becomes  A'B'C'.  The  normals  AA',  BB'  are 
drawn  from  the  surface- element  AB  =  dS  to  the  new  surface.  We  set 
A  A'  =  v  and  obtain  for  the  work  done  by  the  body, 

\vp .  dS=p^v  .  dS—p  .  dv, 

where  dv  denotes  the  total  increase  in  volume  of  the  body.  Equation 
(a)  then  becomes  (b)  J .dQ  =  dU+p .dv.  If  the  state  of  a  body  is 
determined  by  the  independent  variables  p  and  v,  the  definite  values  p^ 


FIG.  132. 


JL  £' 

FIG.  133. 


and  vl  correspond  to  a  point  A  (Fig.  133).  Suppose  the  body  to 
pass  through  a  series  of  states  represented  by  the  curve  ACB;  the 
values  p.  2  and  v.2  correspond  to  the  point  B.  We  then  have  from  (b) 

(c)  JQ=Ut-U1  +  £p.dv. 

Q  is  the  quantity  of  heat  introduced   during   the   change   of  state, 

£/2  -  ZTj  the  increase  of  the  internal  energy,  and  /  p  .  dv  the  work  done. 

The  increase  U2  -  Ul  is  determined  by  the  initial  and  final  values  of  p 
and  v,  or  by  the  position  of  the  points  A  and  B.  The  external  work 
is  measured  by  the  area  of  the  figure  A'  ABB'  A  ;  this  work  therefore 
depends  on  the  process  by  which  the  change  from  one  state  to  the  other  is 
effected.  This  holds  also  for  Q.  Since  U  is  a  function  of  p  and  v,  we 
obtain  (d)  J.dQ  =  'dU/'dp.dp  +  ('dUj'dv+p)dv.  If  the  function  U  is 
known,  it  is  possible  to  find  the  quantity  of  heat  necessary  to  produce 
any  change  in  the  state  of  the  body.  U  is  determined  from  equation 
(c),  by  measuring  the  quantity  of  heat  received  by  the  body  and  the 
quantity  of  work  done  by  it.  Our  knowledge  of  the  quantity  U  is 
still  very  limited. 


270  THERMODYNAMICS.  [CHAP.  xm. 

SECTION  CXI.    IDEAL  GASES. 

Clement  and  Desormes  and  subsequently  Joule  showed  that  the 
temperature  of  a  gas  which  expands  without  overcoming  resistance, 
that  is,  without  doing  work,  remains  unchanged.*  The  initial  and 
final  states  of  a  gas  which  expands  without  doing  work  lie  on  the 
same  isothermal,  that  is,  the  internal  energy  of  a  gas  is  a  function  of 
its  temperature  only,  and  is  therefore  independent  of  its  volume  if 
the  temperature  remains  constant.  If  we  take  the  temperature  0 
and  the  volume  v  of  the  gas  as  independent  variables,  we  have 

J.dQ  =  'dUJW  .de  +  'd  Ufdv .  dv  +p .  dr. 

Now  3 Ufdv  =  0  and  therefore  J.dQ  =  'd  U/W  .d6+p.  dv.  If  the  mass 
of  gas  contained  in  the  volume  v  is  equal  to  unity  then  ?>U/'dO  =  Jcn 
where  c,  denotes  the  specific  heat  of  the  gas  at  constant  volume,  that 
is,  the  quantity  of  heat  which  must  be  communicated  to  its  unit  of 
mass  in  order  to  raise  its  temperature  one  degree  in  such  a  way 
that,  while  its  pressure  changes,  its  volume  remains  constant.  If  the 
specific  heat  of  the  gas  at  constant  volume  is  constant,  its  internal  energy 
must  be  a  linear  function  of  its  temperature. 

For  ideal  gases  the  equation  giving  the  relation  between  pressure, 
volume  and  temperature  is  pv  =  R6,  where  R  is  a  constant.  If  0  and  v 
are  the  independent  variables  of  the  gas,  we  have  (a)  J.dQ  =  Jc,.d6+p.dv. 
From  the  observations  of  Regnault,  c,  is  independent  of  the  pressure 
and  temperature  of  the  gas.  If  6  and  p  are  chosen  as  the  independent 
variables,  v  must  be  considered  as  a  function  of  them,  so  that 

dv  =  "dvfde .  dO  +  'dvj'dp .  dp, 
-and  substituting  this  in  equation  (a)  we  have 

/.  dQ  =  (Jc,  +p .  'dv/Wfie  +p .  "dvfdp .  dp. 
From  the  equation  pv  =  E6,  it  follows  that 

p .  'dv/W  -  R  and  p .  'dv/dp  =  -  v,   and   J.dQ  =  (Jc,  +  B)dO  -  v .  dp. 

In  order  to  obtain  the  specific  heat  cp  at  constant  pressure,  that  is,  the 
quantity  of  heat  which  must  be  communicated  to  the  unit  of  mass 
of  the  gas  to  raise  its  temperature  one  degree,  in  such  a  way  that, 
while  its  volume  changes,  its  pressure  remains  constant,  we  set  dp  =  0 
and  obtain  cp  =  c,  +  ElJ,  (b)  J.dQ  =  Jcp.dO-v.dp.  If  p  and  v  are 

*More  exact  measurements  show  that  the  gas,  in  these  circumstances,  is 
slightly  cooled.  From  this  it  follows  that  there  are  attractive  forces  between 
its  separate  particles. 


SECT,  cxi.]  IDEAL  GASES.  271 

chosen  as  the  independent  variables,  we  have  dO  =  'dOj'dp .  dp  +  Wfdv .  dv. 
It  follows  from  pv  =  R0  that 

R.o0/'dp  =  v,   R.Wj?)v=p,   R.d0  =  v.dp+p.dv, 

and  from  (b)  that  (c)  R.dQ  =  c,v.dp  +  cpp . dv.  If  therefore  the  specific 
heat  cp  and  the  constant  R  are  known,  the  equation  (c)  enables  us  to 
determine  the  specific  heat  for  any  change  of  state  in  the  vp-p\a.ne. 
The  specific  heat  has  an  infinite  number  of  values  for  a  given  state 
in  the  t^-plane,  depending  on  the  direction  in  which  this  change  of 
state  takes  place. 

The  expressions  (a),  (b),  (c)  show  a  noteworthy  peculiarity.  If 
one  of  them,  say  (a),  is  divided  by  0,  we  obtain  by  the  use  of  the 
fundamental  equation  pv  =  R9,  J  .dQ/0  =  Jc,.d0/9  + R.dv/v.  If,  for 
example,  the  gas  passes  from  the  state  A  (Fig.  133)  to  the  state  B, 
and  if  the  temperatures  and  volumes  at  these  points  are  6V  vl  and 
00,  v2  respectively,  we  have  by  integration, 

(d)  /.  \dQjO  =  J.c..  log^flj)  +  R .  log(V»i)- 

Therefore,  while  the  integral  ^dQ  depends  on  the  path  on  which  the  gas 
passes  from  one  state  to  another,  the  integral  ^dQ/d  does  not  depend  on 
this  path. 

Clausius  called  the  quantity  S  =  J .\dQjO  the  entropy.  This  concept 
is  of  great  importance  in  the  theory  of  heat.  If  a  body  passes  from 
one  state  to  another  the  change  of  the  entropy  is  determined  by  the  coordinates 
of  the  initial  and  final  points.  This  theorem  is  here  proved  only  for 
a  gas,  but  holds  also  for  all  bodies. 

If  the  change  of  state  of  a  gas  occurs  along  an  isothermal  curve,  we 
have  from  (a)  J.dQ=p.dv.  Using  the  equation  of  state  and  inte- 
grating, we  obtain 

(e)  JQ =  £p .  dv  =  Rd .  log(^1). 

All  the  heat  communicated  is  therefore  used  in  keeping  the  tempera- 
ture constant.  If  we  set  v2  equal  to  fivv  fj?vv  p?vv  etc.,  in  succession, 
where  ^  is  any  number,  the  corresponding  values  for  Q  are 

Q  =  R0/J.logp,   2R6/J.  logp,   SRe/J.logp,  etc. 

If  the  change  of  state  occurs  along  an  isothermal  curve,  and  ii  the 
quantities  of  heat  introduced  are  in  arithmetical  progression,  the 
volumes,  according  to  equation  (e),  are  in  geometrical  progression ;  at 
the  same  time  the  pressure  changes  proportionally  to  the  density. 

If  the  change  of  state  occurs  along  an  admbatic  curve,  we  have  from 
(c)  cvlogp  +  cplog  v  =  cr  Setting  cp/c,  =  k  we  obtain  (f )  pv*  =  c,  where  c  is 


272  THERMODYNAMICS.  [CHAP.  xm. 

constant.  The  equation  (f)  is  the  equation  of  the  adiabatic  curves.  Com- 
bining this  with  the  equation  pv  =  B6,  we  have  from  (f)  Bdvk~l  =  c. 
If  we  introduce  in  this  formula  the  density  8  =  M/v  of  the  gas,  where 
M  denotes  its  mass,  it  follows  that  its  temperature  is  proportional 
to  the  (&-1)  power  of  the  density  when  the  state  of  the  gas  changes 
along  an  adiabatic  curve. 

Further  we  obtain  the  relation  [CX.  (b)]  fp.dv=Ul-  U.2.  The 
work  is  therefore  done  at  the  expense  of  the  internal  energy,  if  the 
change  of  state  is  adiabatic. 


SECTION  CXII.    CYCLIC  PROCESSES. 

A  simple  reversible  cycle  is  one  in  which  all  changes  occur  in  such 
a  way  that  if  reversed  they  may  be  effected  under  the  same  circum- 
stances. The  body  which  performs  the  cycle  is  called  the  working 
body.  In  the  performance  of  a  simple  reversible  cycle  the  working 
body  must  be  associated  with  two  others,  one  which  communicates 
heat  to  it,  and  another  which  receives  heat  from  it.  In  a  gas  engine 
the  working  body  is  the  gas  in  the  cylinder ;  in  a  steam  engine  it  is 
the  water  or  steam.  The  gases  of  the  fire  and  the  walls  of  the  boiler 
give  up  heat,  the  water  in  the  condenser  receives  heat.  The  gas  or 
steam  passes  through  a  series  of  states  and,  at  least  in  some  machines, 
returns  to  its  original  state;  it  is  then  in  condition  to  repeat  the 
same  process.  Since  the  value  of  the  internal  energy  U  at  the 
beginning  and  end  of  the  process  is  the  same,  we  have  [CX.  (a)] 
(a)  JQ=W, 

where  Q  is  the  difference  between  the  heat  received  and   the   heat 
given  up. 

The  quantity  of  heat  received  by  the  working  body  and  not  given 
up  to  the  colder  body  is  the  equivalent  of  the  work  done.  If  the 
working  body  is  a  gas,  we  have  for  the  cycle  JQ  =  ^pdi:  The 
entropy  of  a  gas  depends  only  on  the  coordinates  and  therefore  has 
the  same  value  at  the  beginning  and  end  of  the  process.  If  S}  denotes 
the  entropy  at  the  starting  point,  the  entropy  at  any  instant  during 
the  process  is  equal  to  Sl  +  \dQjQ.  If  the  integration  is  extended  over 
the  whole  cycle,  the  entropy  returns  again  to  its  value  Sv  and  we 
have  therefore  \dQ/B  =  0. 

We  will  discuss  more  particularly  a  special  case,  the  so-called 
Carnofs  cycle,  which  is  of  great  importance  in  the  theory  of  heat. 


SECT.  CXII.] 


CYCLIC  PROCESSES. 


273 


Suppose  a  gram  of  gas  to  be  in  the  state  represented  by  the  point  B 

(Fig.  134)  in  the  vp-plane.     The  curve  representing  the  cycle  is  in  this 

case  composed  of  two  isothermal   curves  BC  and   ED  and   of  two 

adiabatic  curves  CD  and  BE.     The  gas  first  expands  at  the  constant 

temperature  Or     This  is  accomplished  by  keeping  it  in  contact  with 

the    infinitely   great    body   Ml   at    the    temperature   0lt  and  by   so 

regulating  the  external  pressure  on  the  gas 

that  it  passes  to  the  state  C  along  the  path 

BC.      During   the   change   of  state   BC  the 

quantity    of   heat    Ql    is    absorbed    and    the 

work   represented   by  the   surface  BCC'B1  is 

done.      The  gas  then  expands  adiabatically, 

in  the  manner  represented  by  the  adiabatic 

curve    CD,   and   its   temperature   falls  to  02. 

Then  the  gas  is  brought  in  contact  with  an      I - — 4 »     i — if 


FIG.  134. 


infinitely  great  body  Me,  at  the  temperature 

02   and    compressed ;    during   this   process   it 

gives  up  to  M.2  the  quantity  of  heat  Q.     Its  state  is  represented  by  E. 

Finally  the  gas  is  further  compressed  without  communication  of  heat 

until  it  returns  to  the  original  state  B.     The  integral  ^p  .  dv,  extended 

over  the  whole  cycle,  equals  the  area  BCDE,  and  represents  the  work 

done  by  the  gas.     We  have  [CX.  (a)]  (c)  J(Ql  -  Q2)  =  W. 

When  the  gas  expands  from  B  to  C,  its  entropy  is  increased  by 
QI/OI  ;  it  remains  constant  along  the  path  from  C  to  D,  is  diminished 
by  Q.2/Q-2  along  the  path  from  D  to  E,  and  again  remains  constant  along 
the  path  from  E  to  B.  Since  the  gas  on  its  return  to  B  has  the 
same  entropy  as  at  the  outset  we  have 

From  (c)  and  (d)  it  follows  that 

Therefore  the  work  done  by  this  cyclic  process  is  proportional  to  the 
quantity  of  heat  Q}  absorbed  and  to  the  difference  of  temperature 
Ql  -  6.2,  and  is  inversely  proportional  to  the  absolute  temperature  Olt 
at  which  the  heat  is  absorbed.  The  heat  received  from  the  source  Ml 
is  not  wholly  transformed  into  work,  but  is  divided  into  two  parts,  one 
of  which  is  transformed  into  work,  and  the  other  transferred  to  M0. 

The  efficiency  £  of  the  Carnot's  cycle  is  the  ratio  of  the  heat  trans- 
formed into  work  to  that  communicated  to  the  gas.  We  have 


274  THERMODYNAMICS.  [CHAP.  xin. 

Hence  the  efficiency  depends  only  on  the  temperatures  dl  and  0.2  of 
the  sources  of  heat.  If  we  consider  the  reversed  cycle,  the  state  of  the 
gas  first  changes  along  BE ;  along  the  path  ED  it  receives  a  certain 
quantity  of  heat  from  M*,  and  has  a  certain  quantity  of  external  work 
done  upon  it  along  the  path  DC.  The  heat  received  and  the  work 
done  transformed  into  heat  are  given  up  by  the  gas  to  the  body  M^. 
along  the  path  CB.  In  the  case  of  the  cyclic  process  first  considered 
heat  is  transformed  into  work ;  in  the  reverse  process  work  is  trans- 
formed into  heat. 


SECTION  CXIII.     CARNOT'S  AND  CLAUSIUS'  THEOREM. 

It  was  shown  in  the  preceding  section  that  ^dQ/d  =  Q  for  any 
reversible  cycle,  when  the  body  describing  the  cycle  is  a  gas.  We 
will  now  see  if  this  theorem  holds  when  any  other  body  is  used 
as  the  working  body  instead  of  a  gas.  Let  us  take  the  simple 
case  in  which  the  process  is  carried  out  along  two  isothermal  lines 
BC  and  ED  (Fig.  134),  and  two  adiabatic  lines  CD  and  BE.  Suppose 
the  change  of  state  to  take  place  in  the  sense  given  by  the  letters 
BODE.  If  the  body  at  the  temperature  Ol  expands  from  B  to  C, 
it  receives  the  quantity  of  heat  Q^ ;  when  it  is  compressed  from 
D  to  E  it  gives  up  the  quantity  of  heat  Q.2.  Along  the  paths  CD 
and  EB  heat  will  neither  be  received  nor  rejected.  In  this  process 
the  total  quantity  of  heat  received  by  the  body  is  Ql  -  $„•  Since 
it  returns  to  its  original  state,  the  quantity  of  heat  Q1  -  Q2  is 
equivalent  to  the  work  done,  which  is  therefore  J(Ql  -  Q.2). 

S.  Carnot  published,  in  1824,  a  work  on  the  motive  power  of 
heat,  in  which  he  proposed  an  important  theorem  on  the  connection 
between  heat  and  work.  He  was  of  the  opinion  that  heat  was  a 
fundamental  substance  whose  quantity  remained  invariable  in  nature. 
Applying  this  view  to  explain  the  action  of  the  steam-engine,  he 
supposed  that  the  steam  gave  up  a  quantity  of  heat  Q1  at  the  higher 
temperature  6V  that  this  heat  was  transferred  to  the  condenser  at 
the  lower  temperature  02,  and  that  the  motive  power  of  the  heat 
was  due  to  its  passage  from  the  higher  to  the  lower  temperature. 
The  work  thus  done  by  this  passage  of  heat  from  a  higher  to  a 
lower  temperature  was  considered  analogous  to  that  done  by  n 
falling  fluid  or  by  any  falling  body.  This  latter  is  proportional  to 
the  weight  of  the  falling  body  and  to  the  distance  which  it  falls. 
Hence  for  the  work  done  by  the  heat  Carnot  proposed  the  expression 


SECT,  cxin.]      CAKNOT'S   AND  CLAUSIUS'   THEOREM.  275 

KQl(Ol  -  02)  where  K  is  a  function  of  the  absolute  temperatures  Ol 
and  02.  This  conclusion  of  Carnot  was  confirmed  by  experiment, 
but  did  not  agree  with  the  mechanical  theory  of  heat  in  so  far  as 
it  regarded  heat  as  an  invariable  quantity.  If  for  the  present  we 
disregard  this  error,  we  have  for  the  cycle  just  described 


Since  K  must  be  independent  of  the  nature  of  the  body  doing  the 
work  we  have,  if  the  body  is  a  gas  [CXIL  (e)],  (b)  K^J/0^  It 
therefore  follows  that  (Ql  -  Q^/^  -  02)  =  Q1/6l  and  hence  Ql/0l  =  Q2/02, 
that  is,  if  a  body  traverses  a  Carnot's  cycle  any  number  of  times,  by  being 
placed  alternately  in  contact  with  two  infinite  sources  of  heat,  the  quantities 
of  heat  which  it  receives  from  one  source  and  gives  up  to  the  other  are 
in  the  same  ratio  as  the  temperatures  of  the  sources. 

There  can  be  no  doubt  that  this  theorem  holds  for  a  cycle  of 
the  kind  considered.  The  application  of  the  theorem  in  many 
departments  of  physics  and  chemistry  has  led  to  no  results  which 
are  as  yet  contradicted  by  experiment.  Several  attempts  were  made 
to  give  a  direct  proof  of  the  theorem,  the  first  and  most  important 
of  which  is  due  to  Clausius,  whose  method  may  be  presented  in 
the  following  way  : 

Suppose  a  gas  to  traverse  the  cycle  BCDE  (Fig.  135)  composed 
of  the  isothermal  curves  BC  and  DE,  which  correspond  to  the 
absolute  temperatures  Ol  and  02,  and  of 
the  adiabatic  curves  CD  and  BE.  During 
its  expansion  from  B  to  C,  the  gas  takes 
the  quantity  of  heat  Ql  from  an  infinitely 
great  source  M-^  whose  temperature  Ol  is 
constant.  It  then  expands  from  C  to  D 
without  communication  of  heat.  It  is 
then  brought  in  communication  with  the 
infinitely  great  source  of  heat  M.>  whose 
temperature  00  is  constant,  and  by  com- 


:// 


pression   is   made  to  give  up  to  it  the     '  FIG   135 

quantity    of    heat     Qy       Finally    it    is 

brought  back  to  its  original  state  B.  During  the  cycle  the  gas 
has  received  from  the  source  Ml  the  quantity  of  heat  Qv  which 
is  divided  into  two  parts.  One  of  these  parts  is  transferred  as  a 
quantity  of  heat  Q2  to  the  source  M.2,  the  other  is  transformed 
into  work  and  is  represented  in  amount  by,  the  area  BCDE. 

Suppose  B'C'  and  E'D'  (Fig.  135)  to  be  the  two  isothermal  curves 
corresponding  to  the  temperatures  6l  and  9.2  for  another  body,  say 


276  THERMODYNAMICS.  [CHAP.  xni. 

for  water  vapour.  C'D'  and  B'E'  are  two  adiabatic  curves  so  chosen 
that  the  surface  BC'D'E  equals  the  surface  BODE.  If  the  water- 
vapour  is  subjected  to  a  process  similar  to  that  just  described  for  the 
gas,  the  heat  which  it  will  receive  while  in  contact  with  the  source 
Ml  is  Ql  +  q,  and  during  its  passage  from  H  to  E',  while  in  contact 
with  the  source  M2,  it  gives  up  to  that  source  the  quantity  of  heat 
Q2'.  The  work  done  by  the  vapour  is  equal  to  that  done  by  the  gas, 
because  the  surface  BCDE  is  equal  to  the  surface  B'C'DE',  and  we 
therefore  have  Ql  +  q  -  Q.2'  =  Ql  -  Q2,  and  therefore  Q2  =Q.2  +  q.  The 
vapour  in  expanding  along  B'(J  receives  the  quantity  of  heat  Ql  +  q, 
and  gives  up  the  quantity  Q2  +  q  along  the  path  D'E'. 

The  cycle  described  can  also  be  performed  in  the  opposite  sense. 
For  example,  the  water-vapour  can  expand  along  the  isentropic  curve 
B'E';  it  may  then  be  brought  in  contact  with  the  source  of  heat 
M2,  and  expand  from  E'  to  D',  during  which  expansion  the  quantity 
of  heat  Q2  +  q  is  taken  from  M2.  It  may  then  be  compressed  along 
the  isentropic  curve  Z^C",  and  lastly  along  the  path  (7-6',  while  in 
contact  with  the  source  of  heat  Mv  During  this  compression  it 
gives  up  to  Ml  the  quantity  of  heat  Ql  +  q.  To  carry  out  this  pro- 
cess, a  quantity  of  work  must  be  done  which  is  equivalent  to  the  heat 


this  work  is  represented  in  Fig.  135  by  the  surface  B'C'HE'. 

We  consider  finally  two  engines,  one  of  which  is  a  gas  engine, 
in  which  the  gas  performs  the  cycle  BCDE,  and  the  other  a  steam- 
engine,  in  which  the  steam  performs  the  reversed  cycle  B'C'D'E'. 
The  work  done  by  the  one  is  equal  to  that  supplied  to  the  other, 
if  we  neglect  friction  and  other  resistances.  The  gas  engine  in 
each  revolution  takes  from  the  source  Ml  the  quantity  of  heat  Ql 
and  gives  up  to  the  source  M.2  the  quantity  Q.2  ;  at  the  same  time 
the  steam-engine  takes  from  M2  the  quantity  of  heat  Q2  +  q  and 
gives  up  to  M^  the  quantity  Ql  +  q.  Hence,  in  these  circumstances, 
the  source  of  heat  at  higher  temperature  receives  during  each  revolu- 
tion the  quantity  of  heat  q,  while  the  source  at  lower  temperature 
M2  gives  up  the  same  quantity  of  heat  ;  this  transfer  of  heat  q  from 
the  lower  to  the  higher  temperature  being  effected  without  the  doing 
of  work. 

By  this  process,  therefore,  heat  can  be  transferred  from  a  colder 
to  a  hotter  body.  This  Clausius  declares  to  contradict  experience. 
While  heat  invariably  tends  to  flow  from  hotter  to  colder  bodies, 
in  the  process  described  above  the  opposite  occurs.  The  objection 
has  been  raised  to  this  conception  of  Clausius  that  a  thermoelectric 


SECT,  cxin.]      CARNOT'S  AND   CLAUSIUS'   THEOREM. 


277 


circuit,  in  which  one  junction  is  at  the  temperature  100°  and  the 
other  at  0°,  can  produce  a  current  which  will  heat  a  platinum  wire  red 
hot,  so  that  heat  passes  from  a  colder  to  a  hotter  body,  that  is,  to 
the  red  hot  platinum.  Clausius  answered  this  objection  by  asserting 
that  this  transfer  of  heat  to  a  higher  temperature  is'  compensated 
for  by  the  heat  generated  at  the  points  of  contact. 

Clausius  therefore  proposed  this  theorem :  Heat  can  never  pass 
from  a  colder  to  a  hotter  body  without  the  expenditure  of  work  or  the 
occurrence  of  some  change  of  state.  Hence,  by  this  principle  of  Clausius, 
2  =  0,  and  therefore  for  any  cycle  of  the  sort  described,  whatever 
body  is  used  in  it,  we  have  (a)  Ql/0l  =  Q2/Q2. 

We  can  now  show  that  a  similar  theorem  holds  for  a  cycle  of 
any  sort.  Suppose  that  the  change  of  state  of  a  body  proceeds  from 
B  along  the  curve  EG  (Fig.  13G).  If  the  isothermal  curve  BD  passes 
through  B  and  the  adiabatic  curve  CD  through  C,  we  may  replace 
the  path  BC  by  the  path  BDC,  that  is,  the  body  may  first  expand 
at  constant  temperature  along  BD  and  then  at  constant  entropy,  that 
is,  without  communication  of  heat,  along  DC.  If  BC  is  infinitesimal, 
BD  and  DC  are  so  also,  and  the  change  of  state  BC  may  be  replaced 
by  the  two  changes  BD  and  DC.  On  both  paths  the  body  receives 


FIG.  136, 


FIG.  137. 


the  same  increment  dU  of  internal  energy.  The  external  work  is 
in  the  one  case  BCC'B,  in  the  other  BDCC'B.  But  since  B'C'  =  dv 
is  infinitely  small,  while  B'B=p  remains  constant,  the  surface  BDC 
vanishes  in  comparison  with  the  surface  BCC'B.  If  dQ  represents 
the  quantity  of  heat  supplied,  we  have  J.dQ  =  dU+p.dv  along  the 
path  BC,  as  well  as  along  the  path  BDC. 

Let  BCPDEQ  (Fig.  137)  be  any  cycle,  Be,  Cc',  Ed,  Dd'  isothermal 
curves,    and  BE,  CD,  etc.,  adiabatic  curves.     Let  the  body  receive 


278  THERMODYNAMICS.  [CHAP.  xm. 

the  quantity  of  heat  dQ  along  EC,  and  give  up  the  quantity  dQ.2 
along  DE.  As  we  have  already  seen,  the  body  would  receive 
by  a  change  of  state  along  Be  the  same  quantity  of  heat  dQv  and 
give  up  along  dE  the  same  quantity  dQ.2.  Therefore  we  have  for  the 
cycle  BcdE,  dQl/6l  =  dQ.2/62,  if  6l  and  02  are  the  absolute  temperatures 
corresponding  to  the  isothermals  Be  and  Ed.  In  the  same  way  we 
have  for  Cc'  and  Dd',  etc.,  dQ^/O^dQ^/0^  dQ1a/0l"  =  dQ/l02"i  etc. 
If  Q  and  P  denote  the  points  at  which  the  cycle  touches  two  isothermal 
curves,  we  have  by  addition  (b)  \dQ1/0l  =  Je^/02,  where  dQ1  is  the 
quantity  of  heat  received  along  an  element  of  QBCP  and  6l  the  corre- 
sponding temperature,  dQ.2  the  quantity  given  up  along  an  element 
of  PDEQ  and  0.,  the  corresponding  temperature.  If  the  heat  received 
is  considered  positive  and  that  given  up  negative,  tJie  sum  of  all  infini- 
tesimal quantities  of  heat  received  during  the  performance  of  a  reversible 
cycle,  each  divided  by  the  absolute  temperature  at  which  it  is  received,  equal* 
zero,  that  is,  (c)  ^dQ/0  =  0.  This  is  the  second  law  of  thermodynamics. 
The  theorem  (c)  which  Clausius  first  expressed  in  this  form  may 
be  given  in  another  way.  Let  ABCD  (Fig.  138)  represent  a  cycle, 
so  that  f  dQ/6  =  Q.  We  divide  the  integral  into  two  parts, 

J  A  BCD  A 


of  which  the  first  is  extended  from  A 
over  B  to  C.  the  second  from  C  over 
D  to  Aj  and  have 


If,  therefore,  the  body  passes  from  the 

state  A    to   the  state   C,  the  value  of 

FlG  13g  the  integral  \dQ/6  is  independent  of  the 

path.     If  6l  and  i\  are  the  coordinates 
of  the  point  A,  and  62  and  v2  those  of  the  point  C,  we  have 


Clausius  introduced  a  special  symbol  for  the  entropy  by  setting 
dS  =  J.  dQ/e,  from  which  jCdQ/6  =  S2-  Sr     The  function  S  represents 

the  entropy  of  the  body ;  it  depends  only  on  the  state  of  the  body 
at  any  instant,  and  is  independent  of  all  previous  states. 


SECT,  cxiv.]       APPLICATION  OF  THE   SECOND   LAW.  279 

SECTION  CXIV.    APPLICATION  OF  THE  SECOND  LAW. 

We  have  already  obtained  [CX.  (a)]  the  equation 

(a)  J.dQ  =  dU+p.dv. 

If  the  state  of  the  body  is  determined  only  by  the  independent 
variables  6  and  v,  equation  (a)  may  take  the  form 

(b)  /.  dQ  =  (VUfdff).  .  dO  +  ((3  Ufdv)e  +p)dr, 

where  the  indices  attached  indicate  that  the  quantities  which  they 
represent  remain  constant  during  differentiation.  If  S  denotes  the 
entropy,  we  have 

dS  =  J.  dQ/6  =  1/6  .  (3  UfiO),  .dO+(l/e.  (d  Upv)g  +pi'B)dv. 
Since  S  is  here  a  function  of  v  and  B,  we  may  set 

(3S/30),  =  1  JO  .  (3  £7/30),  ;     (dSf'dv),  =  1/6.  (dUfdv),  +p/0. 
But  for  the  same  reason  we  have  also 


and  further  3(3S/30),/3t;=l/0.  3(3  tf/30),/  30, 

3(3S/3»)9/30=  1/0.  3(3C7/30)e/30-  1/02. 
Whence  it  follows  that  (c)  (3Z7/30)e  =  02.  3(p/0),/30,   since   U  is  also 
a  function  of  0  and  v  only,  and 


The  internal  energy  must  satisfy  the  differential  equation  (c).  The 
second  law  furnishes  the  means  of  determining  the  internal  energy. 
It  follows  from  equations  (c)  and  (b)  that 

(d)  J.dQ  =  (3£//30X  .  d6  +  (<BFd(pl&)Jde  +p)dv. 

Hence  if  the  equation  of  state  and  the  specific  heat  c,=  l/J.  (dU/W)f 
are  known,  the  quantity  of  heat  required  for  a  given  change  in  the 
state  of  the  body  may  be  determined  by  equation  (d). 

The  quantity  of  heat  which  a  body  has  received  is  not  determined 
by  the  state  of  the  body  at  any  given  instant,  and  therefore  cannot  be 
considered  as  a  function  of  the  coordinates.  We  can,  however,  set 

(e)  J(dQ[d0).  =  @Uj-dO)n   J(-dQ/-dv)e  =  (-dU/-d-v)e+p, 

since  (dQ;"dO),  .  dO  is  the  quantity  of  heat  which  is  used  in  raising  the 
temperature  by  dd,  while  the  volume  remains  constant,  and  (dQj'dv)e  is 
the  quantity  which  is  absorbed  during  an  increase  in  volume  by  dv 
at  constant  temperature.  But  'd-QjWdv  is  not  equal  to 
From  equation  (c)  we  obtain 

-  l/J. 


280  THERMODYNAMICS.  [CHAP.  xm. 

The  differential  equation  (c)  is  applied  to  the  relations  of  an  ideal 
gas,  for  which  p/0  =  E/v.  From  this  relation  (dUj"dv)e  =  Qt  which  agrees 
with  the  results  in  CXI. 

If  the  energy  of  a  body  at  constant  temperature  is  independent  of 
its  volume,  its  equation  of  state,  from  (c),  will  have  the  form 

P/e=f(v). 


SECTION  CXV.    THE  DIFFERENTIAL  COEFFICIENTS. 

As  a  rule  the  equation  of  state  of  a  body  is  unknown.  There  are, 
however,  many  bodies  for  which,  within  narrow  limits,  we  know 
approximately  the  relations  of  volume,  pressure,  and  temperature. 
Within  these  limits,  therefore,  an  equation  of  state  may  be  constructed 
of  the  form  (a)  f(v,  p,  0)  =  0.  If  the  pressure  p  is  constant,  we  have 

f(p,  v  +  dv,  6  +  d6)  =  0   and   3//3»  .  dv  +  3//30  .d8  =  Q. 

The  ratio  between  dv  and  dO  is  written  in  the  form  (c3f/3$)p.     The 
volume  v  of  the  body  is  generally  given  in  the  form 


where  t=0-273.     Hence  we  have 

(St;/30),  =  i;(a  +  2/3(0  -273)  +  ...)/(!  +a(<9-273)+  ...). 
If  /3  is  very  small,  we  obtain  (b)  (dv/c)6)p  =  va.     This  formula  can  be 
used  even  when  a  is  a  function  of  6. 

If  the  temperature  0  is  constant  in  equation  (a),  we  have 

"dffdp  .  dp  +  'df/'dv  .dv  =  0. 

From  this  equation  the  change  of  volume  due  to  change  of  pressure 
at  constant  temperature  can  be  determined.     Since  the  volume  always 
diminishes  when  the  pressure  increases,  (dv/'dp)g  is  negative.     From 
the  theory  of  elasticity  (cf.  XXIX.)  we  have  found  that 
XVv/v  =  -'dp  and   6  =  "dv/v  =  -  (1  -  -2k)3?>p/E, 
in  the  case  of  fluids  and  solids  respectively.     Therefore 
(c)  (3r/3p)a=-r/X. 

It  follows  from  the  equation  of  state  of  an  ideal  gas  that 

(dv[dp)e=  -v/p, 
and  hence  \=p. 

If  the  volume  of  the  body  is  constant,  the  pressure  is  increased  by  dp 
by  the  rise  of  temperature  dO  ;  we  obtain  in  this  way  a  third  quantity 
Besides  these  differential  coefficients  we  must  also  notice 


SECT,  cxv.]         THE   DIFFERENTIAL   COEFFICIENTS. 


281 


three  others,  Wj'dv,  ty/'dv,  and  30/3p,  which  are  connected  with  those 

already  mentioned  by  the  following  relations  : 

(d)  (B»/30),.  (30/3^=1,  W9p)..(3p/3»).=  l,  (3p/30)..(a0/3p).=  l. 

Let  LM  and  NP  (Fig.   139)  represent  two  isothermal  curves  cor- 
responding to  the  temperatures  9  and  6  +  d6.     We  then  have 

(3j9/3#)fl  =  tg  AEv, 
if  AE   is    the    tangent   at   the 
point  A  of  the  isothermal  curve 
whose  parameter  is  0.     If  BAD 
is  perpendicular  to  Ov,  and  ^46' 
parallel  to  Ov,  we  have 
tgAEv  =  -AD/DE=  -AS/AC, 
and  hence 

(3p/3r)fl  =  -AB/AC. 
Further,  we  have 


and  hence  we  obtain  (e)  (3p/3w)8  .  (dvfd6)p  .  (d0/ty).  =  -1. 

Equations  (d)  and  (e)  show  that  if  we  know  two  of  these  differential 
coefficients  which  are  independent,  the  others  are  also  known.  Equa- 
tion (e)  may  be  derived  in  the  following  way  :  If  p  is  considered 
a  function  of  v  and  6,  we  have  dp  =  (3p/3v)e  .  dv  +  (dp/W),  .  d8.  Assum- 
ing the  pressure  constant,  so  that  dp  =  0,  we  have 

dv/d6  =  -  (3p/30)./(3p/di;)fl. 

The  quantity  dv/dQ  in  this  equation  is  that  which  has  already  been 
designated  by  (dv/?>6)p  •  we  therefore  again  obtain  equation  (e). 

For  gases  we  have  (dppv)e  =  -p/v;  (dvpO)p  =  E/p  ;  (deity),  =  v/R 
These  values  satisfy  equation  (e). 

For  liquids  and  solids  we  have  (dvj^>Q\  =  av,  (dvfty)e  =  —  v/X,  and 
hence  by  equation  (e),  (f)  (ty/W)v  =  aX. 


SECTION  CXVI.    LIQUIDS  AND  SOLIDS. 
If  6  and  v  are  the  independent  variables,  we  have  [CXI.] 

(a)  J.dQ  =  (3 U/W).  .dO+((d U/dv),  +p)  .  dv. 

From  the  second  law  of  thermodynamics  we  obtain  as  in  CXIV.  (c), 

(b)  (dU/c;v)g  =  62  .  3(^/0),/30  =  0  .  (ty[d6),-p.       Hence   equation    (a) 
takes  the  form  (c)  J.dQ  =  (pU[b9\.  d6  +  6.(dp/W),.  dv.     If  we  designate 


282  THEEMODYNAMICS.  [CHAP.  xm. 

the  specific  heat  at  constant  volume  by  cv  we  have  (d)  /.  cv  =  (dU/'d0')v. 
Equation  (c)  then  becomes  [CXV.  (f)]  (e)  dQ  =  ca.d0  +  6aX  .  dvfJ.  It 
follows  from  equations  (b)  and  (d)  that 

(f  )  /.  3c,/dt>  =  32  Upv'dO  =  6  .  3(3p/30X/30, 

and  therefore  from  CXV.  (f)  that  (g)  .7.3e,/30=0.3(aA.)/30.  This 
may  be  obtained  also  from  equation  (e)  by  the  use  of  the  second 
law.  Equation  (g)  shows  that  c,  is  independent  of  the  volume  if 
aA.  is  independent  of  the  temperature. 

In  order  to  express  the  dependence  of  the  quantity  of  heat  com- 
municated to  a  body  on  its  pressure  and  temperature,  we  set 

dv  =  (dvfdff)p  .  dO  +  (do[dp)9  .  dp, 
and  then  obtain  from  (c) 


j.dQ={(d  Z//30),  +  e  .  (dppe),  .  (dvpe)p}d0  +  e  .  (dppo),  .  (3^/3/4  .  dp, 

or  [CXV.  (e)],  (h)  J.dQ=  {Jcv-  0.(3?>/30)*/(30/3p)e}<Z0-  6.(dvjW)p.dp. 
If  cp  is  the  specific  heat  at  constant  pressure,  we  have 


Now  since  'dv/'dp  is  always  negative,  cp  is  greater  than  cv,  so  long  as 
30/30  is  not  equal  to  zero;  the  case  in  which  30/30  =  0  is  exhibited 
by  water  at  4°  C. 

Introducing   the   values   for  the   differential   coefficients   found    in 
CXV.,  we  obtain 


(i)        J.dQ={Jc,  +  a?X.v8}d6-av6.dp  and   (k)  cp  =  c,  + 

The  temperature  of  a  fluid  or  of  a  solid  is  changed  by  compression. 
If  we  set  dQ  =  0  in  equation  (i)  the  rise  of  temperature  dd  due  to 
the  increase  of  pressure  dp  is  dO=  +av6{cpJ.dp,  that  is,  the  tempera- 
ture rises  with  increasing  pressure  if  a  is  positive,  that  is,  if  the  body 
expands  when  heated  ;  if  a  is  negative  the  temperature  falls  ivith  increasing 


SECTION  CXVII.    THE  DEVELOPMENT  or  HEAT  BY  CHANGE  OF 
LENGTH. 

If  the  pressure  p  is  exerted  on  each  unit  of  surface  of  the  ends  of 
a  solid  cylinder,  each  unit  of  length  of  the  cylinder  is  shortened  by 
p/f,  where  e  is  a  constant.  If  /  denotes  the  original  length  of  the 


SKCT.  cxvii.]  THE   DEVELOPMENT  OF  HEAT.  283 

cylinder  at  0°  C.,  its  length  L  at  the  temperature   0  and  under  the 
pressure  p,  if  the  limits  of  elasticity  are  not  exceeded,  is 

(a)  L  =  l.(l-p/t).  (1  +£(0-273)), 

where  (3  is  the  coefficient  of  expansion. 

If  the  length  of  the  body  increases  by  dL  and  its  temperature 
rises  by  dO,  the  quantity  of  heat  dQ  must  be  supplied  to  it  ;  the  work 
done  by  the  elongation  is  A  .  p  .  dL,  if  A  is  the  cross-section  of  the 
cylinder.  If  M  represents  the  mass  of  the  body  when  the  tempera- 
ture is  6  and  the  length  L,  we  have 

(b)  J.dQ  =  M.'dU/W.d8  +  (M  .  3  Z7/3L  +  Ap)  .  dL. 
Applying  the  second  law  to  this  expression  we  obtain 

-d(M/0  .  3  £7/30)/3£  =  -d(M/6  .  3  UfdL  +  Aplff)fdO 
or  M  .  (d  U/3L)g  =  0*-.A.  "d(pl8)JdO. 

Hence          J.dQ  =  M.(dUfd8)L.d6+8.A.  (op/W)L  .  dL. 
If  6  and  p  are  taken  as  the  independent  variables,  we  have 

dL  =  (dL/W)p  .  d6  +  (3Z/3p)fl  .  dp  and 
J.dQ  =  [M.  (dUfd8)L  +  B  .  A  . 


Since  the  deformation  is  very  small,  we  have,  representing  by  cp 
the  specific  heat  at  constant  pressure, 

(c)  J.dQ  =  JMcp  .dO-BA.  (dL/W)p  .  dp, 

since  by  analogy  with  CXV.  (e)  (dpfd6)L  .  (30/3^  .  (3£/3/>)e  =  -  1  . 
If  the  pressure  on  the  ends  is  increased  by  dp,  so  that  the  total 
pressure  is  A.dp  =  P,  and  if  there  is  no  communication  of  heat,  the 
temperature  of  the  body  increases  by  d9  =  6PL^!JMcp,  or,  if  m  is 
the  mass  of  unit  length,  by  dd=B/3P/Jmcp.  If  the  cylinder  is 
stretched  by  the  force  P  a  corresponding  cooling  will  occur. 


SECTION  CXV  III.    VAN  DER  WAAL'S  EQUATION  OF  STATE. 

The  equation  of  state  of  an  ideal  gas  is  pv  =  B&,  and  its  isothermal 
curve  is  therefore  a  rectangular  hyperbola.  Real  gases,  however, 
at  low  temperatures  and  under  high  pressures,  do  not  conform  to 
this  equation.  Suppose  that  a  certain  quantity  of  gas  at  a  given 
temperature  has  the  volume  OC'  (Fig.  140)  and  is  under  the  pres- 


284 


THERMODYNAMICS. 


[CHAP.  xni. 


PG 


sure  CO'.  If  the  pressure  is  increased  while  the  temperature  remains 
constant,  the  volume  will  be  diminished.  At  last  the  space  in  which 
the  gas  is  contained  becomes  saturated  with  it ;  let  the  corresponding 
pressure  be  DD'.  DD'  is  then  the  pressure  of  the  saturated  vapour  or 
the  vapour  pressure  at  the  given  temperature.  If  the  volume  is  still 
further  diminished,  the  pressure  remains  constant,  while  a  part  of 
the  vapour  passes  over  into  the  liquid  state.  At  last  all  the  vapour 
is  transformed  into  liquid;  let  the  corresponding  volume  be  OF. 
So  long  as  the  vapour  and  liquid  are  in  the  same  space,  the  isothermal  curve 
is  a  straight  line  parallel  to  the  axis  Ov.  If  the  volume  of  the  liquid 
is  now  diminished,  the  pressure  increases  very  rapidly ;  the  corre- 
sponding isothermal  curve  is  represented  by  FG  (Fig.  140).  Andrews 
found,  by  experimenting  with  carbon-dioxide,  that,  as  the  temperature 

rises,  the  line  DF  becomes  shorter, 
and  that,  at  a  certain  temperature, 
which  he  called  the  critical  tempera- 
ture, it  disappears  altogether.  If 
the  temperature  of  the  carbon- 
dioxide  remains  constant,  its  state 
changes  along  curves  which  are 
represented  in  Fig.  141.  The 
abscissas  represent  volumes,  the 
ordinates  pressures.  For  example, 
let  us  examine  the  isothermal 

0  — ^~> —  TV ~£, —     curve  ABCD,   which  corresponds 

to  the  temperature  15'1°C. ;  at 
the  point  A  the  carbon-dioxide  is 

still  in  the  gaseous  state;  at  B  it  may  be  considered  as  saturated 
vapour.  If  the  compression  is  continued,  condensation  begins,  and 
the  pressure  remains  constant  until  the  substance  has  become  liquid, 
that  is,  until  its  state  is  represented  by  C.  From  C  on,  the  pressure 
increases  very  rapidly  as  the  volume  is  diminished.  At  the  tem- 
perature 21-5°C.  the  condensation  begins  at  B,  and  the  horizontal 
part  of  the  curve  is  shorter.  At  31'1°C.  the  horizontal  part  of  the 
isothermal  curve  vanishes ;  the  critical  temperature  has  now  been 
reached.  Isothermals  corresponding  to  higher  temperatures  are 
continuous  curves ;  it  is  therefore  impossible  to  reduce  carbon-dioxide 
to  the  liquid  state  at  a  temperature  higher  than  31-1°C.  At  higher 
temperatures  than  this  there  is  no  apparent  difference  between  its 
liquid  and  gaseous  states.  The  liquid,  at  the  critical  temperature, 
has  the  same  density  as  the  saturated  vapour.  A  gas  can  be  reduced 


SECT,  cxvni.]  VAN  DEE  WAAL'S  EQUATION  OF  STATE. 


285 


to  the  fluid  state  by  compression  only  when  its  temperature  is   lower  than 
the  critical  temperature. 

James  Thomson  substituted  for  the  isothermal  curve  here  described 
a  continuous  curve  CDHEJFG  (Fig.  140);  the  part  DHEJF  cor- 
responds to  an  unstable  state.  It  appears  from  various  investigations 
on  the  relations  between  vapours  and  their  liquids  at  the  boiling 
point,  that  it  is  possible  to  obtain  a  vapour  in  the  states  represented 


FIG.  141. 


by  DH  and  FJ,  while  the  states  represented  by  HEJ  are  always 
unstable,  since  in  these  states  the  pressure  and  volume  change  in 
the  same  sense. 

J.  Clerk  Maxwell  called  attention  to  an  important  peculiarity  of 
these  isothermals  which  may  be  deduced  by  applying  the  laws  of 
thermo-dynamics.  If  a  gas  traverses  the  cycle  FEDHEJF,  in  passing 
along  the  straight  line  from  F  to  D  it  receives  the  quantity  of  heat 


286  THERMODYNAMICS.  [CHAP.  xm. 

L,  and  in  passing  along  the  curve  DHEJF  gives  up  the  quantity 
L'  ;  the  temperature  is  the  same  along  both  paths.  Since  the  gas 
has  traversed  a  complete  cycle,  we  have  \dQ/0  =  L/0  -  L'/O  =  0,  and 
therefore  L  =  L'.  Therefore,  since  no  heat  is  used  in  this  cycle,  no 
work  can  be  done,  and  hence  the  surface  FJE  is  equal  to  the  surface 
DEE.  Thus,  if  the  isothermal  curve  CDHEJFG  is  given,  the 
maximum  pressure  of  the  vapour  can  be  determined,  by  determining 
the  line  FD  so  that  the  surfaces  FJE  and  DHE  are  equal. 

Van  der  Waals  has  proposed  an  equation  of  state  for  gases,  which 
represents,  more  exactly  than  the  simple  one,  the  behaviour  of  the 
gas  and  which  permits  the  calculation  of  the  critical  temperature. 
The  volume  of  a  gas  is  determined  not  only  by  the  external 
pressure  but  also  by  the  attraction  of  its  molecules  ;  we  may  think 
of  this  attraction  as  replaced  by  a  pressure  p  added  to  the  external 
pressure  p.  Since  the  attracting  and  attracted  molecules  approach 
one  another  as  the  density  increases,  p'  must  be  directly  proportional 
to  the  square  of  the  density,  and  therefore  inversely  proportional 
to  the  square  of  the  volume.  Hence  we  set  p'  =  a/v*,  so  that  the 
total  pressure  of  the  gas  is  p  +  a/v2.  Further,  the  molecules  are  not 
free  to  move  everywhere  in  the  region  v,  for  they  themselves  occupy 
part  of  the  region.  Van  der  Waals  assumed  that  the  volume  of  a 
fluid  cannot  fall  below  a  certain  limit  without  the  particles  losing 
their  freedom  of  motion.  In  place  of  the  apparent  volume  v,  he  used, 
as  the  true  or  effective  volume,  v  -  b,  where  b  is  a  very  small  quantity, 
though  much  greater  (about  4-8  times)  than  the  volume  of  all  the 
molecules  of  the  gas.  We  thus  obtain  the  equation  of  state 


If  the  volume  v  is  very  great,  this  equation  becomes  the  equation 
of  state  of  ideal  gases. 

The  positions  of  the  points  H  and  /  (Fig.  140),  at  which  the 
tangents  to  the  isothermal  curves  are  parallel  to  the  r-axis,  are 
obtained  from  the  equation  dpfdv  —  Q  or  (b)  p  +  a/iP  —  2a(v—  fc)/t'3  =  0. 
This  is  the  equation  of  the  curve  which  passes  through  all  points 
at  which  the  tangents  to  the  isothermal  curves  are  parallel  to  the 
axis  Ov.  All  these  isothermal  curves  correspond  to  temperatures  at 
which  the  body  can  be  either  liquid  or  gaseous.  When  the  two 
points  coincide  we  reach  the  critical  state.  But  since  the  two  coin- 
cident points  must  have  a  line  joining  them  parallel  to  the  axis  Oi; 
we  introduce  the  condition  for  the  critical  state  by  setting  dp/dv, 
obtained  by  differentiating  the  foregoing  equation  (b),  equal  to  zero,  or 
•<c)  6rtv  -  bv*  -  4rtV3  =  0. 


SECT,  cxvui.]  VAN  DER  WAAL'S  EQUATION  OF  STATE. 


287 


If  vl  denotes  the  critical  volume,  that  is,  the  volume  of  unit  mass 
of  the  gas  or  liquid  at  the  critical  temperature,  we  obtain  from  the 
last  equation  (d)  vl  =  3b.  The  critical  temperature  6V  and  the  critical 
pressure  plt  which  must  exist  in  order  that  the  fluid  shall  not  boil 
at  a  temperature  which  is  lower  by  an  infinitesimal  than  the  critical 
temperature,  are  (e)  01  =  l/E.8a/27b,  pl  =  a/2762.  These  values  are 
obtained  by  introducing  the  value  of  t\  in  (b)  and  (a). 

Choosing   vv    0:,   and  pl   as    units   of  volume,   temperature,    and 
pressure,  we  may  set 

v  =  V*i  =  V.  3b,  p  =  PPl  =  P.  fl/27ft2,    e  =  T6l  =  T.  8aj27bR, 
and  thus  give  to  the  equation  of  state  the  form 
(f)  (P  +  3/r2)(3F-l)  =  ST. 

From  this  equation  we  obtain  the  following  law  :  If  the  critical  pressure 
is  taken  as  the  unit  of  pressure,  the  critical  volume  as  the  unit  of  volume, 
and  the  absolute  critical  temperature  as  the  unit  of  temperature,  the  iso- 
•thermals  of  all  bodies  are  the  same.  We  will  now  consider  some  applica- 
tions of  formula  (f). 

(a)  Equation  (f)  may  be  written  in  the  form 


The  following  table   shows   how  the   product   PV  depends  on  P 
when  the  temperature  is  constant  : 


1 
F 

r=l-0 

r=i-5 

r=2-o 

r=2-5 

T=3'0 

PV 

P 

PV 

p 

PV 

P 

PV 

P 

PV 

P 

o-o 

2-67 

o-oo 

4-00 

o-oo 

5-33 

o-oo 

6-67 

o-oo 

8-00 

o-oo 

0-2 

2-26 

0-45 

3-69 

0-74 

5-11 

1-02 

6-54 

1-31 

7-97 

1-59 

0-4 

1-88 

0-75 

3-42 

1-37 

4-96 

1-98 

6-50 

2-60 

8-03 

3-21 

0-6 

1-53 

0-92 

3-20 

1-92 

4-87 

2-92 

6-54 

3-92 

8-20 

4-92 

0-8 

1-24 

0-99 

3-06 

2-44 

4-87 

3-90 

6-69 

5-35 

8-51 

6-81 

1-0 

1-00 

1-00 

3-00 

3-00 

5-00 

5-00 

7-00 

7-00 

9-00 

9-00 

1-2 

0-85 

1-01 

3-07 

3-68 

529 

6-35 

7-51 

9-01 

9-73 

11-68 

1-4 

0-80 

1-12 

3-30 

4-62 

5-80 

8-12 

8-30 

11-62 

10-80 

15-12 

1-6 

0-91 

1-46 

3-77 

6-03 

663 

10-60 

9-49 

1538 

12-34 

19-75 

1-8 

1-27 

2-28 

4-60 

8-28 

7-93 

14-28 

11-27 

20-28 

14-60 

26-28 

2-0 

2-00 

4-00 

6-00 

12-00 

10-00 

20-00 

14-00 

28-00 

18-00 

36-00 

The  results  of  this  table  may  be  represented  by  a  figure,  in  which  P 
is  the  abscissa  and  PV  the  ordinate ;  in  the  neighbourhood  of  the 


288 


THERMODYNAMICS. 


[CHAP.  xin. 


critical  temperature  T=l,  PV  is  very  variable,  and  the  departures 
from  the  ordinary  laws  are  in  consequence  very  considerable.  As 
the  temperatures  rise  the  relations  change  rather  rapidly,  PV  becoming 
nearly  constant.  The  product  PV  has  a  minimum  value,  which  in 
some  cases  is  reached  only  when  the  pressure  is  negative.  This 
minimum  is  determined  by  dPVIdF=3/V*-8T/(3V-  1)2  =  0.  PV 
is  therefore  a  minimum  if  3F-1/FW8T/3. 

The   corresponding  values  of  P  and   F",  which  we  may  designate 
by  Pm  and  Vm,  are,  by  the  use  of  the  equation  of  state,  equal  to 

Vm  =  l/( 


Pm  =  3(2v/8773  -  3)(3  - 

and  hence  PmFm=6x/8f/3  -  9. 
Hence  we  obtain  lor 


T=l-0 
Pw=l-09 


1-5 
3-00 

3-00 


2-0 

3-35 

4-86 


2-5 

2-72 

6-49 


3-0. 
1-37. 

7-97. 


As  an  example  of  the  application  of  the  relations  here  developed, 
the  following  table  has  been  calculated  for  18°C.  or  0=291  : 


Crit.  Temp. 
ei 

Crit.  Press. 
P, 

T 

Pm          P=Pmpi 

Hydrogen, 

33 

7-3 

neg.        neg. 

Nitrogen, 

127 

33 

2-29 

3-08  !      102 

Carbon  -monoxide,  - 

132 

36 

2-20 

3-19  |      115 

Oxygen,  - 

155 

50 

1-88 

3-37 

169 

Nitrogen-monoxide, 

179 

71 

1-63 

3-21 

228 

Methane, 

191 

55 

1-52 

3-05 

168 

Ethylene, 

263 

51 

I'll 

1-68 

86 

The  product  of  pressure  and  volume  therefore  diminishes  as  the 
pressure  increases,  if  the  pressure  is  less  than  102  atmospheres.  If 
the  pressure  is  greater  than  this  the  product  increases  with  the  pressure. 
(b)  The  coefficient  a,,  which  is  called  the  coefficient  of  pressure 
(formerly  called  the  coefficient  of  expansion  at  constant  volume),  and 
which  denotes  the  change  of  pressure  for  a  rise  of  temperature  of 
1°  C.  at  constant  volume  of  the  gas,  is  determined  in  the  following 
way :  We  have 

a.  =  1/p .  -dp/W  =  1/^P .  VPfdT=  1/0! .  8/(3PF-  P) 


SECT,  cxvin.]  VAN  DER  WAALS'  EQUATION  OF  STATE.  289 

For  very  great  values  of  V,  a,  =  1/6*,  corresponding  to  an  ideal  gas. 
Further  a,  .  6l  =  l/(T-  f  .  1/F+  1  .  1/F2).  Van  der  Waals  defines  car- 
responding  states  as  those  in  which  the  volumes,  temperatures  and 
pressures  of  both  gases  are  in  the  same  ratio  to  the  same  quantities 
in  the  critical  state,  that  is,  in  which 

v/v,  -  *//<  -  r,  p/Pl  =P'iplf  =  P,  BIOI  =  &/e{  =  T, 

where  the  quantities  v',  »/,  p',  etc.,  refer  to  the  second  gas  ;  thus  we 
have  <i,6l  =  a,'  6^  or  a,/a,'  =  &1'/Bl.  Hence  the  coefficients  a,  and  a,'  in 
corresponding  states  are  inversely  as  the  critical  temperatures. 

(c)  The  coefficient  of  expansion  ap,  which  represents  the  increase 
of  volume  for  a  rise  of  temperature  of  1°C.  at  constant  pressure,  is 
defined  in  the  following  way  : 


With  increasing  values  of  F",  ap  approaches  the  value  ar  If  the  bodies 
are  in  corresponding  states,  we  have  apdl  =  ap'6l'.  If  the  basis  for 
this  method  is  sound  it  should  find  application  in  the  expansion  of 
liquids  by  heat.  Since  changes  of  pressure  have  only  slight  effect  on 
the  volume  of  liquids,  it  is  sufficient  to  compare  the  coefficients  of 
expansion  at  corresponding  temperatures.  The  calculations  made 
by  van  der  Waals  have  shown  that  this  law  is  in  essentials  correct 
for  liquids  whose  critical  temperatures  are  known. 

(d)  The  pressure  of  saturated  vapours.  —  To  determine  the  pressure  of 
saturated  vapours  we  use  the  above-mentioned  theorem  of  Maxwell, 
which  states  that  the  surface  F'FEDD'  (Fig.  140)  is  equal  to  the 
surface  F'FJEHDH,  that  is,  setting 

OF'=TV    OD'=Vv   and  F'F=D'D  =  PV 

we  have  P    F!  -  F   =    *P  .  d  V. 


Effecting  the  integration  and  using  the  equation  of  state  (f)  it  follows 
that 


(g)          fs 

For  the  points  F  and  D  the  following  equations  hold  : 

(Pl  +  3j  F12)(3  ri-l)  =  8T  and  (Pl  +  3/  F22)(3  F2  -  I  )  =  8T. 
If  F!  and  F2  are  eliminated  from  these  equations,  a  relation  is  obtained 
between  the  pressure  Pl  of  the  saturated  vapour  and  the  tempera- 
ture T.  We  may  therefore  conclude  with  van  der  Waals  that  :  If 
for  different  bodies  the  absolute  temperature  is  the  same  multiple  of  the 
critical  temperature,  the  pressure  of  their  saturated  vapours  -is  also  the  same 
multiple  of  their  critical  pressures.  Similar  laws  hold  for  the  relation 


290  THERMODYNAMICS.  [CHAP.  xin. 

between  the  volume  of  the  saturated  vapour  and  its  pressure  and 
temperature. 

On  the  basis  of  van  der  Waals'  investigations,   Clausius  has  pre- 
sented a  slightly  altered  form  of  the  equation  of  state,  viz., 


This  equation  represents  the  actual  relations  better  than  the  other, 
but  leads  to  essentially  the  same  results.  Starting  from  the  equation 
(h)  J.dQ  =  (dUfdO)je  +  ((dU/'dv)e+p)dv,  and  applying  the  second 
law  we  obtain  the  relation  [CXIV.  (c)],  (3  U/^  =  P  .  3(^/0)./30. 
From  Clausius'  equation  p/0  =  E/(v  -&)-  a/  (P(v  +  {$}-,  and  hence 

(i)  (dU/Vv)6=2a/6(v  +  W. 

If  the  temperature  increases  by  dd  and  the  volume  by  dv,  the  internal 

energy  increases  by 

(k)  dU=Jc,.dO  +  2a/e(v  +  py2.dv, 

where  Jc,  =  (3  UfdO),  is  an   unknown   function  of  6  and  v.      If  the 

changes  of  temperature  are  slight,  c,  may  be  considered  constant,  as 

is  shown  by  observation.      If  the  temperature  increases  from   0  to 

0  +  A0,  while  v  increases  from  ^   to  v.2,  the   increment  A?7  of  the 

internal  energy  is  approximately  equal  to 

(1)  A  U=  Jc,  .  A0  +  2a/B  .(Ifa-  l/«2). 

If  the  gas  expands  without  resistance  the  internal  energy  remains 

constant.     Hence,  if  we  set  A  £7  =  0  in  equation  (1),  we  have 

(m)  A0  =  -  2a/Jc,0  .  (Ifa  -  l/v2). 

In  this  case  the  temperature  falls. 


SECTION  CXIX.    SATURATED  VAPOURS. 

If  the  volume  of  a  gram  of  a  certain  substance,  or  its  specific  volume, 
is  called  vl  when  it  is  a  vapour  and  v.2  when  it  is  a  liquid,  the  volume 
v  occupied  by  a  gram  of  liquid  and  vapour  is  (a)  v  =  vlx  +  v2(\-x), 
if  the  volume  contains  x  grams  of  vapour  and  therefore  (1  -x)  grams 
of  liquid.  If  U±  is  the  internal  energy  of  the  vapour,  U2  that  of 
the  liquid,  the  internal  energy  of  the  mixture  of  the  two  is 
(b)  U=UlX+UJl-x). 

So  long  as  the  vapour  is  saturated,  its  internal  energy  and  pres- 
sure depend  only  on  the  temperature;  the  pressure  and  volume 
of  the  liquid  are  also  determined  by  it.  Hence,  for  a  mixture  of 


SECT,  cxix.]  SATUEATED  VAPOUKS.  291 

liquid  and  saturated  vapour,  we  have  (c)  p=f(Q),  where  p  is  the  pres- 
sure of  the  saturated  vapour  at  the  temperature  0,  and  consequently  6  is 
the  boiling  point  of  the  liquid  under  the  pressure  p. 

That  quantity  of  heat  which  is  needed  to  transform  one  gram  of 
liquid  into  vapour  at  constant  temperature  6  and  under  the  corre- 
sponding pressure  p  is  called  the  heat  of  vaporization  L.  This  heat 
is  partly  used  in  increasing  the  internal  energy,  partly  in  doing 
external  work.  If  U.2  denotes  the  energy  of  the  liquid,  U-^  that  of 
the  vapour,  at  the  temperature  6,  the  internal  energy  is  increased  by 
£/!  -  U.-,  by  the  transformation.  Since  the  pressure  p  is  constant, 

the  external  work  is  I  pdv  =p(vl  -  v2).     The  work  needed  to  evaporate 

a  gram  of  liquid  is  (d)  J.L=Ul-  U^+p^-v^. 

If  a  mixture  of  liquid  and  vapour  receives  heat  and  also  changes 
its  volume,  the  increase  of  internal  energy,  since  0  and  x  both 
vary,  is 

<e)  dU=(U1-  U.2)dx  +  ( 

Since,  from  equation  (a), 
(f  )  dv  =  (i\  -  v.2)dx  +  (x  . 

it  follows  from  the  equation  J.  dQ  =  dU+p.  dv  that  the  heat  imparted 
to  the  mixture  is  determined  by 


(  /.  dQ  =  {x(dUl/W+p  .  S^/30)  +  (1  -  x)(dU2/W+p  . 
\ 


Vi-  U.2-pv.2}dx. 

If  in  this  equation  we  set  dd  =  Q,  and  integrate  from  x  =  0  to  x=\, 
we  obtain  equation  (d).  If  equation  (g)  is  brought  into  the  form 
J.dQ  =  Q.dd  +  X.  dx,  it  follows,  from  the  Carnot-Clausius  theorem, 

(JT/0)/30.     But  we  have 
=  1/0.  (3  UJV0  +p  . 
c>(Z/0)/c>0  =  1/0  .  (3  UJW  +p  . 


Hence  it  follows  that  (h)  U^  -  U^(v^  -  »2)  .  0  .  3p/30  -p(v1  -  »2),  and 
thus  the  difference  between  the  internal  energy  of  the  vapour  and 
that  of  the  fluid  is  determined. 

We  obtain  from  equations  (d)  and  (h),  (i)  JL  =  (vl  -vz)B  .  'dp/W. 
Now  we  know  by  observation  that  ^  >  v0,  so  that  3p/30  is  positive. 
The  boiling  point  is  therefore  higher,  the  higher  t)ie  pressure. 

We  may  also  apply  equation  (i)  to  the  process  of  melting.  In 
that  case  i\  denotes  the  volume  of  the  liquid,  vz  that  of  the  solid. 
We  must  here  distinguish  between  two  kinds  of  substances,  those 


292  THERMODYNAMICS.  [CHAP.  xm. 

like  wax,  whose  volume  increases  during  melting,  and  those  like  ice, 
whose  volume  diminishes  during  melting.  For  the  former  i\  >  r.2, 
and  therefore  'dpI'dB  is  positive;  for  the  latter  i'l<r.2,  and  therefore 
3^/30  is  negative.  For  those  substances  whose  volume  increases  during 
melting,  the  melting  temperature  rises  as  the  pressure  increases.  For  those 
substances  whose  volume  diminishes  during  melting,  the  melting  temperature 
falls  as  tlie  pressure  increases. 

If  the  volume  is  always  filled  with  saturated  vapour  only,  it 
follows  from  (g),  since  «=1,  that  (k)  J.dQ  =  (dU1l'd6+p.'drlfd6)d0. 
Hence  dQ  is  the  quantity  of  heat  which  must  be  imparted  to  the 
vapour  that  its  temperature  shall  increase  by  dO  while  it  remains 
saturated. 

From  equation  (d)  we  have  L\-  U^  =  JL~p(vl-v2).  If  c  denotes 
the  specific  heat  of  the  liquid,  and  if  k  is  a  constant,  we  may  set 
If  we  consider  v2  as  constant,  it  follows  that 


-p  . 

From  equations  (i)  and  (k)  we  then  have  dQ  =  (dL/d8-  L/6  +  c)d8. 
Designating  by  h  the  quantity  of  heat  which  must  be  used  in  raising 
the  temperature  of  the  vapour  by  1°C.  while  it  remains  saturated, 
we  have  (1)  h  =  dL/dO  -  L/6  +  c. 


SECTION  CXX.    THE  ENTROPY. 

The  methods  which  have  here  been  applied  to  the  discussion  of 
the  equilibrium  of  a  fluid  and  its  vapour  may  be  used  to  advantage 
in  many  other  cases,  especially  in  connection  with  chemical  problems. 
All  the  methods  are  based  on  the  equation  Jd$/0  =  0  for  a  cyclic 
process.  M.  Planck  has  given  general  formulas  by  which  treatment 
of  such  questions  is  much  facilitated.  The  bodies  whose  chemical 
equilibrium  is  to  be  investigated  are  contained  in  the  volume  V  at 
the  temperature  6,  and  are  subjected  to  an  external  pressure.  A 
change  in  the  chemical  composition,  or  in  the  proportions  of  the 
mixture,  is  accompanied  by  a  change  of  volume  </Fand  a  change  of 
temperature  dO,  and  at  the  same  time  the  quantity  of  heat  dQ  is 
received  from  surrounding  bodies.  If  S  denotes  the  entropy  and  V 
the  internal  energy  of  the  system,  we  have  (a)  dS=(dU +  P.dV)jO^ 

The  state  of  the  system  of  bodies  is  determined  by  the  pressure 
P,  the  temperature  6,  and  certain  other  variables  n,  n^  n.»  etc.  If, 
for  example,  the  space  contains  water  and  saturated  water  vapour 


SECT,  cxx.]  THE  ENTROPY.  293 

and  if  the  whole  mass  equals  M,  we  may  call  the  quantity  of  vapour 
Mn  and  the  quantity  of  liquid  Mnr  where  n  +  n1  =  l.  If  we  are 
dealing  with  a  case  of  dissociation,  we  may  use  n  for  the  number  of 
molecules  of  the  original  gas,  while  n^  and  n2  are  the  numbers  of 
the  dissociated  molecules.  Hence  the  state  of  a  system  depends 
generally  on  the  quantities  0,  P,  n,  nv  n2...  ,  and  we  have 


(b)  dV=  -3VIW  .  dO  +  -dVj-d 

[  dS  =  3S  pB  .dd  +  'dS  /oP  .dP  +  'dS  fdn  . 


From  the  definition  (a)  of  the  entropy  we  have 
l/e.(dU/W  +P. 


l/0.(dUfdF+  P  . 


since  0  and  P  are  independent.      It  follows  from  (a)  and  (b),  since 
6  and  P  do  not  depend  on  n,  nv  n.2...,  that 

'dS/'dn  . 


If  we  set  (c)  &  =  S-(U+PF)/Q,  this  equation  takes  the  form 

(d  )  3*/9n  .  dn  +  'd^fdnl  .  dn^  +  ?&l'dnz  .  dn,2  +  .  .  .  =  0. 

If  the  quantities  n,  nv  n.2...  are  independent  of  each  other,  we  have 

3^/d/i  =  'dSj'dn  -l/6.(d  Uj'dn  +  P  .  3F/3w) 

and  analogous  equations,  which  may  also  in  this  case  be  obtained 
directly  from  (a).  In  general,  there  will  be  some  relation  among 
the  quantities  n,  nv  n.^  ____  As  an  example  of  this  method,  we  will 
consider  the  problem  of  the  change  of  state.  If  a  quantity  of  vapour 
Mn  and  a  quantity  of  liquid  Mn^  are  enclosed  in  a  given  volume, 
we  have  as  above  n  +  itl  =  l.  Then  the  following  equations  hold 

S  =  Mns  +  Mn^,    U  =  Mnu  +  Mn^i^    V=  Mnv  +  Mn^, 
where  $,  it,  v  denote  the  entropy,  the  internal  energy  and  the  volume 
of  the   vapour   respectively,   while   sv  uv  and   i\   denote   the   same 
quantities  for  the  liquid.     From  equation  (c)  we  then  have 

<£  =  Mn(s  -  (u  +  Pv)/0)  +  Mnfa  -  (wa  +  PvJ/6) 
and  0  =  M(s  -  (u  +  Pv)/0)dn  +  ^/(^  -  (MJ  +  Pvl}l6)dnr 

In  addition,  we  have  dn  +  dn^  =  0.  Hence  equilibrium  exists  between 
the  vapour  and  the  liquid  if  (e)  sO  -  u  -Pv  =  s^  -  ul  -  Pvr  Since 
the  quantities  in  this  equation  depend  only  on  P  and  0,  it  may  take 
the  form  P=f(0).  Hence  the  equation  (e)  states  the  way  in  which 
the  pressure  of  the  saturated  vapour  depends  on  the  temperature. 


294  THERMODYNAMICS.  [CHAP.  xin. 

If  the  unit  mass  of  the  substance  is  transformed  from  liquid  to 
vapour  at  the  temperature  0,  the  entropy  increases  by  (f)  s-sl  =  JL/0. 
We  therefore  obtain  from  equation  (e)  JL  =  u-ul  +  P(v-vl),  as  in 
CXIX.  (d).  If  equation  (e)  is  differentiated  with  respect  to  0,  and 
P  considered  as  a  function  of  6,  and  if  we  use  the  equations 

=  1/0 .  (3tt/30  +  P .  30/30) ; 
=  1/0 .  (3MJ/30  +  P .  3rj/30) 

we  obtain  the  relation  J.L  =  (v-v1)6.'dP/W. 

We  may  consider  the  method  here  described  as  having  its  basis 
in  a  certain  tendency  in  nature.  Almost  all  natural  processes  are 
accompanied  by  the  development  of  heat ;  hence  energy  seems  to 
be  especially  inclined  to  assume  the  form  of  heat,  and  heat  tends 
to  pass  from  bodies  of  a  higher  to  those  of  a  lower  temperature. 
In  all  such  transformations  the  total  energy  remains  unchanged,  but 
it  loses  more  and  more  the  capacity  of  transforming  itself  into  kinetic 
energy.  A  body  which  contains  a  quantity  of  heat  Ql  and  whose 
temperature  is  0P  while  the  temperature  of  all  surrounding  bodies 
is  02,  may  yield  [CXII.  (e)]  the  kinetic  energy  ^(0a  -  02)/0r  Hence 
the  lower  01?  the  less  the  kinetic  energy  yielded,  if  02  remains 
constant. 

Clausius  expressed  this  principle  in  the  statement  that  the  entropy 
always  increases  and  tends  toward  a  maximum.  For  example,  if  a  quantity 
of  heat  Q  passes  by  conduction  or  radiation  from  a  body  at  the 
higher  temperature  6l  to  one  at  the  lower  temperature  02,  the  increase 
of  entropy  is  &S=Q/d2-Q/Or  The  entropy  remains  unchanged  only 
in  the  case  of  a  cycle,  in  which  the  bodies  receiving  the  heat  have 
the  same  temperature  as  those  giving  it  up,  and  in  which  the  whole 
system  is  in  neutral  equilibrium,  since  after  the  performance  of  this 
cycle  \dQ/6  =  0  or  A£=0.  This  is  also  the  case  at  any  instant 
during  the  cycle,  if  we  take  into  account  not  only  the  entropy  of 
the  working  body,  but  also  that  of  surrounding  bodies ;  if  the  first 
receives  the  quantity  of  heat  dQ,  the  second  gives  up  the  same 
quantity;  since  the  temperature  of  both  bodies  is  the  same,  the 
entropy  remains  unchanged.  This  holds,  however,  only  for  ideal 
cycles;  in  any  actual  movement  of  heat,  the  heat  passes  from  a 
higher  to  a  lower  temperature,  and  the  entropy  must  therefore 
increase. 

Hence  the  condition  for  a  change  in  the  state  of  a  system  is  an 
increase  of  the  entropy;  changes  in  which  the  entropy  diminishes 
are  impossible.  The  state  of  a  body  in  equilibrium  is  such  that  if 


SECT,  cxxi.]  DISSOCIATION.  995 

it  undergoes   a   small   change,   the   entropy   will   either  increase   or 
remain  constant. 

We  thus  return  to  the  conditions  of  equilibrium  (a).  By  the 
communication  of  the  quantity  of  heat  dQ  the  entropy  of  the  body 
considered  diminishes  by  dS;  hence  the  increase  of  the  total  entropy 
hdS-dQ/6,  and  this  must  be  equal  to  zero.  Since  dQ  =  dU+P.dV 
we  obtain  dS-l/0.(dU+P.dF)  =  0. 


SECTION  CXXI.    DISSOCIATION. 

If  a  compound  gas  is  separated  into  two  or  more  constituent  gases, 
either  by  heating  or  by  diminution  of  pressure,  it  is  said  to  be  dis- 
sociated ;  the  extent  of  the  dissociation  depends  on  the  pressure  and 
temperature.  In  order  to  determine  it,  we  must  determine  the 
function  (a)  3?  =  S  -  (  U  +PV}jQ.  If  n  denotes  the  number  of  molecules 
of  the  original  gas,  n1}  n2...  the  numbers  of  molecules  of  the  pro- 
ducts of  dissociation,  we  have,  as  an  expression  for  the  total  internal 
energy  U,  U=nu  +  nlul  +  n2u2  +  ...,  where  u,  uv  u2  denote  the  energies 
of  the  separate  molecules.  If  m  is  the  mass  of  a  molecule,  and  c, 
its  specific  heat  at  constant  volume,  we  may  set  the  internal  energy 
at  6°  C.  equal  to  Jmc,0  +  /t,  when  A  is  a  constant.  If  we  represent 
Jmc,  by  c,  we  have  (b)  U=n(cO  +  h)  +  nl(clQ  +  h1)  +  ....  If  v  denotes 
the  volume  of  a  molecule  of  a  gas  and  p  its  pressure,  we  have  pv  =  E6, 
where  R  does  not  depend  on  the  nature  of  the  gas.  For  the  sake 
of  simplicity  Planck  sets  R=\.  Since  the  gases  are  uniformly  dis- 
tributed throughout  the  whole  volume,  we  have  nv  =  nlv1  =  ...  =  T, 
and  therefore  pV=  n6,  plV=  r^O,  etc.,  (c)  PV=  (n  +  1^  +  n2+.  .  .)(?.  By 
this  equation  V  is  given  as  a  function  of  P,  0,  n,  nl.... 

The  entropy  of  the  system  is  equal  to  the  sum  of  the  entropies  of 
all  the  gases,  so  that  S=ns  +  nlsl  +  n2s2+  ...,  if  s  denotes  the  entropy 
of  the  molecule.  Using  c  with  the  meaning  given  above,  the  entropy 
of  a  molecule  equals  [CXI.  (d)]  c  .  log  0  +  log  v  +  k.  Here  nv  =  V,  and 
therefore  by  the  use  of  equation  (c), 

s  =  c  .  log  e  +  log(e/P  .  (n  +  n1  +  n2  +  .  ..)/»)  +  k. 
If  we  set 


n.2  +...),  etc.,    (7+  C^ 
we  have  (d) 


296  THERMODYNAMICS.  [CHAP.  xin. 

By  the  use  of  equations  (b),  (c),  and  (d),  (a)  takes  the  form 
n[(c  +  l)(log  8-I)-\ogP-logC  +  k-  h/ff] 


f  * 

I        +  njfa  +  l)(log  0  -  1)  -  log  P  -  log  C,  +  jfcj  -  V*]  +  .  .  - 


But  we  have  3(«.  logC+i^  AogC1  +  ...)/3/i  =  log<7,  and  further 
(f  )  cX£/cto  =  (c  +  1  )(log  e~l)-logP-logC+k-  h/e. 

Similar  expressions  hold  for  the  other  differential  coefficients.     Intro- 
ducing these  values  in  the  condition  of  equilibrium 

3*/3»  .  dn  +  t&rdn^  .  dn^  +  .  .  .  =  0, 

it  follows  that  the  problem  may  be  solved  if  we  know  the  relations 
existing  among  the  quantities  n,  nlt  n2  — 

If  we  investigate,  for  example,  the  dissociation  of  hydriodic  acid 
into  iodine  and  hydrogen,  the  proportions  of  the  gases  in  the 
mixture  can  be  represented  in  the  following  way  :  nJH,  n-^H^  ».*/.>. 
By  dissociation  two  molecules  of  hydriodic  acid  form  one  molecule 
of  hydrogen  and  one  of  iodine,  hence  the  ratio  between  dn,  dnv  dn.2 
is  -  2  :  1  :  1.  If  we  set  generally  dn  :  dn-^  :  dn.2  :  ...  =  v  :  vl  :  v2  :  .  .  .  ,  the 
condition  of  equilibrium  becomes 

(g)  v  .  oQ/dn  +  Vl  .  3*73^  +  v2  .  3*/3»3  +  .  .  .  =  0. 

From  equations  (f)  and  (g)  we  obtain 
(h)  2[v(c  +  l)(log  6  -  1)  -  v  log  P  -  v  log  C+  vk  -  vh/ff]  =  0. 

To  simplify  the  calculation  Planck  assumed  that  the  atomic  heat 
is  constant  even  in  the  compound  gas,  and  that  the  molecular  heat 
is  equal  to  the  sum  of  the  atomic  heats  ;  experiment  shows  that  this 
is  approximately  true  in  all  cases.  If  a,  a1?  a2  denote  the  number 
of  atoms  in  the  molecule  of  each  gas,  we  may  set 
C  =  ya,  Ct  =  yav  c.2  =  ya.2.... 

Since  the  whole  number  of  atoms  is  unchanged  by  dissociation,  the 
sum  na  +  nlal  +  n2a2+  ...  is  constant,  and  hence 

a.dn  +  al.  dn^  +  a2  .  dn.2  +  .  .  .  =  0. 

Consequently  also  va  +  v^  +  v2a2  4-  .  .  .  =  0   and   vc  +  vlcl  +  v.2c.2  +  .  .  .  =  0. 
Further  if  we  set 

v  +  vl  +  v.2  +  v3+  ...  =VQ;    vh  +  v1hl  +  vji2+  ...  = 


it  follows  from  (h)  that 


SECT,  cxxi.]  DISSOCIATION.  297 

Hence  for  hydriodic  acid  we  have 

v0  =  0,    and    ClCzICz  =  k0.h™. 

If  no  hydrogen  or  iodine  is  present  except  that  set  free  by  dis- 
sociation, we  have  nl  =  n2,  and  therefore  Cl  =  C.2  and  C1/C=-Jk0.h01'e. 
Now  (71/C'=n1M.  Hence  the  degree  of  dissociation  is  independent 
of  the  pressure,  but  increases  with  the  temperature.  In  this  case, 
however,  the  dissociation  can  never  become  complete,  since,  for  0  =  o>  , 
we  have  Cl/C=»Jk^. 

From  equation  (i)  the  pressure  has  no  influence  on  the  degree  of 
dissociation  if  the  total  volume  remains  unchanged  by  the  dissociation. 
This  is  the  case  when  v0  =  0.  If,  on  the  other  hand,  the  volume 
increases  during  the  dissociation,  any  increase  of  the  pressure  will 
lessen  the  degree  of  dissociation.  This  occurs  in  the  case  of  nitrogen- 
dioxide,  N004,  in  which  one  molecule  is  broken  up  by  the  dissociation 
into  two  molecules  NO2.  Hence  v=  -  1,  Vj  =  2,  and  therefore 

c^/c=k0.h^.e/p. 

This    equation,   together  with    Cl  +  C=\,  determines   the   degree   of 
dissociation. 

In  order  to  occasion  the  dissociation  determined  by  the  quantities 
dn,  dnv  dn2,  at  constant  temperature  and  at  constant  pressure,  the 
quantity  of  heat  dQ  is  required,  which  is  determined  by 


or,  from  equations  (b)  and  (c),  by  J  .  dQ  =  2(c0  +  h)dn  +  6  .  *2dn.     Since 

-ir  =  0,  the  quantity  of  heat  required  for  the  dissociation  determined 

by  the  quantities  v,  vv  v2...  is  determined  by 

(k)  J.Q  =  vh  +  v1hl  +  vJi2+...  +  v00=v08-logh0. 

We  reach  the  same  result  from  the  equation  J.dQ=B.dS  together 

with  the  relations  (d)  and  (h). 


CHAPTER  XIV. 

CONDUCTION   OF  HEAT. 
SECTION  CXXII.    FOURIER'S  EQUATION. 

IF  the  temperatures  of  the  different  parts  of  a  body  are  different, 
a  gradual  change  goes  on  until  the  temperature  of  all  parts  of  the 
body  is  the  same,  that  is,  until  equilibrium  of  temperature  has  been 
reached.  In  this  statement  it  is  assumed  that  the  body  neither 
receives  heat  from  surrounding  bodies  nor  gives  up  heat  to  them. 
The  rate  at  which  the  condition  of  equilibrium  is  reached  depends 
upon  the  facility  with  which  the  body  conducts  heat.  Without 
making  any  assumptions  on  the  nature  of  heat,  we  may  say  that 
heat  flows  in  a  body  until  a  state  of  equilibrium  is  reached.  We 
define  the  rate  of  flmv  of  heat  in  any  direction,  as  that  quantity  of 
heat  which  passes  in  unit  time  through  unit  area  perpendicular  to 
that  direction. 

Hence,  if  Q  represents  the  rate  of  flow  of  heat  through  an  area 
dS  within  the  body,  the  quantity  of  heat  which  will  pass  through 
that  area  in  the  time  dt  is  Q .  (IS .  dt.  If  Z7,  V,  W  are  the  components 
of  flow  in  the  directions  of  the  coordinate  axes,  the  quantities  of 
heat  which  pass  through  the  elementary  areas  dy .  dz,  dx .  dz,  dx .  dy, 
in  the  time  dt,  are  U.dy.dz.  dt,  V . dx . dz . dt,  and  W.dx.dy.dt 
respectively. 

Using  the  general  equations  (XIV.)  of  fluid  motion,  we  obtain 
for  Q  (a)  Q  =  lU+mV+nllf',  where  /,  m,  n  are  the  direction  cosines 
of  the  normal  to  the  elementary  area  dS. 

Let  00'  (Fig.  H2)  be  a  rectangular  parallelepiped,  whose  edges 
OA  =  a,  OB  =  b,  and  00  =  c  are  parallel  to  the  coordinate  axes.  If 
U,  V,  W  represent  the  components  of  flow  at  the  point  0,  those 
at  A  are  U+'dU/'dx .  a,  F+^Vj'dx.a,  W+'dWJ'dx.a  respectively, 
supposing  a,  b,  c  so  small  that  only  the  first  terms  in  the  expansion 


CHAP.  xiv.  SECT,  cxxii.]    FOUEIEE'S   EQUATION. 


299 


need  be  retained.  The  parallelepiped  receives  the  quantity  of  heat 
U.dt.bc  in  the  time  dt  through  the  surface  OB  AC,  and  loses  the 
quantity  (U+'dU/'dx.  a)bc.  dt,  which  flows  out  through  the  surface 
AC'O'B'  in  the  same  time.  The  parallelepiped  gains,  on  the  whole, 
the  quantity  -  9  U/'dx  .  a.bc.dt.  If  we  take  account  of  the  other 
surfaces,  the  quantity  of  heat  which  remains  in  the  parallelepiped 
is  -(Bl7/aa;+3F/3y+a#7a»).a6e.&J  or,  if  we  set  a.b.c  =  dv, 


-  (d  U/dx  +  "d  Vj-dy  +  3  Wfiz)dvdt. 

This  quantity  of  heat  raises  the  temperature  of  the  parallelepiped 
by  d6,  which,  if  c  denotes  the  specific  heat  of  the  body,  and  p  its 
density,  is  determined  by  the  following  equation, 

(b)  cp  .  dd  =  - 


This  equation  holds  only  if  the  heat  received  is  used  solely  in 
causing  change  of  temperature  and  does  not  produce  any  change 
in  the  state  of  aggregation  or  any  chemical  change.  Sometimes, 
too,  heat  exists  in  the  interior  of  a  body  which  has  not  penetrated 
into  it  in  the  form  of  heat,  but  is  produced  by  friction  or  by  an 
electrical  current  in  the  body ;  and  to  this  the  above  equation  does 
not  apply. 

The  components  of  flow  U,  V,  W  depend  on  the  distribution  of 
heat  in  the  body  and  on  the  nature  of  the  body.  If  the  body 
conducts  heat  equally  well  in  all  directions,  that  is  if  it  is  isotropic, 
we  may  determine  the  rate  of  flow  in  the  following  way.  Let 
A  and  B  be  two  points  within  the  body  infinitely  near  each  other, 
in  which  the  temperatures  are  respectively  9  and  &,  If  dv  denotes 
the  distance  between  the  points  A  and  B,  and  k  the  conductivity  of 
the  body  for  heat,  the  rate  of  flow  of  heat  in  the  direction  AB  is 


300  CONDUCTION  OF   HEAT.  [CHAP.  xiv. 

given  by  Q  =  k(0-  &)/dv.  Hence  the  condi.i<-tint>/  is  the  quantity  of 
heat  which  flows  in  unit  time  through  unit  area  of  a  surface  in  the 
body,  parallel  with  and  between  two  surfaces  whose  temperatures 
differ  by  1°,  and  which  are  distant  from  each  other  by  one  centimetre. 
Now  since  &  =  6  +  ddjdv . dv,  we  have  also  (c)  Q=  -k.dOldv.  We 
obtain  in  like  manner  for  the  components  of  flow  U,  V,  IV,  the 
expressions  (d)  U= -k.W/Vx,  V= -k.'dBj'dy,  fT= -k.'dO/'dz.  In 
actual  cases  the  conductivity  k  is  a  function  of  0  •  but  for  the  sake 
of  simplicity  we  will  assume  that  k  is  constant.  We  obtain  from 
(b)  and  (d) 

(e)  cp .  Wfit  =  k(d-epx2  +  320/9/  +  320/3z2). 

This  equation  was  first  given  by  Fourier,  and  is  therefore  called 
Fourier's  equation.  The  specific  heat  c,  the  density  p,  and  the  con- 
ductivity k  are  functions  of  6;  we  will,  however,  consider  them 
constant.  Fourier's  equation  may  also  take  the  form 

(f)  30/ctf  =  K2(920/9z2  +  320/3/  +  920/^2), 
where  (g)  Kz  =  k/cp. 

In  the  following  table  the  values  of  k  and  K  for  several  metals  at 
the  temperatures  0°  and  100°  C.  are  given  from  the  experiments  of 
L.  Lorenz : 


N> 

*m 

*0 

AT100 

Copper,     .     . 

0,7198 

0,7226 

0,909 

0,873 

Tin,      .     .     . 

0,1598 

0,1423 

0,392 

0,344 

Iron,    .     .     . 

0,1665 

0,1627 

0,202 

0,179 

Lead,  .     .     . 

0,0836 

0,0764 

0,242 

0,222. 

SECTION  CXXIII.     STEADY  STATE. 

The  state  of  the  body  with  respect  to  heat  is  called  steady,  if  the 
temperatures  of  the  different  parts  of  the  body  are  different,  but 
do  not  change  with  the  time.  In  this  case  each  particle  gives  up 
on  the  one  side  as  much  heat  as  it  receives  on  the  other,  and  the 
temperature  is  independent  of  the  time  t  and  dependent  only  on 
the  coordinates  x,  y,  z.  For  the  steady  state,  equation  CXXII.  (f) 
becomes  (a)  320/d.c2  +  320/3/  +  32^2  =  V2#  =  0.  The  components  of 
flow  are  expressed  by  equations  CXXII.  (d). 

Flow  of  lieat  in  a  phite. — We  will  consider  a  thin  plate  whose  faces 
L  and  M  are  parallel  to  the  yz-plane.  The  temperatures  of  the  faces 
are  respectively  6l  and  02.  This  being  so,  the  flow  of  heat  is  parallel 
to  the  .r-axis,  and  the  temperature  0  in  the  vicinity  of  the  .T-axis 


SECT,  cxxin.]  STEADY  STATE.  301 

depends  only  on  x,  so  that  from  (a)  we  have  d20jdx2  =  0.  Hence 
0=px  +  q.  If  the  distances  of  the  faces  L  and  M  from  the  y^-plane 
are  a  and  b  respectively,  we  have  01=pa  +  q,  02=pb  +  q,  and  further 

6  =  (bO}  -  a0z)f(b  -a)-  (61  -  82)x/(b  -  a). 

If  we  represent  the  distance  b  -  a  between  the  faces  by  e,  the  rate 
of  flow  of  heat  U  between  them  is  (c)  17=^-0^/6. 

Every  integral  of  equation  (a)  corresponds  to  a  steady  state  of  heat. 
If  8=f(x,  y,  z)  is  an  integral  of  (a),  01=f(x,  y,  z)  and  02=f(x,  y,  z) 
are  the  equations  of  two  surfaces  of  constant  temperatures,  or  of  two 
isothermal  surfaces,  where  Ol  and  02  are  constant.  If  the  body  is 
bounded  by  the  surfaces  which  are  determined  by  0l  and  6.2,  and 
if  6  is  a  temperature  which  lies  between  0l  and  02,  0=f(x,  y,  z)  is 
the  equation  of  any  isothermal  surface. 

The  flow  of  heat  in  a  sphere. — If  m  and  c  are  constant,  and  if 

r2  =  x2  +  y2  +  z2,    0  =  m/r  +  c 
is  a  solution  of  (a).     Therefore  setting  0l  =  m/rl  +  c,   09  =  m/r.2  +  c,  we 

>-^-^ 

as  the  equation  of  the  system  of  isothermal  surfaces,  which  in  this 
case  are  spheres.  For  the  rate  of  flow  of  heat  U  in  the  direction  r, 
we  have  (e)  U=  -k.  d0/dr  =  k(0l  -  02)rlr2/r2(r<i - 1\).  The  temperature 
and  flow  of  heat  in  a  hollow  sphere,  whose  internal  and  external 
surfaces  are  at  the  temperatures  0l  and  02  respectively,  are  also 
given  by  equations  (d)  and  (e).  The  total  quantity  of  heat  which 
flows  out  through  the  hollow  sphere  is  ^Trr2U=4irk(0l-  0.2)rlr2/(r2  -  ra). 
The  flow  of  heat  in  a  tube. — If  c  and  c  are  constants  and  if  r2  =  x2  +  y-, 
we  have,  from  XV.,  0  =  clogr  +  c'  as  an  integral  of  (a).  Therefore,  if 
we  set  6l  =  c.\ogrl  +  c',  02  =  c . log r.2  +  c',  we  obtain 
(f )  8  =  (0l-  02)log  /-/(log  r,  -  log  rz)  +  (^log  r2  -  02log  r1)/(log  rz  -  log  r,}. 
The  rate  of  flow  U  in  the  direction  r  is  U=k(8l- 82)/r(\ogr2-\ogrl). 
The  quantity  of  heat  which  flows  out  through  a  unit  length  of  the 
tube  is  (g)  2irrCT=2^(01-02)/(log»3-logr1). 


SECTION  CXXIV.     THE  PERIODIC  FLOW  OF  HEAT  IN  A  GIVEN 
DIRECTION. 

If  the  temperature  of  the  body  depends  only  on  one  coordinate, 
say  on  x,  Fourier's  equation  becomes  (a)  'dO/'dt  =  K2 .  'd20/'dx2.  We  will 
hereafter  investigate  in  what  way  this  equation  can  be  integrated. 


302  CONDUCTION  OF   HEAT.  [CHAP.  xiv. 

For  the  present  we  will  consider  special  integrals  which  correspond 
to  simple  yet  important  cases. 

The  temperature  of  the  earth  changes  during  the  year  ;  it  rises 
and  falls  with  the  temperature  of  the  air.  The  time  at  which  the 
maximum  or  minimum  temperature  at  any  point  is  reached  is  later 
as  the  point  lies  further  below  the  surface.  In  the  following  dis- 
cussion we  will  not  take  into  account  the  internal  heat  of  the  earth. 

If  the  temperature  at  the  earth's  surface  is  given  by  (b)  0  =  sin  at, 
we  may  express  the  temperature  at  a  point  in  the  interior  of  the 
earth  by  (c)  6  =  P  .  sin  at  +  Q  .  cos  at,  where  P  and  Q  are  functions  of 
the  distance  x  of  that  point  from  the  earth's  surface.  If  we  substitute 
for  9  in  (a)  the  expression  (c)  we  have 

Pa.  .  cos  at-Qa.  sin  at  =  K2(sin  at  .  d^P/^x2  +  cos  at  .  d^Q/dx2). 

Hence  we  must  have  *c2  .  d2P/dx2  =  -  Qa  and  K*.d2Q/dxz  =  Pa.  Now 
if  we  set  e2  =  a/*2,  we  have  (d),  (e),  d*P/dx*  =  -  e4P  and  Q  =  -  1/c2  .  d2P/dx?. 
In  order  to  integrate  equation  (d)  we  set  P  =  Aepx,  and  obtain 
p  =  f$J  '  -  1.  The  integral  of  equation  (d)  then  takes  the  form 


Since  0  must  equal  0  when  x  =  GO  ,  we  have  A=B  =  Q,  and  hence 


We  obtain  from  equation  (e) 

Q  =  (C£-1+V^*X1V*-  J)e(-l- 

But  from  equations  (b)  and  (c)  we  have  P  —  1  and  Q  =  0  when  x  =  0, 
and  therefore  C=D  =   .     Hence  we  obtain 


p  =  e  -  ex/v/r  ^  cos  (^2)  ;    Q=-e-  «/^  .  sin 
and  6  =  e-  «/^  .  sin  (at  -  ee,>/2). 

Substituting  for  e  its  value,  we  have 

6  =  e  -*•£/«  .  sin  (at  -  */£«/*). 

The  difference  between  the  highest  and  lowest  temperatures  at  the 
depth  x  below  the  surface  is  therefore  2e~xVWK.  This  difference 
depends  on  the  value  of  a.  The  faster  the  temperature  changes 
at  the  surface  the  smaller  the  influence  of  this  change  on  the  tem- 
perature in  the  interior.  For  example,  if  we  set  the  temperature 
at  the  surface  equal  to  0  =  8in(27r2/r),  the  difference  between  the 
highest  and  lowest  temperatures  is  equal  to  '2e-x^^lK,  and  this  is 
very  much  greater  when  T  is  a  year  than  when  it  is  a  day. 


SECT,  cxxv.]  A   HEATED  SURFACE.  303 

The  temperature  relations  within  the  earth  are  actually  different 
from  those  here  described,  because  the  temperature  at  the  surface 
cannot  be  expressed  in  any  simple  way.  The  main  features  of  the 
phenomena,  however,  are  similar  to  those  deduced  in  this  discussion. 


SECTION  CXXV.     A  HEATED  SURFACE. 

Let  the  temperature  in  an  infinite  body  at  the  time  t  =  0  be  every- 
where zero,  except  in  a  plane  in  which  each  unit  of  area  contains 
the  quantity  of  heat  o-.  Fourier  showed  that  at  the  time  t  the 
temperature  6  at  a  point  at  the  distance  x  from  the  heated  plane  is 
given  by 

~ 


where  k  is  the  conductivity  and  K  the  quantity  defined  in  CXXIL  (g). 
We  will  now  examine  whether  this  expression  for  6  satisfies  all  the 
conditions  of  the  problem.  We  will  first  consider  the  differential 
equation  W/dt  =  K2  .  320/3x2.  From  (a)  we  obtain 

(b),  (c), 
(d) 

It  follows  from  (b)  and  (d)  that  the  differential  equation  is  satisfied. 
Since  the  function  ze~^  approaches  zero  as  its  limit  if  z  becomes 
infinitely  great,  it  follows  that  for  /  =  0  we  have  0  =  0,  for  all  values 
of  a-,  with  the  exception  of  the  value  x  =  Q.  If  6  is  determined  by 
the  equation  (a),  we  can  further  show  that  each  unit  of  area  of  the 
heated  surface  S  contains  the  quantity  of  heat  cr  at  the  time  t  =  0. 
The  total  quantity  of  heat  which  is  present  in  the  body  is  given 
by  the  expression 


But  because  (e)  I    e~**dq  = 


the  quantity  of  heat  present  at  any  time  must  be  So-,  and  since  this 
quantity  is  present  on  the  infinite  surface  8  at  the  time  2  =  0,  the 
unit  of  surface  at  that  time  must  contain  the  quantity  o-. 

It  follows  from  (a)  that  6  =  0  for  t  =  0  as  well  as  for  t  =  <x>  ;  there- 
fore there  must  be  a  certain  time  at  which  6  is  a  maximum.  This 
time  is  found  from  (f)  0  =  0,  which  gives  t  =  xz/2K2.  The  corre- 
sponding value  of  0  is  (g)  6  =  \/\/2ire  .  a-fcpx.  It  appears  from  equation 


304  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

(a)  that  the  heat  is  propagated  with  an  infinite  velocity,  since  0  is 
everywhere  different  from  zero  as  soon  as  t  has  a  finite  value. 

We  will  now  determine  the  temperature  at  any  time  in  a  region 
in  which  the  original  distribution  of  heat  depends  only  on  one  of 
the  coordinates.  Let  0=f(a)  when  /  =  0,  where  a  is  the  distance 
from  the  y^-plane.  The  part  S  of  the  region  which  is  bounded  by 
two  parallel  planes  for  which  x  =  a  and  x  =  a  +  da,  contains  the  quantity 
of  heat  So-  =  S.da.  pc  ./(«•)•  Therefore  the  quantity  o-  which  is  present 
in  the  unit  of  area  of  this  sheet  is  a-  =  da  .  pc  ./(«). 

If  the  temperature  of  the  rest  of  the  region  is  zero,  the  heat  flows 
out  from  this  sheet  on  both  sides,  and  at  a  point  whose  distance 
from  the  f/2-plane  equals  x,  and  which  therefore  is  at  the  distance 
x-a  from  the  shell,  the  temperature  by  (a)  is 

,70  -JL 
^fr 

All  other  similar  sheets  emit  heat  according  to  the  same  law,  and 
we  therefore  have 


If  we  set  (i)  q  =  (a-x)/2K,Jt,  we  have 
6  =  1  /N/TT  .  [+ 


The  expressions  (h)  and  (i)  contain  the  complete  solution  of  the  problem 
before  us.  If  we  set  t  =  0  in  equation  (i)  and  make  use  of  (e),  it 
follows  at  once  that  6=f(x). 

For  example,  if  the  initial  temperature  is  constant  and  equal  to 
00  within  the  portion  determined  by  -I  <x<  +1,  but  equal  to  zero 
outside  these  limits,  the  integration  in  (h)  is  effected  between  these 
limits,  so  that 


SECTION  CXXVI.     THE  FLOW  OF  HEAT  FROM  A  POINT. 

Let  us  suppose  that,  at  the  time  1  =  0,  the  temperature  in  an 
infinitely  great  body  is  everywhere  equal  to  zero,  except  at  one 
point,  in  which  the  quantity  of  heat  m  is  concentrated.  We  will 
investigate  the  distribution  of  heat  in  the  body  at  any  subsequent 


SECT,  cxxvi.]  THE    FLOW  OF   HEAT   FROM   A   POINT.  305 

time  t.  This  problem  was  first  handled  by  Fourier,  who  found  that 
the  temperature  8  at  a  point  whose  distance  from  m  is  r  is  given  by 

(a)  0  =  mK2/k  .  (  IjlK^tY  .  e  ~  5*/4*\ 

We  can  show  that  this  expression  satisfies  all  the  conditions  of  the 
problem.  If  the  point  which  contains  the  quantity  of  heat  m  is  at 
the  origin  of  coordinates,  Fourier's  equation 

Wfdt  -  /c2(320/3a;2  +  320/3y2  +  320/as«) 

takes  the  form  (XV.)  (b)  30/3<  =  *2(320/3r2  +  2/r  .  30/3r),  because  0 
is  a  function  of  r  only.  This  equation  may  be  given  the  form 

(c)  -d(r6)pt  =  K*d*(rO)pr2. 
We  obtain  from  (a) 

(  3(r0)/3<  =  (  -  3/2t  +  r2/4jc2/2)  .  r0, 

(d)  J  9(r0)/3/-  =  (l/r-r/2/c20.r0, 

(  3*(r0)/3r3  =  (  -  3/2*2£  +  r2/4K4*2)  .  r0, 

and  by  the  use  of  these  values  prove  first  that  Fourier's  equation 
is  satisfied.  Further,  0  =  0  when  t  =  0.  The  quantity  of  heat  originally 
present  is  ///,  since  the  total  quantity  of  heat  at  any  time  is  given  by 


/  " 

Jo 


r2  .  dr  .  pcd  =      4*1*  .  dr  .  ™(l 

Jo 

If  we  set  q  =  r/'2i<^Jf,  the  integral  takes  the  form 


We  find  by  integration  by  parts,  and  by  the  use  of  CXXV.  (e),  that 
the  value  of  the  integral  is  m. 

The  time  /,  at  which  B  reaches  its  maximum  value,  is  obtained 
from  the  equation  0  =  0,  and  is  from  (d),  t  =  r2j6K2.  The  corresponding 
maximum  value  of  0  is  0=(l/\A|7re)3.'m/e/w3. 


SECTION  CXXVII.    THE  FLOW  OF  HEAT  IN  AN  INFINITELY 
EXTENDED  BODY. 

We  will  now  investigate,  with  the  aid  of  the  results  already 
obtained,  the  flow  of  heat  in  an  infinitely  extended  body,  when  the 
distribution  of  heat  at  a  particular  time  is  given.  Let  0=/(a,  b,  c) 
at  the  time  /  =  0,  where  a,  b,  c  are  the  coordinates  of  a  point  referred 
to  a  system  of  rectangular  axes.  The  quantity  of  heat  contained 
by  a  volume-element  da .  db .  dc  is  dm  =f(a,  b,  c) .  k/K2 .  dadbdc. 

u 


306  CONDUCTION   OF  HEAT.  [CHAP.  xiv. 

If  this  quantity  is  propagated   through  the  body,  it  produces  the 
rise  of  temperature  [CXXVI.  (a)], 

dO=(ll2K,JrtFe-Kx-<tf+<y-by+<*-eW*lft.f(a,  b,  c)dadbdc. 
If  we  take  the  sum  of  all  increments  of  temperature  which   arise 
from  the  distribution  of  heat  considered,  we  obtain  for  the  tempera- 
ture 0  at  the  point  x,  y,  z, 

(a)  6  =  (1/2/cv/^)3  f+"  f+*  [+g-[(*-«)2^-*>)2+(e-c)W2<  .y(a>  ^  c)dadbdc. 

J—  QO    J-oo    J-oo 

This  expression  for  B  is  an  integral  of  the  differential  equation 

(b)  Wf-dt  =  K2(320/3.r2  +  320/3/  +  320/3z2). 

We  notice  that  the  integration  of  this  equation  depends  on  that  of 
the  simpler  equation  (c)  'dX/'dt  =  K-32JT/3x2.  For  if  X  is  a  function 
of  x  and  tt  which  satisfies  equation  (c),  and  if  Y  and  Z  are  functions 
of  y,  t  and  z,  t  respectively,  which  satisfy  equations  analogous  to 

(c)  for  y  and  z,  the  equation  0  =  XYZ  satisfies  (b).     We  have 

YZX+  XZY+  XYZ=  K\YZ  .  92JT/3z2  +  XZ  .  ^Y/^f  +  XY.  ^Z/^). 
It  follows  from  (c)  and  the  analogous  equations   for  y  and  z  that 
this  equation  is  satisfied,  from   CXXV.   (a),  by  X=  l/Jt.  g-C*-«W, 
hence  the  expression 

\IJt.e-b-  «W*^  .Ijjt.e-to-  *>  W  .\IJt.e-to-  c^K'2f 
is  an  integral  of  equation  (b).     Therefore,  also, 

e  =  Cf/+"I+"T"{1^3  '  e-{(x-a?+(>J-b?+(*-c™KU  -/(a,  b,  c,)dadbdc 
is  an  integral  of  equation  (b).      C  is  a  constant,   and  /(a,  b,  c)  an 
arbitrary  function  of  a,  b,  c.     If  we  set 

a  =  (a-x)l-2K^t,   /3  =  (b-y)/2KJt,   y  =  (c-z)!'2K,Jt, 
it  follows  that 

0  =  (2K)3Cf+Xf^f^e  ~  at~^-^f(x  +  Znajt,  y  +  '2><PJt, 

z  +  2K-yJt)dad/3dy. 
If  we  now  assume  t  =  Q,  we  obtain  by  the  help  of  CXXV.  (e), 

e  =  (2K)*.C(j7r)*f(x,y,z). 

If  f(x,  y,  z)  is  an  expression  for  the  temperature  when  t  =  Q,  we 
set  C'=l/(2K>/7r)3,  and  obtain 


The  expressions  (a)  and  (d)  are  identical,  as  may  be  shown  by  the 
substitution  already  employed. 


SECT,  cxxvin.]  THE   FORMATION  OF  ICE.  307 


SECTION  CXXVIII.     THE  FORMATION  OF  ICE. 

Suppose  that  the  temperature  of  a  mass  of  water  is  everywhere 
0  =  0,  and  that  the  surface  of  the  mass  is  in  contact  with  another 
surface  whose  temperature  is  -  00.  60  may  be  either  constant  or 
variable,  but  must  be  always  below  zero.  A  sheet  of  ice  will  be 
formed  under  this  surface,  whose  thickness  e  is  a  function  of  the 
time  t.  The  temperature  Q  of  the  mass  of  ice  is  itself  a  function 
of  t  and  of  the  distance  x  from  the  surface.  For  x  =  e,  0  is  always 
equal  to  zero.  The  equation  (a)  Wfftt**K*&Ofdx*  holds  everywhere 
within  the  mass  of  ice.  New  ice  will  form  continually  on  the  bound- 
ing surface  of  the  ice  and  water.  The  quantity  of  heat  which  flows 
outward  through  unit  area  of  the  lowest  sheet  of  ice  is  given  by 
k~dd/'dx.dt.  During  the  same  time  a  sheet  of  ice,  whose  thickness 
is  de,  is  formed,  and  the  quantity  of  heat  set  free  thereby  is  Lpde, 
where  L  represents  the  heat  of  fusion  of  ice  arid  p  its  density.  When 
x  =  e,  we  have 


(b)  kWfdx  =  Lpd€/dt,   or 

We  may  write  for  6  the  expression 


As  may  easily  be  seen,  this  expression  satisfies  equation  (a).  It  also 
satisfies  the  condition  that  0  =  0  when  x  =  e.  In  order  to  find  whether 
it  satisfies  the  condition  contained  in  (b),  we  differentiate  (c)  with 
respect  to  x,  and  obtain 


When  x  =  e  this  becomes  equation  (b). 

Since,  at  the  surface,   6=  —  60,  it  follows  from  (c)  that 


If  the  thickness  of  the  sheet  of  ice  is  given  as  a  function  of  the 
time  t,  #0  may  be  easily  determined  ;  on  the  other  hand,  if  00  is 
given,  it  is  in  general  difficult  to  determine  e. 

*This  solution  was  communicated  to  the  author  by  L.  Lorenz.  See  also 
Stefan,  Wied.  Ann.,  Bd.  XLIL,  S.  269. 


308  CONDUCTION   OF   HEAT.  [CHAP.  xiv. 

If  00  is  constant,  the  right  side  of  equation  (d)  must  also  be  con- 
stant. This  condition  is  fulfilled  if  €2//c2  =  2p2t,  where  p  is  constant. 
From  (d)  we  then  obtain  the  equation 


which  serves  to  determine  p.     In  order  to  put  the  series  in  (e)  into 
a  finite  form,  we  form  from  (e) 


and  thus  obtain  d(cd0/Lp)/dp  =  1  +  cBJL.  The  integral  of  this  equation 
is  (f)  cejL^pTe-^-^^da.  If  the  thickness  of  the  sheet  of  ice 
increases  in  direct  ratio  with  the  time  t,  that  is,  if  €  =  JK/,  where  q 
is  a  new  constant,  it  follows  from  (d)  that 

rAft  n6fS 

^o/£  =  ^+fVr  2  3  +  "''  or  (g)  cOJL  =  f-\. 
If  e  is  very  small,  we  obtain  from  (d)  c60/L=  1/2/c2.  d(?/dt,  and 
hence  (h)  e2  =  2k/  Lp  .  f  60dt.  This  result  also  follows  if  we  set  the 
flow  of  heat  upward  equal  to  kdjt,  in  which,  however,  we  assume 
that  the  temperature  in  the  ice  increases  uniformly  from  its  upper 
surface  downward.  On  this  assumption,  the  quantity  of  heat  kO^.dtje. 
flows  upward  through  the  ice  in  the  time  (It.  In  the  same  time  a 
sheet  of  ice,  whose  thickness  is  dt,  is  formed,  and  the  quantity  of 
heat  Lp.de  is  set  free.  Hence  we  have  kOQ  .  dt/t  —  Lp  .  de.  This 
equation  leads  to  the  result  we  have  already  obtained.  If  0(}  is  con- 
stant, it  follows  that  (i)  f  =  j2 


SECTION  CXXIX.     THE  FLOW  OF  HEAT  IN  A  PLATE  WHOSE 
SURFACE  is  KEPT  AT  A  CONSTANT  TEMPERATURE. 

It  is  in  general  very  difficult  to  determine  the  variations  of  tem- 
perature in  a  limited  body.  We  will  discuss  a  few  cases  in  which 
it  is  possible  to  solve  this  problem.  Suppose  that  the  temperature 
in  the  interior  of  a  plate  bounded  by  parallel  plane  faces  is  6  =/(«), 
where  x  denotes  the  distance  of  the  point  considered  from  one  of 
the  faces  of  the  plate.  From  the  time  t  =  0  on,  the  surfaces  are 
supposed  to  be  in  contact  with  a  mixture  of  ice  and  water,  or  to 
be  so  conditioned  that  their  temperature  is  kept  at  zero.  The  law 


SECT,  cxxix.]      THE   FLOW  OF  HEAT   IN   A   PLATE.  309 

according  to  which  the  temperature  changes  in  the  interior  of  the 
plate  is  to  be  determined.  Designating  the  thickness  of  the  plate 
by  a,  we  have 

for  t  =  0,  0=f(x) ;   for  t  =  oo  ,   0  =  0; 
for  x  =  0,  0  =  0;       for  x  =  «,     0  =  0. 

The  rate  at  which  the  temperature  changes  at  the  surface  is  infinitely 
great;  just  outside  the  surface  it  is  equal  to  zero,  while  just  within 
it,  at  one  face,  it  is  equal  to  /(O).  At  the  other  face  the  temperature 
outside  the  plate  is  also  zero,  and  within  it  f(a).  The  function  0 
must  satisfy  not  only  these  conditions,  but  also  the  differential  equa- 
tion (b)  'dd/'dt  =  K2320/2te2.  An  integral  of  this  equation  is 

(c)  6  =  e~ m'K2t(A  sinmx  + Bcosrnx). 

From  (a)  B  =  0,  so  that  (d)  0  =  Ae~ mVt  sin  mx.  This  value  of  0  satisfies 
not  only  equation  (b),  but  also  vanishes  for  x  =  0.  Since  0  is  also 
zero  when  x  =  a,  we  must  have  sin  ma  =  0,  and  therefore  ma=  ±pir, 
where  p  is  a  whole  number.  Hence  we  have 

(e)  0  =  Ae-#***«*l<* .  sin  (pxirja). 

If  we  notice  further  that  6=f(x)  when  ^  =  0,  we  have 

(f )  f(x)  =  A  sin  (pTTX/a). 

In  general,  the  function /(x)  can  not  be  represented  by  this  expres- 
sion. To  solve  the  problem  we  use  the  following  method. 

Since  the  expression  (e)  is  an  integral  of  Fourier's  equation,  the 
complete  integral  is  obtained  by  taking  the  sum  of  the  similar  expres- 
sions, which  are  obtained  by  giving  p  all  values  between  1  and  oo . 
The  terms  which  correspond  to  a  negative  value  of  p  differ  from 
those  terms  for  which  p  is  positive  only  in  sign,  and  can  therefore 
be  considered  as  contained  in  the  latter.  Hence  we  set 

(g)  8  =  Al  sin  fax/a) .  e  -  **&!«*  +  A2  sin  (2vx/a) .  e  ~  2ViA/a*  +    _ 
When  t  =  Q,  we  have  0  =/(«),  so  that  for  0<x<a 

(h)  f(x)  =  Al  $in(Trx/a)  +  A2 sin (2Trx/a)  +  ... . 

We  will  now  investigate  whether  a  function  f(x),  which  is  arbitrary 
within  the  given  limits,  can  be  represented  by  a  trigonometrical 
series  of  this  form.  For  this  purpose  we  choose  instead  of  the 
infinite  series  (h)  another  series  with  (n -  1)  coefficients  Av  A^...  AH_V 
which  coincides  with /(x)  at  (n-})  points,  namely,  at  the  points 

x  —  a/n ,   x  =  2ajn,    ...    x  =  (n-l )a/n. 


310  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

We  then  obtain  the  following  n  -  1  equations  : 


f((n  -  l)a/n)  =  Al  sin((w  -  I)TT/TI)  +  A2  sin((n  -  1)2*»  +  .  .. 
+  An_lsin((n-  l)(n  -!)«•/»). 

If  we  multiply  the  first  of  these  equations  by  sin(ir/»),  the  second 
by  sin(27r/?i),  etc.,  and  add  the  right  and  left  sides  of  the  equations 
thus  obtained,  we  have 

/»)  .  sin(27i»  +  ...  +f((n  -  l)a/»)sin((w  -  l)ir/w) 
/w)  +  .  .  .  +  sin2((n  -  l)w/»)] 


+  sin  ((n  -  1  )w/»)  sin  ((TO  -  1  )2ir/»)]  +  .  .  .  . 
Now 

sin2a  +  sin22a  +  .  .  .  +  sin2((?i  -  1  )a) 

=  jf"n-l  -(cos2a  +  cos4a+  ...  +  cos((2/i-2)a))1. 
But  because 


we  have 

sin2a  +  sin22a  +  ...  +sin2((%-  l)a)  =  |[w-  1—  cos%a.  sin(%-  l)a/sina]. 

Substituting  for  a  its  value  irjn,  we  have 

sin2(7r/-n)  +  sin2(27r»  +  .  .  .  sin2((n  -  1  )ir/n)  =  w/2. 
Now  we  can  give  the  factor  of  A2  the  form 

|[cos(7r/?i)  -  cos(37r/n)  +  cos(27r/n)  -  cos(6ir/»)... 
+  cos((n-  l)w/n)  -  cos((n  -  1)3»>)]. 

Applying  the  above  given  summation  formula,  we  find  that  this 
factor  is  equal  to  zero.  In  the  same  way  the  factors  of  A3,  Av 
etc.,  vanish,  and  we  obtain  finally 

Al  =  2/n.  \_f(a/n)  sin  (77/71)  +/(2a/n)  .  sin  (  2ir/n)  +  .  .  . 

+f((n-  l)a/7i)sin((n-  l)»r/»)]. 
In  general  we  have,  for  0<m<w, 

j  Am=  2/n  .  \_f(a/n)  .  sin(mir/n)  +/(2a/»)  .  sin(2m7r/w) 
1  +...+/((»-!  )o/n)  .  sin  (  (»  -  1  )wir/»)]. 


SKCT.  cxxix.]      THE   FLOW   OF   HEAT  IN   A  PLATE.  311 

Hence  it  is  possible  to  so  determine  the  coefficients  Av  Av  ... 
that  f(x)  and  the  trigonometrical  series  coincide  for  (n  -  1)  values 
of  x  between  0  and  a.  The  greater  the  value  of  n,  the  more  values 
will  the  two  functions  have  in  common,  and  when  n  =  GO  ,  one  function 
may  be  replaced  by  the  other  between  the  limits  considered.  The  two 
functions  are  not,  however,  necessarily  identical,  for  their  differential 
coefficients  may  be  entirely  different.  One  of  them  is  related  to 
the  other  in  the  same  way  as  a  straight  line  to  a  zig-zag  line, 
whose  irregularities  are  infinitely  small. 

We  will  now  assume  that  ?i=oo   and  write  the  equation  (i), 

Am  =  2/7T  .  Tr/n  .  [f(a/n)  .  sin  (rmr/n  )  +f(2a/n)  sin 

+f(rajri)  sin  (rimr/n)  +  ...]. 

Setting  r-rr/n  =  y,  ir/n  =  dy,  rafn  =  ay  ITT,  it  follows  that 
(k)  Am  =  2/7T  .  f/(ay/7r)  .  sm(my)dy. 

-0 

Further,  if  we  set  x  =  ay/ir,  we  have 

(1)  Am  =  21  a  .    /fa)  sin  (rmrxla)dx. 


The  same  result  is  obtained  in  another  way  in  XXXVII.  (c). 

Therefore,  within  given  limits,  we  can  replace  the  function  f(x]  by 
a  trigonometrical  series,  and  set,  for  0<£<a, 

f(x)  =  2/a  .  [sin(n-x/a)  . 

+  sin(27ra;/a). 
o 

Introducing  the  values  for  Av  A2...  contained  in  (g),  the  problem 
is  solved,  and  we  obtain 


[  \aQ  =  sin  (iTx^e-****"*  .  [  /(or)  sin  (irx/a)dx 
(n) 


For  example,  if  the  initial  temperature  of  the  plate   is   constant 
and  equal  to  00,  we  have 

I  60sin(nnrx/a)dx  =  (l  -  cos  mrfaOJmir, 
and  therefore 

(o)         \TrO  =  60 .  siu(7rxla)e-(*Kla}*-t  +  £00 .  sin(37ne/a) .  g- <"*/•)*•'+  .... 
When  t  =  0,  we  obtain,  for  0<x<a, 
(p)  ^TT  =  sin (irxja)  +  ^ .  sin (3irx/a)  +  ^ .  sin (5irx/a)  +  ... 


312  CONDUCTION  OF   HEAT.  [CHAP.  xiv. 

SECTION  CXXX.    THE  DEVELOPMENT  OF  FUNCTIONS  IN  SERIES  OF 
SINES  AND  COSINES. 

As  shown  in  the  foregoing  paragraph,  we  may  always  set 

(a)  f(x)  =  Alsin(Trx/a)  +  A2sin('27rx/a)  +  ... , 
in  which  [CXXIX.  (1)] 

(b)  Am  =  2!a .  I  /(a)  sin  (mira/a)da, 

.'o 

where  x  is  replaced  by  a.  This  development  holds  only  for  0  <  x  <  a ; 
it  does  not  hold  for  the  limits  0  and  a,  except  when  f(x)  itself  is 
equal  to  zero  for  these  limits.  The  right  side  of  (a)  is  an  odd  function, 
which  changes  its  sign  with  .r.  The  series  (a)  holds  then  within 
the  limits  -a<a;<0,  when/(z)  is  also  odd.  Setting /(.r)  =  .r,  we  have 

Am  =  2/a .  I  a  sin  (iwra/a)da  =  —  2a  cos  (mir)fmir, 
and  further 

(c)  | .  irx/a  =  sin (irxja)  -  i  .  sin  (2irx/a)  +  J .  sin  (3-.r/«) 

Since  x  is  an  odd  function,  the  series  holds  for  negative  values  of  x 
if  it  holds  for  positive  values.  Further,  since  the  series  holds  for 
x=Q,  it  holds  within  the  limits  -a<x<  +  a.  Setting  Trxja  =  y,  we 
have  for  —ir<y<  +  TT, 

(d)  \y  =  sin y -  £ .  sin  2y  +  J .  sin  3y -  .... 

Further,  if  we  set  (e)  f(x)  =  E0  +  Bl cos (irx/a)  +  B.2 cos (2irar/« )+..., 
multiply  both  sides  of  this  equation  by  cos(rmrx/a),  and  then  integrate 
from  0  to  a,  it  follows,  if  in  and  n  are  whole  numbers,  that 

[  cos  (mirx/a) cos (mrx/a)dx  =  0   and  j  eos2(nnrx/a)dx  =  \a. 
Hence  for  m>0  we  obtain 
(f )  Bm  =  2/a .  I  af(x) .  cos  (mirxla)dx  and  £0=l/a.  ["/(zjdx. 

We  obtain  BQ  by  multiplying  both  sides  of  (e)  by  dx  and  integrating 
from  0  to  a.  If  f(x)  is  an  even  function,  the  series  holds  within  the 
limits  -a<x<a,  since  the  cosine  series  does  not  change  its  sign 
with  x.  But  if  /(.r)  is  an  odd  function,  the  series  (e)  holds  only 
within  the  limits  0  and  a. 

We  therefore  obtain  the  result  (g) 

\a.f(x)  =  sin(«/a)  f/(a)sin(;ra/rt)r/a  +  sin(27r.c/fl)  ^  /(a)sin(27ra/rt)f/a  +  . . . , 


SECT,  cxxx.]      THE   DEVELOPMENT  OF   FUNCTIONS.  313 

.f(x)  =  J .  I  f(a.)da  +  cos(Trx/a) .  |/(a)cos(?ra/a)dfa 


(h)         , 

+  cos(2;r:c/a)  |  f(a)cos(2Tra/a)da 

An  arbitrary  function  f(x)  can  also  be  developed  in  a  series  of 
sines  and  cosines,  so  that  the  development  holds  within  the  limits 
-a<x<a.  To  effect  this,  we  set 

f(x)  =  i  .  [f(x]  +f(  -  x)]  +  \  .  [f(x]  -/(  -  x)], 

in  which  %[/(%)+  f(  -x)~]  is  an  even  function,  because  it  remains  un- 
changed when  x  is  replaced  by  -x.  This  function  §[f(x)+f(  -x)] 
can  therefore  be  represented  by  a  cosine  series.  The  coefficient  of 
cos(mirx/a")  is 


=  I  •  /  f(a)cos(mira/a)da  +  £  .  /  /(  -  a)  cos  (nnra/a)da. 
If  in  the  last  integral  a  is  replaced  by  -  a,  the  integral  is  transformed 
into    -  i  •  /    /(a)cos(w7ra/rt)rfa,    and    the    coefficient    sought    becomes 

i/  /(a)cos(w7ra/a)e?a.     Hence  we  obtain 

(  !«  •  [/(*)  +/(  -«)]  =  £•  f  +/(«)^«  +  cos(^/a)  f+f(a)CO&(ira/a)da 

/•\          I  -'-a  J-a 

<J)   1  r+« 

+  cos(27rx/a)  /  /(a)cos(2iro/a)rfa+  .... 

On  the  other  hand,  the  function  J  .  [/(x)  -/(  -  x)]  is  an  odd  function. 
because  it  changes  its  sign  with  x;  therefore,  by  using  (g),  we  can 
represent  this  function  by  a  sine  series.  The  coefficient  of  sin(imrx/a) 
is 

2  •  /  l/(a)  -/(  -  a)]sin(w7ra/a)da 

=  ^  .  I  /(a)  sin  (rmra/a)da  -  |  .  j  f(  -  a)sm(mira/a)da. 
If  we   replace    —  a   by  a   in    the    last   integral,   it   is   transformed 
into  -i.  I  f(a)sin(mTra/a)da,  and  the  coefficient  becomes 

r  +  i 

J.j  f(a)sm(nnrala)da. 
We  therefore  obtain 

f  i«  •  [/(*)  -/(  -  *)]  -  sin(7r.r/«)  .  f  /(<z)sm(7ra/«ya 

<k) 

+  sin  (2mc/a)  /  /(a)  sin  (2ira/a)da  +  .... 


314  CONDUCTION   OF   HEAT.  [CHAP.  xiv. 

By  the  addition  of  equations  (i)  and  (k)  we  obtain  finally 
(  a  .f(x)  =  1  .   i  +f(a)da  +  (+f(a)CosU(x  -  a^da 

0) 

[  +fj(a)COS(2Tr(x-«)/a)da+..., 

or,  for  -a<z<a, 

(m)    f(x)  =  1  /a  .  £**[£  +  cos  (ir(x  -  a)  /a)  +  cos  (  -lir(x  -  a)/a)  +  .  .  .]/(a)da. 
This  series  is  due  to  Fourier.     It  may  also  be  expressed  in  the  form 
(n)         f(x)  =  l/2a.+f(a)da+l/ 


We  may  now  ascribe  any  value  to  a.  If  a  is  infinitely  great,  and 
if  /  f(a)da.  is  finite,  the  first  term  on  the  right  side  of  the  equation 
(n)  vanishes,  and  we  obtain 


f(x)  =  1/ir  .  2>/o  •     /(a)cos(m7r(2  -  a)/aVo. 

m=l  J-a 

Now  setting  wi7r/a  =  A,  and  therefore  7r/a  =  dX,  it  follows  that 
(o)  /(«)  =  I/TT  .  jf  dx|_+/(a)cos(  A(a;  -  a))^, 

where  -co  <  a;  <  ao  .  Instead  of  this  equation  we  may  often  use 
one  of  the  two  which  are  obtained  from  (g)  and  (h).  The  general 
term  in  (g)  is 

sin  (mirx/a)  .  f  /(a)  sin  (mira/a)da, 
and  hence 

f(x)  =  2/7T  .  IT  I  a  .  ^  sin  (mirx/a)  I  /(a)  sin  (mTra/a)da. 
Now  if  we  set  mir/a  =  A,  and  therefore  via  =  d\,  we  have  for  0  <  x  <  oo  , 
(P)  /(*)  =  2/7T  .  l"d\.  .  sm(Xx)j  /(a)sin(AaXa. 

From  (h)  we  obtain  in  the  same  way  for  0  <  x  <  oo  , 
(q)  /(*)  =  2/7T  .  {™d\  cos(Aa;)  .  [7(a)cos(Aa)^a. 


SECT,  cxxxi.]  APPLICATION  OF  FOURIEK'S  THEOREM.  315 

SECTION  CXXXI.     THE  APPLICATION  OF  FOURIER'S  THEOREM 
TO  THE  CONDUCTION  or  HEAT. 

If  the  temperature  in  a  certain  region  depends  only  on  the  x-co- 
ordinate,  the  temperature  6  must  satisfy  Fourier's  equation 

(a)  30/3*  =  K2320/3a;2. 

From  CXXIX.  (c)  6  =  e~x*K2:(A  sin  Xx  +  Bcoskx)  is  an  integral  of 
equation  (a)  where  A,  A,  B  are  constants.  We  may  also  give  the 
expression  for  9  the  form  6  =  e~^K2t  cos  (\(x  -  a))  .  /(a),  where  /(a) 
is  an  arbitrary  function  of  a,  and  A.  and  a  are  constants,  which  may 
take  all  possible  values.  Any  sum  of  such  terms  satisfies  the  equation, 
and  as  the  integral 

(b)  0  =  I/TT  .  J^Ae-^'[+7(a)cos(A(z  -  a))da 

is  such  a  sum,  it  will  also  satisfy  the  equation.  But  when  /  =  0, 
we  have 

6>=1/7T.  {°°d\f*f(a)cos(\(x-a))da, 

from  which,  by  comparison  with  CXXX.  (o),  we  obtain  6=f(x). 

The  formula  (b)  contains  the  solution  of  the  problem,  to  determine 
the  temperature  in  a  body  at  any  time  t,  when  the  temperature  is 
given,  at  the  time  £  =  0,  by  6=f(x).  This  problem  has  already  been 
solved  in  another  way  in  CXXV.  (h)  and  (i).  We  proceed  to  show 
that  the  solution  hero  given  is  identical  with  the  former  one. 

Since  (c)        0  =  1  /«•  .  f+J(a)da  /"  V  ***  cos  (  X(x  -  a))d\, 
we  first  determine  the  value  of  the  integral 


If  we  develop  cos  (A(z-  a))  in  a  series,  this  integral  is  represented  by 


It  follows  by  integration  by  parts  that 
/  VXU%A*V*A.  =  (2n  -  l)/2*2£ 
and  by  continued  reduction 


316  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

But  because  j  e~fdq  =  \,Jirt 


we  obtain 


The  value  of  the  integral  sought  is  therefore 


or  «-«- 

2*^ 

If  we  replace  the  integral  in  (c)  by  this  value,  we  obtain 


This  expression  for  6  is  identical  with  that  given  in  CXXV.  (h). 

We  will  now  apply  Fourier's  theorem  to  find  the  law  of  penetra- 
tion of  heat  into  a  body.  For  this  purpose  we  will  consider  the 
simple  case  in  which  the  body  is  in  contact  over  a  plane  bounding 
surface  F  with  another  body  whose  temperature  00  is  constant  and 
given.  Let  the  original  temperature  of  the  cold  body  be  zero. 

If  we  proceed  as  before  and  use  CXXX.  (p),  we  obtain 

B  =  00  +  2/?r  .  I  "dX  sin  (  Aa-)<rX2K2<  .  /"/(a)  sin  (  Aa)da. 

This  expression  for  6  satisfies  all  conditions  if  only  we  have,  when  t  =  0, 
0  =  00  +  2/7r.  f°<Usin(Aa:).  [  /(a)sin(Xo)da. 

-0  -0 

This  condition  is  fulfilled  [CXXX.  (p)]  when  /(a)  =  -  00.     Hence  the 
solution  of  the  proposed  problem  is  contained  in 

(e)  6  =  00  -  200/ir  .  f"d\  sin  (  \x)e~x^'  .'  J'sin  (\a)da. 

Using  the  same  method  of  reduction  as  that  by  which  (c)  is  trans- 
formed into  (d),  we  obtain 


from  which  B  =  00  -  -^  •  -!  (V*W?  -  l"e  "  *2dq  }  , 

VTT     lJ-*tevi        Jxftni/t 

and  therefore  6  =  00  {  1  -  1  /Jir  .  f^e^'dq  }  . 

J-XpKVt 


SECT,  cxxxr.]  APPLICATION  OF  FOURIER'S  THEOREM.  317 

Since  e~q~  is  an  even  function,  we  obtain 


or  by  the  help  of  CXXV.  (e), 
(f)  e 


Let  A  and  B  be  two  points  within  the  body,  whose  distances  from 
the  surface  F  are  xl  and  x2  respectively.  The  temperature  &  which 

A  attains  after  the  lapse  of  the  time  /,  is  &  =  '2BJJir  .  I  e~g2dq.     B  attains 

Jr^-'KVti 

the  same  temperature  after  the  lapse  of  the  time  /.„  given  by  the 

f  " 
equation  0'  =  20JJir  .  I  e~g2dq.     Comparing  the  two  integrals,  it  appears 

/VbV«1 

that  Xil*Jtl  =  xzl,Jtt>  or  t.2/tl  =  x22/xl2,  that  is,  the  times  required  for  two 
points  to  attain  the  same  temperature  are  proportional  to  the  squares  of 
the.  distances  of  the  points  from  the  heated  surface  F. 

AVe  will  now  determine  the  quantity  of  heat  which  flows  into 
the  cooler  body  through  unit  area  in  unit  time.  For  this  purpose 
we  give  the  equation  (f)  the  form  0  =  20J,jTr.  [f(ao  )  -f(x/2K,Jt)],  from 
which  follows,  by  (f),  -  kW/'dx  =  kOJitj^t  .  e~zt:^.  Setting  x  =  0,  we 
find  the  quantity  of  heat  U  desired,  (g)  U  =  k60/  K-Jiri. 

By  the  help  of  equation  (g)  we  may  solve  an  important  problem. 
Two  bodies  L  arid  L'  are  in  contact  over  a  plane  surface,  the  tem- 
perature of  one  of  these  bodies  being  T,  that  of  the  other  T.  If 
the  two  bodies  are  brought  in  contact,  one  of  them  is  heated  and 
the  other  is  cooled.  We  can  also  determine  the  temperature  2\ 
of  the  surface  of  contact.  Assuming  that  T0  is  constant,  the  quantity 
of  heat  which  L  receives  in  unit  time  is,  from  (g),  given  by 

U=k(T0~T)lKj*t, 

In  the  same  time,  L'  receives  the  quantity  of  heat 
U'=k'(T,-T')/K'^t, 

where  k'  and  K'  have  the  same  meaning  for  L'  as  k  and  K  for  L. 
But  since  the  infinitely  thin  bounding  surface  can  contain  no  heat, 
U+U'  must  equal  zero,  or  k/n  .  (T0-  T)  =  k'/K  .  (T  -  T0),  from  which 
follows  (h)  *  T0  =  (Tjkcp  +  T*JJMji)l(Jkcp  +  */kVfi).  It  is  thus  shown 
that  the  assumption  is  correct,  that  the  temperature  in  the  bounding 
surface  between  two  bodies  which  meet  in  a  plane  surface  is  constant. 
Strictly  speaking,  the  bodies  in  contact  must  both  be  infinitely  large, 

*  L.  Lorenz,  Lehre  von  der   Warme.  S.   178.      Kopenhagen,   1877. 


318  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

but  the  formula  (h)  may  also  be  applied  to  small  bodies  if  we  only 
consider  them  shortly  after  they  are  brought  in  contact.  We  may 
show  from  (h)  that  the  temperature  of  a  heated  solid  is  very  little 
diminished  by  contact  with  the  air  ;  this  holds  for  the  metals  and 
for  good  conductors  in  general.  It  follows  from  equation  (h)  that 


If  T  represents  the  temperature  of  the  solid,  p  is  always  very  much 
greater  than  p.  Hence  T0-T  is  very  much  greater  than  T  -  T0, 
especially  since  k  is  also  greater  than  k',  while  c  and  c'  are  not  very 
different  from  each  other. 


SECTION  CXXXII.     THE  COOLING  OF  A  SPHERE. 

Let  us  suppose  that  the  temperature  at  a  point  in  the  interior 
of  a  sphere  depends  only  on  the  distance  of  that  point  from  the 
centre  of  the  sphere.  In  this  case  [CXXVI.  (c)]  Fourier's  equation 
takes  the  form  (a)  3(r0)/a*  =  K232(r0)/3r2.  If  m,  A,  B  are  arbitrary 
constants,  an  integral  of  equation  (a)  is 

rO  =  e-"™  .  (A  sin  (mr)  +  B  cos  (mr)). 

But  since  this  equation  leads  to  the  conclusion  that  0  =  oo  when  r  =  0, 
B  must  equal  zero,  and  we  obtain  as  the  integral  of  equation  (a) 

<b)  r0  =  A.e-a**.Bm(mr). 

We  will  first  consider  the  case  of  the  sphere  immersed  in  a  mixture 
of  ice  and  water,  or  so  situated  that  its  surface  is  kept  at  the  tem- 
perature 0°  by  any  means.  If  we  represent  the  radius  of  the  sphere 
by  R,  we  have  0  =  0  when  r  =  E,  and  therefore  sin(w7?)  =  0.  Hence, 
if  p  is  an  arbitrary  whole  number,  we  must  have  mli  =  ±p-.  We 
•can  now  set 


The  constants  Av  A2...are  determined  by  the  help  of  the  tem- 
peratures of  the  different  parts  of  the  sphere  at  the  time  t  =  0.  Let 
these  temperatures  be  given  by/(r).  We  then  have 

r  .f(r)  =  A^  sin  (lerfR)  +  A2  sin  (2wr/.B)  +  .  .  .  . 
From  CXXIX.  (1)  we  obtain  for  Am 

<d)  Am  =  2/fi.     'rf(r)aiu(mirrjE)dr. 


SECT,  cxxxn.]  THE  COOLING  OF  A   SPHERE.  319 

If  the  temperature  is  constant  and  equal  to  #0  at  the  time  t  =  0,  we 
have 

R 

rsin(mirr/R)dr,  aud  hence  Am=  -  200R/mir  .  cos  (rmr). 


Using  these  values  we  obtain  finally 
I  6=2R60/irr.  [sin(7rr/R)  .  e-< 
\  -  i  .  sin  (27rr/R)e-(*"«™  +  £  .  sin  (STr 


The  mean  temperature  &  is 

(f  )  &  =  600/,r2  .  {«-<«/*)*  +  |  . 

This  equation  may  be  applied  to  a  thermometer  which  is  immersed 
in  a  fluid  cooler  than  itself.  The  temperature  of  the  thermometer 
is  then  given,  to  a  close  approximation,  by  the  first  term  of  the  above 
equation.  The  rate  of  cooling  is  (g)  -  dO'/dt  =  Tr2kO'/cpR2. 

We  will  now  consider  another  important  case,  that  of  a  sphere 
in  vacuo  losing  heat  by  radiation.  We  suppose  the  temperature  of 
the  region,  or  rather  of  its  boundary,  to  be  0°.  We  suppose  the 
radiation  to  take  place  according  to  Newton's  law,  and  therefore  to 
be  proportional  to  the  temperature  on  the  surface  of  the  sphere. 
From  (b)  the  integral  takes  the  form  (h)  rO  —  ^Ame~mt>ft  .  sin  (mr). 
If  E  represents  the  coefficient  of  radiation,  the  quantity  of  heat 
which  radiates  in  unit  time  from  an  element  dS  of  the  surface  is 
dS.  E6.  We  will  assume  that  E  is  constant.  The  same  surface- 
element  dS  receives  from  the  interior  of  the  sphere  in  the  same  time 
the  quantity  of  heat  -k.dS.dO/dr.  Since  the  quantity  of  heat 
which  dS  receives  must  be  equal  to  that  which  it  emits,  we  have 
(i)  -k.dO/dr  =  E6  or  -d6jdr  =  h6,  where,  for  brevity,  we  set  h  =  E/k. 
Hence,  for  r  =  R, 

2^TO<rmV2<  .  (m  cos  (mR)/R  -  sin  (mR)/R?)  =  -  KSAne~aAft  sin  (mR)/R, 
or  ^Ame-m^f  .  [mRcos(mR)  -  (1  -  hR)sm(mR)]  =  0. 

If  this  equation  is  to  hold  for  every  value  of  t,  we  must  have 
(k)  mR  .  cos  (mR)  =  (l-hR).  sin  (mR). 

This  equation  must  be  solved  for  m.  We  set  mR  =  x  and  obtain 
(1)  tg  z  =  z/(l  -hR).  If  we  further  set  yl  =  tgx,  yl  and  x  may  be  con- 
sidered as  the  rectangular  coordinates  of  a  curve  (Fig.  143).  This 
curve  has  an  infinite  number  of  branches,  oa,  irb,  27rc...,  to  which 
the  straight  lines  X  =  ^TT,  x  =  %ir...  are  asymptotes.  Further,  if  we 
set  i/.2  =  x/  (I  -  hR)  this  equation  represents  a  straight  line,  such  as 
opq,  which  passes  through  the  origin  of  coordinates. 


320 


CONDUCTION   OF  HEAT. 


[CHAP.  xiv. 


The  constant  h  is  positive,  and  its  value  must  lie  between  0  and  GO  . 
First  consider  the  case  where  h  =  0  ;  then  y.2  =  x,  and  this  equation 
represents  the  straight  line  opq  which  touches  the  curve  oa  at  the 
point  o,  and  cuts  irb  at  p,  "2-rrc  at  q,  etc.  The  abscissas  0,  xv  x2...  are 
roots  of  equation  (1).  We  have  further 

ir<xl<3ir/2  ;    2?7  <  z2  <  577/2  ;    ...nir<xtt<(n  +  ^)ir. 

As  n  increases  xn  approaches  the  superior  limit  (?i  +  i)ir.  Besides 
its  positive  roots  equation  (1)  has  also  negative  roots,  which  are  equal 
in  absolute  value  to  the  positive. 


FIG.  143. 

Next  suppose  that  0<h<l/R,  so  that  0<l-hlt<l,  and  hence 
y2>x.  This  represents  a  line  such  as  oa/3  (Fig.  143)  which  cuts  the 
curve  at  o,  a,  /3.  The  abscissas  0,  x^,  x2',  x3' ...  of  these  points  are 
roots  of  equation  (1),  and  we  have 

Q<Xl'<^ir;  77  <£.,'<  377/2;   2ir <xs'<5Tr/2  ...  (n  -  l)ir<xn'  <(n  -  $)TT. 

As  h  increases,  the  angle  aoir  approaches  the  angle  ITT,  and  the 
roots  approach  their  superior  limits.  If  h  =  l/R,  we  have  x  =  Q,  and 
the  roots  are  0,  \ir,  3^/2,  5:7/2,....  Now  if  l/E<h<<v,  we  have 
?/2=  -xj(hR-\).  The  straight  line  then  has  the  position  ofi'y'.  If 
in  this  case  we  represent  the  roots  of  equation  (1)  by  0,  a^",  x2"..., 
we  have  ^•jr<xl"<ir-  37T/2 <z./<27r....  As  h  increases,  the  roots 
approach  their  superior  limits.  And  if  h  =  oo  ,  we  obtain  ,TI"  =  TT  ; 


SECT,  cxxxn.]  THE   COOLING  OF  A  SPHERE.  321 

We  may  now  consider  the  roots  of  equation  (1)  as  known  and 
determine  m  from  them.  The  values  of  m  corresponding  to  the 
several  roots  are  0,  mv  m.2,....  We  may  neglect  the  negative  roots, 
because  the  terms  in  (h)  corresponding  to  them  may  be  considered 
as  included  in  those  arising  from  the  positive  roots.  We  therefore 
set  (m)  rd  =  A-ie~mi'*"'*t  sin  (m^r)  +  A2e~m*'K*  sin  (m.-,r]  +  ....  If  the  tempera- 
ture at  the  time  t  =  0  is  given  by  6=f(r),  we  have 

(n)  r.f(r)  =  Alsii\(mlr)  +  A.2sin(m2r)  +  .... 

Now  let  ma  and  mb  be  two  roots  of  equation  (k),  and  multiply  both 
sides  of  equation  (n)  by  sin(mar).  It  follows  by  integration  from  0 
to  R  that 

,-R  [R 

(o)  /  r.f(r)sin(mar)dr  =  'S,Ab     sm(mbr)sin(mar)dr. 

Now 


ra 

I  sin  (mbr)  sin  (mar)dr 


=  \  sin  [(mb  -  ma)R]/(mb  -  ma)  -  \  .  sin  [(m.  +  ma)R]l(mb  +  ma) 


I  =  [ma sin  (mbR)  cos  (maR)  -  mb cos  (mbR)  sin  (maR)]/(mb2  -  wa2). 

But  because  we  have  from  (k) 

maR  =  (1  -  A.R)  tg  (maR),   mbR  =  (l-hR)tg  (m^R), 
and  therefore  ma  tg  (mbR)  =  mb  tg  (maR), 

or  7??a .  sin  (?%Z?) .  cos(maR)  =  mh .  sin  (wa.R) .  cos (mbR), 

/K 
sin  (m,br) .  sin  (mar)dr  =  0, 

whenever  ma  and  7W6  are  different  from  each  other.  But  if  they  are 
equal,  (p)  is  indeterminate.  We  find  the  value  of  the  expression  (p) 
in  this  case  by  setting  mb  =  ma  +  e,  where  e  is  a  small  quantity.  We 
reach  the  same  result  more  simply  if  we  investigate  the  value  o 
the  integral 

r  =  2  •  L  [i  "~  cos(2mar)]dr  =  %[R  -  sin(2maR)/2mn]. 
We  thus  obtain 

[R 

Aa=2/R.  I  r.f(r)sin(mar)drl[I  -sin(2mnfi)/2maR]. 
Jo 

Hence  the  complete  solution  of  the  problem  is  contained  in 

(RrO=      sin(w1r)g-""w 
«-T«  /*«   *.\«— w 


322  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

In  the  simple  case  in  which  the  initial  temperature  of  the  sphere 
is  everywhere  equal,  we  have  f(r)  =  00,  and  then  find 

00  .  FT  sin  (mr)dr  =  6Jm?  .  [sin  (mR)  -mE  cos  (mR)], 
or,  by  the  use  of  (k), 

09f*r  sin  (mr)dr  =  hR00sin  (mE)^. 
We  therefore  get  the  result 


(  \    4.  ti-hRfl  ii  22 

O'^m^m^-sin^m^R)]      m.2[2m2R  -  sin(2m2R)] 

If  the  coefficient  of  radiation  E  and  therefore  also  h  are  very  small, 
or  if  the  radius  of  the  sphere  is  small,  the  product  hR  is  a  small 
quantity.  In  this  case,  if  we  neglect  the  higher  powers  when  the 
sine  and  cosine  are  developed  in  series,  we  obtain  from  (k), 

1  -  1  .  m^R2  =  (1  -  hB)(l  -  1  .  m*fl*), 

from  which  follows  ml2  =  3h/R. 

The  other  values  of  m  are  so  very  much  greater  that  the  corre- 
sponding terms  in  (r)  vanish  in  comparison  with  the  first  term.  We 
therefore  obtain  0  =  60.  e~3Hlcl<'B,  or,  substituting  the  value  of  h, 

(s)  0=00e-3JB/<"* 

We  can  derive  this  formula  more  simply.  The  quantity  of  heat 
which  the  sphere  radiates  in  the  time  dt  is  4irR~E6dt.  Thus  the 
temperature  of  the  sphere  increases  by  —  dd  ;  and  the  quantity  of 
heat  given  up  is  -  4?r/3  .  R?cpdd.  Hence  we  have 

=  -  4;r/3  . 


from  which  follows  6  =  00e~3£t'cf>K,  since  the  temperature  of  the  sphere 
is  On  at  the  time  t  =  0. 


SECTION  CXXXIII.    THE  MOTION  OF  HEAT  IN  AN  INFINITELY 
LONG  CYLINDER. 

Let  the  cross-section  S  of  the  cylinder  be  so  small  that  its  tem- 
perature 6  is  constant,  and  let  A  arid  B  be  two  cross-sections  separated 
by  the  distance  dx.  The  quantity  of  heat  -  Sk .  Wfdx .  dt  flows  through 
A  in  the  time  dt,  and  the  quantity  -  SktfO/'dx  +  320/3a;2 .  dx)dt  flows 
through  B  in  the  same  time.  Hence  the  part  of  the  cylinder 
between  A  and  B  receives  the  quantity  of  heat  Sk .  320/3z2 .  dxdt.  A 


SECT.  CXXXIIL]  THE   MOTION  OF  HEAT.  323 

part  of  this  heat  is  given  up  to  surrounding  bodies  by  conduction 
or  radiation.  If  P  is  the  perimeter  of  the  cylinder,  E  a  constant, 
and  if  the  temperature  of  the  medium  around  the  cylinder  is  0,  the 
heat  given  up  by  conduction  or  radiation  is  PEO .  dxdt.  Another 
portion  of  the  heat  received  serves  to  heat  the  cylinder ;  this  portion 
is  S .  dx .  cp .  dO.  Hence  we  obtain  the  equation 

Sk.-d*-efdx'2  =  Spc.'d6/-dt  +  PEe,  or  (a)  90/3/  =  K2.  c>20/3x2- A0, 

if  we  set  K*  =  k/cp  and  h  =  PE/SPc. 

If  the  state  of  the  cylinder  or  rod  has  become  steady,  we  have 
30/3/  =  0,  and  equation  (a)  takes  the  form  *2 . 320/3z2  =  h0.  From 
this  it  follows  that  (b)  6  =  A<?**K  +  Be-xVJtl\  If  the  temperature  of 
the  rod  is  given  at  two  points,  we  obtain  from  (b)  the  temperature 
of  the  intermediate  points.  We  will  assume  that  the  temperature 
of  a  certain  point  in  the  rod  is  00,  and  that  at  a  very  great  distance 
from  this  point  the  temperature  of  the  rod  is  0.  Then,  for  x  =  oo  , 
we  will  have  0  =  0  and  therefore  (c)  0  =  00 .  «~*v'*'1'.  But  if  the  rod  is 
not  in  a  steady  kstate  the  equation  (a)  must  be  used.  If  we  sub- 
stitute in  this  equation  0  =  u .  e~ht,  we  obtain  "dufdt  =  K2 . 3%/Bo;2.  The 
integral  of  this  differential  equation  has  already  been  given.  By  the 
help  of  equations  (c)  and  (d)  we  can  determine  the  cooling  of  any 
rod  which  is  heated  in  any  manner. 

We  will  here  consider  only  the  case  in  which  one  cross-section  S 
of  the  rod  has  the  temperature  00,  while  the  temperature  of  all  other 
parts  of  the  rod  is  zero.  The  heat  flows  from  S  toward  both  parts 
of  the  rod,  and  after  an  infinitely  long  time  the  temperature  at  the 
distance  x  from  S  is  given  by  0  =  00 .  e~'yjtlK.  On  the  other  hand,  the 
temperature  at  the  same  place  at  the  time  t  is  given  by 

<e)  6  =  60.e-'yv<  +  u.e-'*. 

In  this  case  u  must  satisfy  the  following  conditions  : 

(1)  It  must  satisfy  the  equation  dw/3*  =  K2 . 3%/Sx2 ; 

(2)  When  t  =  0  the  equation  holds  0  =  00 .  «-***/«  +  u ; 

(3)  When  x  =  0  we  must  have  u  =  0. 
Conditions  (1)  and  (3)  are  satisfied  by 

%  =  2/7T .  l~dX  sin  (\x)  (/(a)sin(Aa)e-X2K2<</a. 

Jo  ^0 

And  as  u  also  satisfies  (2)  we  have 


+  2/7T.  /  <Usin(Az)//(a)sin(AaX/c 


324  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

It  is  requisite  for  this  that  /(a)  =  -00e~aVr/K.     Hence  the  solution  of 
the  proposed  problem  is 

8  =  eo  .  e-*^'"  -  200/7T  .  e~M  f*dX  sin(Az)  .  f"s 
As  in  CXXXL,  we  can  give  the  expression 

2/7T  .  f°sui(Aa;)  .  si 

Jo 
the  form 


and  this  latter  expression  equals  1/2*7^7.  (e-(*-w*  _  g-(»f«w*«^     Sub- 
stituting this  expression  in  the  equation  for  6,  it  follows  that 

0  =  eXe-**"  -  -^=  .  fV"-0**"'  -  e-^w**)  .  <r«v»>  rfal 

2/Cx/7T<    -/O 

To  simplify  this  expression,  we  set  p  =  (a-  x)/2i<Jt,  a  =  x  +  2i<pjL     It 
then  follows  that 


_  g_«  _  , 

Vft/K  ,00 

_  r  e_^ 

TT      -V^  -  z/2K  VF 


We  obtain  in  the  same  way 


2* 
and  therefore 


(f) 


A  careful  examination  of  this  expression  shows  that  it  represents  the 
flow  of  heat  through  an  infinitely  long  rod.  For  t  =  0  the  lower  limit 
of  the  first  integral  equals  -  oo ,  and  the  value  of  the  integral  itself 
is  then  equal  to  JTT  ;  the  lower  limit  of  the  second  integral  is  in 
the  same  case  oo  ,  and  therefore  the  value  of  the  second  integral 
equals  zero.  Hence  for  t  =  0  we  also  obtain  0  =  0,  which  should  be 
the  case  for  all  cross-sections  of  the  rod,  except  for  the  heated 
section.  Both  integrals  have  the  same  value  for  x  =  Q,  and  there- 
fore 6  =  00.  Both  integrals  vanish  when  t  =  ao ,  and  for  the  steady 
state  of  heat  we  have  the  evidently  correct  result  0=QQ.  e~*^lK. 
Because  h  =  PE/Scp,  h  is  infinitely  small,  if  the  cross-section  of  the 


SECT,  cxxxiu.]  THE  MOTION  OF  HEAT.  325 

rod  is  infinitely  great,  or  if  the  coefficient  of  radiation  E  is  infinitely 
small.     Setting  h  =  0  in  (f)  we  come  back  to  a  case  already  treated, 


This  result  is  also  found  in  CXXXI.  The  expression  (g)  gives  the 
temperature  in  an  infinitely  extended  body,  having,  at  the  time  t  =  0, 
the  temperature  6  =  0  at  all  points,  with  the  exception  of  the  points 
on  the  surface  ce  =  0,  for  which  0  =  60. 

The  solution  (f)  holds  only  for  positive  values  of  x;  and  that  it 
shall  hold  for  those  parts  of  the  rod  which  correspond  to  negative 
values  of  x,  x  must  be  replaced  by  -  x  in  (f  ). 


SECTION  CXXXIV.     ON  THE  CONDUCTION  OF  HEAT  IN  FLUIDS. 

Up  to  this  point  we  have  treated  only  the  motion  of  heat  in 
solids.  The  results  which  have  thus  been  obtained  cannot  in 
general  be  applied  to  fluids,  because  any  difference  of  temperature 
which  causes  a  different  expansion  in  different  parts  of  the  fluid, 
occasions  so-called  convection  currents.  In  general,  differences  of 
temperature  are  more  quickly  equalized  by  these  currents  than  by 
conduction  alone.  The  relations  are  therefore  very  complicated.  We 
will  confine  ourselves  to  developing  the  general  equations  of  motion 
which  will  be  applied  in  some  simple  cases. 

We  use  the  notation  of  hydrodynamics.  The  equation  of  con- 
tinuity, which  expresses  that  the  quantity  of  matter  is  constant, 
becomes  [cf.  XLI.  (d)] 


The  momentum  received  by  the  unit  of  volume  in  the  unit  of 
time  is  equal  to  the  force  acting  on  that  unit  of  volume.  We  have 
therefore  from  XLI. 

A  =  p(du[dt  +  u'du/'dx  +  vdufiy  +  wduj'dz) 

=  'dXJ'dx  +  VXj'dy  +  'dXJ'dz  +  pX, 
B  =  P(dvfdt  +  u'dv/'dx  +  vdvfdy  +  wdvfiz) 
=  VYJ'dx  +  ^YJ-dy  +  -dYfiz  +  PY, 
C  =  p(dwfdt  +  u'dw/'dx  +  vdwfdy  +  vfdw/'dz) 

=  'dZJdx  4-  aZ,/3y  +  'dZfiz  +  pZ. 
The  symbols  A,  B,  C  are  introduced  on  account  of  the  use  to  be 
subsequently  made  of  them. 


326  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

Suppose  the  fluid  which  is  here  considered  to  be  a  liquid  and 
incompressible.  In  this  case  it  contains  energy  only  in  the  form 
of  kinetic  energy  or  heat.  If  the  body,  on  the  other  hand,  i& 
gaseous,  we  suppose  it  to  be  an  ideal  gas,  which  conforms  to  the 
law  of  Boyle  and  Gay-Lussac.  Such  a  gas  can  indeed  be  com- 
pressed, but  the  work  done  by  the  compression  is  transformed  into 
heat,  so  that  the  energy  contained  by  the  gas  is  independent  of 
the  volume,  and  is  determined  only  by  its  kinetic  energy  and 
temperature. 

A  volume-element  of  the  fluid  d<a  =  dxdydz  contains,  at  the  time 
/,  a  quantity  of  energy  which  is  the  sum  of  the  kinetic  energy  and 
the  quantity  of  heat  contained  in  it.  We  multiply  the  latter  by 
the  mechanical  equivalent  /  of  the  unit  of  heat.  If  E  is  the  unit 
of  volume,  and  if  the  unit  of  mass  receives  the  quantity  of  heat  0, 
we  have,  designating  the  velocity  by  h,  E  =  ^ph?  +  JpQ.  During  the 
time  dt  the  volume  element  d(a  receives  the  quantity  of  energy 

(c)  dE/dt .  dtda,  where  dEfdt  =  $d(ph*)/dt  +  J.  d(pQ)/dt. 

The  increment  of  energy  which  the  element  dw  receives  in  the  time 
dt  proceeds  from  the  following  causes : 

1.  From  the  work  done  by  the  accelerating  forces  X,  Y,  Z. 

2.  From  the   kinetic  energy  which,  in  consequence  of  the  flow  of 
the  fluid,  passes  into  the  volume-element  do>  through  its  surface. 

3.  From  the  work  done  by  the  surface  forces  X^   Y^  ...  on  that 
part  of  the  fluid  which  is  situated  on  the  surface  of  the  element  dw. 

4.  From  the  heat  contained  by  that  part  of  the  fluid  which  flows 
through  the  surface-element  d<o. 

5.  From  the  heat  which  passes  into  the  element  da  by  conduction. 
We  will  designate  these  quantities  of  energy  in  order  by  e^wdt, 

e^dudt,  e.jdwdt,  e4d(adt,  and  e.0da>dt;  e1  is  therefore  the  quantity  of  energy 
received  by  the  unit  of  volume  in  the  unit  of  time  only  through 
the  influence  of  the  accelerating  forces.  We  will  now  investigate 
the  values  of  e^  &>, . . . . 

We  determine  the  work  done  by  the  accelerating  forces  in  the 
time  dt  in  the  following  way.  The  volume  element  contains  the 
mass  pdo>  and  moves  in  the  time  dt  through  the  distance  udt  in  the 
direction  of  the  ar-axis.  Thus  the  force  X  does  the  work  pd<o .  Xudt. 
The  work  done  by  the  forces  Y  and  Z  is  determined  in  the  same 
way.  The  work  considered  is  therefore  p(uX+vY+u-Z}dwdt.  We 
have  represented  this  quantity  of  work  by  e^-wdt  and  hence  obtain 

(d)  el 


SECT,  cxxxiv.]    CONDUCTION   OF   HEAT   IN   FLUIDS.  327 

The  kinetic  energy  which  Jw  receives  from  that  part  of  the  fluid 
which  flows  in  the  time  dt  through  the  element  d<a,  is  determined 
thus.  The  mass  which  flows  through  the  surface-element  dydz  in  the 
time  dt  is  p.u.dt.  dydz  ;  the  kinetic  energy  of  this  mass  is  therefore 
^.p.itdt.dydz.h?.  But  if  we  set  U=^puh'2,  U  is  the  component 
of  flow  of  the  kinetic  energy  in  the  direction  of  the  a-axis.  Let 
the  corresponding  components  of  flow  with  respect  to  the  y-  and 
z-axes  be  V  and  W  respectively  ;  we  then  have 

F=|.M2,   W=\.po®' 

By  a  method  similar  to  that  used  in  XIV.,  it  may  be  shown  that 
in  the  time  dt  the  volume-element  da  receives  the  quantity  of  energy 
-(dU/'dx  +  'dV/'dy  +  'dW/'dz)d(adt.  We  represent  this  quantity  by 
e0db)dt  and  hence  obtain 


-  pu^idufdx  +  v'dv/'dx  +  wdwfdx) 

-  pv  (u'du/'dy  +  vdvj'dy  +  wdwj'dy) 
—  pw(udufdz  +  vdv/'dz  +  wdwj?)z). 

It  follows  from  this  by  the  help  of  equations  (a)  and  (b)  that 

e2  =  1  h2  .  dp/dt  +  p(u'du/'dt  +  vdvfdt  +  wdw\"t  t)  -(Au  +  Bv  +  Cw), 
or  more  simply 

e,  =  $hMp/dt  +  IpdWldt  -  (Au  +  Bv  +  Cw), 
(e)      «2  =  2  •  d(ph2){dt  -  (Au  +Bv+  Cic). 

The  quantity  of  energy  which  the  surface  forces  Xx,  Yy,  ...  impart 
to  the  element  du,  may  be  determined  in  the  following  way.  The 
force  -  Xxdydz  acts  on  the  surface-element  dydz  which  bounds  d<D 
on  the  side  lying  in  the  direction  of  the  negative  .r-axis,  in  the 
direction  of  the  #-axis.  The  fluid  particles  which  flow  in  the  time 
dt  through  the  element  dydz  traverse  the  path  udt  in  the  direction 
of  the  a-axis.  Thus  the  force  -  Xx  does  the  work  -  Xxdydz  .  udt.  But 
the  fluid  particles  in  the  surface-element  have  also  tangential  motions. 
Thev  traverse  the  path  wit  in  the  direction  of  the  y-axis  under  the 
influence  of  the  force  -  Yxdydz,  by  which  the  work  -  Yxdy'dz  .  vdt  is 
done.  The  same  particles  also  move  in  the  direction  of  the  2  axis, 
so  that  the  work  -  Zxdydz  .  ivdt  is  done.  The  total  work  done  by 
the  forces  in  the  time  dt  on  the  element  dydz  is  therefore 
Yxv  +  Zxw)dydzdt. 


328  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

We  will  designate  this  flow  of  energy  in  the  direction  of  the  z-axis 
by  U'dydzdt;  let  the  corresponding  flow  in  the  direction  of  the 
y-  and  £-axes  be  V'dxdzdt  and  Wdxdydt  respectively.  We  then  have 

U'=-  (Xxu  +  Yjo  +  Zxw\   V=-  (Xyu  +  Yyv  +  Z,w), 
W'  =  ~(X2u+Yzv  +  Z,w}. 

The  quantity  of  energy  which  d<a  thus  receives  is  determined  as 
in  the  foregoing  case,  and  is  equal  to 

e^udt  =  -  (d  U'j'dx  +  Wfiy  +  -d!P'rdz)d<adt. 
Hence  we  obtain 

Yyv  +  Zju^fdy 


But  using  equations  (b)  it  follows  that 

f  e3  =  X^u/Zx  +  Ypvfdy  +  Z&cfiz 
(f  )  +  Z,(dwfiy  +  Vvpz)  +  XtCdufdz  + 

[  +  YjCdvfdx  +  'dul'dy}  +  (A-  pX)u  +  (B-  pY)v  +(C-  pZ)w. 

The  quantity  of  heat  which  the  separate  parts  of  the  fluid  contain 
is  transferred  with  them  by  the  flow.  During  the  time  dt  the 
mass  pudt.dydz  enters  the  element  dw  through  the  surface  dydz, 
and  brings  with  it  the  quantity  of  heat  pudydzdtQ  or  the  energy 
JpudydzdtQ.  We  determine  in  the  same  way  the  quantities  of  heat 
which  enter  the  element  d<a  through  the  other  bounding  surfaces. 
If  we  use  the  method  given  above  and  set 


the  quantity  of  heat  e4d<adt  received  by  the  parallelepiped  d<a  in  the 
time  dt,  is  given  by 


or  e4= 

Hence,  by  use  of  equation  (a),  we  have 

(g)          e4  =  /.  -d(pQ)pt  -  Jp(de/dt  +  udQrd 

Finally  the   element  do>  receives   heat  by  conduction.     The  com- 
ponents of  flow  of  heat  are  [CXXIL] 

-Jk.VOfdx,    -Jk.Wfiy,    -Jk.-ddj-dz. 

If  we  set  the  quantity  of  energy  thus  received  by  da>  in  the  time 
dt,  equal  to  e^d^dt,  and  assume  the  conductivity  constant,  we  have 
(h)  e.a  =  J  . 


SECT,  cxxxiv.]    CONDUCTION   OF  HEAT  IN   FLUIDS.  329 

The  increase  in  energy  which  d<a  receives  in  the  time  dt  is  given 
by  [%.d(ph*)/dt  +  Jd(pQ)/dt]dudt.  At  the  same  time  the  quantity 
of  energy  («1  +  «2  +  «s  +  «4  +  «6)  dwdt  enters  the  element  do>,  and  we 
therefore  have 

(i)  \  .  d(fW)jdt  +  Jd(PQ)/dt  =  e1  +  e2  +  e,  +  e,+  ey 

Introducing  in   this    equation    the   values   found   for   ev   e.2,   e3,  ...it 
follows  that 


If  internal  friction  exists  in  the  fluid,  we  have  from  XLVII.  (h) 
Xx  =  -p  +  2p.  .  'duj'dx  -  ^(dufix  +  Zv/dy  +  "die  fix) 

ZJ,  =  fu(dwl'dy  +  'dvl'dz),  etc. 
By  the  help  of  these  relations  we  may  give  equation  (k)  the  form 

Jp(dQj 


r 

\          =  -  p(du/dx  +  dv 

+  (dz;/^)2  +  (dw/-dz)*  - 
I 


2  +  (du/oz 

For  the  determination  of  the  motion  and  temperature  of  the  fluid 
we  have  the  five  equations  given  under  (a),  (b),  and  (1).  These  five 
equations  are  not  sufficient  to  determine  the  seven  unknown  quantities 
u,  v,  w,  p,  p,  9,  and  6.  We  obtain  two  other  equations  in  the  follow- 
ing way.  The  total  quantity  of  heat  9  contained  by  the  unit  of 
mass  must  depend  on  0,  and  we  assume  that  (m)  9  =  c0,  where  c 
is  the  specific  heat,  a  constant.  If  the  fluid  considered  is  gaseous, 
c  denotes  the  specific  heat  of  constant  volume. 

The  second  equation  must  express  the  relation  between  density, 
pressure,  and  temperature.  In  the  case  of  liquids,  we  may  set  approxi- 
mately (n)  p  =  pJ(l+aB),  where  pQ  is  the  density  when  0  =  0  and 
a  is  a  constant.  But  for  gases,  if  V  is  the  volume  of  the  unit  of 
mass  at  pressure  p  and  temperature  6,  V$  the  volume  of  the  same 
mass  at  pressure  p0  and  temperature  0°,  we  have  ^F"=j90F0(l  +a0). 
Since  Vp=\  and  J>0=1,  we  have  (o)  p/p  =pJpQ  .  (  1  +  a0).  The 
equation  (o)  in  connection  with  (a),  (b),  (1),  and  (m)  serves  to  deter- 
mine the  unknown  quantities.  The  complicated  equations  which 
determine  temperature  and  motion  in  a  fluid  are  very  hard  to  integrate, 
so  that  up  to  this  time  no  case  has  been  completely  solved. 


330  CONDUCTION  OF  HEAT.  [CHAP.  xiv. 

SECTION  CXXXV.     THE  INFLUENCE  OF  THE  CONDUCTION  OF  HEAT 
ON  THE  INTENSITY  AND  VELOCITY  OF  SOUND  IN  GASES. 

We  have  the  following  equations  [CXXXIV.]  for  the  determina- 
tion of  motion  in  a  gas  in  which  the  temperature  is  variable  : 

1.  The   equation  of  continuity   [CXXXIV.    (a)],  which  may  take 
the  following  form : 

'dp/'dt  +  p(dufdx  +  'dv/'dy  +  'dw/'dz)  +  udpj'dx  +  vdpfiy  +  wdpfdz  =  0. 

2.  The  equations  of  motion  [CXXXIV.  (b)].     We  replace  in  these 
equations  the  forces  Xa  Xy,  ...by  the  values  found  in  XL VII.  (b) 
and  (h),  and  obtain  [cf.  XLVIIL  (a)] 

p(dufdt  +  uduj'dx  +  vdufdy  +  w'duj'dz) 
=  pX-  'dpj'dx  +  ft  V2^  +  If^^ufdx  +  'dv/'dy  +  'dwfdz)l'dx, 
and  analogous  equations  for  y  and  z. 

3.  The  condition  for  the  conservation  of  energy  [cf.  CXXXIV.  (1)]. 

4.  The   connection  between  the  heat  contained  in  the  body  and 
the  temperature  [cf.  CXXXIV.  (m)]. 

5.  The  Boyle-Gay-Lussac  law  [cf.  CXXXIV.  (o)]. 

Let  the  velocity  and  change  of  temperature  be  very  small  quantities ; 
the  same  is  then  true  of  such  differential  coefficients  as  'dp/'dx,  'dQ/'dx, 
etc.,  and  we  will  therefore  neglect  the  product  of  these  quantities, 
that  is,  terms  of  the  form  udpj'dx,  udu/~dx,  itdQjdx,  etc.  The  equa- 
tions 1 — 5  are  then  very  much  simplified.  We  obtain 

(a)  3(log  p)j?>t  +  'du/'dx  +  'dv/'dy  +  'dwj'dz  =  0. 

.If  we  further  set  /t/p  =  //,  equation  (2),  by  use  of  (a),  takes  the  form 

(b)  'du/'dt  + 1  Ip .  'dp/'dx  =  // V2«  -  £/*' .  32(logp)/ar3i5. 

Similar   equations   hold  for  u  and  v,  if  x  is   replaced   by   y  and  s 
respectively. 

Eliminating  0  in  equation  CXXXIV.  (1)  by  means  of  the  relation 
Q  =  cO  and  introducing  the  heat  equivalent  A  of  the  unit  of  work 
for  l/J,  it  follows  that  (c)  cp.Wj'dt-k\7-d  =  Apd(\o«p)l'dt.  We  have 
further  the  equation  [CXXXIV.  (o)],  (d)  p/p=p0fp0.(l  +  a8).  We 
consider  ///,  k,  and  c  as  constants.  We  substitute  p0  for  p,  if  p  or 
l/p  occurs  as  a  coefficient;  we  also  substitute  p0  for  p  in  (c).  In 
these  substitutions  we  neglect  only  infinitely  small  quantities  of  the 
second  order.  Setting  p  =  Po(l+a-))  we  obtain  (e)  log  p  =  log  p0  +  <ry 
because  o-  is  a  small  quantity.  Hence  equation  (d)  takes  the  form 
P=Po(\  +°")(1  +«#)>  °r>  because  0  is  also  a  small  quantity, 
(0  7>=/> 


SECT,  cxxxv.]          VELOCITY  OF  SOUND  IN  GASES.  331 


Equation  (c)  now  becomes  Wfdi  -  k/cp0  .  V2#  =  -dpo/cPo  •  d^/cM,  and 
if  we  set  K-  =  k/cp0  and  9  =^  cpQB/Ap0,  we  obtain  from  the  last  equation 
(g)  Wfdt  -  K2  V20  =  ftr/3*. 

By  the  use  of  (e)  and  (f)  we  may  transform  equation  (b)  into 

'du/'dt  +PQ/PQ  .  'do-fa  +pQa/p0  .  'dB/'dx  =  /*'  .  V2«  -  $/*'  •  ^(r/'dt'dx. 

Introducing  in  place  of  6  the  quantity  0  already  defined,  we  have 


'dv/'dt  +  b2.  'da-f'dy  +  (a2  - 


The  heat  required  to  raise  the  temperature  of  a  gram  of  air 
under  constant  pressure  from  6  to  0  +  dO  is  equal  to  C  .  dd,  if  C  is 
the  specific  heat  at  constant  pressure.  A  part  of  this  heat,  namely 
c  .  dO,  is  used  in  raising  the  temperature,  the  other  part  is  used  in 
overcoming  resistance  during  the  expansion,  by  which  the  woAp*dP 
is  done.  We  have  therefore  C.dO  =  e.dO  +  Ap.dT.  It  follows  from 
the  equation  pF=p0F'0(l  +  ad),  because  p  is  here  constant,  that 
p.dV=pJPr<ft.dOt  and  therefore  (i)  C=c  +  4p0a/pQ,  since  rop0=l. 

Finally,  if  we  set  a2=p0G/p0c  and  b2=p0/p0,  the  equations  (a),  (b), 
and  (c)  take  the  following  forms  : 

'da-j'dt  +  'dul'dx  +  *dvfdy  +  *dw/oz  =  0, 


(k) 


These  equations  are  due  to  Kirchhoff.*  Eeference  may  be  made  to 
KirchhofF s  work  for  the  application  of  these  equations  to  the  more 
difficult  cases  of  the  transmission  of  sound.  We  will  here  investigate 
only  the  influence  of  conduction  and  friction  on  the  motion  of  plane 
sound  waves.  First,  however,  the  physical  significance  of  the  con- 
stants (a)  and  (b)  must  be  determined. 

If  there  is  neither  conduction  nor  friction  in  the  air,  we  have  K  =  0 
and  p.'  =  0 ;  further,  if  the  vibrations  occur  in  the  direction  of 
the  z-axis,  we  also  have  v  =  w  =  0.  Under  these  circumstances  the 
equations  (k)  become 

If  the  second  of  these  three  equations  is  differentiated  with  respect 
to  /,  we  obtain 


•Kirchhoff,  Pogrj.  Ann.,  Vol.   134.     1868. 


332  CONDUCTION   OF   HEAT.  [CHAP..XIV. 

It  thus  follows,  by  the  use  of  the  first  and  last  of  these  equations, 
that  3%/322  =  a2  .  32w/3a;2.     An  integral  of  this  equation  is 


and  this  expression  represents  a  wave  motion  which  proceeds  with 
the  velocity  (1)  a  =  »Jp0C/p(>c  =  b-JC/c.  This  value  for  the  velocity  of 
sound  was  found  by  Laplace.  It  differs  from  the  value  calculated 
in  XXXV.,  which  was  originally  found  by  Newton,  and  which  in 
our  present  notation  is  b  =  JpQ/p0.  The  difference  between  the  two 
formulas  is  due  to  the  fact  that  in  the  first  we  have  taken  into 
account  the  heating  of  the  air  by  compression  and  its  cooling  by 
expansion.  Since  the  ratio  C/c  has  been  determined  by  direct  experi- 
ment, the  true  velocity  of  sound  in  the  air  may  be  calculated.  For 
atmospheric  air  at  0°  C,,  C/c  =  1,405  ;  hence  a  =  33815  cm.  This  value 
agrees  very  well  with  experiment. 

Suppose  that  a  plane-wave  is  propagated  in  the  direction  of  the 
z-axis,  and  that  K  and  /*'  are  not  zero.  The  vibrations  are  parallel 
to  the  x-axis,  so  that  v  =  Q  and  w  =  Q.  Since  u,  0,  and  cr  are  then 
functions  of  x  and  t  alone,  equations  (k)  become 


-dufit  +  62  .  "do-fix  +  (a2  -  V) 

39/3*  -  /c2  .  920/3.s2  --=  fa  fit. 

The  unknown  quantities  u,  0,  and  o-  are  periodic  functions  of  t. 
We  will  represent  by  h  a  real  magnitude,  and  by  u',  0',  and  o-'  three 
magnitudes  which  are  functions  of  x  alone.  It  is  then  admissible 
to  make  the  assumptions  (n)  u  =  u'  .  e*",  0  =  6'.  ehit,  a-  =  ar'  .  e*",  where 
i  =  -J  -  1.  By  the  help  of  these  equations  we  obtain  from  (m) 

hi<r'  +  du'jdx  =  0, 
hiu'  +  b'2  .  da-'/dx  +  (a2  -  tf)dQ'/dx  =  /*'  .  d2u'/dx*  -  ^'hi  .  do-'/dx, 


We  eliminate  a-'  from  these  equations,  and  then  have 

J  _  A  v  +  fci(a2  _  j2)  .  dQ'/dx  =  (b2  +  $p'hi)  .  tPu'/da?, 
\  du'/dx  =  K2d'2Q'/dx2  -  hiQ'. 

If  the  first  of  these  equations  (o)  is  differentiated  with  respect  to  x, 
u'  may  be  eliminated,  and  we  obtain  the  following  differential  equation  : 
(p)      K2(&2  +  *'p'hi)  .  d*&/dx*  +  (h*K'2  +  ^V  -  hazi)  .  d*Q'/dx*  -  h*i&  =  0. 
Since  this  equation  is  linear,  we  set  0'  =  emx,  and  obtain 
(r)  K-'m4(Z»2  +  *  X/w)  +  TO2(A2*2  +  |/*'A2  -  hcM)  -  hsi  =  0. 


SECT,  cxxxv.]          VELOCITY  OF  SOUND  IN  GASES.  333 

We  will  determine  the  exponent  in  only  in  the  case  in  which 
the  conductivity  as  well  as  the  internal  friction  is  very  small.  If 
«  =  0  and  /*'  =  (),  we  have  from  (r)  m=  -hi/a.  If  therefore  we  set 
m  =  (-hi  +  $)/a,  where  8  is  a  small  quantity  whose  higher  powers 
may  be  disregarded,  we  have  from  (r),  if  the  terms  K2//,  *28,  etc.,  are 
neglected,  (s)  8  =  -  [^'h2  +  (1  -  6'2/a2)K2fr2]/2a2.  But  it  follows  from 
(n)  and  (q)  that  one  value  for  0  is  0  =  e8x/a .  ehi(t~xM  •  the  other  is 
obtained  by  substituting  - 1  for  i,  which  gives  0  =  eSx/a .  «-**"-"/•».  Half 
the  sum  of  the  two  values  of  0  satisfies  the  conditions  and  is  at 
the  same  time  real,  since 
(t)  0  =  e-WaM-*»+ (i  -  tw* *n •  *w  m  cos  tyy  _  xja^ 

From  the  exponent  of  e  we  see  that  the  changes  of  temperature  in 
the  wave  diminish  the  further  it  travels ;  at  the  same  time  u  also 
diminishes.  The  sound,  therefore,  becomes  weaker  as  the  wave  travels 
further.  If  T  is  the  period  of  vibration  and  n  the  number  of  vibra- 
tions, we  have  h  =  27r/T—2mr. 

By  using  this  value  of  h  it  follows,  from  equation  (t),  that  the 
higher  tones  lose  their  intensity  more  quickly  than  the  lower  ones. 

The  mathematical  treatment  of  conduction  is  principally  due  to 
Fourier,  who  not  only  developed  the  partial  differential  equation 
which  is  at  the  foundation  of  the  treatment  of  conduction,  but  also 
gave  us  methods  for  the  solution  of  a  great  number  of  problems. 
His  principal  work  is  :  Thdorie  Analytique  de  la  Chaleur,  Paris,  1822. 
Of  the  later  works  on  this  subject  we  mention  Eiemann,  Partielle 
Differentialgleichungen,  edited  by  Hattendorf,  Braunschweig,  1876. 


INDEX. 


Acceleration,  2. 

Centripetal,  11. 

Resultant,  4. 
Action  and  Reaction,  48. 
Adiabatic  Curves,  268. 
Amplitude,  90. 
Attraction,  Universal,  30. 


Biot  and  Savart's  Law,  184. 
Bodies,  Structure  of,  50. 

Rigid,  58. 

Equilibrium  of,  59. 

Motion  of,  59. 

Rotation  of,  GO. 
Boyle's  Law,  267. 


Capacity,  Electrical,  131,  149. 

of  Coaxial  Cylinders,  153. 
of  Condenser,  141. 
of  Spherical  Condenser,140,152. 
of  Parallel  Plates,  151. 
Specific  Inductive. 
(Dielectric  Constant),  156. 
Relation  to  Index  of  Refrac- 
tion, 221. 
Capillarity,  121. 
Capillary  Constant,  121. 

Tubes,  125. 
Carnot's  Cycle,  272. 

Theorem,  274. 
Clausius's  Equation  of  the  State  of  a  Gas, 

290. 

Theorem,  275,  277. 
Collision,  49. 


Condenser,  Cylindrical,  153. 

Parallel  Plate,  151. 
Spherical,  152. 
Conductivity  for  Electricity,  197. 

for  Heat,  300. 
Conductors,  System  of,  147. 

Work  done  on,  150. 
Contact  Angle,  125. 
Continuity,  Equation  of,  105. 
Cooling  of  a  Sphere  by  Conduction,  318. 

by  Radiation,  319. 
Corresponding  States,  289. 
Critical  Temperature,  284. 
Current,  Electrical,  Continuity  of,  191. 
Force  of,  184. 
Force  of  Linear,  190. 
Measurement  of  Con- 
stant, 194. 

Measurement  of  Vari- 
able, 195. 
Potential  of,  185. 
Potential  Energy  of, 

193. 

Currents,Electrical,MutualActionof,192. 
Potential  Energy  of, 

193. 

Systems  of,  186, 190. 
Cycle  (Cyclic  Process),  268,  272. 
Carnot's,  272. 
Efficiency  of  Carnot's,  273. 


Damping  Action,  195. 
Deformation,  74. 

Relation  of,  to  Stress,  79. 
Density,  35. 


336 


INDEX. 


Descartes's  Explanation  of  Mutual  Ac- 
tions, 49. 
Dielectric,  155. 

Equations  of,  219. 
Plane  Waves  in,  221. 
Dielectric  Constant,  156. 

in  Crystals,  246. 
Displacement,  156. 
Dilatation,  76. 

Linear,  77. 

Principal  Axes  of,  78. 
Volume,  77. 

Dirichlet's  Principle,  136. 
Dissociation,  295. 
Dyne,  9. 


Earth,  Temperature  of,  302. 
Elastic  Body,  Equilibrium  of,  82. 
Motion  of,  89. 
Potential  Energy  of,  96. 
Elasticity,  Coefficient  of,  79. 

Modulus  of,  81. 
Electrical  Convection,  161. 

Displacement,  191. 
Distribution,  128. 

on  a  Conductor,  130. 
on  Conductors,  139. 
on  an  Ellipsoid,  133. 
on  a  Plane,  135. 
on  a  Sphere,  132, 137. 
Double  Sheets,  160. 
Energy,  145. 
Force,  Law  of,  128. 
Lines  of,  143. 
Images,  135. 
Oscillations,  215,  223. 
Polarization,  191. 
Potential,  128. 

of  a  Conductor,  131. 
near  a  Surface,  129. 
Electricity,  Theories  of,  127. 
Electrified  Body,  Force  on,  141. 
Electro-kinetic  Energy,  201,  210. 
Electromagnetism,  184. 

Equations  of,  188. 
Electrometer,  Quadrant,  154. 

Thomson's  Absolute,  142. 
Energy,  Conservation  of,  58. 
Kinetic,  14. 

of  a  System,  56. 
Potential,  57. 


Entropy,  271,  272,  278,  292,  294. 
Entropy  Criterion  of  Equilibrium,  294. 
Equilibrium,  58. 

Conditions  of,  58. 
of  Fluid  Surfaces,  12    . 
Equipotential  Surfaces,  23. 

Construction  of,  132. 
Equivalence  of  Heat  and  Energy,  268. 
Equivalent,  Mechanical,  of  Heat,  268. 
Ether,  230,  231. 

Fresnel's  Assumption  Concerning, 

233. 

Neumann's  Assumption  Concern- 
ing, 233. 
Euler's  Equations  of  Motion  of  Fluids,103. 


Falling  Bodies,  Laws  of,  5. 

Flexure,  87. 

Flow  of  Fluid,  108. 

Through  Tube,  119. 
Flux  of  Force,  42. 
Force,  8. 

Centripetal.  11. 

Components  of,  9. 

Line  of,  23. 

Measure  of,  6,  9. 

Normal,  13. 

Tangential,  13. 

Tubes  of,  145. 

Unit  of,  9. 
Forces,  Conservative,  20. 

External,  53. 

Internal,  53,  62. 
Fluid,  Conduction  of  Heat  in,  325. 

Elasticity  of,  81. 

Equilibrium  of,  62,  99. 

Motion  of,  103. 

Viscous,  118. 

Motion  of  Sphere  in,  109. 

Steady  Motion  of,  118. 
Fourier's  Theorem,  312,  315. 
Fresnel's  Formulas  for  Light,  231. 

Failure  of,  235. 
Friction  of  Fluids,  115. 

Coefficient  of,  115. 


Galileo's  Laws  of  Falling  Bodies,  5. 
Gas,  Elasticity  of,  82. 
Ideal,  270. 

Specific  Heats  of,  270. 


INDEX. 


337 


Gauss's  Theorem,  41. 
Gravity,  Acceleration  of,  5. 

Centre  of,  50. 
Gyration,  Radius  of,  60. 

Heat,  Conduction  of,  298. 

in  Fluids,  325. 

Flow  of,  between  two  Bodies,  317. 
from  a  Point,  204. 
from  a  Surface,  303,  316. 
in  a  Cylinder,  322. 
in  an  Infinite  Body,  305. 
in  a  Plate,  308. 
Fourier's  Equation  of,  298. 
Steady  Flow  of,  in  a  Cylinder,  323. 
in  a  Plate,  300. 
in  a  Sphere,  301. 
in  a  Tube,  301. 
Helmholtz's  Transformation  of   Euler's 

Equations,  106. 
Hertz's  Apparatus,  216. 

Form  of  Max  well's  Equations,  219. 
Huygens's  Construction,  262. 

Ice,  Formation  of,  307. 
Impulse,  8. 

Measure  of,  9. 
Inclined  Plane,  24. 
Induction,  Coefficients  of  Magnetic,  149. 

of  Electrical  Currents,  199. 

Coefficients  of,  202. 

Equations  of,  208. 

Law  of,  200. 

Measurement  of    Coefficients 
of,  203. 

Mutual,  201. 

Self-,  200. 

Inertia,  Moment  of,  60. 

Principle  of,  6. 

Isentropic  Curves,  268. 

Isothermal  Curves,  267. 

Joule's  Law,  197. 
Kepler's  Laws,  27. 

Lagrange's  Equations  of  Motion  of  Fluids, 

111. 
Laplace's  Equation,  45. 

Application  of,  46. 


Lenz's  Law,  200. 

Light,   Electromagnetic  Theory  of,  230, 

235,  237. 

Emission  Theory  of,  229. 
Principal  Laws  of,  230. 
Wave  Theory  of,  229. 
Lines  of  Force,  23. 

Electrical  Force,  143. 
Logarithmic  Decrement,  196. 


MacCullagh's  Construction,  263. 
Magnet,  Constitution  of,  163. 
Forces  acting  on,  169. 
Oscillation  of,  170. 
Potential  Energy  of,  171. 
Potential  of,  166. 
Magnetic  Axis,  166. 

Induction,  173,  177,  179. 

Coefficient  of,  180. 
Moment,  165. 
Permeability,  180. 
Poles,  163. 
SheU,  180. 

Strength  of,  180. 
Force,  166. 

due   to  Electrical   Cur- 
rent, 184. 
Law  of,  163. 
Lines  of,  174,  178. 
Tubes  of,  174. 
Magnetism,  163. 

Distribution  of,  165. 
Magnetization,  Intensity  of,  165. 
Magnetized  Sphere,  Potential  of,  168. 
Material  System,  53. 
Maxwell's    Electromagnetic    Theory    of 

Light,  230. 
Theory  of  the  Action  of   a 

Medium,  72. 
Melting,  291. 
Moment,  of  Force,  56. 
of  Inertia,  60. 
of  Momentum,  55, 
Momentum,  49. 

Moment  of,  55. 
Motion,  Constrained.  24. 
Curvilinear,  3. 

Equations  of,  of  a  Particle,  10. 
In  a  Circle,  11. 
Oscillatory,  12. 
Periodic,  1,  12, 


338 


INDEX; 


Motion,  Uniform,  1. 

Variable,  1,  2. 

Newton's  Law  of  Attraction,  30. 

Ohm's  Law,  197. 

Optic  Axes,  249. 

Optic  Axes  of  Elasticity,  247. 

Oscillatory  Motion,  12. 


Path,  1. 

Equation  of,  4. 
Pendulum,  25,  61. 
Period,  90. 

Points,  Electrical  Action  of,  144. 
Poisson's  Equation,  45. 

Application,  46. 
Polarization,  Angle  of,  234. 
Potential,  20,  21. 

Coefficients  of,  148. 

Difference  of,  22. 

of  a  Circular  Plate,  38. 

of  a  Cylinder,  39. 

of  a  Solid  Sphere,  37. 

of  a  Spherical  Shell,  36. 

of  a  Straight  Line,  39. 

of  a  System,  34. 
Poynting's  Theorem,  224. 
Pressure,  64. 

Hydrostatic,  62,  63. 
Principal  Section,  260. 
Projectiles,  7,  11. 


Kays,  Direction  of,  in  Crystals,  256. 

Velocity  of  Propagation  of,  257. 
Reflection  of  Polarized  Light,  232,  237, 

239 

Total,  240,  241. 
Refraction,  Double,  246. 

at  the  Surface  of  a 

Crystal,  261. 
in  Uniaxial  Crystals, 

264. 

In  a  Plate,  242. 
Index  of,  230. 
of  Polarized  Light.  233,  237, 

239. 
Resistance,  198. 

Measurement  of,  205. 


Resistance,  Measurement  of — 

Lorenz's  Method,  207. 
Thomson's  Method,  206. 

Shear,  77. 

Solenoid,  186. 

Solid,  Internal  Forces  in,  62. 

Sound,  Velocity  of,  in  Gases,  330. 

Spherical  Shell  under  Pressure,  83. 

Sphondyloid,  143. 

State  of  a  Body,  266. 

Diagram  Representing,  267. 
Equation  of,  267. 

Stateof  a  Gas,  Clausius's  Equation  of  ,290. 
Van  der  Waal's  Equation 

of,  286. 

Stokes's  Theorem,  21. 
Stress,  64. 

Components  of,  65. 
Equilibrium  under,  67. 
Principal,  69. 
Strings,  Vibrating,  95. 
Surface  Tension,  122. 


Tension,  64. 

Thermodynamic  Relations,  280. 

PropertiesofBodies,281. 
Thermodynamics,  First  Law  of,  268. 

Equation  embodying,  269. 
Second  Law  of,  170. 

Application  of,  279. 
Torsion,  85. 

Coefficient  of,  86. 
Trigonometrical      Series      Representing 

Arbitrary  Functions,  309,  312. 
Tubes  of  Force,  145. 


Uniaxial  Crystals,  259. 

Wave  Surface  in,  259. 
Units,  Absolute,  2,  211. 

Derived,  2. 

Practical,  214. 

Van  der  Waal's  Equation  of  State,  283. 
Vaporization,  Heat  of,  291. 
Vapour,  Saturated,  290. 

Specific  Heat  of,  292. 
Vector,  55. 
Velocity,  1,  2. 

Angular,  1. 


INDEX. 


339 


Velocity,  Components  of,  3. 

Resultant,  4. 
Velocity  Potential,  106. 
Vibrations,  Longitudinal,  91. 

Transverse,  91. 
Vortex  Motion,  107. 
Filament,  109. 

Wave  Surface,  251. 

Reciprocal,  257. 
"Waves  on  the  Surface  of  a  Fluid,  112. 


Waves,  Plane,  90. 

in  Dielectric,  221. 
Spherical,  93. 
Stationary,  93. 
Torsional,  93. 
Velocity  of,  91,  92. 

in  Crystals,  248,  249,  257. 


Weight,  9. 
Work,  14. 

Done  in  Closed  Path,  15, 16. 


QC  20  C462E  1897 
L  006  580  2