This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
)2 + e2"2 (dx 2 ) 2 + e2"3 (dx 3 ) 2 ,
(47)
where the functions, it, p, to, fi2, and fj.3 are now functions of x2 and x3 only. However, as in the case of the metric (8), we have the gauge-freedom to impose a coordinate condition on /z2 and fi3. From a comparison of the metrics (6) and (45), it is clear that we should be able to pass from equations which are valid for time-dependent axisymmetric systems to equations which are valid for time-independent non-axisymmetric systems by some simple rules of transcription. Thus, in place of Equation (12), expressing the conservation of angular momentum per baryon, we now have the conservation equation
"1^"°)=°'
(48)
(It may be noted here that in the Newtonian limit Equation (48) reduces to the Bernoulli integral - a fact one might not have suspected.) Quite generally, it can be shown that by the replacements u°-> — in1, i//->n,
t^-Wu 0 ,
v->p,-
CO-+-W,
u0-*iut,
u1-> — iu0,
B
d
at
o
and Z). Summation over the respective ranges of repeated indices is to be understood; also commas will be used to indicate ordinary partial differentiation with respect to the coordinate specified, while semicolons will be used to indicate corresponding covariant differentiation. Equation (3) must be considered together with the equations (4)
M , _ i ^
+ s (
^ _
2 M )
_ - ^
[ u
o
(
_ ^ /
and u'[ul(€+p)/N].J*0 (5) that ensure the conservation of the baryon number N and of the angular momentum (per baryon). In Eq. (5) u, is the covariant component of the four-velocity in the ip direction. The variational equations. -Starting then from an initial configuration that is stationary and is uniformly rotating with an angular velocity SI {=d
Q2 ^> a n d
?2e> We 3 + 2(1 - a ) 2 H(i
and
®, e
for the quantities defined in the context of the Einstein—Maxwell equations; and x> ?2> *F, and 0 will denote the corresponding quantities for the vacuum (and whose values are given in equations (62)-(64). With the solution for E given by equation (67), the solution for We follows from equation (50); thus ^* 0
^
,(l-|#t|») 1 j n + in l-aE*\*
^'y,
^L
( 1 - W ) |l-fft|2 \1-E*\2 |l-<x£t|2
l-£t|2
-J..-
( 69 )
But i _ ^
= i[(l-a)Zt + (l+a)].
(70)
Accordingly, we may write 4a 2 V9 = ^ V , w
(71)
232
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
where ru2 = \(l-a)Z^
+ (l+a)\2
= (l-a.)2(tF2
= \ (I-a)
(
+ $2) + 2(l-a?)¥+(l+a)2.
(72)
Similarly, from the remaining expressions for
H
-l^
(74)
= ^-^ii^k^-y
Finally,
V(^) *>2 x
(75)
°=~wr=-&*•
I t remains to solve for q2e and (v + fi3) to complete the solution. Equations (71) and (73), relating We and
(77)
From this last equation we deduce t h a t 2J*nHH*t
4a(l — a2) = —^—.—'-{{l-a)[('F2-
+ 2lF)lFQ]
+ (1 + a) [(?P+1)
(78)
Similarly, by combining the equations, (W2)]O = 2 ( l - a ) 2 ( V ^ o + # 0 o ) + 2 ( l - a 2 ) ^ o >
(79)
and ®e,o = —i[™2,o-®(™2),ol
(80)
w which follow from equations (72) and (73), we find 4a ®e 0 = - 1 {[(1 - a ) 2 ( " P - 0 2 ) + 2(l - a 2 ) f + (1 + a ) 2 ]
(81)
On colliding waves in the Einstein-Maxwell theory
233
Inserting equations (78) and (81) in the denning equation (82) «
we find after some considerable reductions that ^e,3 = j ^ { [ ( l - « )
2
(^
2
- n + (l+a)2]^,o + 2 ( l - a ) 2 ^ ^ , o } -
(83)
(84)
Similarly, we find 8 ?2e,o ~
4a2IJ/2
{[(1 -*)2(
Alternative forms of these equations are (1+a)2 / , ?2«
4 a
2
* ll^>!4t-(!P'-*')*. + 2*!P!P,0]> 2 |p
«f/2: - n ,0 +
(1+a)2 S ?2«,0
4 a
2
(85)
4a
(1-a)2 8
ip2
4 a
2
<J/2
[-(W-
+ 2
(86)
I t is now apparent from equations (58) that apart from the factor (1 + a ) 2 / 4 a 2 , the first terms, on the right-hand sides of equations (85) and (86), integrate to give the vacuum solution q2. Accordingly, we may write
(1+a) 4a 2
™
?22 +
*
(1-a) 4a 2
,(e)
(87)
where q2e) is to be determined from the equations 12,0
8 Y, ^ (¥»-*•)*3 +2**-^
(88)
and ?2:)3 = - | i ( ^ - ^
2
)^o+2^^^.
(89)
Again, making use of equations (58), (59), and (63), we can reduce equations (88) and (89) to the forms:
^—(»--*,)«...-2*{A.+ (l-^)[ln(l^^)]J,
(90)
2 ,<> ! 3 = -(«P-^), 2 ,3-2^{, + ( l - , ) [ l n ( l ^ ^ ) ] J ,
(91)
where, according to the solution for q2 given in equation (62) 4pq/i(l-p?j) 02,o-
[(l-pV)'
+ q*/**]*
92,3
^2q[(l-pV)2-q2fi2] [(l-p7})2 + q2/i2]2 '
(92)
The reduction of equations (90) and (91) with the known solutions (63) and (64) for W and
P = q/i,
and
A = 1 — p2rj2 — q2/i2
(93)
234
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
rewriting the basic equations in the forms, X=
*$?• --?<-V>f
r-w-m-M±£ 2q(a2-p2)
Apafi q2
'°
(tf+p2)2'
?2 3
'
{a' +fi*)*'
(94)
and making use of the identities A = 4:a2-2{a2+P2)A-(a2
+ P2)2,
and
(95) a2+fi2 =
2a-A.
I n this manner we find t h a t equation (90) eventually reduces to the simple equation (96)
^7<«) == •^ ( 1 - ^ ^ ^ p ' ' ' [(1— pr/)2 + q2/i2]2 which immediately integrates to give 7 (e)
2q3 p(l -fi2^2 p2 {i—priY + q2^ \+M>
(97)
where/(/t) is a function of fi which is left undetermined at this stage. Simplifying equation (91) in similar fashion, we find
2q(l-^)
=
u
4a 2
{q
2
2
2
(l~/i2)
2
gV2(l-^)
2
(a 2 +/^) 2
a + fi (tf+P )
(98)
On the other hand, differentiating the solution (97) with respect to fi, we find a(e)
¥2,3
2q{\-/i2
_ :
p<
4g> 2 a 2 +y? 2
2 |g
(l-^2) (c^+y?2)
2
g>2(l-^2) (a2 + j82)2 +/,3-
(99)
From a comparison of equations (98) and (99) we find,
/..-£u-v>
(100)
/=^(3-,«>;
(101)
Therefore,
and the required solution for q2e) is given by g2(l-/t2)2 ;+(3-/* 2 ) (l—pr))2 + q2ju,'
(102)
The complete solution for q2e can now be written down by combining q2 (given by equation (62)) and q2e) (given by equation (102)) in accordance with equation (87) (see equation (117) below).
On colliding
waves in the Einstein-Maxwell (b) The solution for
theory
235
(v+/i3)
We now turn to equations (46) and (47) for the solution of (v + /i3). These equations in our present notation are It
7i
1
- T1 — - 1i (" + / * 3 ) , 0 - 17 — ! I ( " + / « B ) , 8 = 75 (*«,0*«,S + ?M.0?2«.8) f
V
Ae
+ 2-^-(Hi0H%
+ H%H:3)
(103)
and 3 1 2'/(l' + /*3),0+2M^ + /«3),3 = Y 3 I 5 + Y—^2 - i ^ [ ( ^ , o )
2
+ fee,o)2] + ^[(Xe,3) 2 + ( ^ , 3 ) 2 ] }
AC
±Xe [JH H*+8H H%l <0 ta y/(AS)
(104)
Considering first equation (103), we can reduce the terms on the right-hand side of this equation by making use of the various definitions and equations in §4 and in the earlier parts of this section and in particular equations (56)-(59) of §4 and equations (72), (74), (75), (76), (83), and (84) of this section. Thus, we find Xe
V(AS) = 2
[H>0H% + H*H
4(l-a wW
WT^r^'^W^ (105)
(y0y,+
and 2 \Xe,oXe,3 Xe
\X
' 5,2e,o5,2e,3/
+ W
™* ) \ X 1
a
2
2
?2,0?2,3
t 1
*,0*,: xfri
+^p«( - ) (^ -^)+( +a)2]^,3 +
2
(1-a)2*^^,3}
x {[(1 - a ) 2 ( < £ 2 - ¥») + (1 + a ) 2 ] * 0 + 2(l -a) 2 ^^^}
236
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
+ {[(1 -a) 2 (
+ (l+a)2](Po
+ 2(l-a)20*F*Fo}l
(106)
2
where in the third line we have added and subtracted q2 Qq2 3/x ( =
(i = 0,3),
(107)
for its derivatives. We find
- 2 ( 1 + a ) <*>„<*> 3 ].
(108)
This expression combined with the one on the right-hand side of equation (105) (in the last line) gives 2 ( 1 _ a ) 2
wW
[2(l-a)^o^3 +2(l+a)^o^3 + (l-a)0(^o03+^30o)]
= ^ ) { ^ [ ( l + a)!C, 0 + ( l - « ) ( ! P ! P | 0 + M 0 ) ] + f0[(l+a)f3 + ( l - a ) ( ^ 1 2X
w V
3
+ <2>$3)]}
CF.W%+Y0W*.),
(109)
where, in arriving at the last step in the reductions we have made use of equation (107). The final result of the reduction of equation (103) is, therefore, - T ^ ^ + Z ^ . o -1 T — ^ + ^sU = r*(X,oX,t + Q»,aq*.») 1 r
7
A
where it may be recalled that x a n d q2 are the solutions appropriate to the vacuum. A similar reduction of equation (104) yields 3
1
--2{^x,o)2+(q^)2]+s[(x,3)2+K3r]}+2n^+^-^.
(in)
We now observe that equations (110) and (111), apart from the terms in w2 on the right-hand sides of these equations, are the same as the equations for the
On colliding waves in the Einstein-Maxwell
theory
237
vacuum in Paper I, equations (39), (44), and (45). We conclude that the solution of equations (110) and (111) is given by v + fiz = the solution for the vacuum + In m2 + constant.
(112)
With the solution for the vacuum given in Paper I, equation (61), we can now write the solution for (v + /i3) in the form ->" + P-3
=
w
(113)
2 (l-7)2)S(l-fi,2VP )i
4a 2
where A is defined in equation (93) and the constant of integration in equation (112) has been so chosen that when a = 1, the solution reduces to that for the vacuum. (c) The expression for the metric With the solution for the metric coefficients completed, we can write the metric in the form ds 2 = w 2 4a (l-^2)J(l-^2);
(cb/)2 l-y2
(d/i)2 l-/i2_ 1
-(l-7i2)l(l-fi2)i[Xe(dx2)2
+ -(dx'-q2edx2n Xe
(114)
where = 1 — p2i]2 — q2ju,2,
4 a 2 (1 —py)2 + q2/i2'
(1+a) 4a
2
2q/i (l-<x)22q/i 2 2 2+ (l—prj) + q /i ' 4a 2 p2
q2(l-fi2)2 :+(3"/* 2 ) (l—p7j)2 + q2/i'
w2 = ( l - a ) 2 ( V 2 + 0 2 ) + 2 ( l - a 2 ) ¥ r + ( l + a ) *F = ^^-[(l-pri)2
+ q2/i2l
and
(115) (116)
(117) (118) (119)
We shall find it convenient to rewrite the metric (114) by replacing rj and ju, by i/r and 6 where 7) — cos^" and fi = cos 6, (120) and defining i+< i-<
l - K |2 Xe
=
| i - < f I2
and
iq2e =
(121)
where 2£ and $ are formally the same as ' Z ' and ' E ' defined in Paper I, equations (39) and (41), and reduce to them, in the limit H = 0. With the foregoing
238
S. Chandrasekhar and B. C. Xanthopoulos
definitions, the metric (114) takes a form similar to what we had for the vacuum (cf. Paper I, equation (80)); thus d 2
* = T~it - / A fla[(<¥)2-(dfl)2] 2 4a (sinyr sine?)*
- ^ y K l - ' ^ + iU + ^ d s T , (122) where A = p 2 s i n 2 ^ + g 2 sin 2 0,
Qo2 + ^a = 1).
(123)
6. T H E D E S C R I P T I O N O F T H E S P A C E - T I M E I N A N E W M A N - P E N R O S E FORMALISM. T H E W E Y L AND THE M A X W E L L
SCALARS
I t is manifest from the metric written in the form (122) that a null-tetrad basis is provided by the vectors: 6
X1
X2
«(1) =+e<«> = (/,) = + - L (U,
-u,
0,
0
),
e[2) = +eW = (ni) = + ~
+ u,
o,
o
),
*
(U,
«(3) = - e ( 4 ) = K ) = - ^ 2 (0,
o,
V2t
-iva.%
),
eii)=-e^
o,
V2._,
+ iV2+
),
=
(mi)=-~(0,
(124)
where, for the sake of brevity, we have introduced LT
VA w , (sin \]r sin 6)1 2a
and S)
-
—
V = (siniijj- sin o)V i
\±S
{\-\S\2f
(125)
The quantities J ± , as defined, satisfy the identity J _ J 2 * + J2*J2 + = 2.
(126)
With the aid of the null basis (124), we find that the non-vanishing spin-coefficients are given by [7(l-|
P=
cot xjr + cot 6 (2V2)C7 '
C/(l-K|2)-v/2' _ fl = +
cot ijr — cot 8 (2V2)U '
On colliding waves in the Einstein-Maxwell theory e =
1 (lnEA e +(lntf). f (2V2) U +7 2(1-
r
239
1
r**,-< r£+#*f9-*fS)
I (lnC7) 0 -(lnC/)^ = (2V2) U + 2 ( 1 - l| * | » ) v "' *"* < * - «""*r*-*** - " - .fl«+< •—.»/j-
(127)
With the foregoing spin-coefficients as the only non-vanishing ones, the Weyl scalars S^ and ^3 and the Maxwell scalar ^>1 vanish identically, a consequence, really, of the existence of the two commuting Killing fields. The remaining scalars follow from the Ricci identities (M.T. Chapter 1, equations 310 (/), (b), (j), (a), (n), and (g), respectively) by noting t h a t in the present context the directional derivatives, D and A, are given by e(2)
= D = — K K ( ^ T +^ \
£V2W
and
30/
eW
= A =-±—l^--^\.
(128)
£ V 2 W 80/
We find !f2 = fip — A
0 p
, 3_oo _
/*
, . . .„. (T(T* 1 / 3 . 3\ , ) l n p - (/O + e + e*)U-^2 \di/r- +30;
I — - — lln/t+(/t + 7 + y*) +
f/V2 \3^"
^
/* >
(129)
On evaluating the scalars with the aid of the spin coefficients (127), we find ^=-^(cot^-cot>0) ¥«
+
+ f&) JL.2 t / ffl)(f*f 2(1_K|2)2
2U'
* + < > * *M+2
+4
(130)
2[7 a (l-|<*T)
(cot ^ - c o t 0) + 2
+-
•r + *
\u
% + s% 1-
}•
(131)
u)
r + i \e,e~%&,
1-1
(132)
240
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
_. *-oo - ==
cosec2ijr + cosec 2 6 ,, . U,A i „v n{U„ cot£ + cotfl +*<** + + «**)-*(-£ + -#)
cot \jr ^vr+cot8( ^ -W^{-
K f + ffll2
+2
1
( c o t ^ + cot6')(l-|
=
(n
o,
l
;
^
;
cot ilr — cot (9 f cosec2 ilr + cosec2 6 ,, , „, „(U H U ,,\ ^ 7 ^ ( cot^-cotfl + i ( c o t ^ - o 0 t g ) + 2(-If-^j ( c o t ^ - c o t ( 9 ) ( l - K | I22 ^) 2' /''
*™ = -2U^-%)k*iCOtf
+ COtd) + 2
'
+ i ( c o t ^ - o o t g ) ^ ± ^ + ^ - M .
(135)
With the specification of the null basis, the spin coefficients, and the Weyl and the Maxwell scalars, we have completed the description of the space—time in a Newman—Penrose formalism.
7. T H E I N T R O D U C T I O N O F T H E N U L L - C O O R D I N A T E S A N D T H E E X T E N S I O N OF T H E S P A C E - T I M E INTO R E G I O N S I I , I I I , AND IV
So far, we have concerned ourselves only with the region of the space—time in which the interaction between the colliding waves — gravitational and electromagnetic - occurs; this is region I in the space-time diagram illustrated in figure 1 and to which the metric (122) we have derived applies. To specify the limits of this region and to extend the space-time beyond these limits, it is convenient to rewrite the metric (122) in terms of two null coordinates, u and v, in place of \]r and 6, denned by the equations (cf. I, equation (96)) 7) = COS^r = U\/(l— V2) + V\/(l— U2)
and /M
= COS6» = M V / ( 1 - « 2 ) - W V / ( 1 - « 2 ) >
(136)
or, equivalently, w = cos±(<9 + ^ )
and v = ain\(6-i/r).
(136')
By this transformation,
m'-m'=V{l_^;X_vr
(i37)
and the metric takes the form ds2
~a*[(l-u*-v*)(l-u*)(l-v*)$dudV 1
fll 2
/i »2
• ! _ m 2 l d - ^ ) d ^ + i(l+«f)dx 2 | 2 ,
(138)
On colliding waves in the Einstein-Maxwell
theory
241
FIGURE 1. The different regions of the space-time that have to be distinguished. In §5, the metric appropriate to region I, in which the interaction between the colliding waves (propagated along the null directions u and v) occurs, is derived. In §§7 and 8, the metric appropriate to regions II, III, and IV is considered. (For a more detailed description of the space—time see figure 2 in §12.) w h e r e xu2, A, a n d $, given in t e r m s of 7] a n d fi in e q u a t i o n s (115)—(118) a n d (121), m u s t be r e w r i t t e n in t e r m s of u a n d v w i t h t h e aid of e q u a t i o n s (136). T h e r a n g e s of t h e c o o r d i n a t e s u a n d v, as we h a v e p r e s e n t l y defined t h e m b y e q u a t i o n s (136), are r e s t r i c t e d t o t h e d o m a i n 0 < w < l
and
0 < v ^ 1.
(139)
H o w e v e r , it is m a n i f e s t from t h e form of t h e m e t r i c (138) t h a t it is singular on t h e surface u2 + v2 = 1.
(140)
I n d e e d , a c u r v a t u r e s i n g u l a r i t y develops on t h i s surface as m a y be verified b y considering die *F2, since (M.T., p . 44, e q u a t i o n (299)) - 2 Me ¥2 = Cim
PnWn1
= Ci]kl mi m? mk ml
(141)
is i n v a r i a n t t o scale t r a n s f o r m a t i o n s , i.e. t o t e t r a d t r a n s f o r m a t i o n s belonging t o class I I I (cf. M.T., §8(g), e q u a t i o n (347J). T h u s , b y e q u a t i o n (130),
UV
'
=
2a2
*R
V{l~u2-v2)-
uv^{l-u2)^(i-v2) 2 2
(i-KI )
(i—u2—v2y
;
(142)
a n d t h i s manifestly diverges o n t h e surface (140). I n c o n t r a s t t o w h a t h a p p e n s onu2 + v2 = 1, we h a v e o n l y c o o r d i n a t e singularities on t h e surfaces, u = 0,
0 ^ « < 1
and
v = 0,
0<w
(143) Vol. 398. A
242
S. Chandrasekhar and B. C. Xanthopoulos
separating region I from regions II and I I I . Following the original prescription of Penrose, we shall extend the metric (138) across these surfaces by the substitutions uH{u)
and
vH(v),
(144)
in place of u and v in the metric coefficients, where H(u) and H(v) are the Heaviside functions that are unity for positive and zero values of the argument and zero for negative values of the argument. By this substitution, we shall obtain a C°-extension of the metric; but ^-function singularities or //-function discontinuities, or both, will occur in the Weyl and the Maxwell scalars. Postponing to § 10 the consideration of the singularities and the discontinuities that occur on the surfaces (143), we shall pass on now to the metrics one obtains for regions II, III, and IV by the substitutions (144). We consider first the metrics one obtains for regions II and III. Since the metric (138) is entirely symmetric in u and v, it will suffice to consider the metric in region II. In this region, rj^-v,
/i-^—v,
A^l—v2,
and
<J->1— v2,
(145)
by the substitution (144) and the various functions defined in equations (115)—(119) become A=l-w2, n
~^
W= l-2pv
[4 + p2(v* +
+ v2,
-2q(l-pv)/p,
6v2-3)-4pv(l+q2+p2v2)] + 2(l-a.2){l-2pv
+ v2) + (l+a.)2,
ru2 1—v2 4a 2 1 — 2pv + v2' and
+ (l-ac)2[4: + 2pv(v2-3)-p2(l-v2)2]}.
(146)
With these definitions, the metric takes the form w2 ds 2 = —r AuAv — (1-
2
Xe(dx*)
+
1
-(dxl-q2edx2)2
(147)
Xe
where w2, xe, and q2e are functions of the one null-coordinate v only. The metric appropriate to region III can be written down by similarly setting 7/ = fi — u in equations (115)—(119). I t is clear that, unlike in the case of the vacuum, the space-time in regions II and III is not flat. As to what their precise nature is, we shall consider in the following section.
On colliding waves in the Einstein-Maxwell theory
243
The extension of the space—time into region IV Since both u and v are negative in region IV, the effect of the substitutions (144) on the metric is to set both u and v equal to zero in the expressions for the different metric coefficients. The resulting metric can therefore be obtained by simply putting v = 0 in the expressions listed in (146). We thus find that ^=±[(l-oc)2q2+p% p
and
g2e = 0,) j
2
Xe
= w /*** = C,
(148)
J
where C is a positive constant. Therefore, by the extension, the metric in region IV becomes ds 2 = 4Cdudv — CfcLc^ + i f c L r 1 ) 8
(149)
which is manifestly Minkowskian. We conclude that the space—time t h a t is causally to t h e past of the instant of collision is flat; clearly a requirement of the circumstances we are envisaging. Finally, it should be noted that by the substitutions (144), we have achieved a unique C°-extension of the space—time in region I to the entire space—time. I t is, of course, conceivable t h a t other manners of extension are possible.
8. T H E N A T U R E O F T H E S P A C E - T I M E I N R E G I O N S I I A N D I I I
As we have seen in §7, the space—time in regions I I and I I I is described by a metric of the form ds 2 = e2v[(dx°)2-(dx3)2]-e^(dx1-q2dx2)2-e2^{dx2)2,
(150)
where v,i/r,q2, and/t 2 are functions only of one or the other of the two null-coordinates x° + x3; and it will clearly suffice to consider only one of the two cases. Choosing x° — x3( = w)as the independent variable, we shall indicate by primes the derivatives with respect to the argument; thus
Since the metric (150) is of the standard form considered in M.T., Chapter 2, we can a t once write down the tetrad components of the Riemann tensor by specializing appropriately the expressions listed on pp. 78-79. We thus find that the non-vanishing components of the Riemann tensor, in the present context, are given by -""1313
=
-"1010
=
-"1310
2v
= e- [ft" + i/r'(T/r'-2v')]-le2*-*,'-2fiz(q'2)2 -'''1332
-'''1002 -^1002
"- ^^11002233
=L
(say),
(152)
^"1302
i^-^-^[ql + q'^f'-2v'-^
=N
(say),
(153)
244
S. Chandrasekhar and B. C. Xanthopoulos
and -^2323
=
^2020
=
^3220
,
= e-" [fi,;+pfo't-2v')-]
+ }e*+-»*-*'(q,)a
=M
(say).
(154)
Using these expressions, we find t h a t the only non-vanishing components of the Ricci tensor are R00 = R33 = R03 = -(L + M). (155) Besides, R = 0 and accordingly, -^00
=
G
00
a n d
^33
=
G
33-
(156)
The solutions (2) and (7) of Maxwell's equations in the earlier context of the metric (1) can be directly specialized to our present context when the metric coefficients are functions of (x° — x3) only and v = /i3. We find ^12 = ^03 = 0,
F01 = F13 = e-*-»A'
and
Fm = FM = e-+-"B'.
(157)
(158)
The corresponding expressions for the Maxwell stress-tensor are: ^oo = ^33 = ^03 = - e ^ - ^ [ ( ^ l ' ) 2 + ( 5 ' ) 2 ] ;
(159)
and the remaining components vanish. From equations (155) and (159) it follows that the Einstein-Maxwell equations provide only the single equation
= -2e^-2"[(i')2+(B')2],
(160)
while N is left unrestricted. We conclude that any metric of the form (150), in which the metric coefficients are functions only of a single null-coordinate, is a solution of the Einstein—Maxwell equations provided L + M < 0; and among the six functions v, xj/, q2, /i2, A, and B there is only the one constraint (160). (This result seems to be known; see Misner et al. (1970) and Bell & Szekeres (1974).) The requirement L + M ^ 0 follows from equation (179) (below) since 0 2 2 = | ^ 2 | 2 ^ 0. We shall presently identify the two invariants of this space—time. For a direct comparison with the equations obtained for region I, it is convenient to define fi=f + Pi, X = e~f+H, (161) and . e, & = X + k2 = 1±g(162) With these definitions, the metric (150) takes the form e'5 ds 2 = e 2 " [ ( d a ; 0 ) 2 - ( d a ; 3 ) 2 ] - - — - — | (l-
(163)
On colliding waves in the Einstein-Maxwell theory In terms of the variables (161) and (162) the expression for M+L equation (161) becomes M+L
/r+|(/n 2 -2.//?'+^l
= e-
245 given in (164)
or, since rfZ
(165)
-2»//?'+(12JfX, 2 2
(166)
and
(1
x =
we can write M+L
/r+K/n
= e~
2
(i-KI )
Also, we may note here for future reference t h a t
M-L+2IN
= e-2" {(^-^r+^+^'HX-vn^v;-^') (
+- ^+i[?J+?i(3^'-2r'-/»;)]} = e
-^"+-(A'-2i'')^"-Ar(^")2 u: A; A-2
(167)
or, substituting for 2£ and ^ in terms of $', we have i / - L + 2iiV
2e-2"(l-
+ (/?'-2/)
(168)
7%e spin coefficients and the Weyl and the Maxwell scalars From a comparison of the metrics (122) and (163), it is evident that with the definitions U = e",
V = e^2
and
J± =
\+S V(l-I
(169)
we can set up for the space—time, we are presently considering, a null-tetrad basis of the same form as (124) in §6. With the basis so chosen, we find that the non-vanishing spin-coefficients are A= - ( V 2 ) e " * ^ — > 1
/* =
~^e-'/]' V2
and 7 =
V2
V+
?*£' — ,
2(i-KI 2 ) J
(170)
and, besides, these are functions of (x° —x3) only. From these facts and from equations (129) (in which 3^ and 3^ are now to be replaced by 9zo and 8^3) we find t h a t only lFi and
(171)
and •[022] = e - > V 2 + ^ + y + y*) + AA*.
(172)
246
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
Evaluating these expressions, we find, on further comparison with equations (166) and (168), that 2e" 2 " [*".] = i - | < f | 2 +(
[r A'-2,v+2^;"
= {^(M-L
+ 2iN),
(173)
and = -e-2*
/r +iW ^V-fw]
= -(M+L).
(174)
We have enclosed the expressions for Wi and
M
( l - « 2 ) mlt
(175)
-(l-v*)%.
(176)
and [li\ =
Since both Wi and
(177)
We shall accordingly write w = 4
(M-L
+
2iN)V(l-v2)
(178)
i-S*
and
-(M+L)V(l-v2
(179)
From the non-vanishing of only Wi and
On colliding 9.
waves in the Einstein—Maxwell
theory
T H E E X P L I C I T FORMS OF THE W E Y L AND THE M A X W E L L FOR
247 SCALARS
THE CASE q = 0
I t will be convenient to have the explicit forms of the Weyl and the Maxwell scalars, in the different regions for the case q = 0, before we proceed to the consideration of the singularities or the discontinuities, or both, t h a t occur along the null boundaries separating them. (We shall not consider in detail the case q =£ 0 in this paper: the relevant expressions for the Weyl and the Maxwell scalars are far too complicated; and it does not appear that they will change any of the principal conclusions.) (a) The Weyl and the Maxwell scalars in region I When q = 0 and p = 1, the expressions for the various metric functions simplify considerably. Thus, in place of equations (115)—(119), we now have A = A = l—n2 = sin 2 \]s, W= sin 6 sin \jr tan 2 \\jr,
m/2a. = a + b sin 8 sin \]s tan2\\jr = X Xe=X2cot2^
and
(say),
q2e = 0,
(180)
where we have written a = (l+a)/2a
and
6=(l-a)/2a.
(181)
Also (cf. equations (121) and (125)) ._Z2cot2|^-l X2 cot 2 i ^ + l
and
U•
sin<^ X. sini0
(182)
And the metric takes the simple form
d52 =
(lr^y X 2 [ W ) 2 _ ( d ^ ) 2 ] — sin 6 sin i/r
(da;1)2 + (Z2cot2^)(da;2)2 X2cot2^
(183)
The task of the reduction and the simplification of the expressions for the Weyl and the Maxwell scalars (which is considerable even in this simpler case) can be lightened if appropriate use is made of the following relations: Xe = b cos 6 sin \jr tan 2 \\jr; Ufi/U = -\ootd
+ Xfi/X;
X^ = 6(2 + cos i/r) sin 6 tan 2 \\jr,
Uj/U
^Icotf
+ X^/X,
(184)
and
\f±S,e
4X 2
(X cot2 | f +1) :
{ - a c o s e c f c o t 2 ! f + 6[sin0+sin(0 + f)]}.
(185)
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
248 We find: W0n =
3 sin* (9 4 J P c o f i f sin§^
sin(0 — \]r) 4 cot |^ sin 2 i/r sin 6
+ ab , C ° t 2 ^ , [sin (d + ilr) + 2 sin 0] 2 sin 6 smyr - 62{[sin ((9 + f) + 2 sin Of + sin 6> sin (6 -
4
3 sin*6> 4 ~4X cot4^sin!^
f)}
„ sin(0 + ^ ) a2 . 2„ . . l c o t 4 | ^ sin ^ sin 0
+ ab .""f 2 ^, [sin (0 - ilr) + 2 sin 0] 2 sin 0 sin ^r -6 2 {[sin (6- xjr) + 2 sin d]2 + sin (9 sin (6» + f)} ^, =
1
sin**?
8X4cot4^sini^
a 2 (3 cosec2 xjr + cosec2 0) cot 4 \ijr
c o t ^•)/^
- 2afe , „ ^ , [(sin2 0 - sin 2 ^ ) + 4( 1 + cos ^ ) sin 2 0] sin 0 sin ^/ + 62[3 (sin2 6 - sin 2 f) + 4( 1 + 2 cos ^ ) sin 2 0]
00 = + o/ • a • 6w a ^ 44— ^ T *oo T t s i n (<9 + rV^) + 2 sin 0] 2 2(sin0sm ^)2X cot 2 |^r
022
+
_ 02 =
_
a6 ^ ^ ^ [ - ( ^ - ^ + 2 sin 0] 2 , 2(sin0sin5^ a6 2 sin g+ ^) 2 ( s i n ( 9 s i n 6 f ) 2 X* c o t t y [ (
+ 2
sinfl][sin(fl-^) + 2 sing], (186)
The limiting values to which the various scalars tend as we approach the null boundary, u = 0,
0
(187)
can be readily ascertained by noting that on this boundary 0 + ^ ^ 7 1 — 0:
sin0 = s i n ^ = V ( l — v2)> s i n ( 0 + f ) = O; cot2^Jr=
sm{6-\jr)
(l+v)/(l-v),
cosi/r = — cos0 = v, =
and
2v^{l-v2), X = a + b(l-v)2.
(188)
On colliding waves in the Einstein-Maxwell theory
249
We find
2ab (l-v)± 'IP {1+vfi'
2ab
0
l-v2)\,
^ 9
2ab t\ X* \l+v)' V,0
V.2
2
2X\\-v2)
i[a-b{l-v)
1 [a-b(l 2XHl-v2)i
][av
+ b(2 +
-v)2][a-b(l
v)(l-v)2],
-v)2{l
+ 2v)].
(189)
An important consequence of the solution for
0
< M <
(190)
1,
can be obtained by simply replacing v by u and interchanging W0 and 9,i and
and
?M
= 0
(191)
>
yhere Z = a + 6(l-v)2
(192)
is the same expression that we defined earlier in equations (181) and (188); and the metric consistent with the notation of § 8 is ds 2 = 4X 2 [(dz°) 2 -(dar 5 ) 2 ]-
X 2 (l + v)2 (dx2)2 + -^{l-vf
(dx 1 ) 2
(193)
where v = (x° — x3). Comparison of the metrics (150) and (193) shows that e" = 2X,
eP=(l-v2},
e* = (l-v)/X,
and
e^ = Z ( l + « ) .
Remembering that now, N = 0, the expressions for M+L equations (164) and (167) give
M-L
= e~
(-)
+(/3'-2v')X Xi
(194)
and M—L given in
36 [a-b{l-v)2], X4
(195)
250
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
and
P" + W? + \(^j-WP'
2ab ~X
(196)
Inserting these expressions in equations (178) and (179) and remembering that $ is now real, we obtain f,4 = ^ [ « - ^ ( l - ^ ) 2 ] V ( l - « 2 )
and
022 = ^ V ( l - « 2 ) .
(197)
We observe that these expressions for Vfi and &22 agree with the values they attain, as w-> + 0 (from the side of region I), on the null boundary separating regions I and I I . They are, therefore, continuous on the boundary. The remaining scalars, !^0, W2,
, and
^ 2ab <2>00 = — V ( l - ^ 2 )
(198)
where Y=a
+ b(l~uf;
(199)
and these will be continuous on the null boundary (v = 0, 0 ^ u < I), separating regions I and I I I . We also note that as we approach the boundary (v^ + 0 (from the side of region II), u < 0), separating regions I I and IV, *Fi and
and
(200)
And WQ and
10.
T H E CHARACTERIZATION OF THE S I N G U L A R I T I E S AND THE D I S C O N T I N U I T I E S ALONG THE NULL B O U N D A R I E S
Singularities or discontinuities, or both, along the null boundaries separating the different regions will arise from the substitutions, u^uH(u)
and
v->vH(v),
(201)
t h a t we have made for extending the metric, derived for region I, into the other regions. An exact description of the resulting singularities and discontinuities can be obtained from the following simple considerations. Let/(%) be a function of u, which at u = 0, has the Taylor expansion f(u) = a + bu + cu2+....
(202)
On colliding waves in the Einstein-Maxwell theory , f{u) = f(u H(u)) be the function obtained from/(w) by the substitution u^uH(u). u = 0, has the expansion
251
Further, let
f(u) = a + buH(u)+cu*H2(u)
(203) Then/(w), at
+....
(204)
By differentiating /(it) successively and making use of the identities H'{u) = 8(u),
u8{u) = 0
and
uS'(u) = -d{u)
(205)
(in the sense of distributions), we find /;(0) = 6
and
f"+(0) = b8(u) + 2c,
(206)
where it should be noted that the validity of these relations depends crucially on the definition of the Heaviside function requiring H(0) = 1 and H(u) = 0 only for u strictly negative. From the relations (206) it follows t h a t f(u H(u)) will have a (^-function singularity in its second derivative at u = 0 and that /;(0)=/*(0)+/'(0)*(«);
(207)
and, in contrast, the first derivative of/(w) will exhibit only discontinuity at u = 0. We shall use these results in deciding which among the Weyl and the Maxwell scalars will exhibit a ^-function singularity and which a simple //-function discontinuity along the null boundaries separating the different regions. We shall consider first the boundary {u = 0;0^v<
1)
and
6 + i/r^n-O,
(208)
separating regions I and I I . By the transformation (136) u = cos±(iJr + d), (c^
+
al)/(M)
= - /
'
( M ) s i n
(209) ^
+
(210)
^
and
(c|-clK)=0-
(211
>
On u = 0, equations (210) and (211) give
d f{u) = f(0) \
and
[&-hh\
= 0.
(212)
M-0
Turning now to the Weyl and the Maxwell scalars listed in equations (130)-(135), we observe that W2 and $ 2 2 do not involve any second derivatives of the metric coefficients; they cannot therefore involve any ^-function singularity: they can at most suffer an //-function discontinuity. The Weyl scalar Wt and the Maxwell scalar &20 involve second derivatives only in the combinations, (3^ —3e)2(f and (51ir — de)(dlir + dg)(o. By equations (212) these scalars cannot also involve any ^-function singularity. However, lF0 involves the term
252
S. Chandrasekhar and B. C. Xanthopoulos
Therefore, by equations (207) and (212) it must have a d-function singularity given by '
2U2(\-\s\2)
"
S(u)
{0^v<
1).
(214)
l + f^-K — 0
Similarly, along (v = 0,0^u
and
0 - ^ - > + O,
(215)
we shall have a ^-function singularity in W given by JL.
2U2{i-\S
S(v)
|2).
(0<M<1).
(216)
For the case q = 0, the factors of the (^-functions in (214) and (216) can be explicitly evaluated by making use of equations (182) and (185); we find that the ^-function singularity in WQ, on the null boundary separating regions I and I I , is given by a-b(l-v)2 (217) 2 2(l-v )[a + b{l-v2)f •
*
(
«
)
;
and the complete solution for !P0 in regions I and I I can be written in the form a-b{l-v)2 d(u) 2(l-v )[a + b(l-v2)]i
V™H{u) +
2
(218)
where the superscript (I) indicates that it represents the solution for region I given in equations (186). Similarly, we may write for *F2, $ 0 0 , and
<&$H(u),
and
(219)
and, as we have directly verified, !f4 and
S(v) =
2 -1(1- -a)a 8(v),
(220)
along the null boundary, v = 0 and u < 0, separating regions II and IV. In addition Wi also experiences an //-function discontinuity of amount specified in equation (200). The Maxwell scalar, <£22, however, experiences only an //-function discontinuity of amount also specified in equation (200). Finally, the singularity in ¥0 and the discontinuity in $ 0 0 along the null boundary, u = 0 and v < 0, separating regions I I I and IV, is the same as for Yt and
On colliding waves in the Einstein-Maxwell theory 11.
253
T H E C O N S I S T E N C Y OF T H E D E R I V E D S O L U T I O N W I T H T H E
R E Q U I R E M E N T S OF M A X W E L L ' S E Q U A T I O N S AND THE A B S E N C E OF CURRENT SHEETS
The characterization of the discontinuities in the Maxwell scalars in § 10 leaves open the question whether the resulting discontinuities in the Maxwell field across the null boundaries are consistent with the requirements of Maxwell's equations. The requirements are of two kinds: the mathematical requirement that the 'jump-conditions' across these surfaces of discontinuity, that follow from Maxwell's equation, ^ , ^ = 0, (221) are satisfied; and the physical requirement that no current sheets occur on the surfaces of discontinuity in accord with the equation Jt = WFtj = 0.
(222)
(The subscripts, i,j, k, etc. (from the latter part of the alphabet) are tensor indices (in conformity with the notation in M.T.).) The reason for this latter requirement is that, since Maxwell's equations, without charges and currents, have been satisfied in the rest of space-time, the occurrence of current sheets, on the null boundaries separating the different regions, would imply the creation of charges from nowhere; and this we cannot tolerate! As we have seen in § 10, the Maxwell tensor Fy suffers //-function discontinuities across the null boundaries separating the different regions. Following Pirani (1964; see also O'Brien & Synge 1952 and Bell & Szekeres 1974) we shall express this fact that, across the boundary separating regions I and I I , and I I I and IV, for example, Ftj has the form Fi^fv + r/rtjHiu), (223) where \jr^ is that part of Fy that discontinuously becomes zero for u < 0 and fy that part which remains continuous at u = 0. We have shown that across this boundary u = 0, 0O suffers an //-function discontinuity. Therefore, only
t224)
Therefore, Maxwell's equation (221) now requires /[«, k] + fin, iti H(u) + fW
u
,fcis(u) = °-
(225)
Since Maxwell's equations have been satisfied in all four regions I, I I , I I I , and IV, f[ij,fc]+ f[i], iti = °
(« > 0; in regions I and III)
and
(226) 0"[y,fcl
=
0
(M < 0; in regions I I and IV);
therefore, the required 'jump-condition', lfrwutk]
on u = 0, is = 0.
(227)
254
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
Now, for the null tetrad chosen (cf. equation (124)) l.dx1 = - ^ - U(difr-d8)
= dv
and nt dxl = — - U{drjr + dd) = du. V2
(228)
Therefore, M 4 is a vector parallel to ni,
(229)
v t is a vector parallel to li
(230)
and
Returning to the jump-condition (227) we have the requirement f w » * ] = 0; (231) and this requirement is manifestly met by i/r^ given by (224). Turning next to the condition t h a t must be met across the boundary separating regions I and I I I , and II and IV, we must now require ^ [ 0 « * l = O,
(232)
where fa = - (02 mt +
(233)
since it is <j>2 that experiences an //-function discontinuity (in v) while 0O is continuous. By (230), the jump condition is equivalent to the requirement
*W*i = °;
< 234 )
and this requirement is, again, met by ^ y given by (233). Turning next to the examination of whether or not current sheets are present at the interfaces separating the different regions, we shall first write down Maxwell's equations, including currents, in a Newman-Penrose formalism. I n a general tetrad formalism, the current vector J ( a ) is given by (235) J(a) = nimc)F{a)m{e)t where the vertical rule signifies 'intrinsic differentiation' (as defined, for example, in M.T., p. 37). In terms of the Maxwell scalars <j)0, <j)v and (f>2, equation (235) gives J(D = (0i + 0 ? ) | i - 0 o | 4 - 0 ? | 3 > J(2) = 0 2 | 8 + 0 * | 4 - ( 0 1 + 0*)|2> J
(3) = 011 3 + 02* 1 - 00 I 2 - 01* 3 = JW •
(236)
Substituting for the intrinsic derivatives their expressions in terms of directional derivatives and spin coefficients, we obtain J (1) = [D^>1 — S*(p0+ (2a — n)
(237) (238)
^(3) = [^0i-^0o + ( 2 T-/*)0o- 27 "0i + cr02] + [Dcf>2-S*
(239)
On colliding waves in the Einstein-Maxwell theory
255
Since we have satisfied all of Maxwell's equations, without currents, in regions I, I I , I I I , and IV except on the boundaries separating them, only current sheets can be present (if at all) at the interfaces as ^-function distributions. By our discussion in §10, ^-function distributions can arise only if the expressions (237)-(239) for the currents involve the second derivatives 3 2 /9« 2 or 3 2 /cV with respect to u and v; but no such singularity will arise by the presence of the mixed derivative d2/dudv. Since none of the expressions (127) for the spin coefficients involve any second derivative of the metric coefficients, their appearance in equations (237)—(239) cannot lead to any ^-function distribution. Also, since
(240)
To show t h a t J ( 3 ) and J (4) also vanish, we must examine the limits lim M->O
[Dtf-Jfa].
(241)
or D - > O
Before we examine the limits (241), we first remark t h a t the expressions for the Maxwell scalars given in equations (133)-(135) determine only #oo = l0o I2.
#22 = 10212 and
<£2O = 0 2 0*-
(242)
Therefore, if we write "o
l^ole 1 ^
and
(243)
the known solutions for
D02 = -A0 o +(p-2e)0 2 and J
(244)
t h a t follow from M.T., p. 52, equations (331) and (332) for the problem on hand. Rewriting the foregoing equations in the forms iD£ 2 = - D In |
10:2 1
^ 0 = - J l n | ^ 0 | - ( A t - 2 y ) + O-{|a|e1
(245)
we observe that all the quantities on the right-hand sides of these equations are known; they can, therefore, be used to determine £0 and £2, separately.
256
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
Returning to the expression (241), we have to examine lim
[e-^(D\^\-i\^2\DQ-e^(A\^0\
+ i\^>0\JQ], (246)
A careful examination of the quantity in square brackets in (246), with the known expressions for &00 and &22, given in equations (133) and (135), and equations (245) governing £2 and £0, shows that it involves second derivatives only in the combination
But as we have seen, mixed derivatives cannot give rise to ^-function distributions. We therefore conclude that Jw = J(i) = 0
on
(M
=0
or v = 0).
(248)
The demonstration is now complete that the discontinuities in the Maxwell field, t h a t arise from the metric we have derived, satisfy the necessary jump conditions and do not also require any current sheets a t the null interfaces between the different regions. The physical interpretation of these facts must be that the discontinuities in the Maxwell field are in some sense ' supported' by the impulsive gravitational waves. 12.
CONCLUDING
REMARKS
In view of the complexity of the analysis and of the many issues that had to be resolved, we shall, in this concluding section, restate the essential physical and mathematical content of the problem t h a t has been studied. The problem originated in an observation of Penrose that colliding waves in general relativity may provide further examples of situations which, generically, may lead to space—time singularities; and a further remark of his in a personal communication t h a t he had "thought a little about Einstein-Maxwell impulsive waves [and] the curious feature t h a t in order t h a t the Ricci curvatures be a ^-function, the Maxwell field must be a kind of' square-root of the ^-function'; and it was never clear to me how to make sense of such things''. This paper is concerned with the extension of earlier treatments of the problem of colliding impulsive pure gravitational waves to the case when gravitational and electromagnetic waves are coupled via the Einstein—Maxwell equations; but the restriction to plane-fronted impulsive waves and space—times with two commuting space-like Killing vectors is retained. An exact solution of the Einstein—Maxwell equations is obtained that is a natural generalization (in a sense we shall presently clarify) of the solutions of Khan & Penrose and of Nutku & Halil. The physical content of the solution is best described by the space-time diagram illustrated in figure 2 and in the detailed legend given below it. I t is especially noteworthy that the discontinuities in the Maxwell field, which occur along the null boundaries separating the different regions, are consistent with the 'jump conditions' t h a t must be satisfied and do not also require the presence of current sheets as ^-function distributions a t the
On colliding waves in the Einstein-Maxwell theory
257
aft er the collision
V*~t(v) ¥". ~H(v),<j>2- -H(v) Vo, <j>a continuous
tf/1 = V 3 =0 1 =O everywhere FIGURE 2. The space-time diagram for two colliding plane impulsive gravitational waves. The colliding waves are propagated along the null directions, u and v. The time coordinate is along the vertical, and the spatial direction of propagation (in space) is along the horizontal. The plane of the wavefronts (on which the geometry is invariant) is orthogonal to the plane of the diagram. The instant of the collision is at the origin of the («, ^-coordinates. The flat portion of the space-time, prior to the arrival of either wave, is region IV. These waves produce a spray of gravitational and electromagnetic radiation which fills regions II and I I I ; and the region of the space—time in which the waves scatter off each other and focus is region I. The result of the collision is the development of a curvature space-time singularity on w2 + v2 = 1; and v = 1 or u = 1 are space—time singularities for observers who do not observe the collision. Singular behaviours in the Weyl scalars, !P4, W2, and !f0 (as Dirac ^-functions or Heaviside (H) step-functions, or both, expressing the shock-wave character of the colliding electromagnetic waves) occur, as indicated, along the null boundaries separating the different regions.
interfaces. T h e a n a l y s i s d i d n o t r e q u i r e t h e i n t r o d u c t i o n of a n y i n a d m i s s i b l e n o t i o n such as t h e ' s q u a r e r o o t of a ^-function'. T h e t w o null 3-surfaces, (v = I, — oo < w < 0) a n d (u = 1, — oo < v < 0), r e p r e s e n t e d b y lines in figure 2, a r e , a s i n d i c a t e d , s p a c e - t i m e singularities. I n t h e case of t h e v a c u u m (when regions I I a n d I I I a r e flat) t h e origin a n d t h e n a t u r e of t h e s e singularities h a v e been a m p l y discussed b y M a t z n e r & Tipler (1984). T h a t t h e y c o n t i n u e t o be space—time singularities in o u r p r e s e n t c o n t e x t (when regions I I a n d I I I are n o longer flat) can b e visualized m o s t simply b y t r a c i n g t h e time-like a n d t h e null geodesies t h a t e n t e r regions I I a n d I I I b y crossing t h e c o o r d i n a t e - a x e s v = 0 or u = 0 a t some n e g a t i v e v a l u e s of u or v. T h e c o n s t r u c t i o n of t h e geodesies in regions I I a n d I I I , in which t h e m e t r i c coefficients are functions of o n l y one of t h e t w o n u l l - c o o r d i n a t e s , v or u, is simple
258
S. C h a n d r a s e k h a r a n d B . C. X a n t h o p o u l o s
since the space—time allows three conserved momenta, px\, px2, and pu or pv; and the energy integral can be immediately integrated. Thus, in the case q = 0, the three conserved momenta in region I I are (cf. equation (150)) pxl = e2^ (x1)' = a,
pxz = e 2 ^ (a;2)- = /?,
and
pu = \ e2vv = y,
(249)
where a, /?, and y are constants and the dots denote differentiation with respect to some affine parameter; and the energy integral gives e^uv-e2^
[(a; 1 )-] 2 -e 2 ^ [(a;2)-]2 = E,
(250)
where E > 0 for time-like geodesies and E = 0 for null geodesies. From the foregoing four conservation laws, we readily obtain the equation du 4y 2 —- = a 2 e - 2 ^ + 2 " + /tf2 e-2^
+ 2v
+ Ee2v.
(251)
With e2", e2'*', and e 2 ^ given in equations (194), equation (251) can be integrated to give Cv dv v Cv 2 2 2 i 2 y (u + c) = a \ [a + b(l-v) ] ——+ j3 + E\ [a + b(l-v)2]2 dv, (252) v v Jo (1 ) 1 ~^~ Jo where u = — c is the point on the negative w-axis at which the geodesic crosses into region II. Accordingly c > 0. We now observe that the right-hand side of equation (252) is positive-definite (for v > 0); and further that the first term diverges to + oo for v -> 1 — 0 with the behaviour a2a4-^-. (253) 1 —v Therefore, no time-like or null geodesic with a # 0 can avoid crossing the v-axis into region I for some 0 < v < 1 and hitting the singularity at u2 + v2 = 1. But for a = 0, there exist geodesies, both time-like and null, which can avoid crossing into region I : these geodesies hit the singularity at v = 1. The situation is exactly the same as in the vacuum; and we may pictorially represent the space—time in the manner of Penrose (see figure 7 in Matzner & Tipler 1984). On the mathematical side, the procedure that was followed in obtaining the solution departs from earlier attempts in which solutions were sought for the metric, for regions of the space-time in which the colliding waves scatter off each other, t h a t will be compatible with carefully formulated initial conditions on characteristic surfaces (cf. Bell & Szekeres 1974). We have started, instead, at the 'opposite end' by first selecting (ad hocl) the metric, for the region of interaction, by an 'aesthetic' criterion and then systematically extending it across null boundaries, where coordinate singularities occur, with the aid of the Heaviside step-function. The fact that by this procedure, we have been able to obtain a physically consistent solution, is to be credited to the criterion for the choice of the solution in the interacting region. In the choice, we were guided by the observation that both the Khan-Penrose and the Nutku-Halil solutions follow from the same simplest solution (cf. Paper I, equation (55)), E = p?j + iq/A,
(p2+q2=l),
(254)
On colliding waves in the Einstein-Maxwell theory
259
of the Ernst equation (for space-times with two commuting space-like Killing vectors) as the Schwarzschild and the Kerr solution do from, formally, the same Ernst equation (now for space—times with one space-like and one time-like Killing vector). We, therefore, sought (and found) for the region of interaction, a solution of the Einstein-Maxwell equations that can be obtained from the Nutku-Halil solution by following the same procedure by which one obtains the Kerr-Newman solution, for the charged black-hole, from the Kerr solution; and, in both cases, the same solution of (formally) the same coupled Ernst equations is involved. The fact that this 'inverted' procedure leads to a physically consistent solution is a further manifestation of the firm aesthetic base of the general theory of relativity. While we have obtained in this paper the solution for the general case when the polarizations of the colliding waves are not parallel, we have analysed in full detail (in §9) only the case when the polarizations are parallel. In a later paper, we intend to analyse in equal detail the more general case of non-parallel polarizations. But there are other directions in which the methods of this paper and of paper I can be extended; for example, to the problem of colliding gravitational waves coupled with hydrodynamic shocks (a problem which we are presently investigating). We are grateful to Professor R. Geroch for some useful discussions. The research reported in this paper has, in part, been supported by grants from the National Science Foundation under grant P H Y 80-26043 with the University of Chicago. B. C. Xanthopoulos's tenure at the University of Chicago during June—September 1984 was supported by a grant from the Senior Scholar Fulbright Program; he also wishes to express his thanks to the Relativity Group at the Fermi Institute of the University of Chicago for their hospitality.
REFERENCES
Bell, P. & Szekeres, P. 1974 Gen. Rel. Grav. 5, 275. Chandrasekhar, S. 1983 The mathematical theory of black holes (referred to as M.T.). Oxford: Clarendon Press. Chandrasekhar, S. & Ferrari, V. 1984 Proc. R. 80c. Lond. A 396, 55-74. Ehlers, J. 1957 Dissertation. University of Hamburg. Harrison, B. K. 1968 J. math. Phys. 9, 1744. Khan, K. & Penrose, R. 1971 Nature, Lond. 229, 185. Matzner, R. A. & Tipler, F. J. 1984 Phys. Rev. D 29, 1575. Misner.C. W.,Thorne,K. S. & Wheeler, J. W. 1970 Gravitation. San Francisco: W. H. Freeman &Co. Nutku, Y. & Halil, M. 1977 Phys. Rev. Lett. 39, 1379. O'Brien, S. & Synge, J. L. 1952 Communs Dubl. Inst. advd. Stud. A 9. Pirani, F. A. E. 1964 In Brandeis Lectures on General Relativity, pp. 269-275. New Jersey: Prentice-Hall. Szekeres, P. 1970 Nature, Lond. 228, 1183. Szekeres, P. 1972 J. math. Phys. 13, 286.
997
IX. The Nonradial Oscillations of Stars in General Relativity The papers in this part concern the work during the last five years of Chandra's life. During this period, notably after completing the development of the theory of colliding waves, Chandra often said he wanted to quit serious science. This did not happen. On the contrary, while occupied with completing his opus on Newton's Principia, he continued to work with Valeria Ferrari, mostly on the nonradial oscillations of stars. As Valeria Ferrari says in her foreword to Vol. 7 of Selected Papers, "The idea underlying the theory of non-radial oscillations of stars is of a very fundamental nature. General relativity teaches us that any distribution of matter generates a space-time curvature, that is, a potential well." This suggests that the scattering of gravitational waves in this potential should provide a suitable framework for studying the oscillations of stars. The first paper in this part confirms this idea. Chandra derives a flux integral from the linearized equations governing the axisymmetric perturbations of static space-times with a Maxwell field or a perfect fluid as sources. This ensures the conservation of energy in the attendant scattering of radiation as well as in the transformation of one kind of radiation into another. The second paper develops a complete theory of the nonradial oscillations of a static spherically symmetric distribution of matter. The starting point, as stated above by Ferrari, is the assumption that the oscillations are excited by the incident gravitational waves. The equations governing the perturbations are formulated in such a way that the equations governing the perturbations of the space-time metric are decoupled from those governing the hydrodynamical variables. Because of this decoupling, the problem of determining the complex characteristic frequencies of the quasinormal modes of the nonradial oscillations is simplified. It reduces to a problem in the scattering of incident gravitational waves by the curvature of space-time with the matter content of the source acting as a potential. In Paper 3, Regge's theory of potential scattering in quantum mechanics is suitably adapted and generalized to the determination of the flow of energy in the form of gravitational radiation through a star in nonradial oscillations. Chandra was extremely happy about this work, as it was a totally unanticipated application of theory of Regge poles that was developed for an entirely different set of problems. The final paper deals with the nonradial oscillations of a Newtonian star, i.e. a star built in the Newtonian framework. It is shown how the oscillations can be treated fully relativistically by considering the scattering of the gravitational radiation by the shallow, curved space-time of the star.
999
IX. The Nonradial Oscillations of Stars in General Relativity 1.
The Flux Integral for Axisymmetric Perturbations of Static Space-Times With V. Ferrari; Proceedings of the Royal Society A428 (1990): 325-49
1000
2.
On the Non-radial Oscillations of a Star With V. Ferrari; Proceedings of the Royal Society A432 (1991): 247-79
1025
3.
On the Non-radial Oscillations of a Star. IV: An Application of the Theory of Regge Poles With V. Ferrari; Proceedings of the Royal Society A437 (1992): 133-49
1058
On the Non-radial Oscillations of a Star. V: A Fully Relativistic Treatment of a Newtonian Star With V. Ferrari; Proceedings of the Royal Society A450 (1995): 463-75
1075
4.
Proc. R. Soc. Lond. A 428, 325-349 (1990) Printed in Great Britain
The flux integral for axisymmetric perturbations of static space-times B Y S U B R A H M A N Y A N CHANDRASEKHAR1,
F.R.S., A N D V. FERRARI 2
1
2
ICRA
University of Chicago, Chicago, Illinois 60637, U.S.A. (International Centre for Relativistic Astrophysics), Dipartimento di Fisica 'G. Marconi', Universita di Roma, Rome, Italy (Received 21 August 1989)
The axisymmetric perturbations of static space-times with prevailing sources (a Maxwell field or a perfect fluid) are considered; and it is shown how a flux integral can be derived directly from the relevant linearized equations. The flux integral ensures the conservation of energy in the attendant scattering of radiation and the sometimes accompanying transformation of one kind of radiation into another. The flux integral derived for perturbed Einstein-Maxwell space-times will be particularly useful in this latter context (as in the scattering of radiation by two extreme Reissner-Nordstrom black-holes) and in the setting up of a scattering matrix. And the flux integral derived for a space-time with a perfect-fluid source will be directly applicable to the problem of the nonradial oscillations of a star with accompanying emission of gravitational radiation and enable its reformulation as a problem in scattering theory.
1. I N T R O D U C T I O N
I t could perhaps be stated that the principal object in studying the perturbations of a static or a stationary space-time is to understand the attendant scattering of radiation and the sometimes accompanying transformation of one kind of radiation into another in terms of the unitarity and the time-reversibility of a suitably defined scattering matrix. The scattering matrix essentially expresses the conservation of energy in the scattering process that takes place. In the case of the reflexion and absorption of radiation by the Schwarzschild, the Reissner-Nordstrom, and the Kerr black-holes, the understanding derives from the existence of one or more Wronskians whose constancy assures the sum of the reflexion and the absorption coefficients to be unity and thus the conservation of energy. In the examples of the Schwarzschild and the Reissner-Nordstrom black-holes, the constancy of the Wronskians expresses no more than the elementary fact that the Wronskian of a solution and its complex conjugate of a one-dimensional Schrodinger equation with a real potential, is a constant. But the deeper significance of this simple result is that it is a very special case of a far more general fact, that for perturbations of static or stationary spacetimes, there exists a three-dimensional vector, E, whose divergence vanishes; and that, therefore, by Gauss's theorem, the existence of a flux integral follows (cf. equations (52) and (53) below). Thus, in the problem recently studied [ 325 ]
326
S. Chandrasekhar and V. Ferrari
(Chandrasekhar 1989) of the scattering and the accompanying transformation of gravitational radiation into electromagnetic radiation (and conversely) by the axisymmetric axial modes of perturbation of two extreme Reissner-Nordstrom black-holes, it was possible to set up a scattering matrix only by making use of the flux-integral (Chandrasekhar 1989, equation (56)) involving certain Wronskians. (In the Appendix, we provide a derivation of this flux integral, more generally than in the context of the particular problem considered.) For treating the scattering of radiation by the same two Reissner-Nordstrom black-holes (or, more generally, by any static, axisymmetric Einstein-Maxwell space-time) by the alternate polar modes of perturbation, an analogous flux integral of the then applicable linearized equations is a necessary prerequisite. Since the number of coupled equations governing the polar perturbations is eight, in contrast to two for axial perturbations, the derivation of the necessary flux integral is neither simple nor straightforward: it requires, in particular, the writing of the linearized equations in a form that manifests the inner relationships among them. The existence and the exhibition of these relationships is one of the prime motivations of this paper. With respect to the flux integral that equations, governing the first-order perturbations of static and stationary space-times of general relativity, allow, some of those whom we consulted (e.g. A. Ashtekar, J. Friedman, R. Sorkin and R. Wald) were aware t h a t the existence of such an integral can indeed be inferred on general grounds (cf. J. Friedman 1978). More recently, Lee & Wald (1990) have established that the existence of such a conserved flux (or, a simplectic current in their terminology) can always be inferred for any field theory derived from a suitably defined Lagrangian action. Also, Burnett & Wald (1990) have shown how an explicit expression for this conserved simplectic-current can be derived for a perturbed space-time of general relativity. The expressions they have derived do indeed reduce to the ones derived in this paper when simplified for the contexts that are considered. If one can thus obtain the flux integrals derived in this paper from a general theory, one may well question the usefulness of obtaining them ab initio from the equations applicable to the separate cases. Although the answer to this question depends on one's point of view, the following facts may be relevant. The reduction of Burnett & Wald's general expression to the forms derived in this paper (equations (50)-(52), (63)-(65), and (132)-(134)) requires the explicit use of what we have called the initial-value equations (equations (6) and (7) and their special forms). These equations, while they are not included in the general theory in a natural way, they are essential ingredients in the ab initio derivations. Besides, it will appear t h a t the inclusion in the flux integral, of terms that derive from different sources (e.g. a Maxwell-field or a perfect fluid), is simple and straightforward, once the expression for the vacuum has been obtained in its simplest form. Apart from these factors, for us the overriding motivation derives from the insight one gains in the intimate relationships that exist among the equations themselves, a motivation that may not have a general appeal. The plan of the paper is the following. In §2 the mathematical background of the problems to which the paper is addressed is described; and the basic equations
Flux integral for axisymmetric perturbations
327
are assembled in forms t h a t we shall need. The paper then divides into two parts. In Part I, the equations are specialized appropriately for Einstein-Maxwell spacetimes ; and the flux integrals for the vacuum and for the Einstein-Maxwell spacetimes are derived (equations (50)-(52) and (63)-(65)). In Part I I , the case when the prevailing source is a perfect fluid is considered. The analysis in this part is directly applicable to the problem of non-radial oscillations of a static spherical star; and the flux integral t h a t is derived (equations (132)-(134)) reduces the associated problem - the emission of gravitational radiation and the attendant damping of the oscillations - to one in scattering theory that can be described by a scattering matrix.
2. T H E F O R M U L A T I O N O F T H E P R O B L E M A N D T H E B A S I C
EQUATIONS
We consider a static axisymmetric space-time. The metric of such a space-time can be written in the form, ds2 = e 2 "(d<) 2 -e^(d
(1)
where v, i/r, /i2, and fi3 are functions of the spatial coordinates, x2 and x 3 , only. (In accordance with the attitude expressed in Chandrasekhar (1988), in writing the metric in the form (1) we have not made use of the available gauge freedom to 'simplify' the form of the metric as is customary.) We consider axisymmetric perturbations of the space-time described by the metric (1) with sources which we leave unspecified for the present. As has been explained in The mathematical theory of black holes (Chandrasekhar 1983, §§24 and 42; this book will be referred to hereafter as M.T.); and in Chandrasekhar (1989, §3) the axisymmetric perturbations of a static space-time fall into two noncombining groups: the axial and the polar. The metric of the axially perturbed space-time is of the form, ds 2 =
e2"(dt)2-e2*(d(p-a)dt-q2dx2-q3dx3)2 -e2^(dx2)2-e2^(dx3)2,
(2)
where v, ijr, ji2, and /i3 retain the values they have in the static space-time, while w, q2, and q3, describing the perturbation, are functions of a;2, x3, and t. Such axial perturbations of an Einstein-Maxwell space-time, in the context of the two-centre problem, have been considered in Chandrasekhar (1989); and the flux integral appropriate to that problem was derived (Chandrasekhar 1989, equation (56)). The derivation of the flux integral, without the specialization to the two-centre problem, follows along the same lines and is given in the Appendix. In the main text of the paper, we shall be concerned only with the more difficult problem of the polar perturbations. Under such perturbations, the metric retains the same form (1) but the functions v, 1/r, ji2, and /JL3 suffer infinitesimal increments, bv, 8 ^ , 8/t2, and 6fi3, respectively. We shall further suppose - no loss of generality is implied by this supposition - that these perturbations (and others) have a common time dependence given by wt i%\
328
S. Chandrasekhar and V. Ferrari
where a is a constant. In view of the applications contemplated in Part II, a is allowed to be complex though we shall normally restrict it to be real (unless stated otherwise). The equations governing the perturbations, bv, h\jr, 8/i2, and d/i3 can be readily written down by appropriately linearizing the equations listed mM.T. (pages 141 and 142, equations (4) and (5)). In writing down the various linearized equations, we shall adopt the following definitions: a 0 = r/r + a
/i2+/i3-v;
2 = ft + v+/i3-[i2,
fti
cc3 = \/r + v + fi2-u3,
= i/r+/i3 + v;
fi3
= ifr + {i2 + v;
A
2 = f ,2/ i 3,2+/ M 3,2 l ',2 + l ',2^,2>
A
ft,3ll2.3+fl2,aV,a
3 =
+ V,3ir,3>
SA2 = i/r 2b(^3 + v)2+/i325{v 8
^ 2 = e**A2 \ X
3 =
^Ai^
+ }Jf) i + v<25(i/f + / t 3 ) 2 ;
(4)
^ 3 = V ir ,3 5 (/*2 + 1 '),3+/ W 2,3 8 ( 1 ; + ^ ) , 3 + l ' , 3 8 ( ^ + / « 2 ) , 3 ;
8X2 = ea*(A2 8a 2 + &42),
8X 3 = e"*(A3 8a 3 + 5A3);
8
/ 2 = ^ , 2 8 ( " + / M 3 ) + / * 3 , 2 8 ( ^ + ^) + »',2 8 (V ! r +/*3);
8
/ 3 = V i f ,3 8 ( , '+/*2)+/*2,3 8 (V ! r + l , ) + , ' , 3 8 ( ^ + / t 2 ) -
With these definitions, the equations governing the perturbations are: [ e * ( 8 ^ 2 + f,
2 5a s )], 2 +
[e"»(8^,, + i/r 3 8a,)], 3
2
+
2
+ ?>(i/r +
v+/i2+/i3)R11]v/-g;
+ [e^Sc. 3 + v , 8 a 3 ) ] | S
2 a
+ o- e «8(f +/i2 + /i3) = +[8R0O + 5{ft + i' + [ea3(8/i2 3+/i2,38a3)]3
/i2+/J,3)R0O]V-g
(5ii)
+ [e^(8/? 22 + /? 2 , 2 8a 2 )] > 2 - 2 8 X ,
2 a
+ o- e »8/M2 = - [ 8 i ? 2 2 + 8 ( ^ + ^ + ^ 2 + ^ 3 ) ^ 2 2 ] [eaM8/«3,2 + /*3,2 Sa 2 )], 2 + [ea3(S/?3,3 + A, 3 8 a 3 ) ] , 3 2 a
+cr e »8/*3 = - [ 8 ^ 3 3 + 8 ( ^ + ^ + ^ 2 + ^ 3 ) ^ 3 3 ] a
(5i)
{e '[?>{ft+ti3)i2 + (ijr + ti3)
V-g; 28
(5iii)
-^ 3
V-g;
(5iv)
28cc2]}2-&X2
+ {e"»[8(^+/«,), 3 + ^ + / » , ) , 8 8a,]}, 8 - & X 8 = -[^G00 + d(iJf + v + /i2+/i3)G00]V-g-
(5v)
In addition, we have the initial-value equations, eai?>(f+fi3)2
+ if2?,(i/f-/i2)+/i325(/i3-u2)-v25(ir
+ /i3)]i0 = -2e^ + 2 " + "38 J R n
(6)
e a 3 [8(^+/*2),3 + ^ , 3 8 ( ^ - / « 3 ) + / t t 2 , 3 8 ( / M 2 - / * 3 ) - l ' , 3 8 ( f +/*«)], 0 = -2e^ + 2 " + ^8i? 0 3 .
(7)
Flux integral for axisymmetric perturbations
329
We call these the initial-value equations since, as we shall find in §§4 and 11, 8i?02 and &R03 are in turn time derivatives of functions describing the perturbations of the prevailing sources. The equations accordingly allow immediate integration with respect to time. Alternative forms of equations (6) and (7) which we shall find useful are: e^[8(^ + /t,) i 2 + ( ^ + ^ 3 ) f 2 8 a 2 ] i 0 - ( e - ^ / 2 ) i 0 = -2e*+*"+»°5R02,
(8)
e ^ S ^ + ^ . + ^ + ^ ^ S a . ] , 0He"'8/8).0 = - 2 e W * S i ? 0 3 .
(9)
P A R T I. T H E F L U X I N T E G R A L F O R P O L A R P E R T U R B A T I O N S OF STATIC E I N S T E I N - V A C U U M AND EINSTEIN-MAXWELL SPACE-TIMES 3. L I N E A R I Z E D M A X W E L L ' S
EQUATIONS
In static Einstein-Maxwell space-times, the only non-vanishing components of the Maxwell tensor can be F02 and F03 related by the equation, ( e v + ^ o 2 ) , 3 - ( e ^ ^ o 3 ) , 2 = 0.
(10)
For the polar perturbations of the space-time, we are presently considering, besides the first-order changes, 5F02 and Si^g in F02 and F03, we must also allow for a first-order Si^g induced by the perturbations. The equations governing these perturbations follow directly from M.T., p. 220, equations (118). Thus, letting e ^ i ^ = krF,
(11)
in accordance with our general assumption (3), that all quantities describing the perturbation have the common time-dependent factor e1
find:
5F M = -FmW+pt)
+ e-*-*Yt,
(12)
5^.2 = -FotSW+ra)-e-*-*Ya, {e"+*>[e-^Y2
+
(13)
F035(v+p3-ir-p2)]},2
+ {e"+^[e-^-^Y3-F028(u+fi2-}/f-fi3)]}i3
= -
4. L I N E A R I Z E D E I N S T E I N ' S
(14)
EQUATIONS
For the static Einstein-Maxwell space-time considered, the non-vanishing components of the Ricci tensor (by Einstein's equations) are: •"00
=
-"ll
=
-^02 "'"-''03 >
-"22
=
—
-"33
=
^03
—
^02 >
(1°)
and for the polar perturbations of the space-time: 8R00 = + 5Rn = 2F03 8F03 + 2F02 8F02;) (16) Si?,, ''22 = - 8fl M = 2Fn„ BFn, - 2Fno 8Fm 8RO2 = +2F036F23
and
5R03 =-2F02bF23.
(17)
330
S. C h a n d r a s e k h a r a n d V . F e r r a r i
The linearized Einstein's equations for the problem on hand now follow from equations (5i-v) by substituting from (16) for the terms on the right-hand sides of these equations. With the further definitions,
|
(18)
V3 = 2e-+»Fn Y2,J ft,5a3)],,
= -(0% + 4>a) + (Yt-¥a)-tr'el>*#;
(19i)
[ea*(8v, 2 + v , 8a 2 )], 2 + [ea3(5i^ , + v , 6a,)] f , = +(
(19ii)
[e a 3(8/i 2|3 +/i i! , 3 5a3)] i 3-l-[e a H5/3 2 , 2 +A 2i2 8a 2 )] >2 -25Z 2 = +((P2-03)-(!?2+!f3)-(r2e^2; a2
(19iii)
a3
[e (Oytt 3 , 2 +/*3,2 K ) ] , 2 + [e (5/?3, a+^3,38^3)1,3-25X3 = -(
(19iv)
{e^[8(^+/i3), 2 + (^+ / tt3), 2 8a 2 ]}, 2 + {e 0I '[5(^+ / « 2 ) 3 + (f+^ 2 ), 3 8a 3 ]},3 = (8Z 2 + 8Z 3 )-(
(19v)
Since, in accordance with equations (11) and (17), 8fl M = 2 i « r e - ^ ^ 0 8 7 )
8i?03 = - 2i
(20)
the initial-value equations (8) and (9) now give e«°[8(^+/g,2 + ( f + / t , ) i , 8 a , ] - e - 5 / 2 = - 2 e " + ^ 0 3 7 ,
(21)
e^[8(^+/»,),,+ (^+/*,),, 8 a , ] - e - 8 / , = + 2 e ^ ^ 0 2 y .
(22)
From these equations, it follows t h a t {e".[8tyr+/*,),, + ( ^ + / t 8 ) , , 5a 2 ]},, - (e-8/,),, + {e"3[8(^+ /t2 ),3 + (^+ ytt2 ),38 a3 ]},3-(e a 38/ 3 ),3 = - 2 ( e " + ^ 0 3 F),, + 2 ( e " + ^ 0 2 F),, = W2 - ¥,. Together with equation (19v), we obtain the useful identity: (e-5/,),, + (e*8/,),, = 6X2 + SX, - («P, +
(23) (24)
Alternative forms of equations (19iii, iv) Returning to equation (19hi) and remembering the definition of /?2 (in equations (4)) we can write, [ea3(8/t2,3 +/* 2 , 3 8a 3 )], 3 + {e a H8(^+/* 3 ), 2 + (^ +/^a), 2 8a 2 ]}, 2 + [e*(Si>i, + i; i ,8a,)] i 2 = 2SX2 + (
¥,)-o*
d-8/i,.
(25)
Flux integral for axisymmetric perturbations
331
Replacing the second term, {} 2, on the left-hand side of equation (25) by the term, {ea3[2->3]} 3, with the aid of equation (19v), we obtain: [ e ^ S ^ a + ^ a 8 a 3 ) ] i 3 - { e ^ [ 5 ( ^ + / t 2 ) i 3 + (^+^2) 3 8a 3 ]} | 3 + [ea'(5vt 2 + v >2 8a 2 )] i2 = 8 X , - 8 X , + 2 # 2 - 2 5 ^ - t r ^ S / v
(26)
On the other hand, by the initial-value equation (22), -{e a »[8(^ + /t 2 ) > , + ( ^ + ^ 2 ) i , 8 a 8 ] } i , = -(e-»8/, + 2 e ^ ^ 0 2 y ) , s = -(e a 3 5/3),3-2(e^ J F 0 2 ) ) 3 Y- W2.
(27)
Inserting this last expression for {} 3 in equation (26), we obtain the required form of equation (19 hi): [e"'(8fi2,3+/i2t38a3)]3-(ea°8f3)t;i
+ [ea°(Si>i2 + vi2Sa2)]t2
= 8 Z 2 - 8 Z 3 + 2 0 2 - W2 + 2 ( e ' + ^ 0 2 ) , 3 F - o - 2 e-8^ 2 .
(28)
By an analogous sequence of transformations, we obtain from equation (19iv): [e^S/^g, 2 + / * 3 , 2 5 a 2 ) ] 2 - ( e ^ 8 / 2 ) i 2 + [6^(8^3 + ^ 3 8a 3 )] > 3 = 8X3 - 5Z 2 + 2 ^ 3 + Y3 - 2(e" + ^ 0 3 ), 2 Y - a* e ^ , .
5. T H E E V A L U A T I O N O F C E R T A I N
(29)
WRONSKIANS
In preparation for deriving the flux integral of the linearized Einstein's and Maxwell's equations, we shall obtain equations for the Wronskians, [A,A*\l=A
(* = 2,3)
(30)
of the metric functions, 8^, 8/t2, 8fi3, and 5v describing the perturbations. The procedure, as we shall apply it to the different equations, will always be the same. Thus, considering equation (19i), we multiply the equation by 8 ^ * and subtract from it the complex conjugate of the equation multiplied by 8^\ By this procedure, we obtain {e* 2 [5^,S^*] 2 } i2 + {e*3[5^,8^*]3},3 + 5 ^ [ ( e ^ 2 5 a 2 ) , 2 + ( e ^ 3 8 a 3 ) , 3 ] - c . c . = - 5 ^ * [ ( ^ 2 + ^ 3 ) - ( ^ - ^ 3 ) ] + c.c., (31) where c.c. denotes the complex conjugate of the expression immediately preceding. An alternative form of equation (31) which we shall find useful is {ea*[8^, 8^*] 2 }, 2 + [e*^, 2 8^*8(^ + ^ 3 - ^ ) - c.c], 2 + {e*3[S^, 8^*] 3 }, 3 + [ e a ^ . 3 5^*8(v +/i 2 - / t 8 ) - c . c ] , , — ( e a ^ 2 8^f*2 8a 2 + e"*3^ 3 8y!r*3 8<x3) + c.c. = -8jfr*[(
(32)
332
S. Chandrasekhar and V. Ferrari
In simplifying equations such as the foregoing, we shall make repeated use of the following elementary relations: A{A+B + C+...)*-c.c.=A(B
+
C+...)*-c.c] \ (33) and AB*-c.c. =-A*B+c.c. J Next, by combining two of the terms in the first line of equation (29) with the aid of the relation, /«3,2oa 2 -5/ 2 = /t3,25(/t3—/*2) —^5r 2 8(y+/* 3 ) —j/ 2 5(^+/* 3 ),
(34)
and applying to the resulting equation the same procedure that we have described in the context of equation (19i), we obtain: {e^^g.S^y^ + S ^ e ^ ^ a S ^ g - ^ ) - ^
2 8(^
+ / i 3 ) - v 2 5 ( f + /t 3 )]}, 2 -c.c.
+ 5/i*[e*°(5^3 + l > i 3 5a 3 )] i 3 -c.c. = 8fi*[203+*F3-2(e^FO3)
2F]-c.c.
+ dju,*(8X3-8X2)-c.c.
(35)
We rewrite the second term in the first line of equation (35) in the manner: {e^8/i*\ji3y 2 S(/*3 - / t 2 ) - ifrt 2 8{v + fi3) - v^ 2 d(ijr+/i3)] = - c.c.}, 2 -ea*8/i*2[/iZ28(/j,3-fi2)-ip= -{6^8/4(^3 2 8/* 2 + f
2?>(}/r+fi3)]
28(V+/J,3)-V
+ c.c.
2&v+i> 2 8 ^ ) - c c . } 2
-e a =8 / «* 2 [/< 3 ] 2 8(/t 3 -/t 2 )-^ 2 8(v + / « 3 ) - v 2 8 ( ^ + / i 3 ) ] + c.c.
(36)
We thus obtain: {e«*[5/i3,8fi*]2} 2-[e"*5/i*(/i32d/i2 -e^8/i*
2[p3i 2 8(^ 3 -n2)
+ iJr 28v+v
2di/r)-c.c.]
2
- xjr 2 8(v +fi3) - v 2 8 ( ^ + ^3)] + c c .
+ b~/i*[e**{?>v 3 + v 3 8 a 3 ) ] ) 3 - c . c . = ^*(6X3-bX2)
+ 5fc*[2
(37)
By adding to equation (37) the equation which follows from it by the transformations, {2-3, Y^-Y, and ¥^-*F}, (38) we obtain the equation, {e**[8/t3,8/t*]2} 2 + {ea3[8/*2, d/i*]a}t 3 + 5fi*[ea"(8v 3 + V)35a3)]3 + b/i*[e^{bPi2 + - [e"*&[i*(/j,3 2 8/t2 + f
2 5v
a
- [e *5fi*(/i2,3 bfi3 + f
3
v2da2)]2-c.c.
+ v 2 8^r) - c c ]
dv + v 3 8^) - cc]
2
3
- eaa8/4* 2[/i3i 2 8(/*3 -n2) - ^ , 2 b~(v+/i3) - vt 2 8(^r+^3)] + c . c -ea3?>/i*3[ji23?,(/j,2-/i3)-^ = 8/4[2
35(V+/A,2)-V
2
35(f+fi2)]
+ c.c.
Y] + 8/**[2
+ (8/*3 - 8 ^ 2 ) * ( 8 Z 2 - 8 Z 2 ) - c.c.
(39)
Flux integral for axisymrnetric perturbations
333
We simplify equation (39) further by making use of the following identity which follows from applying our standard procedure to equation (19v): 8(i/r + fi2 +/is)*{[ea*(Sv<2 + y 2 8a 2 )] t 2 + [e^Su, 3 + v
35a3)]_3}-c.c.
= 8(^+^2+^,)*[(
(40)
Using this identity, we can rewrite the terms in the second line of equation (39), successively, in the manner: 8/t*[ea*(8v 2 + v 2 8a,,)], 2 + S/**[ea3(Sv 3 + v 3 8a3)]3 - c.c. = -{8(f+/* 3 )*[e^(8 l ', 2 + ^ 2 8 a 2 ) ] , 2 + 8(^+ / M 2 )*[e^(8^ 3 + l ' ] 3 8a 3 )] i 3 -c.c.} + 8 ( ^ + ia2 + /t,)*[(
2
-{e a 38(^ + /*2)*[8v 3 + v 3 8 ( f - ^ 3 ) ] - c . c . }
3
+ S(f+ju,3)*2 e"*(8i> 2 + vt 2 8a 2 ) + 8 ( f H - ^ ) ^ e ^ S ^ 3 + v 3 8a 3 ) - c.c. + 8(^+ / « 2 + / { 3)*[(^ 2 + * 3 ) - ( ^ 2 - ^ 3 ) ] - c . o .
(41)
Replacing now the terms in the second line of equation (39) by the second equality in equation (41) and adding the resulting equation to equation (32), we obtain our principal equation:
-{e^8(yjr+/t 3 )*[8i^ 2 + ^ 2 8 ( v - / t 2 ) ] | 2 - c . c . } 2 + 8 ( ^ + ^ 3 ) * e ^ ( 8 ^ 2 + ^ 2 8 a 2 ) - c . c . -{e^8(i/r+^2)*[^+v
+ 6(}/f+/i2)*3ex'(5v:3 + i>
38(v-^3)]3-c.c.}3
36a3)-c.c.
2
+ [e" ^ 2 di/r*8(v+fi3 — /i2) — c.c.] 2 — e^xjr 2 8^*2 8a 2 + c.c. + [e"3^",3 8i/r*8(v+ju2—/t3) — c.c] 3 — e^ifr 3 8^* 3 8a 3 + c.c. - [ea*5/4(/*3 2 8/i2 + i/rt 2 8v + v 2 ?>f) - c.c] a
-[e °b/i*{/u,2 38/i3 + iJr,3§v+v
2
38^r)-cc.] 3
-e^5/i*2yi328(/j,3-fi2)-i/r
28(v+/i3)-i>
28{i/r+/i3)]
-ea'8/i*3[ju,235(/j,2-/i3)-ijr
35(v+fi2)-v3d(}/f
+ c.c.
+/t2)] + c.c.
a
+ 8 ( ^ 3 - / t 2 ) * { e ^ 2 8 ( f + ^) + 8 ^ 2 ] - e ^ 3 8 ( ^ + ^ + 8 ^ 3 ] } - c c . = - 2 8 f *[(
(42)
6. T H E F L U X I N T E G R A L F O R T H E V A C U U M
I t is convenient at this stage to consider equation (42) for the vacuum when there will be no terms on the right-hand side of this equation. The terms in the
334
S. Chandrasekhar and V. Ferrari
resulting ' homogeneous' equation are of two kinds: the ' inside' terms that are enclosed in the brackets, {} t and [ ] t (i = 2,3), and require to be differentiated and the 'outside' terms which are not so enclosed and do not require to be differentiated. Leaving aside the terms in the Wronskians and the terms underlined in the second and the third lines of equation (42) and picking out of the inside terms those which occur with the factor e"2, we have [-8(i/r+/i3)*v28(v-/i2)
+ 5^*^,2
5(v+/i3-/i2)
-S/*?(/*8,2 8/^2 + ^,2 8 " + v,2 8 ^ ) ] - c-c->
(43)
or, after some rearrangements, = {-8fi2[}/ri2bi/r*+/iatld/i*-vi25(^r
+ /ia)*]
-(fr + v) 25(ft + v)8[i*-(ir-v)i2&ftbv*}-c.c.
(44)
A similar reduction applies to the terms with the factor e"3. Thus the inside terms in equation (42) reduce to [e«K[8/*3,8/*3*]2 + M, 8^*] 2 - [Bvt 2 5(^r+/* 3 )* - c.c] - 8/t 2 [^, 2 5 ^"* +/*3>2 8/t? - v 2 5(i/r + /i3)*] + c.c.}l
2
- { e ^ [ ( ^ + v) > I 8(^ + v)8/t* + ( ^ - v ) i 2 6 ^ 5 i ; * ] - c . c . } > 1 + terms ( 2 : ^ 3 ) .
(45)
We keep the terms in [ ] {(i = 2,3) as they are, but expand by differentiation the terms { } t (i = 2, 3) by making use of the equations,
W - " ) , J , . + W - ^ ) ,3], 3 = 0/ [e«*(i/r + v)2]2
= e«*A2-e"*A3,
3
•
(46)
a
K ( ^ + ^),3],3 = e ^ 3 - e a 2 , valid for the background space-time. We find:
= e«*(ifr + v)2[S(ir + v)bfi*l2
+ e«W + v),3[5(Tfr +
+ ( e M a - e%43) 8(/ia-/i2)*8(ir
V)S/i*l3
+ v)
(47)
and {e**(f -v)
26i/r8v*} 2 +
{ea*(iJr-v}35ifrdv*}
3
= ea*(\jr-v)i2(Sijr8v*)
2
+ ea3(^-^),3(5^5^*)3.
(48)
Including these terms along with the outside terms already present in equation
Flux integral for axisymmetric perturbations
335
(42), we find, after some obvious cancellations, that the terms, with the factor e"2 that we are left with, are: - ( f + v ) i 2 [ 5 ( ^ + ^ 2 5 f t * + 8 ( ^ + v)5 ft * 2 ] + c.c. — {ijr — v) 2(8^-8^*2 + 8 ^ 2 8y*) + c.c. + 8(v^+/i3)*28v_2 + 8{f+/i3)%v
28(i/r
+
v)-c.c.
+ S(/*3—/* 2 )*[^, 25(/*3 + 1 '),2+ /*3,2°> + ^ ) , 2 ] - C - C - 8 / * * 2 L « 3 , 2 8 ( / * 3 - / « 2 ) - f , 2 5 ( , ; + / * 3 ) - l ' , 2 8(^+ / M3)] + C.C.
-^5r 28^*28(vir + »' + / tt3-/* 2 ) + C.C.
(49)
On simplifying these terms, appealing several times to the initial-value equation (6), with the right-hand side set equal to zero, we find that they vanish identically! Therefore, after all these reductions, we are left only with the terms included in [ ] A (i = 2,3) in equation (45); and the term with the factor — b/i2 in the second line of this equation, again by virtue of the initial-value equation, becomes, + 8/t 2 8(f+ /* 3 )* 2 -c.c. Thus, with the definitions, E2 = e«K[5/*3,5/4L + [ 5 ^ , 8^*] 2 - [8vt 2 8(^+/* 3 )* - c.c] + [ 8 ^ 8 ( ^ + ^ , ) * - c . c . } ) (50) a
Es = e 3{[8/*2,8/.*]3 + [ 8 ^ , 8 ^ * ] 3 - f 8 ^ 3 8 ( ^ + / M 2 ) * - c . c . ] + [8/* 3 S(^+/* 2 )*-c.c.]} (51) we obtain the flux integral, ^ 2 , 2 + ^ 3 , 3 = 0.
(52)
We call this the 'flux integral' because, by Gauss's theorem f (E2dx3-E3dxi) = I ( ^ d ^ - ^ d x 2 ) , Jc1 Jc2
(53)
where C1 and C2 are any two closed contours, one inside the other, in the (x2,x3)plane, provided no singularity of E (if any) occurs inside the area included between Cj and C 2 ; and, therefore, the outward normal fluxes of E across Cx and C2 are equal.
7. T H E F L U X I N T E G R A L F O R T H E E I N S T E I N - M A X W E L L
SPACE-TIME
We find that the flux integral must now include the Wronskians, [Y, Y*]t (i = 2,3), besides those already included in equation (42). The required equation follows from equation (14) by applying to it our standard procedure, namely, by
336
S. Chandrasekhar and V. Ferrari
multiplying it by 7* and subtracting from it the complex conjugate of the equation multiplied by 7. We find: {ev+to-f-M,[Y, F*] 2 } i2 + { e " + ' ' ^ - ^ [ y , 7*]3}_3 + Y*[e'+»Fm8(v + _Y*[e^Fo28(v
pa-#-/it)lt-c.c.
+ fc2-^-^)l3-c.c.
= 0,
(54)
or, alternatively, 2le«ie-2*[Y,Y*]2-e-^F03[YS(v + 2leHe-W[Y,Y%
+
+ e-^F02[Y&(v
/,3-1/r-fi2)*-c.c.]}\2
+
p2-ir-/i3)*-C.c.]}\3
= [rtZ(v+pt-i/r-fla)*-'PaS(v+pa-i/r-fli)*]-c.0.
(55)
We add this equation to equation (42). The resulting equation will have on its right-hand side the terms, - 25^*[ + (
?>/l*[-(
+ fy*[ + (02-03)-V3
+ 2(e>'+»*FO2),3Y]-c.c.
+ [Vt&(v + /i2-f-fia)*-¥ab{v
+ ji3-
(56)
The reduction of the inside terms of the combined equations (42) and (55) will proceed as before and in place of (45), we will now have: [ e ^ f e , 8/4] 2 + [5f, 5^*] 2 + 2 e-2*[Y, Y*]2 - [8v 2 h{\js+/i3)* - c.c] - 8ju,2[f 2 5^*+/i 3 2 5/i* - v 2 8(^+^3)*] + c.c. - 2 e - ^ F 0 3 [ 7 5 ( l > + /*3-^-/*2)*-c.c.]}],2 -{&*[($+
v)^ty
+ v)hi4-(-ijr-v)^\lrbv*~\-c.c.}
^
+ terms ( 2 ^ 3 , 7 ^ - 7 ) .
(57)
In expanding the terms {} 4, (i = 2,3), by differentiation, we must now use the equations [ e a ^ - v ) , 2 ] , 2 + [e^ (v ir-v),3] > 3 = - 2 ( ^ 2 + ^ 3 ) V - ^ , l l^(^
+ v)t2l2 = e^A2-e^A3
+ (Fl2-Fl3)
V-9,
(58)
in place of equations (46) valid for the vacuum. The use of equations (58) will lead to additional terms in the reduction of the outside terms. Additional terms will also arise from the use of the inhomogeneous initial-value equations (21) and (22) instead of the homogeneous equations (appropriate for the vacuum). A careful scrutiny of the reduction of the outside terms, in the case of the vacuum, shows that in the present context, the reduction, besides all the terms which vanish (as in the case of the vacuum), contributes to the right-hand side of the combined
Flux
integral for axisymmetric
perturbations
337
equations (42) and (55) the following terms, in addition to the terms (56) we already have:
+ m
+ v) 5 / t
*_5(^
+ v)
-28fi5v*(Fl2+Fl)V
S^*]^-^) V -g-o.c. -g + cc
- 2[8/t* 2 e ^ 3 ^ 0 3 Y- 8 < 3 e" + "^ 02 7] + c.c. + 2[8(f + ^ ) i 2 e " + ^ i ? ; 3 y * - 5 ( ^ + i'),3e , ' + ^ J F 0 2 r*]-c.c.
(59)
By a fairly straightforward reduction, we find that the terms (56) and (59) combine, surprisingly, to give: -[e'+^Fm5(i/r + [e^F02
+v+
/ia)*Y-c.o.lt
S ( f + ^ + / * 2 ) * 7 - c.c.], 3 .
(60)
Therefore, the result of the reduction of the combined equations (42) and (55) is: |B-{[5^„
5/**]2 + [Stfr, 5 f *] 2 + 2 e-**[Y, Y*]2
-[8",25(f+/t3)*-c.c.]-{8/(2[^25^+/«3|28/«*-^25^+/(3)*]-c.c.} + 2 e-^*F03[8(iJr + v + /i3)*Y- C.C] -2e-++>*Fm[8(v+ti3-#-p2)*Y-c.c.]}lt
+ [ 2 ^ 3 , r ^ - r i 3 = o.
(6i)
The term in the second line of equation (61), which occurs with the factor 8/t2, can be simplified by appealing once again to the initial-value equation. The term bsconiGS + {5/M25(^+/*3)*2-2e"+^03y5/t*}-c.c.
(62)
Inserting the expression in (61) and simplifying, we obtain the flux integral, #2,2+^3,3=0,
(63)
where now, E2 = e ^ f c , 5/4L + [Sjfr, 8^*] 2 + 2 e~^[Y, Y*]2 ~[8v 2d(i/r+/i3)*-8p%5(i/r+fi3)]
+
[8/j128{i/r+/i3)*2-d/i*5(i/f+/i3)2]
+ 4F03 e-*+*[Yfy*7*8^]}, E3 = eM[8^ a , b/i*]a + [Sr/r, 8^*] 3 + 2 e~^[Y, Y*]3 -[&vtZ&(i/r + /i2)*-8v%8(i/r + /i2)] + [8/is8(i/r + /i2)*3-8/i*8(ir -4F02e^+'liYdi/r*-Y*8i/f]}.
(64) + /i2)<3] (65)
The remarkable simplicity of these final expressions for E2 and E3 is noteworthy; in particular, the parallel patterns in which the terms representing the vacuum and the Maxwell field occur. Application
to the Reissner-Nordstrom
space-time
The application of the flux integral (65) to the polar perturbations of the Reissner-Nordstrom space-time has greater interest than as a mere exercise. The
338
S. C h a n d r a s e k h a r a n d V . F e r r a r i
interest derives from the fact that, in this instance, we are presented with two independent one-dimensional potential-scattering problems. There are thus two functions, Zx and Z2 which satisfy one-dimensional Schrodinger equations with real potentials and, correspondingly, two Wronskians, [ZvZ*]r and [Z2,Z*]r which are constants (for details seeM.T., §42 (d), pp. 230-235). The question that arises is how two independent flux integrals can result from the single integrand (53). To answer this question, we must first clarify the circumstance which gives rise to this situation. The two functions, Z1 and Z2, are related to the amplitudes, H1 and H2, of the two kinds of radiation - electromagnetic and gravitational - that prevail. Thus (cf. M.T., p. 259, equation (347)) Z1 = # 1 c o s ! F + # 2 s i n ! F j and
(66)
Z2 = H2 cos Y- H1 sin V, J
where where
tan V ~ tan V -
m
+
2 Q . V ( l - l ) ( l + 2) V[Mfa + 4Q,{j_1)(j
+ 2)]-
(b7)
Since Z1 and Z2 satisfy independent equations, we can set either,
or
Z1 = 0,
H1 = -H2 tan ¥,
Z2 = H2sec W,
and
[Z2, Z*\m = const.,
(68)
Z2 = 0,
H2 = +H1 tan V,
Zx = H1 sec 5F, and
[Zt, Z*] r> = const.
(69)
A consequence of the two cases, (68) and (69), that we have distinguished (surprisingly not explicitly noted before) is that the radiation field that prevails (when, for example, an arbitrary superposition of electromagnetic and gravitational waves is incident on the black hole) is a mixture of the two kinds of radiation - photons and gravitons - with correlated phases and amplitudes, as specified in equations (68) and (69). This character of the prevailing radiation field was noted in the context of the scattering of radiation by two extreme Reissner-Nordstrom black-holes (Chandrasekhar, 1989, §8); but it is present quite generally. We now turn to the application of the flux integral (63). Since the Reissner-Nordstrom space-time is spherically symmetric, we can identify x2 with the radial coordinate r and x3 with [i = cos 8 (where 6 denotes the polar angle) and apply equation (53) when Cl and C2 are circles of radii r^ and r2 both exceeding the radius r, of the event horizon. We infer that
I
Erd/j, = const.
(r>r+).
(70)
1
The metric coefficients of the static Reissner-Nordstrom space-time in the chosen coordinate system (r, fi = cos 6) are e2" = e~2^ = A/r\
e 2 ^ = r2/(l-/i2), A = r2-2Mr
+ Q%,
e2* = r2{l-fi2),
(71) (72)
Flux
integral for axisymmetric
perturbations
339
where Q* denotes the charge of the black hole. And the only non-vanishing component of the Maxwell tensor is Foi = -QJr\ (73) I t will be noticed that v, /i2, and ft + /i3 are independent of ft and that xjr r = ju,3 r. The associated separated solutions for the perturbations, 8i>, 5^", 8/i2, 8/i3, and Y(= e ^ S i ^ / i o " ) are given by (M.T., p. 231, equations (158)-(160)): 5v = N(r)P{^); *Hi=HX)Pl(ii);
Sfi =
TWPM-VWpP^,
8fi3 =
5 ^ + 5/^3 = [2T(r)-KV(r)]P,(p),
[TW-KVWPM+VWM
K = 1(1+1),
and Y = -(re*/2Q,)Bm(r)PliltV(l-p')We also have the relations, T=V-L+Bm, and
X = \(l-l)(l
8(^+/i3)
(74)
(75) + 2)V = nV,\
= -2(L+X-B23)Pl.
(76)
j
With the solutions given in equations (74)—(78), we find: [o>3, o>*] 2 + [ 8 ^ , 8ifr*]2 = {2[T, T*]t + K*[V, V*]2-K[T,
+ 2[V,
V*]2&P'1(/I)T-2K[V,
2e-*+[Y,
[-Svt2b(ijr+/i!i)*
+
V*]2 + K[T*,
V]2}P\
V^fiP^,
(77)
Y*]2 = WZQlKB^BtMl-fUPi)*,
(78)
&/i28(iJr+fia)*2]-c.c. = 2[NJL+X-B23)*-L(L+X-B23)*2]P*-c.c.
(79)
Remembering that F03 = 0, we find in accordance with equations (77)-(79): •+i
Erd/n i
= 27^![-[X,X*} 2 2l+l[n*L '
J2
+ 2[L+X-B 23,L*+X*-B* 3]2 L 23 '
23J2
+— 2£:
[B23,B*3]2 *
+ 2[JVir(L+Z-fi„)*-iV*(L+Z-JBM)] -2[L(L+X-B23)*2-L*(L+X-B23)^.
(80)
The further reduction of this expression requires a discriminating use of the solutions for L, N, X, and B23 (due to B. Xanthopoulos) given in M.T., p. 235, equations (190)-(196) as well as equations (173)-(179) given on p . 233. The result of the reduction (after ' miraculous' cancellations) is A. Air
4-K"
(27TT)7^^ =2 7 T T ^ ^
<' = 1 ' 2 ) '
(81)
for the two cases (68) and (69). The flux integral thus gives the same two conserved quantities.
340
S. Chandrasekhar and V. Ferrari P A R T II. T H E F L U X I N T E G R A L FOR POLAR P E R T U R B A T I O N S OF A STATIC S P A C E - T I M E W I T H A PREVAILING F L U I D SOURCE 8.
INTRODUCTION
The polar perturbations of a static spherical star leading to damped non-radial oscillations with accompanying emission of gravitational radiation is one of the central problems of relativistic astrophysics. Its study was initiated by Thorne & Campolattaro (1967) and continued by Thorne and others, notably, by Detweiler & Ipser (1973) and Lindblom & Detweiler (1983, 1985; the first of these papers gives a fairly complete bibliography). But it does not seem to have been noticed that there is a flux integral governing these oscillations which, besides providing a useful additional constraint, recasts the problem of the non-radial oscillations of a star as a problem in scattering theory. In this Part II, we shall be concerned only with the derivation of the flux integral. We postpone to a later paper the reformulation of the general problem of the non-radial oscillations of a star in the light of the flux integral. We do not also consider the problem of the axial perturbations: it is known that the scattering of gravitational waves, belonging to the axial modes (described by the Regge-Wheeler equation outside the star) does not affect and, in turn, is not affected by the axial modes of perturbation of the fluid (characterized by motions purely in the (p-direction).
9. T H E E Q U A T I O N S G O V E R N I N G
EQUILIBRIUM
Our restriction to axisymmetric space-times with a perfect-fluid source, effectively requires us to consider only a spherically symmetric distribution of matter described in terms of an isotropic pressure (p) and an energy-density (e), i.e. a star. The metric of the space-time, both in the interior of the star and in the vacuum outside the star, is well-known and can be found in any text book. We shall simply quote the principal equations in a notation that we shall use. The metric is of the standard form (1), where in the present context, e2^ = r2{l-/i2),
e 2 ^ = r 2 / ( l -fi2),
(82)
r = x2 is a radial coordinate and fi = cosd. Inside the star, with its centre at r = 0, the metric functions v and fi2 are given by Vr
=
-pr/(e
where
+
p),
2
e-
M(r) =
^=l-2M(r)/r,
(83)
er2 dr.
(84)
Jo The equation of hydrostatic equilibrium is [l-2M{r)/r]p
r
= ~{e + p)[M(r)/r2 + rp].
(85)
Flux
integral for axisymmetric
perturbations
341
Three additional relations we shall find useful are 2e = r-*[ + l - e - ^ ( l - 2 r / t 2 i f . ) ] , 2
2
2p = r' [-l+e- ^{l
(86)
+ 2rvrr)],
(87)
e+p = r'1e-2^(v+fi2)r.
and
(88)
Outside the star (where e = p = 0) it follows from equations (83) and (88) that v = -/i2
and
where
e-2^=l-2M/r,
(89)
er2dr
M =
(90)
Jo
is the mass of the star and R its radius. The metric is of course that of Schwarzschild.
10.
T H E E Q U A T I O N S OF H Y D R O D Y N A M I C S AND T H E I R
LINEARIZATION
The equations of hydrodynamics, in the tetrad frame that is adopted in this paper, follow from the intrinsic divergence of the energy-momentum tensor,
where we have enclosed the indices in parentheses (temporarily!) to denote that they are tetrad indices. These equations are: (e+p)umu(a)m and
= pAa)-u(a)uib)pAb)
(92)
e a uw + (e+p) uia\(0) = 0,
(93)
where the vertical rule in u(a)^b) denotes intrinsic differentiation as defined inM.T., p. 37. Making use of the Ricci rotation coefficients listed in M.T. (p. 82, equation (91)), we find that the explicit form of equations (92) and (93) (when the only nonvanishing components of the four-velocity u(a), are um, it(2), and w(3)) are: (e + p)-^ - e~"[u2(2) + wf 3) ]p t 0 + e"^ M ( 0 ) U(2) p = +e
2+
e'^um
w(3) pt 3}
LM(0) M(0),0 ' /*2,0 M (2) +/*3,0 M (3)J
- e -/,2 % (2) [>, 2 M(0) + M(o), 2 ] - e ^ W f s j K s «(o> + M
2
[ui2) w (2)i 2 + v 2 M ( 0 ) — /*3 2 M ( 3 ) ]
+ e um\ji20w(2)
+w ( 2 ) ? 0 J — e
3
w ( 3 ) [/i 2 i 3 w(2) + M ( 2 ) 3 J;
(e+^3)~ 1 { + e-ft[wf0) - wf2)] p 3 - e~"uw um p 0 + e~^u(3) « ( 2 ) p 2} = —e
3
Lw(3) w (3)i 3 + y 3 w ( 0 ) — /t 2 3 ii(2)J
+ e""M(0)L«3, 0 M(3) +
M
(3), 0] -
e
"''2W(2)[/*3, 2 % ) +
M
(3), 2].
(94)
1017
342
S. Chandrasekhar and V. Ferrari e"1e,0M(0) + (e + p)[M(0),0 + ( f + / * 2 + ^ 3 ) , 0 M ( 0 ) ] } -e-^{e2u(2)
+ (e + p)[u{2)i2+(ir
-e-i'z{ei3u(3)
+ (e+p)[u{3h3
+
v+/i3)2u{2)]}
+ (i/r + v+/i2)y3ui3)]}
= 0.
(95)
(a) The linearized equations In the static spherically symmetric background, the only non-vanishing component of the four-velocity is u(0)(= 1); and besides, P , 3 = ^ , 3 = / * 2 , 3 = 0.
(96)
In the perturbed space-time « (0) continues to be 1 (i.e. it differs from unity only in the second order) and w(2) and quantities of the first order. We shall write, M
<2) =
§M
2 = £ 2 ,o
a n d
w
(3) =
8M
3
= £3,0.
(97)
where £2 and § 3 are the associated Lagrangian displacements in the radial and in the ^-directions. From equations (94) we deduce that e ^ [ ( e + p)- 1 P,, + >'1,] = K , o = £2,0,0 1
and
e ^ 5 [ ( e + p ) - p , 3 + v 3 ] = 8W3,0 = ^ 0 , 0 ,
(98) (99)
or, explicitly, 8pt2 + 8(e + p)v2 + (e + p)dv
2
= -
2
and
+
5p3 + (e + p)8i':3 = -a (e + p)e-'' ^£,s,
(100) (101)
where we have assumed, as hitherto, that all the perturbed quantities have the common time-dependent factor, eiat. Turning next to equation (95), we find: 5e+(e+p)b(ijr
+ /i2+ii3) = e"-^{e2£2 + (e + p)[£2,2 + (Tjr + v + p3),2£2]} + e-".(e+i>) (£,,, + j f r . k ) .
(102)
(b) The equation that follows from the conservation of baryon number To complete the system of hydrodynamic equations, we need an additional equation. I t is provided by the conservation of the baryon number N which is assumed to be some known function of e and p. In other words,
and
N = N{e,p)
(103)
[Nu +#«<»>, („, = 0.
(104)
From equations (103) and (104) it readily follows that (cf. equation (102)) Sp + yp5{i/r+/i2 + /is) = e'-^ip^^
+ ypl^
+^ +
v+fis)^^]} + e"-"»y2>(g8i, + ^ i , £ , ) >
where
y =
p M/dp
N-(e +
p)^
(105) (106)
Flux integral for axisymmetric perturbations
343
The physical content of equation (105) is no more than that the changes in the pressure and in the energy density of a fluid element, as it moves, take place adiabatically. With the further definition, Q=(e+p)/yp,
(107)
equation (105) takes the alternative form, Q8p + (e+p)8W+/it+fia)=e*-i*{Qpia£a
+ (e+P)[£tta
+
(}fr+v+fta)
+ er*(e+P)(ga<3
+ ifr
(108)
An eliminant of equations (105) and (108) is 8e = Q 5p + er*{et
2
- Qpt 2) £,.
(109)
From this equation, we obtain the physically meaningful relation. (beg* - 5e*£2) = Q(hpE* - hp%2). 11. T H E L I N E A R I Z E D E I N S T E I N ' S
(110)
EQUATIONS
For a perfect-fluid source, the components of the Ricci tensor (by Einstein's equations) are: -\Rm
= (e + p)u20-±(e-p);
-\R11
=
-\B22
= (e+p)u\ + \(e-p);
-\B33
= {e+p)u\ +
-±B02 = (e+p)u0u2;
l(e-p), \{e-p), (Ill)
-\B0Z
= [e + p)u0u3,
-\B23
=
(e+p)u2u3.
For a static spherically symmetric space-time subject to polar perturbations, the components of the Ricci tensor, inclusive of terms of the first order, are B00=-(e
+ 3p);
G00 = - 2 e ;
Bn=B„=Bn=-(e-p),\
B02 = -2(e + p)u2=-2(e+p)^0,
Bm=-2(e
(112)
+ p)ua = -2(e+p)£ai0.
J
We can now write down the linearized Einstein's equations as we have in §4, equations (19i-v), for the Einstein-Maxwell space-time. The left-hand sides of these equations remain unchanged while, on the right-hand sides, the terms in (&2±
+ v+/i2+fis)]V-9,
-[8(e + 3p) + (e + Sp)8(f + v+/ia+tia)W-g,
(H3i) (113")
+ [8(e-p) + (e-p)8(ir
+ v+/i2+/i3)]V-9,
(113iii)
+ [8{e-p) + (e-p)5{f
+ v+/ia+/ia)]V-9,
(113 iv)
+ 2[8e + e5(ft + v+/i2+/i3)]V-g-
(H3v)
344
S. Chandrasekhar and V. Ferrari
I t is convenient to have the initial-value equations (6) and (7) written out explicitly. We have: e a *[5(f+ /it3)i2 + f > 2 5 ( f - , K 2 ) + /i3t2 H/i3-fi2)-v<25(i/r
+ fi3)] =
eai&(i/r+/i2)t3 +
ftz3S(i/r-/i3)+/i2t35(/i2-/i3)-vta8(i/r
ge^+^+^fe + p ) ^ ,
(114)
+ ji2)] = 2e»*+2'+ft(e + j ) ) ^ .
(115)
Again, we need to rewrite equations (113iii) and (113iv) as in §4. By following the same procedure, we find that in place of equations (28) and (29) we now have: [e^(5/i 2 3+/i 2 3 5a 3 )] 3 - ( e ^ 5 / 3 ) 3 + [e^(8^ 2 + r 2 8 a 2 ) ] i 2 _ 2[ e*+*-*.(e + P) 4 ]
3
= SX2 - SX3 -
-[S(e + p) + (e + p)b(ir + i> + /i2+/i3)]V-g and
(116)
[ea«(8/*3t 2 +/t 3 , 2 8a2)]_ 2 - (e^S/,,), 2 + [e^(8v_ 3 + y 3 8a3)]_ 3 -2[e*+tv+^(e
+ p)itlt
5X3-8X2-a2e«°S/i3
=
-[5(e + p) + (e + p)d(i/r + p + fi2+/i3)]V-9.
(117)
And finally in place of equation (41), we have, 8/t*[e^(Sf
2
+ i/ 2 8a 2 )] 2 + 8/M*[ea'(8^3 + ^3Sa3)]
3
= - 8 ( ^ + / i 3 ) * t e ^ ( 5 ^ 2 + i' i 2 8a 2 )] i 2 -8(^+/* 2 )*[e a 3(8y 3 + v 3 8 a 3 ) ] 3 -[8(e + 3p)8(ijr + pa+jia)*
+ (e + 3P)8W+/ii+/ia)*8v]V-g.
12. T H E F L U X
(H8)
INTEGRAL
We consider the same combination of Wronskians as in equation (42) but evaluated with the aid of the linearized Einstein's equations of §11 (instead of those of §4). It is clear that the left-hand side of the equation will be unaffected; and we find that on the right-hand side (the last three lines of equation (42)) are replaced by 25fi*[e^+2,,+^(e + P)£,2l2 + 2dfi*[e'lr+2''+^(e + + 2{[d(i/f+/t2+/i,3)*(5p+pdv)
p)£,3]i3-c.c.
+ 5i/r*[8e + e5(p+{i2+fi3)]}V-g-c.c.
(119)
The reduction of the 'inside' terms (on the left-hand side of equation (42)) will result in the same equation (45). But the expansion (by differentiation) of the terms {} t should now be effected by making use of the equations, [e^(ir-v)t2]2
+ [e^(i/r-v):3l3
= 2(e +
p)V-g,\
[ e a # + ^), 2 ], 2 = e M 2 - e " » i 3 - 2 p V - 9 ,
W+"),i],3 = ^ . - ^ 2 - 2 p V - y ,
(120)
J
Flux integral for axisymmetric perturbations
345
instead of equations (46). The use of equations (120) will lead to additional terms in the further reduction of the outside terms. Additional terms will also arise from the use of the initial-value equations, (113) and (114), instead of the homogeneous equations appropriate to the vacuum. A careful scrutiny of the reduction of the outside terms, in the case of the vacuum, shows that in the present context, the reduction, besides all the terms which vanish, contributes to the right-hand side of equation (42) only the following terms, in addition to the terms (119) we already have: 2(e+p)(8ft?>v*-5ft*8v)V-g -2p[W
+ v)Sfa+fia)*-W +
+ 2[e* ^(e+p)£2fyl2
+ +
+ ef *'' ^(e +
+2v+
+ 2[e+
v)*8(pt+pi)]V-g
+
p)£a8fit3]-c.c.
+2
^(e + p ) ^ + v)%£2 + e* ^{e+p)h{\lr
+ v)*3£,3\-c.c.
(121)
Combining the terms (119) and (121), we have 2[e*+^3(e+p)£2fy*i2
2[e^^(e+P)£38/i*l3-Cx.
+
+ 2{5(i/r+/i2 + /i3)*(5p+pdv)
+ 8}fr*[Se +
e8(v+fi2+fia)]
— (e+p)5ijr*?>v—p8(ijr + v)&(/j,2+[is)*]}^—g
— c.c.
+ 2[e*+2v+^(e + p)£25(i/r + v)% + e*+2"+^(e + p)£38{ft + v)*3]-c.c.,
(122)
or, after some rearrangements, 2[e^2^(e
+ p)^25^l2
2[e^2^(e+p)^8/j,*l3-o.c.
+
+ 2[e*' + ! ' + ft(e+p)5 2 5(f + i')*2 + e ^ M ' ' ( e + l ' ) 4 5 ( f + i')* 3 ]-c.c. + 2[8i/r*[5e+(e+p)5Ui2+/i3)]
+ 8p5(f+/i2
+ /ia)*}V-g-c.c.
(123)
By equation (102), the underlined term in the foregoing expression is 2le*+2>+"*{et2£2 + (e + p)[£2t2 + (ir + v + /i3)2£2]} +
e^^(
+ p)(£ 3 i 3 + ^ 3 g 3 ) | 5 f * - c . c .
e
(124)
Inserting this expression in (123), we find after some lengthy but straightforward reductions, that we are left with 2[e*+2*+K(e + p)£26(i/r + p3)*]t2 + + 2 e ^+2"+/<3( e + p)
g2 5,,* + 2 ef+f+^e +
2[e*+2>+^e+p)£38(ifr+fi2)*],3-c.c. p)£,38v*3-c.c.
+ 25p8(ir + fi2 + fi3)*V-9-cc.
(125)
The terms in the second line of equation (125) can be simplified by making use of the equations, (e+p)^*dv2-c.c. and
=-SS[dptS
(e + p)£*5v3-c.c.=—!;3¥5p3
+ 8{6+p)v
(126)
346
S. Chandrasekhar and V. Ferrari
which follow directly from equations (100) and (101); and we find, successively, _ 2e ^ +2 " + /<3£ 2 [5p* + §(e+p)*v
2]-2
e^ +2 " + ^| 3 8p% + c.c.
= - 2(e^ + 2 " + ^ 2 §p*) 2 - 2(e^ +2 " + ^g 3 5p*) >3 + c.c. + 2[(e^ +2 " + ^£ 2 ) i 2 + (e«Sr+2"+^3)] 3] 5p* - c.c. _2 e ^+2>.+ ft £ 2 5( e + :p)y2 =
+ c_C]
_2(ef+2"+^ 2 8 J9 *) ]2 _2(e* ir+2, ' + ^ 3 5p*) T 3 + c.c. + 2 { e ^ 2 ^ [ f 2 2 + (^ + v + /< 3) i2 g 2 ] + e ^ 2 ^ ( ^ , 3 + ^ 3 g 3 ) } 8 y * - c . c . -2et+2"+»°E>2Se*v2
+ c.c,
(127)
where, in passing from the first to the second equality, we have made use of the equilibrium equation, p2 = — v2/(e + p). The terms in the last but one line of equation (127) can be simplified by making use of equation (108) ' contracted' with dp*. In this manner, the terms in the second line of (125) reduce to -2(e' i r + 2 , ' + ^| 2 5^*) 2 -2(e^ + 2 , , + ^|3 8j9*)?3 + c.c. _ 2 e^+2"+ftg2[5e*i'i 2 + Qp, 2(e -p^bp*]
+ c.c.
+ 28p*5(i/r + /i2+ju,3)V-g-c.c.,
(128)
where the terms in the second line vanish by virtue of the relation Pti = —v 2{e + p) and equation (110). Also the terms in the last lines of (125) and (128) cancel each other. Therefore the reduction of the entire equation (42), with the terms (119) on the right-hand side, is to provide the same terms (cf. equation (45)), [eM[5/*3,8/**]2 + M, 8 ^ * ] 2 - [ S v - 8 > 2 { ^ y28i/r*+/i325/i*-v:
28(f+/i3)*-c.c]
28(i/r+/i3)*}
+ c.c.\2
+ P^3],3,
(129)
on the left-hand side of the equation and the terms, 2te^ + 2 " + "3( e + : P )| 2 8(^+ / «3)*], 2 + 2 [ e ^ 2 ^ ( e + y ) g 3 5 ( ^ + / a 2 ) * ] , 3 - c . c . - 2 ( e ^ + 2 " + ^ 2 8 p * ) | 2 - 2 ( e ^ + 2 ' ' + ^ 3 8 ? ) * ) i 3 + c.c.,
(130)
on the right-hand side of the equation. And finally the term which occurs with the factor — 8fi2 in (129) can be replaced by + 8^2[8(^+/t3)*-2e^(e+^)g*]-c.c,
(131)
by making use of the initial-value equation (114). With this replacement, we obtain the flux integral, once again, in the form ^2,2+^3,3 = 0,
(132)
Flux
integral for axisymmetric
perturbations
347
where E2 = e«*[[o/*3,5/**]2 + [ o f , 5 f * ] , - [ 8 i ; i S 6 ( ^ + ^ 3 ) * - 6 * v
2 S(^+/* 3 )]
-2e" + ^{[(e + ^ ) 5 ( ^ + / . 3 - / < 2 ) * - 8 p * ] ^ - c . c . } ]
(133)
and
+ [8/*3 S(^r +/i2)*3 -8fi* h(f + /*2) 3 ] -2e"+H[(e+p)8(t/r+/i2-/i3)*-8p*]^-c.c.}l
13. C O N C L U D I N G
(134)
REMARKS
The derivation of the flux integral, for the axisymmetric perturbations of a static space-time, directly from the appropriate linearized equations, demonstrates vividly that the ' magic' of the exact theory extends to its linearized versions. I t is specially noteworthy that allowance for different prevailing sources is accomplished parallelly with such ease. I t is equally remarkable how every single equation governing the source plays its own equal role, a feature particularly evident in the reductions of §12 for a perfect-fluid source. As we have stated in the introductory section, the present investigation arose when we wished to extend our earlier study (Chandrasekhar 1989) of the scattering of radiation by two extreme Reissner-Nordstrom black-holes by axial modes to include polar modes. The flux integral of § 7 now provides the necessary means for this extension. Similarly, as we have indicated, the flux-integral derived in § 12 will enable a reformulation of the problem of the non-radial oscillations of a star. We are presently studying these matters. We are grateful to Greg Burnett, John Friedman, Rafael Sorkin, and Robert Wald for helpful and critical discussions of the various questions - both concrete and general - that arise in the context of the problem considered. The research reported in this paper has, in part, been supported by grants from the National Science Foundation under Grant PHY-84-16691 with the University of Chicago. We are also grateful for a grant from the Division of Physical Sciences of the University of Chicago which has enabled our continued collaboration by making possible periodic visits by Valeria Ferrari to the University of Chicago.
A P P E N D I X . T H E FLUX INTEGRAL FOR THE AXIAL P E R T U R B A T I O N S A STATIC E I N S T E I N - M A X W E L L
OF
SPACE-TIME
For axial perturbations, the metric functions, v, ft, ji2, and JX3 and the components F02 and -F03 of the Maxwell-tensor retain their values in the static space-time while the perturbation is represented by the first-order quantities, w, q2, and q3 in the metric (2) and the component F01 of the Maxwell-tensor. In Chandrasekhar (1989) the relevant perturbation equations have been written 12
Vol. 428.
A
348
S. Chandrasekhar and V. Ferrari
down in the particular context of the two-centre problem. Without this specialization, one readily finds that equations (41) and (42) of Chandrasekhar (1989) are replaced by e-H-"+f2+^Xt 00
= [ e -^+'-/'.+h(I |2 -4ei , ' + ''»i*;3 7 ) ] i 2 + [e-^^-^h{X^3
+ 4 e ^ + ^ 2 Y)\a,
(A 1)
_erW+»+f>*F02(Xi 3 + 4 e*+*Fm Y) + e-2*+"+^F03(X ^ 2 - 4 e*+"FM Y), where
X = ez*+"-^*Q23
and
Y = e*+"Fn.
(A 2) (A3)
Applying to equations (A 1) and (A 2) our standard procedure, we obtain {e-»++*-*+i>,[X, X*]2}2 + {e-a*+"~^+^[X, X*] 3 }, 3 + A{X*(e~2*+"^F02 Y)i3-X(e-^+,>+^F02 -X*{e~2^"^F03
F*)3
Y)2 +X(e-2*+,,+^F03
F*)_ J = 0
(A 4)
and {e"+/<3-f-/<2[F; 7 * ] 2 } 2 + {e" + ^-> ! '-ft[7,
-4:e-2^+^F02{Y*X3-YX%)
Y*]3}3
+ 4:e-^+''+^F03(Y*Xi2-YX*2)
= 0.
(A 5)
By adding equation (A 4) to four times equation (A 5), we obtain, {e-^+"-^^[X,X*]2
+ 4 e - ^ ^ - ^ I T , Y*]2-4e-^+^F^YX*
- F*X)}>2
+ {e- 3 ^ + "-^ + ^[X,Z*]3 + 4e-t ! r + , ' + ^-^[F, F*] 3 + 4 e - 2 ^ + ^ . F , 0 2 ( F X * - F * J 0 }
3
= 0. (A 6)
Equation (A 6) represents the flux integral for the problem on hand. Application to the Reissner-Nordstrom space-time In accordance with the solution for the axial perturbations of the ReissnerNordstrom space-time given inM.T. (equations (131), (132), (137), (138), and (143) on pp. 228 and 229) we have: x
Y =
= e*++»-H-HQn
^'F«
=
= rH2C$t{ji),
-VUl-M+2)lH*C*M>
(A 7) (A8)
where CjJ2 and Cj^ are the Gegenbauer functions in their standard normalizations. From equations (A 7) and (A 8) we find: e-^~"^[X,X*]2 Ae-i>+v+H-My
F*l =
= -2[H2,H*]r[^(^2 -
— \H
H*~\ ^itiWi
(A 9) ,\ JQN
Flux integral for axisymmetric perturbations
349
or, after integration over /*, J
{e-3^+"-^+^[X, X*]2 + 4 e-^"^-h[Y,
Y*]2} dju,
= (/ + |)a + 2 ) ' + l ) ^ - l ) { ^ ' ^
+
^'
i /
^
}
= C°nSt-
(A
">
Since F03 = 0 for the Reissner-Nordstrom space-time, it is manifest t h a t equation ( A l l ) is consistent with the required constancy of the Wronskians for the two cases (68) and (69).
REFERENCES
Burnett, G. & Wald, R. 1990 Proc. R. Soc. Lond. A 430. (In the press.) Chandrasekhar, S. 1983 The mathematical theory of black holes. Oxford: Clarendon Press. Chandrasekhar, S. 1988 Proc. R. Soc. Lond. A 415, 329-345. Chandrasekhar, S. 1989 Proc. R. Soc. Lond. A 421, 227-258. Detweiler, S. L. & Ipser, J. R. 1978 Astrophys. J. 185, 685-207. Detweiler, S. L. & Lindblom, L. 1985 Astrophys. J. 292, 12-15. Friedman, J. L. 1978 Communs math. Phys. 62, 247-278. Lee, J. & Wald, R. 1990 (In the press.) Lindblom, L. & Detweiler, S. L. 1983 Astrophys. J. Suppl. 53, 73-92. Thome, K. S. & Campolattaro, A. 1967 Astrophys. J. 149, 591-611.
On the non-radial oscillations of a star B Y S U B R A H M A N Y A N C H A N D R A S E K H A R 1 A N D V A L E R I A FERRARI 2 1
2
ICRA
University of Chicago, Chicago, Illinois 60637, U.S.A. {International Centre for Relativistic Astrophysics), Dipartimento di Fisica 'G. Marconi', Universita di Roma, Rome, Italy
A complete theory of the non-radial oscillations of a static spherically symmetric distribution of matter, described in terms of an energy density and an isotropic pressure, is developed, ab initio, on the premise that the oscillations are excited by incident gravitational waves. The equations, as formulated, enable the decoupling of the equations governing the perturbations in the metric of the space-time from the equations governing the hydrodynamical variables. This decoupling of the equations reduces the problem of determining the complex characteristic frequencies of the quasi-normal modes of the non-radial oscillations to a problem in the scattering of incident gravitational waves by the curvature of the space-time and the matter content of the source acting as a potential. The present paper is restricted (for the sake of simplicity) to the case when the underlying equation of state is barotropic. The algorism developed for the determination of the quasi-normal modes is directly confirmed by comparison with an independent evaluation by the extant alternative algorism. Both polar and axial perturbations are considered. Dipole oscillations (which do not emit gravitational waves),are also treated as a particularly simple special case. Thus, all aspects of the theory of the non-radial oscillations of stars find a unified treatment in the present approach. The reduction achieved in this paper, besides providing a fresh understanding of known physical problems when formulated in the spirit of general relativity, provides also a basis for an understanding, at a deeper level, of Newtonian theory itself.
1. Introduction During the vintage years of relativistic astrophysics, the two central problems were the radial and the non-radial oscillations of spherical stars: the solutions to the linearized versions of the relevant exact equations of the theory provided unique insights into the physical consequences that may derive from the general theory of relativity without the ambiguity and the uncertainty of ad hoc approximative treatments. Thus, the theory of radial oscillations (Chandrasekhar 1964a, b) revealed a global instability of relativistic origin - an instability that is ultimately the cause for the gravitational collapse of massive stars - while the theory of the non-radial oscillations, initiated by Thorne and his collaborators (Thorne & Compolattaro 1967 and subsequent papers: for a fairly complete bibliography see Lindblom & Detweiler 1983) provided an adequate base for a rigorous resolution of the vexed question of Proc. R. Soc. Lond. A (1991) 432, 247-279 Printed in Great Britain
247
248
S. Chandrasekhar and V. Ferrari
the emission of gravitational radiation by non-stationary sources and the damping that ensues. The theory of radial oscillations is so direct and so simple that its implications were immediately appreciated. But the completion of the theory of non-radial oscillations, on the base provided by Thorne, was slow in coming: only in 1983 did Lindblom & Detweiler succeed in bringing the analytical framework of the theory to a satisfactory enough stage to allow a tabulation of the real and the imaginary parts of the characteristic frequencies of the quasi-normal quadrupole (1 = 2, by Lindblom & Detweiler 1983) and higher order (I ^ 2, by Cutler & Lindblom 1987) modes of oscillation for a range of stellar models sufficient for a comparison with the observations on neutron stars. Pursuant to a suggestion made in an earlier paper (Chandrasekhar & Ferrari 1990, §§8 and 13; this paper will be referred to hereafter as Paper I), we develop in this paper, ab initio, a complete unified version of the theory of the non-radial oscillations of a spherical distribution of matter (a star!) that provides not only a different physical base for the origin and the nature of these oscillations, but also simpler algorisms and a simpler set of equations for the numerical evaluation of the complex frequencies of the quasi-normal modes (I ^ 2) and the real frequencies of the dipole oscillations (I = 1). We consider first the conceptual aspects of the present formulation in the following section.
2. The Newtonian and the relativistic views on the origin of non-radial oscillations On the Newtonian view, we imagine that an initially static spherically symmetric distribution of matter, representing a star, is disturbed non-radially and set into oscillation by some unspecified external agent. We analyse the perturbation in spherical harmonics (Ylm) and consider, individually, the modes of oscillation belonging to the different Is and ms. (In view of the spherical symmetry of the background static distribution, there will be no loss of generality in assuming that the perturbations belonging to the various modes are axisymmetric and are described by appropriate Legendre polynomials - a simplification that is vital in the relativistic context (see §§4 and 11).) In the mathematical treatment of the problem, the variable that is singled out is the Lagrangian displacement, £, which an element of mass experiences as we follow it during its motion: for, in its terms, the accompanying changes in the density, the pressure, and the gravitational potential, can all be expressed uniquely. By further supposing that the time dependence of the perturbations is given by ei
On the non-radial oscillations of a star
249
unspecified the agent responsible for setting the star in oscillation, it would be natural to suppose that they are excited by the incidence of gravitational waves of a specified kind - polar or axial, belonging to particular Legendre or Gegenbauer polynomials - as in the treatment of the perturbations of the Schwarzschild spacetime. The incident gravitational waves will be scattered (i.e. reflected and absorbed) by the curvature of the space-time; and the problem can be reduced to one in scattering theory. The solution to the problem of the quasi-normal modes (in which we are, of course, primarily interested) can be deduced from the general theory. By developing the theory of the scattering of gravitational radiation by a spherical distribution of matter in strict analogy with the theory of the scattering by the Schwarzschild black-hole, as set out in The mathematical theory of black holes (Chandrasekhar 1983, §§24 and 26, pp. 142-152 and 160-163; this book will be referred to in the sequel asilf.T.), we find that the perturbation equations allow, very directly, a decoupling of the equations governing the metric perturbations from the equations governing the hydrodynamical perturbations. Once the equations governing the metric functions have been solved, the solutions for the hydrodynamical variables follow algebraically without any further ado. The reduction of the problem of the non-radial oscillations of a star to one purely of the scattering of gravitational waves by the static space-time of the star, without any reference to the motions that may or may not be induced in the star (as in the case of the axial perturbations considered in § 11) is consonant with the physical base of the general theory of relativity. Clearly, this relativistic picture of the origin and the nature of the non-radial oscillations of a star is different from the Newtonian picture. Nowhere is this difference more manifestly brought out than in the total absence of the Lagrangian displacement from the equations determining the quasinormal modes. The plan of the paper is the following. In §3, the equations governing the static space-time of the star are quoted in the forms they are written in Paper I | . In §§4-10, the theory of the polar perturbations (the only class of perturbations that are normally considered) is developed; in §4, the linearized versions of the relevant equations of hydrodynamics and of Einstein are separated and a suitable set from them is selected. In §5 it is shown how the decoupling of the equations governing the metric variables from the equations governing the hydrodynamic variables can be effected; and the equations are specialized for the case when the equation of state governing the star is barotropic. (In the rest of the paper we restrict ourselves to this barotropic case.) An attempt to determine, in §6, the behaviour of the solutions near r = 0, reveals that the equations derived in §5 are not linearly independent at the origin. The origin of this novel circumstance is explored in §6 and it is shown how by defining a different set of independent variables, which satisfy equations that are linearly independent at the origin, we are led to a determinate indicial equation; and the behaviours of the solutions for the chosen variables at r = 0 are derived. In §7 it •(" The conventions and definitions that are used in this paper are the following: Q(a)(b) = +2^(o)(6) and
W
T
250
S. Chandrasekhar and V. Ferrari
is shown how a pair of singularity free solutions, allowed by the indicial equation, satisfy the requirements at r = 0, and exactly suffice to determine the unique solution (apart from a constant of proportionality) that satisfies the necessary conditions at the boundary of the star. In §8, the manner in which the solution, found for the interior, is to be extended into the vacuum outside the star is considered and how the asymptotic behaviour at infinity of the solution so extended enables the determination of the quasi-normal modes. And finally, in §9, we confirm by considering a specific example - a relativistic polytrope of index 1 . 5 - t h a t the algorism provided in this paper for determining the complex characteristic frequencies of the quasi-normal modes, yields values in almost exact agreement with those given by the different algorism of Lindblom & Detweiler. In § 10, we consider the dipole oscillations of a star. Even though these oscillations are not accompanied by gravitational radiation, the problem is nevertheless amenable to the present treatment as a simple special case. An illustrative numerical example is provided. And finally we consider the axial perturbations in §11. I t is known that the incidence of axial gravitational waves cannot excite any motions in the star. On this account, the reduction of the problem to one of scattering by a potential barrier (see figure 2) is plainly to be expected. Indeed, we find that the wave equation one obtains is a simple generalization of the ' Regge-Wheeler' equation that governs the axial perturbations of the Schwarzschild black-hole.
3. The equations governing the static space-time The metric of the static space-time of a spherically symmetric distribution of matter, described in terms of an energy density e and an isotropic pressure p, and the equations determining the structure of a star, have been written down in Paper I. (It may be recalled that both in M.T. and in Paper I the basic equations are expressed in a tetrad frame.) For convenience of reference, we shall write them out again. The metric of the space-time, both in the interior of the star and in the vacuum outside, is of the standard form, ds 2 = e ^ d ^ - e ^ f d r f - e ^ f d r ^ - e ^ f d i p ) 2 .
(1)
where, in the present context, e2*> = r2,
e2* = r2sm2d,
(2)
r = x2 is a radial coordinate, and 6 = x3 is a polar angle. Inside the star, with its centre at r = 0, the metric functions v and determined by v,r = -P.r/(e+P) e~2^= l-2M(r)/r,
and where
er 2 dr.
M{r) =
(3) (4) (5)
Jo
The equation of hydrostatic equilibrium is [i-2M{r)/r]pr Proc.R. Soc. Lond. A (1991)
= -(e+p)[pr+M(r)/r2].
(6)
On the non-radial oscillations of a star 2
Besides
2er = + 1 — e~
2/,2
251
(l — 2r/i2
2pr2 = - l + e " 2 ^ ( l + 2 r v r ) ,
(7)
(e+p)r = e~2/1>(v + fi2)ir.
and
(8)
I t should be noted that in obtaining the solution for v by the integration of equation (3), we must allow for a constant of integration v0: thus, P,r ;dr+v0. (e + p)
(9)
The constant v0 is to be determined by the condition that, at the boundary r = rlt of the star (where p and, generally, also e vanish), (e2Vr1 = ( e - 2 / i 2 U 1 = l - 2 M / r 1 , where
if =
p.
er 2 dr.
(10) (11)
Jo
denotes the inertial mass of the star. By this choice of v0, we ensure that the spacetime outside the star is described by the Schwarzschild metric in its standard form.
4. The equations governing the polar perturbations of the star For the polar perturbations of the star - we postpone to § 11 consideration of the axial perturbations - the metric continues to be diagonal and the quantities which describe the perturbation are the amplitudes, 5^, 5i/r, &/i2, 8/*3, 8e
and
bp,
(12)
lcr(
of the oscillations, with a time dependence e where cr is a constant (not necessarily real), in the metric functions v, \jr, /i2, and ji% and the hydrodynamical variables e and p. Also, we express the tetrad components of the four velocity, u^ and n^, of the motions that are induced in the meridian planes, in terms of the Lagrangian displacements, £2 and £3 in the manner (cf. Paper I, equations (97)): um = 8w2 = g2,o = io"^2.] U(3)
(
= 8w3 = £ 3 i 0 = icr£g.J
(a) The linearized hydrodynamical
i
equations
The linearized hydrodynamical equations for the problem on hand are (cf. Paper I, equations (100) and (101)): -cr 2 (e+p)e"" + ^g 2 = §p^ + §(e + p)v^ + (e+p)$v^t 2
+
-a (e+p)e-" "^3
= bp
(14) (15)
while the equation governing the conservation of baryon number is (cf. Paper I, equations (107) and (109)): 8e = Q&p + e>-*(eir-Qptr)it, (16) where
Q = (e+p)/yp,
and y denotes the adiabatic exponent (defined in Paper I, equation (106)). Proc. R. Soc. Land. A (1991)
(17)
252
S. Chandrasekhar and V. Ferrari
(b) The linearized Einstein equations In Paper I, §11, we have written down the linearized Einstein equations in the forms needed for the purposes of that paper, namely, the derivation of the flux integral that the equations admit (cf. Paper I, equations (132)-(134) with the difference in sign convention noted on p. 249). But for the purposes of this paper, it is more convenient to write down the inhomogeneous equations corresponding to the homogeneous equations (M.T., pp. 145, 146, equations (31)-(34)) that govern the perturbations of the Schwarzschild black-hold. We have : \r
(18)
r
(8i/r + 8/i2) g + ( 5 ^ — 5/^)cot# = — 2(e+p)£3.?+!*$ e' (8i/r + 8v)
rJ)
+ (8i/r — 8/i3) rcotd + i vr — \8v
r1
(19) (20)
1 •2|- + 2vr)^a" ' ) r yr
-2ft>
+ \[(8f + 8v)
\28f
+ 8v- 8/i3) e cot 6 +28fi3] - e " 2 l 5 ^ + 5 / « 3 ) i0 , 0 = + 2 8 p ,
(21)
and e-2^|-25/t2
V
\-+v,r)(v,r-/J'2,r)
.r,r +
+ (5^ + 5i^)
+ e~2>l3(o-v + b~/i2)t3cotd-e-2''{oi/r
+ O{i2)i0i0 = 2b~p. (22)
(Equation (22) is not included among the equations listed inM.T.; it derives from the £r 33 -component of the perturbed field equations.) (c) The separation of the variables Equations (13)—(16) and (18)-(22) can be separated by the substitutions (cf. M.T., p. 147, equations (36)-(39), originally due to J. Friedman): 5^ = N(r)Pl (cos 6), ty, = T(r) Pl + V(r) P^e, Ur,9)
= Ur)Pt(oosd),
op = n(r) Pt (cos 6),
8/i2 = L(r)P, (cos 6), 8^ = T(r) Pt + V(r)
£3(r,d) = and
Pli()0otd,
Ur)Pli(),
(23)
5e = E(r) Px (cos 6).
According to these substitutions, 8 ^ + 5/t, = [2T-l(l+ bfi6 + {bf-0H3) 8i/rJ}t6+(28iJr-8/i3)<ecotd Proc. R. Soc. Lond. A (1991)
1)7] Plt
cot 0 = (T-
(24)
V)Pli$,
+ 28/i3 = -{l-l){l
+ 2)TPl.
On the non-radial oscillations oj a star
253
With the further definitions, 2(e + p)e"+^£,2(r,d) = U(r)Plt "1 2(e+P)e"+^3(r,d)
=
(25)
W(r)Pl:f),j
the hydrodynamical equations (14) and (15) give: U = 2(e + p)ev+^i2
= -—2e^[{E
+ II) v r + 77\ r +(e+p)N r],
W = 2(e + p)e"+"^3 = -^e2v[n+(e+P)Nl o~
(26) (27)
Making use of equations (24), we find from equations (18)-(22): d (1 dr \r
^,
r +
(2T-KV)--L
= -U,
(28)
r T-V+L = -W,
(29)
( r - F + J V ) , r - ( i - v r ) ^ - ( i + v r ) L = 0,
(30)
( V r ) ( 2 T - . F ) , r - ^ + 2,r)L -4(2%T+/cA0e 2 ' t 2 + cr 2 e- 2 , ' + 2 ^(2T-KF) = 2e 2 ^/7,
(31)
F r r + f- + v r - / « 2 r V r + ^ ( i V + L ) + (7 2 e^- 2 , 'F = 0,
(32)
where K = 1(1+1),
and 2n=(l-l)(l
+ 2) = K-2.
(33)
Equation (32) is in many ways a remarkable equation. I t was first encountered in the form, /. \ 2fti Vrir + 2(-+vn\Vr + ~(N+L) + a2e-i"V=0 (34) in the context of the perturbations of the Schwarzschild black-hole (M.T., p. 148, equation (51)). I t was next encountered in the context of the perturbations of the Reissner-Nordstrom black-hole (M.T., pp. 232-233). (In these two later cases, v = — [i2, and the reduction of equation (32) to the simpler form (34) is clearly required.) We now encounter equation (32) once again in the present context of the perturbations of the spherically symmetric distribution of matter. The remarkable feature of equation (32) is that it seems to govern quite generally the perturbations of spherically symmetric static space-times independently of the nature of the source. A derivation of equation (34) is provided in Appendix A. Transcribing equation (16) similarly, we obtain -2/. 8
E = Qn+
e {e
2^+Y) -'~QV
Proc. R. Soc. Lond. A (1991)
(35)
254
S. Chandrasekhar and V. Ferrari
Next, rewriting equations (27) and (28) in the forms n
r
+ (E + n)p
+ (e+p)N
r
2
and
r
= -la22e~2"U,
(36)
L
2
- n = \(T e- "W+{e + p)N;
(37)
and eliminating E and 77 from equation (35) with the aid of equations (34) and (36), we obtain: e
2/1
'v
e+p
——^(er-Qp
U=(la*e-^W),r + (Q+l)VJlor2e-^W)
+ (eir-Qpr)N.
(38)
Returning to equation (31), we can eliminate 77 from the right-hand side of the equation, in favour of the scalars representing the metric functions, by writing equation (36) in the equivalent form (cf. equation (30)), n=l
(39)
we find: -Nr
+ (-+v
T)(2T-KV)
^(2nT+KV)
r~(l+2v
rr)L
+ 2e2^(e+p)N+a2e2^2"[T~{K-l)V-L]
= 0.
(40)
r
5. The decoupling of the metric perturbations from the hydrodynamical perturbations: the barotropic case By replacing equation (31) by equation (40) and eliminating from equation (28) with the aid of equation (38), we can clearly decouple the equations governing the metric perturbations from the equations governing the hydrodynamical perturbations. In developing this new approach to the theory of non-radial oscillations, it is convenient, in the first instance, for the sake of simplicity, to restrict our consideration to the barotropic case, i.e. when the pressure is a unique function of the energy density and the generally applicable relation,
F
\" u /entropy=const.
g _(- p (\p
becomes,
y =
— (p = p{e)). (42) p de The relation (42) is applicable in the most important contexts that one has in view in developing this theory, namely, white dwarfs and neutron stars. For an initially static spherical distribution of matter with a barotropic equation of state, equation (42) is equivalent to y = {e+p)pT/er. (43) For y given by equation (43), Q as denned in equation (17) becomes, Q = e,rlV,r, Proc.R. Soc. Lond. A (1991)
(44)
On the non-radial oscillations of a star
255
and equation (38) reduces to the very simple form (45)
U=W
r
= Wr +
(46)
(Q-l)vrW.
Equations (29), (30), (32), (40), and (46) now provide a basic set of equations for the four functions N, L, T, and V that describe the metric perturbations. In considering these equations, it is convenient to eliminate T in favour of W and choose as our independent variables the functions, N,L,X
= \(l-l)(l
+ 2)V = nV
and
W = -(T-L
+ V).
(47)
In making this choice, we are following the treatment of the perturbations of the Schwarzschild black-hole as set out in M.T. 24(6) (pp. 145-152); only in this latter case W = 0. In other words, while W is simply related to the Lagrangian displacement £3, it should be considered, here and in the sequel, as an abbreviation for the combination of their scalars — (T— V+L), describing the metric perturbations. With the choice of variables (47), we obtain, in place of equations (30), (32), (40), and (46), the following set of equations: L
.r + (--».r)L
+
N.
-'
-Ul+rv
x n{w, + 0-"') r ,r + (l+V,r)L
' )
r)L r
1
{ne2^-i)L
^1*2
[l+rvtr)Wir /2 + [- + v T-/iiT)X ' ' \r ' / rr
W\ = 0,
•W=0,
(48) (49)
N
•(e+p)r
+ 2v rL +
o-2e2^-v)rL
X+o-2 e 2 ( ^"VZ
1 + rv r) X r +
X
2 -(Q+l)v r
L
r n+l
-e2f2
N,
x +
e^W+\o-2e2^")rW\ afyi2
+ n—(N+L) ' '
r
= 0,
(50)
+ o-2e2i»*-,')X = 0.
(51)
Equations (48)-(51) provide a system of four coupled linear equations for the four functions N, L, X, and W which describe the perturbations of the space-time. They are of total degree 5 and not 4 as in the reduction of Lindblom & Detweiler in their (conventional) treatment of the problem. I t does not however appear that the degree of the system of equations (48)-(51) can be reduced further: the reason derives from the special character of equation (51) to which we have already made reference in §4 following equation (32); and, possibly, also for a more fundamental reason: e.g. the linear dependence of the equations (48)-(51) at r = 0 pointed out in §6. Proc. R. Soc. Lond. A (1991)
256
S. Chandrasekhar and V. Ferrari
6. The transformation of equations (48)-(51) to a system linearly independent at the origin: the indicial equation and the behaviour of the solutions at r = 0 A novel feature of equations (48)—(51) - t h e i r linear dependence at the origin becomes manifest when we examine the behaviour of the solutions at r = 0. For establishing the behaviour of the solutions for r->0, it is, of course, necessary to know the behaviours of the various quantities that describe the static space-time of the star. We list below the expansions that we shall need. e = e0-e2r2+...,
2
e-2^ = i-2M(r)/r
3C0'
= Q0 + O(r2);
cr0 = ere""",
4
= l - f e 0 r + fe 2 r , e 2 ^ - ") = e_2,'»[l + ( 6 - a ) r ! ! ] )
622V2),
2
a = p0 + le0, "
Q = er/pr
+ b2rit
l+br2
/i2r = r(6 +
+ ...,
M(r) = |e 0 r3 - £e2 r5 + ...,
Qo = ed-Pv
e2A2 =
p = p0-p2r*
(52)
v
a2 = |e0(|>0 + | e 0 ) - ( ^ 2 + | e 2 ) , 9C0
u
2
5C2'
u
9c0
2
5C2'
We may, parenthetically, note the relations, a + b = e0+p0
and
a — b = p0 — |e 0 .
(53)
We now seek the behaviour of the solutions, N, L, X, and W of equations (48)-(51) via an indicial equation for the exponent x in the substitution, (N,L,X, W) = (N0,L0,X0, W0)r* + O(r*
(54)
By this substitution, we obtain a system of linear homogeneous equations for N0,L0, X0, and W0; and the required vanishing of the determinant of the system leads to the indicial equation:
0 x— 1 x—(n+l) n
x+2 -(x + 1) — x+(n—l) n
x+1 0 -(x+1) x(x+ 1)
|x+l —X
—x+n 0
= 0.
(55)
The determinant vanishes identically and any value of the exponent x is allowed! The reason for this paradoxical result is that equations (48)-(51) are not linearly independent at the origin (though they are at finite distances from the origin). This possible feature of a system of coupled linear equations does not seem to have been considered in the extant standard treatments of the subject. However, in the present context, the impasse can be overcome by a judicious choice of independent variables when we do obtain equations which are linearly independent at the origin. We proceed as follows. First, separating the terms in equation (50) that are of the lowest order in r (as Proc. R. Soc. Lond. A (1991)
257
On the non-radial oscillations of a star
r ^ O ) (and which contribute to the indicial equation (55)) from the rest, we can rewrite it in the form, n+1 r
(e2/i2 — 1) — e2/i2(e +p) rN
Lr-n-^L\-
rv rL
( e 2 ^ - l ) L + 2 v r L + cr 2 e 2( ''^VL
r
-X\- rv X r + - (e2^ -1) X + o-2 e2<*-v)rX
Xr +
wr--w\-
7)
rv rWr
( e ^ - l ) W + 4 o - 2 e 2 ^ - v ) r W = 0,
(56)
or, in the abridged form, N
: r
- ^ N - U
r
- ^ L ) - \ X ,
r
+ ±x\-(w,r--W)-III(say)=0
(57)
Similarly, rewriting equation (49) in the form Nr--N-(L ' r
r
\ '
+ -L\-WT r J
+ vr{N-L)
(58)
= 0,
and subtracting it from equation (57), we obtain -n(N-L-W)-(rX
= r[III + v r(N-L)]
= r 2 ^(say).
(59)
From this last equation, it follows that (rt9)tT
= -n(N-L-W)ir-{rX
(60)
+ ZXir).
On the other hand, according to equations (49) and (51), (N-L-W)
r
=
(61)
-{N+L)-vr(N-L).
and 2
9?
71
X r > r + - X r + - ( i V + L ) + - ( e 2 ^ - l ) ( i V + L ) + ( ^ r - / « 2 j r ) Z r + o- 2 e 2 ^-'' ) Z = 0.(62) Substituting from these last two equations for the terms on the right-hand side of equation (60), we obtain (r2
+ r(vtr-{i2ir)Xr
+ (r2e^-'')rX.
(63)
Returning to ^ as denned in equations (57) and (59), namely, $ = [III+ vr(N-L)]r-\
(64)
and evaluating it accordingly, we find 9 = vr(L+X+W)tr
+ -2{e2»*-l)[(n+l)N-n(L
+ W)+X]
+ ^(N+L)-e2^(e+p)N+
+ lW).
(65)
258
S. Chandrasekhar and V. Ferrari
We shall find it suitable for our purposes to simplify the foregoing expression for <& with the aid of the equation (cf. equation (61)) v r{N-L-W)
+ (v r)2{N-L)
r-^(N+L)
= 0.
(66)
r
We find: 0 = v r[(N+X)
r+
vr(N-L)]
-Je2^-l)[(n+l)N-n(L+W)+X]
+
-e^{e+p)N+a2e2^-"){L+X
+ lW).
(67)
We shall now show that by considering N,L,X,
and 0
(68)
as the independent variables, we can obtain a set of equations for them which are linearly independent at the origin. For the required elimination of W we make use of the equation (cf. equation (59)) W = -<$ + (N-L) + -(rX n n
r+X).
(69)
Substituting for W from this equation in equation (67), we obtain after some additional simplifications, <S = J,r[(N+X)r
+ vtr(N-L)]
\(e2^-l)(N-rXr-r2
+
2
-e2^(e+p)N+la2e2^-v)\N+L
+ -_ ^i + -[rXir+{2n+l)X]\. n n'
(70)
Equations (51), (63), and (70) provide three of the required equations. A fourth equation is obtained by eliminating W r from equations (48) and (49) and substituting for W from equation (69). We find in this fashion, (L+N+2X)r
+ (--vr)(-N+3L
+ 2X)
= 0. (71) --(Q + i)v, {N-L + -g + -(rXr+X)} [ n n ' J r The basic set of equations, when considering X, 'S, N, and L as the independent variables, is ,„ > X r r + f - + i ' r - ^ 2 r]X r + ^e2^(N+L) + o-2e2^-^X = 0, (72)
+
\T
'
' /
T
7}
(r2<$) r = nv r(N-L)+-(e2»*-l) r
_VrNtr
(N+L) + r(v r-/it •
= -g + v
Proc. R. Soc. Lond. A (1991)
+
r)X r
+ o- 2 e 2 ^-"VZ,
(73)
\(e2^-l)(N-rX,r-r2g) + -g + -[rX,r + (2n+l)X]\,
(74)
On the non-radial oscillations of a star Lr
= (N+2X)
T+
[--v
r)(-N+3L
259
+ 2X)
+
N-L
•(<2+l)>\
+ - $ + -(rX
r+X)
(75)
The indicial equation By the manner of their derivation, we may expect that equations (72)-(75) are linearly independent at r = 0. Indeed, as we shall now show, they lead to a determinate indicial equation which provides the required number of linearly independent solutions at the origin. Thus, by assuming a behaviour of the form, (X,<S,N,L) = (X0,%,N0,L0)r*
+ O(r*+2),
(76)
and making use of the expansions listed in equations (52), we find for the exponent x the indicial equation,
(x+l)
x(x+ 1) (a — b)x + a\ o~2 (a-b)x + -±(x + 2n+l) In 1(1+1)
0 -(x + 2)
n n(a + b)
n — n(a-
-1
a(x-l) + \(rl
2°0
0
n
n
= 0,
(77) where in reducing the determinant to the foregoing form, we have made use of the first of the two relations (53). By expanding the determinant, we find na(x+l)(x-l)2(x
+ l+l)2
= 0.
(78)
We conclude that equations (72-(75) allow two linearly independent singularity-free solutions with the behaviour rl at the origin. Indeed, for x = I, the homogeneous system of equations for X0, ^0, N0, and L 0 , that follows from equation (77), are: l(l + l)X0 + n(L0+N0) [(a-b)l+al]X0-(l (a-b)l
+ -±(P + 2n
= 0,
+ 2)% + n(a + b)N0-n(a-b)L0
\ = 0,
(79)
2l-l) X0 - % + a(l - 1) N0 + \p*(N0 +L0) = 0;
and a fourth equation which is a repetition of the first. From these equations, it readily follows that we may take for the two linearly independent solutions with the behaviour rl at the origin, the ones derived from L0 = 0,
X0
1(1+1) °'
% = +\(l-l)\a and
*o
+
b-j^)[(a-b)l+o-l^N,-
(80)
j~r)[(a-b)l+al^L0.
(81)
= 0
- X° = -W+T)L°' % = -\(l-l)\a-b
Proc. R. Soc. Lond. A (1991)
+
260
S. Chandrasekhar and V. Ferrari
Expansions including terms of 0(rl+2) can readily be found by standard procedures In particular, by assuming an expansion of the form (X,9,N,L)
= (X0,%,N0,L0)rl
+ (X2,%,N2,L2)rl+2
+ O(rM)
(82)
we find (l + 2)(l + 3)X2 + n(N2+L2) + l(a-b)X0 + nb{N0+L0) + a20X0 = 0, (l + 2)(a-b) + cr2]X2-(l + 4:)% + n{a + b)N2-n(a-b)L2 + n(a2 + b2)N0-n(a2-b2)La + l(a2-b2V)X0 + (T20(b-a)X0 = 0, - ^
+
[{l+i)a
+ \a2]N2 + \alL2 +
{l
+ 2)(a-b)
+ [a2l + a2 + b2-b(e0+p0) + (1 + 3) N2+L2
ti
1±HX2
-^L
+ ^(b-a)(l2
a\Q0N0-(Q0-2)L0
where we may insert for (X0,@0,N0,L0)
[
+
(e2+p2)+la2{b-a)]N0
+
+ l(a2-b2)
x
+
2l-l)
Xo + (£-b)% = 0,
l)^±i' X \
+ 2 + (/+l)
0
+ -90 = 0.
either of the two solutions (80) or (81).
7. The interior solution A solution that describes the interior of a star must satisfy certain requirements at the centre, (r = 0), and at the boundary, (r = r^), where the pressure p of the static star vanishes. At the centre the conditions are that 5e and 8^» (i.e. E and 77) both vanish. (A somewhat milder requirement may suffice: but the stronger conditions that E = 77 = 0 and singularity free, are certainly sufficient.) By equations (35) and (37), the conditions are met by the requirement that W and N vanish at r = 0. Of the solutions allowed by the indicial equation (78), only the two linearly independent solutions with the r'-behaviour at the origin are compatible with the physical requirements. These solutions are explicitly determined by the expansions specified in equations (80)-(83). (It may be noted here t h a t the r'-behaviour of the functions N,L,X, and W assures the required flatness of the space-time at the origin.) At the boundary, r = rlt of the star two conditions must be met: 77 must vanish and we must also ensure that the space-time is continuous with the vacuum t h a t prevails outside the star. (Strictly we need only require that 77 vanishes on the displaced moving boundary of the oscillating star. This distinction between the static boundary and the moving boundary is relevant only if e ^ 0 on r = r1. Allowance for this contingency is readily made: but we shall not digress now to consider this eventuality.) The vanishing of 77 clearly requires (under the circumstances envisaged, namely that e and p vanish on r = rx) W = -{T-V+L)=0
at
r=
Tl
(84)
and this is also one of the conditions that the metric perturbations are required to satisfy (cf. M.T., p. 147, eqn (43)) in order to match continuously with the exterior metric perturbations - we consider the remaining conditions in §8 below. The condition (84) on W follows from either of the equations (27) and (37). More stringent conditions on FT follow from equations (26) and (45). The former equation, Proc. R. Soc. Lond. A (1991)
On the non-radial oscillations of a star
261
relating U with the Lagrangian displacement £2, requires U to vanish on the boundary of the star. At the same time, it follows from entirely general considerations, explained in Appendix C, that v r tends to a finite limit v'v and
Q = er/pr->QJ{r1-r)
)
(r^r1-0),J>
where Qx is some constant. By virtue of these behaviours, equation (45) gives, U->Wr + ^^W (r-n\-0). rl — r Therefore, the governing equations already require that W^const.
(r1 — r)
for
(86)
r-^r^ — 0;
(87)
a result that is directly derived in Appendix C. We must in addition require (as a boundary condition that we must impose) Wr = 0
at
r = rx.
(88)
To satisfy this further condition at r = rlt a superposition of the two linearly independent solutions, belonging to the double root x = I of the indicial equation, is needed. For, along the solutions belonging to this double root, W r tends to different constants as r^r1 — 0 - a fact that is established in Appendix C. We can therefore find for each, assigned a2 ( > 0), a superposition of the two solutions, unique apart from a single constant of proportionality, that satisfies the condition (84) on the boundary. In obtaining the required superposition of the two linearly independent solutions, belonging to the double root, x = I, we shall adopt the following prescription. Denoting by
[No=i,Lo
= 0]
[L0 = l;N=0l
and
(89)
the two linearly independent solutions derived from the expansions (80)-(83) at the origin, we determine the constant, q in the superposition, q[N0 = 1 ;L0 = 0] + [L0 = 1 ;N0 = 0],
(90)
by the requirement that Wr = 0 at r — rx. Solutions for the interior, 0 < r ^ rx, obtained in this fashion, satisfying the necessary boundary conditions at r = 0 and r = rx for different assigned a2, are 'normalized' so that L0 = 1
(91)
for all of them.
8. The exterior solution and the procedure for determining the quasi-normal modes In the vacuum that prevails outside the star (r > rx), the equations that govern the metric perturbations are (M.T., pp. 147-150, equations (43), (58), (62) and (63)): T-V+L=0
(92)
- an equation which we have already considered in § 7 as a boundary condition to be imposed on the interior solution for r^r1 — 0- and the Zerilli equation, -^. + dr%
Proc.R. Soc. Lond. A (1991)
J
V<+)Z, where
Z =
L—(^.X-rL\. nr + 3M\ n
)
(93) K '
262 yi+)
=
S. Chandrasekhar and V. Ferrari 2A 2 s + 3Mn2r2 + 9M2nr + 9M3], rb(nr + 3M)2;[n (n+l)r
(94) d A d / 2M\ d J = r -2Mr, —— = —— = 1 — dr dr.,. r^dr \ r Jar and I f is the mass of the star enclosed inside r = rx. In integrating equation (93) into the vacuum, we must, to ensure the continuity of the space-time at r = rv use as starting values for Z and Z , the values they have as r^r1 — 0 from the interior, i.e. .
2 2
Z(r = r,) = lim . r^Ti_0{nr and
(r = rj = [ 1
lim ^1 / r ^ r , - 0
f^Z-rZ
„„,. + 3M\ n
(95)
'3Jf X-rL nr + 3M \ n
(96) ,r
where Z {r = r1)
1 — (nr + 3M)rXr + — X ' n (wr + 3M)s n -(nr + GM)rL-r2(nr
+ 3M)L
(97)
With the starting values for Z and Z , at r = rv determined, as described, from the interior solution, we must integrate the Zerilli equation forward to determine its asymptotic form. One readily finds from the Zerilli equation that for r -* oo, Z must have the asymptotic form (cf. Chandrasekhar & Detweiler 1975, equations (51) and (52); Lindblom & Detweiler 1983, equations (A 37)-(A 40)): Z
n+l/30 1 » ( » + l ) a 0 - p f t r ( l +-)/?„ cr r " 2 ^ n+ 1 a 0 1 %(«,+ !)/? 0 + pfo- 1 + - a 0 r "2^
~ /?„ + a
2
+ ... [• cos a-r,
2
+ ... 1-sin err,.
(98)
appropriate for standing waves and real
On the non-radial oscillations of a star
263
pp. 603-611), that aJ5 + /?Q, as a function of cr ( > 0, say) must exhibit a deep minimum at some determinate cr0 with the behaviour, <** + $ = const.
[(
(T=1M),
(99)
in the neighbourhood of cr0. The origin of this behaviour is simply that, if the star were set in oscillation with some general frequency cr, very little of the energy of excitation will be converted into the mechanical energy of oscillation and almost all of it will escape to infinity as gravitational waves. If on the other hand, the star were set in oscillation with its 'resonant' natural frequency cr0, most of the energy of excitation will be converted into the mechanical energy of oscillation and very little will escape to infinity. (There are some further restrictions for the validity of the formula (99): e.g. the non-existence of other resonant frequencies in the neighbourhood of cr0 and cri <^ cr0 - restrictions which obtain in the contexts considered.) A method for determining the complex frequency of the quasi-normal modes that is suggested, then, is to determine al + fil as a function of cr ( > 0) and match it with the predicted behaviour (99). Examples are provided in §9 below.
9. A direct numerical confirmation of the algorism developed in this paper The algorism developed in this paper for determining the quasi-normal modes of oscillation of a star, is founded on a physical base sufficiently different from that of extant methods, that a numerical confirmation may be useful. We have, therefore, carried through the algorism, numerically, to its completion, for a specific model. The model chosen was a polytrope of index 1.5 with a 0 ( = e0/p0) = 9. For this polytrope, ^ = 1.576412485,
i f =0.20541,
and
e = m/r\
= 0.1573,
(100)
e
where r1 is the radius of the polytrope in the unit V o- (The need for rx to be known to so many decimals is explained in Appendix D.) For this model, a,\ + ji\ was numerically evaluated for various initially assigned values of cr by the procedures described in §§7 and 8 for the quadrupole mode 1 — 2: they are listed in table 1 and a 2 + fi\ as a function cr ( > 0) is further exhibited in figure 1. Table 1 and figure 1 also include a comparison with the matching parabola, al + fil = const. [(o--0.3248) 2 +(1.026 x 10"4)2]
(101)
in the neighbourhood of its minimum. The agreement of the computed curve with the matching parabola is within the accuracy of our calculations. We conclude that for the model considered ^ = 0 i 3 2 4 8 a n d ^ = 1 . 0 2 6 x 1 0 ^ . (102) Dr Lee Lindblom very kindly undertook to determine c 0 and cri for this same model and for I = 2 by the method described by him and Detweiler in their paper on The quadrupole oscillations of neutron stars (1982); and he finds cr0 = 0.3248
and
cri = 1.00 x 10" 4 ,
(103)
in astonishingly good agreement with the values (100).
10. The dipole oscillations The oscillations of a star with a dipole character presents a singular case in the theory of non-radial oscillations: they are not accompanied by the emission of gravitational radiation for the simple reason that gravitational waves with dipole symmetry do not exist. Proc. R. Soc. Lond. A (1991)
264
S. Chandrasekhar and V. Ferrari
0.003
0.002
-
0.001
-
+
0.32425
0.32450
0.3200
0.32475
0.32525
Figure 1. A comparison of the variation with a of the flux of radiation, (c^ + y?2,), of standing gravitational waves at infinity with a matching parabola. The deep minimum occurs at the real part of the characteristic frequency of the quasi-normal mode of oscillation and the width of the resonance is a measure of the reciprocal of the imaginary part. (The calculations are for the quadrupole oscillations of a relativistic polytrope of index 1.5 and e0/p0 = 9.)
Table 1. The parabolic variation of
«s+#t
ccl + fil* 2
0.3006 x 1 0 0.2037 x 10"2 0.1263 x l O 2 0.6876 x l O 3 0.3065 x 10~3
2
0.2988 x 10" 0.2026 x l O 2 0.1259xl0" 2 0.6853 x 10"3 0.3065 x 1 0 3
cr
«! + $ *
*8+flt
0.3248 0.3249 0.3250 0.3251 0.3252
0.1219xl03 0.1317 x l O ' 3 0.3349 x 10~3 0.7323 x 10"3 0.1316 xlO" 2
0.1219 xlO" 3 0.1317 x 10"3 0.3357 x 10"3 0.7341 x 1 0 3 0.1327 xlO" 2
* As derived from the asymptotic behaviour of the solutions at infinity. | As derived from the matching parabola (101). Dipole oscillations in relativistic astrophysics were first considered by Campolattaro & Thorne (1970) and by others since. B u t the theory was completed satisfactorily only recently by Lindblom & Splinter (1989). In the context of the present development, the theory of these oscillations follows as a particularly simple special case. Because of the singular character of these oscillations, it is necessary t h a t we start, ab initio, with the equations (48)-(51) and specialize them to the case 1=1,
n = 0
and
Clearly X = 0,
K = 2.
(104) (105)
265
On the non-radial oscillations of a star and we are left with the three equations:
L^ +
w\ = o,
^-V^L+^W^ --(Q+l)v.r
(106)
r
*.r-(~".r)tf'
—
,r + {-r + v , ) L -w,r = o,
L
(107)
2
(e + p)r i V - | ( l + w r ) L r + ( - + 2 ^ J L + o- e 2(fi2-
e2/*2
-"ViJ
r
-{(l+rv_r)Wi
r
+ lo-2e2^-"hW}
= 0.
(108)
(It is important to observe that for the present case, I — I, equation (32) for V, as an independent equation, does not exist, as follows from its derivation in Appendix A.) Equations (106)—(108) are linearly dependent at r = 0. However, the equation, i v | - ( e 2 ^ - l ) + ^ r - r ( e + p ) e 2 ^ | + r | i ' i J L i r + - j + (r 2 e 2 ( ^-'' ) L| + r(v,r W,r + fo-2e2(^-">W0 = 0,
(109)
obtained by subtracting equation (107) from (108), together with equations (106) and (107), do provide a system linearly independent at r = 0. Thus, by considering at the origin, a behaviour of the form (N,L,W)
= (N0,L0,W0)rx
+ O(rx+2)
(r->0),
(110)
| a ( x - l ) 2 ( a ; + 2).
(Ill)
we obtain the indicial equation, 0 x—l 0
x +2 \{x + 2) — (x+1) —x a(x + 1) + O-Q ax + \v\
We therefore obtain two singularity-free solutions corresponding to the double root x = 1. And we find that from the corresponding homogeneous equations for (N0, L0, W0) that N0 is undetermined and W0 = —2L0. (112) The two linearly independent solutions at r = 0 become determinate when we seek expansions including terms of 0(r3). Thus, by assuming the expansions, (N,L,W) = (N0,L0,W0)r+(N2,L2,W2)rs
+ O(r!i)
(r-0),
(113)
we find that N2, L2, and W2 are determined in terms of N0, L0, and W0 by the equations, 2L2 + W2 = -laQ0L0, 2N2-4L2-3W2 2
= a(L0-N0),
(±a +
) (e2+p2)]N^
(114)
266
o. Chandrasekhar and V. Ferrari
Explicitly, we may take for the two linearly independent solutions at r = 0 the following: 1. Ln 0, N0=l: + (e2+p2)]r3
N = r- — [a* + (b2 + az)-(e0+p0)b
(115) L
+ ai)-
= ^[{b2
W = --[(bt and
(eo +Po)b + (e2 +P2)]r3
+ aJ-(e0+Po)b
2. L 0 = l ,
+ (ea+Pa)]i*
N0 = 0: (116)
L = +r-l(3a + |^)Q0r3, W = -2r + l(±a +
Turning next to the behaviour of the solutions of equations (106), (107), and (109) at the boundary, r = rls of the star, we conclude by the same arguments as in Appendix C, that it is of the form, (L,N,W)^(L1,N1,yW1)e^
(y =
r1-r^0),
(117)
where a. is the root of the characteristic equation, -a-v'1 + 2/r1 + a — v'l — l/rl »"iK( - « + 1/rj) + or2 e"4"']
0 —cc + v'1 — l/r1 v[ + (e~^ - 1 )/r x
W + Qiv'i) 1
= 0.
(118)
Therefore, quite generally, (L,N, Wy-1) -> finite limits {LVNV WJ
as
•^-0.
(119)
Besides the conditions, W=Wr
=0
at
(120)
r = r.
that we must impose, as in §7 (equations (84) and (88)), we must also require in the present instance (1=1) that N = L = 0 at r = r, (121) to ensure that, consistently with the dipole character of the oscillations, no perturbations in the space-time extend into the vacuum outside the star. The vanishing of L,N, and W at r = r1 is sufficient to ensure the identical vanishing of the perturbations in the space-time for r > rlt since from the linear homogeneous character of the governing equations, (L,N,W)
=0
at
:iVr=W% = 0
r
at
r = rv
(122)
The remaining requirements, L=N=Wr
=0
at
r = rlt
(123)
would appear to be one too many. But this is not the case: for, as we shall now show, Proc. R. Soc. Lond. A (1991)
267
On the non-radial oscillations of a star
Table 2. Illustrating the algorism for determining the fundamental characteristic frequency of dipole oscillations for a relativistic polytro-pe of index 1.5 and e0/p0 = 9 (For each o% the first row gives the values of Wr, L, and N, at the boundary, r = rlt for the solution [Nn = 1; L0 = 0] and the second row gives the corresponding values for the solution [L0 = 1; •^o = 0]- (Note the simultaneous vanishing of Wr for the two solutions for a = 0.4786.)) a
L
N
0.2
0.1972 xlO 0.6262
1
-0.5735 -0.4726
0.2989 0.1150
0.4786 -0.1362x10-* -0.5422 xlO'5
0.3
0.1102 x 101 0.2363
-0.4981 -0.3973
0.3934 0.2294
0.479
0.45
0.1083 -0.1883 xlO" 2
-0.3708 -0.2840
0.5580 0.4162
0.478
0.1972 xlO" 2 -0.1356 xlO" 3
-0.3480 -0.2657
0.5883 0.4490
W.r
a
L
N
-0.3475 -0.2654
0.5890 0.4497
-0.1331 x l O 2 0.1015 xlO" 3
-0.3472 -0.2651
0.5894 0.4502
0.5
-0.6300 x l O 1 -0.7270 xlO" 2
-0.3307 -0.2523
0.6117 0.4740
0.6
-0.1795 0.6140 x l O 1
-0.2600 -0.2015
0.7098 0.5764
W.r
any solution of equations (106), (107), and (109) for which W = W r = 0atr = r^N and L become determinate multiples of one another. The result stated follows from the fact that for any solution for which W = W r = 0 at r = rv equations (106), (107), and (109) tend to the vacuum equations given in M.T., pp. 151 and 152. therefore, the solutions f o r i andN, in these cases, approach those for the vacuum. In particular, fromM.T., equations (65), (71) (or (73)) and (75) on pp. 151 and 152, it follows t h a t e" 3M L = const. - = —<£,
and
N=\M
Therefore,
„ I N=
e
„ . M2 + cr2ri G
(124)
K-
(125)
\
(^ r i-°)-
X—7W^ ~T
<126)
i.e. one becomes a determinate multiple of the other, as stated. (We have numerically verified the validity of the relation (126).) By virtue of the relation (126), we are left only with the boundary conditions, L
or
N=0
and
Tf r = 0
at
r = rx,
(127)
to satisfy, since W = 0 at r = r1 for all solutions. A solution satisfying the boundary conditions (126) can be found as follows. We have shown that for any assigned a2, along all solutions of equations (106), (107) and (109) and, therefore, in particular along the two singularity-free linearly independent solutions derived from the expansions (115) and (116), N, L, and Wy'1 will tend to determinate finite limits, Nlt Lx, and Wx as y ( = r1 — r) tends to zero (cf. equation (117)). By varying cr2, we can find solutions, as in §7, for which W r vanishes. Since, in specifying the solutions for which W r vanishes, no member of the twoparameter family of the singularity-free solutions is distinguished. W r will vanish simultaneously, for the two linearly independent solutions derived from the expansions Proc. R. Soc. Land. A (1991)
268
S. Chandrasekhar and V. Ferrari
(114) and (115). (This conclusion, while it is intuitively manifest, can be established directly with the aid of equation (118).) And, finally, by a linear superposition of the two linearly independent solutions for which Wr vanishes, simultaneously, at r = rx, we can find one for which Nand, therefore, also L vanishes at r = rx; and the solution will be completed. An illustrative example The algorism we have described for determining the characteristic frequencies of dipole oscillations of a star, is illustrated in table 2 for the same polytropic model (n = 1.5 and e0/p0 = 9) considered in §8. The values attained by W r, L, and Nat the boundary, r = r 1; are tabulated for the two solutions. [N0 = 1 ;L0 = 0]
and
[L0 =l;N0
= 0],
(128)
derived, respectively, from the expansions (114), (115) and (116). I t will be noticed that, as predicted, W r vanishes simultaneously for a particular a ( « 0.4786 in the example considered). We readily find from the tabulated values that for the solution, [N0= 1;L0 = 0 ] - 1 . 3 0 9 [L0 = l;iV0 = 0]
for
a = 0.4786,
(129)
the boundary values attained by W r, L, and N are: JF r = - 5 . 2 x 1 0 - * ,
L%4xl0"4,
and
iV«5xl0"4.
(130)
We conclude that the solution (129) satisfies all the required boundary conditions as accurately as one would wish. Dr Lindblom and Dr Splinter have (at our request) determined, for the same polytrophic model, by their alternative algorism, the characteristic frequency a = 0.4786 in ' e x a c t ' agreement with our determination.
11. The axial perturbations We now turn to the axial perturbations of a star. For such perturbations, the metric is of the form, ds 2 = e2"(dt)2-e2'lr(dcp-o)dt~q2dx2-q3dx3)2-e2^(dx2)2-e2^(dx3)2,
(131)
where v, \jr, /i2, and /i3 retain their unperturbed values (given in equations (2)-(9)) while a), q2, and q3, defining the perturbations, are functions of t ( = x°), x2 ( = r), and x3 (= 6). Besides, these perturbations, by definition, are not accompanied by any motions in the r- and the ^-directions. Accordingly, the components of the fourvelocity, ii(2) and « (3) (in the tetrad-frame) vanish identically. In the first instance, only the (^-component, u(1), is allowed to be non-zero (though as we shall presently show, it too vanishes). Since for the perturbations considered, ^(l)(2) = (e+P)UWU(2)
= T(l){3) = (e+P)«(l)M(S) = 0 .
(132)
it follows from the field equations (M.T., p. 143, equations (11) and (12)) that
where
QAB
=
( e 3 ^ - ^ 3 Q 2 3 ) , 0 + e 3 ^ - ^ Q O 2 | O = 0,
(133)
[e*++<>-»rHQm)
(134)
QA,B-°B,A
Proc. R. Soc. Lond. A (1991)
and
QAO = QA,O~U,A
(-4 = 2,3).
(135)
On the non-radial oscillations of a star
269
The integrability condition of equations (133) and (134) is ( e 3 ^ - ^ Q 0 2 U o + ( e ^ - ^ - ^ o s U o = 0For a time-dependence e (
lcrt
e
(136)
of the perturbations, equation (136) is equivalent to
3 ^ ^
0 8
)
r
+ (e^-"+^-^Q03)te
= 0.
(137)
By the (Ol)-component of the field-equations (M.T., p. 141, equation (d)), equation (137) implies that Bma) = +2T ( 0 ) ( 1 ) = +2(e+p)uwu(1) = 0. (138) In other words, ^ = Q (139) We conclude that incident gravitational waves do not excite any fluid motions in the star. The star simply scatters the incident radiation by the potential barrier (derived from the curvature of the space-time of the star) that it presents. Making use of the relation (cf. equation (110)), ^20,0
^B30,r
=
(923
liil.O
=
^23,0'
(14U)
we derive from equations (133) and (134) the wave-equation, [e-3!^-A!+ft(e3^-ftr7.3£23)
r ] rJr[e-H^+H-h(Q^+"-H-HQ23)
e]^
= <2 23i0i0 .
The further reduction of this equation proceeds exactly as in M.T., 143-144. Thus, equation (141) can be separated by the substitution, e^^-HQM
(141)
§24(a), pp.
= X(r) CfUO),
(142)
where Cn denotes the Gegenbauer function as defined inM.T., p. 144, equation (20). We find: r
2
e ^ / ^ I
f
j
- 2 ? i ^ I + e r 2 I = 0,
\ji2 = 2n = (I-1) (1 + 2)],
(143)
where we have made use of the relations, g-s^-ft,-^ _ gv+z^-4 c o g e c 3 Q
and
e-w+>>-/'l+/'a
= e"-^r-i
cosec3 d.
(144)
Equation (143) can be reduced to the form, e 2/*2
XITIT
,,2
{2 + r2[e-p-GM(r)/r3]}Xir-e2^X+o-2e2^-")X
= 0.
(145)
I t is of interest to notice t h a t by defining the variable, rt=
e -"+^dr
(146)
Jo analogous to the variable r, in the treatment of the perturbations of the Schwarzschild black-hole (M.T., p. 144, equation (25)) and letting X = rZ, (147) equation (145) can be brought to the form, ~^)Z=VZ, Proc. R. Soc. Land. A (1991)
(148)
S. Ghandrasekhar and V. Ferrari
10 -
y
/
2M 0.5
1.5
1.0 r
2.0
Figure 2. A comparison of the potential barriers, presented to incident axial gravitational waves, by a relativistic polytrope of index 1.5 and e0/p0 = 9 and by a black hole of the same mass. In the vacuum exterior to the boundary of the star (at r ^ r±) the potentials are the same. They are different for r < rx\ the potential for the black hole vanishes at its horizon r = 2M while that of the polytrope tends to infinity for r^O.
where
V = -£ [{/i2 + 2)r +
r3{e-p)-6M(r)].
(149)
Outside the star, where p = e = 0 and M(r) = M, the potential V reduces to the ' Regge-Wheeler' potential of the Schwarzschild black-hole for axial perturbations (see equation (151) below). I t should, however, be noted that, unlike in the case of the black hole when r, ranges from — oo to +oo, the variable r„, as defined in equation (146) ranges only from 0 to + oo. The potential Vdefined in equation (149) is strictly a central field; and equation (145) describes a scattering problem for a 'soft-core' Coulomb-potential. (See figure 2 in which a comparison is made of the potential for the relativistic jDolytrope, n = 1.5 and e0/p0 = 9, with that for the Schwarzschild black-hole of the same mass.) A study of the scattering of axial gravitational waves by a star, with the aid of equation (145), for the interior of the star (r ^ r^), and the 'Regge-Wheeler' equation, / J2 \
^ where
+ cr2\Z=V^Z,
F(-) = _[( / t 2 + 2 ) r - 6 M ] ,
for the vacuum exterior to the star (r > rx), presents no difficulty. Proc. R. Soc. Land. A (1991)
(151)
On the non-radial oscillations of a star
271
The solution of equation (145), free of singularity at the origin, has the expansion, X = rl+* + 2{J+3){(l
+ 2)[U2l-l)e0-p0]-
+ ....
(152)
With the aid of this expansion, equation (145) can be integrated forward to the boundary, r = rv of the star. The integration can then be continued into the vacuum outside the star with the Regge-Wheeler equation (150) with the starting values, Z(r = rj = lim (X/r)
and
2M\
Z r ( r = r1) = | l
r
lim
l I r^r^-0
(153)
\(rXtr-X)
(154)
The integration of equation (150) must be continued to a sufficiently large r that, matching with the asymptotic expansion (cf. equation (98)), Z^ + \a0
nA-1 /?
1
w
1
V—^-—2 - 7 - 2 ^ NM ^ + +l )!)a ooc-0-3Ma/3 3 M ^ 0o]-+...\ ] -
/?o + — — -^-2[n(n+l)/S0
cos ar,
+ 3Maa0]-2+...\smar^ r J
(155)
will enable us to determine a 0 and /?0. Determining a0 and /?„, in the present context, is equivalent to determining the 'phase-shift' in the standard terminology. To exhibit the differing scattering properties of a star for the axial and the polar perturbations, we present in figure 3, for comparison with figure 1, the variation of (ajj + /?j|) with a, for I = 2 and the same polytropic model. I t is not an uncommon view that the axial perturbations of a star present but an uninviting subject for study since the incidence of axial gravitational waves does not excite any fluid motions in the star. But this view overlooks certain facts germane to the 'conformity of the p a r t s ' of general relativity 'to one another and to the whole.' I t is known, for example, that the scattering of axial and polar gravitational waves by the Schwarzschild black-hole (and, indeed, also by the Reissner-Nordstrom black-hole) is isospectral in that the reflection and the transmission coefficients for both types of waves are the same. While we do not, of course, expect that the scattering of axial and polar gravitational waves by spherical distributions of matter will be as closely related as they are in the case of black holes, it would, nevertheless, strike a discordant note if the scattering of axial gravitational waves depended only on the static space-time and it was not so for polar gravitational waves. It is consonant with the character of general relativity that the two cases share the same property in this important respect. Also, the particular simplicity of the analysis for the barotropic case must be attributed to the circumstance that in this case the treatment of the perturbation problem requires no additional information (such as, for example, the conservation of baryon number) beyond that needed for the specification of the static space-time.
Proc. R. Soc. Land. A (1991)
S. Chandrasekhar
272
and V.
Ferrari
+5 +4 +3 O
+2
+ +1 o
o
0
-1
0
0.5
J
I
I
1.0
1.5
2.0
cr
Figure 3. The variation with cr of the flux of radiation, (<xj^+ /?„), at infinity of standing axial gravitational waves for the same relativistic polytrope considered for polar perturbations in figure 1.
12. Concluding remarks We shall dispense with what one normally expects in the concluding section of a paper: a restatement of the underlying ideas, an outline of the development, and a final summing. The two long introductory sections, the detailed summary, and the substantive comments in §§10 and 11, amply take their place. And we shall also not enumerate the many directions in which the present investigation can be extended: they are numerous and they are obvious. We shall confine our remarks, instead, to an aspect of the theory to which we have not made any reference so far: a comparison with the Newtonian theory. The simplicity of the algorism developed in this paper for determining the complex characteristic frequencies of the non-radial modes of oscillation of a star, relative to the traditional ones, requires no emphasis in the relativistic context. What is astonishing is its simplicity relative to the corresponding Newtonian algorism (cf. Chandrasekhar 1964; Hurley et al. 1966). Indeed, by contrasting with the Newtonian treatment of non-radial oscillation, one becomes aware of a source of even greater puzzlement. On the relativistic theory, the frequencies of oscillation of the non-radial modes (as we have shown) depend only on the distribution of the energy-density and the pressure in the static configuration and the equation of state only to the extent of its adiabatic exponent. If this is a true representation of the physical situation, then it must be valid in the Newtonian theory as well: the true nature of an object cannot change with the mode and manner of one's perception. In the relativistic picture, the independence of the frequencies of the non-radial modes of oscillation of a star, on Proc. R. Soc. Lond. A (1991)
On the non-radial oscillations of a star
273
anything except its characterization in terms of its equilibrium structure, is to be understood in terms of the scattering of incident gravitational waves by the curvature of the static space-time and its matter content acting as a potential. But what are the counterparts of these same concepts in the Newtonian framework ? Perhaps they lie concealed in the meanings that are to be attached, in the Newtonian theory, to the four metric functions {and their perturbations) that describe a spherically symmetric static space-time {and their polar perturbations). I t is known that the Newtonian gravitational potential, in some sense, replaces the metric function gtt. Are there similar meanings to be attached to grr, ggg, and gw ? That is the predominant question to which the present investigation seems to lead. We are grateful to several colleagues for their patience in discussing with us different aspects of the problem considered in this paper: to John Friedman and Robert Wald on the physical aspects and to Norman Lebovitz and Sotirios Persides on the linear dependence, at the origin, of the basic system of differential equations; to Kip Thorne for extremely valuable decisive critical remarks; and to Jesus Ibanez for a careful scrutiny of the entire analysis. We are also very grateful to Bernard F. Whiting who independently pointed out to us a serious error which vitiated an earlier version of this paper. But our greatest indebtedness is to Lee Lindblom for generously sharing with us his experience with his alternative treatment of the quadrupole oscillations of neutron stars; and particularly for determining for us afresh by his methods the quasi-normal quadrupole mode of oscillation of the polytrophic model considered in §9. We are similarly grateful to Lee Lindblom and Randall Splinter for determining the characteristic frequency for dipole oscillations for the same polytrophic model. The research reported in this paper has, in part, been supported by grants from the National Science Foundation under Grant PHY-89-18388 with the University of Chicago. We are also grateful for a grant from the Division of Physical Sciences of the University of Chicago which has enabled our continued collaboration by making possible periodic visits by Valeria Ferrari to the University of Chicago.
Appendix A. The derivation of equation (32) We have already remarked in §4c the absence of any explicit derivation (or an outline of a derivation) of equation (32) in spite of its remarkable character: its apparent independence of the nature of the source of the static space-time. We start with equation (22). Making use of the relation, = 2pe2*>
vtTtT + \- + vMvtT-iiiir)
(A 1)
(which follows from the unperturbed equation, G33 = 2p) and substituting from equations (23) for the various quantities describing the perturbation, we find after some rearrangement of the terms: \-4pLe** {
+ {T+N)
rr
' '
+ (v r-/i2r){T+N)
r
'
+
-T. r '
+ ^+vr){N-L)r-2e2^n+(T2e2^-''){T + \v,r,r + (-+v,r-/*2,r)v,r
+ ^(N+L)
+ L)\pi + a2e2^->)v\piiecotd
= 0, (A 2)
i.e. a linear combination of terms with the angular factors Pl and P, 9 c o t # vanishes. Proc. R. Soc. Lond. A (1991)
274
S. Chandrasekhar and V. Ferrari
We can accordingly equate separately the terms with the two factors. (Exactly the same remarks are made in Chandrasekhar & Xanthopoulos (1979) following equation (71) in the context of the perturbations of the Reissner-Nordstrom black-hole.) Perhaps an explanation for this statement is needed. Consider quite generally an equation in the form A(r)PM-B(r)/,Plt/1
=0
(,* = cos0).
(A3)
By making use of the known recurrence relations among the Legendre polynomials, we can rewrite equation (A 3) in the form (A-lB)Pl-B(r)Pl_lt/l
= 0.
(A 4)
Since Pi-it/l is a polynomial of degree I — 2 in ft, it is orthogonal to Pl: and, therefore, A-lB and
=0 A-B
We conclude that
A=B
and
B =0
=0 =0
(1>1),\
(1=1). if
(A 5)
J
1>1.
(A 6)
Applying this last result to equation (A 2) we conclude that /2
\
v
,r,r + [-+v,r-H,M,r
e2/42
+ ^{N+L)
+ aie2^-'')V=0,
(A 7)
so long as I > 1. I t is to be particularly noted that equation (AT), as an independent equation, does not exist for 1=1. I t can be readily verified that the vanishing of the terms in equation (A 2) with the factor P ; (I > 1) contains no more information than equation (30).
Appendix B. Relativistic polytropes Relativistic polytropes have been considered extensively, in various contexts, since the early sixties (cf. Chandrasekhar 1964; Tooper 1964). Nevertheless, some of the essential relations that we needed for the numerical illustrations in §§9 and 10 were not readily available in the extant literature (at least in the forms we needed them). The present Appendix is provided on that account. The fundamental assumption of the theory is that the energy density e and the pressure p are expressible in the manner, e = eo0n
and
p = po0n+1,
(B 1)
where e0 and p0 denote the values of e and p at the centre and n is the polytropic index (not to be confused with the 'n' defined in equation (25)). By measuring r (and t) in the unit Veo> the relativistic equation of hydrostatic equilibrium (equation (6)) takes the form: 1
2M(r)
(n+l)0tr
= -(ao + 0)
*£)+!*-,
(B2)
where, in accordance with our present convention regarding units, e = 0n,
p = 0n+1oc„\
M(r)=\
6>Vdr, Jo
Proc. R. Soc. Lond. A (1991)
and
oc0 = ejp0.
(B 3)
On the non-radial oscillations of a star Equation (3) for v
275
now gives .. ,r
^ = _ e+p
( n + 1 )
_^_ aQ + 0
(B4)
or, after integration,
The constant of integration, y0, is to be determined by the condition (cf. equation (10)) foe +i\2
At r = rx, 0 vanishes linearly as (r1 — r) with its derivative tending to a finite negative limit 0\: <9 ^ 16>il ( ^ - r ) (r-^-0). (B8) The value of \0'^\ is an important constant of the theory. I t is best determined by the equation, (
*9)
a0 + 2 ( n + l ) r 1 | 6 ) ; | '
for M (which follows directly from equation (B 2)), since M(r) attains its limiting value long before r1 (to a sufficient accuracy). Using equation (B 9), we can rewrite equation (B 6) more conveniently in the form, / e *".=
\2(»+l)
-?i2_]
^2
.
(BIO)
From equations (B 4) and (B 7) it follows that vr-(n + l)^il=K
(say).
(B 11)
a0 Also, by the choice of v0 (by equations (10) and (B 6)), li2^-v, /i2,r^--v,r, and e "^e "i = l-2M/r1 (r-^-0). Besides, in accordance with equations (B 7) and (B 8) 2
2
Q^Qi/{ri-r) (r-^-0), where
Q1 = na0/(n+
1)\0[\.
A useful relation which follows from equations (B 11) and (B 15) is Q1 v\ = n. Proc. R. Soc. Lond. A (1991)
(B 12) (B 13) (B 14) (B 15) (B 16)
276
S. Chandrasekhar and V. Ferrari
Finally, we may note the following series expansions for 0, e, and p at r = 0: 0 = i+6> 2 r 2 + 6> 4 r 4 +..., e = e0{l+nd2r2 + [ndi + ln(n-l)dl]ri+...}, ) p = p0{l + (n+l)d2r2 + [(n + i ) 0 4 + i ( w + l ) w ^ ] r 4 + . . . } ,
(B 17)
where
"'-
e
°2(»+l)U + «a 1
1 ,
*—Jtyi+l)3V
^J^
*>+l
(B18)
°4(n+l)
(In writing the foregoing expansions, we have, for convenience, abandoned the convention \/e0 = 1.)
Appendix C. The behaviour of the solutions at the boundary of the star In obtaining the behaviours of the solutions of equations (48)—(51) for r tending to the boundary rl of the star, it is important to note that quite generally, vr tends to a finite limit as r^r1—
0.
(CI)
Proof. By combining equations (3) and (6), we have the relation, 'r
pr+M(r)/r2 l-2M(r)/r'
(C2)
and therefore, M/r\ l-2M/r1
say) for
r^r1
— 0.
(C 3)
(It can be verified that v\ given by this equation agrees with equation (B 11) for polytropes.) Also, by the choice of v0 in equation (9), /i2 = — v
and
l^i
&
t
r = rx
On the other hand, we may expect on entirely general grounds that Q ( = which occurs in equation (48) has the behaviour, Q^Q1/(r1~r)
(r-^-0),
(C 4) er/pr) (C 5)
where Q1 is some constant. For polytropes, Ql has the value Q1 = nocJ(n+l)\0'1\
while
Q1v'1 = n.
(C 6)
From the behaviours (C 3)-(C 5) of the various coefficients that occur in equations (48)-(51), we conclude that near the boundary, r = rv the solutions take the forms, (L,N,X,W)^(L1,N1,X1,yW1)e*u
(y = r1-r),
(C 7)
where a, L1,N1,X1, and W1 are constants. (We have already remarked in §7 on the necessity of W vanishing at the boundary.) We may note t h a t according to the substitution (C 7), W^^-W^ (y^O). (C8) Proc. R. Soc. Land. A (1991)
On the non-radial oscillations of a star
277
Making the substitution (C 7) in equations (48)-(51) we obtain, by virtue of equations (C 3)-(C 5) and (C 9), the following characteristic equation for a:
Lx -a-v'1 + 2/rl + a, — v[ — l/r1 -a(l+rv[) + 2v'1 + o-2e"4\-(«e"2'-l)/r1 ne - 2 " 1 /^
# i
0 — a, + v[ — l/r1 + cc + (n+l)e-2^/r1 ne^2"1/r1
x, — a. — v'1+i/rl 0 -a.(\+rv'1) + e~2^/rl + a2 e - 4 "^ 2 r 1 [ a - 2 a ( ^ + l/r 1 )] + cr2e-4"ir1
Wr -i(i+0i"i I
+ (l + n>i)
= 0.
(C 10)
0
This is a quintic for a. In terms of its five roots, tx,} (j = 1,..., 5), the behaviour of the solutions at the boundary will be given by a superposition of the basic solutions, (LpNpXpyW})e^\
(C 11)
where (LpNpXp W^ represents a characteristic vector belonging to a.p I t follows that in general L,N,X, and W r tend to finite determinate constants as r^r1 — 0.
Appendix D. Some details on the procedures adopted in the numerical integrations In this Appendix we shall give some details on the procedures adopted in the numerical integrations to achieve the necessary accuracy in the derived characteristic frequencies of the quadrupole (I = 2) and the dipole (1=1) modes of oscillation: they may be useful for others who may wish to use the algorisms developed in this paper. The integration of the equations for the polytrope n = 1.5 and e0/p0 = 9 was performed by a simple version of the standard Runge-Kutta routine with a constant integration step; and all the calculations were carried out in double precision. The integration was started at r0 = 0.01 by using the series expansions given in §§6 and 8. Since all the metric functions have the behaviour, rl(ct +
fir2),
(D 1)
the percentage error that is made by neglecting the next term, yr 4 (say), in the series expansion is y r 4 / a « 10 - 8 . (D 2) (a) The quadrupole case Starting from r0 (=0.01) we integrate the two linearly independent solutions derived with the expansions (80)-(83). Our object, of course, is to find by superposition of these two solutions (by the prescription (90)) a solution for which W r vanishes at the boundary r = r1 where the polytropic function 0 vanishes. And we need also the values of the functions X and L and their derivatives at r = r,. Proc. R. Soc. Land. A (1991)
278
S. Chandrasekhar
and V.
Ferrari
Table A 1. (For explanation see text.
(a) (b) 1.5764 (a) (b) 1.5764123 (a) (b) 1.55
X
X,
-0.5501568 -1.0166848 -0.5589986 -1.0472447 -0.5590026 -1.0472589
-0.3401758 -1.1600380 -0.3846175 -1.1549196 -0.3297308 -1.1549169
w -0.1419385 +0.2857576 -0.1563445 +0.2341168 -0.1563465 +0.2340938
-0.7029284 -1.8964062 -0.1685542 -1.8805155 -0.1550421 -1.8716880
w,
-0.3022427 x 10"1 1.2871274 -0.7022729 x 10"1 2.5341350 -0.9675857 x 10"5 0.7800915 -0.3306309x10-" 2.6514366 -0.1419701 x lO"6 0.7665871 -0.4897413 x 1 0 6 2.6426765
Table A 2. (For explanation see text. dr = 10" dr = 10dr = 10-
r= 1.576412485: r= 1.5764124853: = 1.57641248534:
(a) (b) (a) (b) (a) (b)
LT Lr Lr Lr Lr Lr
= -0.1533122, = -1.8705368, = -0.1531958, = -1.8704600, = -0.1531701, = -1.8704437,
Wr = 0.7648572, Wr = 2.6415263;
0.7647808, wr == 2.6414495;
W
,r
wr == 0.7647151, 2.6414332.
Wr
I t was found that the determination of the values which the various functions attain at r = r1 is a very delicate matter. And to check whether the desired accuracy was reached, the range of integration, r,0n
(1) r0 < r < ra;
ra= 1.55,
dr = 10
(2) ra^r^rb;
rb = 1.5764,
dr = 10"6
(3) r 6 s$ r ^ r c ;
rc = 1.5764123,
dr = 10~8
(4) r c < r < r i ;
^ = 1.5764124853,
dr = 10" 11 .
(D3)
The criterion that was adopted in the choice of the step-size dr, for each interval, was that the same integration carried out with a step ten times smaller affected the values of the functions and their derivatives, at the end points of the intervals, only in the eighth significant figure. An illustrative example is provided in table A 1. The entries in the rows (labelled (a) and (&)) are the values of X, L, and W and their derivatives at the end points of the first three intervals (with the respective step-sizes 10~4, 10"6 and 10~8) for the two solutions derived, respectively, with the expansions (80)-(83). I t will be observed that by the end point of the third interval W « 0 and X and L have attained their limiting values to the desired accuracy. Thus, (a) X = -0.5590027,
L = -0.1563465,
(6) X = -1.0472591,
L = 0.2340934,
W ~ 10"7 • W~ 10
P}
(D4)
But it will be noticed (cf. table 1 A) that L r and Wr continue to change. We must therefore continue to integrate beyond the end point of the third interval. The results of further integrations with decreasing step-sizes are shown in table A 2. From the results in table A 2 we conclude that a further reduction in the step-size will affect only the sixth significant figure and that a step-size 1 0 _ u is needed in the last interval. With the necessary accuracy achieved for the functions and their derivatives at r = rv the integration of the Zerilli equation for r > r1 was carried out with a step-size dr, = 0.03. For r. > 25/
On the non-radial
oscillations
of a star
279
(98), is well established and the determination of a„ + /?o does not present any difficulty. All the values given in table 1 are therefore reliable to the number of decimals retained. (b) The dipole case We integrate the two independent solutions with the initial conditions (115) and (116) using the same procedure described for the quadrupole case. In this case the characteristic frequency is identified by the simultaneous vanishing of Wr for both solutions at the boundary. And it turns out t h a t in order to obtain a result accurate up to the sixth significant figure we again have to use the same values for the integration steps used in the quadrupole case.
References Campolattaro, A. & Thome, K. S. 1970 Astrophys. J. 159, 847-858. Chandrasekhar, S. 1964a Astrophys. J. 140, 417-433. Chandrasekhar, S. 19646 Phys. Rev. Lett. 12, 114-116; 437-438. Chandrasekhar, S. 1964c Astrophys. J. 139, 664-674. Chandrasekhar, S. 1983 The mathematical theory of black holes. Oxford: Clarendon Press. Chandrasekhar, S. & Detweiler, S. 1975 Proc. R. Soc. Lond. A 344, 441-452. Chandrasekhar, S. & Ferrari, V. 1990 Proc. R. Soc. Lond. A 428, 325-349. Chandrasekhar, S. & Xanthopoulos, B. C. 1979 Proc. R. Soc. Lond. A 367, 1-14. Cutler, C. & Lindblom, L. 1987 Astrophys. J. 314, 234-241. Hurley, M., Roberts, P. H. & Wright, K. 1966 Astrophys. J. 143, 535-551. Landau, L. D. & Lifshitz, E. M. 1977 Quantum mechanics: non-relativistic theory. London: Pergamon Press. Lindblom, L. & Detweiler, S. 1983 Astrophys. J. Suppl. 53, 73-92. Lindblom, L. & Splinter, R. J. 1989 Astrophys. J. 345, 925-930. Thome, K. S. & Campolattaro, A. 1967 Astrophys. J. 149, 591-611. Thorne, K. S. 1969 Astrophys. J. 158, 1-16. Tooper, R. F. 1964 Astrophys. J. 140, 434-459. Received 23 July 1990; accepted 27 September 1990
Proc. R. Soc. Lond. A (1991)
On the non-radial oscillations of a star. IV An application of the theory of Regge poles BY SUBRAHMANYAN
C H A N D R A S E K H A R 1 AND V A L E R I A
FERRARI2
1
2
ICRA
University of Chicago, Chicago, Illinois 60637, U.S.A. (International Centre for Relativistic Astrophysics), Dipartimento 'G.Marconi', Universita di Roma, Rome, Italy
di Fisica
I t is shown how the flow of energy in the form of gravitational radiation, through a star in non-radial oscillations, can be determined by a suitable adaptation of Regge's theory of potential scattering in the quantum theory. While the resonant scattering of axial gravitational waves can be treated by Regge's theory, essentially, in its original form, the treatment of the resonant scattering of polar gravitational waves requires a generalization of the theory beyond its standard range and the incorporation of a 'flux integral' derived from independent considerations. Illustrative examples of the applications of the theory are provided.
1. I n t r o d u c t i o n We have shown in a recent series of papers (Chandrasekhar & Ferrari 1991 a-c, referred to hereafter as Papers I, I I , and I I I ) t h a t the non-radial oscillations of a star, viewed as a problem in the scattering of gravitational waves by the curvature of the space-time, is, in all essential respects, the same as the problem of resonant scattering encountered in atomic and nuclear physics. As is well known, resonant scattering always occurs when a system is in a quasi-stationary state t h a t decays (as a radioactive nucleus with the emission of an a-particle) or in the context of potential scattering, when the potential including the centrifugal potential has a deep enough minimum followed by a maximum. In atomic and in nuclear physics, resonance is indicated when the cross section for scattering, as a function of the frequency a (or, equivalently, of the energy E in the quantal context) exhibits a sharp peak a t some cr0 with a half-width ai ( crQ) in accordance with the Breit-Wigner formula, cross section =
const. -z . (o--<x0)2 + crf
(1)
An alternative description of the phenomenon (more useful in the context of the scattering of gravitational waves) t h a t provides at the same time an algorism for determining cr0 and <7j, is in terms of the square of the asymptotic amplitude, a 2 + /?2, of the standing waves t h a t prevail a t infinity, when the boundary condition at the centre and a t such other interfaces as m a y exist, are satisfied. And the resonance at
(2)
[133]
Numerous examples illustrating this behaviour (in the context of the scattering of gravitational waves by a star) will be found in the papers to which reference has been made. According to equation (2), the amplitude of the standing-wave a t infinity has, in the neighbourhood of cr0, the behaviour, a-t-i/? « const, x (cr — cr0 — icril
(3)
(It may be noted, parenthetically, that the amplitude of the standing wave at infinity as we have now defined (and implicitly also in Papers I—III) appears to have played no role in earlier discussions of this problem; but it is crucial for our considerations (for additional comments on this and related matters see §5)). In so writing, the standing-wave amplitude, A(cr), is extended to complex cr; and one assumes, or deduces from more general principles, that A(cr) is an analytic function in the complex cr-plane. On this premise, A(cr) in the neighbourhood of cr0, on the real axis, will have the behaviour, l"3^)! 3cr
[ o - f a . + krj)];
(4)
and relation (2) will directly follow: \A(o-)\2 = ** + p2 =
3.4(0-) do-
[(o—
(5)
At cr = cTu + icTj, we have a pole in the complex cr-plane corresponding to the occurrence of resonance at cr0 with a half-width cri in the Breit-Wigner formula (1). An approach to the theory of potential scattering, alternative to the one we have described and developed by Regge (see Alfaro & Regge 1963), is applicable when the wave equation is separable and the wave function is expressible as a product of a radial function, Z(r)/r, and a Legendre polynomial, Pt(cosd) (see equation (7) below). In these cases, scattering by states belonging to different integral values of I can be considered separately, and the standing-wave amplitude A (cr) is defined for each I. In Regge's theory, both a and I are considered as complex variables; and the standingwave amplitude, now a function of cr and I, is assumed to be analytic in both variables. In this paper, we shall show how an application of Regge's theory enables us to close some essential gaps in our present understanding of the non-radial oscillations of stars in general relativity. The plan of the paper is the following. In §2, we give a brief account of Regge's theory adapted to our needs. In §3, we show how the theory as described finds an immediate application to the resonant scattering of axial gravitational waves considered in Paper I I I . And in §4, we extend the theory to apply to the scattering of polar waves.
2. A n elementary account of Regge's theory The elements of Regge's theory are best described in the context of the theory of potential scattering in central fields. The radial wave equation t h a t underlies the theory is d2Z
[134]
1(1+1)
•U(r)
Z = 0,
(6)
where h'tr) i» the central potential (usually restricted to be of short-range, i.e. fJ(r) < olr l) for r->oo) and l(l+l)/r2 is the centrifugal potential derived from the angular factor Pt(cos0) in the separated wave function, ^ = ^P;(cos0).
(7)
VVhile /, aa introduced, is allowed all positive integral values, in Regge's theory / = 0 in excluded; and the characteristic values of I considered are Z = Z 0 = 1,2,3, etc.
(8)
I t should be noted t h a t for the scattering of gravitational waves by the curvature of the s p a w - t i m e and the matter content of the star - the problem in which we are primarily interested - the angular factor need not be P,(cos0): it is, for example, the Gegenbaucr polynomial, C^2(cos0), for the problem considered in §3. Besides, the .should also be excluded, essentially for the reason t h a t dipole oscillations c a H e 1=1 emit no gravitational waves. And, finally, even though the scattering of gravitational waves always occurs in a Coulombic potential, the restriction to a short-range potential in equation (6) does not affect the applications we have in view since we shall ' a b s o r b ' the logarithmic singularity, in the arguments of the sine and the cosine in the behaviour of the solutions at infinity, in a redefinition of the independent variable: r + in the place of r (see equation (30) below). Returning to equation (6), we first observe t h a t freedom from singularity at r = 0, requires t h a t Z has, a t the origin, the behaviour, Z = rl + 0(rl+t)
(r-+0).
(9)
Solutions, for different initially assigned real values of
(10)
where a(
(11)
In the theory of potential scattering, we are accustomed to considering A (
dA(a0;l) dl
[J-(J 0 + Hi)];
(12)
[135]
and we deduce M(
(a 2 +/? 2
[(l-kf + ni
(13)
[(o—
(14)
Comparison with the equation (cf. equation (5)) 2
M{cr;l0) 9
(a 2 +/? 2
derived from the existence of the pole at
=
M(a;l0) cV
(15)
Equation (13) can clearly be made the basis for a determination of £4 for a value of
+
„ . _ * & + ! ) _ i7 ( r )
z^u^^plz+oiif).
(17)
Multiplying equation (17) by Z* and subtracting from the resulting equation the complex-conjugate of equation (17) multiplied by Z we derive in the usual manner that 2/ -I- 1 d_ (18) L = 2Ul^q-±\Z\\ r dr-[Z,Z*) where
[Z,Z*]r =
Z
(19)
is the Wronskian of Z and Z* with respect to r. From equations (9) and (18) it follows that [Z,Z*]r = 2H,(2i 0 + 1)T%\Z\\ (20) Jo r where it will be noted t h a t the integral on the right-hand converges for r-> oo. Finally, we may note t h a t the complex solution for Z, belonging to cr and Z + iZ4 for |/J ^ I, can be obtained explicitly either as in Paper I I by separating the real and the imaginary parts of equation (17) or more directly from the analyticity of Z(a;l) in the complex Z-plane. Thus, seeking a solution, Z c = Z + iZ1I
(21)
of equation (6) for the complex Jo = f + Hi. (22) (where I is not necessarily an integer) and separating the real and the imaginary parts of the equation, we find (as in Paper I I I , §3) t h a t the real part, Z, of the solution (21) is the real solution of the same equation (6) for the real part of lc = I. Denoting by Z = Z(r;
[136]
(23)
the solution of equation (6) for assigned real values of <x and I, the imaginary part Z{ of the solution is given by Z 1 = / 1 Z,(r;or,J) + 0(ZJ).
(24)
With these definitions, the required complex solution belonging to (
zc (r; °"o> k+ui) = z(r; °\» k)+ilx 3ZZ(r;
(26)
The Wronskian of this solution Zc and its complex-conjugate is given by (as may readily be verified) tZc,Z*]r = 2«1[Zl,Z]r, (27) where Z ; and Z have the same meanings as in equation (26). Equation (20) now gives dr [ Z , , Z ] r = (2Z0 + l) I -5-[Z(r;o- 0 ,I 0 )]
(28)
3. T h e resonant axial m o d e : the growth of the flow of gravitational energy through the star The theory described in §2 is immediately applicable to the resonant axial modes of oscillation of a compact enough star. As shown in Paper I I I , they present a case of pure potential scattering by a central field governed by a wave equation of the same form (6). Thus (Paper I I I , eqs (l)-(3)): d2Z drl~
o-2
where
T)t
U[r)
=
-l(l+i)-U{r) Z=--0,
(29)
e-"+"s dr,
(30)
lo
-PY
GM(r)
(31)
and e and p denote the energy density and the pressure in the star; and resonant axial modes, derived from the Regge poles, Z0 + iZ1 (Z0 = 2, 3,...) can be determined exactly as in Paper I I I , §3, by integrating equation (29) for the square of the amplitude, a 2 + /?2, of the standing waves at infinity for a value of I (for a chosen cr0, determined already) and varying I in the neighbourhood of Z0 and ascertaining whether its (a 2 +/? 2 ) dependence on Z confirms the prediction (13). In this manner, the Regge poles for the homogeneous model, rJM = 2.3 a n d M — 0.496557, considered in Paper I I I (see table 2) were determined. The results of the calculations are summarized in table 1 and illustrated in figure 1 and 2.
[137]
Table 1. Resonant oscillations induced by axial gravitational waves (Homogeneous star: rJM = 2.3; M = 0.496557, r, = 1.142. The units in which the various quantities are expressed are the same as in Papers I and III.) 108(a2 + /?2)
I
fO.473524 8.982 {0.473525 8.945 [0.473526 8.936 a2+/p = A[(cr-o-<))> + o-l];A = 140
iov 2 +/? 2 ;
(1.99999 8.954 o-0 = 0.473525 -,'2.0 8.945 [2.00001 9.039 a2 + /?2 = B[(l-l0)2 + l*];B = 5.15 l0 = 2; I, = 7.32 x 70-"; o-, V(4/.B) = 7.30 x 10~
/0 = 2
a2 + /F
I
d' + i
[2.99999 4.464 x 10-9
f 0.65619 1.303x10"' \ 0.65620 2.427 x 10-" 10.65621 1.332x10"' (<x2+/P) =A[(
(a) TAe growth of the flow of gravitational energy through the star We turn next to equation (17) which in the present context takes the form, d2Zc drl
= ^(2/0+1) — Zc.
(32)
From this equation, we readily deduce (cf. equations (18) and (19)):
— and
[Zc,Zt]r^2ili(2l0+l)^\Zc
(33)
-r\Ze\*drm,
(34)
Zc = Z{U;
(35)
[Z c ,Z c *] r = 2 ^ ( 2 / 0 + 1)
where (see equation (26))
An equivalent form of equation (34) t h a t one obtains with the aid of equations (30) and (35) is (cf. equation (28)) e - " + * [ Z I > Z ] r = (2* 0 +l)
f e"+'"2
[Z(r;
(36)
where Z — Z(r+; (T0, Z0) and Z , = [Z ; (r + ; o-0, Z)];_( as in equation (35). Also, it should be noted t h a t the quantity on the right-hand side of equation (36) is a measure of the total energy in the form of gravitational radiation t h a t crosses the sphere of radius r. The vanishing of the Wronskian, [Z c ,Z*] r < , for Z4 = 0 signifies no more than t h a t there is no net flux of radiation associated with the real part of the solution for Z belonging to er0 and l0 and representing, as it does, standing waves. Conversely, the non-constancy of the same Wronskian wh< n Z4 (and, therefore, also ov) is different from zero implies, according to the right-hand side of the equation (34), a monotonic growth of the outward flow of radiation from zero at the centre to a maximum upper
[138]
9x10'
0.473517 Figure 1. The determination of the poles cr, and lx in the complex cr- and i-planes for the resonan waves by a homogeneous model of mass M = 0.496557 and r/M = 2.3. (a) The potential V(r) inc location of the quasi-stationary state, (fc) The dependence of a'+fi2 on cr for l„ = 2. (c) The depe For further details see table 1.
2.5x10 1.0x10
+ 2x10
t 0.65614
t 0.65620
t 0.65626
Figure 2. The same as figure 1 but for I = 3. (a) The potential V(r) including the centrifugal poten o- for la = 3. (c) The dependence of a2 + ^ on I for o"0 = 0.65620. For further de
2x10
1x10
1x10
r,= 1.142
= 1.142
Figure 3. The growth of the flow of gravitational energy through the star for the homogeneous model illustrated in figures 1 and 2: (a) for quadrupole (/ = 2) waves; and (b) for sextupole (I = 3) waves. Note that the energy flow continues to increase (albeit very little) beyond r = rl. (The function that is illustrated is: {2l0+ 1) J o 'e* + ^[Z(r;o- 0 ,/ 0 )]*dr/r 2 (cf. equation (36).)
limit for r*->-oo. This behaviour of [Zc,Z*]r^ is in conformity with the purely outgoing character of the radiation t h a t defines the quasi-normal mode (by condition (c) in the enumeration of the boundary conditions in Paper I I , §2)f. But an attempt to evaluate the growth in the flow of energy with the aid of the solution, I I , eq. (18), namely (cf. the analogous definition (26)) Zc = Z(r, ; o-0, Z0) + i
(37)
will fail, since the damped time dependence, exp(icr0 — (Ti)t,
(38)
of the quasi-normal mode implies an exponentially diverging radial dependence, (39)
exp(io- 0 + a i ) r # ,
a fact t h a t we have encountered (to our chagrin) in other contexts (e.g. Paper I I I , eq. (A 7)). I t is therefore important to draw attention to the fact that only by recourse to the Regge theory are we able to circumvent the obstacle of divergent radial integrals and evaluate explicitly the growth of the flow of gravitational radiation through the star and exhibit an essential physical aspect of the emission of gravitational radiation by a star in non-radial oscillations. The growth of the flow of gravitational radiation in the homogeneous model considered in table 1, for both the quadrupole and the sextupole modes, was determined with the aid of equation (36); and they are exhibited in figure 3. I t should be noted t h a t we did not find any detectable differences (beyond numerical f This interpretation of the non-constancy of the Wronskian is not different from the one that is given, in quantum theory, of the equation,
a
e
\z\*
(i) - \Z\2 - i — [Z, Z*]r = 21,(21 +1) — dt dr r2 which follows from the time-dependent Schrodinger equation that is the counter-part of equation (6). Equation (i) is equivalent to the equation given in Landau & Lifshitz (1977, bottom of page 588); and they write 'the right-hand side of this equation' is positive definite and signifies 'the emission of a new particle in the field volume'.
[141]
uncertainties) in the fluxes evaluated with the expression on the left-hand side and the volume integral on the right-hand side of equation (36) - a welcome check on both the theory and the numerical calculations.
4. The resonant polar mode: the growth of the flow of gravitational energy through the star The quasi-stationary states t h a t come into being when polar gravitational waves are incident on a star derives from the fluid motions t h a t they excite, in much the same way t h a t an excited compound nucleus is formed when high-energy protons or neutrons are incident on an atomic nucleus. This analogy between a star and a compound nucleus, in the present context, is not a far-fetched one. For, in considering the scattering of a high-energy neutron (for example) by a nucleus, one envisages t h a t the neutron, on entering the nucleus, rapidly distributes its excess energy to the other constituent nucleons with the result t h a t an excited intermediate compound nucleus is formed; and t h a t it is the quasi-stationary states of the compound nucleus t h a t are responsible for the resonant scattering (in accordance with the Breit-Wigner formula) t h a t one observes. In the same fashion, the incidence of polar gravitational waves, by inducing fluid motions in the star, creates an 'intermediate' state characterized by quasi-stationary states which are responsible for the resonant scattering and the quasi-normal modes of oscillation. Even apart from the analogy we have described, the resonant scattering of polar gravitational waves by a star is not, in any sense, a case of conventional potential scattering governed by a radial wave equation of the standard form (6) (or (29) as in the case of the scattering of axial waves considered in §3). Nevertheless, to describe the scattering process by a Regge pole in the complex Z-plane does not transgress any more rules t h a n describing it in terms of a pole in the complex cr-plane. A Regge pole, l0 + ilit t h a t corresponds to the pole craJricri can be found from the same equations, I (72)-(75), by determining the square of the asymptotic amplitude, a 2 +/? 2 , of the standing waves at infinity, as a function of real I in the neighbourhood of a chosen lQ and for a value of cr = cr0 t h a t has already been determined; and comparing it (i.e. a 2 +/? 2 ) with the predicted dependence (13). In this manner, the Regge pole was determined for the polytropic model considered in Paper I, §9. The results of the calculations are summarized in table 2 and illustrated in figure 4. The problem of evaluating the growth of the flow of gravitational radiation through a star in polar oscillations is of far greater complexity than the one considered in §3 in the context of axial oscillations and requires the resolution of issues outside the range of normal Regge theory. In the context of the axial oscillations, the non-constancy of the Wronskian [Z,Z*]r# of Z and Z* belonging to ((T0;l0 + ili) could be derived directly from the governing equation (32) and related to a positive-definite volume integral (34). In contrast, in the present context of polar oscillations, we are presented with a none too simple a set of four coupled equations (see equations (50)-(53) below) in which no mention of a potential occurs. We do not also have any simple analogue of the theorem t h a t assures the constancy of the Wronskian of any two linearly independent solutions of a second-order linear differential equation. Nevertheless, we shall show how these difficulties can be circumvented by appealing to the flux integral derived in a recent paper (Chandrasekhar & Ferrari 1990; this paper will be referred to hereafter as Paper IV).
[142]
Table 2. Resonant oscillations induced by polar gravitational waves (Polytropic model: n = 1.5, e0/p0 = 9; rl = 1.57641; M = 0.20541. The units in which the various quantities are expressed are the same as in Paper I, fig. 1.)
a
ioV2+/?2)
i
( 0.32484 1.027 \ 0.32485 1.025 (.0.32486 1.044 a' + fi, = A[{
iov 2 +/? 2 )
(1.9999 1.045 ~
I t was shown in P a r t I I of Paper IV (eqs (132)-(134)) t h a t the polar perturbations of a star allow a flux integral of the form, #2.2+^3,3=0.
(40)
F
Separating the contributions, Ef^ and E[ \ derived, respectively, from the perturbations of the metric coefficients and of the hydrodynamical variables, respectively, we have, in the notation in Paper I, Ei = E{1G)+El/\
(41)
where i?iG> = e-{[5 A t „8/t*] 2 + [ 8 ^ , 5 ^ * ] , - [ 8 i ; i l 8 ( ^ + / t , ) * - c . c . ] + [8 / * 2 5(^+ / * 3 )* 2 -c.c.]}, E
(42) (43) (44)
and 8 ^ , 8/*2, 8/*3, Sv, 8p, £2, and their complex-conjugates, 8y>*, b/i*, etc., are the solutions of the same equations I, eqs (48)-(51) (or, equivalently, I, eqs (72)-(75)) for real crs and integral Is. (It should be noted t h a t Ef) as defined in equation (43) is opposite in sign to the corresponding terms included in the expression for E2 in IV, eq. (132). This difference arises from the different conventions adopted in the two papers as noted in the footnote on p. 249 of Paper I.) We have not written out the term E3 since, in the present context, the flux integral after the separation of the variable is equivalent to (cf. IV, eq. (70))
where
| ; < £ 2 C > + ^ > > = 0,
(45)
<£ 2 > = (E{a) + E?> > = P r2(£
(46)
Jo
The required expression for <J£ 2 G) > can be written down directly by simply replacing —B23 by W in the expression on the right-hand side of IV, eq. (80) (as follows from a comparison of the definitions of T in IV, eq. (76) and I, eq. (29)) and ignoring the term [B23,B*3]2 derived from the Maxwell field. We thus have
( C)> =
^
27TTr2 e"""2fe[X'X*]*+[F' F*]*+{N- *F* ~N**F) ~ (L'F*2 ~L*Fa))' (47)
[143]
1.8x10
2.0x10 03.
+
+
1.2x10
1.2x10
0.324780
0.324840
-
0.324900
Figure 4. The determination of the poles
where
F = L+X+W
(48
and a misprint - (2K/TO2) in place of (K/2W) - in the expression on the right-hand side of IV, eq. (80) has been corrected. The expression for {E'p} can, similarly, b< reduced to the form
<EiF)> =
r2e
2^
^{[{e+p){2T~KV~L)*~n*]'^~c-c}
4 (FU*-F*U) + \(LU*-L*V) 2 21+1 r e*-*»
+n l \ _^{IIU*-II*U) • (49) 2(e+p)'
In the further reduction of <7i^G)> and (E2F)) we shall find it convenient to rewrite I, eqs (50), (48), (49), and (32), respectively, in the forms ZV r - (1 + rv r) F r + ( - e2"* - <x2r e 2 *"^) F »+l
- e2"* (N+ X) - ( - + 2 v_ T J L - r e2"' 77 = 0, (50)
Fr + (±-v
+ jL-\U = 0,
N
.r-(^-v.r)N-Fr+Xtf^
(r2e."~^Xr)
(51)
+ v^L = 0, + (T2r2^-"X^
0,
(52) (53)
by making use of the following definitions (I, eqs (37), (45), and (47)):
X = nV,
- 7 7 = (e+p)N+\a-2e-2"W!
(54)
U=WtT + (Q-l)viTW, W= -(T+L-W),
(55)
n = \{l-i){l + 2)
(56)
and K = 1(1+1) = 2(n+l).
We shall find the following lemmas, which follow directly from equations (50)-(53), useful: [F,F]t = --{LF* -L*F) + \{UF* -U*F), r (l-rv,r)[F,F*]t + (L*Fr-LF%) = %r(U*F
r)[F,F*]2
= (n+l) — r
(57) (58)
[(N+X)F*-(N+X)*F]
+ (^+2vir)(LF*-L*F)
+ re2"*(HF*-n*F).
(59)
Returning to equations (47) and (49), we find that (E[G) + EiF)} = <7£2>, after some
[145]
considerable simplifications resulting from the use of the foregoing lemmas, can be reduced to the form,
+ &vir(UF*-U*F)
+ re2?'(nF*-n*F)+—l—{nU*-n*U)\.
(60)
2{e + p)
J
The Wronskian [X,X*]r t h a t appears on the right-hand side of this equation can be replaced by a volume integral with the aid of equation (53). Thus multiplying equation (53) by X* and subtracting from the resulting equation the complexconjugate of equation (53) multiplied by X, we obtain 4-{r2e"-^[X,X*]2} dr
= -ne"+/i>[(N+L)X*-{N+L)*F];
(61)
or, after integration, -^-r2e"-">[X,X*}T
= -{n+\)
\ e"+*'[(N+L)X*-(N+L)*X]dr.
(62)
Inserting this expression on the right-hand side of equation (60), we obtain, \(2l+l)(E2)
= -(n+l)
\re"+*>[(N+L)X*-(N+L)*X]dr
+ e*+**{(n+l)r[(N+X)F*-(N+X)*F] + r2e"-^hrvr{UF*-U*F)
+—
+
r!>(nF*-n*F)}
(77f7*-77*C/)|.
(63)
Before we consider how we may use equation (63) to evaluate the growth of the flow of energy in the form of gravitational radiation through the star, we recall first t h a t the equation is applicable, as it stands, to complex-conjugate solutions of the same equations (50)-(53) (or equivalently of I, eqs -(48)-(51) or (72)-(75)) when both n(=\(l—l)(l + 2)) and a are real; and (E2y (see equation (45)) has the meaning of a conserved (constant) flux of radiation t h a t obtains under the circumstances. But the boundary conditions of the problem (enumerated in Paper II, §2) do not allow physically meaningful complex-conjugate solutions for real crs and Is. And when the problem is solved, by the algorism described earlier in this section for real values of n and er0, the derived solutions for N, X, etc., represent standing waves; and in conformity with this fact <2?2> = 0. On the other hand, on the assumed analyticity of the perturbations in the metric functions and in the hydrodynamic variables, we can consider (E2~) as analytically extended into the complex Z-plane by substituting for N, X, etc., the solutions Nc = N(r; cr0,10) + ilt[Nt ,(r;
(W)
Xc = X(r • a0, l0) ± i/it-Y j(r;
[146]
We have thus evaluated £2> for the polytropic model considered in table 2. (It may be noted, parenthetically, t h a t in the evaluation the following relations were found useful: for any two functions, Fir-v^ FG*-GF*
+ il&F^r-a^l)}^ = iil^iG-Q^F)
and
G(r;
= 2U&F, G\.)
J
The results of the evaluation is exhibited in figure 4 c. I t will be observed t h a t
(r&rj.
(66)
This constancy of (E2y for r > rx follows from the relation, i(2Z+l)i- =
{n+l){-[(N+L)X*-(N+L)*X]
+ ^r[(N+X)F*-(N+X)*F]\
= 0,
(67)
t h a t is valid for r > rx when the fluid variables U and 77 vanish and v+/i2 = 0. (It is not difficult to verify equation (67) with the aid of equations (50)-(53) specialized for the vacuum.) The constancy of (,E2} for r > rx is to be contrasted with the continued growth of [Z, Z*],. ( - albeit a very little - beyond the boundary of the star at r = rx (see figure 3). This difference between the polar and the axial modes must be traced to the fact t h a t while the resonant scattering of axial waves results from the character of the potential (illustrated in figures 1 a and 2 a), the resonant scattering of the polar waves results solely from the induced fluid motions and must in consequence cease to be operative beyond r = rx. 5. Concluding remarks One should not, perhaps be surprised t h a t the methods t h a t had been developed for the treatment of potential scattering in the context of the Breit-Wigner formula and the Regge poles are the natural methods for the treatment of the non-radial oscillations of stars in general relativity. But to obtain a correct perspective, it is important to emphasize t h a t the arenas of the quantum, theory and of classical general relativity are different; and the transfer of methods appropriate for one terrain to another requires a re-orientation of objectives and some essential generalizations as well. We have referred to some of them already in §2. In order not to interrupt the logical development of the subject, we passed over several major aspects of the differences in the two approaches. We return to them in this concluding section. A major cause for reorientation derives from the fact t h a t even in the simplest cases when a radial Schrodinger equation is found to govern the problem (as in the case of the axial modes considered in §3), the equation is not as simple or as direct a consequence as the radial wave equation t h a t governs the scattering by a spherically symmetric potential. In the latter case, the kinetic energy E replaces u 2 in equation (29) where it will be recalled a2 results from the assumed common timedependent factor e1
[147]
of complex-conjugate poles in the physical scattering amplitude analytically extended across a square-root type cut along the real axis in the complex ^-plane, a cut necessitated by the required analyticity of the 5-matrix. Of the two poles, one is near the 'physical' region while the other is far away. On this account, only the nearer one contributes significantly to the physical scattering amplitude. In contrast, in the relativistic context, the boundary conditions of the problem do not restrict consideration to purely outgoing or purely ingoing waves in conformity with standing waves as the starting point of our consideration. Consequently, at resonance, both the poles at + a and — a contribute significantly. A further basic difference in the strategy of the two approaches is exemplified in the replacement of the scattering amplitude (basic to the notion of the scattering matrix central to the quantum theory) by the asymptotic amplitude, A(
[148]
a p r o b l e m in classical g e n e r a l r e l a t i v i t y s h o u l d r e q u i r e n o t o n l y t h e i n c o r p o r a t i o n b u t a l s o g e n e r a l i z a t i o n s of t h e m e t h o d s d e v e l o p e d in t h e f r a m e w o r k of t h e quantum t h e o r y , d e c a d e s a f t e r t h e f o u n d i n g of t h e t h e o r y of r e l a t i v i t y . We are grateful to Roland Winston for stimulating conversations in which the possibility of applying the Regge theory to the problem on hand first occurred; and to Peter Freund for extended discussions in which he clarified for us various aspects of the Regge theory as formulated in quantum theory. The research reported in this paper has, in part, been supported by grants from the National Science Foundation under Grant PHY-89-18388 with the University of Chicago. We are also grateful for a grant from the Division of Physical Sciences of the University of Chicago which has enabled our continued collaboration by making possible periodic visits by Valeria Ferrari to the University of Chicago.
References Alfaro, V. de & Regge, T. 1963 Chandrasekhar, S. & Ferrari, V. Chandrasekhar, S. & Ferrari, V. Chandrasekhar, S. & Ferrari, V. Chandrasekhar, S. & Ferrari, V. Landau, L. D. & Lifshitz, E. M.
Potential scattering. Amsterdam: North Holland Press. 1990 Proc. R. Soc. Lond. A428, 325-349. (Paper IV.) 1991a Proc. R. Soc. Lond. A432, 247-279. (Paper I.) 19916 Proc. R. Soc. Lond. A434, 635-641. (Paper II.) 1991c Proc. R. Soc. Lond. A434, 449-457. (Paper III.) 1975 The classical theory of fields. New York: Pergamon Press.
Received 27 September 1991; accepted 21 November 1991
[149]
On the non-radial oscillations of a star. V. A fully relativistic treatment of a Newtonian star B Y SUBRAHMANYAN C H A N D R A S E K H A R 1 AND VALERIA F E R R A R I 2 1
University of Chicago, Chicago, IL 60637, USA JCRA (International Centre for Relativistic Astrophysics), Dipartimento di Fisica 'G. Marconi', Universitd di Roma, Rome, Italy
2
It is shown how the non-radial oscillations of a Newtonian star (i.e. a star built in the Newtonian framework) can be treated fully relativistically as the result of the scattering of gravitational radiation by its own, shallow, curved spacetime. It has the distinguishing characteristic that no gravitational radiation emerges.
1. Introduction In a series of papers published during the years 1991-93, we have developed a new approach to the theory of the non-radial oscillations of a star on the premise that the oscillations can be induced by the scattering of incident gravitational waves by the spacetime of the static star. (For a review of these papers, see Ferrari (1993).) The method as described is applicable to all stars independently of their sizes relative to their Schwarzschild radii. Indeed, Ferrari & Germano (1994) have shown that the method applied to a polytropic star of index n = 3 and 2.362 x 105 Schwarzschild radius predicts with great precision the frequencies of all the first 21 normal modes of oscillation, belonging to I — 2, in agreement with the Newtonian calculations. Nevertheless, the question remains whether the strict Newtonian limit arises in an exactly specified context of the fully relativistic theory as developed in Chandrasekhar & Ferrari (1991; referred to hereafter as Paper I) with suitable boundary conditions. That such a limit should exist appears reasonable since the curvature of the spacetime remains finite, though small and no matter how dispersed the matter may be: the curvature has only to be measured in the unit {G/c2) cm g - 1 (= 1.476 km/M 0 ); and the scattering of radiation must certainly be possible even by such shallow gravitational fields. However, as we shall show that by the scattering of gravitational waves, closed Newtonian systems cannot give rise to any emergent gravitational radiation even as the dipole oscillations of a star (cf. Paper I, §9). This result should not be confused with the finite radiation damping of the oscillations of Newtonian stars in a higher approximation. These conceptual matters are more profitably discussed in § 8 after the analystical developments in the following sections are completed. 2. The origins of the non-radiating Newtonian limit in the fully relativistic theory The static spacetime of a star built on the Newtonian theory arises from the limiting forms of the exact equations of equilibrium of general relativity. These equations
[463]
are (cf. Paper I, eqns (3)-(6)): 3-2M*
=
i _
2
GM(r) cr
,r
(1)
r
p,T _ G M{r)/r2 + M / c 2 e + p~ c2 l - 2 M ( r ) / r c 2 '
~
(2)
M(r) = — / er2 dr ~ / pr 2 dr,
(3) 2
where in equation (3) we have replaced the energy-density e by pc as appropriate in the context. (Note also the absence of the factor 4ir in our present definition.) Prom these equations it follows that for a Newtonian star, i.e. a star built on the Newtonian equations of equilibrium, G M(r) c2
r2
G K
=
M{r) a n d
C2
/=
in
e
d
r
5 = 7 -p-
and finally
^
r2
d
f
(6)
i
r
M(r)"
d M(r) dr r
(5)
;
M(r)
M2,r — "• ,
or
where
r2
dr M2,j
(4)
^JVI
= Kg
(say),
,
-2«r/. -1 The following elementary identities among / and 3 may be noted: Q2M2
f + g = rp
and
/ - g = -r/,r.
(7) (8) (9) (10)
At this point, it is important to note that in the Newtonian theory, the frequency a of the normal modes of oscillation of a star are expressed in the unit, GM R3
-|V2
(11)
which, in the units, G = c = 1, adopted in Paper I, a is measured in the unit c. Hence, in transcribing the equations of Paper I, in forms appropriate for the Newtonian limit, we may write 2 2 M (12) (7 = KO where
a- 2
M._ 'R? ~
1
S/T2
(13)
and p denotes the mean density.
[464]
3. T h e equations governing the non-radial oscillations in the Newtonian limit We now turn to the equations describing the scattering of gravitational radiation by a star in the fully relativistic theory; and we consider them, as written in Paper I equations (72)-(75): X,r,r +(l+
",r ~ M2,r) X,r + £&*<» (N + L) + a2f?ll»-»)X
-L,r = {N + 2X),r
+
+[--VtT (Q + \)v,r
= 0,
{-N + 3L + 2X) N-L+-G
+ -{rX,r + X)
(r 2 £), r = nvAN -L) + -(e2^ - 1)(N + L) + r{v,T - /i 2 , r )X, r + a2e2^-v)rX, -(v,rN,r)
(14)
(15) (16)
= -G + v,r[X.r + "AN ~ L)\ + T ( e 2 w - l)(N - rX
- e 2 «(e +p)N + iaV*""-"* IN + L + — Q + -\rXT [ n n where, it may recalled, that (cf. I, eqns (3) and (17))
+ (2n + l)X]
(17)
P, = _ (18) - - — = -q (say), IP e + p 7p a quantity that is well defined in flat space. A remarkable property of equations (14)-(17), which we failed to notice before and on which our subsequent analysis depends, is that they include the pair of equations (14) and (15) which survive in the Minkowskian limit when vT and fi2>r vanish and v and fj. can be set equal to zero; and the pair (16) and (17) which vanish identically in the same limit. Thus, in the Minkowskian limit, equations (14) and (15) reduce to Qv,r
and
X,rr+-Xir+^(N r r2-
{L + N + 2X),T + J(3L - N + 2X) + (* + q
(19)
+ L)=0
N-L+-(rX,r
+ X)
0. (20)
These equations are in fact included in the equations which govern the propagation of free gravitational waves in Minkowski space (cf. Chandresekhar & Ferrari 1993, eqn (5)). In contrast, equations (16) and (17) vanish identically in the same limit since all the terms in these equations have as factors i/ r , fj.2,r> e2M2 — 1> or a1 (= KO2 by equation (12)); in other words, equations (16) and (17) have an overall factor «. Therefore, after the removal of this factor, they survive in finite forms. However, for our purposes it will suffice to consider the identity (A 4) established in Chandresekhar & Ferrari (1993), namely, [r(II - III)],r - nil - rIV - crVl = 0,
[465]
(21)
with the definitions, I = i/ r (L + X) + i(Q + l)!/ r W, II = -u,r(N - L), n + 1, III ( e ^ - l ) - ( I / + M 2 ), r N + u,r[r(L + X + 2
2
2( v)
+ -(e »* - 1)[X - n(L + W)\ + a r[e ^' IV = - ( 1 / -
W
W),r+2L]
(22)
- 1](L + X + ±W),
) r X, r - ^ ( e 2 ^ - 1)(7V + L) - <7 2 [e 2( «-"' - 1]X.
It is manifest that the terms in the identity have K as a factor; and after the removal of this factor the identity will survive in the Minkowskian limit. We shall obtain the explicit form of the simplified identity in § 4. Meantime, it is useful to adjoin I, eqns (58) and (69) in the Minkowskian limit, to the equations we have already derived, namely, W = N-L+-(rXr + X), (23) n ' Wr = (N-L),r--(L
+ N), (24) r where it will be recalled that W (= L — T — X/n) vanishes identically in the vacuum (cf. Chandrasekhar 1992, p. 147, eqn (43)). 4. The reduction of equations (19), (20) and (21) We consider first the reduction of equations (19) and (20). These equations can be rewritten in the forms:
±(N + L) = - (x,rr + lX,r) = -H(rX,r + X)
(25)
and (L + N)
r
+ -{L + N) + -(L-N) + 2(xr + -X r r \ r
+
N-L+-(rX,r + X) + qW = 0, (26)
where in the last term in equation (26) we have replaced by W (in accordance with equation (23)) the terms which occur as the factor of (q + 2/r) in equation (20). Simplifying equation (26), we obtain [r(L + N)],r + 2 ^ i ( r X n
r
+ X) + rqW = 0.
(27)
Now writing, F = rXtT + X, when equation (25) takes the form,
(28)
F,r = --(N + L), r equation (27) can be transformed to give
(29)
(r 2 F r ) i 7 . - 2(n + 1)F - nqrW = 0.
(30)
[466]
Turning next to the identity (21), we find, making use of equation (18), \ = -\qW + 0(K2).
(31)
The terms II, III and IV, apart from the factor K, and, ignoring terms of higher order, are manifestly, II = -f(N - L); III = f{r(L + X + W),r + 2L\ + [2(n + 1 ) / - (/ + g)]N +2f[X - n(L + W)] = f[r(L + X),r + r(N - L\T - (N + L) + 2L] +[2(n + 1 ) / - (/ + g)N] + 2f[X -nN{rX,r + X)} = -gN + fL + rf(N,r-X>r); TV = -{f-g)X,r-2-f(N
r Making use of these equations, we find:
(32)
+ L).
(33) (34)
II - III = r/,rAT - rfN,r + rfX,r, nil - rIV = nf(3N + L) - r2fX>r. The identity (21) now provides [r2(f,rN - fN
(35)
or, on simplification, making use of equation (25), we find [r2(f,rN - fN,r)],r + 2nfN + \c2r2qW
= 0.
(36)
Equations (30) and (36) provide a pair of coupled second-order differential equations for F and N, since W is expressible in terms of them: W = 2N - (N + L) + -F = 2N + -(F + rFr), n n
(37)
or, alternatively, W =
2N+-(rF)r. n
(38)
5. The boundary conditions for t h e non-radiating Newtonian normal modes of oscillation We shall now show that solutions of equations (30) and (36), describing nonradiating Newtonian normal modes of oscillation, can be found which satisfy the necessary conditions at the boundary, r = r\, of the star. The necessary conditions can be inferred from the example of dipole oscillations considered in § 9 of Paper I. Dipole oscillations cannot, of course, be accompanied by the emission of gravitational radiation; and this requirement is assured by the relevant equations I, (106)-(108) allowing solutions that satisfy the following conditions: (1) They allow, for each real a, two linearly independent singularity free solutions at the centre.
[467]
(2) Along each of the two solutions W = 0 at the boundary r = r\.
(39)
(3) There exist discrete real values of a for which Wj. vanishes, simultaneously, at the boundary, r = r\, of both the two linearly independent solutions.
(40)
These are the characteristic values of a that belong to the normal modes of oscillation that do not radiate. The requirement that both W and W,r vanish at the boundary is to assure that W vanishes indentically beyond the boundary, consistently with the prevalence of the vacuum outside the star. Besides, we must also require that N and L also vanish at the boundary for no space-time perturbations to extend into the vacuum outside the star. And this we can accomplish by finding the superposition of the two solutions belonging to a particular characteristic value a for which N (and therefore also L by I, eqns (124)-(127) vanishes. How all these conditions are satisfied is illustrated in table 2 and eqns (128)-(130) of Paper I. We shall show in § 7 how similar conditions appropriate for the present somewhat different circumstances, can be satisfied in the context of equations (30) and (36). 6. The existence of two linearly independent singularity-free solutions (o) Alternative forms of the equations and an identity It is first convenient to replace -2
F
by
G=^-F 2n
(41)
in equation (30), when it becomes {r2G,T),T-l{l
+ \)G-\a2rqW
= 0,
(42)
where we have substituted 1(1 +1) for 2(n +1). Similarly, by subsituting (I — l)(l + 2) for 2n in equation (36), we obtain [r2(f,rN - fN,r)],r + (I - W + 2)fN + \d2r2qW
= 0.
(43) 2
An important identity follows from equations (42) and (43) by eliminating d qW. We find [r2(/,r7V - /JV P )]. r + (I - 1)(Z + 2)fN + r[(r 2 G, r ),r - l{l + 1)G] = 0.
(44)
We obtain a form of equation (43) which removes some of its mystery by expanding the equation. A straightforward reduction yields ^.«- - -T-Mfsr + 2rf,r)N + - NT - ^ N - a2\w = 0. rzf r r* j The second term in / can be simplified by the following sequence of steps: r 2 /, r r + 2r/, r = iLr 2 /,r;
(45)
(46)
[468]
f=m,
dr
= p-2-f, r
l r
r'
r2f,r =
r2P-2rf;
(47)
{r2p - 2rf) = r2p,r + 2rp - 2 / - 2r/, r r*p,r + 2rp-2f-2rlp-
-f (48)
; 2
r p,r + 2/.
On inserting the last result in equation (45) and substituing for W the expression (cf. equation (38)) (49)
W = 2 N+—(rG),r we find JV rr + -JV,r -
N-a'
f
-JN = 0,
N+^(rG),
(50)
remarkably similar to equation (42): r—G —
G, rr H—G,r
= 0.
(51)
(b) The series expansions We now assume that at the centre, the series expansions obtain: N = N0rx + N2rx+2 + N4rx+4 + ... , G = G0rx + G2rx+2 + G4rx+i + ... ,
(52)
where x is the indicial exponent that is to be determined. The required expansion for / follows from the assumed behaviour of the density distribution at the centre: p = p0 + p2r2 + p4r4 + ...
(53)
Thus / = 5Por + ip 2 r 3 + ip 4 r 5 + . . . . /.r = \Po + zPir2 + f p 4 r 4 + •.. . And finally, with the expansion, P = Po + P2T2 + p4rA + ...
(54)
(55)
for the pressure, we derive q=H
= H r + 0{r (56) IP 7Po Together with the expression (53) and (54), we find for the remaining coefficients in equation (50):
i /
[469]
= + -^_ + 0(r»),
7PoPo
£F«+5£ + 0 (r»). /
Po
(57)
(c) The indicial exponent and the expansion coefficients Inserting the expansions (52) and (54) in equation (44) and retaining all terms inclusive of order r I + 3 we find: [r2(lrN - fN,r)],r + (I - 1)(J + 2) f N + r[(r2G,r),r - 1(1 + 1)G] = -±p0N0(x - l)(x + 2)r x + 1 + (x + 4)[±p2(3 - x)N0 - \Po(x + l)N2}rx + lPoN0(l -l)(l
+ 2)r* +1 + (I - l)(l + 2)(\p0N2 +
\p2NQ)rx+3
G0[x{x + 1) - 1(1 + l)]r I + 1 + G2[(x + 2)(x + 3) - 1(1 + l)]r I + 3 + ... . (58) We observe that the terms in No and Go of order rx+1 vanish identically if x = l.
(59) x+3
It is therefore a double indicial root. The remaining terms of order r (l + 4)[\p2(3-l)N0-\p0(l
provide
+ l)N2]
+ (l- 1)(J + 2)(ip0AT2 + ±p2JV0) + G2[(l + 2)(l + 3) - 1(1 + 1)] = 0. On collecting the terms and simplifying, we find
(60)
ip2AT0[(Z + 4)(3 - 1) + (I - l)(l + 2)] - ±p0N2[(l + 4)(Z + 1) - (I - l)(f + 2)] + 2G2(2Z + 3 ) = 0 ,
(61)
or
2p 2 7V 0 -|p 0 Ar 2 (2/ + 3) + 2G2(2/ + 3 ) = 0 ,
(62)
i.e.
N2 =p-G,+ ,J*+ 3)p0 No(21 0
(63)
Now returning to equations (50) and (51) and expanding them inclusive of order rl+2, we obtain (21 + 3)G2 - ^-{a2N0 7Po
+ (l + 1)G0] = 0, (64)
(21 + 3)N2 - -^-[a2N0 + (l + 1)G0] - ^-NQ IPoPo Po i+4 Similarly, by retaining terms of order r , we obtain 4(2Z + 5)G4 = —[a2N2 7Po
+ (I + 3)G2] +
4p4 Po
2pf Pl\
= 0.
[a2N0 + (l + l)G0],
4(2/ + 5)7V4 = 6P2 r_2 d2N2 + (l + 3)G2] IPoPo 2p| 6p2p2 2 +-7PoPo 4p4 Pi ' 5po [a N0 + (l + 1)G0] ^iV2 + LPo
+
2
6_P4
Po
9P2
\N
(65) It is manifest that two linearly independent solutions free of singularity at the origin follow from setting, for example, either G0 = 1 and N0 = 2, or G0 = 2 and N0 = 1.
(66)
[470]
7. The boundary conditions and the proof that the characteristic frequencies of oscillation belong to non-radiating solutions It has been shown in Paper I (eqns (84), (86) and (87)), that quite generally, at the boundary, r = rx, of the star, p, 6p, and W vanish together, linearly with the distance, r\ — r, from the boundary. In particular W ~ const, x (ri — r)
and
p ~ const, x (ri — r)
as
{ri — r) —> 0.
(67)
In consequence, the vanishing of W 'compensates' for the singualrity of q (as defined in equation (18)) resulting from the behaviour of p near the boundary, namely, const. , > q~ as r—* ri, (68) ri — r and, therefore, qW —> const, (r —> rx), (69) a fact which is crucial for the integrations of equations (42) and (43) (or, equivalently, equations (50) and (51)). It remains to formulate the boundary conditions which will endure the prevalence of the vacuum outside the star and the absence of any emerging gravitational radiation. As we have stated in § 5, the necessary and sufficient condition for the prevalence of the vacuum outside the star is that, in accordance with equation (38), W
2N+~{rF)r n
(70)
-{rF). (71) n both vanish at the boundary of the star to ensure the identical vanishing of W outside the star. We shall now prove the following theorem. W.=
2Nr +
Theorem 7.1. The characteristic frequencies, a, of the periods of oscillation are determined by the requirement, that for determinate superpositions of the two linearly independent singularity-free solutions belonging to a, N and W
(72)
With the same constant C(a), evaluate the corresponding superposition, [jyr(
so
= {WM*ir1)
+ C{5)(W,r)u(&-,r1).
(73)
The function [W ,r(5';7*i)]s constructed will not in general vanish. We seek a value (or values) a for which it does. Values of a so determined and the solutions belonging to them will satisfy the conditions of the theorem, i.e. at their boundarie, W
[471]
We shall now show that no gravitational radiation emerges from the solutions determined by the algorism described. From equations (70) and (71), it follows that the vanishing of W, WtT, N, and NT at the boundary implies that (rF), r = (rF),„ = 0;
(74)
and, therefore, (rF),r=0 Integrating this last equation, we find,
for
F=—
r > rr.
(r > ri),
(75)
(76)
where D is a constant with no constant of integration for the convergence of F at infinity. From equation (29), we conclude (since N = 0), -~L
= F,r = --2;
(77)
and, accordingly, ,
D ^ (78) K nr ' The Coulombic behaviours of F and L for r ^ 7^ confirm the non-radiative character of the solutions determined. L =
8. A direct numerical confirmation of the algorism described in § 7 We have sought a numerical confirmation of the algorism described in § 7 for the determination of the normal modes of oscillation of a Newtonian star by considering a polytrope of index 3 and 7 = | . Quite generally, for a polytrope of index m, the density and the pressure given by p = Po&m
and
p = p0em+1
(79)
(where p 0 and p0 denote the central density and pressure) are determined by the Lane-Emden function 9 governed by the equation, 1 d A2d0\ satisfying the conditions, and
— = 0 at
£ = 0,
(81)
^-"-i^r
<*>
where
With p, p, and r measured in the units p0, po, and K, the mass is given by d6
M(t) = -?%d?
(83)
[472]
Table 1. The values N and Vv,r at the boundary £1 (= 6.8968486) of the two linearly independent solutions, I and II, for three selected values of a are listed N
Wl(
a
I
II
I
II
2.4514
-628.38708470 -628.37437592 -628.36166689
-294.16185033
-3.3857943988 -3.3836669258
--1.5849944298 --1.5839625104
-3.3815347040
--1.5829320157
2.4515 2.4516
-294.15533461 -294.14881877
Table 2. The determination of a o
W,z
C{a)
2.4514 2.4515 2.4516
0.6283D-04 -0.7557D-05 -0.7017D-04
-2.1361950369 -2.1361991505 -2.1362032644
while /(0 = - ^ |
and
g(0 = m)-tem-
(84)
In the same units, the radius and the mass of the star are fi = 6.8968486
and
M(£) = 2.01824.
(85)
Equations (50) and (51) were integrated for a polytrope of index 3 and 7 = 5, for different values of the frequency a, starting from £ = 10~2 with the series expansion provided in §66 and the initial values (66). In table 1 the values N and Wt^ at the boundary ^ , of the two linearly independent solutions, I and II, for selected three 'consecutive' values of a are listed. In table 2 the solutions iVj and Nu are superposed so that Ni(Z1\
(86)
The derived constants C(a) are listed in the last column. With the same constants C(a), (W,€)(&;
a = 2.4515,
(88)
which therefore belongs to a normal mode. The characteristic values, <7, for other normal modes, determined in the same manner are listed in bold face in table 3; and compared, in the parallel column with
[473]
Table 3. A comparison of the characteristic frequencies determined by relativistic algorism of this paper with the values determined directly by a standard Newtonian method
order
by the relativistic algorism
a by a standard Newtonian method
2.4515 3.4751 4.5395 5.6411 6.7610 7.8889 9.0197 10.1515
2.424894 3.465915 4.537138 5.640526 6.760854 7.888755 9.019690 10.151484
values determined directly by a standard Newtonian algorism.f It will be observed that the agreement between the two sets of values are entirely within the limits of accuracy of our calculation of the theory described in this paper. 9. Concluding remarks In this last section we wish to emphasize that our principal interest in this series of papers on the non-radial oscillations of stars has been to lay bare the new insights that one gains by exploring new avenues of approach towards even well-worn problems in general relativity. Thus, taking the problem of non-radial oscillations of fluid spheres as a prototype, we have shown how viewing them as the result of the scattering of incident gravitational waves with the aid of the Breit-Wigner formula, several features of the phenomenon emerged that had not even been envisaged earlier: the existence of resonant axial modes of oscillation; their coupling with the polar modes by slow rotation and the associated Lense-Thirring effect; the relevance of the Regge theory of scattering in the domain of the complex angular momentum for determining the growth of the flux of gravitational radiation through the body of the star; and the role of the flux integral that perturbations of static and stationary spacetimes allows. Yet, in our view, one basic conceptual problem remained unresolved: Should one not be able to consider the oscillations of a Newtonian star - i.e. a star built on Newtonian laws - as a result of the scattering of gravitational waves by its curved spacetime albeit its shallowness. The curvature of spacetime in the presence of matter can never be abolished: 'In honoured poverty, thy voice [must] weave songs' (Shelley). We explored this question for two or more years, but in vain. The clue to the final resolution came from casually noticing a remarkable feature of the four basic equations that we had derived already in our first paper (eqns (72)-(75) in Paper I) - a feature that stares in the face, once noticed! The four equations split into two pairs: a pair that survives in the Minkowskian limit when all the terms depending f We are indebted to Dr J0rgen Christensen-Dalsgaard for providing us with the values of 5 determined by a Newtonian algorism.
[474]
on the curvature of the spacetime are ignored - equations which in essence describe the propagation of free gravitational waves in Minkowski space; and the other pair which vanishes identically as each of the terms in this pair is directly dependent on the curvature expressed by z/r, fj,2,r, or (e2M2 — 1). If one removes the common proportionality factor G/c2 of these terms, the equations remain finite after ignoring terms that are of second and higher orders in the curvature. In this manner, we obtain a pair of coupled second-order linear differential equations in two variables which combine in a common scheme both the Minkowskian and the Newtonian limits. The equations derived in the manner described, determine the exact Newtonian characteristic frequencies of non-radial oscillations by allowing two linearly independent singularity-free solutions at the centre and satisfying the boundary conditions that will ensure that no gravitational radiation emerges (even as they do not for dipole oscillations of fully relativistic stars). In other words, the distinguishing characteristic of gravitational waves by Newtonian stars is that no gravitational radiation emerges. Several questions occur in contexts larger than in the example considered: Should one expect, quite generally, that equations describing the scattering of gravitational waves by the spacetimes of closed Newtonian systems will combine in the same manner in a common scheme both of the Minkowskian and the Newtonian limits? Can one expect that no gravitational radiation will emerge from such a Newtonian system as a theorem of general validity? We are grateful for a grant from the Division of the Physical Sciences of the University of Chicago.
References Chandrasekhar, S. 1992 The mathematical theory of black holes. Oxford: Clarendon Press. Chandrasekhar, S. & Ferrari, V. 1991 Proc. R. Soc. Lond. A 432, 247-249. (Referred to as I.) Chandrasekhar, S. & Ferrari, V. 1991 Proc. R. Soc. Lond. A 4 4 3 , 445-449. Ferrari, V. 1992 Phil. Trans. R. Soc. Lond. A 335, 340. Ferrari, V. & Germano, M. 1992 Proc. R. Soc. Lond. A 444, 389-398. Received 3 April 1995; accepted 16 June 1995
[475]
1089
X.
Miscellaneous Writings
In this final part of this volume, I have collected together articles semipopular in nature. They complement Chandra's scientific and technical writings, and illuminate his wide-ranging interests pertaining to the history and philosophy of science. In his later years, Chandra was seriously interested in questions concerning differences in the patterns of creativity between practitioners in the arts and practitioners in the sciences. He also became interested in understanding aesthetic motivations and the perception of beauty in the pursuit of scientific knowledge. Drawing on a varied and rich source of literary and scientific writings, Chandra presents his views in articles 7, 11 and 12. Articles on the theory of general relativity (8, 9, 10, 14 and 15) go beyond the usual popular expositions. They provide rare historical insights and poetic interludes, for instance the fable of the dragonfly at the end of the Halley Lecture (article 15). In article 6, Chandra traces briefly Fermi's fundamental contributions to physics, and, more importantly, presents a discussion of some historical significance concerning the process of discovery in physics. Fermi describes how the idea of using a block of paraffin instead of lead came to him in his crucial experiments on the neutron-induced radioactivity. The idea occurred to him, he says, "without advance warning, no conscious, prior, reasoning." But, according to Maurice Goldhaber, Fermi, during his visit to the Cavendish Laboratory, had been shown neutron experiments with the use of paraffin to slow the neutrons. The idea at the time was probably implanted and pushed away somewhere at the back of his mind. There is also, at the end of this brief article, a moving account of Fermi's reaction on learning that he had terminal cancer and had about six months to live. Chandra had a lifelong interest in the life and works of the Indian mathematical prodigy Srinivasa Ramanujan. Articles 18 and 19 are devoted to his recollections of Ramanujan. Articles 20 and 21 tell us about Chandra's last undertaking, namely the writing of Newton's Principia. The late Lyman Spitzer once said, "It's a rewarding aesthetic experience to listen to Chandra's lectures and study the development of theoretical structures at his hands. The pleasure I get is the same as I get when I go to an art gallery and admire paintings." Reading these articles, some of them more than once, I have had similar feelings.
1091
Miscellaneous Writings The Scientist The Works of the Mind (The University of Chicago Press, 1947), pp. 159-79
1093
The Case for Astronomy Proceedings of the American Philosophical Society 108, no. 1 (1964): 1-6
1114
Astronomy in Science and in Human Culture Jawaharlal Nehru Memorial Lecture, 1968
1120
Ellipsoidal Figures of Equilibrium — An Historical Account Communications on Pure and Applied Mathematics XX, no. 2 (1967): 251-65
1138
A Chapter in the Astrophysicist's View of the Universe The Physicist's Conception of Nature, ed. J. Mehra (D. Reidel, Dordrecht), pp. 39-44
1153
Remarks on Enrico Fermi The Physicist s Conception of the Universe, ed. J. Mehra (D. Reidel, Dordrecht), pp. 800-02
1164
Shakespeare, Newton, and Beethoven or Patterns of Creativity The Nora and Edward Ryerson Lecture, 1975 (The University of Chicago Center for Policy Study)
1167
Einstein and General Relativity — Historical Perspectives Oppenheimer Memorial Lecture, 1978 (Los Alamos Scientific Laboratory, LASL-78-91)
1206
The Aesthetic Base of the General Theory of Relativity Karl Schwarzschild Lecture; Sounderdruck aus "Mitteilungen der Astronomischen Geselshaft Nr. 67" (Hamburg, 1986)
1222
The General Theory of Relativity: Why "It Is Probably the Most Beautiful of All Existing Theories" 1253 Wolfgang Pauli Lecture; Journal of Astrophysics and Astronomy 5 (1984): 3-11
1092 11. The Pursuit of Knowledge Paper read at the Regents' Fellows Colloquium on "The Pursuit of Knowledge," arranged by the Smithsonian Institution in Washington, D.C., on 5 May 1980
1262
12. The Perception of Beauty and the Pursuit of Science Presented at the 209th Annual Meeting of the American Academy of the Arts and Sciences, 10 May 1989
1270
13. The Pursuit of Science: Its Motivations Current Science 54, no. 4 (1985): 161-69
1278
14. A Commentary on Dirac's Views on "The Excellence of General Relativity" Festschrift for Val Telegdi, ed. K. Winter (Elsevier, 1988)
1287
15. The Increasing Role of General Relativity in Astronomy Halley Lecture for 1972, The Observatory 92, no. 990 (1972): 160-74
1295
16. Why Are the Stars as They Are? Proceedings of the International School of Physics "Enrico Fermi," Course 65 (1975), Physics and Astrophysics of Neutron Stars and Black Holes, ed. R. Gianconni and R. Rutherford (North-Holland, New York, 1978), pp. 1-14
1310
17. Black Holes: The Why and the Wherefore New Horizons of Human Knowledge
1324
18. How One May Explore the Physical Content of the General Theory of Relativity Proceedings of the Yale Symposium in Honor of the 150th Anniversary of the Birth of J. Willard Gibbs (American Mathematical Society and American Physical Society), pp. 227-51
1339
19. On Ramanujan Ramanujan Revisited (Academic, Boston, 1988), pp. 1-6
1364
20. Reminiscences and Discoveries: On Ramanujan's Bust Notes and Records, Royal Society of London 49, no. 1 (1995): 153-57
1368
21. On Reading Newton's Principia at Age Past Eighty Current Science 67, no. 7 (1994): 495-96
1373
22. Newton and Michelangelo Current Science 67, no. 7 (1994): 497-99
1379
23. On Stars, Their Evolution and Their Stability Nobel Lecture, 1983 (includes a brief biographical sketch); Reviews of Modern Physics 56, no. 2 (1984): 137-47
1387
The Scientist S.
CHANDRASEKHAR
confess at the outset to a feeling of apprehension at being included in this series on "Works of the Mind," as I am deeply aware of my shortcomings to speak either with assurance or with authority on a subject as wide and comprehensive in its scope as a discussion of the creative works of the scientist must be. But while I have these misgivings about the appropriateness of my representing the scientist in this series, I have no such misgivings in the choice of astronomy and astrophysics to represent the exact sciences. For astronomy among the exact sciences is the most comprehensive of all, and perfection in its practice requires scholarship in all its many phases. In another respect also, astronomy holds a unique position among the sciences, as it is the only branch of the ancient sciences which has come to us intact after the collapse of the Roman Empire. Of course, the level of astronomical studies dropped within the boundaries of the remnants of the Roman Empire, but the traditions of astronomical theory and practice were never lost. On the contrary, the clumsy methods of Greek trigonometry were improved by Hindu and Arabic astronomers, and new observations were constantly compared with those of Ptolemy and so on. This must be paralleled with the total loss of understanding of the higher branches of Greek mathematics before one realizes that astronomy is the most direct link connecting the modern sciences with the ancient. Indeed, the
I
MUST
159
THE
WORKS
OF
THE
MIND
works of Copernicus, Tycho Brahe, and Kepler can be understood only by constant reference to the ancient methods and concepts, while the Greek theory of irrational magnitudes or the Archimedes method of integration were understood only after being independently discovered by the moderns. The sponsors of this series have indicated that each contributor will demonstrate the value of his art or profession, by elucidating its nature, formulating its purpose, and explaining its techniques. But before I discuss these questions, I want to draw your attention to one broad division of the physical sciences which has to be kept in mind: the division into a basic science and a derived science. You will notice that my distinction is not between a "pure science" and an "applied science." I shall not be concerned with the latter, as I do not believe that the true values of science are to be found in the conscious calculated pursuit of the applications of science. I shall, therefore, be concerned only with what is generally called "pure science"; and it is the division of this into a basic science and a derived science that I wish to draw your attention. While an exact or a sharp division between the two domains cannot be made or maintained, that it exists all the same will become apparent from the examples I shall presently give. But, broadly speaking, we may say that basic science seeks to analyze the ultimate constitution of matter and the basic concepts of space and time. Derived science, on the other hand, is concerned with the rational ordering of the multifarious aspects of natural phenomena in terms of the basic concepts. Stated in this manner, it is evident, first, that the division is dependent on the state of science at a particular time and, second, that there may be, and indeed are, different levels in which natural phenomena can be analyzed. For example, there is the domain of the Newtonian laws in 160
THE
SCIENTIST
which an enormous range and variety of phenomena find their direct and natural explanation. And then there is the domain of the quantum theory in which other types of problems receive their solution. When there are such different levels of analysis, there exist criteria which will enable us to decide when one set of laws is appropriate and the other set inappropriate or clumsy as the case may be. But to return to the division itself. I do not believe that there exists a better example of a basic discovery than Rutherford's discovery of the large angle scattering of a-particles. The experiment he performed was quite simple. Using a source of high-energy a-particles emitted by a radioactive substance, Rutherford allowed them to fall on a thin foil and found that sometimes the a-particles were actually scattered backward—rarely but certainly. Recalling this later in his life (1936), Rutherford said:"This was quite the most incredible event that has ever happened to me in my life." He has described his immediate reactions in the following words: "It was almost as incredible as if you fired a 15-inch shell at a tissue paper and it bounced back and hit you." He further records: "On consideration I soon realized that this scattering backwards must be the result of a single collision and when I made calculations I saw that it was impossible to get anything of that order of magnitude unless you took a system in which the greater part of the mass of the atom was concentrated in a minute nucleus. It was then that I had the idea of an atom with a minute massive center carrying a charge. I worked out mathematically what laws scattering should obey and I found that the number of particles scattered through a given angle should be proportional to the thickness of the foil, the square of the nuclear charge and inversely proportional to the fourth power of the velocity. These deductions were later verified by Geiger and Marsden 161
THE
WORKS
OF
THE
MIND
in a series of beautiful experiments." And so it was that the nuclear model of the atom which is at the basis of all science was born. A single observation and its correct interpretation led to a revolution in scientific thought unparalleled in the annals of science. I would suppose that the discovery of the neutron by Chadwick is in the same category, as it is now believed that these together with protons are the basic constituents of all nuclei. But you should not suppose from these two examples that all basic scientific facts are necessarily to be gathered only in the field of atomic physics. Indeed, the first example of a law which may be called "basic" is astronomical in origin. I refer to the discovery by Kepler of the laws of planetary motion after long and patient analysis of the extensive observations of Tycho Brahe. These laws of Kepler led Newton to his celebrated laws of gravitation, which occupied the central scientific arena for over two hundred years. I shall return to this matter presently in a somewhat different context, but the example suffices to indicate that only in the gravitational domain can astronomy lead directly to results of a basic character. As a further illustration of this I may cite the fact that the minute departures exhibited by the motion of the planet Mercury from the predictions based on Newtonian laws indicated and later confirmed the far-reaching changes in our concept of space and time implied by the general theory of relativity. In the same way, it is not impossible that the discovery by Hubble of the recession of extra-galactic nebulae with velocities proportional to their distance may lead to further modifications in our basic concepts. The brief illustration in terms of examples, which I have just given, might suggest that the true values of science are to 162
THE
SCIENTIST
be found in those pursuits which lead directly to advances which I have termed "basic." Indeed, there are many physicists who seriously take this view. For example, a very distinguished physicist, apparently feeling sorry for my preoccupation with things astronomical and with an intent to cheering me, said that I should really have been a physicist. T o my mind this attitude represents a misunderstanding of the real values of science. And the history of science contradicts it. Thus, from the time of Newton to the beginning of this century, the whole science of dynamics and its derivative celestial mechanics have consisted entirely in amplification, in elaboration, and in working out the consequences of the laws of Newton. Halley, Laplace, Lagrange, Hamilton, Jacobi, Poincare—all of them were content to spend a large part of their scientific efforts in doing exactly this, that is, in the furtherance of a derived science. A derision of the derived aspects of science, implying as it would a denial of the values which these men have so earnestly sought, is to my mind sufficiently absurd to merit no further consideration. Indeed, it must be apparent to an impartial observer that there is a complementary relationship between the basic and the derived aspects of science. The basic concepts gain their validity in proportion to the extent of the domain of natural phenomena which can be analyzed in terms of them. And, in limiting the domain of validity of these concepts, we recognize the operation of other laws more general than those we have operated with. Looked at in this way, science is"a perpetual becoming, and it is in sharing its progress in common effort that the values of science are achieved. With these remarks, I think that I may state in a more formal manner what I regard as the true values of science which a scientist in the practice of his profession seeks to attain. Scientific values consist in the continual and increasing 163
THE
WORKS
OF
THE
M I N D
recognition of the uniformity of nature. In practice this only means that the values are attained in a larger or smaller measure in extending, or equivalently limiting, the domain of applicability of our concepts relating to matter, space, and time. In other words, a scientist seeks continually to extend the domain of validity of certain basic concepts. In so doing, he attempts to discover the limitations, if any, of these same concepts, and in this way he tries to formulate concepts of wider scope and generality. These values which are the quest of a scientist take in practice one of three distinct forms which I shall discuss under the headings: "Universality of the Basic Laws," "Predictions Based on the Basic Laws," and "Identifications Resulting from the Basic Laws." Let me illustrate each of these by some examples. UNIVERSALITY OF THE BASIC LAWS
In some ways the universality of the laws of nature is best illustrated by showing how this was achieved with regard to the law of gravitation. It is found that, all over the earth, objects are attracted toward the center of the earth. How far does this tendency go? Can it reach as far as the moon? These were the questions which Newton asked himself and answered. Galileo had already shown that uniform motion is as natural as rest and that deviation from such motion must imply a force. If, then, the moon were relieved of all forces, it would leave its orbit and go off along the instantaneous tangent to the orbit. Consequently, if the motion of the moon is due to the attraction of the earth, then what the attraction really does is to draw the moon out of the tangent into the orbit. As the period and the distances of the moon are known, it is easy to compute how much the moon falls away from the tangent in one second. Comparing this with the speed of falling bodies on the 164
THE
SCIENTIST
earth, Newton found the ratio of the two speeds to be about as 1 to 3600. As the moon is sixty times as far from the center of the earth as we are, this implies a force that decreases with the square of the distance. The next question Newton asked himself was how universal this property is. In particular, does a similar force reside in the sun, keeping the planets in their orbits as the earth does the moon? The answer is to be found in Kepler's laws. Newton showed that Kepler's second law—that planets describe equal areas in equal times—implies a central force, that is, a force directed toward the sun; that the first law—that the planetary orbits are ellipses with the sun at one focus— is a consequence of the inverse square law of attraction; and, finally, if the same law holds from planet to planet, then the periods and distances should be related as in Kepler's third law. It was in this manner that Newton was able to announce his law of gravitation that every particle in the universe attracts every other particle with a force that varies inversely as the square of their distance apart and directly as the masses of the two particles. You will notice Newton's use of the word "universe" in his formulation and the clear indication that its importance arises from the universality. One further related observation. In 1803 William Herschel was able to announce from his study of close pairs of stars that in some instances the pairs represented real physical binaries, revolving in orbits about each other. Herschel was further able to show that the apparent orbits were ellipses and that Kepler's law of areas was also valid. In other words, this observation extended the validity of the laws of gravitation from the solar system to the distant stars. It is difficult for us to imagine how tremendous the impression was which this discovery of Herschel made on his contemporaries. A good deal of the progress of astronomy since Newton's 165
THE
WORKS
OF
THE
MIND
laws were announced has been concerned with their application to the motions in the solar system. Newton himself pointed out many of their chief consequences. T o mention only two of these, he found the correct explanation of the phenomenon of ocean tides, and also for the precession of the equinoxes, a phenomenon which had been discovered twenty centuries earlier by Hipparchus. The application of Newton's laws throughout the solar system is a task of incredible difficulty and one that has taxed the powers of such giants as Lagrange, Laplace, Euler, Adams, Delauny, Hill, Newcomb, and Poincare. I have already referred to the fact that the motion of the planet Mercury cannot be fully accounted for on Newtonian theory. Departures are found in the sense that there appears to be a slow revolution of the orbit as a whole at a rate which is in excess of what can be accounted for on Newtonian theory by 40 seconds of arc per century. It would seem that this is now satisfactorily accounted for in terms of Einstein's theory of general relativity. There are still many fields of astronomy to which Newton's laws could be profitably applied. The newest of these relates to the motions in the galaxy as a whole and the new branch of dynamics called "stellar dynamics," which is rapidly growing in range and scope. I may have a few things to say about this later. Let me turn from this classic example of the universality of nature's laws to a more recent advance which in some ways is as striking. I hardly need point out to an audience in 1946 that the phenomenon of nuclear transformation (more commonly called "atom-smashing") has been studied extensively in recent years. Using the information obtained from such studies, Bethe was able to announce a few years ago that certain nuclear transformations involving carbon and 166
THE
SCIENTIST
nitrogen can lead indirectly to the synthesis of a helium nucleus from four protons. He was further able to show that, under the conditions for the interior of the sun which had been derived earlier by astrophysicists and with the crosssections for the reactions found in the laboratory, we can account in a most satisfactory way for the source of energy of the sun—a striking example of the synthesis of many types of investigations. Let me consider one further example. In 1926 Fermi and Dirac were led to a reformulation of the laws of statistical mechanics as they applied to an electron gas and showed that departures from classical laws should be expected at high densities and/or low temperatures. The nature of the departures predicted was this. According to classical laws, the pressure is proportional to the concentration and temperature. If at a given temperature we increase the concentration, then departures set in, in the sense that the pressure begins to increase more rapidly with the concentration and eventually becomes a function of the concentration only. When such a state is reached, one says that the electron gas has become degenerate. These new laws have found extensive applications in the theory of metals and are of the greatest practical importance. But the first application of the new laws was found in an astrophysical context by R. H. Fowler, who used the laws of a Fermi-Dirac gas to elucidate the structure of very dense stars such as the companion of Sirius. These dense stars, commonly known as white dwarfs, have densities of the order of several tons per cubic inch. The most extreme example is a star discovered some years ago by G. P. Kuiper which is estimated to have a density of 620 tons per cubic inch. Fowler immediately recognized that under these conditions the electrons must be degenerate in the sense of the Fermi-Dirac statistics. And 167
THE
WORKS
OF
THE
MIND
with this discovery of Fowler it became possible to work out the constitution of the white dwarf stars. This subject of the structure of the white dwarf stars has interested me personally, and I may be pardoned if I dwell on it a little longer. Extending Fowler's discussion, it soon became apparent that the laws of Fermi and Dirac required further modifications to take account of the fact that, at the high densities prevailing in the white dwarf stars, there will be a considerable number of electrons moving with velocities comparable to that of light. When modifications resulting from such high velocities are included, it was found that there is an upper limit to the mass of dense stars. This upper limit is in the neighborhood of 1.5 solar masses. The reason for the appearance of this upper limit is that for larger masses no stable equilibrium configuration exists. The recognition of this upper limit raises many questions of interest concerning stellar evolution. And it is not impossible that the occurrence of the supernova phenomenon is in some ways related to this. I shall not go into these matters now, but I mention these only to draw your attention to the manner in which the domain of validity of certain basic laws is continually being extended. In the three illustrations I have given, I have discussed the applicability of the same laws. But sometimes we have the application of the same set of ideas to problems which may appear entirely unrelated at first sight. For example, it is surprising to realize that the same basic ideas which account for the motions of microscopic colloidal particles in solution, also account for the motions of stars in clusters. This basic identity of the two problems which is far-reaching is one of the most striking phenomena I have personally encountered, and I would like to say a few words about it. The phenomenon of what is called "Brownian motion" was discovered by the English botanist Brown in 1827, who 168
THE
SCIENTIST
observed that when small particles (in his case pollen) are suspended in water, then instead of settling down, they get into a state of perpetual agitation. It may be amusing to recall that at first this perpetual motion was thought of as due to the life in the pollens, but Brown soon showed that this cannot be, for even the fine dust gathered from the Sphinx in Egypt exhibited the same behavior! We know now that the Brownian motion arises in consequence of the collisions which the colloidal particles suffer with the molecules of the surrounding fluid. Since even the smallest colloidal particle is several million times more massive than the individual molecules, it is apparent that a single collision can hardly make an impression on the colloidal particle. But the cumulative effect of a large number of collisions can become appreciable, and it is indeed the cause of Brownian motion. It is remarkable that the same methods which have been used to study Brownian motion can also be applied to study the motions of stars in a cluster like the Pleiades. The reason why we can do this is the following: When two stars in a cluster pass each other, the direction and magnitude of the motion of each of the stars change. But on account of the inverse square character of the forces between the stars, each such passage alters the motions only by a very minute amount. And again it is the cumulative effect of a large number of such passages which produces an appreciable change. The analogy with Brownian motion is apparent, and the theory of the motions of stars can be developed along the lines of the theory of Brownian movement. Indeed, the theory in the stellar case gives a more complete picture of Brownian motion than even the motion of colloidal particles provides. I may further mention that the development of this theory enables us to predict in a general way the evolution of star clusters and provides estimates for the time scale of the universe. 169
THE
WORKS
OF
THE
M I N D
PREDICTIONS BASED ON THE BASIC LAWS
I shall now pass on to another aspect of scientific investigations in which predictions are made on the basis of laws derived from other evidence and confirmations are later sought for these same predictions. I suppose that the most spectacular of such predictions made in recent times and which was later confirmed is that of Halley. In 1705 Edmund Halley communicated to the Royal Society his memoir on Astronomiae Cometicae Synopsis. This classic paper surveys the subject from the earliest times to those of Newton. Next, on Newtonian principles Halley calculated parabolic elements for twenty-four properly observed comets from A.D. 1337 to 1698. It is captivating to turn the pages of this memorable and deeply interesting contribution to knowledge in which no pains are spared for accuracy and completeness. And it is also in this contribution that Halley reflects on the possibility, or rather the probability, that comets move rather in highly elliptical orbits than in parabolic trajectories. In the latter case comets would come from infinity and go to infinity. However, in the former case, they would be members of the solar system and would be expected to return after a lengthy period of years. It was indeed in view of this probability that Halley undertook the immense computational task, in order that if a new comet should appear we could, on comparing its elements with the computed ones, see whether or not it was an old one returned. He says, further, that many things persuaded him to the belief that the comet of 1531 was the same that was observed in 1607 and that he himself had observed in 1682 and which, moreover, he considered confirmed by one which was roughly observed in 1456. After which he adds: "Whence I would venture to predict with confidence a return of the same anno scil in 1758." Such was the origin of Halley's comet, the 170
THE
SCIENTIST
most celebrated of the comets. It came as predicted, the year after Halley's death and has appeared on two further occasions since. A more recent example of such a prediction, which was later confirmed, is that of Dirac concerning the positron. In 1928, by a stroke of singular genius, Dirac wrote down the equation of an electron which predicted a number of things in agreement with experiments. But his equation also predicted that an electron should have states in which it should have negative energy—an unheard-of possibility! However, Dirac was so convinced, on general grounds, of the correctness of his equation that he concluded that these states cannot be denied their reality. And to get over the difficulty of all electrons dropping to states of negative energy and producing a bizarre world around us, Dirac suggested that normally all states of negative energy are occupied so that the few remaining electrons with positive energy cannot get into such states, normally speaking, that is. However, he pointed out at the same time that under certain conditions an electron in a state of negative energy could be raised to a state of positive energy, creating in this way an electron and an absence of an electron in the infinite distribution of negative energy. And this hole in the distribution of negative energy will behave as though it were a perfectly sensible particle with a positive energy, but with a positive charge. This is the positron, and the phenomenon he had considered is the creation of an electron pair. Dirac even worked out a theory for the probability of such pair creations. Some three years later all these predictions were verified and confirmed his conviction in the absolute correctness of his equation. A third and final example I shall consider in this group is Einstein's prediction of the deflection of light in a gravitational field and how it was later confirmed. In telling this 171
THE
WORKS
OF
THE
MIND
story, I shall quote from a lecture by Eddington, who was chiefly responsible for the verification. "The most exciting event I can recall in my own connection with Astronomy is the verification of Einstein's prediction of the deflection of light at the eclipse of 1919. The circumstances were unusual. Plans were begun in 1918 during the war, and it was doubtful until the eleventh hour whether there would be any possibility of the expeditions starting. But it was important not to miss the 1919 eclipse, because it was in an exceptionally good star field: none of the subsequent expeditions have had this advantage. Two expeditions were organized at Greenwich by Sir Frank Dyson, the late Astronomer Royal, the one going to Sobral in Brazil and the other to the isle of Principe in West Africa. [Eddington was in charge of the expedition which went to Principe.] It was impossible to get any work done by the instrument-makers until after the armistice; and as the expeditions had to sail in February there was a tremendous rush of preparation. The Brazil party had perfect weather for the eclipse; through incidental circumstances their observation could not be reduced until some months later, but in the end they provided the most conclusive confirmation. I was at Principe. There the eclipse day came with rain, and cloud-covered sky, which almost took away all hope. Near totality, the sun began to show dimly; and we carried through the program, hoping that the conditions might not be so bad as they seemed. The clouds must have thinned before the end of totality, because amid many failures, we obtained two plates showing the desired star images. These were compared with the plates already taken of the same star field at a time when the sun was elsewhere, so that the difference indicated the apparent displacement of the stars due to the bending of the light rays in passing near the sun. 172
THE
SCIENTIST
"As the problem then presented itself to us, there were three possibilities. There might be no deflection at all; that is, light might not be subject to gravitation. There might be a 'half deflection' signifying that light was subject to gravitation as would follow from Newtonian law. Or there might be a full deflection confirming Einstein's instead of Newton's law. I remember Dyson explaining all this to Cottingham who gathered the main idea, that the bigger result the more exciting it would be. 'What will it mean if we get double the deflection.' 'Then,' said Dyson, 'Eddington will go mad and you will have to come home alone.' "Arrangements had been made to measure the plates on the spot, not merely from impatience, but as a precaution against mishap on the way home; so one of the successful plates was examined immediately. The quantity to be looked for was large as astronomical measures go, so that one plate would virtually decide the question, though confirmation from others would be sought. Three days after the eclipse, as the last lines of the calculation were reached, I knew that Einstein's theory had stood the test and the new outlook of scientific thought must prevail. Cottingham did not have to go home alone." IDENTIFICATIONS RESULTING FROM THE BASIC LAWS
I now want to consider a third aspect of scientific investigations which is in a sense intermediate to the kinds I have already described. During the eighteenth century the idealist philosopher Bishop Berkeley and his followers claimed that the sun, the moon, and the stars are but "so many sensations in our mind" and that it would be meaningless to inquire, for example, as to the composition of the stars. And yet, only a few decades later, Kirchhoff was to announce in 1860 his mo173
THE
WORKS
OF
THE
MIND
mentous chemical interpretation of the Fraunhoffer lines showing the presence of the familiar metals as glowing vapors in the sun's atmosphere. From this time onward, to speak of the composition of stars was no longer in the realm of idle dreaming. It became a problem of intense practical interest. It is hard to believe mat, in the eighty years which have followed, the whole problem of the interpretation of the innumerable spectra which have been observed, both in laboratory and in stellar sources, should have reached a stage so near completion. The story of these investigations forms one of the most romantic chapters in the history and methods of science: Indeed, for most part it cannot be distinguished from the history of physics, chemistry, and astronomy of the last fifty or more years. And if I select from this whole wide field two incidental details for special consideration, it is not because I attach any undue importance or significance to them, but only because I happen to be specially interested in them. What I want to say relates to the two atoms, which are, next to hydrogen, the simplest. These are the atoms with two electrons: helium and the negative ion of hydrogen. First, regarding helium. Until March, 1895, helium was known only as a chromospheric element on the sun. It had, in fact been detected during the total eclipse of the sun which occurred in August, 1868, by the French astronomer, Jansen. What Jansen observed was, that in the chromospheric spectrum which appears during the flash at the instant of totality, there is a bright yellow line at X 5876 A near the wellknown lines of sodium. At first it was thought that this line may also be due to sodium, but it was Sir Norman Lockeyer who first realized that this cannot be so and that this new line could not be identified with the lines produced by any of the then known terrestrial elements. He, therefore, concluded that a new element was involved, and, since it had been de174
THE
SCIENTIST
tected on the sun, he called it "helium." In 1895, a quarter of a century later, the well-known chemist, Sir William Ramsay, while studying the gases evolved by certain uranium minerals, examined their spectra. He found that in the spectra thus obtained there was a brilliant yellow line exactly at the place of the chromospheric line. Further investigation confirmed that in both cases the same source was involved, and it was thus that the element which had been first detected on the sun was later isolated on the earth. The story of the negative ion of hydrogen is in some ways equally fascinating. That an atom composed of a proton and two electrons can exist in a free state was established by Bethe and Hylleraas on theoretical grounds. The calculations of Bethe and Hylleraas are such straightforward consequences of the quantum theory that there can be no doubt either of its stability or of its ability to exist in a free ssate under suitable conditions. But all the same, the fact remains that so far the negative hydrogen ion has not been isolated as such in the laboratory. However, it was pointed out by Wildt some years ago that negative ions of hydrogen must exist in the free state and in considerable numbers in the atmosphere of the sun. The question arises, "Can we detect it?" In order that we may do this, it is first necessary to know the manner in which the negative ion of hydrogen would absorb light and further the effect such an absorption would have on the solar spectrum. The theoretical problem of determining how the negative ion of hydrogen would absorb light turns out to be an exceptionally delicate matter. But the underlying physical problem has now been solved, and it is possible to predict with a fair degree of certainty the effects which may be observed in the solar spectrum. The nature of these effects is so clear cut, and they are so fully borne out by the observations, that it is no exaggeration to say that this atom consisting of a 175
THE
WORKS
OF
THE
MIND
proton and two electrons which was predicted by the quantum theory to have a stable existence, has now been identified on the sun. So far I have considered only the sort of things a scientist seeks in the practice of his profession. And I have left to the last the consideration of his motivations. There are several schools of thought here. I reject the view that the motivation springs from a conscious or a subconscious belief that everything he does will eventually find use for the amenities of daily life. And I reject also the corollary which insists that a scientist must always consciously integrate his efforts with the social needs and urgencies of his time. But I do not also accept the view that scientists are urged on in their work by a "holy passion" for truth and a "burning curiosity" to unravel the "secrets" of nature. I do not believe that in the daily practice of his profession a scientist has much in common with the Buddha who renounced his princely life to contemplate on the ethical and moral values which give meaning and significance to life. And, I am afraid, he has very little in common also with Marco Polo. What actually does give substance and reality to the efforts of a scientist is his desire actively to participate in the progress of his science to the best of his ability. And if I have to describe in one word what is the prime motive which underlies a scientist's work, I would say systematization. That may sound rather prosaic, but I think it approaches the truth. What a scientist tries to do essentially is to select a certain domain, a certain aspect, or a certain detail, and see if that takes its appropriate place in a general scheme which has form and coherence; and, if not, to seek further information which would help him to do that. This is perhaps somewhat vague, particularly, the use of the words "appropriate," 176
THE
SCIENTIST
"general scheme," "form," and "coherence." I admit that these are things which cannot be defined any more than beauty in art can be defined; but people who are acquainted with the subject have no difficulty in recognizing or appreciating it. Let me try, however, to explain what I mean by considering very briefly two examples. The phenomenon of radioactivity was discovered by Henri Becquerel in 1896. We know now that there are three radioactive families; that in a radioactive transformation one or more of three distinct types of radiation can be emitted; that the laws of radioactive displacements are involved; that the recognition of the existence of isotopes and isobars is involved; and that a novel theory of spontaneous disintegration of atoms is involved. You can imagine the enormous variety and complexity of the phenomena which radioactivity must have presented to those who did not know any of these things. And yet, when Rutherford's first edition of Radioactivity appeared in 1904, the essentials of the phenomenon had already been unraveled. This was largely because the problem had been systematically investigated with that energy, orderliness, and thoroughness which are the characteristics of Rutherford. To take another example. During the period including the first World War and the twenties, the very immense task of unraveling complex atomic spectra was undertaken by the physicists—a task which could not have been completed without conscious efforts toward what I have called "systematization." And it was in this way also that the principles of the quantum theory came to be established during the late twenties. And, as the sponsors of this series have expressly asked each contributor to speak from personal experience, I may be permitted to add that the method I have adopted in my own 177
THE
WORKS
OF
THE
MIND
work has always been first to learning what is already known about a subject; then to see if it conforms to those standards of rigor, logical ordering, and completion which one has a right to ask; and, if it does not, to set about doing it. The motivation has always been systematization based on scholarship. And I venture to suppose that this is true quite generally. In any case, it would seem to me that only in that way can a healthy scientific life be led and the real scientific values be achieved. I am afraid I have left myself no time to discuss a very important phase of scientific work, namely, its co-operative nature. I shall, therefore, content myself with just one quotation from Rutherford: "It is not in the nature of things for any one man to make a sudden, violent discovery; science goes step by step and every man depends on the work of his predecessors. When you hear of a sudden unexpected discovery—a bolt from the blue, as it were—you can always be sure that it has grown up by the influence of one man on another, and it is the mutual influence which makes the enormous possibility of scientific advance. Scientists are not dependent on the ideas of a single man, but on the combined wisdom of thousands of men, all thinking of the same problem and each doing his little bit to add to the great structure of knowledge which is gradually being erected." That is the opinion of one of the greatest—I would almost say, the greatest—physicist of our time. And you can, therefore, understand why it is that scientists are always internationalists and why it is also that so much apprehension is being expressed by them now at die prospective limitations of the freedom of science. 178
THE
SCIENTIST
And, finally, one may ask, "What is the case for the life of a scientist?" The case is this: that he has added something to knowledge and helped others to add more and that these somethings have a value which differs in degree only, and not in kind, from that of the creations of the great scientists or of any of the other artists, great or small, who have left some kind of memorial behind them.
179
1114
THE CASE FOR ASTRONOMY S. CHANDRASEKHAR Professor of Theoretical Astrophysics, University of Chicago {Read April 19, 1963) "When Science does not advance, the fault may be in our stars and not so much in ourselves."—Lord Adrian
O. NEUGEBAUER has pointed out that "astronomy is the only branch of the ancient sciences which survived almost intact after the collapse of the Roman Empire"; and further that "the tradition of astronomical theory and practice was never completely lost." In view of this tradition and in view of the present enlarging interest in the spatial environment of the sun and the earth, a restatement of the values of astronomy, in current terms, may be useful in restoring perspective to these values as a part of the heritage which has been handed down to us from ancient times. Throughout its long history, astronomy, taking its place among the physical sciences, has sought its values as a physical science. II Now the study of the physical sciences is generally carried out at two levels: a primary level at which one seeks to formulate the basic elemental laws which govern the behavior of all physical systems and a secondary level at which one seeks to analyze and interpret particular complexes of phenomena in terms of the basic laws. Thus, Newton's law of gravitation 1 that "every particle of matter in the universe attracts every other particle with a force whose direction is that of the line joining the two and whose magnitude is directly as the product of their masses and inversely as the square of their distance from each other" is an example of a basic law formulated at the first level; and the application of this same law, by Newton himself, to the interpretation of Kepler's laws of planetary motion is an example of the analysis at the second level. 1
Philosophic Naturalis Principia Mothematica, Book III, Propositions I-VII. The formulation as quoted is in Thomson and Tait's Natural Philosophy, Part II, p. 9.
The primary level at which one seeks to formulate the basic laws governing the idealized elementary situations is the domain of physics as commonly understood. The secondary level at which one seeks to apply these same laws to the understanding of particular complexes of phenomena is the domain of the various special branches of the physical sciences. And astronomy is one such branch.
Ill While the separation of the study of the nature of the physical world into physics and its various branches is a convenient one, it is not, of course, a sharp one. The various lines of separation are necessarily diffuse: there is no general law of physics which has not been derived from the study of particular physical phenomena ; and, conversely, there is no physical phenomenon which does not, in some measure, illuminate or illustrate some general physical law. Astronomy, then, as a study devoted to an aspect of the physical world, must illuminate and illustrate specific laws of physics. The relationship between physics and astronomy which thus emerges is a fundamental and a reciprocal one. It is a relationship which can be traced in many of the current developments of astronomy. A few concrete examples may illustrate the reciprocity of this relationship. IV It was an established fact of astronomy in 1924 that there are stars (the so-called whitedwarf stars) in which matter attains densities which are two thousand and more times the density of platinum. Analyzing this fact in terms of the physics of that time, Eddington arrived at the paradoxical conclusion: "a star which has once got into this state can never get out of it: it is losing heat but without sufficient
PROCEEDINGS OF THE AMERICAN PHILOSOPHICAL SOCIETY, VOL. 108, NO. 1, FEBRUARY, 1964
Reprint
Printed in U.S.A.
1115
S. C H A N D R A S E K H A R
PROC. AMER. PHIL. SOC.
energy to grow cold"! This paradox arose from applying the gas laws of Boyle and Charles to the A similar relationship between the developstellar conditions which had been disclosed. And ments in physics and in astronomy can be traced Eddington's whimsical formulation can be re- in the way an understanding of the physical conphrased as follows: if one applies the gas-kinetic ditions in the solar atmosphere (more precisely, theory of Maxwell and Boltzmann (which pro- the solar photosphere) has been achieved. One vides the molecular basis for the laws of Boyle of the striking aspects of the solar disc is its and Charles) to matter at the densities and tem- systematic darkening towards its circumferential peratures which occur in the white-dwarf stars, edge (the limb) which is apparent on any direct then one finds that the thermal energy of motions photograph of the sun. This darkening towards is less than the energy that would be needed to the limb can be observed and measured not only overcome the electrostatic binding of the electrons in the integrated white light but, also, in the and the atomic nuclei at those densities. Conse- different wave lengths (colors). A little conquently, if such dense matter were removed sideration shows that the observed darkening of from the white-dwarf stars and relieved of the the solar limb is very directly related to the containing pressure, we would find (on the stated nature of the opacity of the solar atmosphere to premises) that it is unable to expand and resume light. Indeed, it was shown by E. A. Milne a "normal state" of high temperatures. This in 1922 that the observed extents of the darkenwould be most curious if it were true. But R. H . ing in the different wave lengths enable one to Fowler resolved Eddington's paradox in 1926 deduce in a relatively straightforward manner on the basis of certain new developments in statistical mechanics which had occurred a few the dependence on wave length of the opacity months earlier. These new developments, due of the solar atmosphere. And Milne deduced that the opacity must show a very pronounced to Fermi and Dirac, required a modification of the classical gas laws when matter is compressed maximum at about 8,500 A and a deep minimum to the state it is in, in the white-dwarf stars. at 16,000 A. And other astrophysical investigations on the colors of stars established that a In such a dense state, the kinetic energy of motions of the electrons will be very greatly similar dependence must be present in the opacity underestimated, if the dependence on tempera- of the atmospheres of other stars as well. Howture of that energy is taken as a measure of its ever, two decades elapsed before it was possible amount (as we may at ordinary densities). And to identify the substance in the atmospheres Fowler showed that the large temperature-inde- of the sun and the stars which was responsible pendent reservoir of energy (the "zero-point" for the deduced dependence of the opacity on energy) in such a highly "degenerate" electron wave length. It was only in 1938 that R. Wildt gas is entirely ample to overcome the electrostatic suggested that the substance in question might well be the negative ion of hydrogen, i.e., a binding and provide for the matter, relieved of its containing pressure, an exceedingly high tem- hydrogen atom to which an electron is stably perature. This explanation of Fowler in terms bound. The remarkable aspect of this suggestion of the (then) newest developments in physics was that at that time the possible stable existence provided at the same time the foundations for a of such an atomic species had been established physical theory of the white-dwarf stars. only on theoretical ground's by H . A. Bethe and E. A. Hylleraas. But no doubts regarding its It is perhaps worth recalling in this connection existence could be entertained on that account. that Fowler's clarification and resolution of the However, the properties of the negative ion paradox of the white-dwarf stars provided the of hydrogen, with respect to its absorptive and first illustration of the deviations from the classi- emissive characteristics, had to be derived on the cal gas laws required by the theory of Fermi foundations of atomic physics and quantum theory and Dirac; and this is contrary to the general before Wildt's suggestion could be tested and belief that the first illustrations were provided confirmed. This was largely accomplished in by Pauli and Sommerfeld in the context of the the forties. And in the fifties, the negative ion 2 electron theory of metals. of hydrogen was isolated, in the free state, in the laboratory by L. Branscomb; and he was able to 2 For a more complete account of this history see S. confirm that its absorptive and emissive charChandrasekhar, "Ralph Howard Fowler," Astrophysical acteristics were in agreement with the requireJour. 101 (1944) : 1-5.
1116
VOL. io8, NO. i, 1964]
T H E CASE F O R A S T R O N O M Y
ments both of physical theory and of physical deductions.
astro-
VI A third example is provided by the manner in which progress in our understanding of the source of stellar energy has been achieved. From the study of the requirements of hydrostatic and of thermal equilibrium of configurations with the masses and the radii of the stars, it can be shown that, unless something extraordinary happens, the central temperatures of stars similar to the sun cannot differ from ten million degrees by any very substantial factor. And in the twenties there was a tendency among the physicists to consider temperatures of this order as "too low" to allow for the generation of energy at the observed rates. But the physical and the mathematical basis for the derived temperatures (and densities) appeared so secure that the expressions of such doubts elicited from Eddington the retort: " W e do not argue with the critic who urges that the stars are not hot enough for this process: we tell him to go and find a hotter place." 3 Nevertheless, the question as to the particular physical processes which could take place at the derived internal temperatures and densities of the normal stars and liberate the requisite amounts of energies remained unanswered for some fifteen years: experimental nuclear physics had to advance to the point where one had some systematic knowledge of the protoninduced nuclear reactions among the lighter elements. The author recalls that during the weeks and months following Cockroft and Walton's first success in disintegrating the nucleus of the lithium atom by protons accelerated to one hundred kev energies in the laboratory, Rutherford used to emphasize, both in conversations and in lectures, that the source of stellar energy must be sought in the net conversion of hydrogen into helium as a result of nuclear reactions effected by protons in their collisions, under the thermal conditions in the interiors of stars, with the nuclei of the lighter elements. And so it turned out. However, some six years had to elapse before it was possible for Bethe to outline the principal chain of nuclear reactions (the so-called carbon cycle) which effects the conversion of hydrogen into helium (through the medium of carbon and nitrogen acting as cata-
3 A. S. Eddington, The Internal Constitution of the Stars (Cambridge University Press, 1926), p. 301.
3
lysts) and is responsible for the energy production in the normal stars. For the energy production in stars of masses smaller than the sun (and in which central temperatures are appreciably less) another chain of nuclear reactions must be invoked; and the principal reaction of this chain is (as Bethe also showed) the fusion (by a /J-process) of two protons to form a deuteron. While the cross section for this most elementary of all ^-processes has been reliably evaluated in terms of established physical theory, the only empirical evidence that the reaction does take place, according to the principles of the physical theory, is still the astronomical one. VII As a last example of a somewhat different kind, we shall briefly consider the recent recognition (through the work of I. Shklovsky and others) of the important role which synchrotron radiation (which is the radiation emitted by electrons, with energies exceeding their rest energies by several orders, as they execute spiral orbits in a magnetic field) plays in diverse astronomical contexts. The polarization characteristic of this synchrotron radiation is very striking and its presence often identifies the mechanism. The strongly polarized continuous radiation (i.e., radiation of all wave lengths) from the remnants of supernovae (such as the Crab Nebula) is synchrotron radiation. And it is believed that the cause of the so-called nonthermal radio emission from galaxies and radio sources is also synchrotron radiation. And most recently, the identification, by Sandage and others of synchrotron emission as one of the principal components of the radiation from the enormous expanding filamentary mass of gas enveloping the galaxy M82, has brought to light evidence that an explosive event occurred in the center of this galaxy about a million years ago with an integrated intensity comparable to a million or more supernovae. These new discoveries have brought astronomy to the verge of a new great revolution. And yet, the necessary basic theory of the physics of synchrotron emission was fully worked out and published in 1912 by G. A. Schott in a well-known (but neglected!) treatise on Electromagnetic Radiation. (On this last account, the current name, synchrotron radiation, is somewhat misleading since it suggests that the phenomenon and its theory are of relatively recent origin.)
1117
4
S. C H A N D R A S E K H A R
VIII Many more examples of the kind we have described can be given. And let it not be forgotten either that the only crucial empirical basis for the aesthetically most satisfying physical theory conceived by the mind of man—Einstein's general theory of relativity—is the astronomical one derived from the motion of Mercury. The type of relationship between physics and astronomy which the examples we have described illustrate is in many ways an obvious and a necessary one: for astronomy is a physical science and is concerned with physical phenomena. The principal values of astronomy, then, are the same as those of any of the physical sciences; about this one can be certain. But with hesitation one can perhaps go a little further and ask whether there may not be some special values to be sought in astronomy in the sense that its study may reveal patterns of order and patterns of behavior which cannot be discerned in physical phenomena manifested in any smaller scale. The intent of this question requires some clarification. IX Consider physical phenomena which are governed by the laws of thermodynamics and statistical mechanics. The essential characteristic of a physical system which is governed by these laws is that it consists of loosely coupled assemblies each of which in turn consists of a very large number of practically independent identical "particles" or "elements." And because of the large number of elements in each assembly and because also of the weak coupling ("practical independence") between the elements (belonging to the same or different assemblies) a number of simplifications arise. W e can, for example, describe the system (in the framework of thermodynamics and statistical mechanics) in terms of far fewer variables (e.g., the partial pressures, the partial densities, and the common temperature) than a detailed microscopic specification of the state of the system will require. With this specification of the system in terms of the fewer thermodynamic variables a pattern of behavior with elements of irreversibility (implied in the second law of thermodynamics) emerges which has no counterpart in the laws governing the individual elements and their mutual interactions. Consider, next, systems which are a "scale"
[PROC. AMER. PHIL. soc.
larger than those to which we normally apply thermodynamics and statistical mechanics. Do these systems exhibit a new and a different pattern of behavior ? T h e question can be made more specific by considering hydrodynamical systems in which fluid motions occur. As is well known the equations of hydrodynamics are founded on statistical mechanics and thermodynamics. And in terms of these hydrodynamical equations we can, indeed, account for the normal behavior of fluid motion if under "normal" we include motions which are described as "laminary." T h e essential characteristic of systems which exhibit laminar motions is that they become determinate (at least, in the majority of cases) by the conditions at the boundaries. On the other hand, we are familiar with the circumstance that even the simplest of the laminar patterns of motion are not realized if the geometrical and the other parameters defining them (or, more strictly certain combinations of them) surpass certain critical values: for the laminar states are then unstable. T h e conditions which must be satisfied for the motions in a hydrodynamical system to be laminar are essentially conditions which restrict the linear scale of the systems. The patterns of fluid motion which are realized when the critical conditions are far surpassed are those which are realized under the natural conditions of meteorology, oceanography, and astrophysics. And the characteristics of the motions which prevail under these circumstances are certain randomness and disorder which one includes under the name of turbulence. While a comprehensive physical theory of turbulence is not yet available, it would appear that systems which exhibit turbulence must be described in terms of parameters which bear the same kind of relationship to the thermodynamic variables as the thermodynamic variables do to the Hamiltonian variables of microscopic physics. The question which now occurs i s : would we discern other patterns of behavior by going to systems which are a scale larger than those which exhibit turbulence? More generally, we may ask: do the patterns of behavior of physical systems arrange themselves in a hierarchical order determined by their scales? A n d if scale is the determining factor, then the role of astronomy in discerning and in ascertaining the hierarchy in the natural order of things becomes unique. While these considerations may border on the cogitative and the speculative, we may briefly
1118
T H E CASE FOR ASTRONOMY
VOL. 108, NO. 1, 1964]
pause and ask ourselves whether astronomy gives any indications of patterns of order and patterns of behavior which are not manifested in smaller systems. X Consider the simplest of the hydrodynamical problems, namely, the propagation of sound. It is known that in a homogeneous medium sound (or, more precisely, a small disturbance in pressure or density at some point) is propagated with a velocity which is the same for all frequencies (pitch). It is this independence of the velocity of propagation on pitch that makes possible a large auditorium in which music can be sounded and heard without distortion. Consider now the same problem in a medium as extensive as one encounters in astronomy, say, the interstellar medium. If we should assume that the medium is uniform and allow, as we must, for the gravitational effects of the condensations and rarefactions which take place concurrently with the propagation of a sound pulse, then, we should find (as Jeans first showed) that the velocity of propagation depends on the frequency and, further, that all notes of sufficiently low pitch "howl" : the music of the spheres, far from being harmonious, is cacaphonous! What we have to conclude from Jeans's fundamental result is that the concept of a medium which is strictly uniform is untenable if its extension requires us to include the gravitational effects of a sound pulse. While the concept of strict uniformity should be abandoned, the question arises whether it cannot be replaced by some other concept which will allow for a constant formation and decay of condensations and rarefactions (in which gravitational effects play an essential role) and which when averaged over sufficiently large spatial and temporal domains will lead to some notion of a statistically steady state of "supra-turbulence." And if such a state of supra-turbulence can be properly formulated and described then we shall have gone one scale, higher in the hierarchy of orders. XI Let us consider other empirical aspects of systems of the dimensions of the galaxies. Three principal elements characterize them. These a r e : the presence of rotation, the prevalence of magnetic fields, and the containment of high-energy particles such as the cosmic rays. And one must
FIG. 1. Two external galaxies: N.G.C. 2253 • (top) and N.G.C. 1398 (bottom). -From the Hubble Atlas of Galaxies (Carnegie Institution of Washington, 1961). add to these elements the possibility (indeed, the inevitability) of a constant exchange of energy between the three forms. The possibility of such exchanges of energy was first contemplated by Fermi, who showed that they can take place via the medium of the extensive magnetic fields which prevail in interstellar space. Clearly, we have here all the ingredients for the emergence of entirely new patterns of order and of behavior which have no counterparts in any scale smaller than that provided by astronomy. Quite apart from the general considerations of the preceding paragraphs, one's feeling, that patterns of order determined by the large scales of astronomy occur, is confirmed by objects such as are illustrated in the figure. That these objects exhibit striking patterns of order cannot be denied; and it cannot be denied, either, that they are manifested in the scales of astronomy. The
1119
6
S. C H A N D R A S E K H A R
existence of these orders and the need to know how they arise are the particular case for astronomy. XII To summarize: the principal case for astronomy is the same as the case for any of the physical sciences. No less; but, perhaps, more : for, only
[PROC. AMER. PHIL. SOC.
in the scales provided by astronomy can we discern the largest in the natural order of things. And it is possible that the case for astronomy is even more; for, as Eugene Wigner has expressed : "Physics can teach us only what the laws of nature are today. It is only Astronomy that can teach us what the initial conditions for these laws are."
ASTRONOMY IN SCIENCE AND IN HUMAN CULTURE Professor S. Chandrasekhar
It is hardly necessary for me to say how deeply sensitive I am to the honour of giving this second lecture in this series founded in the memory of the most illustrious name of independent modern India. As Pandit Jawaharlal Nehru has written, "The roots of an Indian grow deep into the ancient soil; and though the future beckons, the past holds back." I hope I will be forgiven if I stray for a moment from the announced topic of my lecture to recall, how forty-one years ago, I was one of thousands of students who went to greet young Jawaharlal (as we used to call him at that time) on his arrival to address the National Congress meeting in Madras that year. I recall also how the dominant feeling in all of us at that time was one of intense pride in the men amongst us and in what they inspired in us. Lokamanya Tilak, Mahatma Gandhi, Lala Lajpat Rai, Motilal Nehru, Jawaharlal Nehru, Sardar Patel, Sarojini Naidu, Rabindranath Tagore, Srinivasa Ramanujan—names that herald the giants that lived amongst us in that predawn era. The topic I have chosen for this lecture, "Astronomy in Science and in Human Culture," is so large that I am afraid that what I can say on this occasion can at best be a collection of incoherent thoughts. In the first part of the lecture I shall make some general observations on ancient Hindu 5
astronomy, particularly with reference to the way it relates Hindu culture to the other cultures of antiquity. I am not in any sense a student of these matters. My knowledge is solely derived from the writings of a distinguished historian of science, Professor Otto Neugebauer, who has kindly helped me in preparing this part of my lecture. In the second part of the lecture I shall say something about the particular role of astronomy in expanding the realm of man's curiosity about his environment. One aspect of astronomy is certain: it is the only science for which we have a continuous record from ancient times to the present. As Abu 1-Qasim Said ibn Ahmad wrote in 1068 in a book entitled, "The Categories of Nations": "The category of nations which has cultivated the sciences form an elite and an essential part of the creation of Allah." And he enumerated eight nations as belonging to this class : "The Hindus, the Persians, the Chaldeans, the Hebrews, the Greeks, the Romans, the Egyptians, and the Arabs." Chronologically, the interactions between the leading civilizations of the ancient world are far more complex than this simple enumeration suggests. And a study of these interactions provides us with a most impressive testimony to man's abiding interest in the universe around him. We know today that Babylonian astronomy reached a scientific level only a century or two before the beginning of Greek astronomy in the fourth century B.C. The development of Hellenistic astronomy, after its early beginnings, to its last perfection by Ptolemy in 140 A.D. is largely unknown. Then about three centuries later Indian astronomy, manifestly influenced by Greek methods, emerges. This last fact raises the question as to the way in which this transmission of information from Greece to India took place. Its answer is made particularly difficult since it implies possible 6
Persian intermediaries. Some centuries later, in the ninth century, Islamic astronomy appears influenced by Hindu as well as Hellenistic sources. While the Greek astronomy rapidly became dominant in the eastern part of the Muslim world from Egypt to Persia, the methods of Hindu astronomy persisted in Western Europe even as late as the fifteenth century, as I shall indicate later. As far as Babylonian astronomy is concerned, we know very little about its earlier phases. But it appears that a mathematical approach to the prediction of lunar and planetary theory was not developed before the fifth century B.C. : that is to say barely prior to the corresponding stage of development of Greek astronomy. It is, however, generally agreed that the development of Babylonian astronomy took place independently of the Greeks. An important distinction between the Babylonian and the Greek methods is this: Babylonian methods are strictly arithmetical in character and are not derived from a geometrical model of planetary motion; the Greek methods, on the other hand, have invariably had a geometric basis. This distinction enables us to identify their influence in Hindu astronomy, for example. Let me make a few remarks on Greek astronomy as it is relevant to my further discussion. The earliest Greek model that was devised to account for the appearance of planetary motion is that of Eudoxus in the middle of the fourth century B.C. On this model planetary motion was interpreted as a superposition of uniform rotations about certain inclined axes. In spite of many glaring inadequacies, this model had a profound impact on subsequent planetary theory. The culmination of Hellenistic astronomy is, of course, contained in Ptolemy's "Almagest"—perhaps the greatest book on astronomy ever 7
written; and it remained unsurpassed and unsuperseded until the beginning of the modern age of astronomy with Kepler. Ptolemy's modification of lunar theory is of special importance for the problem of the transmission of Greek astronomy to India. The essentially Greek origin of the Surya Siddhanta—which is the classical textbook of Hindu astronomy—cannot be doubted: it is manifested in the terminology, in the units used, and in the computational methods. But Hindu astronomy of the North does not appear to have been influenced by the Ptolemaic refinements of the lunar theory; and this appears to be true with planetary theory also. This fact is of importance: a study of Hindu astronomy will give us much needed information on the development of Greek astronomy from Hipparchus in 150 B.C. to Ptolemy in 150 A.D. In early Hindu astronomy, as summarized by Varaha Mihira in the Panca Siddhantika, we can distinguish two distinct methods of approach: the trigonometric methods best known through Surya Siddhanta and the arithmetical methods of Babylonian astronomy in the astronomy of the South. The Babylonian influence has come to light only in recent years; and I shall presently refer to its continued active presence in the Tamil tradition of the seventeenth and the eighteenth centuries. I should perhaps state explicitly here that the fact that Hindu astronomy was deeply influenced by the West does not by any means exclude that it developed independent and original methods. It is known, for example, that in Hindu astronomy the chords of a circle (of radius R) were replaced by the more convenient trigonometric function Sin <x=(R sin a). Before I conclude with some remarks on the simultaneous existence of two distinct astronomical traditions in India, I should like to illustrate my general remarks by two specific 8
illustrations which are of some interest. In 1825 Colonel John Warren, of the East India Company stationed at Fort St. George, Madras, wrote a book of over 500 quarto pages entitled Kala Sankalita with a Collection of Memoirs on the Various Methods According to Which the Southern Part of India Divided Time. In this book, Warren described how he had found a calendar maker in Pondicherry who showed him how to compute a lunar eclipse by means of shells placed on the ground and from tables memorized as he stated "by means of certain artificial words and phrases." Warren narrates that even though his informer did not understand a word of the theories of Hindu astronomy, he was nevertheless endowed with a memory sufficient to arrange very distinctly his operations in his mind and on the ground." And Warren's informer illustrated his methods by computing for him the circumstances of the lunar eclipse of May 31—June 1, 1825 with an error of + 4 minutes for the beginning, —23 minutes for the middle, and —52 minutes for the end. But it is not the degree of accuracy of his result that concerns us here; it is rather the fact that a continuous tradition still survived in 1825, a tradition that can be traced back to the sixth century A.D. with Varaha Mihira, to the third century in the Roman Empire and to the Seleucid cuneiform tablets of the second and the third centuries. A second instance I should like to mention is an example of the survival of Hindu astronomy in parts of the Western world that were remote from Hellenistic influences during the medieval times. A Latin manuscript has recently been published which contains chronological and astronomical computations for the year 1428 for the geographical latitude of Newminster, England. It used methods manifestly related to Surya Siddhanta. Obviously one has to assume 9
Islamic intermediaries for a contact of this kind between England of the fifteenth century and Hindu astronomy. While Surya Siddhanta manifests Greek influence, Babylonian influence has recently been established in the post-Vedic and pre-Surya Siddhanta period. For example, in the astronomy of that period, the assumptions of a longest day of 18 muhurtas and a shortest day of 12 muhurtas were made. This ratio of 3:2 is hardly possible for India. But it is appropriate for Mesapotamia; and possible doubts about the Babylonian origin of this ratio were removed when the same ratio was actually found in Babylonian texts. In addition a whole group of other parallels between Babylonian and Indian astronomy have since been established. Thus, the most characteristic feature of Hindu time reckoning—the tithis—occurs in Babylonian lunar theory. Clearly all these facts must be taken into account in any rational attempt to evaluate the intellectual contacts between ancient India and the Western world. This problem of the foreign contacts is by no means the only, or even the most important, fact that is to be ascertained. One must consider the Dravidic civilizations of the South on a par with the history, the language, and the literature of the Aryan component of Indian culture. It is, as Neugebauer has emphasized, this dualism of Tamil and Sanskrit sources that will provide for us, eventually, a deeper insight into the structure of Indian astronomy. In his book, "Rome Beyond Imperial Frontiers," Sir Mortimer Wheeler comes to the conclusion that "the far more extensive contacts with South India have been a blessing to the archeologists" but he adds that "these contacts had no influence on these cultures themselves." Hindu astronomy provides an example to the contrary. Exactly as it is possible to distinguish between commercial contacts 10
which India had through the Punjab or through the Malabar and Coromandal Coast, it is possible to distinguish the astronomy of the Surya Siddhanta on the one hand and the Tamil methods on the other. This distinction is indeed very marked. The Surya Siddhanta is clearly based on prePtolemaic Greek methods while the Tamil methods, in their essentially arithmetical character, manifest the influence of Babylonian astronomy of the Seleucid—Parthian period. One must not, of course, conclude that the Tamil methods were imported directly from Mesapotamia while the geometric methods came to the North via the Greeks and through, possibly, Persian intermediaries. And as I stated earlier, the fact that the Surya Siddhanta appears not to have been influenced by the Ptolemaic refinements, provides an important key to the development of Hellenistic astronomy between the times of Hipparchus and Ptolemy. A proper assessment of the role of Hindu science in the ancient world has yet to be made. The problem is made more difficult, than is necessary, by the tendency of the majority of publications of Indian scholars to claim priority for Hindu discoveries and to deny foreign influence, as well as the opposite tendency among some European scholars. These tendencies on both sides have been aggravated by the inadequate publication of the original documents; this is indeed a most pressing need. Since no astronomy at an advanced level can exist without actual computations of planetary and lunar ephimerides, it must be the first task of the historian of Hindu astronomy to search for such texts. Such texts are indeed preserved in great numbers, though actually written in very late periods. But the publication of this material is an urgent need in the exploration of oriental astronomy. Let me conclude this somewhat incoherent account, 11
bearing on the ancient culture of India, by emphasizing that its principal interest lies not in the sharing or in the apportioning of credit to one nation or another but rather in the continuing thread of common understanding that has bound the elite nations of Abu 1-Qasim Said ibn Ahmad in man's constant quest to comprehend his environment. The pursuit of astronomy at the more sophisticated level of modern science, since the time of Galileo and Kepler, is concerned with the same broad questions even though that fact is often obscured by the technical details of particular investigations. Questions that may naturally occur to one often appear to be meaningless in the context of current science. But with the progress of science, questions that appear as meaningless to one generation become meaningful to another. It is to this aspect of the development of astronomy in recent times that I should like to turn my attention now. The first question that I shall consider concerns the assumption that is implicit in all science: Nature is governed by the same set of laws at all places and at all times, i.e. Nature's laws are universal. That the validity of this assumption must be raised and answered in the affirmative was the supreme inspiration which came to Newton as he saw the apple fall. Let me explain. Galileo had formulated the elementary laws of mechanics governing the motions of bodies as they occur on the earth; and the laws he formulated were based on his studies of the motions of projectiles, of falling bodies, and of pendulums. And Galileo had, of course, confirmed the Copernican doctrine by observing the motions of the satellites of Jupiter with his telescope. But the question whether a set of laws could be formulated which governed equally the motions of all bodies, whether they be of stones thrown on the earth or 12
of planets in their motions about the sun, did not occur to Galileo or his contemporaries. And it was the falling apple that triggered in Newton's mind the following crucial train of thought. All over the earth objects are attracted towards the center of the earth. How far does this tendency go? Can it reach as far as the moon? Galileo had already shown that a state of uniform motion is as natural as a state of rest and that deviations from uniform motion must imply force. If then the moon were relieved of all forces, it would leave its circular orbit about the earth and go off along the instantaneous tangent to the orbit. Consequently, so argued Newton, if the motion of the moon is due to the attraction of the earth, then what the attraction really does is to draw the moon out of the tangent and into the orbit. As Newton knew the period and the distance of the moon, he could compute how much the moon falls away from the tangent in one second. Comparing this result with the speed of falling bodies, Newton found the ratio of the two speeds to be about 1 to 3600. And as the moon is sixty times farther from the center of the earth than we are, Newton concluded that the attractive force due to the earth decreases as the square of the distance. The question then arose: If the earth can be the center of such an attractive force, then does a similar force reside in the sun, and is that force in turn responsible for the motions of the planets about the sun? Newton immediately saw that if one supposed that the sun had an attractive property similar to the earth, then Kepler's laws of planetary motion become explicable at once. On these grounds, Newton formulated his law of gravitation with lofty grandeur. He stated: "Every particle in the universe attracts every other particle in the universe with a force directly as the product of the masses of the two particles 13
and inversely as the square of their distance apart." Notice that Newton was not content in saying that the sun attracts the planets according to his law and that the earth also attracts the particles in its neighbourhood in a similar manner. Instead with sweeping generality, he asserted that the property of gravitational attraction must be shared by all matter and that his law has universal validity. During the eighteenth century, the ramifications of Newton's laws for all manner of details of planetary motions were investigated and explored. But whether the validity of Newton's laws could be extended beyond the solar system was considered doubtful by many. However, in 1803 William Herschel was able to announce from his study of close pairs of stars that in some instances the pairs represented real physical binaries revolving in orbits about each other. Herschel's observations further established that the apparent orbits were ellipses and that Kepler's second law of planetary motion, that equal areas are described in equal times, was also valid. The applicability of Newton's laws of gravitation to the distant stars was thus established. The question whether a uniform set of laws could be formulated for all matter in the universe became at last an established tenet of Science. And the first great revolution in scientific thought had been accomplished. Let me turn next to the second great revolution in the explicit context of astronomy that was accomplished during the middle of the last century. During the eighteenth century the idealist philosopher Bishop Berkeley claimed that the sun, the moon, and the stars are but so many sensations in our mind and that it would be meaningless to inquire, for example, as to the composition of the stars. And it was an oft-quoted statement of Auguste Compte, a positivist philosopher influential 14
during the early part of the nineteenth century, that it is in the nature of things that we shall never know what the stars are made of. And yet that very question became meaningful and the center of astronomical interest very soon afterwards. Let me tell this story very briefly. You are familiar with Newton's demonstration of the character of white light by allowing sunlight to pass through a small round hole and letting the pencil of light so isolated fall on the face of a prism. The pencil of light was dispersed by the prism into its constituent rainbow colors. In 1802 it occurred to an English physicist, William Wollaston, to substitute the round hole, used by Newton and his successors to admit the light to be examined with the prism, with an elongated crevice (or slit as we would now say) l/20th of an inch in width. Wollaston noticed that the spectrum thus formed, of light "purified" (as he stated) by the abolition of overlapping images, was traversed by seven dark lines. These Wollaston took to be the natural boundaries of the various colors. Satisfied with this quasi-explanation, he allowed the subject to drop. The subject was independently taken up in 1814 by the great Munich optician Fraunhofer. In the course of experiments on light, directed towards the perfecting of his achromatic lenses, Fraunhofer, by means of a slit and a telescope, made the surprising discovery that the solar spectrum is crossed not by seven lines but by thousands of obscure streaks. He counted some six hundred and carefully mapped over three hundred of them. Nor did Fraunhofer stop there. He applied the same system of examination to other stars; and he found that the spectra of these stars, while they differ in details from that of the sun, are similar to it in that they are also traversed by dark lines. The explanation of these dark lines of Fraunhofer was sought widely and earnestly. But convincing evidence as 15
to their true nature came only in the fall of 1859 when the great German physicist Kirchhoff formulated his laws of radiation. His laws in this context consist of two parts. The first part states that each substance emits radiations characteristic of itself and only of itself. And the second part states that if radiation from a higher temperature traverses a gas at a lower temperature, glowing with its own characteristic radiations, then in the light which is transmitted the characteristic radiations of the glowing gas will appear as dark lines in a bright background. It is clear that in these two propositions we have the basis for a chemical analysis of the atmospheres of the sun and the stars. By comparisons with the spectral emissions produced by terrestrial substances, Kirchhoff was able to identify the presence of sodium, iron, magnesium, calcium, and a host of other elements in the atmosphere of the sun. The question which had been considered as meaningless only a few years earlier had acquired meaning. The modern age of astrophysics began with Kirchhoff and continues to the present. And we all know that one of the major contributions to our understanding of the spectra of stars and the physics of stellar atmospheres was made in our own times by Meghnad Saha. Now I come to a question that man has always put to Nature: Was there a natural beginning to the universe around us? Or to put the question more directly: How did it all begin? All religions and all philosophical systems have felt the need and the urge to answer this question. Indeed, one may say that a theory of the universe, a theory of cosmology, underlies all religions and all myths. And one of the earliest cosmologies, formulated as such, occurs in the Babylonian epic Enuma Elish in the second millenium B.C. The poem opens with a description of the universe as 16
it was in the beginning. When a sky above had not been mentioned And the name of firm ground below had not been thought of; When only primeval Apsu, their begetter, And Mummu and Ti'amat—she who gave birth to them allWere mingling their waters in one; When no bog had formed and no island could be found; When no god whosoever had appeared, Had been named by name, had been determined as to his lot, Then were gods formed within them. Whether the question of the origin of the universe can be answered on rational scientific grounds is not clear. It might be simplest to suppose that in all aspects the astronomical universe has always been. Or, alternatively, following Compte we might even say that it is in the nature of things that we shall never know how or when the universe began. Nevertheless, recent discoveries in astronomy have enabled us for the first time to contemplate rationally the question: Was there a natural beginning to the present order of the astronomical universe? A related question is: If the astronomical universe did have a beginning, then are we entitled to suppose that the laws of Nature have remained unchanged? The two questions are clearly related. Let me take the second question first. Have the laws of Nature remained the same ? Can the universality of Nature's laws implied by Newton in his formulation of the laws of gravitation, be extended to all time in a changing universe? It is clear that over limited periods of time the laws of Nature can be assumed not to have changed. After all, the motions of planets have been followed accurately over the 17
past three centuries—and less accurately over all historical times—and all we know about planetary motions has been accounted for with great precision with the same Newtonian laws and with the same value for the constant of gravitation. Moreover, the physical properties of the Milky Way system have been studied over most of its extent—and its extent is 30,000 light years. It can be asserted that the laws of atomic physics have not changed measurably during a period of this extent. And on the earth geological strata have been dated for times which go back several hundreds of millions of years. In particular the dating of these strata by the radioactive content of the minerals they contain assume that the laws of physics have not changed over these long periods. But if during these times the astronomical universe in its broad aspects has not changed appreciably, then the assumption that the laws have not changed appreciably during these same periods would appear to be a natural one. The questions that I have formulated, to have a meaning, must be predicated on the supposition that there is a time scale on which the universe is changing its aspect. And if such a time scale exists, the first question is: What is it? That a time scale characteristic of the universe at large exists was first suggested by the discoveries of Hubble in the early twenties. There are two parts to Hubble's discoveries. The first part relates to what may be considered as the fundamental units or constituents of the universe. It emerged unequivocally from Hubble's studies that the fundamental units are the galaxies of which our own Milky Way system is not an untypical one. Galaxies occur in a wide variety of shapes and forms. The majority exhibit extraordinary organization and pattern. To fix ideas, let me say that a galaxy contains some ten thousand million or more stars; its dimension can be measured 18
in thousands of light years: our own galaxy has a radius of 30,000 light years. Further, the distance between galaxies is about 50 to 100 times their dimensions. The second part to Hubble's discovery is that beyond the immediate neighbourhood of our own Milky Way system, the galaxies appear to be receding from us with a velocity increasing linearly with the distance. In other words, all the galaxies appear to be running away from us as though, as Eddington once said, "we were the plague spot of the universe." Hubble's law that galaxies recede from us with a velocity proportional to the distance was deduced from an examination of their spectra. Now suppose that we take Hubble's law literally. Then it follows that a galaxy which is twice as far as another will be receding with a velocity twice that of the nearer one. Accordingly, if we could extrapolate backwards, then both galaxies would have been on top of us at a past epoch. More generally, we may conclude that if Hubble's relation is a strict mathematical one, then all the galaxies constituting the astronomical universe should have been together at a common point at a past calculable epoch. Whether or not we are willing to extrapolate Hubble's law backward in this literal fashion, it is clear that the past epoch calculated in the manner I have indicated does provide a scale of time in which the universe must have changed substantially. Current analysis of the observations suggests that the scale of time so deduced is about seventy thousand million years. With the time scale established, the questions I stated earlier can be rephrased as follows: Have the laws of Nature been constant over periods as long as say thirty or forty thousand million years? And, what indeed was the universe like seventy thousand million years ago? These questions cannot be answered without some underlying theory. While there are 19
several competing theories that are presently being considered, I shall base my remarks on the framework provided by Einstein's general theory of relativity. This theory appears to me the most reasonable. This is clearly not the occasion to digress at this point and describe the content of the theory of relativity. Suffice it to say that it is a natural generalization of Newton's theory and a more comprehensive one. On Einstein's theory, applied to the astronomical universe in the large, it follows that at each instant the universe can be described by a scale of distance which we may call the radius of the universe. At a given epoch it measures the furthest distance from which a light signal can reach us. This radius varies with the time. Its currently estimated value is ten thousand million light years. But the most important consequence that follows from the theory is that this radius of the universe was zero at a certain calculable past epoch some seventy thousand million years ago. In other words, the conclusion arrived at by a naive extrapolation backwards of Hubble's law, interpreted literally, is indeed a valid one. That the theory predicts such a singular origin for the universe is surprising; but it has been established rigorously, with great generality, by a young English mathematician, Roger Penrose. And finally, it is an exact consequence of the theory that the ratio of the wavelengths of an identified line in the light of a distant galaxy to the wavelength of the same source as measured here and now is the same as the ratio of the radius of the universe now and as it was when the light was emitted by the galaxy. During the past few years a dozen or more objects have been discovered for which the ratio of the wavelengths I mentioned is about three. Precisely what has been found is the following. In a laboratory source hydrogen emits a line 20
with a wavelength that is about a third of the wavelength of the visible extreme violet light. But this same line emitted by the stellar object in the remote past and arriving here on earth now is actually observed in visible light. The fact that all the identifiable spectral lines in these objects are shifted by the same factor of about three means that the radius of the universe at the time light left these objects was three times smaller and the density was some twenty-seven times greater than they are now. And a careful analysis of the spectrum shows that during this span of time at any rate the laws of atomic physics have not changed to any measurable extent. To have been able to see back in time when the density of the universe was thirty times what it is now is, of course, a considerable advance. But even this ratio is very far from what it would have been if we take the relativistic picture and go further back in time when the radius of the universe was say ten thousand million times smaller, not merely three times or a thousand times smaller. Does it appear that this extrapolation is meaningless and fanciful? But the general theory of relativity gives a theoretical meaning to such a question since a state of affairs attained by such extrapolation is predicted as an initial state for our present universe. In other words, the question is meaningful, and one can reasonably ask: Is there anything we can observe now that can be considered as the residue or the remnant of that initial singular past? But to answer this question we must take the relativistic picture seriously and determine what it has to say about that remote past. Such a determination has been made by Robert Dicke and his associates at Princeton. Dicke calculated that at the time the radius of the universe was 1010 times smaller, the temperature should have been some ten thousand million degrees—in other words a veritable fireball. And as the universe expanded, radiation of this very 21
high temperature, which would have filled the universe at that time, would be reduced. For example, its temperature would have fallen to ten thousand degrees after the first ten million years. As the universe continued to expand beyond this point the radiation will cool adiabatically, i.e. in the same manner as gas in a chamber will cool if it is suddenly expanded. And Dicke concludes that the radiation from the original fireball must now fill the universe uniformly, but that its temperature must be very low—in fact, 3° Kelvin, a temperature that is attainable in the laboratory only by liquefying helium. It corresponds to radiation at a temperature of 270° of frost! How can we detect this low temperature radiation? It can be shown that this radiation at 270° of frost should have its maximum observable intensity at wavelengths in the neighbourhood of 3 millimeters, i.e. the radiation must be present in the microwave region. The remarkable fact is that radiation in these wavelengths has been detected; it comes with incredible uniformity from all directions; and they have all the properties that one might, on theoretical grounds, want to attribute to such fossil radiation from the original fireball. With these discoveries I have described, astronomy appears to have justified the curiosity that man has felt about the origin of the universe, from the beginning of time. As I said at the outset, man's contemplation of the astronomical universe has provided us with the one continuous thread that connects us with antiquity. And might I add now that it has also inspired in him the best.
22
COMMUNICATIONS ON PURE AND APPLIED MATHEMATICS, VOL. XX, 2 5 1 - 2 6 5 (1967)
Ellipsoidal Figures of Equilibrium— An Historical Account* S. C H A N D R A S E K H A R University of Chicago
1. N e w t o n T h e study of the gravitational equilibrium of homogeneous uniformly rotating masses began with Newton's investigation on the figure of the earth (Principia, Book I I I , Propositions X V I I I - X X ) . Newton showed that the effect of a small rotation on the figure must be in the direction of making it slightly oblate; and, further, that the equilibrium of the body will demand a simple proportionality between the effect of rotation, as measured by the ellipticity, equatorial radius — polar radius the mean radius and its cause, as measured by centrifugal acceleration at the equator mean gravitational acceleration on the surface (2) _ WR _ D.2R3 2 ~ GM/R ~~GM ' where G denotes the constant of gravitation a n d M is the mass of the body. More precisely, Newton established the relation (3)
e = \m
in case the body is homogeneous. T h e arguments by which Newton derived this relation are magisterial; and they are worth recalling. Newton imagined a hole of unit cross-section drilled from a point on the equator to the center of the earth and a similar hole drilled from the pole to the center; and he further imagined that the "canals" so constructed were filled with * This paper is a somewhat expanded version of the lecture given at the Courant Institute of Mathematical Sciences on the occasion of the Conference to dedicate Warren Weaver Hall in March, 1966. Reproduction in whole or in part is permitted for any purpose of the United States Government. 251
252
S. CHANDRASEKHAR
a fluid (see Figure 1, after Newton's original illustration in the Principia). From the fact that the fluid in the canals will be in equilibrium, Newton concludes that the "weights" of the equatorial a n d the polar columns of the fluid must be equal. However, along the equator the acceleration due to gravity is "diluted" by the centrifugal acceleration; and since both these accelerations in a homogeneous body vary from the center proportionately with the distance, the "dilution factor" remains constant a n d is given by its value at the boundary, namely m.
Figure 1
If a denotes the equatorial radius, the weight of the equatorial column is given by (4)
weight of equatorial column = \agevi&t0T(\
— m) ,
where gequator i s t r i e acceleration due to gravity at the equator. Similarly, if b denotes the polar radius, (5)
weight of polar column = %bgpole .
And since the two weights must be equal, ^equator!1 -
(6)
m) = bgvole
.
But for a slightly oblate body Newton knew that
(7)
-i*2!l_
=
i +
ie
+ 0 ( e ") .
^equator
Equations (6) a n d (7) and the definition of e ( = 1 — bja) now give (8)
1 - m = (1 -
e)(l + \e) + 0 ( e 2 ) = 1 - fe + 0(e») ;
and Newton's relation (3) follows. I t was known already in Newton's time that (9)
m = zio .
ELLIPSOIDAL FIGURES OF EQUILIBRIUM
AN HISTORICAL ACCOUNT
253
Therefore, Newton concluded that if the earth were homogeneous, it should be oblate with an ellipticity (10)
€ = i T^To ~
Tr^o .
This prediction of Newton was contrary to the astronomical evidence of the time and "two generations of the best astronomical observers formed in the school of the Cassini's struggled in vain against the authority and reasoning of N e w t o n " (I. T o d h u n t e r [1], page 100). T h e opposing ideas of Newton and Cassini are strikingly illustrated in the accompanying old caricature (Figure 2). However,
o Pole
N£WTOM
Pole
0
CASSINI Pole
Figure 2
geodetic measurements m a d e in Lapland by Maupertius and Clairaut (1738) afforded data which conclusively showed the flattening of the earth at the poles. As T o d h u n t e r has written ([1], page 100), " T h e success of the arctic expedition may be ascribed in great measure to the skill and energy of Maupertius; and his fame was widely celebrated. T h e engravings of the period represent him in the costume of a Lapland Hercules having a fur cap over his eyes; with one hand he holds a club and with the other he compresses the terrestrial globe." And Voltaire, then Maupertius' friend, congratulated him warmly for having "aplati les poles et les Cassini." Later Maupertius a n d Voltaire became involved in a heroic-comic controversy and Voltaire wrote "Vouz avez confirme dans les lieux pleins d'ennui Ce que Newton connut sans sortir de chez lui." W e know now that the actual ellipticity of the earth (.—^z) is substantially smaller than Newton's predicted value (~2To) j a n d this discrepancy is interpreted in terms of the inhomogeneity of the earth.
2. M a c l a u r i n . T h e next advance (1742) in the theory was due to Maclaurin who generalized Newton's result to the case when the ellipticity caused by the rotation cannot be considered small.
254
S. CHANDRASEKHAR
Maclaurin had solved earlier the problem of the attraction of an oblate spheroid at an internal point; and he had shown in particular that the acceleration due to gravity at the equator and the poles have the values
^equator = 2-TrGpa
-e2
VI
[ s i n " 1 e — eV I — e2 1 ,
fll) £poie = 4irGpa
VT 3
«' [e — V l — e2 sin" 1 e\ ,
e
where p is the density of the spheroid, a its semi-major axis, and e its eccentricity. And since both the centrifugal acceleration in the equatorial plane and the acceleration due to gravity vary linearly with the coordinates, Newton's argument applies to this case equally well and we can write (^equator ~ aQ.2) = £ p 0 l e V l - e2 , . e Q2 ^equator ffpole * * a
(12)
Inserting the expressions for Equator Maclaurin's formula 13
( )
Q2 -FT= •nGp
an
d £poie from equations (11), we obtain
V l - e2 6 g—2(3-2«»)sin-i«--(l e3
el
-e2).
Maclaurin realized that the foregoing derivation does not establish that a rapidly rotating mass will necessarily take the figure of an oblate spheroid. But he did show: "1) that the force which results from the attraction of the spheroid and those extraneous powers compounded together acts always in a right line perpendicular to the surface of the spheroid, 2) that the columns of the fluid sustain or balance each other at the center of the spheroid, and 3) that any particle in the spheroid is impelled equally in all directions." T o appreciate the foregoing qualifications of Maclaurin, one must remember that there was as yet no theory of hydrostatic equilibrium which provided sufficient conditions; so Maclaurin had to content himself with showing that all the conditions which had been recognized as necessary for equilibrium were satisfied. Considering then, the state of knowledge in his time, one can only admire Maclaurin's achievement in deriving the exact relation (13). And as Todhunter remarks ([1], page 175), "Maclaurin well deserves the association of his n a m e with that of the great master in the inscription which records that he was appointed professor of mathematics at Edinburgh ipso Newtono suadente." A remarkable feature of Maclaurin's relation was noticed by Thomas Simpson (1743): for any angular velocity less than a certain maximum value there
ELLIPSOIDAL FIGURES OF EQUILIBRIUM
AN HISTORICAL ACCOUNT
255
are two and only two possible " o b l a t a " . This result is noteworthy in that we cannot deduce from the fact of a small equatorial angular velocity that the spheroid departs only slightly from a sphere; for as Q? —*• 0, we have two solutions: a solution which, indeed, leads to a spheroid of small eccentricity and a second solution which leads to a highly flattened spheroid. It is generally believed that d'Alembert was the first to notice this feature of Maclaurin's solution; but as Todhunter has remarked ([1], page 181): "although d'Alembert may have first explicitly published the statement, yet Simpson gives a table which distinctly implies the fact." 3. J a c o b i For nearly a century after Maclaurin's discovery of the spheroids (known after his name) it was believed that they represent the only admissible solution to the problem of the equilibrium of uniformly rotating homogeneous masses. T h e supposed generality of Maclaurin's solution was never questioned even though Lagrange in his Mecanique Celeste (1811) considered formally the possibility of ellipsoids with unequal axes satisfying the requirements of equilibrium. However, after obtaining two governing equations, in which the two equatorial axes occur symmetrically (see equation (17) below), Lagrange infers that the two axes must be equal even though only the sufficiency (not the necessity) could be concluded. Jacobi (1834) [3] recognized the inadequacy of Lagrange's demonstration 1 as he remarked, " O n e would make a grave mistake if one supposed that the spheroids of revolution are the only admissible figures of equilibrium even under the restrictive assumption of second degree surfaces." In making this last statement, Jacobi refers to the fact that while Maclaurin's solution provides, in the limit Q 2 —*• 0, two solutions, one with e —>• 0 and another with e —>• 1, Legendre h a d shown that if one supposes that the figure is nearly spherical so that the attraction at a point on its surface can be expanded in powers of the departure from sphericity, then one obtains only the first of the two solutions " n o t in any approximation but with absolute geometrical rigor". According to Jacobi, the conclusion one must draw from Legendre's demonstration is that figures of equilibrium may exist that cannot be surmised from what one can establish in the limit of spherical figures. And Jacobi concludes "in fact a simple consideration shows that ellipsoids with three unequal axes can very well be figures of equilibrium; and that one can assume a n ellipse of arbitrary shape for the equatorial section and determine the third axis (which is also the least of the three axes) and the angular velocity of rotation such that the ellipsoid is a figure of equilibrium." T h e existence of these ellipsoids of Jacobi can be established and the relations governing them can be determined by a simple extension of Newton's original argument. 1 Rather as Dirichlet [4] states in his Gedachtnissrede auf Carl Gustav Jacob Jacobi, Jacobi's suspicion was aroused by the qualification "necessary" in an account of Lagrange's considerations by the author of a "well-known textbook".
256
S.
CHANDRASEKHAR
At the time Jacobi m a d e his discovery, it was known that the components of the attraction, gi, i = 1, 2, 3, along the directions of the principal axes of an ellipsoid can be expressed in the manner (14)
gi
= -2nGpAiXt
,
where f°°
(15)
At = aia2a3
du ———— , Jo (a? + u)/\
and A 2 = (al + u)(al + u)(al + u) .
(16)
T h e formulae for the components of the attraction in the foregoing forms were apparently first derived by Gauss (1813) and by Rodrigues (1815), independently. However, in the less symmetrical forms in which one generally writes them for purposes of reducing them to the standard elliptic integrals of the two kinds, they were known much earlier: they are (as Legendre has said) effectively included in Maclaurin's writings; but explicitly, for ellipsoids with three unequal axes, they occur for the first time 2 in Laplace's Theorie du Mouvement et de la Figure Elliptique des Planetes (1784). Returning to the extension of Newton's argument to the case of tri-axial ellipsoids, we may imagine that three "canals" are drilled along the directions of the three principal axes from the surface to the center and further that they are all filled with a fluid. From the equilibrium of the fluid in the three canals, we m a y infer the equality of the weights of the three columns (per unit cross-section). W e thus have (17)
2AX
- ^
D2
al = 2A2a$
iGp
"*> == '>At <
TTLrp
These relations require (if a1 ^ a2 ^ a3) O2 (18) {
'
TTGP
A,a* - A9ai f °° = 2 - ^ — = 2fl,a,0 3 a*-a*
1 2
Mo
u du
{al + u){a\
+ u)
And we also have the purely geometrical condition al a\ 2 ([1], page 417), (19) As Todhunter has pointed out A1--\A = the formulae Aa--{At, themselves appear in the writing a
of d'Alembert though "he deliberately rejects them . . . this is perhaps the strangest of all his (d'Alembert's) strange mistakes." And with regard to Laplace's derivation, Todhunter says ([2], page 32) "thus Laplace values and appropriates the treasure which d'Alembert deliberately threw away."
ELLIPSOIDAL FIGURES OF EQUILIBRIUM
AN HISTORICAL ACCOUNT
257
«Ma f i-4a1 = 4V
(20)
t - \
This last relation explicitly has the form (21)
du 2A
3
J o (fli + B)(a* («» ++ «) A «)A
H (^ + «)A Jo
Equations (18) and (21), in exactly these forms, are given in Jacobi's paper. And as Jacobi further states, for any assigned ax and a2 , equation (21) allows a solution for a3 which satisfies the inequality
(22)
1
1
1
1 > -fl 2 + -, > fl 3
l
«2
and that when ax = fl2 equations (18) and (21) determine a configuration common to the spheroidal and the ellipsoidal sequences. Referring to this discovery of Jacobi, Thomson, a n d T a i t in their Natural Philosophy ([5], Volume I I , page 530) say "this curious theorem was discovered by Jacobi in 1834 and seems, simple as it is, to have been enunciated by him as a challenge to the French Mathematicians." In Todhunter's "History" there is no reference to Jacobi having issued a "challenge". But T o d h u n t e r ([2], page 381) does refer to a communication by Poisson to the French Academy on November 24, 1834 and states "Poisson begins by referring to a letter recently sent by Jacobi to the French Academy in which two results were enunciated. O n e was what we call Jacobi's theorem, namely, that an ellipsoid is a possible form of relative equilibrium for a rotating fluid; the other related to the attraction of a heterogeneous ellipsoid • • • Poisson's note related to the second result."
4. M e y e r a n d L i o u v i l l e In his short and brief paper on the subject, Jacobi did not seriously examine the relationship of his ellipsoids to the Maclaurin spheroids. C. O . Meyer (1842) was the first to do so. Meyer's principal result was to show that the Jacobian sequence "bifurcates" (in the later terminology of Poincare) from the Maclaurin sequence a t the point where the eccentricity e = 0.81267. This result can be readily deduced from Jacobi's equations (18) and (21). Thus, by setting a1 = a2 in these equations we obtain the relations
258
S. CHANDRASEKHAR
and (24)
4f
H
CO
du
_ af"
{a\ + u)W~^+~u ~ H
du K + « ) K + H)3/2'
where Q.2lnGp on the left-hand side of equation (23) must now be identified with Maclaurin's function (13). I t can be shown that both equations (23) a n d (24) are simultaneously satisfied when e = 0.81267,
where Q^jirGp = 0.37423 .
Since it is known that the maximum value of QPjirGp along the Maclaurin sequence is 0.4493, it follows thatfor QP/nGp < 0.37423 there are three equilibrium figures possible: two Maclaurin spheroids a n d one Jacobi ellipsoid; for 0.4493 > Q?/TTGP > 0.3743 only the Maclaurin figures are possible; and finally, for Q2/TTGP > 0.4493 no equilibrium figures are possible. This enumeration of the different possibilities is due to Meyer. I n 1846 Liouville restated Meyer's result using the angular momentum, instead of the angular velocity, as the variable; and he showed that while the angular m o m e n t u m increases from zero to infinity along the Maclaurin sequence, the Jacobian figures are possible only for angular momenta exceeding a certain value (namely, that at the point of bifurcation along the Maclaurin sequence).
5. D i r i c h l e t , D e d e k i n d , a n d R i e m a n n T h e fact that no figures of equilibrium are possible for uniformly rotating bodies when the angular velocity exceeds a certain limit raises the question: W h a t happens when the angular velocity exceeds this limit ? Dirichlet addressed himself to this question during the winter of 1856-57; and though he included this topic in his lectures on partial differential equations in J u l y 1857, he did not publish any detailed account of his investigations during his lifetime. Dirichlet's results were collated from some papers he left a n d were edited for publication by Dedekind [6]. " I n this p a p e r , " as R i e m a n n wrote, "Dirichlet opened u p an entirely new approach to investigations bearing on the homologous motions of self-gravitating ellipsoids in a most remarkable way. T h e further development of this beautiful discovery has a particular interest to the mathematician even apart from its relevance to the forms of heavenly bodies which initially instigated these investigations." T h e precise problem which Dirichlet considered in his paper is the following: U n d e r what conditions can one have a configuration which, at every instant, has an ellipsoidal figure a n d in which the motion, in an inertial frame, is a linear function of the coordinates ? Dirichlet formulated the general equations governing this problem (in a Lagrangian framework) and solved them in detail for the case when the bounding surface is a spheroid of revolution. Dirichlet did not seriously investigate the figures of equilibrium admissible under the general circumstances
ELLIPSOIDAL FIGURES OF EQUILIBRIUM—AN HISTORICAL ACCOUNT
259
of his formulation. I n the latter context, Dedekind ([7] in an a d d e n d u m to Dirichlet's paper) proved explicitly the following theorem (though, as R i e m a n n remarks, it is already implicit in Dirichlet's equations): Let a homogeneous ellipsoid with semi-axes ax , a2 , and as be in gravitational equilibrium with a prevalent motion whose components, resolved along the instantaneous directions of the principal axes of the ellipsoid and in an inertial frame, are given by
(25)
,(0)
"•31
12
«13
22
a
32
«33
*iK
*i/«i
#2/^2
23
= A
*2/a2
x
zlaz
x
3la3
then the same ellipsoid will also be a figure of equilibrium if the prevalent motion is that derived from the transposed matrix A*, i.e. u<0)1' given by
(26)
u(0)t
=
31
x1ja1
32
#2/^2
33
*s/fls
x a
il i
= At xja2 x3\a3
W e shall call the configuration with the motion derived from A1^ as the adjoint of the configuration with the motion derived from A. Dedekind considered in particular the configurations which are congruent to the Jacobi ellipsoids and are their adjoints in the sense we have defined. T h e motion of a Jacobi ellipsoid rotating uniformly with a n angular velocity Q about the * 3 -axis can be represented in the m a n n e r
(27)
,(0)
0
-Qa2
0
x1ja1
Qa x
0
0
xja2
0
0
0
x
3la3
T h e motion in the adjoint configuration will be given by 0
(28)
u«»t =
-Qa2 0
D.ai
0
x1/a1
0
0
X
0
0
xja3
2I&2
or, in terms of components, (29)
< >it' = — * x2, a,
uf
= - —-2 Xl,
u3»r = 0 ;
and this motion clearly satisfies the condition (30)
tt
.g r a d (f|
+
f|
+
f|)
=
o
260
S. CHANDRASEKHAR
required for the preservation of the ellipsoidal boundary. represented by (29) is one of uniform vorticity
(31)
£
„_
Q
(5
+
Also, the motion
^«_Q1±^.
These ellipsoids of Dedekind, while they are congruent to the Jacobi ellipsoids, are stationary in a n inertial frame and they maintain their ellipsoidal figures by the internal motions which prevail. (Lamb erroneously attributes to Love the discovery of this relation between the Jacobi and the Dedekind ellipsoids.) It is also clear that the ellipsoids of Dedekind bifurcate from the Maclaurin spheroids at the same point that the Jacobi ellipsoids do. T h e complete solution to the problem of the stationary figures admissible under Dirichlet's general assumptions was given by R i e m a n n [8] in a paper of remarkable insight and power. Riemann first shows that under the restriction of motions which are linear in the coordinates, the most general type of motion compatible with an ellipsoidal figure of equilibrium consists of a superposition of a uniform rotation SI and internal motions of a uniform vorticity X, (in the rotating frame). More precisely he showed that ellipsoidal figures of equilibrium are possible only under the following three circumstances: (a) the case of uniform rotation with no internal motions, (b) the case when the directions of Si. and £ coincide with a principal axis of the ellipsoids, and (c) the case when the directions of Si a n d JJ lie in a principal plane of the ellipsoid. Case (a) leads to the sequences of Maclaurin and Jacobi. Case (b) leads to sequences of ellipsoids along which the r a t i o / = £/Q remains constant (the Jacobian and the Dedekind sequences are special cases of these general " R i e m a n n sequences" f o r / = 0 a n d oo, respectively). A n d finally, case (c) leads to three other classes of ellipsoids. R i e m a n n wrote down the equations governing the equilibrium of these ellipsoids and specified their domain of occupancy in the a1} a2, <23-space. (A more detailed description of the properties of these ellipsoids will be found in the Epilogue.) R i e m a n n also sought to determine the stability of these ellipsoids by an energy criterion. But his criterion, as has recently been shown by Lebovitz [9], is erroneous and Riemann's conclusions, with the notable exception of those pertaining to the Maclaurin and the Riemann sequences for/" ^ — 2, are false. While Riemann's paper made an impressive start towards the solution of Dirichlet's general problem, it left a large number of questions unanswered. Indeed, even the relation of Riemann's ellipsoids to the Maclaurin spheroids which they adjoin was left obscure. Nevertheless these questions were to remain unanswered for more than a hundred years. T h e reason for this total neglect must, in part, be attributed to a spectacular discovery by Poincare (see Section 6 below) which channeled all subsequent investigations along directions which appeared rich with possibilities; but the long quest it entailed turned out, in the end, to be after a chimera.
ELLIPSOIDAL FIGURES OF EQUILIBRIUM
AN HISTORICAL ACCOUNT
261
6. P o i n c a r e a n d C a r t a n T h e investigations relating to the equilibrium and the stability of ellipsoidal figures of equilibrium, for which Dirichlet and Riemann had laid such firm foundations, took an unexpected turn (from which it was not to be diverted for the next seventy-five years) when Poincare [10] discovered in 1885 that along the Jacobian sequence a point of bifurcation occurs similar to the one along the Maclaurin sequence and that even as the Jacobian sequence branches off from the Maclaurin sequence, a new sequence of pear-shaped configurations branches off from the Jacobian sequence. This result of Poincare is equivalent to the statement (in current terminology) that along the Jacobian sequence there is a point where the ellipsoid allows a neutral mode of oscillation belonging to the third harmonics. A corollary which was also enunciated by Poincare is that along the Jacobian sequence there must be further points of bifurcation where the Jacobian ellipsoid allows a neutral mode of oscillation belonging to the fourth, fifth, and higher harmonics. And Poincare conjectured " t h a t the bifurcation of the pear-shaped body leads onward stably and continuously to a planet attended by a satellite, the bifurcation into the fourth zonal harmonic probably leads unstably to a planet with a satellite on each side, that into the fifth harmonics to a planet with two satellites on one and one on the other and so o n " (Darwin). It was further conjectured by Darwin that one may look for the origin of the double stars in similar instabilities; the "fission theory" of the origin of double stars arose in this fashion. T h e grand mental p a n o r a m a that was thus created was so intoxicating that those who followed Poincare were not to recover from its pursuit. I n any event, Darwin, Liapounoff, and Jeans spent years of effort towards the substantiation of these conjectures; and so single minded was the pursuit 3 that one did not even linger to investigate the stability of the Maclaurin spheroids and the Jacobi ellipsoids from a direct analysis of normal modes. Finally, in 1924 Cartan [11], [12] established that the Jacobi ellipsoid becomes unstable at its first point of bifurcation and behaves in this respect differently from the Maclaurin spheroid which, in the absence of any dissipative mechanism, is stable on either side of the point of bifurcation where the Jacobian sequence branches off. And at this point the subject quietly went into a coma.
Epilogue T h e subject of the allowed ellipsoidal figures of equilibrium of homogeneous masses a n d their stability, left incomplete by Riemann, has now been completed 3 For example, the question whether along the Dedekind sequence a neutral point occurs similar to the one along the congruent Jacobian sequence does not appear to have been considered or even raised.
S. CHANDRASEKHAR
262
(Chandrasekhar [13], [14], and [15]). This is clearly not the place to attempt even an outline of the methods by which the various aspects of the problem have been solved (however, for a brief account, see Chandrasekhar and Lebovitz [16]); we shall content ourselves by simply exhibiting the results of the analysis. Figure 3 depicts the domains of occupancy of the ellipsoidal figures in the a^\ax, a3/a1-plane, and Figures 4 and 5 present results of the stability analysis.
Figure 3.
The domain of occupancy of the Riemann ellipsoids in the aj^ as/deplane.
,
The stable part of the Maclaurin sequence is represented by the segment 02S on the line ax = a2 . At 0 2 the Maclaurin spheroid becomes unstable by overstable oscillations. The Riemann ellipsoids of type S (for which the directions of rotation and vorticity coincide with the x3-axis) are included between the selfadjoint sequences represented by SO and 020. Along the arc Xis>0 the Riemann ellipsoids of type S become unstable by an odd mode of oscillation belonging to the second harmonics. The Riemann ellipsoids, in which the directions of rotation and vorticity do not coincide but lie in the a a , a 3 -plane, are of three types—I, II, and III—with the domains of occupancy shown. Type I ellipsoids adjoin the Maclaurin sequence and are bounded on one side (SRj) by a selfadjoint sequence. Along the locus RiRu , which limits the domain of occupancy of the type II ellipsoids, the pressure is
ELLIPSOIDAL FIGURES OF EQUILIBRIUM
AN HISTORICAL ACCOUNT
zero. And along the loci X%0 and X% 0", limiting the domain of occupancy of the type III ellipsoids, the directions of £2 and % coincide with one of the principal axes (the a3-axis in the case a2 > ax and the a2-axis in the case a2 < a^). The locus X'zO' (for the case a2 > ax) is transformed into X^s)0 if the roles of ax and a2 are interchanged; and simultaneously the domain of occupancy A X^O similarly becomes transformed into the domain AXls>0. The dotted curve X^lll)0' defines the locus of configurations, among the type III ellipsoids, that are marginally overstable by a mode of oscillation belonging to the second harmonics.
Figure 4. The Riemann ellipsoids of type S (for which the directions of rotation and vorticity coincide with the x3-axis) can be arranged in sequences along which f = £/Q is a constant. The stable part of the Maclaurin sequence is represented by the segment 0 2 S of the line a2 = 1. At 0 2 the Maclaurin spheroid becomes unstable by overstable oscillations and at Af2 the Jacobian and the Dedekind sequences bifurcate (labeled by"0, ±oo"). The different Riemann sequences are labeled by the values of/ to which they belong; these sequences are bounded by the two selfadjoint sequences (the dotted curves labeled x = — 1 and x = +1) along w h i c h / = / t = T(fli + a\)laiai • The sequences belonging t o / i n the range —2 ^ / ^ + 2 form a nonintersecting
263
1151
264
S. CHANDRASEKHAR family of continuous curves which join points on the line 0%S to the origin. The sequences belonging t o / < —2 a n d / > + 2 are represented by curves which consist of two parts: a part which joins a point on the line SM2 (or M%0^) to a point of the selfadjoint sequence for x = — 1 (or x = + 1 ) and a part which joins the point on the selfadjoint sequence to the origin. Along the selfadjoint sequence x = — 1, instability by a mode of oscillation belonging to the second harmonics sets in at the point indicated by Xi and the locus of points at which instability by this mode sets in is the curve which joins X2 to the origin. The curve labeled AND is the locus of neutral points, belonging to the third harmonics, along the Riemann sequences for —2 ^f^ + 2 ; and the curve labeled BN^C is the corresponding locus for configurations adjoint to the Riemann ellipsoids represented in the domain included between the same sequences f— —2 a n d y = + 2 . The continuations of the loci AND and BN^C into the domains included between the sequences x = — 1 a n d / = —2 (and, similarly, between the sequences x = + 1 and_/" = + 2 ) are represented by curves (not shown) joining the points A and B to Aj — ' on the sequence x = — 1 (and, similarly, by curves joining the points D and C to the point Xl3 + ' on the sequence x = + 1); X^' and X{3+' are the neutral points, belonging to the third harmonics, along the selfadjoint sequences x = — I and x = + 1 , respectively.
Figure 5.
The loci of marginally stable configurations in the a2\a^,
aja^plane.
The type S ellipsoids are bounded by two selfadjoint sequences {SO and 020) and the stable part of the Maclaurin sequence represented by S0a . Along the arc X^O the type 5 ellipsoids become unstable by a mode of oscillation belonging to the second harmonics; and along this same arc the stability passes to the type III ellipsoids whose domain of occupancy is AX^'O. The shaded region included between X^u)0 and X^s)0 represents the domain of stability for type III ellipsoids with respect to oscillations belonging to the second harmonics. The type I ellipsoids occupy the triangle SMcR ; and the region of the stable members is included in the two domains marked "stable". The domain SO%X™ of stable ellipsoids adjoining the stable Maclaurin spheroids is to be expected, but the domain D OR including disklike ellipsoids along D R is unexpected. All type II ellipsoids are unstable.
ELLIPSOIDAL FIGURES OF EQUILIBRIUM
AN HISTORICAL ACCOUNT
265
Bibliography [1] Todhunter, I., History of the Mathematical Theories of Attraction and the Figure of the Earth, Vol. I, Constable & Company, London, 1873; reprint edition, Dover Publications, Inc. New York, 1962. [2] Todhunter, I., History of the Mathematical Theories of Attraction and the Figure of the Earth, Vol. II, Constable & Company, London, 1873; reprint edition, Dover Publications, Inc. New York, 1962. [3] Jacobi, C. G. J., Vber dieFigur des Gleichgewichts, Poggendorff Annallen der Physik und Chemie, Vol. 33, 1834, pp. 229-238; reprinted in Gesammelte Werke, Vol. 2, pp. 17-72, Verlag Von G. Reimer, Berlin, 1882. [4] Dirichlet, G. Lejeune, Geddchtnissrede auf Carl Gustav Jacob Jacobi gehalten in der Akademie der Wissenschaften am 1 July 1852, Gesammelte Werke, Vol. 2, p. 243, Verlag Von G. Reimer, Berlin, 1897. [5] Thomson, W., and Tait, P. G., Treatise on Natural Philosophy, Part II, Cambridge University Press, Cambridge, 1883, pp. 324-335. [6] Dirichlet, G. L., Untersuchungen uber ein Problem der Hydrodynamik, J. Reine Angew. Math., Vol. 58, 1860, pp. 181-216. [7] Dedekind, R., Zusatz zu der vorstehenden Abhandlung, J. Reine Angew. Math., Vol. 58, 1860, pp. 217-228. [8] Riemann, B. Untersuchungen uber die Bewegung einesfliissigen gleichartigen Ellipsoides, Abh. d. Konigl, Gessel. der Wis. zu Gottingen, Vol. 9, 1860, pp. 3-36. [9] Lebovitz, N. R., On Riemanris criterion for the stability of liquid ellipsoids, Astrophysical Journal, Vol. 145, 1966, pp. 878-885. [10] Poincare, H., Sur Vequilibre d'une massefiuide anime'e d'un mouvement de rotation, Acta Math., Vol. 7, 1885, pp. 259-380. [11] Cartan, H., Sur lespetites oscillations d'une massefiuide,Bull. Sci. Math., Vol. 46,1922, pp. 317-352; 356-369. [12] Cartan, H., Sur la stabilite ordinaire des ellipsoides de Jacobi, Proc. International Math. Congress, Toronto, 1924, Vol. 2, University of Toronto Press, 1928, pp. 2-17. [13] Chandrasekhar, S., The equilibrium and the stability of the Dedekind ellipsoids, Astrophysical Journal, Vol. 141, 1965, pp. 1043-1055. [14] Chandrasekhar, S., The equilibrium and the stability of the Riemann ellipsoids. I., Astrophysical Journal, Vol. 142, 1965, pp. 890-921. [15] Chandrasekhar, S., The equilibrium and the stability of the Riemann ellipsoids. II, Astrophysical Journal, Vol. 145, 1966, pp. 842-877. [16] Chandrasekhar, S., and Lebovitz, N. R., On the ellipsoidalfiguresof equilibrium of homogeneous masses, Astrophysica Norvegica, Vol. 9, 1964, pp. 323-332.
Received November 1966.
/ . Mehra (ed.), The Physicist's Conception of Nature, 34-44. All Rights Reserved Copyright © 1973 by D. Reidel Publishing Company, Dordrecht-Holland
A Chapter in the Astrophysicist's View of the Universe S. Chandrasekhar At the outset, I must state quite frankly that I have neither the knowledge nor the competence to undertake a coherent and a reasoned account of the developments that have led to our current ideas concerning the 'astrophysical universe'. But it is safe to assume that there are many facets to the astrophysical universe, and I shall therefore limit myself to only one facet, namely, the developments that have led to our present ideas concerning gravitational collapse and black holes. Even here I cannot be fair to ail aspects though I can perhaps claim, by virtue at least of longevity, that I have been personally associated with related problems for a longer period than anyone else in this audience.1 In the first instance, one may be surprised that a problem of such current lively interest should have been a matter for serious discussion some fifty years ago; but the question concerning the last stages of stellar evolution occurs to one even if one has no precise knowledge of the physical processes that are responsible for the energy of the stars. Indeed, the question occurs to one almost inevitably: for no matter what the source of energy is, it must be exhausted sooner or later; and sooner or later the question must be confronted. The question was in fact formulated by Eddington in 1926 in one of his famous aphorisms: lA star will need energy to cool.1 Let me rephrase the question in a less oracular fashion. By 1925 it was known through the work of W. Adams on the binary system of Sirius that stellar masses with densities in the range of 105 to 106 grams cm - 3 exist. And Eddington estimated that the central temperatures of such stars should be 107 to 108 degrees, on the assumption that ordinary perfect-gas laws hold and that matter is ionized to bare nuclei and electrons. On these same assumptions let Ev denote the negative electrostatic energy per gram. And let EK denote the kinetic energy of the particles per gram. If such matter were released of the pressure to which it is subject, then it could resume the state of ordinary un-ionized matter only if EK > Ey. Eddington's paradox2 was this (though Eddington at a later time disclaimed to this particular formulation): if the quantities EK and Ev are evaluated for the densities and temperatures expected in the interiors of the white dwarfs, then one finds that ,EK (perfect %&%)<EV. In other words, on these premises, the matter will not be able to resume its ordinary state - an extraordinary situation if true. R. H. Fowler's resolution of this paradox in 1927 in terms of the then very new statistical mechanics of Fermi and Dirac is, in my opinion, one of the great landmarks in the development of our ideas on stellar structure and stellar evolution. May I spend a few minutes on it.
A CHAPTER IN THE ASTROPHYSICIST'S VIEW OF THE UNIVERSE
35
Dirac's paper, which contains the derivation of what has since come to be called the 'Fermi-Dirac distribution', was communicated by Fowler to the Royal Society on 26 August 1926.3 On 3 November Fowler communicated a paper of his own in which the application of the laws of the 'new quantum theory' to the statistical mechanics of assemblies consisting of similar particles is systematically developed and incorporated into the general scheme of the Darwin-Fowler method. And by 10 December (i.e., before a month had elapsed) his paper entitled 'Dense Matter' was read before the Royal Astronomical Society. In this paper Fowler drew attention to the fact that the electron gas in matter as dense as in the companion of Sirius, must be degenerate in the sense of the Fermi-Dirac statistics. Thus, to Fowler belongs the credit for first recognizing afieldof application for the then 'very new' statistics of Fermi and Dirac, though among physicists this credit is generally given to Pauli for his explanation of the paramagnetic susceptibility of the alkali metals. Fowler's resolution of Eddington's paradox is simply this: since at the temperature and densities in the white dwarfs the electron assembly will be highly degenerate, EK should be evaluated by using the formulae appropriate in this limit. He showed that when EK is so evaluated, it is indeed much greater than Ev. And Fowler concluded his paper4 with the following statement. The black-dwarf material is best likened to a single gigantic molecule in its lowest quantum state. On the Fermi-Dirac statistics, its high density can be achieved in one and only one way, in virtue of a correspondingly great energy content. But this energy can no more be expended in radiation than the energy of a normal atom or molecule. The only difference between black-dwarf matter and a normal molecule is that the molecule can exist in a free state while the black-dwarf matter can only so exist under very high external pressure.
The limiting form of the equation of state which Fowler used in his paper is the standard one which is familiar in solid-state physics, namely,
where kt is a constant and ne is the concentration of the electron. The energy (=3/2 pV) associated with this pressure is zero-point energy; and the point of Fowler's paper is that this zero-point energy is so great that under normal circumstances a star could be expected to settle down into a state in which all of its internal energy is in this form. And it appeared for a time that all stars will have the necessary energy to cool. Fowler's arguments can be stated more precisely in the following manner. On the basis of the equation of state, one can readily determine the structure which a configuration of an assigned mass M will assume when in equilibrium under its own gravity. One finds5 that an equilibrium state is possible for any assigned mass: one finds in fact a mass-radius relation of the form M = constant/? -3 . Accordingly, the larger the mass, the smaller is its radius. Also the mean densities of these configurations are found to be in the range of 105 to 106 grams cm - 3 when the mass is of solar magnitude. These masses and densities are of the order one meets in
S. CHANDRASEKHAR
36
the so-called white-dwarf stars. And it seemed for a time that the white-dwarf stage (or rather the 'black-dwarf stage as Fowler described it) represented the last stage of stellar evolution for all stars. Since afinitestate seemed possible for any assigned mass, one could rest with the comfortable assurance that all stars will have the 'necessary energy to cool'. But this assurance was soon broken when it was realized that the electrons in the centres of degenerate masses begin to have momenta comparable to mc where m is the mass of the electron. Accordingly, one must allow for the effects of special relativity. These effects can be readily allowed for and look harmless enough in thefirstinstance: the correct equation of state, while it approximates to that by the equation given earlier for low enough electron concentrations, it tends to
p = M«,) 4/3 as the electron concentration increases indefinitely. (In the foregoing equation, k2 is another atomic constant.) This limiting form of the equation of state has a dramatic effect on the predicted mass-radius relation; instead of predicting afiniteradius for all masses, the theory now predicts that the radius must tend to zero as a certain limiting mass is reached. The value of this limiting mass is Mlimit = 5.76/,; 2 0, where nt denotes the mean molecular weight per electron and O stands for the solar mass. For the expected value fie=2, M„mit = 1.440. These results were obtained in 1930 and were published early in 1931.6-7 The existence of this limiting mass means that a white-dwarf state does not exist for stars that are more massive. In other words, 'the massive stars do not have sufficient energy to cool.' But to return to 1931. Once the existence of the critical mass was established, the question that was puzzling was how one was to relate its existence to the evolution of stars which start as gaseous masses. If the stars had masses less than M,jmit then the assumption that they would eventually settle down to a white-dwarf stage seemed natural enough. But what if their masses were greater? By a simple comparison of the limiting forms of the equations of state, it can be readily shown that if at a given p, Q, and T, the fraction 1 - fi =
\aT*(= radiation pressure) v -- > 0.092, 4A ia T + $Rer( = total pressure)
then matter cannot become degenerate. It is not difficult to convince oneself by fairly rigorous arguments that for perfect-gas stars < 1 — /J>average increases with mass. In fact, one can estimate that for M> 6.6fi~ 2 O, matter can never become degenerate. And this gave rise to the question formulated already in 19328:
A CHAPTER IN THE ASTROPHYSICIST'S VIEW OF THE UNIVERSE
37
Given an enclosure containing electrons and atomic nuclei (total charge zero) what happens if we go on compressing the material indefinitely ? But these questions so clearly raised in 1932 failed to attract any attention. In 1934 the exact mass-radius relations for degenerate configurations were derived.9 Fig. 1 exhibits the mass-radius relation that was deduced on the basis of the exact equation of state. At this time the significance of existence of the limiting mass was so clear that to draw attention to the result as emphatically as possible, the conclusion was formulated in the following terms. 9a The life history of a star of small mass must be essentially different from the history life of a star of large mass. For a star of small mass the natural white-dwarf stage is an initial step towards complete extinction. A star of large mass cannot pass into the white-dwarf stage and one is left speculating on other possibilities.
5-5 50 4-5 40
\ »
K 30
or 25 20 1-5 10 0-5 O 0
01
02
0-3 0-4
0 5 0-6 0 7
OB 0 9
10
Fig. 1. The full-line curve represents the exact (mass-radius)-reIation for completely degenerate configurations. The mass, along the abscissa, is measured in units of the limiting mass (denoted by M3) and the radius, along the ordinate, is measured in the unit h = 7.72//J-1 x 108 cm. The dashed curve represents the relation (2) that follows from the equations of state (1); at the point B along this curve, the threshold momentum po of the electrons at the centre of the configuration is exactly equal to mc. Along the exact curve, at the point where a full circle (with no shaded part) is drawn, po (at the centre) is again equal to mc; the shaded parts of the other circles represent the regions in these configurations where the electrons may be considered to be relativistic (po > mc). (This illustration is reproduced from S. Chandrasekhar, Mon. Not. Roy. Astron. Soc. 95, 219 (1935).)
38
S. CHANDRASEKHAR
Statements very similar to the one I have just quoted from a paper written thirty-eight years ago frequently occur in current literature. But why, it may be asked, were these conclusions not accepted forty years ago? The answer is that they did not meet with the approval of the stalwarts of the day. Thus Eddington 10 commenting on the foregoing conclusion stated: The star apparently has to go on radiating and radiating and contracting and contracting until, I suppose, it gets down to a few kilometres radius when gravity becomes strong enough to hold the radiation and the star can at last find peace.
If Eddington had stopped at that point, we should now be giving him credit for having been the first to predict the occurrence of black holes - a topic to which I shall return presently. But alas! he continued to say: I felt driven to the conclusion that this was almost a reductio ad absurdum of the relativistic degeneracy formula. Various accidents may intervene to save the star, but I want more protection than that. I think that there should be a law of nature to prevent the star from behaving in this absurd way.
In spite of the then prevalent opposition, it seemed to me likely that a massive star, once it had exhausted its nuclear sources of energy, will contract and in the process eject a large fraction of its mass; and further that if by this process, it reduced its mass sufficiently, it could find a state in which to settle. A theoretical advance in a different direction suggested another possibility. It is that as we approach the limiting mass along the white-dwarf sequence, we must reach a point where the protons at the centre of the configuration become unstable with respect to electron capture. The situation is this. Under normal conditions, the neutron is /J-active and unstable while the proton is a stable nucleon. But if in the environment in which the neutron finds itself (as it will in the centre of degenerate configurations near the limiting mass), all the electron states with energies less than or equal to the maximum energy of the /J-ray spectrum of the neutron are occupied, then Pauli's principle will prevent the decay of the neutron. Under these circumstances the proton will be unstable and the neutron will be stable. At these high densities, the equilibrium that will obtain will be one at which, consistent with charge neutrality, there will be just exactly the right number of electrons, protons, and neutrons with appropriate threshold energies that none of the existing protons or neutrons decay. At these densities the neutrons will begin to outnumber the protons and electrons by large factors. In any event it is clear that once neutrons begin to form, the configuration essentially collapses to such small dimensions that the mean density will approach that of nuclear matter and in the range 10 13 to 1015 grams c m - 3 . These are the neutron stars that were first studied in detail by Oppenheimer and Volkoff n in 1939 though their possible occurrence had been suggested by Zwicky some five years earlier. Based on the work of Oppenheimer and Volkoff it appeared likely that a massive star during the course of its evolution could collapse to form a neutron star if during the process of contracting it had reduced its mass sufficiently. The process would clearly be cataclysmic, and it seemed likely that the result would be a supernova phenomenon. If the degenerate cores attain sufficiently high densities (as is possible
A CHAPTER IN THE ASTROPHYSICIST'S VIEW OF THE UNIVERSE
39
for these stars) the protons and electrons will combine to form neutrons. This would cause a sudden diminution of pressure resulting in the collapse of the star onto a neutron core giving rise to an enormous liberation of gravitational energy. This may be the origin of the supernova phenomenon. But the formation of a neutron star, as the result of the collapse, will depend on whether a star, initially more massive than the limiting mass for the white-dwarf stars, ejects just the right amount of mass in order that what remains is in the permissible range of masses for stable neutron stars. While the question of the ultimate fate of massive stars with all its implications was not faced until recently, the theory of the white-dwarf stars, based on the relativistic equation of state for degenerate matter, gained gradual acceptance during the forties and fifties. The principal astronomical reasons for this acceptance were twofold. First, the number of the known white-dwarf stars had, in the meantime, increased very substantially, largely through the efforts of Luyten; and the study of their spectra, particularly by Greenstein, confirmed the adequacy and in some cases even the necessity of the theoretically deduced mass-radius relation. And second, since a time scale of the order of ten million years, for the exhaustion of the nuclear sources of energy of the massive stars, requires the continual formation of these stars, one should be able to distinguish a population of young stars from a population of old stars. And spectroscopic studies provided evidence that the chemical composition of the young stars differs systematically from the chemical composition of old stars; and in fact the difference is in the sense that the young stars appear to have been formed from matter that has been cycled through nuclear reactions. And this last fact is consistent with the picture that during the course of the evolution of the massive stars a large fraction of their masses is returned to interstellar space. And it also seemed likely that this returning of processed matter to the interstellar space was via the supernova phenomenon. While all these ideas became a part of common belief, it remained only as belief. Their full implications were not seriously explored before the discovery of the pulsars. The discovery, in particular, of a pulsar (with the shortest known period) at the centre of the Crab nebula added much credence to the views that I have described since the Crab nebula is itself the remnant of the supernova explosion. The discovery of the association of further pulsars (of longer periods) with what are believed to be the remnants of more ancient supernova explosions strengthens one's conviction. The story of the pulsars and their identification with neutron stars are matters of common knowledge, but I should like to refer to only one aspect that is related to the theory that I have been describing. As we have seen, the construction of degenerate configurations on the Newtonian theory of gravitation leads to a limiting mass for these configurations; and at the limit the radius literally tends to zero unless, that is, no other factor intervenes. And what intervenes is general relativity. Mentioning general relativity I should like to briefly recall the following incident. During my first year in Cambridge (1930-31), R. H. Fowler, who was my official supervisor, was absent in the United States. On this account, there were occasions when I consulted Dirac. My interests at that time were in the theory of white dwarfs.
40
S. CHANDRASEKHAR
While Dirac did evince encouraging interest in these matters, he nevertheless told me that if he should be interested in astrophysics he should rather be working on general relativity and cosmology. That was in 1930. It has taken me well over thirty years to take the first part of his advice! A natural problem for someone with training in classical astrophysics is the radial pulsation of stars and the related question of their dynamical stability. In other words, the re-examination of the pulsation problem that was Eddington's first consideration in the modern theory of the internal constitution of the stars that he initiated. The discussion of the radial pulsations of stars in the framework of general relativity happens to be a particularly simple problem. Indeed, Eddington could have solved it in 1918. The principal result of general relativity here is that stars that would be considered stable in the framework of the Newtonian theory can become unstable in general relativity. More precisely, this instability of relativistic origin would set in whenever the ratio of the specific heats is close to four-thirds. 12 This is the case for degenerate configurations near the limiting mass; and the application of the relativistic criterion shows that they become dynamically unstable before the limiting mass is reached. Precisely what happens is the following. On the Newtonian theory, it can be shown that the period of radial pulsation decreases monotonically to zero as we approach the limiting mass; but in the framework of general relativity, because of the instability it causes, the period attains a minimum and then tends to infinity prior to the limiting mass, where the sequence becomes unstable. In other words, while general relativity does not modify to any appreciable extent the structure derived from the Newtonian theory, it changes qualitatively the period mass relation; it exhibits a minimum period that was absent in the Newtonian theory. And this minimum period is about seven-tenths of a second. Since pulsars are known to have periods much shorter than this minimum value, the possibility of their being white-dwarf configurations was ruled out. The principal conclusions that follow from these theoretical and observational studies can be summarized very simply. Massive stars in the course of their evolution must collapse to dimensions of the order of ten to twenty kilometres once they have exhausted their nuclear source of energy. In this process of collapse, a substantial fraction of the mass will be returned (as processed matter) to the interstellar space. If the mass ejected is such that what remains is in the permissible range of masses for stable neutron stars, then a pulsar will be formed. The exact specification of the permissible range of masses for stable neutron stars is subject to uncertainties in the equation of state for neutron matter; but it is definite that the range is narrow: the current estimate is between 0.3 to 1.0 solar mass. While the formation of a stable neutron star could be expected in some cases, it is clear that their formation is subject to vicissitudes. It is not in fact an a priori likely event that a star initially having a mass of say ten solar masses, ejects during an explosion, subject to violent fluctuations, an amount of mass just sufficient to leave behind a residue in a specified narrow range of masses. It is more likely that the star ejects an amount of mass that is either too large or too little. In such cases the residue will not
A CHAPTER IN THE ASTROPHYSICIST'S VIEW OF THE UNIVERSE
41
be able to settle into afinitestate; and the process of collapse must continue indefinitely till the gravitational force becomes so strong that what Eddington concluded isareductio ad absurdum must in fact happen: 'the gravity becomes strong enough to hold the radiation.' In other words, a black hole must form; and it is to these that I now turn. Let me be more precise as to what one means by a black hole. One says that a black hole is formed when the gravitational forces on the surface become so strong that light cannot escape from it. That such a contingency can arise was surmised, already, by Laplace in 179813. Laplace argued as follows. For a particle to escape from the surface of a spherical body of mass M and radius R, it must be projected with a velocity such that \v% > GM/R; and it cannot escape, if v2 < 2GM/R. On the basis of this last inequality, Laplace concluded that if R<2GM/c2 = RS (say), where c denotes the velocity of light, then light will not be able to escape from such a body and we should not be able to see it. By a curious coincidence, the limit Rs discovered by Laplace is exactly the same that general relativity gives for the occurrence of a trapped surface around a spherical mass. (A trapped surface is one from which light cannot escape to infinity.) While the formula for R looks the same, the radial coordinate r (in general relativity) is so defined that 4nr2 is the area of the 3-surface of constant r; it is not the proper radial distance from the centre. That for a radial coordinate r = Rs, the character of space-time changes is manifest from the standard from the Schwarzschild metric that describes the geometry of spacetime external to a spherical distribution of a total (inertial) mass M located at the centre. The metric is given by dj»,-c»(l..ggW+
\
crJ
2 2 2 2 £ + sm 0d
For a mass equal to the solar mass, the Schwarzschild radius Rs has the value £ s = 2.5km. At one time, the thought that a mass as large as that of the sun could be compressed to a radius as small as 2.5 km would have seemed absurd. One no longer thinks so: neutron stars have comparable masses and radii. Also one must remember that one need not associate high density with the occurrence of trapped surfaces: for M=4x 108 O, for example, i?s = 1014 cm and the mean density of the enclosed matter is only 10" 1/2 grams cm - 3 . From what I have said, collapse of the kind I have described must be of frequent occurrence in the galaxy; and black holes must be present in numbers comparable to,
42
S.CHANDRASEKHAR
if not exceeding, the pulsars. While the black holes will not be visible to external observers, they can nevertheless interact with one another and with the outside world through their externalfields.But one important generalization is necessary and essential. It is known that most stars rotate. And during the collapse of such rotating stars, we may expect the angular momentum to be retained except for that part of it which may be radiated away in gravitational waves. The question now arises as to the end result of the collapse of such rotating stars. One might have thought that the inclusion of angular momentum would make the problem excessively complicated. But if the current ideas are confirmed, the expected end result is not only simple in all essentials, it also provides n possibilities for the astrophysicist. In 1963, Kerr14 discovered the following solution of Einstein's equations for the vacuum which has two parameters M and a and which is also asymptotically flat: ds2 = -—(dt-a
sin20dcj>f + ^ ^ l> 2 + «2)d> - «
Q
+ e^dr2 +
dt
T
Q
2 Q
de2,
A
where
Q2 = r2 + a2 cos2 0 and A = r2 - 2Mr + a2.
(The solution is written in units in which c = G = l ; and in a system of coordinates introduced by Boyer and Lundquist.) Kerr's solution has rotational symmetry about the axis 0 = 0: none of the metric coefficients depend on the cyclic coordinate 4>. It is, moreover, stationary: none of the metric coefficients depend on the coordinate t which is time for an observer at infinity. And Kerr's solution reduces to Schwarzschild's solution when a=0. A test particle describing a geodesic in Kerr's metric at a large distance from the centre will describe its motion as in the gravitational field of a body having a mass M and an angular momentum aM (as deduced from the Lens-Thirring effect). It is now believed that the end result of the collapse of a massive rotating star is a black hole with an external metric that will eventually be Kerr's, all the asymmetries having been radiated away. I shall not attempt to explain the reasons for this belief except to say that they derive, principally, from a theorem of Carter15 (see also Chandrasekhar and Friedman16) which essentially states that sequences of axisymmetric metrics, external to black holes, must be disjoint, i.e. have no members in common. The Kerr metric, like Schwarzschild's, has an event horizon: it occurs at r = Gi[M + (M2 - a 2 ) 1 / 2 ].
(£)
In writing this formula, I have assumed that a<M; if this should not be the case, there
A CHAPTER IN THE ASTROPHYSICIST'S VIEW OF THE UNIVERSE
43
will be no event horizon and we shall have a 'naked singularity', i.e. a singularity that will be visible and communicable to the outside world. For the present, I shall restrict myself to the case a<M. Trajectories, time-like or null, can cross the event horizon from the outside; but they cannot emerge from the inside. In this respect also the Kerr black hole is like the Schwarzschild black hole. But unlike the Schwarzschild metric, the Kerr metric defines another surface external to the event horizon whose equation is given by r = G-2[M + (M2 - a2 cos2 0) , / 2 ].
(S)
This surface touches the horizon at the poles; and it intersects the equator (0 = n/2) on a circle whose radius (=2GMjc2) is larger than that of the horizon. On the surface, an observer who considers himself as staying in the same place must travel with the local velocity of light: like Alice, he must run as fast as he can to stay exactly where he is! Light emitted by such an observer must accordingly appear as infinitely red-shifted to one stationed at infinity. The occurrence of the two separate surfaces (E) and (S) in the Kerr geometry gives rise to unexpected possibilities. These possibilities derive from the fact that in the space between the two surfaces - termed the ergosphere by Wheeler and RufBni - the coordinate r, which is time-like external to the surface (S), becomes space-like. Therefore, the component of the four-momentum in the ^-direction, which is the conserved energy for an observer at infinity, becomes space-like in the ergosphere; it can accordingly assume here negative values. In view of these circumstances, we can contemplate a process in which an element of matter enters the ergosphere from infinity and splits here (in the ergosphere) into two parts in such a way that one part, as judged by an observer at infinity, has a negative energy. Conservation of energy requires that the other part acquire an energy that is in excess of that of the original element. If the part with the excess energy escapes along a geodesic to infinity while the other part crosses the event horizon and is swallowed up by the black hole, then we should have extracted some of the rotational energy of the black hole by reducing its angular momentum. The possibility that such processes can be realized was first pointed out by Penrose17. The fact that energy can be extracted from a Kerr black hole raises the question of its secular stability since it is known that rotating systems can become unstable in the presence of dissipative mechanisms. In the case of the Kerr geometry, dissipation by the emission of gravitational waves is clearly a possibility. The only case of secular instability that is fully understood is that which occurs along the Maclaurin sequence of rotating homogeneous masses. Along this sequence secular instability by viscous dissipation occurs at the point where the Jacobian sequence bifurcates. In the absence of viscous dissipation, no instability occurs at this point; but if viscous dissipation is present, then instability occurs (with an e-folding time depending on the magnitude of the prevailing viscosity) by the neutral mode which transforms the Maclaurin spheroid into a Jacobian ellipsoid at the point
44
S. CHANDRASEKHAR
of bifurcation 18 .Whenthissameproblemisconsideredbyallowingfor radiation-reaction terms that result from the emission of gravitational waves, one findslfl that the Maclaurin spheroid does become secularly unstable for the same value of the eccentricity (for which viscosity causes instability in the Newtonian framework) but the mode by which the instability sets in is not the one which leads to the Jacobian sequence: instead, it is the one which leads to the Dedekind sequence20. (The Dedekind ellipsoid, in contrast to the Jacobi ellipsoid, is stationary in the inertial frame and derives its ellipsoidal shape from internal motions of uniform vorticity.) One may therefore ask whether along the Kerr sequence a Dedekind-like point of bifurcation occurs. A criterion for such occurrence has recently been established21. If by the application of this criterion it can be shown that there is a point along the Kerr sequence where secular instability does set in then, the astrophysical consequences will be immense in view of the existence of mechanisms (such as those of Penrose) by which energy of far larger amounts than by nuclear processes can be extracted from the rotational energy of the black hole. REFERENCES 1. For accounts somewhat similar to the present one but with differing emphasis, see S. Chandrasekhar, Amer. J. Phys. 37, 577 (1969) and Observatory 92, 160 (1972). 2. A. S. Eddington, Internal Constitution of the Stars, Cambridge University Press (1926), p. 173. 3. R. H. Fowler, Proc. Roy. Soc. A, 113, 432 (1926). 4. R. H. Fowler, Mon. Not. Roy Astro.% Soc. 87, 114 (1926). 5. S. Chandrasekhar, Phil. Mag. 11, S92, 594 (1931). 6. S. Chandrasekhar, Astrophys. J. 74, 81 (1931). 7. S. Chandrasekhar, Mon. Not. Roy. Asiron. Soc. 91, 456 (1931). 8. S. Chandrasekhar, Z. Astrophys. 5, 321 (1932). 9. S. Chandrasekhar, Mon. Not. Roy. Astron. Soc. 95, 207 (1935). 9a. S. Chandrasekhar, Observatory 57, 373, 377 (1934). 10. A. S. Eddington, Observatory 58, 38 (1935). 11. J. R. Oppenheimer and G. M. Volkoff, Phys. Rev. 55, 374 (1939). 12. S. Chandrasekhar, Astrophys. J. 140, 417 (1964). 13. P. S. Laplace, Systime de Monde, Book 5, Chapter VI (as quoted by A. S. Eddington in Internal Constitution of the Stars, Cambridge University Press (1926); p. 6). 14. R, P. Kerr, Phys. Rev. Letters, 11, 237 (1963). 15. B. Carter, Phys. Rev. Letters, 26, 331 (1971). 16. S. Chandrasekhar and John L. Friedman, Astrophys. J. 177, 745 (1972). 17. R. Penrose, Nuovo Cimento, Scrie I, 1, 252 (1969) and Nature, 236, 377 (1972). 18. S. Chandrasekhar, Ellipsoidal Figures of Equilibrium, Yale University Press (1969), p. 98. 19. S. Chandrasekhar, Astrophys. J. 161, 561 (1970). 20. S. Chandrasekhar, Ellipsoidal Figures of Equilibrium, Yale University Press (1969), p. 124. 21. S. Chandrasekhar and John L. Friedman, Astrophys. J. (1973) in press.
Remarks on Enrico Fermi S. Chandrasekhar While I appreciate the courtesy of the organizers of this symposium in asking me to talk about Enrico Fermi, I am not altogether certain that I am the most appropriate person: there are others here in the audience who knew him much better and much longer than I did or who were associated with him in some of his more well known contributions. I am probably unique in that I was associated with him in the least known of his work; and that may be an advantage since I can reflect on Fermi's broader interest in physics and not only on those special aspects which have made him well known. That Fermi was a great physicist requires no elaboration; and this is hardly the occasion to attempt an evaluation of his many contributions to many phases of physics. Let it suffice then to enumerate only those of his contributions that are permanently associated with his name: the Fermi transport in general relativity, the Fermi-Dirac statistics, the Fermi-Thomas atom, the Fermi resonance (in C0 2 ), the Fermi theory of /J-decay, Fermi's work on slow neutrons, Fermi's leadership in the construction of the first self-sustaining nuclear reactor, the Fermi mechanism for the acceleration of charged particles in random magnetic fields, and the discovery of the first of the resonances in the proton-pion scattering. Many of Fermi's friends have recalled one or other of these contributions. I was associated with him in none of these, but in one of his side interests which in 1952 was the role of magneticfieldsin astrophysical problems. It was apparently one of Fermi's methods to find someone who might have some knowledge of an area which he wished to learn; and get introduced to the problems of the area by conversations and discussions. And so it came about that in the fall of 1952 and the winter and spring of 1953 I saw Fermi regularly: we met for two hours every Thursday morning and we discussed a variety of astrophysical problems bearing on hydromagnetics and the origin of cosmic radiation. These were problems in which he had not worked before. Nevertheless, I was constantly amazed by the insight with which he used to penetrate to the heart of a problem. In the manner in which he reacted to new problems and new physical situations, he always gave me the impression of a musician who when presented with a new piece of music at once plays it with a perception and a discernment which often surprises even the composer. The fact, of course, was that Fermi was instantly able to bring to bear on any physical problem with which he was confronted his profound and deep feeling for physical laws: the result invariably was that the problem was illuminated and clarified.
/ . Mehra (ed.). The Physicist's Conception of Nature, 800-802. All Rights Reserved Copyright © 1973 by D. Reidel Publishing Companv, Dordrecht-Holland
REMARKS ON ENRICO FERMI
801
I believe that for most of you assembled here, Fermi is only a name that is attached to many important concepts in physics. For that reason, I should like to narrate a few incidents which might give, to those of you who did not know him, some idea as to the manner of man he was. Let me start by describing the way he used to discuss problems with me. It often happened that during our meetings I would state some result. Sometimes he would react by saying that the result did not seem correct to him. He would then start off by saying, 'Let me assume that you are wrong; and let me argue on that basis. If I am wrong in my assumption, then at some point it must become clear that I am wrong and that you are right.' The discussion would then proceed and a point would invariably come when it became clear to both of us that Fermi was indeed right. And at that point he would change the subject. He never took pleasure in drawing attention to the fact that he was right. I do not know if this is the experience of others who have worked with Fermi; but certainly in my experience I have rarely known of any other scientist who showed that same degree of personal generosity in scientific discussions. Still, there was one occasion that I got the better of him. I was once asked to talk to a seminar and when I expressed my doubts as to what I should talk about, Fermi advised, 'If I were you, I would not be technical.' And I asked, 'Do you mean if I were you, or you were me?' This baffled him. Others with greater competence have written about Fermi's fundamental contributions to physics. But his own account of the critical moment when the effect of the slowing down of neutrons on their ability to induce nuclear transformations was discovered is perhaps worth recording. I described to Fermi Hadamard's thesis regarding the psychology of invention in mathematics, namely, how one must distinguish four different stages: a period of conscious effort, a period of 'incubation' when various combinations are made in the subconscious mind, the moment of 'revelation' when the 'right combination' (made in the subconscious) emerges into the conscious, and finally the stage of further conscious effort. I then asked Fermi if the process of discovery in physics had any similarity. Fermi volunteered and said: 'I will tell you how I came to make the discovery which I suppose is the most important one I have made.' And he continued: 'We were working very hard on the neutron induced radioactivity and the results we were obtaining made no sense. One day, as I came to the laboratory, it occurred to me that I should examine the effect of placing a piece of lead before the incident neutrons. And instead of my usual custom, I took great pains to have the piece of lead precisely machined. I was clearly dissatisfied with something: I tried every "excuse" to postpone putting the piece of lead in its place. Whenfinally,with some reluctance, I was going to put it in its place, I said to myself, "No: I do not want this piece of lead here; what I want is a piece of paraffin." It was just like that: with no advanced warning, no conscious, prior, reasoning. I immediately took some odd piece of paraffin I could put my hands on and placed it where the piece of lead was to have been.' As you all know, Fermi became ill during the summer of 1954 while he was in Italy. When he returned to Chicago in the fall, all of his friends at the University were
802
S. CHANDR ASEKHAR
shocked to see how ill he looked. At first, the doctors could not diagnose what was wrong with Fermi. It finally became clear that the cause of the illness was either congestion of the oesophagus or cancer of the stomach and intestines. And the matter would be settled by surgery. Dr L. R. Dragstedt, who performed the surgery, told me of his conversation with Fermi the night preceding the operation. Dragstedt told him that if the problem was with the oesophagus the surgery would be complicated and would take a long time; but if it was cancer of the stomach and intestines, then they could probably do very little about it and the operation would be only of short duration. And the next day on returning from surgery, Fermi opened his eyes and noticed that he had not been in surgery for long. He turned to Dragstedt and asked him, 'Has the mitosis set in?' The answer was yes. Then he asked, 'How many more months?' And Dragstedt replied, 'Some six months.' And Fermi went back to sleep. The following day Herbert Anderson and I went to see Fermi in the hospital. It was of course very difficult to know what to say or how to open a conversation when all of us knew what the surgery had shown. Fermi resolved the gloom by turning to me and saying, 'For a man past fifty, nothing essentially new can happen; and the loss is not as great as one might think. Now you tell me, will I be an elephant next time?' Among the great physicists it has been my privilege to know, Fermi was unique in that he was interested, genuinely, seriously, and deeply, in all aspects of the physical world around us: all of physics interested him and he drew no lines of demarcation.
1167
5S
The Nora and Edward Ryerson Lecture
I
NawTOH AND OR BaathovaN
1S
PHTTaKNS OF QRaHTlVlTY S. Chandrasekhar
nX
The University of Chicago Center for Policy Study
55
Shakespeare, Newton, and Beethoven or Patterns of Creativity
Prefacing a somewhat derogatory criticism of Milton, T. S. Eliot once stated that "the only jury of judgement" that he would accept on his views was that "of the ablest poetical practitioners of his time." Ten years later, perhaps in a more mellow mood, he added: "the scholar and the practitioner, in the field of literary criticism, should supplement each others' work. The criticism of the practitioner will be all the better, certainly, if he is not wholly destitute of scholarship; and the criticism of the scholar will be all the better if he has some experience of the difficulties of writing verse." By the same criterion, anyone, who is emboldened to ask if there are discernible differences in the patterns of creativity among the practitioners in the arts and the practitioners in the sciences, must be a practitioner, as well as a scholar, in the arts as well as in the sciences. It will not suffice to be a practitioner in the arts only, or in the sciences only. Certainly, a wanderer, often lonely, in some of the by-lanes of the physical sciences, has simply not the circumference of comprehension to address himself to a question which encompasses the arts and the sciences. I, therefore, begin by asking your forbearance. Allowing, as we must, for the innumerable individual differences in tastes, temperaments, and comprehension, we ask: Can we in fact discern any major differences in the patterns of creativity among the practitioners in the arts and the practitioners in the sciences? The way I propose to approach this question is to examine, first, the creative patterns of Shakespeare, Newton, and Beethoven, who, by common consent, have,
5
each in his own way, scaled the very summits of human achievement. I shall then seek to determine whether, from the likenesses and the differences in the patterns at these rarified heights, we can draw any larger conclusions which may be valid at lower levels. I I begin with Shakespeare. Shakespeare's education was simple, as Elizabethan education was. While it sufficed and stood him in good stead, Shakespeare was never persuaded by scholarship as such. He clearly expressed his attitude in Small have continual plodders ever won Save base authority from others' books or O, this learning, what a thing it is! Even so, when Shakespeare arrived in London in 1587, at the age of twenty-three, he had none of the advantages of a London background that Lodge and Kyd had, or the advantages of years at Oxford or Cambridge that Peele, Lyly, Greene, Marlowe, and Nashe had. There can be little doubt that Shakespeare was acutely aware of his shortcomings and his handicaps. He overcame them by reading and absorbing whatever came his way. The publication of the revised second edition of Holinshed's Chronicle History of England was particularly timely: it provided Shakespeare with the inspiration for his Chronicle plays, yet to come. By 1592, Shakespeare had written his three parts of Henry VI and his early comedies, Titus Andronicus and The Comedy of Errors. His success with these plays produced Robert Greene's vicious attack on him in that year. Greene was six years older than Shakespeare; and he was among the most prominent figures in the literary life of London at that time. As it happened, Greene's attack was posthumous, as he had died somewhat earlier as the result of a fatal banquet, it is said, "of Rhenish wine and
pickled herrings." It was therefore "a time bomb which Greene left." His attack in part read: For there is an upstart crow, beautified by our feathers, that with his "Tiger's heart wrapped in a player's hide," supposes he is as well able to bombast out a blank verse as the best of you, and being an absolute Johannes Factotum, is in his own conceit the only Shake-scene in a country, Greene's attack brings out very clearly that Shakespeare was considered an outsider and an intruder: he had no university background and he did not belong to the aristocratic court circles. In spite of his early successes, life for Shakespeare, as a player and a playwright, was fraught with uncertainties with the recurring years of the plague and the periodic closing of the theaters in London. But in 1590, Shakespeare found a patron, a friend, and love. Shakespeare's patron was the young Earl of Southampton who came of age in 1591. The intensity of Shakespeare's emotional experience, in the four years that followed, was decisive for the development of his art and for the opportunities that opened up for him. Shakespeare's genius matured and flowered with an unexampled outburst of creative activity. Besides the plays already mentioned, he wrote The Taming of the Shrew, Love's Labour's Lost, and Richard III. The two splendid narrative poems, Venus and Adonis and The Rape of Lucrece, dedicated to the Earl of Southampton, belong to this same period. During 1592-95, Shakespeare wrote his sonnets as a part of his services for Southampton's patronage. The sonnets are the most autobiographical ever written. They throw a flood of light on Shakespeare's attitude to himself and his art; and they also reveal the extent of his dependence on Southampton's friendship and patronage. The course of the friendship between Southampton and Shakespeare was by no means smooth. There was the difference in their ages; there was the disparity in their stations, as the aristocratic patron and a player poet; and besides, there was the complication of Shakespeare's mistress—the dark lady of the sonnets—turning her attention away from Shakespeare to
the responsive Earl. Shakespeare poured his feelings with poignant sincerity into the sonnets: When, in disgrace with fortune and men's eyes, I all alone beweep my outcast state, And trouble deaf heaven with my bootless cries, And look upon myself and curse my fate: Against that time, if ever that time come, When I shall see thee frown on my defects, When as thy love hath cast his utmost sum, Called to that audit by advised respects: Against that time when thou shalt strangely pass, And scarcely greet me with that sun, thine eye, When love, converted from the thing it was, Shall reasons find for that settled gravity: Against that time do I ensconce me here Within the knowledge of mine own desert, And this my hand against myself uprear, To guard the lawful reasons on thy part: To leave poor me thou hast the strength of laws, Since why to love I can allege no cause.
Their relationship, at least as perceived by Shakespeare, was so fragile that he even considers the possibility of death: No longer mourn for me when I am dead Than you shall hear the surly sullen bell Give warning to the world that I am fled From this vile world with vilest worms to dwell:
And Shakespeare feels that his life cannot last longer than Southampton's love and that it will come to an end with it. But do thy worst to steal thyself away, For term of life thou art assured mine; And life no longer than thy love will stay,
1172
For it depends upon that love of thine. Then need I not to fear the worst of wrongs, When in the least of them ray life hath end; I see a better state to me belongs Than that which on thy humour depend. Thou canst not vex me with inconstant mind, Since that my life on thy revolt doth lie: O, what a happy title do I find, Happy to have thy love, happy to die! But what's so blessed-fair that fears no blot? Thou mayst be false, and yet I know it not. In s p i t e of t h e u n c e r t a i n t y w h i c h p e r v a d e s t h e e n t i r e s o n n e t seq u e n c e , S h a k e s p e a r e ' s p r o p h e t i c confidence in h i s o w n p o e t r y o c c a s i o n ally e r u p t s . T h u s , i n t h e f a m o u s s o n n e t 55, w e h a v e t h e o u t p o u r i n g : Not marble, nor the guilded monuments Of princes, shall outlive this powerful rhyme; But you shall shine more bright in these contents Than unswept stone, besmeared with sluttish time. When wasteful war shall statues overturn, And broils root out the work of masonry, Nor Mars's sword nor war's quick fire shall burn, The living record of your memory. M e a n t i m e , M a r l o w e a p p e a r s as a d a n g e r o u s rival to S o u t h a m p t o n ' s p a t r o n a g e . To offset S h a k e s p e a r e ' s V e n u s a n d A d o n i s , M a r l o w e b e g a n writing his Hero and Leander. Shakespeare expresses his uneasiness with this rivalry w h i l e c o n c e d i n g M a r l o w e ' s s u p e r i o r i t y : O, how I faint when I of you do write, Knowing a better spirit doth use your name, And in the praise thereof spends all his might, To make me tongue-tied speaking of your fame! But since your worth, wide as the ocean is, The humble, as the proudest sail doth bear, My saucy bark, inferior far to his,
9
On your broad main doth wilfully appear. Your shallowest help will hold me up afloat, Whilst he upon your soundless depth doth ride; Or, being wrecked, I am a worthless boat, He of tall building and of goodly pride. Then if he thrive and I be cast away The worst of this: my love was my decay. M a r l o w e d i e d in 1593 in a n u n h a p p y b r a w l w h i c h S h a k e s p e a r e clearly h a d in m i n d w h e n h e m a d e T o u c h s t o n e in A s You Like It, say: When a man's verses cannot be understood, nor a man's good wit seconded with the forward child Understanding, it strikes a man more dead than a great reckoning in little room. In t h e s a m e play, S h a k e s p e a r e also p a i d M a r l o w e t h e u n u s u a l t r i b u t e of a d d r e s s i n g h i m as " D e a d s h e p h e r d " a n d q u o t i n g h i s line: Who ever loved that loved not at first sight? A n d before long, t h e u n h a p p y e p i s o d e w i t h t h e " d a r k l a d y " also e n d e d : I am perjured most For all my vows are oaths to misuse thee, And all my honest faith in thee is lost. W i t h t h e last s o n n e t of t h e S o u t h a m p t o n s e q u e n c e , S h a k e s p e a r e emerges triumphant: No, let me be obsequious in thy heart, And take thou my oblation, poor but free, Which is not mixed with seconds, knows no art But mutual render, only me for thee. Yes! " p o o r b u t free," " n o t m i x e d w i t h s e c o n d s " a n d " o n l y m e for t h e e . "
In 1594, the Earl of Southampton gave Shakespeare some such amount as £100 to acquire a share in Chamberlain's Company when it was formed. With the future thus assured, Shakespeare's natural spirits rose and his genius matured. A Midsummer Night's Dream, which he wrote in that year, was the first of his great masterpieces. Soon Romeo and Juliet, As You Like It, and Much Ado About Nothing followed. Then Shakespeare turned again to his chronicle plays: King John, the two parts of Henry IV, and Henry V. The one hero in all these chronicle plays is England; and in them Shakespeare gives lasting expression to "the very age and body of the time." Many consider the two parts of Henry IV as the twin summits of Shakespeare's achievement in his chronicle plays. They are certainly superlative plays made more memorable by the character of Falstaff. It has been said that "in a totally different way, Falstaff is to English literature what his contemporary Don Quixote has been to the Spanish." The great "middle period" of Shakespeare begins with A Midsummer Night's Dream and ends with Hamlet (1600-01). In Hamlet Shakespeare gives expression to his thoughts on the theater and also his reaction to the rising rivalry with Ben Jonson and the Blackfriar's theater with their appeal to wit and fashion. Thus, in his instruction to the players (in the play within the play), we find Hamlet saying: For anything so overdone is from the purpose of playing, whose end, both at the first and now, was and is to hold, as 'twere, the mirror up to nature, to show virtue her own feature, scorn her own image, and the very age and body of the time his form and pressure. Shakespeare is here asserting that "the very age and body of the time" can be expressed in drama—as, indeed, he had expressed his own age in his chronicle plays. There is perhaps a hint of admonition to Ben Jonson and the "reformers" in O it offends me to the soul to hear a robustious periwig-
pated fellow tear a passion to tatters, to very rags, to split the ears of the groundlings, who for the most part, are capable of nothing but inexplicable dumb shows and noise O there be players that I have seen play and heard others praise . . . have so strutted and bellowed that I have thought some of nature's journeymen had made men, and not m a d e t h e m well, t h e y i m i t a t e d h u m a n i t y so abominably . . . O reform it altogether.
The plays that followed Hamlet—All's Well That Ends Well and Measure /or Measure—provide indications that, at this time, Shakespeare's "nerves were on edge": he appears disillusioned with men and things— perhaps, a proper frame of mind to embark on his great tragedies. As A. L. Rowse, the distinguished Elizabethan and Shakespearian scholar, has written, the great tragedies "show evidences of strain and exhaustion"; and he continues: As in all significant work, we have a convergence of factors, on the one side literary, on the other personal . . . If Shakespeare were to compare with his rival Ben Jonson he must do so now in tragedy. With the tragedies he was to make the grandest efforts, extend his powers to his fullest capacity and thus fulfill his destiny as a writer . . . There is cumulative evidence that so far from not caring about his fame and achievement as a writer, his ambition was the highest. The argument has come full circle: here is a personal consideration.
When Shakespeare's work was complete, Ben Jonson was able to compare him only with the great tragedians, Aeschylus, Sophocles, and Euripides. The years 1604-08 saw in succession the plays Othello, King Lear, Macbeth, Antony and Cleopatra, and Coriolanus. It staggers one's imagination to realize that these great plays, so utterly different from one
a n o t h e r , c o u l d h a v e b e e n w r i t t e n , in s u c c e s s i o n , w i t h s u c h u n f a l t e r i n g inspiration. H e r e is H a z l i t t ' s s u m m i n g u p of t h e t r a g e d i e s : Macbeth and Lear, Othello and Hamlet, are usually reckoned Shakespeare's four principal tragedies. Lear stands first for the profound intensity of the passion; Macbeth for the wildness of the imagination and the rapidity of action; Othello for the progressive interest and powerful alternations of feeling; Hamlet for the refined development of thought and sentiment. If the force of genius shown in each of these works is astonishing, their variety is not less so. They are like different creations of the same mind, not one of which has the slightest reference to the rest. This distinctness and originality is indeed the necessary consequences of truth and nature. Hazlitt d o e s n o t i n c l u d e A n t o n y a n d C l e o p a t r a a m o n g t h e great t r a g e d i e s . But n o w a d a y s it is c o n s i d e r e d b y m a n y as e q u a l l y great. As T. S. Eliot in a r e m a r k a b l y s e n s i t i v e a n a l y s i s of A n t o n y a n d C l e o p a t r a h a s said: This is a play for mature actors and for a mature audience, for neither on the stage nor in the audience can immature people enter into the feelings of these middle-aged lovers . . . The peculiar triumph of Antony and Cleopatra is in the fusion of the heroic and the sordid, in the same characters in one vision of life. Marlowe could have made them seem equally majestic. Dryden in his later play on the subject almost does so. But only Shakespeare could have made them at once majestic and human in their weakness; and without the human weaknesses we should not have the greatness and the terror of tragedy. And the reason is that Shakespeare had learned to say things in poetry which no one else could have said in prose.
It has sometimes been suggested that the plays which followed the great tragedies—Timon of A t h e n s , Pericles, P r i n c e of Tyre, and
Cymbeline—all show signs of nervous fatigue. As A. L. Rowse has remarked: "there seems to be a hiatus here, a pause, if not something more, during these years." But a contrary view has been expressed by T. S. Eliot. He has said: The last plays are more difficult. Our astonishment in reading and hearing Antony and Cleopatra might often in many places be expressed by the words, "I should never have thought that that would be said in poetry." Our moments of astonishment in the later plays could better be expressed by the words, "I should never have thought that that could be said at all." For in the last plays, and I mean especially Cymbeline, The Winter's Tale, Pericles, and The Tempest, Shakespeare has abandoned the realism of ordinary existence in order to reveal to us a further world of emotion . . . In any event, Shakespeare's last three plays—The Winter's Tale, The Tempest, and Henry VIII—are more accessible—at least, Shakespeare's natural poise is more evident. Thus, Winter's Tale is a most beautiful and moving play. Hazlitt describes it as "one of the best acting of our author's plays," while the well-known Shakespearian scholar Q. writes: "Winter's Tale is beyond criticism and even beyond praise." In his penultimate play, Shakespeare, ever searching for something new, deals with a profound theme which continues to be vexatious down to this day: in his creation of Caliban, he concretely states for us a central issue of the present age. But the mood of The Tempest is one of farewell: Our revels are now ended. These our actors, As I foretold you, were all spirits, and Are melted into air, into thin air: And like the baseless fabric of this vision, The cloud-capped towers, the gorgeous palaces, The solemn temples, the great globe itself, Yea, all which it inherit, shall dissolve, And, like this insubstantial pageant faded, Leave not a rack behind.
And finally, in his last play, Shakespeare returns to his chronicle of the English story, which he began with Henry VI and Richard III, and completes the cycle with Henry VIII and the birth of Elizabeth. The concluding speech by the Archbishop of Canterbury opening with the incantation: This royal infant—Heaven still move about her— Though in her cradles, yet promises is a form of prophesy of what the Elizabethan age was to be. It gave Shakespeare the splendid opportunity to pay his tribute to the Queen, he had not eulogized at her death in 1603, and to sum up the Elizabethan age now only an imprint on time. And as A. L. Rowse concludes his biography of Shakespeare And this too was Shakespeare's end. But like a splendid coiled snake, glittering and richly iridescent—emblem alike of wisdom and immortality—his work lay about him rounded and complete. Ben Jonson's tribute, included with the first folio He was not of an age, but of all time! has been prophetic. Let me conclude by quoting two contemporary writers. Virginia Woolf, after a vain effort imagining how Shakespeare "coined his words," writes in her diary: Indeed, I could say that Shakespeare surpasses literature altogether, if I knew what I meant. And T. S. Eliot sums up Shakespeare as follows: The standard set by Shakespeare is that of continuous development from first to last, a development in which the
choice both of theme and of dramatic and verse technique in each play seems to be determined increasingly by Shakespeare's state of feeling by the particular stage of his emotional maturity at the time . . . We may say confidently that the full meaning of any one of his plays is not in itself alone, but in that play in the order in which it was written, in its relation to all of Shakespeare's other plays, earlier and later: we must know all of Shakespeare's work in order to know any of it. No other dramatist of the time approaches anywhere near to this perfection of pattern . . . It seems to me to correspond to some law of nature that the work of a man like Shakespeare, whose development in the course of his career was so amazing, that it should reach, as in Hamlet, the point at which it can touch the imagination and feeling of the maximum number of people to the greatest possible depth and that, thereafter, like a comet which has approached the earth and then continued away on its course, he should gradually recede from view until he tends to disappear into his private mystery. II I now turn to Beethoven with more qualms: I am even more painfully aware of my shortcomings to discourse on him. When Beethoven came to Vienna in 1792, at the age of twenty-two, his attitude must have been one of caution; and his studies with Haydn, Schenk, Albrechtsberger, and Salieri were, we may assume, primarily for finding out if there were things he could learn from them. He clearly absorbed what they had to teach him without distorting his own musical ideas. In any event, once he found that he could overpower everyone in Vienna by the sheer virtuosity of his improvisations on the pianoforte, he became impatient and sometimes, even defiant. Thus, Haydn's unfavorable opinion of the third of his three trios, Opus 1, only confirmed Beethoven's own opinion that it was the best of the three and that Haydn's contrary view was due to jealousy and malice.
At this time, Beethoven desired great fame; and he seems to have been convinced that his sheer strength was sufficient to protect him against all misfortune. This attitude is clearly expressed in his letter to von Zmeskall: The devil take you! I do not know anything about your whole system of ethics. Power is the morality of men who stand out from the rest, and it is also mine. This supreme confidence in himself, derived from this morality of power, was soon destined to be tried most sorely. The first signs of his deafness appeared, already, when Beethoven was twenty-eight years. His initial reaction was one of rage at what he considered as the senselessness of the affliction. As he wrote to Karl Amenda three years later (1801) Your Beethoven is most unhappy and at strife with nature and Creator. I have often cursed the latter for exposing his creatures to the merest accident, so that often the most beautiful buds are broken or destroyed thereby. Only think that my noblest faculty, my hearing, has greatly deteriorated. But his fortitude was unshaken, for he continued I am resolved to rise superior to every obstacle . . . I am sure my fortune will not desert me. With whom need I be afraid of measuring my strength . . . I will take Fate by the throat. We obtain a proper appreciation of the state of Beethoven's mind at this time from his famous Heiligenstadt testament written in 1802 but discovered among his papers only after his death. The Heiligenstadt testament is so transparently sincere that it should really be read in its entirety: but the following extract must suffice:
But how humiliated I have felt if somebody standing beside me heard the sound of a flute in the distance and I heard nothing, or if somebody heard a shepherd sing and again I heard nothing—Such experiences almost made me despair, and I was on the point of putting an end to my life—The only thing that held me back was my art. For indeed it seemed impossible to leave this world before I had produced all the works that I felt urged to compose. Beethoven's confession that he contemplated suicide and that it was the power of his unfulfilled art that saved him finds an echo in what he wrote twenty years later: I live only for my art and to fulfill my duties as a man. It is clear that Beethoven's growing deafness shattered his earlier ethics of the morality of power. But like a phoenix it rose only to sustain the realization of his creative powers. Thus, by the time (1807) he came to writing his third Rasoumowsky quartet, his resignation to his affliction appears to be complete: for we find him writing in the margin Let your deafness no longer be secret even for art . . . And the work on the grand scale in which his conflict with fate is taken for granted and ignored is his seventh symphony. This "middle period" of intense creativeness lasted for some ten years. By his early forties, Beethoven had composed his eight symphonies, his five piano concertos, his one violin concerto, his twenty-five piano sonatas, his eleven quartets, his seven overtures, his one opera, and his one mass. At the age of forty-two with this magnificent pile of compositions behind him, Beethoven practically stopped composing for the next seven years. The fruits of his meditation—so they must have been—came after this period of quiescence in a manner that is perhaps without parallel in musical history.
From the first symphony written in 1801 to the eighth symphony written in 1812, it is essentially the same Beethoven: it is, in fact, the Beethoven of the common understanding. But the Beethoven of the ninth symphony, of the mass in D, of the last four piano sonatas, and, most of all, the last five quartets is an altogether different Beethoven. Beethoven's own pupil, Czerny, did not understand his music of this last period; and he tried to explain it away as due to Beethoven's deafness. Thus, he writes Beethoven's third style dates from the time when he became gradually completely deaf . . . Thence comes the dissimilarity of the style of his last three sonatas . . . Thence many harmonic roughnesses . . . By all accounts, Beethoven's last quartets are a Mount Everest of an achievement. Here is a sample of what has been said about them: They are peerless They are beyond description or analysis in words The last quartets are unique, unique for Beethoven, unique in all music. But this much may certainly be said: Nobody can say what the quartets really mean; we can only be sure that they express ideas nowhere else to be found. Wordsworth's description of Newton's mind "as voyaging through strange seas of thought alone" applies equally to Beethoven's mind of this last period. Beethoven's last complete work, the quartet No. 16 in F major, provides a noble ending to his great sequence. Of this quartet, J.W.N. Sullivan has written It is the work of a man who is fundamentally at peace. It is the peace of a man who has known conflicts, but whose conflicts are now reminiscent. This quality is most apparent in the last movement with its motto, "Muss es sein? Es muss sein!" (Must it be? It must be!)
Reviewing the life and work of Beethoven, Sullivan sums him up as follows: One of the most significant facts, for the understanding of Beethoven, is that his work shows an organic development up until the very end . . . The greatest music Beethoven ever wrote is to be found in the last string quartets, and the music of every decade before the final period was greater than its predecessor. It is striking how close this summing of Beethoven is to T. S. Eliot's summing of Shakespeare which I quoted earlier. The way Shakespeare and Beethoven overcame the crises of their early years, the continual growth of their minds, the organic unity of their works spanning their entire lives, their great masterpieces towards the end, and even the moods of farewell in The Tempest and in the sixteenth quartet, all these are indeed most striking. Ill I now turn to Newton. Isaac Newton, a posthumous child, born with no father on Christmas Day 1642, was, as Maynard Keynes has aptly written, "the last wonder child to whom the Magi could do sincere and appropriate homage." One of the most remarkable aspects of Newton's most remarkable life is the explosive outburst of his genius. He was not an infant prodigy; and it is probable that when he came to Cambridge in 1661, he knew little more than elementary arithmetic. And it must be remembered that the new outlook on scientific thought that we associate with the names of Galileo, Kepler, and Descartes had hardly yet penetrated the walls of Oxford and Cambridge. Nevertheless, by 1664, when Newton was in his twenty-third year, his genius seems to have flowered. Thus, Newton recalled in his old age that he had "found the method of Infinite Series at such time (1664-65)." Newton, in fact, wrote out his notes as a connected essay entitled, "On Analysis of Equations with an Infinite Number of
Terms" and allowed Barrow to send it to Collins, stipulating, however, that he remain anonymous. This stipulation was withdrawn later; but we encounter here the first indication of a trait which was later to become an obsession with Newton. By the summer of 1665, when Cambridge was evacuated on account of the plague and Newton had gone to Woolsthorpe, his genius was fully in flower. It manifested itself in a manner unsurpassed in the history of scientific thought. But it was not until many years later that the world was to know what happened during the two years that Newton was at Woolsthorpe. For here at Woolsthorpe, Newton at the age of twenty-three made three of the greatest discoveries in science: the Differential Calculus, the Composition of Light, and the Laws of Gravitation. Writing towards the end of his life, Newton recalled his discovery of the laws of gravitation thus: In the same year (1666} I began to think of gravity extending to the orb of the moon . . . I deduced that the forces which keep the planets in their orbs must be reciprocally as the squares of their distances from the centers about which they revolve; and thereby compared the force requisite to keep the moon in her orb with the force of gravity at the surface of the earth, and found them answer pretty well. All this was in the two plague years 1665 and 1666, for in those days I was in the prime of my age for invention, and minded mathematics and philosophy more than at any time since. Notice, first, his statement that "in those days . . . I minded mathematics and philosophy [meaning science] more than at any time since." Notice also the curious words "answer pretty well" to the agreement he had found with respect to the acceleration experienced by the moon in its orbit and as deduced—on the basis of his inverse-square law—from the acceleration experienced by bodies on the earth, i.e., the falling apple. Newton does not appear to have felt any urgency to verify if his prediction "answers" more than "pretty well." Indeed, he does not seem to have
experienced any special delight in having discovered so fundamental a law of nature. In actual fact, he dismissed the entire matter from his mind for a decade and more. Newton returned to Cambridge early in 1667; and in 1669 he was appointed to the Lucasian Chair of Mathematics in succession to Barrow who had relinquished the Chair on Newton's behalf. Soon after his return to Cambridge, Newton appears to have completed to his satisfaction his experimental investigations on the composition of light and constructed his first reflecting telescope to avoid the chromatic aberrations of the then extant refracting telescopes. But he did not publish any of these results of his investigations for several years. The news of Newton having constructed a telescope on a new principle soon spread and Newton was urged to exhibit it at the Royal Society. It is known that Newton sent at least two telescopes to the Royal Society and that the second of them was exhibited in 1671. Newton was elected to the Royal Society in January 1672. Stimulated perhaps by this recognition, Newton acceded to the request by Oldenburg, then the Secretary of the Royal Society, to communicate to the Society an account of his discoveries and in particular the principles underlying the construction of his telescope. In two successive letters, Newton replied to Oldenburg as follows; I shall endeavour to testify my gratitude by communicating what my poor and solitary endeavours can effect towards the promoting your philosophical designs. (January 6, 1672)
In the next letter he suggests communicating an account of his optical discoveries rather than a description of his telescope. He writes An account of a philosophical discovery . . . which I doubt not but will prove much more grateful than the communication of that instrument, being in my judgement the oddest, if not the most considerable detection, which has hitherto been made in the operation of nature. (January 18, 1672)
I should like to draw your attention especially to the words, "the oddest, if not the most considerable detection." This is the first and the only time that Newton expresses a trace of enthusiasm with respect to any of his discoveries. But what followed the publication of Newton's account of his experiments on the composition of light was nothing short of a disaster. A vigorous controversy ensued; and Newton appears to have been irritated beyond endurance by the inability of his critics even to comprehend what it was he had experimentally demonstrated. This lack of c o m p r e h e n s i o n is apparent, for example, from Huygens—even Huygens—arguing that there "would still remain the great difficulty of explaining by mechanical principles, in what consists the diversity of colours, even supposing that Newton's decomposition of white light into the colours of the spectrum is correct." At first Newton tried to persuade by clarifying his method: For the best and safest method of philosophizing seems to be, first to enquire diligently into the properties of things, and of establishing those properties by experiments, and then to proceed more slowly to hypotheses for the explanation of them. For hypotheses should be subservient only in explaining the properties of things, but not assumed in determining them; unless so far as they may furnish experiments . . . (Parenthetically, we may notice that Newton is, here, enunciating what he was to formulate later in his famous aphorism Hypotheses non fingo—I frame no hypotheses.) Newton's failure to persuade resulted in the aversion he now formed to scientific publication, discussion, and arguments. Thus, he wrote to Oldenburg: I have long since determined to concern myself no further about the promotion of philosophy. (December 5, 1672)
I see I have made myself a slave to Philosophy, but if I get free of Mr. Linus' business I will resolutely bid adieu to it eternally, except what I do for my private satisfaction, or leave to come out after me. For I see, a man must either resolve to point out nothing new or to become a slave to defend it. (November 18, 1676) This aversion to scientific publication, discussion, and argument was to find repeated expressions in later years. Here are two examples: For I see not what there is desirable in public esteem, were I able to acquire and maintain it. It would perhaps increase my acquaintance, the thing which I chiefly study to decline. I am grown of all men the most shy of setting pen to paper about anything that may lead into disputes. (September 12, 1682) Soon after the publication of his optical discoveries, Newton receded into himself and we do not know very much as to how he occupied himself during the following decade. But we do know that in 1679, Newton had proved for himself that under the influence of a central inversesquare attractive force an object will describe an elliptical orbit, with the center of attraction at one of its foci. But, again, he kept the result to himself. At long last, in 1684, an incident, not of Newton's making, was to change the course of scientific history. In January of that year, at a meeting in London between Christopher Wren, Robert Hooke, and Edmund Halley, the question arose as to the nature of the orbit a planet would describe under the influence of an inverse-square attractive gravitational force. Since none of them knew how the question could be resolved, Halley went to Cambridge in August of that year to inquire if Newton had any suggestions to offer. To Halley's inquiry, Newton replied at once that the orbit would be an ellipse; and that he had established this result for himself some seven years earlier. Halley was overjoyed and he wished to see Newton's proof. On Newton finding that he had mislaid the piece of
paper on which he had written out the proof, he promised to rework it and send it to him shortly. The reworking of this old problem seems to have aroused Newton's interest in the whole area. By October, he had worked out enough problems to serve as a basis for nine lectures which he gave during the Michaelmas term under the title De Moto Corporum. Halley, on receiving Newton's promised proof at about this time and hearing also of Newton's lectures, went to Cambridge once again, this time to persuade Newton to publish his lectures. By now Newton's mathematical genius seems to have been fully aroused; and Newton appears to be caught in its grip. Newton now entered upon a period of the most intense mathematical activity. Against his will and against his preferences, Newton seems to have been propelled inexorably forward, by the pressure of his own genius, till, at last, he had accomplished the greatest intellectual feat of his life, the greatest intellectual feat in all of science. Let us pause for a moment to take full measure of the magnitude of the feat. By Newton's own account, he began writing the Principia towards the end of December 1684; and he sent the completed manuscript of all three books of the Principia to the Royal Society in May 1686, i.e., in seventeen months. He had solved two of the propositions in the first Book in 1679; and he had also proved eight of the propositions in the second Book in June and July 1685. There are ninety-eight propositions in the first Book; fifty-three in the second; and forty-two in the third. By far the larger proportion of them was, therefore, enunciated and proved during the seventeen consecutive months that Newton was at work on the three Books. It is this rapidity of execution, besides the monumental scale of the whole work, that makes this achievement incomparable. If the problems enunciated in the Principia were the results of a lifetime of thought and work, Newton's position in science would still be unique. But that all these problems should have been enunciated, solved, and arranged in logical sequence in seventeen months is beyond human comprehension. It can be accepted only because it is a fact: it just happens to be so! It is only when we observe the scale of Newton's achievement that comparisons, which have sometimes been made of other men of science
to him, appear altogether inappropriate both with respect to Newton and with respect to the others. In fact, only in juxtaposition with Shakespeare and Beethoven is the consideration of Newton appropriate. Now, a few remarks concerning the style of the Principia. Quite unlike his early communications on his optical discoveries, the Principia is written in a style of glacial remoteness which makes no concessions to his readers. As Whewell aptly wrote: . . . As we read the Principia, we feel as when we are in an ancient armoury where the weapons are of gigantic size; and as we look at them, we marvel what manner of men they were who could use as weapons what we can scarcely lift as a burden . . . It is, however, clear that the rigid and the lamellated style of the Principia is deliberate. For after the publication of the Principia, Newton is reported to have told Rev. Dr. Derham: To avoid being baited by little smatterers in mathematics, I designedly made the Principia abstruse; but yet so as to be understood by able mathematicians who, I imagine, by comprehending my demonstrations would concur with my theory. Although Newton was only forty-two years of age when he finished writing the Principia and was, quite literally, at the height of his mathematical powers and was to remain in full possession of his faculties for another forty years, he never again seriously concerned himself with scientific investigation. He turned to an utterly different way of living. And in time he became one of the principal sights of London for all visiting intellectuals: the Sir Isaac Newton of popular tradition. No account of Newton's life, however brief, can omit some indication of the manner of man he was. The subject is a complex and a controversial one. But this much can fairly be said: Newton seems to have been remarkably insensitive: impervious to the arts, tactless, and with no understanding whatsoever of others.
N e w t o n ' s m o s t r e m a r k a b l e gift w a s p r o b a b l y h i s p o w e r s of c o n c e n t r a tion. As K e y n e s w r o t e His peculiar gift was the power of holding continuously in his mind a purely mental problem until he had seen straight through it. I fancy his pre-eminence is due to his muscles of his intuition being the strongest and most enduring with which a man has ever been gifted . . . I believe that Newton could hold a problem in his mind for hours and days and weeks until it surrendered to him its secret. B e s i d e s , as De M o r g a n h a s said . . . So happy in his conjectures as to seem to know more than he could possibly have any means of proving . . . But t h e c e n t r a l p a r a d o x of N e w t o n ' s life is t h a t h e d e l i b e r a t e l y a n d systematically ignored his supreme mathematical genius and through m o s t of h i s life n e g l e c t e d t h e o n e activity for w h i c h h e w a s gifted b e y o n d a n y m a n . T h i s p a r a d o x c a n be Tesolved o n l y if w e realize t h a t N e w t o n s i m p l y d i d n o t c o n s i d e r s c i e n c e a n d m a t h e m a t i c s as of a n y great i m p o r t a n c e ; or, as K e y n e s h a s said t h e s a m e t h i n g , s o m e w h a t differently, . . . It seems easier to understand . . . this strange spirit, who was tempted by the Devil to believe, at the time when within these walls [of Trinity College] he was solving so much, that he could reach all the secrets of God and Nature by the pure power of mind—Copernicus and Faustus in one. A n d finally, I c a n n o t desist r e p e a t i n g N e w t o n ' s oft-quoted e v a l u a t i o n of himself. I do not know what I may appear to the world, but to myself I seem to have been only like a boy playing on the sea-shore, and diverting myself in now and then finding a
smoother pebble or a prettier shell than ordinary, whilst the great ocean of truth lay all undiscovered before me. In view of Newton's insensitiveness to others, doubts have sometimes been raised about the sincerity of this statement. I do not believe that such doubts are warranted: only someone, like Newton, who can view knowledge from his height, can have the vision of an "ocean of undiscovered truth." As an ancient proverb of India says, "Only the wise can plumb the wells of wisdom." IV From the foregoing accounts of the creative patterns of Shakespeare, Beethoven, and Newton, though very brief and very inadequate, two facts emerge with startling clarity: the remarkable similarity in the creative patterns of Shakespeare and Beethoven on the one hand and their stark contrast with that of Newton on the other. Are the similarity and the contrast accidental? Or, are they manifestations of a general phenomenon which in the case of these giants only happens to be very sharply etched? Consider in juxtaposition the following statements that have been made concerning the creativity of mathematicians and of poets. G. H. Hardy, an outstanding English mathematician of this century, in his essay A Mathematician's Apology—an essay which has been described by C. P. Snow as "the most beautiful statement of the creative mind ever written or ever likely to be written"—G. H. Hardy in his essay writes No mathematician should ever allow himself to forget that mathematics, more than any other art or science, is a young man's game . . . Galois died at twenty-one, Abel at twenty-seven, Ramanujan at thirty-three, Riemann at forty. There have been men who have done great work a good deal later; . . . [but] I do not know an instance of a major mathematical advance initiated by a man past fifty. . . . A mathematician may still be competent enough at sixty, but it is useless to expect him to have original ideas.
And with respect to Ramanujan's early death, Hardy has further written The real tragedy about Ramanujan was not his early death. It is, of course, a disaster that any great man should die young; but a mathematician is comparatively old at thirty, and his death may be less of a catastrophe than it seems . . .
Place beside these statements of Hardy the following one of A. L. Rowse on the death of Christopher Marlowe at the age of twenty-nine: What would he not have achieved if he had lived!—his was the greatest of all losses to English Literature.
Or, of Desmond King-Hele on the death of Shelley at the age of thirty: The rule that a poet is at his best after the age of 30 might have applied as well to him as to Shakespeare, Milton, Wordsworth, Byron, Tennyson, and indeed almost every major English poet who lived to be over 30.
In a more negative vein, there is the statement attributed to Thomas Huxley that a man of science past sixty does more harm than good. I do not doubt that these statements will be challenged or, at least, subjected to qualifications. But consider this. In 1817, at the age of forty-seven, when the long period of meditation, during which Beethoven composed very little, was coming to an end, he said to Cipriani Potter with transparent sincerity, "Now, I know how to compose." I do not believe that there has been any scientist, past forty, who could have said, "Now, I know how to do research." And this to my mind is the center and the core of the difference: the apparent inability of a scientist to continually grow and mature.
V If one should wish to establish with some degree of certainty that a contrast does exist in the patterns of creativity among the practitioners in the arts and the practitioners in the sciences, then one should undertake a survey of an extent and a depth which is far beyond my resources. At the same time it does not seem entirely proper that I leave the matter without some further examples. I shall consider four examples taken from science. My first example is James Clerk Maxwell who is generally considered the greatest physicist of the nineteenth century. Maxwell's principal contributions to physics are his founding of the kinetic theory of gases and the dynamical theory of the electromagnetic field. The new physical concepts which Maxwell introduced in formulating his equations of the electromagnetic field—Maxwell's equations which every student of physics knows—have been described by Einstein as "the most fruitful and profound that physics has experienced since the time of Newton." The four great memoirs which encompass Maxwell's contributions to the two areas were published during the five years 1860-65 when he was between the ages of thirty and thirty-five and was a professor at King's College, London. At the end of this period of intense activity, Maxwell resigned his professorship in London and retired to his country home in Glenlair in Scotland. (Maxwell's biographers never really "explain" why Maxwell felt it necessary to take these actions.) In Glenlair, for the following six years, Maxwell seems to have lived in quietness, occupied, principally, with the planning of his two-volume treatise on Electricity and Magnetism (which was eventually completed and published in 1873). In 1871, Maxwell was persuaded to leave his retirement in Glenlair and to return to academic life in Cambridge as the first Cavendish Professor of Experimental Physics. He died in 1878 at the age of forty-eight. Maxwell's eight years in Cambridge were devoted mostly to editing the scientific papers of Henry Cavendish, organizing and establishing the Cavendish Laboratory, and other diverse university matters. While Maxwell's early death was a tragedy, it must be admitted that his work did not rise again to the heights it had in his early thirties. My second example is George Gabriel Stokes. Stokes was elected to
the Lucasian Chair of Mathematics (in Cambridge) in 1849, when he was just past thirty. He held this Chair until his death in 1903—a Chair that was once held by N e w t o n . Stokes is one of the great figures of nineteenth-century physics and mathematics; and his name continues to be associated with several current notions and concepts. Thus, we have the Navier-Stokes equations governing viscous flow in hydrodynamics; Stokes' law giving the asymptotic rate of fall of small spherical bodies in a viscous medium—a law which provides the basis for Millikan's "oil-drop experiment" for determining the charge on the electron; Stokes' parameters for characterizing polarized radiation which are relevant to most current measurements in radio-astronomy; Stokes' law of fluorescence, that the wavelength of the fluorescing light must exceed that of the exciting light; and Stokes' theorem which in addition to being a very fundamental theorem provides a key element for modern developments in the calculus of differential forms. Now Stokes' scientific papers are collected in five medium-sized volumes. The first three volumes contain all the important concepts and notions that I have enumerated and covers the ten-year period 1842—52; and the remaining two volumes suffice to cover his entire scientific work of the following fifty years. G. Evelyn Hutchinson (the distinguished zoologist at Yale University), whose father was a close associate of Stokes during his last years, makes the remarkable statement: "Stokes, however, quite possibly, emulated his great predecessor [in the Lucasian Chair] consciously. . . . What Newton did, Stokes deemed appropriate for him to do also." My third example is Einstein. The year 1905 was the annus mirabilis both for Einstein and for physics. It was in that year that Einstein, at the age of twenty-six, published three papers, each epoch-making in its own way: the first laid the foundations for his special theory of relativity with remarkable clarity, conciseness, and coherence; the second provided a rational molecular basis (independently of Smoluchowski) for accounting for Brownian motion; and the third carried Planck's hypothesis of the quantum to its logical limit to formulate the concept of the light quantum. In the decade that followed, Einstein was constantly preoccupied with the resolution of the basic inconsistency between Newton's laws of gravita-
tion, with its postulate of instantaneous action at a distance, and his own special theory of relativity, with its postulate that no signal can be propagated with a velocity exceeding that of light. After many detours and false starts, Einstein finally arrived triumphantly to his general theory of relativity in 1916. As Hermann Weyl later expressed, Einstein's general theory of relativity is "one of the greatest examples of the power of speculative thought." In the years following the founding of his general theory of relativity, Einstein made a number of important contributions to the further ramifications of his own general theory as well as to certain aspects of statistical physics. But already by 1925, Einstein was letting the newer developments in the quantum theory, initiated by Heisenberg, pass him by. Thus, Heisenberg records that at the Solvay Congress in 1927, Paul Ehrenfest, Einstein's friend, said to him, "Einstein, I am ashamed of you: you are arguing against the new quantum theory just as your opponents argue about relativity theory." Heisenberg adds sadly that this friendly admonition went unheeded. As Einstein's great admirer Cornelius Lanczos observes From 1925 on his interest in the current affairs of physics begins to slacken. He voluntarily abdicated his leadership as the foremost physicist of his time, and receded more and more into voluntary exile from his laboratory, a state into which only a few of his colleagues were willing to follow. During the last thirty years of his life he became more and more a recluse who lost touch with the contemporary developments of physics. I should like to conclude with an example which in some ways appears counter to Hardy's general rule. I wish to consider the case of Lord Rayleigh, perhaps the greatest pillar of classical mathematical physics. Rayleigh's productivity was remarkably steady and uniform all through his fifty years of scientific publication. His scientific work is encompassed in a two-volume treatise on The Theory of Sound and six large volumes of his Scientific Papers. In a memorial address, delivered in Westminster Abbey in December
1 9 2 1 , J. J. T h o m s o n e v a l u a t e d R a y l e i g h ' s scientific c o n t r i b u t i o n s in t h e following terms: Among the 446 papers which fill these volumes [his six volumes of his Scientific Papers], there is not one that is trivial, there is not one which does not advance the subject with which it deals, there is not one which does not clear away difficulties; and among that great number there are scarcely any which time has shown to require correction . . . Lord Rayleigh took physics for his province and extended the boundary of every department of physics. The impression made by reading his papers is not only due to the beauty of the new results attained, but to the clearness and insight displayed, which gives one a new grasp of the subject . . . T h i s is a r e m a r k a b l e t e s t i m o n y ; a n d a n y o n e w h o h a s h a d o c c a s i o n to u s e R a y l e i g h ' s Scientific P a p e r s w i l l testify to its a c c u r a c y . But w h y w a s R a y l e i g h so different from M a x w e l l a n d E i n s t e i n ? P e r h a p s , t h e c l u e is to be f o u n d in w h a t T h o m s o n said in t h e s a m e memorial address: There are some great men of science whose charm consists in having said the first word on a subject, in having introduced some new idea which has proved fruitful; there are others whose charm consists perhaps in having said the last word on the subject, and who have reduced the subject to logical consistency and clearness. I think by temperament Lord Rayleigh belonged to the second group. A n d p e r h a p s t h e r e is a c l u e also in R a y l e i g h ' s r e s p o n s e to h i s son (also a d i s t i n g u i s h e d physicist) w h e n h e a s k e d h i m to c o m m e n t o n H u x l e y ' s r e m a r k I q u o t e d earlier, " t h a t a m a n of s c i e n c e past sixty d o e s m o r e h a r m t h a n g o o d . " Rayleigh w a s sixty-seven at t h a t time; a n d h i s r e s p o n s e w a s That may be, if he undertakes to criticize the work of younger men, but I do not see why it need be so if he sticks to the things he is conversant with.
Perhaps there is a moral here for all of us! VI I now pass on to some cognate matters. First, may I say that I am frankly puzzled by the difference that appears to exist in the patterns of creativity among the practitioners in the arts and the practitioners in the sciences: for, in the arts, as in the sciences, the quest is after the same elusive quality: beauty. But what is beauty? In a deeply moving essay on "The Meaning of Beauty in the Exact Sciences," Heisenberg gives a definition of beauty which I find most apposite. The definition, which Heisenberg says goes back to antiquity, is that "beauty is the proper conformity of the parts to one another and to the whole." On reflection, it does appear that this definition touches the essence of what we may describe as "beautiful": it applies equally to King Lear, the Missa Solemnis, and the Principia. There is ample evidence that in science, beauty is often the source of delight. One can find many expressions of such delight scattered through the scientific literature. Let me quote a few examples. Kepler: Mathematics is the archetype of the beautiful. David Hilbert (in h i s m e m o r i a l a d d r e s s for H e r m a n n M i n k o w s k i ) : Our Science, which we loved above everything, had brought us together. It appeared to us as a flowering garden. In this garden there were well-worn paths where one might look around at leisure and enjoy oneself without effort, especially at the side of a congenial companion. But we also liked to seek out hidden trails and discovered many an unexpected view which was pleasing to our eyes; and when the one pointed it out to the other, and we admired it together, our joy was complete.
Hermann Weyl (as quoted by Freeman Dyson): My work always tried to unite the true with the beautiful; but when I had to choose one or the other, I usually chose the beautiful. Heisenberg (in a discussion with Einstein): If nature leads us to mathematical forms of great simplicity and beauty—by forms I am referring to coherent systems of hypothesis, axioms, etc.—to forms that no one has previously encountered, we cannot help thinking that they are "true," that they reveal a genuine feature of nature . . . You must have felt this too: the almost frightening simplicity and wholeness of the relationships which nature suddenly spreads out before us and for which none of us was in the least prepared. All these remarks, that I have quoted, may appear vague or too general. Let me try to be concrete and specific. The discovery by Pythagoras, that vibrating strings, under equal tension, sound together harmoniously, if their lengths are in simple numerical ratios, established for the first time a profound connection between the intelligible and the beautiful. I think we may agree with Heisenberg that this is "one of the truly momentous discoveries in the history of mankind." Kepler was certainly under the influence of the Pythagorean concept of beauty when he compared the revolution of the planets about the sun with a vibrating string and spoke of the harmonious concord of the different planetary orbits as the music of the spheres. It is known that Kepler was profoundly grateful that it had been reserved for him to discover, through his laws of planetary motion, a connection of the highest beauty. A more recent example of the reaction of a great scientist, to this aspect of beauty at the moment of revelation of a great truth, is provided by Heisenberg's description of the state of his feelings when he found the key that opened the door to all the subsequent developments in the quantum theory.
Towards the end of May 1925, Heisenberg, ill with hay fever, went to Heligoland to be away from flowers and fields. There by the sea, he made rapid progress in resolving the difficulties in the quantum theory as it was at that time. He writes: Within a few days more, it had become clear to me what precisely had to take the place of the Bohr-Sommerfeld quantum conditions in an atomic physics working with none but observable magnitudes. It also became obvious that with this additional assumption, I had introduced a crucial restriction into the theory. Then I noticed that there was no guarantee that. . . the principle of the conservation of energy would apply . . . Hence I concentrated on demonstrating that the conservation law held; and one evening I reached the point where I was ready to determine the individual terms in the energy table [Energy Matrix] . . . When the first terms seemed to accord with the energy principle, I became rather excited, and I began to make countless arithmetical errors. As a result, it was almost three o'clock in the morning before the final result of my computations lay before me. The energy principle had held for all the terms, and I could no longer doubt the mathematical consistency and coherence of the kind of quantum mechanics to which my calculations pointed. At first, I was deeply alarmed. I had the feeling that, through the surface of atomic phenomena, I was looking at a strangely beautiful interior, and felt almost giddy at the thought that I now had to probe this wealth of mathematical structure nature had so generously spread out before me. I was far too excited to sleep, and so, as a new day dawned, I made for the southern tip of the island, where I had been longing to climb a rock jutting out into the sea. I now did so without too much trouble, and waited for the sun to rise.
May I allow myself at this point a personal reflection? In my entire scientific life, extending over forty-five years, the most shattering experience has been the realization that an exact solution of Einstein's equations
of general relativity, discovered by the New Zealand mathematician, Roy Kerr, provides the absolutely exact representation of untold numbers of massive black holes that populate the universe. This "shuddering before the beautiful," this incredible fact that a discovery motivated by a search after the beautiful in mathematics should find its exact replica in Nature, persuades me to say that beauty is that to which the human mind responds at its deepest and most profound. Indeed, everything I have tried to say in this connection has been stated more succinctly in the Latin mottos: Simplex sigillum veri—The simple is the seal of the true and Puichritudo splendor veritatis—Beauty is the splendour of truth.
VII But I must return to my question: why is there a difference in the patterns of creativity among the practitioners in the arts and the practitioners in the sciences. I shall not attempt to answer this question directly; but I shall make an assortment of remarks which may bear on the answer. First, I should like to consider how scientists and poets view one another. When one thinks of the attitude of the poets to science, one almost always thinks of Wordsworth and Keats and their oft-quoted lines: A fingering slave, One that would peep and botanize Upon his mother's grave? A reasoning self-sufficing thing, An intellectual All-in-all! Sweet is the lore which Nature brings;
Our meddling intellect Misshapes the beauteous forms of things: We murder to dissect. (Wordsworth) Do not all charms fly At the mere touch of cold philosophy? There was an awful rainbow once in heaven: We know her woof, her texture; she is given In the dull catalogue of common things. Philosophy will clip an Angel's wings. (Keats) These lines, perhaps, find an echo in a statement of Lowes Dickinson, "When Science arrives, it expels Literature." It is to be expected that one should find scientists countering these views. Thus, Peter Medawar counters Lowes Dickinson by The case I shall find evidence for is that when literature arrives, it expels science . . . The way things are at present, it is simply no good pretending that science and literature represent complementary and mutually sustaining endeavours to reach a common goal. On the contrary, where they might be expected to cooperate, they compete. It would not seem to me that one can go very far in these matters by pointing accusing fingers at one another. So, let me only say that the attitudes of Wordsworth and Keats are by no means typical. A scientist should rather consider the attitude of Shelley. Shelley is a scientist's poet. It is not an accident that the most discriminating literary criticism of Shelley's thought and work is by a distinguished scientist Desmond King-Hele. As King-Hele has pointed out, "Shelley's attitude to science emphasizes the surprising modern climate of thought in which he chose to live" and that Shelley "describes the mechanisms of Nature with a precision and a wealth of detail unparalleled in English poetry." And here is A. N. Whitehead's testimony:
1202
Shelley's attitude to Science was at the opposite pole to that of Wordsworth. He loved it, and is never tired of expressing in poetry the thoughts which it suggests. It symbolizes to him joy, and peace, and illumination . . . I s h o u l d like to r e a d t w o e x a m p l e s from S h e l l e y ' s p o e t r y w h i c h s u p p o r t w h a t has b e e n said a b o u t h i m . T h e first e x a m p l e is from h i s C l o u d w h i c h "fuses t o g e t h e r a creative m y t h , a scientific m o n o g r a p h , a n d a gay p i c a r e s q u e tale of c l o u d a d v e n t u r e " : I am the daughter of Earth and Water, And the nursling of the Sky; I pass through the pores of the ocean and shores; I change, but I cannot die. For after the rain when with never a stain The pavilion of Heaven is bare, And the winds and sunbeams with their convex gleams Build up the blue dome of air, I silently laugh at my own cenotaph, And out of the caverns of rain, Like a child from the womb, like a ghost from the tomb, I arise and unbuild it again. T h e s e c o n d e x a m p l e is from P r o m e t h u s U n b o u n d , w h i c h has b e e n des c r i b e d b y H e r b e r t R e a d as " t h e g r e a t e s t e x p r e s s i o n e v e r g i v e n to h u m a n i t y ' s d e s i r e for i n t e l l e c t u a l light a n d s p i r i t u a l l i b e r t y " : The lightning is his slave; heaven's utmost deep Gives up her stars, and like a flock of sheep They pass before his eyes, are numbered and roll on! The tempest is his steed, he strides the air; And the abyss shouts from his depth laid bare, Heaven, hast thou secrets? Man unveils me; I have none. Let m e t u r n to a s l i g h t l y different a s p e c t of t h e matter. W h a t are w e to m a k e of t h e f o l l o w i n g c o n f e s s i o n of C h a r l e s D a r w i n :
39
Up to the age or thirty, or beyond it, poetry of many kinds, such as the works of Milton. Gray, Byron, Wordsworth, Coleridge, and Shelley, gave me great pleasure; and even as a school boy I took intense delight in Shakespeare. especially historical plays . . . 1 have also said that formerly pictures gave me considerable, and music very great delight. But now for many years I cannot endure to read a line of poetry; I have tried lately to read Shakespeare, and found it so intolerably dull that it nauseated me. I have almost lost my taste for pictures or music . . . My mind seems to have become a kind of machine for grinding general laws out of large collections of facts, but why this should have caused the atrophy of that part of the brain alone on which the higher tastes depend, I cannot conceive. Or, c o n s i d e r this: F a r a d a y d i s c o v e r e d t h e l a w s of e l e c t r o m a g n e t i c i n d u c tion a n d h i s d i s c o v e r i e s led h i m to f o r m u l a t e c o n c e p t s s u c h as " l i n e s of force" a n d "fields of force" w h i c h w e r e foreign to t h e t h e n p r e v a i l i n g m o d e s of t h o u g h t . T h e y w e r e i n fact l o o k e d a s k a n c e by m a n y of h i s c o n t e m p o r a r i e s . But of F a r a d a y ' s i d e a s , M a x w e l l w r o t e w i t h p r o p h e t i c discernment: The way in which Faraday made use of his idea of lines of force in coordinating the phenomenon of magneto-electric induction shows him to have been in reality a mathematician of a very high order—one from whom the mathematicians of the future may derive valuable and fertile methods. We are probably ignorant even of the name of the science which will be developed out of the materials we are now collecting, when the great philosopher next after Faraday makes his appearance. A n d yet w h e n G l a d s t o n e , t h e n t h e C h a n c e l l o r of t h e E x c h e q u e r , interr u p t e d F a r a d a y i n h i s d e s c r i p t i o n of h i s w o r k o n electricity b y t h e i m p a t i e n t i n q u i r y , " B u t after all, w h a t u s e is i t ? " F a r a d a y ' s r e s p o n s e w a s , " W h y , Sir, t h e r e is every p r o b a b i l i t y t h a t y o u w i l l s o o n b e able to tax it."
A n d Faraday's response has always been quoted most approvingly. It s e e m s to m e t h a t to D a r w i n ' s confession a n d to F a r a d a y ' s r e s p o n s e , w h a t S h e l l e y h a s s a i d a b o u t t h e c u l t i v a t i o n of t h e s c i e n c e s in h i s A Def e n c e of P o e t r y is a p p o s i t e : The cultivation of those sciences which have enlarged the limits of the empire of man over the external world, has, for w a n t of the p o e t i c a l faculty, p r o p o r t i o n a t e l y circumscribed those of the internal world; and man, having enslaved the elements, remains himself a slave. Lest y o u t h i n k t h a t S h e l l e y is n o t sensitive to t h e role of t e c h n o l o g y in m o d e r n society, let m e q u o t e w h a t h e h a s said in t h a t c o n n e c t i o n : Undoubtedly the promoters of utility, in this limited sense, have their appointed office in society. They follow the footstep of poets, and copy the sketches of their creations into the book of common life. They make space, and give time. S h e l l e y ' s A Defence of Poetry from w h i c h I h a v e just q u o t e d is o n e of t h e m o s t m o v i n g d o c u m e n t s in all of E n g l i s h literature. W. B. Yeats has c a l l e d it " t h e p r o f o u n d e s t essay o n t h e f o u n d a t i o n of poetry in t h e E n g l i s h l a n g u a g e . " T h e essay s h o u l d be r e a d in its entirety: b u t a l l o w m e to r e a d a selection: Poetry is the record of the best and happiest moments of the happiest and best minds. Poetry thus makes immortal all that is best and most beautiful in the world . . . arrests the vanishing apparitions which haunt the interlunations of life . . . Poetry is indeed something divine. It is at once the centre and circumference of knowledge; it is that which comprehends all science and to which all science must be referred. It is at the same time the root and blossom of all other systems of thought.
Poets are the hierophants of an unapprehended inspiration; the mirrors of the gigantic shadows which futurity casts upon the present; the words which express what they understand not; the trumpets which sing to battle, and feel not what they inspire; the influence which moves not, but moves. Poets are the unacknowledged legislators of the world. On reading Shelley's A Defence of Poetry, the question insistently occurs why there is no similar A Defence of Science written by a scientist of equal endowment. Perhaps in raising this question I have, in part, suggested an answer to the one I have repeatedly asked during the lecture. I began this lecture by asking your forbearance for addressing myself to matters which are largely outside the circumference of my comprehension. Allow me then to conclude by quoting from Shakespeare's epilogue to the second part of his Henry IV: First, my fear; then my curtsy; last my speech. My fear, is your displeasure, my curtsy, my duty, and my speech, to beg your pardon.
EINSTEIN AND GENERAL RELATIVITYHISTORICAL PERSPECTIVES Einstein's place in the physics of the 20th century is generally considered unique. And one may ask, "Why?" For one could name several whose fundamental contributions to the physicist's common stock of knowledge may be considered even more relevant than Einstein's—at any rate, comparable to his. Here are the names of some: Lorentz, Poincare, Rutherford, Bohr, Fermi, Heisenberg, Dirac, and Schrodinger. Einstein's contributions to that part of physics with which all students of physics, without exception, would be familiar are those derived from his three famous papers of 1905: dealing with his founding of the special theory of relativity, the theory of Brownian motion, and the concept of the photon. While all these contributions, singly, and even more, together, place Einstein among the foremost physicists of our time, one cannot be confident that, on these accounts, his place is one of exceptional uniqueness. After all, Lorentz, and even more Poincare, were not that far behind Einstein in formulating the principles of special relativity; and it is to Minkowski that we turn for the deepest formulation of the concepts of special relativity. Smoluchowski, independently of Einstein, discovered the theory of Brownian motion; and it is to Smoluchowski that we turn for the unravelling of all the multifarious aspects of the theory. And in the formulation of his concept of the photon, Einstein was preceded by Planck and followed by Bohr. And let us not forget the great figure of Poincare who looms behind so much of the mathematics of the 20th century and of physics indirectly. Why, then, is Einstein unique? To this question the answer undoubtedly is that besides all his contributions that have been enumerated, he was the sole and the lonely discoverer of the general theory of relativity. With that assessment, I agree. But Einstein's unique fame deriving from his development of the general theory of relativity has many paradoxical aspects. Perhaps the most striking of these is the exalted place which Einstein was given for his discovery of the general theory of relativity by some of the early investigators who were eminent men of science themselves and the benign neglect to which his theory was consigned by the professional scientific community for some fifty years, not to mention the active hostility to which his theory has been subjected over the years. The unravelling of the many conflicting strands of opinion l
with respect to general relativity is not an easy task. It is made somewhat easier for me since I share and endorse Hermann Weyl's description of general relativity "as the greatest example of the power of speculative thought."
Let me begin by describing in the most general terms the basic ideas which led Einstein to his theory of gravitation by the sheer power of his speculative thought. But it should be emphasized first, that Einstein's replacement of the Newtonian theory of gravitation by his own theory did not arise in any of the normal ways in which new physical theories emerge. It is almost invariably the case that new theories of physics, or novel generalizations of the old, result from a definite conflict with experience; and the ideas for the new theory are distilled from the need to incorporate the facts which appear conflicting with what is already known into a harmonious whole. Further, the successes of the new ideas are judged by the extent to which they can account for new phenomena. The general theory of relativity did not originate in this fashion. Einstein started with the premise that Newton's theory required a reformulation, since it was in manifest conflict with his own special theory of relativity. The basic tenet of the special theory of relativity is that in physics there can be no instantaneous action at a distance; and that no signal of any kind can be propagated with a velocity exceeding that of light. And, of course, Newton's laws of gravitation postulate instantaneous action at a distance. Besides, at the base of Newton's laws of gravitation is an enigmatic fact—well established, but not understood, before Einstein. The enigmatic fact goes back to Galileo's well-known demonstration from the leaning tower at Pisa, that all bodies, large or small, are accelerated equally in the local gravitational field of the earth. From this equality of acceleration of different masses, one concluded that mass as a measure of the quantity of matter and mass as a measure of its weight are identically the same. This identity is commonly referred to as the equality of the inertial and the gravitational mass of a body. But this equality has no theoretical basis: it is an empirical fact which requires experimental evidence. Newton was well aware of the need for experimental evidence for this crucial fact. Thus, in the opening paragraph of his Principia, Newton wrote:
"Air of double density, in a double space, is quadruple in quantity; in a triple space, sextuple in quantity. It is this quantity that I mean hereafter, everywhere, under the name of body or mass. And the same is known by the weight of each body, for mass is proportional to the weight, as I have found by experiments on pendulums very accurately made." (May I parenthetically note that this equality of inertial and gravitational mass to which Newton makes reference, has in recent years been established to an accuracy of one part in several billions.) The extremely general facts I have stated provided Einstein with the entire basis for his formulating the general theory. Even given these bases for generalizing Newton's theory, there really was no compelling need for an exact theory of gravitation. For, on general grounds, one could have argued that the Newtonian theory should be valid so long as the velocity of a planet or a satellite is small compared to the velocity of light. Even Mercury, the planet closest to the Sun, describes its orbit around the Sun with a velocity which is 6000 times smaller than the velocity of light. Accordingly, any departures from the predictions of the Newtonian theory can be estimated to be no more than a few parts in a billion. On this account, it would have been entirely sufficient to generalize the Newtonian theory to allow for such small departures which may arise from the finiteness of the velocity of light since we expect the Newtonian theory to be exact if the velocity of light could be considered as infinite. And that would have been the normal way. But Einstein did not proceed in that way. He searched for an exact theory which would be valid even if the velocities of the gravitating bodies approached that of light. Certainly, an exact theory, even if one should succeed in formulating it, could never be confirmed by experimental or observational features which, as I have stated, must be minute, by all criteria, in the solar system; and moreover, when Einstein sought for a new theory he had no prior conception as to what the nature of the departures may be that the new theory would be asked to account for. But as was stated by one of his early associates, Cornelius Lanczos, by a combination of constructive mathematical thinking, philosophical imagination, and a nonerring aesthetic sense,
Einstein arrived at his exact equations governing the theory of gravitation, a theory in which the three fundamental entities, space, time, and matter were unified. As a rule, Einstein generally refrained from any emotional exclamation marks in his publications; but he overcame his reticence in the concluding sentence of his first communication in November 1915 to the Berlin Academy in which he announced the basic equations of his theory. He wrote: "Scarcely anyone who has fully understood this theory can escape from its magic; it represents a genuine triumph of the method of the absolute differential calculus of Christoffel, Ricci, and Levi Civita." Hermann Weyl and Arthur Eddington, who wrote the first serious expositions of relativity, responded to Einstein's magic. Thus, in the preface to his Space, Time, and Matter, published in the spring of 1918, Weyl wrote: "It is as if a wall which separated us from the truth has collapsed. Wider expanses and greater depths are now exposed to the searching eye of knowledge, regions of which we had not even a presentiment. It has brought us much nearer to grasping the plan that underlies all physical happening." And in Eddington's Space, Time, and Gravity, published in 1920, the opening sentence reads: "By his theory of relativity, Albert Einstein has provoked a revolution of thought in physical science." Others of comparable eminence, who have studied general relativity and made contributions to it have written similarly/Thus, Landau and Lifshitz in their well-known book on Classical Theory of Fields in introducing the general theory of relativity, state that it "represents probably the most beautiful of all existing physical theories." And Dirac has said that Einstein's generalization of the special theory of relativity to include gravitation "is probably the greatest scientific discovery that was ever made." From these statements of the eminent men of science who have studied the theory of relativity and made important contributions to it, one might conclude that the general theory of relativity is an accepted theory and that only cranks would doubt its validity. But that 4
is not the case. A great number of eminent men have either given faint praise or have considered Einstein's theory as just plainly incorrect. Let me quote some varying shades of opinion. Max Born, who was an assistant of Einstein's in Berlin during the very years when Einstein was developing his general theory of relativity, in 1955, on the occasion of the 50th anniversary of Einstein's great paper on special relativity, stated: "The foundation of general relativity appeared to me then, and still does as the greatest feat of human thinking about Nature, the most amazing combination of philosophical penetration, physical intuition, and mathematical skill. But its connection with experience is slender. It appealed to me like a great work of art, to be enjoyed and admired from a distance." But what are we to make of this seeming praise of general relativity? Has it only to be admired from a distance? Does it not then require study and development like any other branch of the physical sciences? And a cynic might add that the description of Einstein's work as a work of art is often the cloak in which physicists disclaim the relevance of general relativity to the advance of physics. Here is J. J. Thomson: "Einstein has given a second theory known as 'general relativity' which includes the theory of gravitation. This involves much very abstruse mathematics, and there is much of it I do not profess to understand. I have, however, a profound admiration for the masterly way in which he has attacked a problem of transcendent difficulty." And here is Rutherford: "The theory of relativity of Einstein, quite apart from its validity, cannot but be regarded as a magnificent work of art." The description of general relativity as a work of art is doubleedged. One senses that in describing the theory in this way, one is trying not to dissociate oneself from the general acclaim that is accorded to Einstein. But the matter is not that simple. It is, in fact, the case that the literature dealing with theories alternative to Einstein's are as numerous as the positive contributions devoted to exploring the content of the theory itself. It is not merely that cranks and 5
pseudoscientists have written tracts disputing Einstein. Several eminent men of science, whom we all respect, have also considered Einstein's theory as scientifically unsound. Let me list the names of some of those who have written books and tracts presenting theories which they consider as viable alternatives to Einstein's theory. Alfred North Whitehead, the distinguished philosopher-mathematician; George Birkoff, the distinguished mathematician; E. A. Milne, the distinguished astrophysicist; and Hoyle and Narlikar. Besides, several earlier adherents of Einstein's theory have now discerned either flaws or grave crises; e.g., Nathan Rosen and Christian Miller. While I shall not give any account (I could not do so dispassionately), I will quote a few sentences from Whitehead, whose many writings on science and philosophy many of us have admired. In 1922, Whitehead published a book entitled, The Principle of Relativity, an alternative to Einstein's theory. Whitehead starts by quoting with approval an aphorism of J. J. Thomson: "I have no doubt whatever that our ultimate aim must be to describe the sensible in terms of the sensible." After this quotation, Whitehead goes on to say, "I do not agree with Einstein's way of handling his discovery," meaning that Einstein's theory does not describe sensible things in a sensible way. Here are some other quotations from Whitehead which will give you a flavor of his reasoning: "So many considerations are raised that we are not justified in accepting blindfolded the formulation of principles which guided Einstein." Or, again: "In the comparative absence of applications, beauty, generality, and even Truth, will not save a doctrine from neglect in scientific thought. ... To expect to reorganize our ideas on Time, and Space, and Measurement without some discussion, which must be ranked as philosophical, is to neglect the teaching of history and the inherent possibilities of the subject." Convinced that Einstein's formulation of space-time as a Riemannian manifold with a metric is invalid on philosophical grounds, Whitehead goes on to develop a detailed theory of his own. But unfortunately for Whitehead, some of the implications of his theory have been shown to be blatantly contrary to experience in several instances in which Einstein's theory succeeds admirably. Whitehead's 6
philosophical acumen has not served him well in his criticisms of Einstein. I said a little earlier that I was making an exception of Whitehead's criticisms of Einstein. But I cannot desist quoting the reaction to relativity of one of my personal friends, the late Professor E. A. Milne. In developing his kinematical theory of relativity, Milne stated: "Einstein's law of gravitation is by no means an inevitable consequence of the conceptual basis given by describing phenomena by means of a Riemannian metric. I have never been convinced of its necessity .... General relativity is like a garden where flowers and weeds grow together. The useless weeds are cut with the desired flowers and separated later." Milne goes on to say, referring to his own theory: "In our garden we grow only flowers."
I think that I have stated enough to convince you that the general theory of relativity did not receive acceptance from many respected scholars, who have apparently tried to understand the theory. But one may ask, "What was the attitude of a serious physicist (of the period say from 1925 to 1965) to general relativity?" Born's remark that the connection of general relativity with experience is slender is representative. I have been told that at a dinner in honor of Einstein's 70th birthday, held at the Institute for Advanced Study in 1949, Oppenheimer made remarks to the effect that general relativity had been singularly without influence in the development of physics during the period 1925-1950. Indeed, general relativity as a discipline in physics was simply ignored or at any rate neglected benignly in most institutions devoted to its study. As an illustration of this fact I might refer to the circumstance that from 1936, when I joined the faculty of the University of Chicago, to 1961, no courses in general relativity, not even for one single quarter, were given at the University. And the University of Chicago is not atypical. And I am not sure how well the principles of general relativity, as laid by Einstein, are appreciated by the common physicists of today. Where, then, does Einstein's fame come from?
7
It will be presumptuous of me to suggest an answer of my own to the question I have just raised. But I will give an answer given by Rutherford, during a conversation 45 years ago, at which I was present. The conversation took place in the Senior Combination Room in Trinity College, after dinner, during the Christmas recess of 1933. During the Christmas recess, very few people normally dine in the College. On this particular occasion there were only five of us: Lord Rutherford, Sir Arthur Eddington, Sir Maurice Amos (at one time, during the 1920s, the Chief Judicial Advisor to the Egyptian government), Dr. Patrick DuVal (a distinguished geometer), and myself. After dinner, we all sat around a fire and in the ensuing conversation Rutherford was in great form. At some point during the conversation, Sir Maurice Amos turned to Rutherford and said: "I do not see why Einstein is accorded a greater public acclaim than you. After all, you invented the nuclear model of the atom; and that model provides the basis for all of physical science today and it is even more universal in its applications than Newton's laws of gravitation. Whereas, granted that Einstein's theory is right—I cannot say otherwise in the presence of Eddington here—Einstein's predictions refer to such minute departures from the Newtonian theory that I do not see what all the fuss is about." Rutherford, in response, turned to Eddington and said: "You are responsible for Einstein's fame." And more seriously, he continued: "The war had just ended, and the complacency of the Victorian and Edwardian times had been shattered. The people felt that all their values and all their ideals had lost their bearings. Now, suddenly, they learned that an astronomical prediction by a German scientist had been confirmed by expeditions to Brazil and West Africa and, indeed, prepared for already during the war, by British astronomers. Astronomy had always appealed to public imagination; and an astronomical discovery, transcending worldly strife, struck a responsive cord. The meeting of the Royal Society, at which the results of the British expeditions were reported, was headlined in all the British papers; and the typhoon of publicity crossed the Atlantic. From 8
that point on, the American press played Einstein to the maximum." I could see from Eddington's reaction that he agreed with Rutherford; and he, in turn, recalled some events of that time. Let me go back a little to tell you about the circumstances which gave rise to the planning of the British expeditions. I learned of the circumstances from Eddington (in 1935) when I expressed to him my admiration of his scientific sensibility in planning the expeditions during "the darkest days of the war." To my surprise, Eddington disclaimed any credit on that account—indeed, he said that, left to himself, he would not have planned the expeditions since he was fully convinced of the truth of the general theory of relativity! And he told me how the expeditions came about. In 1917, after more than two years of war, England enacted conscription for all able-bodied men. Eddington, who was then 34, was eligible for draft. But as a devout Quaker, he was a conscientious objector; and it was generally known and expected that he would claim deferment from military service on that ground. Now the climate of opinion in England during the war was very adverse with respect to conscientious objectors: it was, in fact, a social disgrace to be even associated with one. And the stalwarts of Cambridge of those days—Sir Joseph Larmor (of the Larmor precession), Professor H. F. Newall, and others—felt that Cambridge University would be disgraced by having one of its distinguished members a declared conscientious objector. They therefore tried through the Home Office to have Eddington deferred on the grounds that he was a most distinguished scientist and that it was not in the long-range interests of Britain to have him serve in the army. (The case of H. G. J. Moseley, who discovered the concept of atomic number and who was killed in action at Gallipoli, Turkey, was very much in the minds of the British scientists at that time.) And Larmor and others nearly succeeded in their efforts. A letter from the Home Office was sent to Eddington, and all he had to do was to sign his name and return it. But Eddington added a postscript to the effect that, if he were not deferred on the stated grounds, he would claim it on conscientious objection anyway. This postscript, naturally, placed the Home Office in a logical quandary: a confessed conscientious objector "had to be sent to a camp." Larmor
and others were annoyed. Eddington told me that he could not understand their annoyance; and as he expressed himself, many of his Quaker friends found themselves in camps in Northern England "peeling potatoes" for holding the same convictions and he saw no reason why he should not join them. In any event, at Sir Frank Dyson's intervention—as the Astronomer Royal, he had close connections with the Admiralty—Eddington was deferred with the express stipulation that if the war should have ended by 1919, he should lead one of two expeditions that were being planned for the express purpose of verifying Einstein's prediction with regard to the gravitational deflection of light. In any event, Eddington clearly realized the importance of verifying Einstein's prediction with regard to the deflection of the light from distant stars as it grazed the solar disk during an eclipse. It is best that I continue the story in Eddington's own words. "In a superstitious age a natural philosopher wishing to perform an important experiment would consult an astrologer to ascertain an auspicious moment for the trial. With better reason, an astronomer today consulting the stars would announce that the most favorable day of the year for weighing light is May 29. The reason is that the sun in its annual journey round the ecliptic goes through fields of stars of varying richness, but on May 29 it is in the midst of a quite exceptional patch of bright stars—part of Hyades—by far the best starfield encountered. Now if this problem had been put forward at some other period of history, it might have been necessary to wait some thousands of years for a total eclipse of the sun to happen on the lucky date. But by strange good fortune an eclipse did happen on May 29, 1919. "Attention was called to this remarkable opportunity by the Astronomer Royal (Sir Frank Dyson) in March 1917; and preparations were begun by a committee of the Royal Society and the Royal Astronomical Society for making the observations. "Plans were begun in 1918 during the war, and it was doubtful until the eleventh hour whether there would be 10
any possibility of the expeditions starting. Two expeditions were organized at Greenwich by Sir Frank Dyson, the one going to Sobral in Brazil and the other to the Isle of Principe in West Africa. Dr. A. C. D. Crommelin and Mr. C. Davidson went to Sobral; and Mr. E. T. Cottingham and the writer went to Principe. "It was impossible to get any work done by instrument makers until after the armistice; and as the expeditions had to sail in February, there was a tremendous rush of preparation. The Brazil party had perfect weather for the eclipse; through incidental circumstances, their observations could not be reduced until some months later, but in the end they provided the most conclusive confirmation. I was at Principe. There the eclipse day came with rain and cloudcovered sky, which almost took away all hope. Near totality, the sun began to show dimly; and we carried through the program hoping that the conditions might not be so bad as they seemed. The clouds must have thinned before the end of totality, because amid many failures we obtained two plates showing the desired star-images. These were compared with plates already taken of the same starfield at a time when the sun was elsewhere, so that the difference indicated the apparent displacement of the stars due to the bending of the light-rays in passing near the sun. "As the problem then presented itself to us, there were three possibilities. There might be no deflection at all; that is to say, light might not be subject to gravitation. There might be a "half-deflection," signifying that light was subject to gravitation, as Newton had suggested, and obeyed the simple Newtonian law. Or there might be a "full deflection," confirming Einstein's instead of Newton's law. I remember Dyson explaining all this to my companion Cottingham, who gathered the main idea that the bigger the result, the more exciting it would be. 'What will it mean if we get double the deflection?' 'Then,' said Dyson, 'Eddington will go mad, and you will have to come home alone.' "Arrangements had been made to measure the plates on the spot, not entirely from impatience, but as a precaution
against mishap on the way home, so one of the successful plates was examined immediately. The quantity to be looked for was large as astronomical measures go, so that one plate would virtually decide the question, though, of course, confirmation from others would be sought. Three days after the eclipse, as the last lines of the calculation were reached, I knew that Einstein's theory had stood the test and the new outlook of scientific thought must prevail. Cottingham did not have to go home alone." It was some months before the two expeditions returned to England and the participants were able to measure their plates and collate their results. But rumors of the successful confirmation of Einstein's prediction reached Einstein in early September 1919. And on September 22, 1919, the Dutch physicist, Hendrik Antoon Lorentz, sent Einstein a telegram confirming the rumors to which Einstein replied (also by telegram), "Heartfelt thanks to you and Eddington. Greetings." Einstein's own satisfaction with the outcome of the British expeditions is shown by the postcard (dated September 27, 1919) to his ailing mother in Switzerland. It said: "Dear Mother: Good news today. H. A. Lorentz has wired me that the British expeditions have actually proved the light deflection near the Sun..." There is a further anecdote relative to Einstein's reaction to the news from England which I should like to recall. Use Rosenthal-Schneider, a student of Einstein in 1919, recalls that Einstein showed her the cable from Eddington informing him of the successful verification of his prediction. And she asked him, what if there had been no confirmation of his prediction? Einstein's response was: "Then I should have been sorry for the dear Lord; but the theory is correct." The Times of London for November 7, 1919, carried two headlines: "The Glorious Dead, Armistice Observance. All trains in the Country to Stop," and "Revolution in Science. Newtonian Ideas Overthrown." The second of these headlines referred to the meeting of the Royal Society in London on November 6 at which Dyson had reported on the results of the British expeditions. 12
Alfred North Whitehead's account of this meeting of the Royal Society has often been quoted. It is worth quoting once again. "The whole atmosphere of tense interest was exactly that of the Greek drama: we were the chorus commenting on the decree of destiny as disclosed in the development of a supreme incident. There was dramatic quality in the very staging—the traditional ceremonial, and in the background the picture of Newton to remind us that the greatest of scientific generalizations was now, after more than two centuries, to receive its first modification. Nor was the personal interest wanting: a great adventure in thought had at length come safe to shore." The meeting of November 6, 1919, of the Royal Society also originated a myth that persists even today (though in a very much diluted version): "Only three persons in the world understand relativity." Eddington explained the origin of this myth during the Christmas recess conversation with which I began this account. Sir J. J. Thomson, as President of the Royal Society at that time, concluded the meeting with the statement: "I have to confess that no one has yet succeeded in stating in clear language what the theory of Einstein's really is." And Eddington recalled that as the meeting was dispersing, Ludwig Silberstein (the author of one of the early books on relativity) came up to him and said: "Professor Eddington, you must be one of three persons in the world who understands general relativity." On Eddington demuring to this statement, Silberstein responded, "Don't be modest Eddington." And Eddington's reply was, "On the contrary, I am trying to think who the third person is!" The myth that general relativity is a difficult theory to understand originated at this time. It is a myth which has done immeasurable harm to the development of the theory. The fact is that the theory of general relativity is no more difficult than many other branches of physics. General relativity, at the time it was founded, required familiarity with a mathematical discipline which physicists had not encountered before that time. But that has also been the case with several other branches of physics, including quantum mechanics. I cannot conclude this account, relating to the verification of Einstein's prediction concerning the deflection of light by a gravitational field and how that was responsible for his becoming "a focus of 13
widespread adoration," without remarking that history might well have been very different. In 1911, Einstein had calculated, on the basis of his principle of equivalence, the deflection that light grazing an object, such as the Sun, will experience. The principle of equivalence correctly accounts for the slowing of a clock in a gravitational field; but it gives for the deflection of light only half the value predicted by general relativity. Roughly speaking, one might say that one-half of the predicted effect is the result of the slowing down of the time-measuring process; and that the other half is due to the spatial curvature of space-time. The latter effect is an essential aspect of general relativity. The German astronomer, Finlay Freundlich, had planned to test Einstein's prediction of 1911 at an eclipse of the Sun which occurred in Russia in 1914. But the war intervened; and Freundlich was unable to make the observations he had planned. As Hoffmann and Dukas have said: "Suppose the war had not come and Finlay Freundlich had been able to observe the 1914 eclipse and had found a deflection of 1.7 seconds of arc at a time when Einstein was predicting a deflection of only 0.83 seconds of arc. Imagine how tame Einstein's 1915 calculation of 1.7 seconds of arc would have seemed. ... He would have been belatedly changing the value after the event, having first been shown to have been wrong ... and the deflection of light would have lost the tremendous impact that it had as a prediction." But the war had in fact intervened; and the predicted deflection had been confirmed under circumstances described by Rutherford. And as Eddington wrote to Einstein in December 1919: "All of England has been talking about your theory. It has made a tremendous sensation. ... It is the best possible thing that could have happened for scientific relations between England and Germany."
Are we to conclude then that the unique place in physics which Einstein is accorded is an accident of circumstances? I do not think so. The testimony for his uniqueness comes from those who, serious students of science themselves, are caught in the web of the magic of 14
Einstein's theory and feel, as Hermann Weyl felt, that a wall obscuring truth collapses when we explore the richness of his theory. In saying this, I do not wish you to conclude that those who marvel at the content of Einstein's theory form a cult of some sort. That would be the case if one's admiration for the theory was derived from a distance, as Born has stated, or as a "work of art," as Rutherford and J. J. Thomson have stated. The simple fact is that Einstein's theory is incredibly rich in its content and presents glittering faces at every turn. Let me be specific and illustrate what I mean in a concrete way.
I am sure that all of you are familiar with the role black holes have been publicized to play in current astronomical developments. Let me say at once that I do not associate myself with those who consider black holes as exotic objects predicted by general relativity. Exotic means grotesque, bizarre: and there is nothing grotesque or bizarre about black holes. We are all aware that a body projected from earth, for example, cannot escape the earth's gravitational field unless it is projected with a sufficient speed; otherwise, it will simply fall back. Once we grant that light is deflected by a gravitational field—as, indeed, we must—then it is a matter of simple arithmetic to calculate how strong the gravitational field must be if a particle, projected with a velocity equal to that of light, cannot escape. This calculation was, in fact, made by Laplace as long ago as 1798, even though he had no reason to suspect that light is affected by gravity. We have seen that light grazing the Sun is deflected by the minute amount of 1.7 second of arc. But if the Sun with its present radius of 700,000 km should be compressed to a sphere of radius 2-1/2 km, then at that radius, the gravitational field would be strong enough to prevent light from escaping from the surface; and we should cease to see it: it would have become a black hole. The contraction of a star with a solar mass to a radius of 2-1/2 km does not require us to postulate physical conditions with which we are not familiar. The mean density of matter at that radius is no different than in ordinary atomic nuclei. The physical conditions required for the occurrence for stellar masses to become black holes are, therefore, entirely within the realm of reason. The question is, rather, whether 15
such physical conditions can be realized in Nature and in the natural course of events. Let me categorically state that very simple and quite elementary considerations relating to the last stages in the evolution of massive stars—i.e., stars with masses exceeding say five solar masses—require the formation, barring accidents in every case, of black holes. This is a story which is fascinating in itself, but it is not the story I wish to tell now. I want, rather, to turn to what the general theory of relativity has to say with regard to black holes. One might think that if all that is required for black holes to occur are sufficiently strong gravitational fields, then black holes of diverse shapes, forms, and sizes should be possible. For example, the external shape and size that a gravitating object can have in Newtonian theory are infinitely diverse: they will depend on the mass, the stratification of density, and temperature in the interior, whether it is rotating or not; and if it is rotating, whether it is rotating uniformly or not, and how fast it is rotating; and a whole variety of other factors. But very remarkably, according to the general theory of relativity, black holes belonging to one family and of only one kind can occur. This family of solutions was discovered by Roy Kerr, a New Zealand mathematician, in 1962. Kerr's solutions provide the basis for an exact representation of all black holes that can occur in the astronomical universe. It is not only that Kerr's solution is unique: there is an explicit formula which one can write for it. This is a startling and an unexpected consequence of the general theory of relativity. Besides, the geometry of the space-time around a Kerr black hole has many remarkable features. For example, in Kerr geometry all the standard equations of mathematical physics can be solved exactly; and we are also led to mathematical identities of a kind one had never suspected. I do not know to what extent the foregoing remarks appear convincing to you. But the point I wish to make is that the general theory of relativity is incredibly rich in its content; and as I said, one finds a glittering face at almost every turn. Einstein was certainly correct in his prediction of 1915 that "anyone who genuinely understands the theory" cannot escape from its magic. If some remote and presently unforeseen development requires a modification of Einstein's theory within its own well-defined framework of validity, then we should indeed have reason to "be sorry for the dear Lord." 16
KARL SCHWARZSCHILD LECTURE
THE AESTHETIC BASE OF THE GENERAL THEORY OF RELATIVITY
S.
Chandrasekhar
University of Chicago Chicago.
Illinois 60637
U.S.A.
May
I
begin
by
expressing
my
gratitude
to
the
officers
and
councillors of the Astronomische Gesellschaft for their courtesy in asking to
give
this
lecture
in
the
series
established
in
memory
of
me Karl
Schwarzschlld.
I.
Karl
Schwarzschlld
scientists of
this
century.
staggering:
they
cover
is.
The the
of
course,
breadth entire
and
range
one
of
range of
of
the his
physics,
towering
physical
contributions
are
astronomy,
and
astrophysics of his time.
In
physics,
they
range
from
electrodynamics
and
geometrical
optics to the then newly developing atomic theory of Bohr and Sommerfeld. In electrodynamics, the
electron.
aberrations
In
he derived a variational geometrical
in optical
in clarity and rigour
optics,
instruments
he
(described
by later work")
base for developed
Lorentz's equations the
by Max Born
theory as
of
of the
"unsurpassed
and formulated the principle
underlying
the
optics
of
the
he worked out. and
of
the
Schmidt
And
in his last published
Desiander
molecules.
telescope.
term
in the
in
paper,
the
theory.
the theory of the Stark
rotational-vibrationai
(In this last paper he i n t r o d u c e d ,
of action and angle
Bohr-Sommerfeld
spectra
of
for the first t i m e ,
effect
diatomic
the
notions
variables.)
In Astronomy
and
Astrophysics,
Schwarzschild's
contributions
are
so many and so varied that I shall mention only those of his discoveries
to
which
in
his
name
photographic theory
of
convective velocities, for
is
attached.
photometry,
radiative
instability, and.
describing
and of static
the
transfer,
the
have
the
Schwarzschiid
Schwarzschild-Mllne the
the
of c o u r s e ,
We
Schwarzschiid
Schwarzschiid
integral criterion
ellipsoidal
exponent
equation for
the
distribution
of
the Schwarzschiid solution of Einstein's
space-time
external
to
a
spherical
in
the
onset
of
stellar
equations
distribution
of
mass
black-holes.
And all of these in a brief twenty years!
It is possible some
of
you:
it
is
that the not
announced
addressed
lectures in this series have b e e n : any
astronomical
Schwarzschild's discern
them
overtone.
attitude from
his
and
to
title
any
of
my lecture
concrete
topic
has
as
puzzled
the
earlier
and I am afraid that it will scarcely
My
lecture,
approach
to
published
papers:
his solution of the equations of general
however,
scientific and
it will
will
have
bear
problems,
as
bear
directly
very
I
on can on
relativity.
II. I shall to my m i n d .
consider
The first was
three
examples
from
Illustrate his approach to scientific
discovered
almost at o n c e ,
by
Schwarzschild's
relates to his work on s t a r - s t r e a m i n g . J.C.
Kapteyn:
and
it
work
was
The
phenomenon
adequately
interpreted,
by Eddington on the basis of his and Kapteyn's
of two s t a r - s t r e a m s .
Schwarzschiid
expressed
which.
problems.
his reaction to this
hypothesis hypothesis
as follows : The magnitude along
with
would
appear
motion
of the proper
them,
presumably
also,
be
equal.
The
each
other,
must,
to
through
and It Is problematical I
have,
therefore,
Eddington
has
of the stars their
stars
average In
the
therefore,
believed
that
the
himself stellar
these
same
should
of
argument:
grounds.
stellar
the
common
reworked
Schwarzschild
discussions
systems.
a second
Schwarzschild. addressed
At
a
himself
three-dimensional
What
example.
meeting to
of
the
space
their
fluctuations;
material
on
a
on
which
more
unified
formulated
is
his
ellipsoidal
and this formulation has
bearing
on stellar
remarkable
to
motions
me,
of
I shall this
a
Gesellschaft
question astronomy
take
whether might
be
still
the
is
his
however,
in
earlier
hoc.
publication
1900,
the
been
and
a description of nature must be natural: it cannot be ad
As
sun
during
about.
observational
be
from
and
motions.
basis of all subsequent
dynamics
distances
share
distribution of the peculiar velocities of stars; the
In the two streams
two streams,
how this can be brought
based
hypothesis concerning
On
motions
of
Schwarzschild
geometry
non-Euclidean.
of He
the stated
the problem as follows. As must be known to you, one
has
developed
during
non-Euclidean
this century geometry
the chief examples
of which are the so-called
spaces.
wonder
We can
how the world
pseudo-spherical
geometry
with possibly
would
find
oneself.
If one
know
whether
does
then not
realized
in
the
will.
beauty
[meaning
(besides spherical
radius
and
this
fairyland
century] geometry),
pseudo-spherical
in a spherical
or a
of curvature..
..
One
fairyland;
and
one
In a geometrical of
19th
Euclidean
would appear
a finite
the
may
not
In
fact
be
nature.
We
can
only
curiosity in addressing the founding
marvel
of general
simple imagination.
at
Schwarzschild's
scientific
imagination
himself to such a question some fifteen years relativity.
But to Schwarzschild.
it was
more
and
before than
He actually estimated limits to the radius of curvature of
This and the other translations from German of the originals Einstein's "GedSchtnisrede" in the appendix) are the author's.
(including
the t h r e e - d i m e n s i o n a l
space with the astronomical data available at his time
and concluded that if the space is hyperbolic
its radius of curvature
cannot
be less than 64 light years and that if the space is spherical its radius of curvature must at least be 1600 light years.
We is
far
need
more
not
argue
relevant
that
about
Schwarzschild's
Schwarzschild
particular
allowed
his
estimates.
It
imagination
to
contemplate a world that may have features of a fairyland!
My solution central
third
of
example
Einstein's
spherical
bears
on
Schwarzschild's
vacuum-equations
distribution
of
appropriate
mass
-
discovery
to
undoubtedly
the
of
exterior
the
most
the of
a
important
discovery in relativity after its founding.
Schwarzschild's communicated just
about
by
two
paper
Einstein months
to
in
the
after
which
he
Berliner
Einstein
derived
Akademie
himself
had
equations of his theory in a short communication full
derivations
theoretical
was
rate
still
of the
six
months
precession
in the
of the
-
future
his
in
solution
1916
was
January
published
13.
the
basic
his detailed paper -
and
perihelion
of
had
with
deduced
Mercury
and
of
the the
magnitude of the deflection a light ray will experience as it grazes the
limb
of the sun.
1916
In acknowledging Schwarzschild's
paper.
Einstein wrote on
January 9. / have could
read
your
paper
with
obtain
the
exact
solution
treatment
of the problem
The famous
at
the
this
period
relativity
-
were
front
a fatal of
with
a
that
second
to me
one
small and
the
spring
Schwarzschild
and
army at the
technical
with
not
expected
simply.
Schwarzschild
he died
dealt
so
that
The
one
analytical
splendid.
which
During
I had
problem
In the German
disease:
illness
the
interest, the
under
heroic.
was serving
Eastern
pemphigus,
of
appears
circumstances
solution
Schwarzschild
greatest
staff
summer
1916
wrote
his
two
now
of
1915. While
May
equilibrium
his
Eastern front.
Schwarzschild
on
the
derived
contracted
1 1 . It was papers of
a
on
during general
homogeneous
mass and showed that no hydrostatic equilibrium is possible if the radius of the
object
is
less
than
9/8
of
the
Schwarzschild
radius.
2 GM/c2
-
and
the fundamental one on the Bohr-Sommerfeld theory to which i have already
referred. About
Schwarzschlld's
last
Illness.
Eddington
wrote
in
a
moving
obituary notice: His end Is a sad story of long suffering
from a terrible
the field,
patience.
borne with great courage
Parenthetically. the
late
thirties
that
I may add a footnote.
he
Eastern front while he,
had
met
as
Karl
Illness contracted
distinguished
In
Richard Courant told me in
Schwarzschlld
proceeding
as a member of the general staff,
retreating from the same front; someone
and
to
the
was with a party
and Courant said that he was surprised that
as
Karl
Schwarzschlld
would
be
proceeding
towards a front that was considered too dangerous for the general staff!
Let for
me
seeking
earlier
an
by
return exact
an
statement:
It
Is
Is
even
uniqueness be
always
Mr.
below.
his
Einstein argued
approximate
Schwarzschlld. realized
was
"undaunted" afterwards the
in
of
an
the
exact
of
present
the
his
Einstein
reasons
had
began
solution
solved
with
In a
Instance,
to
the
simple
have
the
whatever doubts there may
problem,
since,
in the nature of this problem,
in
his
letter
of
that there can in
the
for some
solving
whished
fundamental view
and
as
It
will
to establish
the
procedure.
undaunted,
(and
which
and remove
procedure
a
to obtain
treatment
paper
Schwarzschlld
established
It is difficult,
While
problem
In
Einstein's
earlier)
the
Important,
validity of an approximate
(quoted
to
original
procedure.
satisfying
more
of the solution
concerning
appear
solution
approximate
form.
It
to Schwarzschlld's
one
acknowledgement
the
equations,
it
to solve exactly the in
great
decades,
to
Schwarzschlld
be no doubts about the validity of
the
newly
much-ado to the
is
problem
formulated
that
was
detriment
significant
to
which
theory. be
I
made
of the theory)
"difficulty" of Einstein's theory in general and of finding exact
that he said soon about
solutions
In particular.
I
shall
understanding
return
of general
subject of my Lecture.
later
to
relativity.
the
role
of
exact
But I must pass on
solutions
for
now to the
the main
III. The
general
theory
extremely beautiful theory; (as
by
Rutherford
and
of
relativity
has often
been
described
as
an
and it has even been compared to a work of art by
Max
Born,
for
example).
In
the
same
vein.
statements like the following ones by Dlrac are not uncommon: What
makes
against
the
theory
the principle
The
Einstein
own.
(1978)
These
and
acceptable
of simplicity,
theory
similar
so
of
to physicists,
In spite
Is Its great mathematical
gravitation
has
characterizations
a
character
of the general
of
Its
beauty.
of
going
(1939)
excellence
of
theory of relativity
its
raise
the following questions:
What
Is
what
extent
the
aesthetic is an
formulation
and
base
of
aesthetic solution
the
theory?
sensibility
of
to
problems
And.
its
more
excellence
which
will
importantly, relevant
lead
to
to
to
a
the
deeper
understanding of the theory?
To answer first
necessary
theory
of
to
relativity
experiment;
and
this question
appreciate with the
without descending
the
present
respect
reasons
to for
peculiar
its its
to dilettantism.
position
confirmation inspiring
by
of
the
It is
general
observation
confidence
in
and
spite
of
inadequate empirical support.
During the past twenty years a great deal of commendable has
been
expended
to
verify
the
lowest
first-order
departures
Newtonian theory that the general theory of relativity predicts. have
been
differing deflection the
of
and
the
time-keeping
predictions in
light ray experiences
to the
eccentric
and
a
consequent
finally,
been
successful
rates
time
delay;
slowing
to
of
locations
when the
down of the
the of
traversing
precession orbital
theory
differing
from
period
a
relating
to
the
gravity:
to
the
Kepler
of a
field
orbit;
binary
star
orbit by virtue of the emission of gravitational radiation,
confirmed
within
uncertainties.
But
the all
limits these
of
observational
effects
relate
and to
the
These efforts
a gravitational of
effort
departures
in
an
have all
experimental
predictions of the Newtonian theory by a few parts in a million;
and and.
errors
from
the
and of no
more than three or four Einstein field-equations.
parameters
in a post-Newtonian
And. so far.
expansion
of the
rvo predictions of general relativity,
the limit of strong gravitational fields,
have received any confirmation;
in and
none seems likely in the foreseeable future.
Should generalizes
a
one
not
argue
theory
as
well
Newtonian theory, of the which
theory,
tested
confirmation
in
its
domain
of
a
theory,
of
validity
rather
than
Would
to small
first-order
the status of
Dirac's
departures
from
the
which as
the
aspects theory
theory of the electron,
for
be what it is today if its only success consisted in accounting for
Paschen's Ionized
a
should refer to predictions which relate to major
it replaces?
example,
that
1916 measurements of the fine structure splittings of the lines of
helium?
The
real
confirmation
confidence was the discovery,
of
Dirac's
theory
that
in accordance with the theory,
positron pairs In cosmic-ray showers.
Similarly,
inspired
of electron-
would our faith in Maxwell's
equations of the electromagnetic field be as universal as it is without Hertz's experiments
on the propagation
velocity of light and without transformations?
In
the
of electromagnetic
Polncare's
same
way.
the theory,
and only of that theory,
proof of their a
theory of relativity will be forthcoming
waves with precisely
real
invarlance
confirmation
of
to
the
relativity
of their in
any
evolution real
is not a confirmation
sense.
The
notion
that
of a light
of
The occurrence of black
holes as one of the final equilibrium states of massive stars in the course
Lorentz general
only if a prediction characteristic
is confirmed.
the
prediction
cannot
of
escape
natural general from
a
sufficiently strong gravitational field is an inference not based on any exact prediction of the theory; affected by gravity.
it depends only on the empirical fact that light is
On the other hand, since the general theory of relativity
provides an exact description of the space-time around black holes, confirmation
of
the
metric
of
the
space-time
around
black
considered as "establishing* the theory in any real sense. that the Kerr solution with two parameters stationary black holes that can occur
holes
only a can
be
It is well known
provides the unique solution for
in the astronomical
confirmation of the metric of the Kerr space-time
(or.
universe.
But a
some aspect of It)
cannot even be contemplated in the foreseeable future.
Perhaps.
I may digress here to indicate how one may eventually
have a confirmation of the space-time around a rotating Kerr black hole.
If
one Imagines the
a Kerr black hole with an accretion
equatorial
plane,
then
the
polarization
of
disc of free electrons
the
light
emerging
after traversing the strong gravitational field of the black hole, so
non-uniform
Nature
a
distribution
be generous
enough
that to
one
should
provide
be
a clean
able
to
example
from
will
it.
manifest
map
which
in
it.
Will
will
enable
such a mapping? I am afraid that this is the only time my talk will bear on an astronomical observation.
IV As
I have said,
we
have,
as yet.
no
relativity that has been confirmed by observation; In the foreseeable future. theory? have,
One that
should our
exact
feature
of
general
and none appears
feasible
Why then do we have faith and confidence in the
respond
confidence
more
explicitly
derives
from
than
the
merely
"beauty
to
of
say.
the
as
some
mathematical
description of nature which the theory provides".
So
said
To the solid
ground
Wordsworth.
There
relativity.
On what then
of Nature
is
do we
no
trusts
solid
build
our
the mind
ground trust?
for
that builds
the
for
general
We build
aye!
theory
our trust
on
of the
internal consistency of the theory and on its conformity with what we believe are
general
physical
requirements;
and.
above
all.
on
its
freedom
from
contradiction with parts of physics not contemplated in the formulation of the theory.
Let me illustrate by some examples.
The causal complete
initial
determined In-going formally
in
null
data the
character on
a
space-time
rays emanating
stated:
of the
the
basic
laws of physics
space-like
from
3-surface,
domain the
equations
bounded
boundary
of
initial-value formulation which determines
any
requires
the by
of the
physical
future the
that, Is
given
uniquely
future-directed
spatial slice.
theory
must
More
allow
uniquely the future development
the entire domain of dependence of the initial data on a spatial slice. field equations proof
of
this
Lichnerowicz.
of general fact
is
relativity do allow such a formulation
not straightforward:
it was
provided
only
though in
1944
an in The the by
As
a second
central to physics. that
is
static
globally or
energy,
implies of
that
fiat
asymptotically
consider
The
fact,
emit
But
for the
flat
at
then one
to
mass-function infinity
the
isolated
waves
expect
that, one
(extending
if
(i.e..
which
(e
so
is
not
to
the
have a
local
space-time
should
to infinity)
as
and,
further,
be
is
able
a total
to
energy
we
go
-
the
to
infinity
Bondi
of gravitational
waves.
positive finally emerged
tensor,
T|j,
is
along
mass -
that
that the rate of decrease of
But a proof only
1 9 8 1 , and Horowitz & Perry
For a perfect f l u i d ,
is
energy
contribute
expect to
sense)
space
body
that
should
null-Infinity
form
that
Bondi was able to show that if the s p a c e - t i m e
energy-momentum
Tjj =
an
can define a mass-function
remains
& Yau 1 9 8 1 , Wltten that
energy
is exactly equal to the rate at which energy
in the
mass always
of
well-defined
is a decreasing function of t i m e : this
notion
relativity we cannot
entire
In 1962.
that
gravitational
one
some
the
one is accustomed to define a local
in general
(in
globally,
that is positive.
null-rays).
will
energy.
asymptotically define,
conserved.
stationary
definition
example,
In physics,
satisfy
is
that the
in recent years 1982).
some
radiated
The proof "energy
Bondi
(Schoen requires
conditions".
for which + p)
UjUj -
pg|j .
the required conditions are equivalent to e >
|p|
.
The foregoing two examples, relativity.
Illustrate
obvious or
not
internal
consistency
-
a
consistency
by
no
means
self-evident.
An relativity
its
deep in the structure of the general theory of
even
is that
more
remarkable
it does
contemplated long
in
not violate
its
the
formulation,
as one does
feature laws
such
as
not transgress
of of
the
other
general branches
thermodynamics
of or
quantum
so
theory.
(I shall return presently to the meaning I attach to the phrase
black hole.
It
described is known
by Dirac's equation.
that
one
can
of validity of
the "the
relativity".)
My first example derives from a consideration electron waves,
of
physics
theory,
domain of validity of the general theory of
the domain
theory
extract
of the
behaviour
In the s p a c e - t i m e of a the
rotational
energy
of
of
Kerr the
black
hole
by processes
More precisely,
which
result
if we have waves,
in the
slowing
with a time
(t)
down of its rotation.
and an
azimuthal-angle
(
(m = 0.
with a frequency cr(>0)
± 1 . ±2.
.. .).
less than the critical value
as = - a m / 2 M r +
(m = - 1 . - 2 .
...).
where a and M are the Kerr parameters and r+ horizon,
then
reflection radiance
one
has
coefficient is a
for
super-radiance the
necessary
incident
(by
which
waves
consequence
of
is the radius of the event one
exceeds
a
theorem
means
unity). due
that
The
to
the
super-
Hawking
that
every interaction of a black hole with an external source must always result in an increase of the surface area of the event horizon provided only that the energy-momentum
tensor
positive-definite character reflection
of
well-defined
Oirac
of the external
of the energy.
waves
mathematical
by
the
algorism
violated.
But one
Dirac waves, energy we
soon
should
black
of the
realizes that the
have
Had
had
the
standard
a contradiction
relativity and the
when one considers finds, do
the
by
not
a
exhibit
Hawking's theorem is
momentum
tensor
of the
does not satisfy the positive-
algorism
of the
one
that they
then.
energy
between
premises
is compatible with the
hole,
theory,
Apparently,
provided by the quantum theory,
requirement.
theory of
However,
Kerr
the phenomenon of super-radiance.
source
predicted
the
super-radiance.
premises
quantum
of the
theory.
general
But no
such
contradiction occurs!
Let that,
when
me
consider
one
a
considers
second the
example.
curvature
classical potential for electron
(or photon)
of
must
the
black
quantum hole,
Planck)
an
theory,
one
emission
distribution
at a
of
at which energy
(or
an entropy.
that
notion
of
from
the
photons)
determined
as
event with
by the
one
derives
1975
providing
a
of the particles,
horizon Fermi
a
is
entirely
consistent
known laws of thermodynamics and of statistical mechanics.
of
a
(or
a
surface-
and the
one can
When one pursues this line of reasoning,
entropy
In
constant
Associated with this temperature
is lost by the emission
formally, the
electrons
space-time
showed
scattering according to the rules
observe,
temperature
gravity of the event horizon.
of
Hawking
rate
define.
one finds
with
all
the
Thermodynamics
and
statistical
mechanics
were
formulation of the general theory of relativity; follow
from
statistical
the
theory
not
violate
the
not
the that
laws
of
contemplated
thermodynamics
and
of
mechanics.
The
foregoing
consistency
of
domain
physics
of
probably
do
in
and yet the consequence
the
illustrations
general
sufficient
theory
disclose
but
also
not
its
only
consistency
outside
the
realm
originally
grounds
for
one's
confidence
the with
contemplated. and
faith
internal the
entire
These in
are
Einstein's
general theory of relativity.
V.
There aesthetic
is
another
feature
of
the
theory
that
Every valid physical theory is circumscribed to
it.
that
Thus,
the
bodies
light.
Newtonian
should
be
The classical
limited
theory
moving
quantum
of
with
related
of
that the
action,
gravity
is limited small
by the
to
its
Likewise,
we
are
may
inherent
requirement
compared
and of electrodynamics
relevant actions
h.
by limitations
velocities
laws of mechanics
by the requirement
Planck's
is
base.
to are
that
of
similarly
large compared expect
the
to
general
theory of relativity to be limited by the requirement that the intervals of time and
of
distance
are
(-5.4X10-44 sec.)
Any overcoming the
two-body
In
whose the
problem
quantitative one-electron
theory
an
an
earlier
gravitation,
laws.
its
the
exact
departures
of the
from the Balmer series of hydrogen.
Pickering
of
are
peculiar
by to
basis
for
its
solution
of
the
solution
Similarly.
derivation
a
theory
Bohr's
Balmer's
provides
a
theory
of and
the
exact
that
provide
(fi=h/2ir).
and
from
an
of
will
example:
Kepler's
respectively
circumstances
(ftG/c5)1*
scales.
formula
nucleus
provides
Planck
replaces
description
Newtonian
for
the
an exact basis for determining the ratio of the masses of the electron the
systems
which
will envisage exact
provides
explanation
to
(~1.6xl0-33 cm),
theory,
limitations,
and
confirmation.
compared
and ( n G / c 3 ) " 4
physical
its
theory
large
series
of
ionized
helium
We
now
ask:
what
is the
essentially
new feature
theory of relativity? And what are the circumstances features
essential
precise
notions
notions
generalize, of
features
regarding in
relativity.
of
the
space
and
magisterial
We
ask:
general time
theory
which
fashion,
those
there
physical
are
these new notions of the theory are manifested space-time
around
black
holes
provides
the
of
general
it
relativity
those
underlie
The
is simple
it involves only two parameters:
the
and
angular
space-time physical
of
momentum
all
known
theories
characteristic general
of
theory
satisfying
space-time
that
have and
relativity
aspects;
and
of
test
itself
of
around
the
black
circumstances
a
arena.
appears
to
leads
the
me
This
of
examine,
its
The
general
specified: black
None a
most
it
hole
in
the
of
the
problem
feature
more
which
magnificent
behaviour
provides
as one
to
The
predicted.
complete.
These special
in
with
uniquely
and
hitherto
so
me
is
the
purity?
the mass of the
exactly
explored
solution
holes
hole:
particles is
been
this
black
the
in their pristine
requisite
completeness.
are
Incorporates.
that
theory of relativity solves the problem of these s p a c e - t i m e s
and
the
unambiguously?
The
theory
of
which will reveal
of
so the
aesthetically
generally,
the
aesthetic base of the theory.
To aesthetic beauty, and
examine
appeal
is
a
physical
beset
with
theory
and
difficulties.
to
Like
state all
the
source
discussions
relating
it is subject to the tastes and the temperaments of the
it is difficult,
if not impossible,
to achieve
objectivity.
Nevertheless,
as a practitioner of the
theory
more years,
of
relativity for
aspects
the
aesthetic
and
solution
of
the
the
Ingredients of
past twenty
theory
appeal
of the
problems
that
and
to
my aesthetic
theory lead
influence to
a
I can
sensibility
and
direct
deeper
its to
individuals;
seems to me that the question is relevant:
what
of
ask and
the
it
general myself: how
do
formulation
understanding
of
the
physical and the mathematical content of the theory?
I have already remarkable theory
of
fact, relativity
to
referred
which provides,
I
to the theory of the
have for
solution with just two parameters.
also
Isolated
made
black
reference,
stationary
black
holes.
that
the
holes,
a
It is a general unique
As I have said on another o c c a s i o n .
Black
holes
masses
are
to millions
as stationary them,
macroscopic of solar
exactly
to that extent,
have of an exact description
a
variety
of
the only elements
concepts
of
space
most perfect general
and
theory of relativity
that
is
there
provides
not
of
all,
variety
all!
of
electromagnetic,
gravitational,
derived
theories. our
In basic
by definition,
the
And since
two-parameter objects
every
we
objects,
holes are
universe.
unique
to
physics,
physical
almost
in the
a single
Contrary
one of
Macroscopic
of black
thus,
are
solar
considered
every single
object.
they are the simplest
standard equations of mathematical scattering
a
They are,
solutions for their description,
But
to
objects
a few
by a variety of forces,
In the construction
time.
macroscopic
from
This Is the only Instance
are governed
approximations
contrast,
they are
solution.
of a macroscopic us,
varying
To the extent they may be
by the Kerr
as we see them all around from
with masses
masses.
and isolated,
described
objects
as
family of
well.
expectation,
the
relating to the propagation
and
and
prior
the
the
Dlrac-electron
waves.
as well as the geodesic equations of particles and of polarized photons, of
them,
can
be separated
of these equations
of the circumstances be
separated
an example. electron Dlrac's
In
and
and
solved
exactly.
The
has led to a re-examination when partial differential
solved:
and
a
rich
manner
of
separation
of the c e n t u r y - o l d
equations
mathematical
problem
in two variables
theory
has arisen.
I may refer to the separation of Oirac's s p i n o r - e q u a t i o n Kerr
geometry.
spinor-equation
special relativity -
in
As
a
corollary,
spheroidal
it
led
coordinates
to
in
the
can As
of the
separation
Minkowski
all
of
geometry of
a separation that had been considered impossible
before.
VI.
I
now
address
myself,
aspects
of
turn
namely,
the theory
physical significance. am
not
to
to
descend
the
most
how
sensitiveness
enables
the
difficult
question to
the
formulation
and
In answering this question. to
dilettantism.
That.
I
am
to
which
I
mathematically solution I should afraid,
of
wish
aesthetic
problems
be precise will
to
require
of if I a
somewhat more technical language than I have used so far.
There are two major areas
in general
relativity in which
progress
has
been
made
in
recent
years:
the
mathematical
black
theory of colliding
from the gravitational
collapse of massive stars in the late stages of
are well
collision
or
known.
scattering
But the
of
waves
relevance by
waves
Black
of
and the mathematical
evolution,
waves.
theory
holes,
of the theory in
general
as
holes
resulting stellar
bearing
on
relativity
the
requires
explanation.
In the general theory of relativity one can construct gravitational waves confined per
unit
area;
gravitational note
that
For.
a
and.
waves
one
6-functlon
a
cannot
we
construct
for
can.
6-function
profile
profile
6-function
therefore,
with
6-functlon
between two parallel
of
the
the
limit,
impulsive
energy
field
planes with a finite
energy-profile.
such
the
in
construct
energy
impulsive
Parenthetically.
waves
will
imply
variables:
plane-fronted
and
in a
I may
electrodynamics. square-root
the
square-root
of
a
of
a
is simply not permissible for physical description.
In 1 9 7 1 , Khan & Penrose considered the collision of two impulsive gravitational result
of
unlike
the
waves
the
collision
singularity
acquainted. the
this
parallel Is the
in
the
context,
polarizations.
development
the
interior
This phenomenon
theory:
colliding waves, In
with
of
no
of
the
they
showed
nothing
short
of
space-time
singularity
not
which
are
with
exact
relativity
progress
mathematical should
physical
theory not
of
by
a
focussing
solution
of
the
In any event,
constructing
the
in
this
a
colliding
will
for
area
waves.
that
exploration. before
foreseen.
This
the
Clearly,
problem
will
the o c c u r r e n c e
However,
one
mathematical
theory
of
structure
of the theory emerge -
fact
two theories
be as closely
mathematical
mathematical theory of black holes, implications
of
of
suggested to Penrose that a new realm in the remained
have thought
circumstances
developing
we
in any linearized version of
singularity,
an
the
holes
is.
there
was
that
the
realized
mathematical theory of black holes is structurally very closely
one
that
in no way depends on the amplitude of the waves.
general
substantial
a
is not manifested
occurrence
a singularity in this example, of
of
black
suffice to disclose the new phenomenon.
physics
And
in
dealing
related to the
itself,
with
surprising:
such
disparate
related
as they
are.
Indeed,
by
colliding
waves
with
a
to
architecturally
similar
one finds that a variety of new implications
one simply could
view to
the
physical not
have
VII.
A description was accomplished
of
how
the
development
to
which
I have
referred
is not possible without some familiarity with the
of general relativity,
language
any more than an analysis of a musical composition is
possible without some familiarity with musical notation.
We axisymmetric scattering
are
concerned
black of
holes
with
and
plane-fronted
space-times
space-times waves.
In
that the
that
describe
describe former
the
stationary
collision
case,
the
coefficients are independent of the time,
t,
about the axis of rotation;
only on the two remaining
coordinates, case, x
a
radial
they depend
coordinate
r
and
and
metric
and of the azimuthal angle,
the
polar
angle
e.
In
spatial
the
latter
the metric coefficients are independent of two space-like coordinates,
and rr.
both ranging from -oo to +00; they depend only on the time,
and the remaining spatial coordinate,
It can
be
shown that the
1
x . normal to the (x .
metric
appropriate
t.
2
x )-planes.
to a
description
of
stationary axisymmetric black-holes can be written in the form: ds 2
[x(dt)2 -
= y(A6)
~
(do -
wdt) 2 ]
- e^2"*"**3 / A U4n>_ 2
+
Idiii- 2 ].
(1)
where A = T )
2 -
1 .
n2 = sin 2 e
6 = 1 -
(ji = cose)
.
(2)
T) is a radial coordinate (measured in a suitable unit)
and x-
H3
be
are
metric
functions
to
be
directly related to the angular
determined.
It
may
u.
noted,
momentum of the black hole:
and f-z • that
u
is
It is zero for
the Schwarzschild black hole which is static. In writing the metric for
the
occurrence,
at
T\ =
In the form 1,
of
a
(1),
we have already
null-surface
that
will
arranged
eventually
be
Identified with the event horizon of the black hole. The central problem of the theory is to solve for x and <»•
once
one
has solved for
them,
the
remaining
metric
function,
n2
+
f-3'
follows
"conjugate
metric"
by a simple quadrature. Associated obtained by the
with
the
metric
(1),
we
have
a
transformation
t -> +l
and
By this "conjugation", X = ^
<)> •> - I t .
(3)
x and u are replaced by
2
and
w = ^ z -
.
(4)
For the reduction of the physical problems,
it is essential that we
in place of x and u ,
¥ and 4>, where
consider,
the pair of functions,
V(A6)
* = ^~Z and *
(5)
,
Is a potential for o> defined by _6 =
*'T)
A
X7
u
a n d
'/i
One can similarly define *
*'M
and *
=
""x
2
""^
In terms x and S.
In the mathematical theory of black holes, functions *
and *
ZT
= *
and *
+ I*
and *
one combines
the
into the pairs of complex functions,
ZT
and
<6)
•
= *
+ I* ,
(7)
and defines E
T
t_i = ~Z— ZT+1 z
and
E
T
Z1-! = ~^Z ZT-1
.
(8)
Both these functions satisfy the Ernst equation.
<1 -
|EI2{[
= -
Turning colliding
waves,
( 1 - n2)
2E* [ ( 1
next we
to
- n2)
E.^].^ (E.^)2 -
space-times
envisage
-
the
[(1 (1 -
p.2)
appropriate
collision
of
two
n2)E.M].n} CE.^)2]
.
to
description
the
plane-fronted
(9)
of
impulsive
gravitational
waves
shock-waves
with
-00.
Prior
to
accompanied, the
same
the
instant
approaching
wave-fronts
space-time
that develops
boundary
conditions
is
in
general,
fronts, of
collision,
flat.
with
approaching
We
the
are
gravitational
each
collision
fronts
Is
not
a
and
from
space-time
principally
after the instant of collision
at the
other
other
+00 and
between
concerned
the
with
the
(though satisfying
the
negligible
the
part
of
problem) .
The metric of the s p a c e - t i m e after the instant of collision can
be
written in the f o r m , ds2 = -
y(A6)
[x
(dx2)2 + i X
(dx 1 -
+ e " * " 3 VA [
where,
q^dx2)2]
,
(10
now. TI2
A = 1 -
T) measures
.
the time
measures
the
distance
and
x.
ix2 .
6 = 1 -
(In
collision,
a suitable
normal
<\z- and
unit)
to the
v+n3
are
(11)
from
colliding metric
the fronts
functions
instant
of
at the
collision,
instant
the
gravitational
waves:
it
is
zero
when
the
jx
of
the
to be determined.
may be noted that q 2 is directly related to the varying plane of of
)
plane
of
It
polarization
polarization
is
unchanging.
In account,
writing
the
aposteriori,
metric
in
the fact that,
the
form
(10).
solution solved
to for
in the
the
the
case
Einstein metric
of
stationary
field-equations
functions
x
have
as a result of the c o l l i s i o n ,
or a coordinate singularity develops when T)=1 and
As
we
and
be r
<\z ° -
into
a curvature
ji=±l.
axisymmetrlc
can
taken
space-times,
completed
once
equivaientiy.
for
we *
the have
and
related to x and q 2 by
* = and
J4A8L-
(12)
*
36 and
* - T ) = j ? la,/* In
the
present
*•/* = x 5 - *' 2 ' 1 ' 1
case,
we
need
(13)
•
not
consider
the
process
of
"conjugation" since it corresponds to a simple interchange of the roles of x and x 2 .
We now combine
the functions
x
and q 2 and *
and *
into the
pair of complex functions, Z = x + iq 2
and
Z1" = *
and
Ef =
+ i« .
(14)
and define
E = frj
Z
Z+l
z
V"1 • t
(15)
+ 1
We find that both E and E* satisfy the same Ernst-equation
When we turn to the consideration
(9).
of charged black-holes or the
collision of gravitational waves coupled with electromagnetic waves, supplement with
the
Einstein's
two
equations
symmetries
we
with are
Maxwell's
equations.
considering,
the
For
Maxwell
we must
space-times field
can
be
expressed
in terms of a single complex-potential
H;
and the entire set of
equations
governing
be
reduced
the
problem
can
eventually
to
a
pair
of
coupled equations for H where * u
or
q2.
and
Z1" = *
I HI2.
+ I* +
Is defined as in equations defined
similarly
as
in
(5)
(16) and (12)
equations
(6)
and * and
is a potential for (13)
but
including
additional terms in H on the right-hand sides.
There are two cases when the pair of equations governing Z* H can be reduced to a single Ernst equations. Case ( i ) :
H = Q(Zf
where Q is some real constant; Case ( I I ) :
Zt
= 1.
+ 1)
These are:
,
(17)
and *
= 0,
and *
= 1-
|H|2 .
(18)
and
37 In case ( i ) .
t
with the definition,
zM
ET
= ^ ZT + 1
.
(19)
we find that E* satisfies the equation. (1 -
4Q 2 -
|E|2) {[(1-r,2) = -2E*[(1
for
b_oth types
can
of
vacuum,
a
solution
equations. should
we
are
-
[(1
jt2)(E.M)2],
(1 -
presently
n2)E.M].^}
-
considering.
(20)
Moreover.
is a solution of the Ernst equation
(9)
for
It the
then EEi.Ma.
is
n2)(E.T1)2-
-
space-times
be shown that If E v a c
E.,,].,,
V(1-4Q2)
= Evac.
of
equation
(21)
(20)
appropriate
for
the
Einstein-Maxwell
(It should be noted that in the stationary axisymmetric
also
consider
the
process
of
conjugation
"tilded" variables will satisfy the same Ernst
In case
(ii).
we find that
when
the
case,
we
corresponding
equation.)
H satisfies the
Ernst equation
(9)
for
the vacuum so that we c a n write H = Evac
and
The completion consider,
particularly
elaborate
analysis.
not
needed
for
*
of the solution
In the theory of
We
= 1 -
shall
exhibiting
not
the
for the various
colliding
describe
structure
|Evac. I2.
any
and
(22)
problems we
waves,
often
of
analysis
the
that
coherence
shall
requires
of
fairly
since the
it
is
entire
theory.
VIII.
The origin of the structural similarity of the mathematical theory of black
holes
and
of
colliding
waves
stems
from
the
circumstance
that
both cases the Einstein and the Einstein-Maxwell equations are reducible the
same
Ernst
equation;
and,
indeed,
as we
shall
see.
even
the
in to
same
38 solution.
This identity is obtained only by the special choice of coordinates
that assures the occurrence for
black
colliding that
of an event
holes
and
the
waves.
The
richness
and
in
of
are
described,
development
spite
horizon at a radial distance
of
the this
a
singularity
diversity identify,
of
at
the
time
T)=1 for
physical
results
from
t)=l
situations
the
different
combinations of the metric functions which can be associated with the same solution of the Ernst equation.
We
shall
consider
first
the
solutions
equations.
The solution of the Ernst equation
describing
the
diverse
physical
situations
derived
(9).
from
the
vacuum
from which the solutions
follow.
Is
the
simplest
one.
namely. E = pT) + iqji .
(23)
where p and q are the two real constants restricted by the requirement. p2 + q 2 = 1 .
(24)
In the theory of black i.e.. The
to the Ernst equation for solution that follows
solution
when
p
Schwarzschild
=
and
1
a
special
q
Kerr
books and generally known. to
the solution pri + iqji. applies
+ i*
=
belonging to the conjugate
Kerr. 0.
It reduces
The
black holes
to the
resulting
are
to E^ metric.
Schwarzschild
space-times
adequately
described
of
the
in
text
I shall mention only that these solutions belong Petrov
classification.
Solutions belonging to this type have many special properties.
It is to these
properties
algebraic
*
is that of and
the
holes,
that
we
type,
owe
namely,
the
type
separability
of
O in the
all
the
standard
equations
of
mathematical physics in Kerr geometry.
Turning solution
next
to
the
Is that of Khan and
theory
of
colliding
waves,
Penrose which describes
the
purely impulsive gravitational waves with parallel polarizations. the
solution.
E=pr) + iqji. general
case
polarizations.
E =
TJ. of
leads
to
when Thus,
the
Ernst
equation
the
Nutku-Halil
the
colliding
solution impulsive
for
x
+
which waves
fundamental
the collision of two
'q2-
It follows from Tne
describes have
solution. the
more
non-parallel
the Khan-Penrose and the Nutku-Halil solutions play the
same
role
In the theory
of
colliding
waves,
as the
Schwarzschild
and
the
Kerr solutions play in the theory of black holes. The c o m b i n a t i o n . the
same
Ernst
pr) + iq/x for
singularity, are
the
equation;
E'.
unexpected:
a
+ l
we
are
metric
functions,
Invited
to
also
consider
leads
the
horizon
develops
a common
when
waves
extend the s p a c e - t i m e
when belief
collide.
t)
=
that
In
beyond n = 1 and
I n I = 1.
mirror
image
left
place
of
a
curvature
instance,
we find that the extended s p a c e - t i m e one that was
in
space-like
this
made,
of the
1.
we
entirely
curvature
singularities
must,
therefore.
When this extension Is
includes a domain which
behind
to
solution
The solution that follows has properties that were
violating
rule
*
and
a further
domain
is a which
includes hyperbolic a r c - l i k e singularities reminiscent of the ring singularity in the
interior
of
the
Kerr
black-hole.
It
is
remarkable
that
resulting from the collision of gravitational waves should resemblance Looking our
to Alice's
Glass:
passage
"It as
anticipations
[the
far
passage
as
you
In
can
with
respect
the
see,
to the
Looklng-Glass only
It
a
space-time
bear such a world
House]
may
be
quite
the
solution
close
Through
the
Is
very
like
different
on
beyond."
The E'
= pt)
+
space-like
foregoing
Iqn,
apply
curvature
remarks, only
concerning
when
singularity
q
*
0.
q
=
=
from
0
and
p
at
T) = 1 ;
and
the
space-time
that
the
solution
develops
When
derived
1,
a
cannot be extended Into the future.
Finally. E* = pr] +
iqfi
it is
of
should type
be
noted
D and
shares
all
the
derived
mathematical
from
features
of
s p a c e - t i m e s belonging to this type.
Turning solutions
next to the Einstein-Maxwell equations,
appropriate
E = pr) + iq/i accordance
with
to charged
black-holes
( p 2 + q 2 = 1-4Q 2 ) equation
(21).
for the Ernst equation
We obtain
the
were
conceptual
"elementary" solution of the
difficulties
Einstein-Maxwell
in
the
(20)
for
Relssner-Nordstrom
when q = 0 and the Kerr-Newman solution when q *
There
we are led to the
when we consider
Et.
in
solution
0.
obtaining
equations
solution
the
corresponding
for colliding
waves.
Penrose had raised the question: its
associated
6-function
6-functlon
singularity
would an impulsive gravitational wave with
singularity
In
the
In
the
Weyl
energy-momentum
tensor
imply
tensor?
If
a
similar
that
should
happen,
then the expression for the Maxwell tensor would involve the square
root
the
of
6-functlon;
interpret
such
satisfying
the
these
a
many
accounts,
formulated
"one
efforts
conditions
and
the
would
Besides,
boundary
all
initial
Khan-Penrose
and
function".
be
there
at
was
a
loss
the
conditions
at the various
to
solutions
obtain
failed.
However,
Nutku-Halll
know
null
it was
followed
how
to
problem
of
boundaries.
compatible
when
solutions
to
formidable
with
realized
from
On
carefully that
the
the
simplest
solution of the Ernst equation for x + l q 2 -
it was natural to seek a solution
of
will
the
Einstein-Maxwell
solution
when
the
Maxwell
straightforward
one:
equations,
do
we
equations field
since,
not
which is
in
have
switched
the
an
reduce
Ernst
off.
The
framework equation
to
of at
the
problem
the
the
technical
and
a solution
conditions
problems can
and
that
are
presented
be obtained
physical
which
can
be
satisfies
requirements.
That
of
can
not
the
a
metric IHI2.
+ i* +
successfully
all the we
Is
Einstein-Maxwell
level
functions x and Qz'- we have one only for E^ derived from * The
Nutku-Halll
overcome
necessary
boundary
obtain
physically
a
consistent solution by this "inverted procedure" is a manifestation of the firm aesthetic base of the general theory of relativity.
Since we do have an Ernst equation for * consider
the
equation
(20)
solution
for the
the
vacuum
solution for
E =
E*.
vacuum
solution,
pr|
+
When we it
iq/x
Q
=
(p 0.
2
+
have described
develops
a
q
this
2
+ I* 2
=
1-4Q ).
solution
earlier:
horizon
+
and
will
IHI2, for
we can
the
reduce
Ernst to
the
and we find that,
like
subsequently,
timelike
singularities.
In our consideration have
distinguished
essential time.
ways:
two
when
in case ( I ) ,
of the Einstein-Maxwell
cases:
case
while
considered
hitherto belong to case
in
case
electromagnetic
the vacuum.
and field
case
(ii).
Is switched
in SVII.
They off,
we
differ
the
in
space-
reduces to a non-trivial solution of the Einstein v a c u u m -
equations.
complex
(I)
the electromagnetic
equations
(II).
potential.
it
becomes (I). H.
flat.
The
solutions
As we have s e e n , satisfies
the
Ernst
we
have
in case
(ii)
the
equation
(9)
for
We naturally ask the nature of the s p a c e - t i m e that will follow
from the
simplest
solution,
one then obtains
(discovered
a very remarkable one: scalar,
is
impulsive
confined
the impulsive waves, of
the
of the
Ernst equation.
waves.
the
other
6-functlon words,
profile
except
for
the s p a c e - t i m e is conformally flat.
in
accompanying
Impulsive
solution
as manifested by a non-vanishing
to
In
Einstein-Maxwell
space-time
The
by Bell and Szekeres by different methods)
gravitation,
exclusively
gravitational
solution
pri + iq/t.
which
equation,
plane-fronted
gravitational
we
describing the
a
collide
and
of
as an exact
conformally
electromagnetic
waves,
the
presence
Thus,
have
Is
Weyl
flat
shock-waves.
develop
a horizon.
A further feature of the Bell-Szekeres solution is that the solution for q = 0 is
entirely
solution,
equivalent in
this
to
the
solution
framework,
which
situation than the Bell-Szekeres the
simplest
advantage
solution
of
of a transformation
one-parameter
family
equation.
therefore
We
E = pt) + iq/ithe
E'
= pr) + i q n .
features
of
of
will
*
0.
Therefore,
describe
solution,
Ernst
q
a
more
general
For
this
consider
the
any
given
Ehlers
we
of
the
solution
for
It Is remarkable
of
of
space-times
with
this
the
vacuum
derived
the
the
from
that we should obtain
abundant
structure
by
of
take
us to obtain
solution
transform
a
physical
purpose,
due to Ehlers which enables from
obtain
we must go outside the range
equation.
solutions
to
a
a
Ernst
solution.
We find that the resulting solutions are of type D and
all
family
the
for
the
have
solution
one-parameter
applying
the
Ehlers
transformation to the Bell-Szekeres solution.
In Table 1 . we describe more fully the various solutions that have been derived for
black
holes and for colliding waves.
The
pictorial
pattern
of this table is a visible manifestation of the structural unity of the subject.
The inner relationships theory context
of
colliding
when
u
waves
= 0 and
between the theory of black holes and the
is equally
visible
q 2 = 0.
In this
which the solutions for both theories d e p e n d , [(1
This equation
- n2)
can
(Ig*).^]^
be solved
-
(see
Table
case,
the
2).
in the
basic
the two theories are listed in Table 2.
and
on
is
[<1 - j i 2 ) ( l g * ) . ^ ] . M = 0 .
exactly
simpler
equation,
the solutions
that
are
(25)
relevant
in
1245 Table 1 Solution for Ernst Equation for i t
Killing Vectors
Field Equations
E
a,.
a„
Einstein-vacuum
does not exist
8
a
*
Einstein-vacuum
does not exist
a,,
a,
Einstein-Maxwell
does not exist
TI^(1-4Q2)
8
o„
Einstein-Maxwell
does not exist
pn+iqM." p 2 +q 2 =l-4Q 2
T.
interchanges x1 and x2
f
f
V
V
Einstein-vacuum
O x i-
o^
Einstein-vacuum
pn+iq**; p 2 +q 2 =l
Einstein-Maxwell
does not exist
V
3
x2
Elnsteln-vacuum
V
3
x2
Einstein-vacuum
V
3
x2
Einstein-Maxwell
does not exist
a 1- 3 2 x' tr
Einstein-Maxwell
does not exist
E*
n
pTi + iq*i: p 2 +q 2 =1
interchanges x1 and tr
E r ( E v a c = p n + lqM> xy( 1-4CT)
interchanges x1 and x2
T)
pn+iq*i: p 2 +q 2 =i
TH/(1-4Q2)
pTi+lq^t; p 2 +q 2 =l-4Q 2
V
a
x2
Einstein-Maxwetl
does not exist
V
3
x2
Einstein-Maxwell (H=E vac )
does not exist
pn+iqM;
V
ax2
Elnsteln-Maxweli (H=E vac )
does not exist
Ehlers transform of Pn+iqM
Einstein-hydrodynamics («=p)
pn+iqn
a •), a 2
interchanges x1 and x2
interchanges x1 and x
interchanges x1 and >r
interchanges x1 and x2
r\
interchanges x1 and x2
p 2 +q 2 =i
interchanges x1 and x interchanges x1 and x2
interchanges x1 and x^
1246 Table 1
Solution
Description
Schwarzschild
Black hole; static; spherically symmetric event horizon space-like singularity at centre type 0 ; parameter: mass
Kerr
Black hole; stationary, axlsymmetric ovent & Cauchy horizons; ergosphere time-like ring-singularity in equatorial plane type 0 ; parameters: mass and angular momentum
Relssner-Nordstrom
Charged black-hole; static: spherically symmetric event and Cauchy horizons time-like singularity at centre type 0 ; parameters; mass and charge
Kerr-Newman
Charged event & time-like type D;
Khan and Penrose
Collision of Impulsive gravitational waves parallel polarizations develops space-like curvature singularity
Nutku-Halll
Collision of impulsive gravitational waves non-parallel polarizations develops space-like curvature singularity (weaker than Khan-Penrose)
Chandrasekhar a n d Xanthopoulos
Collision of impulsive gravitational waves and accompanying gravitational and electromagnetic shock waves non-parallel polarizations develops space-like curvature singularity
Chandrasekhar a n d Xanthopoulos
Collision of impulsive gravitational waves and accompanying gravitational shock-waves parallel polarizations develops very strong space-like curvature singularity type D
Chandrasekhar a n d Xanthopoulos
Collision of impulsive gravitational waves and accompanying gravitational shock-waves non-parallel polarizations develops a horizon and subsequent time-like arc-slngularltles type D
Chandrasekhar a n d Xanthopoulos
Collision of Impulsive gravitational waves and accompanying gravitational and electromagnetic shock-waves parallel polarizations develops a horizon and subsequent three-dimensional time-like singularities type D
Chandrasekhar a n d Xanthopoulos
Collision of Impulsive gravitational waves and accompanying gravitational and electromagnetic shock-waves non-parallel polarizations develops a horizon and subsequent time-like arc-slngularltles type D
Bell-Szekeres
Collision of impulsive gravitational waves and accompanying electromagnetic shock-waves parallel polarizations space-time conformally flat develops a horizon; permits extension with no subsequent singularities
Bell-Szekeres
Same as above
Chandrasekhar Xanthopoulos
and
Chandrasekhar a n d Xanthopoulos
black-hole; stationary, axlsymmetrlc Cauchy horizons; ergosphere ring-singularity in equatorial plane parameters: mass, charge and angular momentum
Collision of Impulsive gravitational waves and accompanying gravitational and electromagnetic shock-waves develops a horizon and subsequent time-like arc singularities type D Collision of impulsive gravitational and accompanying gravitational shock-waves and null-dust (R|j=Ckjkj) non-parallel polarizations develops weakened space-like singularity transforms null-dust Into a perfect fluid with e=p
Table
Basic Equations
2
[ ( 1-r, 2 ) ClglO . ^ 3 . ^
-
[ (l-/i2) (lg»). ^
.^ = 0
Killing Vectors
Field equations
Solution
Remarks
at.
a*
Einstein
|g*= TCI
Schwarzschil symmetric
»t-
3
Einstein vacuum
|g*= D = l + E AnPn(M)P„
Distorted bla (when EA 2 n (Weyl's sol
Einstein
ig*=
a •,. x
*
a
2
X
X
8
X
vacuum
x
a -,. a
8
vacuum
2
Einstein vacuum
i9*= r^
a
1'
a
+
?AnPn(M)Pn(n)
1±D_ l-n
2 x
Einstein-Maxwell
ig*=
2
Einstein-Maxwell
ig*= ^ l-ri
X
Khan-Penros impulsive w polarization
1-T)
X
V
i±n_ 1-TJ
Collision of waves with tational sho parallel pol Collision of waves with magnetic s formally fla polarization
+ Jf A „ P n ( ^ ) P n ( T , )
Collision of waves with tational and waves; par
As equations
the
foregoing
share
equations.
many
The
of
only
discussion
the
distinctive
source,
coupled with gravitation,
demonstrates,
other
features
than
a
of
the Rlcci tensor,
Einstein
leads to equations which retain,
energy density
(e)
U'UJ
Einstein-Maxwell
field,
vacuum-
which
at least,
when
some of
is a perfect fluid with the
= pressure
(p).
For such a fluid,
in accordance with Einstein's equation,
= -46
R')
the
Maxwell
the distinctive features of the vacuum equations, equation of state,
the
is given by
,
(26)
where u1 denotes the four-velocity of the fluid.
On the colliding waves, fluid with e = p. gravitational
assumption,
that
in the
region
after the instant of collision,
of the
interaction
of
the
we have as source a perfect
we find that prior to the Instant of collision, the impulsive
waves
must
have
been
accompanied
by
null-dust
with
an
energy-momentum tensor of the form. TU
where
_
Ekikj
_^RU
=
(27)
E is some positive scalar function and k' denotes a null vector.
other
words,
dust
(i.e..
under
the
massless
circumstances
particles
envisaged,
dlscrlblng
null
a transformation
trajectories)
fluid (whose stream lines follow time-like trajectories) of collision.
perfect
In the first instance.
But as Roger Penrose and Lee Lindblom have pointed out.
transformation special
a
in
relativity
question though
can this
take fact
place, does
equally.
not
seem
In null
occurs at the Instant
That such a transformation is required is.
surprising.
into
of
In
the
to
frame-work
have
been
the of
notified
before.
In theory
of
developing
black
holes,
the
theory
of
colliding
we
have,
in
effect,
waves
examined
consequences of adopting for the Ernst equation, simplest
solution
(or.
approach
may
disclosed
possibilities
for
example.
appear
the
in as
one an
case,
its
exceedingly
that one could development
of
formal
horizons
parallel
with
the
systematically
the
in Its various contexts,
Ehlers
not have,
in
transform).
one.
it
has
in any way. and
While
its this
nevertheless foreseen,
subsequent
singularities or the transformation of null dust into a perfect fluid.
as
time-like In this
Instance,
then,
exploring
general
relativity,
sensitive to its aesthetic
base,
has led to a deepening of our understanding of the physical content of the theory. In 1915,
his
first
announcement
of
his
field
equations
In
November
Einstein concluded with the statement:
Anyone who fully comprehends
At least,
to one practitioner,
this theory cannot escape
Its
magic.
the magic of the theory is in the harmonious
coherence of its mathematical structure.
REFERENCES The
nature
conventional
of
the
style.
lecture
precludes
However,
the
giving
list
of
papers
of
Karl
Zeitsch.
4.
particular
a
references
in
Schwarzschlld
quoted explicitly in the text are:
1.
"Bemerkungen zur Elektrodynamik",
2.
"Ober die Elgenbewegungen der Fixsterne",
Phys.
3.
"Ober das zulassige
4.
"Ober
431
(1903).
GSttinger Nachrlchten.
p. 614
(1907). KrummungsmaQ
Astronomischen Qeselischaft, das
Gravltatlonsfeld
Theorie".
eines
des Raumes", 35,
337
der
(1900).
Massenpunktes
Sltzungsberichte
Vlerteijahrsschrift
(Berliner
nach
der
Akademle),
Einsteinschen p. 189
(1916
February 3 ) . 5.
"Ober
das
Gravltatlonsfeld
nach
der
Akademle), 6.
"Zur
einer
Kugel
Einsteinschen
aus
Theorie",
Inkompresslbler
FlUsslgkeit
Sltzungsberichte
(Berliner
p.424 (1916 March 2 3 ) .
Quantenhypothese".
Sltzungsberichte
(Berliner
Akademle).
p. 548
(1916 May 4 ) .
Referring
to Schwarzschild's
Lanczos
(in
London, of
the
"The
p. 152
Einstein
(1974))
relatlvlstlcally
paper
on
Decade".
comments.
Invariant
proper
"Bemerkung
zur
Academic
Press:
"Schwarzschild's formulation
Elektrodynamik".
of
New
remarkable the
principle
York
C. and
anticipation of
least
47 action
Is not recorded
in any history
of
physics".
For an account of Schwarzschlld's contributions to the notions of action angle variables,
Sommerfeld.
and
see
A..
Atombau
(Braunschweig:
und
Spektrallinlen.
Friedr.
Vol.1.
Vieweg & Sohn.
5th e d . .
pp. 6 5 9 - 6 6 3
1931).
The discussion of the theory of black holes and of colliding waves in SS VI to VIII is based o n :
Chandrasekhar.
S. .
"The
Mathematical
Clarendon Press. Khan.
K.
Nutku. Bell.
& Penrose.
Y.
& Halll.
P.
M. . Phys.
S.
P.. &
of
Black
Holes"
(Oxford:
1983).
R. . Nature.
& Szekeres.
Chandrasekhar.
Theory
Lond.
Rev.
Lett.
229.
185
(1971).
39.
1379
(1977).
Gen.
Rel.
Grav.
Ferrari,
V. .
Proc.
5.
275
Roy.
(1974). Soc.
Lond.
A396.
55
(1984), Chandrasekhar. 223
S.
(1985);
(1986); new
& Xanthopoulos. 37
(1985):
Proc.
Roy.
Soc.
A403,
189
(1986):
"On colliding waves that develop t i m e - l i k e
class
press);
A402.
B.C.,
of
solutions
of
the
Einstein-Maxwell
Lond.
A398,
A408.
175
singularities: equations"
a (in
and other papers in preparation.
APPENDIX
On
29
June
Schwarzschild measure
1916.
Einstein
at a meeting
of Schwarzschild;
gave
a
brief
of the Berliner and
memorial
Akademle.
I have thought
address
on
It presents a
it worthwhile
Karl proper
to append
the
following translation of Einstein's address.
On old,
was
many-sided all
by
May
11
death
a scientist
his astronomer
of
this
snatched
year away.
Is a grievous
and physicist
[/9f6], This loss
friends.
Karl
early not only
Schwarzschild,
demise to this
of
so
body,
42
years
gifted
and
but also
to
What
Is specially
Is his easy command which
he
could
questions.
penetrate
Rarely
reasoning
has
from The
theoretical
quests
seem
among
delight
discerning
In
understandable mechanics,
that a
established
equilibrium
science
fluid
the
are
of
investigations
objects
are
understanding
of Kapteyn's
indebted
Schwarzschlld
equilibrium radiative the
pressure
basis
the
Here,
solar
light
while
to
to
him
electrodynamics.
for
theory of gravitation: calculation life,
the
for
Sun
is grateful
also
and
for for
belongs
spherical
i.e.,
a
part
belongs.
In
and
of
on the of a
this
area,
widening
their
in
interesting
of theoretical
considerations
his beautiful
particles
fell
to on
questions,
contributions
to
In his last year he became
by a
relating
which provided
purely
in obtaining,
disease,
on the
Investigations
These Investigations
questions
physics to
his investigations
by astronomical
he succeeded
weakened
astronomical
the observations
deepening
in the new theory of gravitation. much
statistics,
to relate
his deep knowledge one
Interested
his
Besides,
on
theory of the
type of stars to the structure
atmosphere
on small
to be
his investigations
Schwarzschlld's
stellar
which
him
they were motivated
Schwarzschlld
Indebted
his
to
To this area
of
on
methods
for Arrhenlus' theory of comet tails.
physics, led
of transfer.
firmly
discovery.
directed
of the Sun.
celestial
more
and Poincare's
of
and the spectral
many
astronomers
the theory
important
by statistical
the velocity,
system
therefore
in
are
I may recall
problem
artists
masses.
most
his
which seeks
In this area,
inner
an
is
were
foundations
restless
deeper
from It
contributions
with many
his
the
than
to
mathematical
in
patterns.
whose
of the three-body
of rotating
luminosity, large
earliest
science
physical
adapted
of
learn
Nature
Schwarzschlld's of
account
to
of
or
been
motivations
curiosity
mathematical
solutions
contributions
a
aspects
way In
that he grappled
on
work
casual
astronomical erudition
away
delicate
branch
Among
of
Schwarzschlld's
from
theoretical
essence
mathematical
different
than any other.
the periodic
Schwarzchild's
and the almost
shrank of
less
the
about methods
And so it was,
others
mainsprings
relationships
the
much
reality.
which
difficulties.
to
so
about physical
problems
astonishing
of mathematical
an
In seem
In physics. the
exact
theoretical to have We
are
foundations
of
Interested
In the new
for the first time,
an exact
And In the very last months of he
yet succeeded
in
making
49 some profound
To
contributions
SchwarzschllcTs
investigations
on
of optical
will remain
an enduring
edifice
Wlen;
at
Director
of
the
astronomical sclentlfc
a
physics,
Images
photographic
new
be
He
also
for measuring
at an
of
this
had
contributions.
for the
visual methods,
modest
man
he has,
Now bitter
has
strength.
out 24,
Assistant
long
1909
In as of
leader
of
led him to advance
his
as
He
a
discovered.
how the blackening
Idea
of of
photometry
using
become
been
In of by
extra-focal
only through this Idea
did
feasible.
a
member
in this short time
circumstances
he
series
him,
fruits and have an enduring
which he devoted all his
an
and
purposes
brilliant
age
and after
observation.
the
of
studies
carried
A
his
astronomy.
From as
Potsdam. observer
their brightness:
besides
1912,
but his work will bear
1896-99
theory
These
were
his lively spirit
used
to whose Sitzungsberichte
many important
as
methods
can
photometry,
Since Akademie
importance.
what has been named after
plate
methods.
of stars
Institute
efforts
belong
the
of the tools of
astronomer.
Interruption:
Moreover,
by charting
photographic
photographic
his
observations.
field
experimental
to
also
refined
of the GOttingen Observatory;
Astrophysical
testify
he
Investigations
as a practical
without
as Director
contributions
which
of astronomical
theoretical
observatories
investigations
In
for the perfection
with his efforts
1901-09
theory.
theoretical
optics
instruments
SchwarzschllcTs simultaneously
great
geometrical
aberrations
worked
to quantum
of
this
communicated
have taken him away:
Influence
on Science
for
J. Astrophys. Astron. 5 (1984): 3-11
The General Theory of Relativity: Why "It is Probably the most Beautiful of all Existing Theories" S. C h a n d r a s e k h a r Laboratory for Astrophysics and Space Research, Enrico Fermi Institute, University of Chicago, 933 East 56th Street, Chicago,, Illinois 60637 L.S.A.
Received 1983 October 24
By common consent, the general theory of relativity has a special aesthetic appeal to those who have studied it. I have chosen to quote Landau and Lifschitz from their Classical Fields in the title of my talk since their magnificent series of volumes, encompassing the whole range of physics, gives to their assessment a special authenticity. Others besides Landau and Lifschitz have applied the epithet 'beautiful' to general relativity. Thus, Pauli, in his well-known article on 'The Theory of Relativity' in the Encyclopadie der Mathematischen Wissenschafiien (1921) has written This fusion of two previously quite disconnected subjects—metric and gravitation—must be considered as the most beautiful achievement of the general theory of relativity. And in a similar vein, Dirac has written There was difficulty reconciling the Newtonian theory of gravitation with its instantaneous propagation of forces with the requirements of special relativity; and Einstein working on this difficulty was led to a generalization of his relativity—which was probably the greatest scientific discovery that was ever made In this lecture, 1 shall attempt to examine the origins and the reasons for the continuing belief that the general theory of relativity represents a beautiful scientific structure; and in this examination 1 shall try to be as objective as possible.
I 1 shall begin with some remarks on the aesthetic impact which a discovery sometimes makes on the discoverer. That Einstein himself felt this aesthetic impact, when he finally arrived at his field equations, is evident from the concluding remark in his first
002
S. Chandrasekhar
preliminary announcement of his equations: Scarcely anyone who fully understands this theory can escape from its magic. Einstein's reaction to his discovery is not very different from Heisenberg's reaction to his discovery that his matrix representation of the position and the momentum coordinates together with his commutation relation led to the correct energy levels of the simple harmonic oscillator. He has written . . . one evening I reached the point where I was ready to determine the individual terms in the energy table [Energy Matrix]. . . . When the first terms seemed to accord with the energy principle, I became rather excited, and 1 began to make countless arithmetical errors. As a result, it was almost three o'clock in the morning before the final result of my computations lay before me. The energy principle had held for all the terms, and I could no longer doubt the mathematical consistency and coherence of the kind of quantum mechanics to which my calculations.pointed. At first, 1 was deeply alarmed. 1 had the feeling that, through the surface of atomic phenomena, I was looking at a strangely beautiful interior, and felt almost giddy at the thought that I now had to probe this wealth of mathematical structure nature had so generously spread out before me. Heisenberg has recalled a meeting with Einstein soon after his discovery of quantum mechanics; and his remarks, at that meeting, as he has recounted them, illuminate the role of aesthetic sensibility in the discerning of great truths about nature: If nature leads us to mathematical forms of great simplicity and beauty . . . we cannot help thinking that they are "true," that they reveal a genuine feature of nature. . . . You must have felt this too: the almost frightening simplicity and wholeness of the relationships which nature suddenly spreads out before us and for which none of us was in the least prepared. It may be argued that the aesthetic sensibility which may have guided Einstein or Heisenberg reflects only on their individuality; and that in any event the aesthetic appeal of a scientific insight is not really relevant to a judgement of its significance. One may indeed contend that the usefulness of any scientific discovery—be it theoretical or experimental—is to be measured only by its consequences. I shall not argue with those who take this pragmatic view. But it is fair to point out that the usefulness' of what one does is not always the prime motive for what one chooses to pursue. For example, Freeman Dyson has quoted Hermann Weyl as having said My work always tried to unite the true with the beautiful; but when 1 had to choose one or the other, I usually chose the beautiful. I shall return later to give some examples of how, in the case of Hermann Weyl at any rate, his choice of the beautiful did eventually turn out to be true as well. But the task 1 now wish to set myself is to explain, as objectively as I can, why the general theory of relativity has had so strong an aesthetic appeal. In this attempt, I wish to be as serious as one is in literary or art criticisms of the works of Shakespeare or Beethoven.
General theory of relativity
003
II At the outset one encounters a curious paradox. It is that the very beauty of the general theory of relativity is sometimes used as an argument for not pursuing it! Thus, Max Born has written It [the general theory of relativity] appeared to me like a great work of art, to be enjoyed and admired from a distance. I am frankly troubled by Bom's remark that the general theory of relativity is to be admired only from a distance. Is one to conclude that the theory does not require study and further development like the other branches of physical science? I find it equally difficult to interpret a statement such as this of Rutherford: The theory of relativity by Einstein, apart from any question of its validity, cannot but be regarded as a magnificent work of art. Apparently, beauty and truth are not to be confused! For the present, I shall not concern myself with the question whether there can be beauty without truth. 1 shall turn instead to consider why a study of the general theory of relativity conduces in one a feeling not dissimilar to one's feelings after seeing a play of Shakespeare or hearing a symphony of Beethoven. But in attempting this task, it is useful to have some definite criteria for beauty in spite of the following view expressed by Dirac and shared by many: [Mathematical beauty] cannot be defined any more than beauty in art can be defined, but which people who study mathematics usually have no difficulty in appreciating. I shall adopt the following two criteria for beauty. The first is that of Francis Bacon There is no excellent beauty that hath not some strangeness in the proportion! ('Strangeness', in this context, has the meaning 'exceptional to a degree that excites wonderment and surprise.') And the second is that of Heisenberg: Beauty is the proper conformity of the parts to one another and to the whole. Ill That the general theory of relativity has some strangeness in the proportion, in the Baconian sense, is manifest. It consists primarily in relating, in juxtaposition, two fundamental concepts which had, till then, been considered as entirely independent: the concepts of space and time, on the one hand, and the concepts of matter and motion on the other. Indeed, as Pauli wrote in 1919, "The geometry of space-time is not given; it is determined by matter and its motion." In the fusion of gravity and metric that followed, Einstein accomplished in 1915 what Riemann had prophesied in 1854, namely, that the metric field must be causally connected with matter and its motion.
004
S. Chandrasekhar
Perhaps the greatest strangeness in the proportion consists in our altered view of space-time with metric as the principal notion. As Eddington wrote: "Space is not a lot of points close together; it is a lot of distances interlocked." There is another aspect of Einstein's founding of his general theory of relativity which has contributed to its uniqueness among physical theories. The uniqueness arises in this way. We can readily concede that Newton's laws of gravitation require to be modified to allow for the finiteness of the velocity of light and to disallow instantaneous action at a distance. With this concession, it follows that the deviations of the planetary orbits from the Newtonian predictions must be quadratic in v/c where v is a measure of the velocity of the planet in its orbit and c is the velocity of light. In planetary systems, these deviations, even in the most favorable cases, can amount to no more than a few parts in a million. Accordingly, it would have been entirely sufficient if Einstein had sought a theory that would allow for such small deviations from the predictions of the Newtonian theory by a perturbative treatment. That would have been the normal way. But that was not Einstein's way: he sought, instead, an exact theory. And his only guides in his search for an exact theory were the geometrical base of his special theory of relativity provided by Minkowski and the principle of equivalence embodying the equality of the inertial and the gravitational mass. The empirical equality of the inertial and the gravitational mass, assumed to be exact, is at the base of the Newtonian theory of gravitation; and Newton gave it its supreme place by formulating it in the opening sentences of his Principia. But the equality, as Weyl has stated, is an 'enigmatic fact'; and Einstein wished to eliminate this enigma. The fact that Einstein was able to arrive at a complete physical theory with such slender guides has been described by Weyl as "one of the greatest examples of the power of speculative thought." There is clearly an element of revelation in the manner of Einstein's arriving at the basic elements of his theory. One feels, as Weyl has expressed, "it is as if a wall which separated us from Truth has collapsed."
IV The general theory of relativity thus stands beside the Newtonian theory of gravitation and motion, as the only examples of a physical theory born whole, as a perfect chrysalis, in the single act of creation of a supreme mind. It is this feature of the general theory of relativity, more than any other, that is normally in one's mind when one describes the theory as a "great work of art to be admired from a distance*'. But for a serious student of relativity, the aesthetic appeal derives even more from discovering that at every level of further understanding, fresh strangenesses in the proportion emerge always in conformity of the parts to one another and to the whole, even as an iridescent butterfly emerges from a chrysalis. I should like to give some illustrations of this feature of the theory. But to the extent they are illustrations, they may reflect my own perspective of the theory. I am sure that others will choose other illustrations. My first illustration will relate to the solutions which the general theory of relativity provides as a basis for the description of the black holes of nature. It is now a matter of common knowledge that black holes are objects so condensed that the force of gravity on their surfaces is so strong that even light cannot escape from them. The most elementary physical ideas combined with the most rudimentary facts
General theory of relativity
005
concerning stars, their sources of energy and their evolution, dictate their occurrence in very large numbers in the astronomical universe. This is not the occasion, and I do not have the time either, to elaborate on these astrophysical matters. I shall turn instead to what the general theory of relativity has to say about them. For this purpose, it is necessary to give a somewhat more precise definition of a black hole than I have given. A black hole partitions the three-dimensional space into two regions: an inner region which is bounded by a smooth two-dimensional surface called the event horizon; and an outer region external to the event horizon which is asymptoticallyflat;and it is required that no point in the inner region can communicate with any point of the outer region. This incommunicability is guaranteed by the impossibility of any light signal, originating in the inner region, crossing the event horizon. The requirement of asymptoticflatnessof the outer region is equivalent to the requirement that the black hole is isolated in space, which means only that far away from the event horizon the space-time approaches the customary space-time of terrestrial physics. In the general theory of relativity we must seek solutions of Einstein's vacuum equations compatible with the two requirements I have stated. It is a startling fact that compatible with these very simple and necessary requirements, the general theory of relativity allows for stationary (i.e., time-independent) black holes exactly a single, unique, two-parameter family of solutions. This is the Kerr family, in which the two parameters are the mass of the black hole and the angular momentum of the black hole. What is even more remarkable, the metric describing these solutions is simple and can be explicitly written down. I do not know if the full import of what I have said is clear. May I explain. As I have already stated, there are innumerable black holes in the present astronomical universe. They are macroscopic objects with masses varying from a few solar masses to millions of solar masses. To the extent they may be considered as stationary and isolated, they are all—every one of them—described exactly by the Kerr solution. This is the only instance we have of an exact description of a macroscopic object. Macroscopic objects, as we see them all around us, are governed by a variety of forces derived from a variety of approximations to a variety of physical theories. In contrast, the only elements in the construction of black holes are our notions of space and time. They are thus, almost by definition, the most perfect among all the macroscopic objects we know. And since the general theory of relativity provides a single unique two-parameter family of solutions for their description, they are the simplest objects as well. As I have said on another occasion, Kerr's discovery of his solution is the only astronomical discovery comparable to the discovery of an elementary particle in physics; but in contrast to elementary particles, the black holes are pristine in their purity. V Again we need not be content with the discovery of the Kerr solution. We can study its properties in a variety of ways: by examining, for example, the manner of the interaction of the Kerr black hole with external perturbations such as the incidence of waves of different sorts. Such studies reveal an analytic richness of the Kerr space-time which one could hardly have expected. This is not the occasion to elaborate on those
006
5. Chandrasekhar
technical matters. Let it suffice to say that, contrary to every prior expectation, all the standard equations of mathematical physics can be solved exactly and explicitly in the Kerr space-time. The Hamxiton-Jacobi equation governing the motion of test particles, Maxwell's equations governing the propagation of electromagnetic waves, the gravitational equations governing the propagation of gravitational waves, and the Dirac equation governing the motion of electrons, all of them can be separated and solved explicitly in Kerr geometry. And the solutions predict a variety of physical phenomena which black holes must exhibit in their interaction with the outside world. Let me illustrate by one particular process, discovered by Roger Penrose, which can take place in such interactions. It is that one can extract, under suitable conditions, the rotational energy of a black hole. When this phenomenon was first investigated, one found that such extraction of energy was accompanied by an increase in the surface area of the black hole. Generalizing this result, Hawking was able to prove an 'area theorem' to the effect that any interaction, experienced by a black hole, in which energy is exchanged, must result in an increase in its surface area. This fact suggests that the surface area of a black hole is in some sense analogous to thermodynamic entropy which has also the monotonic property of always increasing. By considering the quantum mechanics of pressure-free gravitational collapse, Hawking soon showed that this is more than an analogy and that one can, without ambiguity, define not only the entropy of a black hole but a surface temperature as well; and also that there is a flux of radiation from the surface of a black hole with a Planck distribution for the temperature that was assigned. I stated earlier that one of the remarkable features of Einstein's formulation of general relativity was its bringing into a direct relationship the geometry of the spacetime with its content of matter and motion. It is "this fusion of two previously quite disconnected notions" that Pauli found as the "most beautiful achievement of the general theory of relativity." We now find in Hawking's synthesis a still grander fusion of geometry, matter, and thermodynamics. There is clearly no lack in the strangeness in the proportion which a further study of relativity does not reveal.
VI Let me consider one last illustration. It relates to certain singularity theorems proved by Penrose and Hawking. The theorems state, in effect, that the occurrence of singularities in space-times is generic to general relativity. Roughly speaking, what this statement means is that during the course of evolution of material objects, there exist 'points of no return' such that the trespassing of these points will necessarily lead, inexorably, to singularities. This theorem provides in fact the strongest reason for our present belief that our universe started with an initial singularity. The reason is that from the existence of the three-degree microwave-radiation, we can conclude that the universe retained its present homogeneity and isotropy when its radius was some one thousand times smaller than the present. It follows from this result and some additional astronomical facts that the universe was already then (or a little earlier) at a point of no return; and the inference of an initial past singularity cannot be avoided. The problems associated with the conditions just preceding the initial singularity thus become a necessary part of current investigations both in cosmology and in physics.
General theory of relativity
007
VII So far I have considered the aesthetic appeal of the general theory of relativity in the manner of its founding and in the matter of its implications. But Poincare, who has often emphasized the role of beauty in the motivations for scientific pursuits, has also stated that the "value of a discovery is to be measured by the fruitfulness of its consequences." I shall therefore consider some of the "fruitful consequences" of the general theory of relativity. Since astronomy is the natural home of general relativity, we must seek for its consequences in astronomy. I shall consider two such consequences. Both of them relate to certain crucial respects in which considerations of relativity have altered the astrophysicist's views relative to the stability of stars and stellar systems. It is well known that in the framework of the Newtonian theory, the condition for the dynamical stability of a star, derives from its modes of radial oscillations and, that for stability the average ratio of the specific heats y (defined as the ratio of the fractional changes in the pressure and in the density for adiabatic changes) must exceed 4/3. Alternatively, a star will become dynamically unstable if y, or some average of it, is less than 4/3. This Newtonian condition is changed in the framework of general relativity: a star with an average ratio of specific heats y, no matter how high, will become unstable if its radius falls below a certain determinate multiple of the Schwarzschild radius, Rs = 2GM/c2 (where M denotes the mass of the star, G is the constant of gravitation, and c is the velocity of light). It is this fact which is responsible for the existence of a maximum mass for stable neutron stars. I may parenthetically point out that this important result is closely related to an early deduction of Karl Schwarzschild that a star in hydrostatic equilibrium must necessarily have a radius exceeding §RS: this is the radius at which a star, with a ratio of specific heats tending to infinity, becomes unstable. This instability of relativistic origin, discovered some twenty years ago, plays a central role in all current discussions pertaining to the onset of instability during the course of evolution of massive stars prior to gravitational collapse. There is another consequence of general relativity for the stability of neutron stars. The instability to which I now refer was discovered some ten years ago and derives from a dissipative phenomenon which general relativity naturally builds into the theory of non-axisymmetric oscillations of gravitating masses. The dissipation results from the emission of gravitational radiation with accompanying loss of energy and angular momentum. The manner in which this mode of dissipation of energy and angular momentum induces instability is in some ways similar to the manner in which viscous damping sometimes induces instability. It now appears, especially from the work of John Friedman, that this mode of instability sets a limit to the rotation of pulsars and bears on the stability of fast pulsars like the ones that have recently been discovered. It is clear, then, that there are fruitful consequences of the general theory of relativity for the astronomer's view of the universe. He need not be content with admiring general relativity from a distance.
IX I now turn to a somewhat more general question concerning the relation of truth to beauty in science.
008
S. Chandrasekhar
I made a reference earlier to a statement of Weyl's to the effect that in his work he always tried to unite the true with the beautiful and that, when he had to make a choice he generally chose the beautiful. An example which Weyl gave was his gauge theory of gravitation, developed in his Raum, Zeit, und Materie (Space, Time, and Matter, 1918). Weyl became convinced that his theory was not true as a theory of gravitation; but he nevertheless kept it alive because it was beautiful. But much later, it did turn out that Weyl's instinct was right after all: the formalism of gauge invariance was incorporated into quantum electrodynamics. A second example is provided by the two-component relativistic wave-equation of the massless neutrino. Weyl discovered this equation and the physicists ignored it for some thirty years because it violated parity invariance. And again it turned out that Weyl's instinct was right: he had discerned truth by trusting to what he conceived as beautiful. A similar example is provided by Kerr's discovery of his solution. Kerr was not seeking solutions that would describe black holes. He was seeking instead solutions of Einstein's equation which had a very special algebraic property. But once he had discovered his solution, he could show quite readily that it did indeed describe a black hole. But its uniqueness for representing black holes was established only ten years later by Edward Robinson. The foregoing examples provide evidence that a theory developed by a scientist with an exceptionally well-developed aesthetic sensibility can turn out to be true even if at the time of its formulation, it did not appear relevant to the physical world. It is, indeed, an incredible fact that what the human mind, at its deepest and most profound, perceives as beautiful finds its realization in external nature. What is intelligible is also beautiful. We may well ask: how does it happen that beauty in the exact sciences becomes recognizable even before it is understood in detail and before it can be rationally demonstrated? In what does this power of illumination consist? These questions have puzzled many since the earliest times. Thus, Heisenberg has drawn attention, precisely in this connection, to the following thought expressed by Plato in the Phaedrus: The soul is awestricken and shudders at the sight of the beautiful, for it feels that something is evoked in it, that was not imparted to it from without by the senses, but has always been already laid down there in the deeply unconscious region. The same thought is expressed in the following aphorism of David Hume: Beauty in things exists in the mind which contemplates them. Kepler was so struck by the harmony of nature as revealed to him by his discovery of the laws of planetary motion that in his Harmony of the World, he wrote: Now, it might be asked how this faculty of the soul, which does not engage in conceptual thinking and can therefore have no prior knowledge of harmonic relations, should be capable of recognizing what is given in the outward world.... To this, I answer that all pure Ideas, or archetypal patterns of harmony, such as we are speaking of, are inherently present in those who are capable of apprehending them. But they are not first received into the mind by a
General theory of relativity
009
conceptual process, being the product, rather, of a sort of instinctive intuition and innate to those individuals. More recently. Pauli, elaborating on these ideas of Kepler, has written: The bridge, leading from the initially unordered data of experience to the Ideas, consists in certain primeval images pre-existing in the soul—the archetypes of Kepler. These primeval images should not be located in consciousness or related to specific rationally formulizable ideas. It is a question, rather, of forms belonging to the unconscious region of the human soul, images of powerful emotional content, which are not thought, but beheld, as it were, pictoriaily. The delight one feels, on becoming aware of a new piece of knowledge, arises from the way such pre-existing images fall into congruence with the behavior of the external objects. . . . Pauli concludes with One should never declare that theses laid down by rational formulation are the only possible presuppositions of human reason. It is clear that following these thoughts one is dangerously led into the path of the mystical. I shall desist following this path but conclude instead by quoting two ancient mottoes: The simple is the seal of the true and Beauty is the splendour of truth.
THE PURSUIT OF KNOWLEDGE" S. Chandrasekhar I I am naturally pleased and honored to be among the first three Regents' Fellows in the fellowship program inaugurated by the Smithsonian Institution last year.
The election to
a Regents' Fellowship has provided me with the enviable opportunity to spend three months at the Smithsonian-Harvard Center for Astrophysics and to be associated with so many active astrophysicists of youth, of promise, and of accomplishment.
Besides, it has enabled me to bring to the
Center, for the period of my stay, a young and brilliant collaborator of mine, Basilis Xanthopoulos from Greece. My present visit to the Smithsonian-Harvard Center recalls to me my first visit to Harvard 44 years ago during the winter of 1936.
Except for two or three very brief
visits, this will be my first, of any duration, since the one of 44 years ago.
A return, after almost an entire life-
span, to a place one has visited as a youth, with all the expectations of youth, is not without its anxieties.
What
these anxieties may be are stated with clarity by T.S. Eliot in his Family Reunion.
In this play, Harry, Lord Monchensey
Paper read at the Regents' Fellows Colloquium on "The Pursuit of Knowledge' arranged by the Smithsonian Institution in Washington, D.C. on May 5, 1980.
1263
is returning to his estate at Wishwood after many years of Wandering in the tropics Or, against the painted scene of the Mediterranean. In anticipation of his return, his mother Amy, Lady Dowager Monchensey has seen to it that Nothing has changed at Wishwood, Everything is kept as it was when he left. But Amy's sister, Agatha, is apprehensive. Yes!
As she says
I mean at Wishwood he will find another Harry.
The man who returns will have to meet The boy who left.
Round by the stables,
In the coach house, in the orchard, In the plantation, down the corridor That led to the nursery, round the corner Of the new wing, he will have to face him And it will not be a jolly corner. When the loop of time comes - and it does not come to everybody The hidden is revealed, and the spectres show themselves. I shall leave these anxieties aside and address myself to the topic of this colloquium, namely 'The Pursuit of Knowledge'. II May I frankly confess at the outset that I find the juxtaposition of the two words 'pursuit' and somewhat ambiguous.
'knowledge'
Let me explain the nature of my difficulties.
1264
'Pursuit' has the common meaning of a chase in hunting; and, characteristic of our times, we are also familiar with the word in the combination
'pursuit-plane'.
Is this as-
sociation which the word 'pursuit' evokes intended?
And if
it is, are we to conclude that knowledge, like the fox in a chase, or the enemy plane in a hot pursuit, is something whose existence we know of in advance and our pursuit is to aim for it.
Of course, some aspects of what we include
under knowledge are in this category.
Thus, to unearth the
fossil remains of creatures of long past, or the relics of bygone civilizations, to scale the highest peaks, or to fathom the deepest oceans, are all endeavors of high human aspiration. But one may still ask:
Is knowledge, then, something
which we seek to attain in the same spirit as the mountaineer who aspires to ascend the Everest 'because it is there'?
If
that is the case, what are we to understand when we are told that research is a quest after the unknown to charter territories of whose existence we may not even be aware when we started?
Kepler did not know, when he began his long and
arduous analysis of the centuries of accumulated observations on planetary motion^ that buried in that mass of details were the simple laws he discovered.
Neither did Newton know, be-
fore he saw the apple fall, that Kepler's laws could be understood so simply in terms of his laws of motion and of gravitation.
Perhaps
I shall be accused of quibbling.
For one may,
in truth, say that in the pursuit of scientific knowledge if I may be pardoned for that combination of words! - while one does not aim for something which is material and concrete, one does aim for the enlargement of that order and that harmony which are the key signatures of Nature.
Indeed, for a
scientist, the order, the harmony, the uniformity, and the universality of the laws of Nature are as real as the peak of the Everest is for the mountaineer.
And so it is in the
other branches of abstract knowledge. But is that all there is to our quest for knowledge? Do we wish, for example, to quantify new knowledge by the extent to which others can share in it and still others can make use of it for human delight or for human welfare? And if that is our wish, what value do we attach to the refinement of one's own perception and to the enlargement of one's own vision?
Is there no real content to Wordsworth's
famous lines on Newton: The marble index of a mind for ever Voyaging through strange seas of Thought alone. Indeed, there is ample evidence that the very greatest artists, in their ennobled maturity, withdraw into themselves. Here, for example, is T. S. Eliot commenting on the late plays of Shakespeare.
It seems to me to correspond to some law of nature that the work of a man like Shakespeare, whose development in the course of his career was so amazing, that it should reach, as in Hamlet, the point at which it can touch the imagination and feeling of the maximum number of people to the greatest possible depth and that, thereafter, like a comet which has approached the earth and continued away on its course, he should gradually recede from view until he tends to disappear into his private mystery. What T. S. Eliot has said of Shakespeare applies equally to Beethoven.
In his last compositions and
especially in his late quartets, Beethoven was indeed 'voyaging through strange seas of Thought alone' voyaging, in fact, to enlarge his own personal vision. This attempt by the greatest minds to enlarge their personal vision is, I believe, manifest also in the remoteness and in the laminated and glacial style of Newton's Principia.
The lasting value of the Principia
lies more in Newton's vision of the universe which it manifests than even in the surpassing quality of his dis coveries which it summarizes and organizes.
Ill So far, I have addressed myself, only tangentially, to the questions we were asked in the context of this Colloquium. The questions relate to the rewards and the satisfactions in a life devoted to research and to the attitudes one may develop in the twilight of one's life.
I do not find these
questions easy to answer; but I shall try. First, with regard to the rewards and the satisfactions. Perhaps, I should eliminate, ex cathedra, that the rewards of scholarship consist in celebrity and in public honors. The supreme Magisters of Castalia, in Hermann Hesse's Glass Bead Game, had learned to renounce them.
I suppose that one
does renounce them in the end - at least, one feels, that one ought to transcend them. as that.
But the matter is not as simple
None of us are so immune to human sensibilities that
we are not all, to some extent, sensitive to the approval of our colleagues whom we respect.
And I am sure that all of us
hope, each in his own way, that posterity will assign to us our due and humble places so long as we persist and persevere at the limits of our capabilities. But posterity can be harsh.
Here, for example, is John
Ruskin's criticism of the English painter Sir Joshua Reynolds: Why did not Sir Joshua - or could not - or would not Sir Joshua - paint Madonnas?
... Yet!
While
we acknowledge the discretion and simple heartedness of these men ... we have to remember .... that amiable
discretion is not the highest virtue, nor to please the frivolous, the best success ... There is probably some strange weakness in the painter, and some fatal error in the age, when in thinking over the examples of their greatest work, for some type of culminating loveliness, or veracity, we remember no expression either of religion or heroism, and instead of reverently naming a Madonna di San Sisto, can only whisper modestly, "Mrs. Pelham feeding chickens". If one recognizes that one can never come to painting a Madonna, what, then, are the satisfactions and the rewards. I suppose that one must count them in those brief moments of sudden insight which occurs to one on rare occasions. may never come to painting a Madonna.
One
But, perhaps, in
capturing on canvas, the rugged lines in the face of Mrs. Pelham, etched by the toils of her life, the painter may have experienced a sudden insight into the sadness of the human condition which he may cherish all his life.
And so
it is in all other walks of creative effort. While one may grant that the rare moments of illumination one experiences during the course of one's life are the precious rewards, one is still troubled: for, one may ask: Is it the fate of a creative artist or scientist, in his declining years, only to harken back to those rare and brief moments?
Is he condemned to live only in the memory of his
past moments of illumination?
The answer, it seems to me,
is that, if one is not to be so condemned, one must seek in one's enlarged and refined perception a source for quiet contemplation.
In other words, 'Voyage through strange seas
of Thought alone'.
But even this may not always be possible
as Hermann Hesse poignantly describes in his portraiture of Joseph Knecht in his Glass Bead Game - a book to which I have already made one reference. The Glass Bead Game is the wondrous tale of one, who in his youth aspired to the ideals of Castalia and who, in time, rose to become the supreme Magister Ludi.
But in
the end, in deep doubts, he renounces his office, exiles himself from Castalia, only to be drowned accidentally in a Swiss lake. IV I am afraid that, once again, I have spoken only tangentially to the questions that were addressed.
But
let me try to answer directly the question why one, fully aware of one's inherent and often insurmountable limitations, nevertheless devotes oneself to scholarship and to a life of constant endeavor, many failures, and fewer successes? The answer is, as T. S. Eliot has given in his 'Confidential Clerk': It is strange, isn't it, That a man should have a consuming passion To do something for which he lacks the capacity?
1270
Stated Meeting Report
many an unexpected view which was pleading to our eyes.
T h e P e r c e p t i o n of B e a u t y a n d t h e P u r s u i t of Science
It is also not uncommon that when a great scientist, by virtue of his genius, gains an insight or an understanding of depth and rotundity, he perceives, and is perhaps even linded by, the beauty that has been revealed to him. Thus Einstein, in the concluding sentence of his first announcement of his equations of the general theory of relativity. wrote:
Subrahmanyan Chandrasekhar
T h e subject I have chosen — the perception of beauty and the pursuit of science — is not an easy one to talk about if one is to avoid the pitfalls of being obvious and trivial or being pedantic and dilettantish. It is difficult to say anything original on this subject. What T.S. Eliot once said in a lecture entitled "Versification in Shakespeare" applies equally well to my subject: "One can always save the subject by magnificent quotations." What I shall present will, in effect, be a collage of what some great men of science have said. If there is any novelty in what I shall be saying, it will be in the selection, the sequence, and the arrangement of the quotations.
I T h e source of one's aesthetic pleasure in the study of the natural sciences is fairly obvious. As Heisenberg has stated, "The beauty of Nature is also reflected in the beauty of natural science." And Poincare was even more explicit: It is because simplicity and vastness are both beautiful that we seek by preference simple facts and vast facts; that we take delight, now in following the giant courses of the stars, now in scrutinizing with a microscope that prodigious small ness which is also a vastness, and now in seeking in geological ages the traces of the past that attracts us because of its remoteness. In the context of mathematics, Hilbert, in his memorial address for Minkowski, said: It appeared to us as a flowering garden. In this garden there were well-worn paths where one might look around at leisure and enjoy oneself without effort. . . . But we also liked to seek out hidden trails and discovered 14
Scarcely anyone who has fully understood this theory can escape from its magic; the theory represents a genuine triumph of the methods of the absolute differential calculus founded by Gauss, Riemann, Cristoffel. Ricci, and Levi-Civita. This "shuddering before the beautiful" is evident, for example, from the following conversation with Einstein that Heisenberg records: If nature leads us to mathematical forms ol reat simplicity and beauty . . . that no one as previously encountered, we cannot help thinking that they are "true," that they reveal a genuine feature of nature. . . . You must have felt this too: the almost frightening simplicity and wholeness of the relationships which nature suddenly spreads out before us and for which none of us was in the least prepared. It is clear that Heisenberg is here referring to his own discovery of the laws of quantum mechanics and Einstein's discovery of the general theory of relativity. II At this point, I want to digress for a moment to the question of whether there is an\ real dichotomy in the perception of beautv in nature by a poet and a scientist. There are, for example, the well-known lines of Wordsworth and of Keats which emphasize this difference A fingering slave, One that would peep and botanize Upon his mother's grave? Sweet is the lore which Nature brings; Our meddling intellect 15
1271
Misshapes the beauteous forms of things: We murder to dissect. (Worasworth) Do not all charms fly At the mere touch of cold philosophy? There was an awful rainbow once in heaven: We know her woof, her texture; she is given In the dull catalogue of common things. Philosophy will dip an Angel's wings. (Keats) Perhaps Feynman, without being aware of Wordsworth's lines, effectively responds as follows in his own style: I have a friend who's an artist. He'll hold up a flower and say, "Look how beautiful it is," and I'll agree. But then he'll say, "I, as an artist, can see how beautiful a flower is. But you, as a scientist, take it all apart and it becomes dull." First of all, the beauty he sees is available to other people—and to me, too, I believe. Although I might not be quite as refined aesdietically as he is, I can [also] appreciate the beauty of a flower. But at the same time, I see much more in a flower than he sees. I can imagine the cells inside, which also have a beauty. There are other complicated actions of the cells, and other processes. The fact that the colors in die flower have evolved in order to attract insects to pollinate it is interesting; that means insects can see colors. That adds a question: does this aesthetic sense we have also exist in lower forms of life? There are all kinds of interesting questions that come from a knowledge of science, which only adds to the excitement and mystery and awe of a flower. It only adds. I don't understand how it subtracts. And Feynman continues: The same thrill, the same awe and mystery, comes again and again when we look at any question deeply enough. With more knowledge comes a deeper, more wonderful mystery, luring one on to penetrate deeper sail. I should not, however, like to leave you with the impression that all poets share the views of Wordsworth and Keats. Here are some lines from Shelley with a different perception:
16
The lightning is his slave; heaven's utmost deep Gives up her stars, and like a flock of sheep They pass before his eye, are numbered, and roll on! The tempest is his steed, he strides the air; And the abyss shouts from her depth laid bare, Heaven, hast thou secrets? Man unveils me; I have none. (Shelley, Prometheus Unbound) III So far I have concerned myself with the perception of beauty in the physical sciences. It is more difficult to decide the extent to which one's sensibility to the aesthetic aspects of science is a factor in one's own pursuit of science. T h e practice of mathematics appears to have more affinity with the practice of the arts. In any event, practitioners of mathematics are more explicit on this question. T h e following statement by G.H. Hardy is particularly illuminating. A mathematician, like a painter or a poet, is a maker of patterns. If his patterns are more permanent than theirs, it is because they are made with ideas. A painter makes patterns with shapes and colours, a poet with words. A painting may embody an "idea," but the idea is usually commonplace and unimportant. In poetry, ideas count for a good deal more; but, as [A.E.] Housman insisted, the importance of ideas in poetry is habitually exaggerated: "I cannot satisfy myself that mere are any such things as poetical ideas. . . . Poetry is not the thing said but a way of saying it." Not all the water in the rough rude sea Can wash die balm from an anointed King. Could lines be better, and could ideas be at once more trite and more false? The poverty of the ideas seems hardly to affect the beauty of the verbal pattern. A mathematician, on the other hand, has no material to work with but ideas, and so his patterns are likely to last longer, since ideas wear less with time than words. The mathematician's patterns, like the painter's or the poet's, must be beautiful: the
17
ideas, like the colours or the words, must fit together in a harmonious way. Beauty is the first test: there is no permanent place in the world for ugly mathematics. . . . The beauty of a mathematical theorem depends a great deal on its seriousness, as even in poetry the beauty of a line may depend to some extent on the significance of the ideas which it contains. I quoted two lines of Shakespeare as an example of the sheer beauty of a verbal pattern; but After life's fitful fever he sleeps well seems still more beautiful. The pattern is just as fine, and in this case die ideas have significance and the thesis is sound, so that our emotions are stirred much more deeply. The ideas do matter to the pattern, even in poetry, and much more, naturally, in mathematics; but I must not try to argue me question seriously. And John Von Neumann expressed very similar views in a lecture which he gave in Chicago in 1946 and which I had the pleasure of attending: One expects a mathematical theorem or a mathematical theory not only to describe and to classify in a simple and elegant way numerous and a priori disparate special cases. One also expects "elegance" in its "architectural," structural makeup. Ease in stating the problem, great difficulty in getting hold of it and in all attempts at approaching it, then again some very surprising twist by which the approach, or some part of the approach, becomes easy, etc. Also, if the deductions are lengtiiy or complicated, diere should be some simple general principle involved, which "explains" die complications and detours, reduces the apparent arbitrariness to a few simple guiding motivations, etc. These criteria are clearly those of any creative art, and die existence of some underlying empirical, worldly motif in die background— often in a very remote background—overgrown by aestiieticizing developments and Followed into a multitude of labyrinthine variants—all this is much more akin to die atmosphere of art pure and simple man to that of the empirical sciences.
How, in fact, does one decide which things in mathematics are important and which are not? Ultimately, die criteria have to be aesdietic ones. There are otiier values in matfiematics, such as depth, generality and utility. The ultimate values seem simply to be aesdietic; diat is, artistic values such as one has in music or painting or any odier art form. . . . But, although in other subjects the criteria of importance are not quite the same as in mathematics, often they are aesthetic criteria again. .. . One must also appreciate that mere are really several radier different kinds of aesthetic criteria even widiin mathematics itself. For example, die question as to whether die proof of a theorem is beautiful. .. . And the beauty in die content of a theorem or the attractiveness of the abstract ideas contained in a proof are different again from elegance in presentation of a result. . . . What... constitutes elegancef?] It is naturally a difficult question, but I think Professor Atiyah made some remark [to the effect] diat elegance is more or less synonymous with simplicity. . . . Elegance and simplicity are certainly things diat go very much together. But nevertheless it cannot be quite die whole story. 1 dunk perhaps one should say it has to do with unexpected simplicity, where one imagines mat tilings are going to be complicated but suddenly they turn out to be much simpler than expected. It is not unnatural that this should be pleasing to the mind.... So I think that unexpected simplicity is quite a significant thing.... mere is the element of connecting together different ideas diat one had not expected to be related. This is something that frequendy occurs in mathematics, and it has a special beauty of its own. . . . Two problems which seem unrelated can both involve one in a certain amount of work; but as soon as it is realised that the two problems are really basically die same problem, the amount of work may be halved—and the pleasure of unexpected simplicity is consequendy felt! Penrose concludes widi a question which has been one of my own principal concerns:
And Roger Penrose, perhaps the most outstanding relativist of our time, is even more explicit. I shall quote from him rather extensively.
There is also a more subde role played by aesthetics, in connection with research, namely as a means for obtaining results. This is really an important aspect of die role of aesdietics which I do not think has been discussed very much. It is a mysterious thing, in fact, how something which looks attractive
18
19
1273
may have a better chance of being true than something which looks ugly. IV Hardy, Von Neumann, and Penrose, who effectively span the mathematics of this century, are clear, explicit, and concordant in their views. When we turn to theoretical physics, one cannot obtain a comparable consensus, perhaps because the practice of physics is rather more remote from the arts than mathematics. The manner in which Maxwell expresses himself when he writes about Faraday is indicative of this difference. We have, first, the careful observation of selected phenomena, then the examination of the received ideas, and the formulation, when necessary, of new ideas; and, lasdy, the invention of scientific terms adapted for the discussion of the phenomena in the light of the new ideas. The high place which we assign to Faraday in electromagnetic science may appear to some inconsistent with the fact that electromagnetic science is an exact science, and that in some of its branches it had already assumed a mathematical form before the time of Faraday, whereas Faraday was not a professed mathematician; and in his writings we find none of these integrations of differential equations which are supposed to be of the very essence of an exact science. Open Poisson and Ampere, who went before him, or Weber and [Von] Neumann, who came after him, and you will find their pages full of symbols, not one of which Faraday would have understood. It is admitted that Faraday made some great discoveries, but if we put these aside, now can we rank his scientific method so high without disparaging the mathematics of these eminent men? It is true that no one can essentially cultivate any exact science without understanding the mathematics of that science. But we are not to suppose that the calculations and equations which mathematicians find so useful constitute the whole of mathematics. The geometry of position is an example of a mathematical [concept] established without the aid of a single calculation. Now Faraday's lines of force occupy the same position in electromagnetic science that pencils of lines
20
do in the geometry of position. They furnish a method of building up an exact mental image of the thing we are reasoning about. The way in which Faraday made use of his idea of lines of force in co-ordinating the phenomena of magneto-electric induction shows him to have been in reality a mathematician of a very high order—one from whom die mathematicians of the future may derive valuable and fertile methods. We are probably ignorant even of the name of the science which will be developed out of die materials we are now collecung, when the great philosopher next after Faraday makes his appearance! Maxwell is, of course, the "great philosopher next after Faraday." But he is too self-effacing to say this. While Maxwell's views are implicit in what I have read, Heisenberg is more explicit. He begins with the definition, "beauty is the proper conformity of the parts to one another, and to the whole," ana continues: Understanding of the colorful multiplicity of phenomena comes about by recognizing in diem unitary principles of form, which can be expressed in die language of mathematics. By this, too, a close connection is established between die intelligible and the beautiful. For if the beautiful is conceived as a conformity of the parts to one anodier and to die whole, and it, on the other hand, all understanding is first made possible by means of this formal connection, die experience of the beautiful becomes virtually identical with die experience of connections either understood or at least guessed at. It is characteristic of Heisenberg that in this same context he should have said: Pythagoras is said to have made the famous discovery that vibrating strings under equal tension sound together in harmony if their lengths are in a simple numerical ratio. The mathematical structure, namely die numerical ratio as a source of harmony, was certainly one of die most momentous discoveries in die history of mankind. He concludes with two Latin mottoes which, in translation, read "The simple is the seal of the true" and "Beauty is die splendour of truth." 21
1274
At age 60: I wish now to address myself to the most difificult pan of my lecture: the question of whether one's aesthetic motivations, which derive from one's early perceptions, are positive guides for one's later pursuits. I shall consider two examples: the cases of Dirac and of Einstein. Dirac's insistence on aesthetic criteria in formulating fundamental laws of physics is well known. Perhaps the most often quoted saying of his is "It is more important to have beauty in one's equations than to have them fit the experiments." Here, Dirac amplifies this statement: [The] research worker, in his efforts to express the fundamental laws of Nature in mathematical form, should strive mainly for mathematical beauty. He should take simplicity into consideration in a subordinate way to beauty. . . . It often happens that the requirements of simplicity and beauty are the same, but where they clash the latter must take precedence. Dirac made these statements in the context of his own early work. Thus, he writes: My own early work was very much influenced by Bohr orbits, and 1 had the basic belief that Bohr orbits would provide the clue to understanding atomic events. That was a mistaken belief. . . . I had to have some more general basis for my work, and the only reliable basis I could think of which was sufficiendy general. . . was to set up a principle of mathematical beauty: to say mat we don't really know what the basic equations of physics are, but they have to have great mathematical beauty. We must insist on this, and that is the only feature of the equations that we can have confidence in and insist on.... And he maintained these views with remarkable consistency in his later years. Thus, at age 36: As time goes on, it becomes increasingly evident that the rules which the mathematician finds interesting are the same as those which Nature has chosen.
22
I mink it's a peculiarity of myself that I like to play about with equations, just looking for beautiful mathematical relauons which maybe don't have any physical meaning at all. Sometimes they do. At age 78: A good deal of my research work in physics has consisted in not setting out to solve some particular problem, but simply examining mathematical quantities of a kind that physicists use and trying to fit them together in an interesting way, regardless of any application that the work may nave. It is simply a search for pretty mathematics. The remarkable aspect of these various statements of Dirac is tnat they were all made subsequent to the great discoveries that he made prior to 1933. A letter that Freeman Dyson wrote to me in this context is revealing. The question remains, why does Dirac believe so strongly in mathematical beauty as a guide to physical truth? I believe one can explain this as a reaction to the stumbling way in which he discovered the Dirac equation . Like Einstein, he was led to the equation not by mathematical elegance but by physical necessity. However, Dirac's physical reasoning was based on three postulates: (1) The equation must be relativistically invariant, (2) The equation must be linear in the time-derivative, so as to reduce to the Schr6dinger equation in the non-relativisuc limit, (3) The equation must lead to a conserved positive-definite quantity which can be interpreted as a probability-density. These postulates led him uniquely to the Dirac equation. However, it turned out a few years later that Pauli and Weisskopf could construct an equally good relativisuc theory of particles with zero spin based on the second-order wave-equations. Pauli and Weisskopf abandoned Dirac's postulates 2 and 3 and still their theory was physically reasonable. So diere was no longer any compelling physical logic leading to die Dirac equation. I believe diat it was at this time
23
1275
(1934) that Dirac shifted his ground and began to feel that the Dirac equation must be rignt because it was uniquely beautiful. The physical logic turned out to be flawed, but die equation was right anyway. Therefore one should not trust the physical logic but should only trust the mathematical beauty. It is a noteworthy fact that Dirac's own judgment of his subsequent efforts in science was not positive. He stated, I have spent many years searching for a Hamiltonian to bring into the theory and have not yet found it. I shall continue to work on it as long as I can, and other people, I hope, will follow along such lines. Even more explicitly, at age 69 Dirac summed u p his work as follows: My own contributions since [the] early days have been of minor importance. And the judgments of those far more competent than I nave agreed with Dirac's assessment. Thus, Pais has written, With Dirac's Solvay report (1933) his exquisite burst of creativity at the outer fronuers of physics, spanning eight years, comes to an end. . . . He continued to show his high mathematical inventivity and craftsmanship but no longer that almost startling combination of novelty and simplicity that mark his heroic period. [And] from the early 1950s on, Dirac went his own lonely way. VI The case of Einstein is remarkably similar. I quoted earlier Einstein's sentiment with regard to his theory of relativity at the moment of its consummation: "Scarcely anyone who has fully understood this theory can escape from its magic. . . ." But Einstein's appreciation of the underlying geometrical base of his theory was an acquired taste. Einstein's earlier reaction to Minkowski's geometrical formulation of his special theory of relativity was not a very positive one. He is quoted by Pais to have said that Minkowski's formulation was "uberfliissige Gelehrsamkeit." Indeed, even at the time he derived his basic equations, Einstein's mastery 24
of Riemannian geometry was not complete It will take me too far to substantiate this statement. Instead, I shall quote what Pais and Dyson have said in this connection. I do not consider it blasphemous to suggest . . . that mathematics was not Einstein's strong point. He said so more than once to me. One might go further and add that when. in later years, E did accept beauty (or simplicity) as the guiding principle, he then had litde to show for it. (Pais) When one looks at the actual history of General Relativity, it is clear that Einstein was driven reluctandy into elegant mathematics by the demands of the physics. He would have avoided all this beautiful mathematics if he had found a way to do without it. (Dyson) I do not think it is an exaggeration to say that Einstein "escaped the magic of his theory": his own contributions to the theory after its founding were at best marginal. It is, of course, well known that Einstein spent all his later years in a fruitless search for a unified field theory. And he seems to have corresponded regularly with Besso, a friend of his youth, explaining his ideas. Of this correspondence, which has been published, Jeremy Bernstein has written: During this period, Einstein was pursuing his unified field theory. From our perspective, it is difficult to understand what this program was really about. It looks, at least to me, like an all but random shuffling of mathematical formulas, with no physics in view. . . . Besso took all his old friend's attempts extremely seriously, and Einstein gave him detailed explanations of his various formal manipulations. It was a dialogue that somehow reminds me of the plays of Samuel Beckett. VII T h e characteristic (illustrated by the examples of Dirac and Einstein) that men of undoubted genius should have found that the illumination which resulted in their great discoveries was of no avail in their further endeavors appears to be a general phenomenon. T h e case of Yukawa is instructive. There is 25
1276
much that is revealing in his book Creativity and Intuition. And Kemmer summarizes it succinctly: Hideki Yukawa's mind throughout his life remained very much an Oriental one. We learn that of the ancient Chinese writings that he discovered early in life, the ones that made the greatest and most lasting impression on him were those of Taoism in general and of Chuangtse in particular. . . . To Yukawa the awareness of nature, in a much more intuitive way than any Westerner would accept as part of scientific thinking, appeared to be a vital ingredient in creativity. He felt not only that his own success in moving theoretical physics a step further owed something to mis way of thinking, but diat an element of it can be seen in such creative acts as Heisenberg's formulation of the uncertainty principle. While accepting the fact that his later mental struggles to discern the nature of particles did not lead to any breakthrough, he expresses the conviction mat a more Oriental approach is a better way to deeper understanding than the present pursuit of ever greater detail with an ever greater mass of facts produced by more and more sophisticated experimentation. VIII I am sorely puzzled by the examples of Einstein, Dirac, and Yukawa. Perhaps it is my prejudice. But it seems to me that one learns a different lesson from the examples of Beethoven and Shakespeare. There is much that one can say about either of them. But I must be content with a few brief quotations. First, with respect to Beethoven: By his early forties, Beethoven had composed his eight symphonies, his five piano concertos, his one violin concerto, his twentyfive piano sonatas, his eleven quartets, his seven overtures, his one opera, and his one Mass. At the age of forty-two, with this magnificent pile of compositions behind him, Beethoven practically stopped composing for the next seven years. The fruits of his meditation—so they must have been—came after mis period of quiescence in a manner mat is perhaps without parallel in musical history. From the first symphony written in 1801 to the eighth symphony written in 1812, it is 26
essentially the same Beethoven: it is. in fact. the Beethoven of the common understanding. But the Beethoven of the ninth svmphony, of die Mass in D, of the last four piano sonatas, and, most of all, the last five quartets, is an altogeth«r different Beethoven. And yet, in 1817, at the age of forty-seven, when the long period of meditation, during which Beethoven composed very little, was coming to an end, he said to Cipriani Potter with transparent sincerity, "Now, I know how to compose." I do not believe that there has been any scientist, past forty, who could say, "Now, I know how to do research." And this, to my mind, is the center and the core of the difference: the apparent inability of a scientist to continually grow and mature. Reviewing the life and work of Beethoven, J.W.N. Sullivan sums him u p as follows: One of the most significant facts, for the understanding of Beethoven, is that his work shows an organic development up until the very end. . . . The greatest music Beethoven ever wrote is to be found in the last string quartets, and the music of every decade before the final period was greater than its predecessor. Next, with respect to Shakespeare. Here is J. Dover Wilson on Shakespeare's tragedies. From 1601 to 1608 he is absorbed in tragedy; and the path he treads during these eight years may be likened to a mountain track which, rising gendy from the plain, grows ever narrower, until at the climax of the ascent it dwindles to the thinnest razor-edge, a glacial arete, with the abyss on either hand, and then once again grows secure for foothold as it broadens out and gradually descends into the valley beyond. Eight plays compose this tragic course. The first, Julius Caesar, written a little before the tragic period proper, is a tragedy of weakness, not of evil. In Hamlet, the forces of evil are active and sinister, though still the prevailing note is weakness of character. Othello gives Shakespeare's earliest creation of a character wholly evil, and at the same time Iago's victim is blameless—human weakness is no longer allowed to share the responsibility with heaven. King Lear carried us right to the edge of the abyss, for here horror is piled upon horror and pity on pity, 27
1277
to make the greatest monument of human misery and despair in the literature of the world. . . . Shakespeare came very near to madness in Lear. I began with a quotation from T.S. Eliot. Let me conclude with Eliot on Shakespeare: The standard set by Shakespeare is that of continuous development from first to last, a development in which the choice both of theme and of dramatic and verse technique in each play seems to be determined increasingly by Shakespeare's state of feeling by the particular stage of his emotional maturity at the time. . . . We may say confidently that the full meaning of any one of his plays is not in itself alone, but in that play in the order in which it was written, in its relation to all of Shakespeare's other plays, earlier and later: we must know all of Shakespeare's work in order to know any of it. No other dramatist of the time approaches anywhere near to this perfection of pattern. . . . It seems to me to correspond to some law of nature that the work of a man like Shakespeare, whose development in the course of his career was so amazing, that it should reach, as in Hamlet, the point at which it can touch the imagination and feeling of the maximum number of people to the greatest possible depth and that, thereafter, like a comet which has approached the earth and then continued away on its course, he should gradually recede from view until he tends to disappear into his private mystery.
Laboratory for Astrophysics and Space Research, University of Chicago. His communication was presented at the 209th Annual Meeting of the Academy, held at the House of the Academy on May 10, 1989.
In contrast, let me quote Einstein and Dirac, which bears on what I have said. First, Einstein: A person who has not made his great contribution to science before the age of thirty will never do so. And here is a quatrain attributed to Dirac: Age is, of course, a fever chill that every physicist must fear. He's better dead than living still, when once he's past this thirtieth year! Juxtaposed with what I have quoted with respect to Shakespeare and Beethoven, the contrast in attitude is indeed striking. Subrahmanyan Chandrasekhar is Morton D. Hull Distinguished Service Professor Emeritus at the 28
29
1278 Reprinted from Current Science, February 20, 1985, Vol.54, No.4, pp 161-169 THE PURSUIT OF SCIENCE: ITS MOTIVATIONS* S. CHANDRASEKHAR The University of Chicago, Chicago, Illinois 60637, USA. am grateful to President Ramaseshan for assigning to me the privilege of addressing you on this occasion celebrating the Golden Jubilee of this Academy. But for the tragedy which engulfed the country on October 31 st of last year, this celebration should have taken place on November 7th, the 96th birthday of the Academy's illustrious founder and its President for its first 35 years. That this occasion is a moving one for my wife—a former student of Professor Raman—and myself must be obvious to you; and I shall not expand on it. That the occasion is also precious to me, on scientific grounds, derives not only from my having been a fellow of this Academy for all of its fifty years, but equally from the purposes to which this Academy is dedicated. Unlike other national academies, it has not sought, nor has it strived, to influence the public policies of the national government; nor has it followed the practice of awarding innumerable prizes, lectureships, and medals. The purpose of this Academy, as enunciated by its founder, is to promote the pursuit of science in this country by providing journals and periodicals for the publication of scientific papers of the highest standards, and avoiding the necessity to seek foreign avenues.
1
During Professor Raman's lifetime, the Proceedings of the Academy, both series A and B, were published without lapse and without delay. Since 1970, when the affairs of the Academy passed on to his successors, this prime objective has so remained. By the efforts, principally of President Ramaseshan, the number of journals which are presently sponsored by the Academy has increased almost ten-fold. And all these journals do indeed maintain the highest standards by strict refereeing. I cannot wish for the future of the Academy any better than to * Inaugural address at the Golden Tubilee Meeting of the Indian Academy of Sciences, Bangalore, 6 February 1985.
express confidently the hope that the steadily increasing standards of the past ten years will be sustained during the next fifty years. I I now turn with some apprehension to the subject of my address, 'The Pursuit of Science: Its Motivations'. This is a difficult subject if one is to avoid the common and the banal. The difficulty derives in large measure from the variety and the range of the motives of the individual scientists; they are varied and they are diverse: they are as varied as the tastes, the temperaments, and the attitudes of the scientists themselves. Besides, the motivations are subject to substantial changes during the lifetimes of the scientists. Indeed, it is difficult to discern a common denominator. What is it then, one can usefully say? I shall skirt the problem and restrict myself to reflections—perhaps, disjointed—on the lives and the accomplishments of some of the great scientists of the past. And I shall, whenever possible, base my remarks on what they have themselves said or written. But reflecting on the motives and the attitudes of great men is beset with grave semantic difficulties of communication: the words and phrases that language allows in these contexts may suggest criticism or judgement. Therefore, let me make it clear from the outset, that my remarks, at no place, are to be construed as having overtones or undertones of criticism or judgement. Indeed, I have no rights to such criticism or judgement in the contexts I shall be speaking. I should also make it clear that during the course of my reflections that I shall present, thoughts derived from my own personal experience have been entirely absent. Since it may not be possible for me to emphasize these points at every stage, I shall begin with a quotation and a narration. The quotation is from the concluding pages of Turgenev's On the Eve in which there is a
1279
statement by that silent but indomitable character Insarov: We are speaking of other people; why bring in yourself? My narration is of a conversation between Majorana and Fermi in the middle twenties when both of them were also in their middle twenties. (The conversation was reported to me by one who was present on the occasion.) MAJORANA: There are scientists who 'happen' only once in every 500 years, like Archimedes or Newton. And there are scientists who happen only once or twice in a century, like Einstein or Bohr. FERMI: But where do I come in, Majorana? MAJORANA: Be reasonable, Enrico! I am not talking about you or me. I am talking about Einstein and Bohr. Since I shall be talking principally about scientists in the class of Einstein, Bohr, and Fermi, I should indeed be 'reasonable'. One final reservation. The circumference of my comprehension does not extend beyond a very limited circle of the physical sciences. This is, of course, a most serious limitation when I am addressing myself to so general a theme; but I must abide by my limitation. II For a discussion of the motivations which impel one to pursue the goals of science, no example is better than that of Johannes Kepler. Kepler's uniqueness derives from the position he singly occupies at the great crossroads where science shed its enveloping dogmas and the pathway was prepared for Newton. Kepler, in his inquiries, asked questions that none before him, including Copernicus, had asked. Kepler's laws differ qualitatively from earlier assumptions about planetary orbits: the assertion that planetary orbits 'are ellipses' in no way resembles the kind of improvements that his predecessors had
sought. In his analysis of the motions of the planets, Kepler was not preoccupied with geometrical questions; he asked, instead, questions such as 'what is the origin of planetary motions'? 'If the sun is at the centre of the solar system, as it is in the Copernican scheme, should not that fact be discernible in the motions and in the orbits of the planets themselves?'. These are questions in physics; not in some preconceived geometrical framework. While Kepler's approach to the problem of planetary motions was radically different from that of anyone before him, his work is preeminent for the manner in which he extracted general laws from a careful examination of the observations. His examination was long and it was arduous: it took him twenty and more years of constant and persistent effort; but he never lost sight of his goal. For him, it was a search for the holy grail in a very literal sense. Before I describe the manner of Kepler's search, I should like to say that I am in no sense a scholar of medieval astronomy. My knowledge of Kepler is in fact mostly derived from Arthur Koestler's The Sleep-walkers: a History of Man's Changing Vision of the Universe, some collateral reading, and some discussions with scholars who know very much more. Koestler's sensitive account of Kepler, his life and his achievements, includes numerous quotations from Kepler's own writings. My remarks are largely based on these quotations. From the outset Kepler realized that a careful study of the orbit of Mars will provide the key to planetary motions because its orbit departs from a circle the most; and it had defeated Copernicus; and further that an analysis of the accurate observations of Tycho Brahe was an essential prerequisite. As Kepler wrote: Let all keep silence and hark to Tycho who has devoted thirty-five years to his observations. . . . For Tycho alone do I wait; he shall explain to me the order and arrangement of the orbits. . . . Tycho possesses the best observations, and thus so-to-speak the material for the building of the new edifice. . . .
1280
. . . I believe it was an act of Divine Providence that I arrived just at the time when Longomontanus was occupied with Mars. For Mars alone enables us to penetrate the secrets of astronomy which otherwise would remain forever hidden from us. . . . Indeed, Kepler went to extraordinary lengths to acquire the observations of Tycho which he so badly needed. It is not an exaggeration to say that he commited larceny; for, as he confessed: I confess that when Tycho died, I quickly took advantage of the absence, or lack of circumspection, of the heirs, by taking the observations under my care, or perhaps usurping them. . . . and as he explained: The cause of this quarrel lies in the suspicious nature and bad manners of the Brahe family, but on the other hand also in my own passionate and mocking character. It must be admitted that Tengnagel had important reasons for suspecting me. I was in possession of the observations and refused to hand them over to the heirs. . . . With Tycho's observations thus acquired, the question which Kepler constantly asked himself was: If the sun is indeed the origin and the source of planetary motions, then, how does this fact manifest itself in the motions of the planets themselves? Noticing that Mars moved a little faster when it is nearest the sun than when it is farthest, and 'remembering Archimedes', he determined the area described by the radius vector joining the sun to the instantaneous position of Mars, as we follow it in its orbit. As Kepler wrote: Since I was aware that there exists an infinite number of points on the orbit and accordingly an infinite number of distances [from the sun] the idea occurred to me that the sum of these distances is contained in the area of the orbit. For I remembered that in the same manner Archimedes too divided the area of a circle into an infinite number of triangles.
This was the way Kepler discovered in July 1603 his law of areas. This is the second of his three great laws in Newton's enumeration that has been adopted ever since. The establishment of this result took Kepler some five years; for, already prior to the publication of his Mysterium Cosmographicum in 1596, Kepler had sought for such a law in connection with his association of thefiveregular solids with the existence of the six planets known in his time. The law of areas determined the variation of the speed along its orbit; but it did not determine the shape of the orbit. A year before he had arrived at his final statement of the law of areas he had in fact discarded circular orbits for the planets: and in October of 1602 he had written: The conclusion is quite simply that the planet's path is not a circle—it curves inward on both sides and outward again at opposite ends. Such a curve is called an oval. The orbit is not a circle, but an oval figure. Even after having concluded that the orbit of Mars is an 'oval', it took him an additional three years to establish that the orbit was in fact an ellipse. And when that was established he wrote: Why should I mince my words? The truth of Nature, which I had rejected and chased away, returned by stealth through the back door, disguising itself to be accepted. That is to say, I laid [the original equation] aside, and fell back on ellipses, believing that this was a quite different hypothesis, whereas the two, as I shall prove in the next chapter, are one and the same. . . . I thought and searched, until I went nearly mad, for a reason, why the planet preferred an elliptical orbit [to mine]. . . . Ah, what a foolish bird I have been! At long last in 1608 his monumental Astronomia Nova was published. As Koestler has written: It was a beautifully printed volume in folio, of which only a few copies survive. The Emperor [Rudolph] claimed the whole
1281
edition as his property and forbade Kepler to sell or give away any copy of it 'without our foreknowledge and consent'. But since his salary was in arrears, Kepler felt at liberty to do as he liked, and sold the whole edition to the printers. Thus the story of the New Astronomy begins and ends with acts of larceny, committed ad majorem Dei gloriam. Ten more years were to elapse before Kepler discovered his third law: that the squares of the periods of revoution of any two planets is in the ratio of the cubes of their mean distances from the sun. The law is stated in his Harmonice Mundi completed in 1618. Here is how Kepler describes his discovery: On 8 March of this present year 1618, if precise dates are wanted, [the solution] turned up in my head. But I had an unlucky hand and when I tested it by computations I rejected it as false. In the end it came back again to me on 15 May, and in a new attack conquered the darkness of my mind; it agreed so perfectly with the data which my seventeen years of labour on Tycho's observations had yielded, that I thought at first I was dreaming. Thus ended Kepler's long and arduous search for his holy grail. In his first book, Mysterium Cosmographicum, Kepler exclaimed: Oh! that we could live to see the day when both sets of figures agree with each other. Twenty-two years later, he added the following footnote to this exclamation in a reprint edition of Mysterium Cosmographicum after he had discovered his third law and his poignant cry had been answered: We have lived to see this day after 22 years and rejoice in it, at least I did; I trust that Maestlin and many other men will share in my joy! Ill In his novel, The Redemption of Tycho Brahe, Max Brod, the Czech writer who is known for his
publishing, posthumously, the great works of Franz Kafka, portrays and contrasts the characters of Tycho Brahe and Kepler. While Brod's novel is historically, grossly inaccurate, yet the following imagined perception of Kepler by Tycho is an artist's idealizaion of what a scientist like Kepler might have been: Kepler now inspired him [Tycho] with a feeling of awe. The tranquility with which he applied himself to his labours and entirely ignored the warblings of flatterers was to Tycho almost superhuman. There was something incomprehensible in its absence of emotion, like a breath from a distant region of ice. . . . Is the tranquility and the absence of emotion which Brod attributes to his imagined Kepler, ever attained by a practising scientist? May I digress a little at this point to say that Max Brod, when he wrote his novel, The Redemption of Tycho, was one of a small group in Prague that included Einstein and Franz Kafka. It has been said that Brod's portrayal of Kepler was influenced by his association with Einstein. Thus Walter Nernst is reported to have said to Einstein, "You are this man Kepler."
IV As I have stated, the most remarkable aspect of Kepler's pursuit of science is the constancy with which he applied himself to his chosen quest. To use a phrase of Shelley's, his 'was a character superior in singleness'. But does the example of Kepler provide any assurance of success for a similar constancy in others? I shall consider two examples. First, the example of Michelson. His main preoccupation throughout his life was to measure the velocity of light with increasing precision. His interest came about almost by accident, when the Commander of the Naval Academy asked him (then an instructor at the Academy) to prepare some lecture demonstrations to illustrate Foucault's refinement of Cornu's determination of the velocity of light. That was in 1878; and it led to Michelson's first determination of the
1282
velocity of light in 1880. On the 7th of May 1931, two days before he died and fifty years later, he dictated the opening sentences of a paper, that was posthumously published and which gave the results of his last measurement. Michelson's efforts resulted in an improvement in our knowledge ofthe velocity of light from 1 part in 3000 to 1 part in 30,000, i.e. by a factor 10. But by 1973 the accuracy had been improved to 1 part in 10 10 , a measurement that made obsolete, beforehand, all future measurements. Were Michelson's efforts over 50 years then in vain? Leaving that question aside, one must record that during his long career, Michelson made great discoveries derived from his delight in 'light waves and their uses'. Thus, his development of interferometry leading to the first direct determination of the diameter of a star is wonderous. And who does not know the Michelson-Morley experiment which, through Einstein's formulation of the special and the general theory of relativity, changed and changed irrevocably our understanding ofthe nature of space and time? But it is a curious fact that Michelson himself was never happy with the outcome of his experiment. Indeed, it is recorded that when Einstein visited Michelson in April 1931, Mrs. Michelson felt it necessary to warn Einstein "Please don't get him started on the ether." A second example is Eddington who devoted the last 16 years of his life to developing his 'fundamental theory'. Of this prodigious effort he said, a year before he died: "At no time during the past 16 years have I felt any doubt about the correctness of my theory." Yet, his efforts have left no trace on subsequent developments. Is it wise then to pursue science with a single objective and with a singleness of purpose?
V While Kepler provides the supreme example of sustained scientific effort leading to great and fundamental discoveries, there are instances in which great thoughts have seemingly occurred
spontaneously. Thus, Dirac has written that his work on Poisson brackets and on his relativistic wave equation ofthe electron were consequences of ideas . . . which had just come out of the blue. I could not very well say just how it had occurred to me. And I felt that work of this kind was a rather 'undeserved success'. Dirac's recollection, that his ideas underlying his work on Poisson brackets and his relativistic wave equation ofthe electron came to him 'out of the blue', is an example of what is apparently not a unique phenomenon: Those who have made great discoveries seem to remember and cherish the occasions on which they made them. Thus, Einstein has recorded that When in 1907 I was working on a comprehensive paper on the special theory of relativity . . . there occurred to me the happiest thought of my life . . . that 'for an observer falling freely from the roof of a house there exists—at least in his immediate surroundings—no gravitational field: This 'happy thought' was, of course, later enshrined in his principle of equivalence that is at the base of his general theory of relativity. A recollection in a similar vein is that of Fermi. I had once the occasion to ask Fermi, in the context of Hadamard's perceptive 'Essay on the Psychology of Invention in the Mathematical Field' what the psychology of invention in the realm of physics might be. Fermi responded by narrating the occasion of his discovery of the effect of slow neutrons on induced radioactivity. This is what he said: I will tell you how I came to make the discovery which I suppose is the most important one I have made. We were working very hard on the neutron-induced radioactivity and the results we were obtaining made no sense. One day, as I came to the laboratory, it occurred to me that I should examine the effect of placing a piece of lead before the incident neutrons. Instead of my usual custom, I took great
6 pains to have the piece of lead precisely machined. I was clearly dissatisfied with something; I tried every excuse to postpone putting the piece of lead in its place. When finally, with some reluctance, I was going to put it in place, I said to myself: 'No, I do not want this piece of lead here; what I want is a piece of paraffin.' It was just like that with no advance warning, no conscious prior reasoning. I immediately took some odd piece of paraffin and placed it where the piece of lead was to have been. Perhaps the most moving statement in this general context is that of Heisenberg relating the moment when the laws of quantum mechanics came to a sharp focus in his mind. He has written, One evening I reached the point where I was ready to determine the individual terms in the energy table, or, as we put it today, in the energy matrix, by what would now be considered an extremely clumsy series of calculations. When the first terms seemed to accord with the energy principle, I became rather excited, and I began to make countless arithmetical errors. As a result, it was almost three o'clock in the morning before the final result of my computations lay before me. The energy principle had held for all terms, and I could no longer doubt the mathematical consistency and coherence of the kind of quantum mechanics to which my calculations pointed. At first, I was deeply alarmed. I had the feeling that, through the surface of atomic phenomena, I was looking at a strangely beautiful interior, and felt almost giddy at the thought that I now had to probe this wealth of mathematical structures, nature had so generously spread out before me. I was far too excited to sleep, and so, as a new day dawned, I made for the southern tip of the island, where I had been longing to climb a rock jutting out into the sea. I now did so without too much trouble, and waited for the sun to rise. There is no difficulty for any of us to share in Heisenberg's exhilaration of that supreme
moment: we all know of the difficulties and paradoxes that beset the 'old' Bohr-Sommerfeld quantum-theory of the time; and we also know of Heisenberg's long puzzlement over these difficulties and paradoxes with Sommerfeld, Bohr, and Pauli. And he had already published at that time his paper with Kramers on the dispersion theory—a theory which in many ways was the precursor to the developments that were to follow. But what is our reaction to Heisenberg's account of his ideas on the theory of elementary particles that he developed some thirty years later, after his tragic experiences during the war and his disappointments and frustrations of the post-war years? Mrs. Heisenberg, in her book on her husband, has written, One moonlight night we walked all over the Hainberg Mountain, and he was completely enthralled by the visions he had, trying to explain his newest discovery to me. He talked about the miracle of symmetry as the original archetype of creation, about harmony, about the beauty of simplicity, and its inner truth. And she quotes from one of Heisenberg's letters to her sister at this time: In fact, the last few weeks were full of excitement for me. And perhaps I can best illustrate what I have experienced through the analogy that I have attempted an as yet unknown ascent to the fundamental peak of atomic theory, with great efforts during the past five years. And now, with the peak directly ahead of me, the whole terrain of interrelationships in atomic theory is suddenly and clearly spread out before my eyes. That these interrelationships display, in all their mathematical abstraction, an incredible degree of simplicity, is a gift we can only accept humbly. Not even Plato could have believed them to be so beautiful. For these interrelationships cannot be invented; they have been there since the creation of the world. You will notice the remarkable similarity in the language and in the phraseology of this descrip-
7 tion with the description of his discovery of the basic rules of quantum mechanics some thirty years earlier. But do we share in his second vision in the same way? In the earlier case, his ideas won immediate acceptance. In contrast, his ideas on particle physics were rejected and repudiated even by his long time critic and friend Pauli. But it is moving to read what Mrs. Heisenberg writes towards the end of her biography. With smiling certainty, he once said to me: 'I was lucky enough to look over the good Lord's shoulder while He was at work.' That was enough for him, more than enough! It gave him great joy, and the strength to meet the hostilities and misunderstandings he was subjected to in the world time and again with equanimity, and not be led astray. VI A different aspect ofthe effect a great discovery can have on its author is provided by the autobiography entitled The Traveler by Hideki Yukawa. The book was written when Yukawa was past fifty. One would normally have expected that an autobiography entitled The Traveler by one whose life, at least as seen from the outside, had been rich and fruitful, would be an account of his entire life. But Yukawa's account of his 'travels' ends with the publication of his 1934 paper describing his great discovery with the sombre note I do not want to write beyond this point, because those days when I studied relentlessly are nostalgic to me; and on the other hand, I am sad when I think how I have become increasingly preoccupied with matters other than study. VII While all of us can share in the joy of the discoveries ofthe great men ofscience, we may be puzzled by what those many, very many, less perceptive and less fortunate, are to cherish and remember. Are they, like Vladimir and Estragon,
destined to wait for Godot as in Samuel Beckett's play; or, are they to console themselves with Milton's thought 'they also serve who stand and wait'? VIII I now turn to the role which approbation and approval play in one's pursuit of science. The example of Newton 'voyaging through strange seas of thought alone' is not one that any of us can follow. I have referred to Eddington's lonely efforts in pursuing his fundamental theory. In spite of his expressed confidence in the correctness of his theory, Eddington must have been deeply frustrated by the neglect of his work by his contemporaries. This frustration is evident in his plaintive letter to Dingle written a few months before he died: I am continually trying to find out why people find the procedure obscure. But I would point out that even Einstein was considered obscure, and hundreds of people have thought it necessary to explain him. I cannot seriously believe that I ever attain the obscurity that Dirac does. But in the case of Einstein and Dirac people have thought it worthwhile to penetrate the obscurity. I believe they will understand me all right when they realize they have got to do so—and when it becomes the fashion 'to explain Eddington.' The lack of approval by one's contemporaries can have tragic consequences when they are expressed in the form of sharp and violent criticisms. Thus, Ludwig Boltzmann, greatly depressed by the violence of the attacks directed against his ideas by Ostwald and Mach, committed suicide 'as a martyr to his ideas', as his grandson Flamm has written. And George Cantor, the originator of the modern theory of sets of points and ofthe orders of infinity, lost his mind because of the hatred and the animosity against him and his ideas by his teacher Leopold Kronecker, and was confined to a mental hospital during the last many years of his life.
1285
8
IX A case very different from the ones I have considered so far is that of Rutherford. Consider his record. In 1897 he analyzed radioactive radiations into three types: aparticles, j3-rays and y-rays, a nomenclature that has survived to this day. In 1902 he formulated the laws of radioactive disintegration: the first time a physical law was formulated in terms of probability and not certainty: a forerunner of the probability interpretation of quantum mechanics that was to become universal some 25 years later. In 1905-1907 he formulated, together with Soddy, the laws of radioactive displacement and identified the a-particle as the nucleus of the helium atom; and, together with Boltwood, initiated the determination of the ages of rocks and minerals by their radioactivity. In 1909-1910, there were the experiments of Geiger and Marsden, the discovery of the large angle scattering of a-rays and Rutherford's formulation of his law of scattering and the nuclear model of the atom. Then in 1917 he effected the first laboratory transformation of atoms: that of nitrogen-14 into oxygen-17 and a proton by a-ray bombardment. In the twenties he was associated with the clarification of the relationship between the a-ray and the y-ray spectra. And 1932—the annus mirabilis as R. H. Fowler called it—saw the discovery of the artificial disintegration of Li into two a-particles by Cockroft and Walton, of positrons in cosmic-ray showers by Blackett, and the neutron by Chadwick—all of them in Rutherford's Cavendish. In the following year Rutherford, together with Oliphant, himself discovered hydrogen-3 and helium-3. Altogether, then, an accomplishment unparalleled in this century. Rutherford's attitude to his own discoveries is illustrated by his response to a remark of one who was present at the moment of one of his great discoveries: 'Rutherford, you are always on the crest of the wave.' To which Rutherford responded: T made the wave, didn't I?' Somehow from Rutherford's vantage point everything he said seems right, even including his remark, T do not let my boys waste their time' when he was
asked if he encouraged his students to study relativity! Rutherford was a happy warrior if ever there was one. X So far, I have tried to illustrate some facets of the pursuit of science by drawing on incidents in the lives of some great men of science. I shall turn now to some more general matters. I shall start with an example. It has been reported that when Michelson was asked towards the end of his life, why he had devoted such a large fraction of his time to the measurement of the velocity of light, he replied 'it was so much fun'. There is no denying that 'fun' does play a role in the pursuit of science. But the word 'fun' suggests a lack of seriousness. Indeed, the Oxford Dictionary gives to 'fun' the meaning 'drollery'. We can be certain that Michelson did not have that meaning in his mind when he described his life's main interest as 'fun.' If not, what precisely is the meaning we are to attach to 'fun' in the context in which Michelson used it? More generally, what is the role of pleasure and enjoyment? While 'pleasure' and 'enjoyment' are often used to characterize one's efforts in science, failures, frustrations, and disappointments are equally, if not more, the common ingredients of scientific experience. Overcoming difficulties, undoubtedly, contributes to one's final enjoyment of success. Is failure, then, a purely negative aspect of the pursuit of science? A remark of Dirac's describing the rapid development of physics following the founding of the principles of quantum mechanics in the middle and the late twenties is apposite in this connection. It was a good description to say that it was a game, a very interesting game one could play. Whenever one solved one of the little problems, one could write a paper about it. It was very easy in those days for any second-rate physicist to do first-rate work. There has not been such a glorious time since then.
1286
9 Consider in the context of these remarks, J. J. Thomson's assessment of Lord Rayleigh in his memorial address given in Westminister Abbey: There are some great men of science whose charm consists in having said the first word on a subject, in having introduced some new idea which has proved fruitful; there are others whose charm consists perhaps in having said the last word on the subject, and who have reduced the subject to logical consistency and clearness. I think by temperament Lord Rayleigh belonged to the second group. This assessment of Rayleigh by J. J. Thomson has sometimes been described as double-edged. But could one not conclude, instead, that Rayleigh by temperament chose to address himself to difficult problems and was not content to play the kind of games that Dirac describes in his characterization of the 'glorious time' in physics as a time 'when second-rate physicists could do first-rate work'? This last question concerning Rayleigh's temperament raises the further question: after a scientist has reached maturity, what are his criteria for his continued pursuit of science? To what extent are they personal? And to what extent are aesthetic criteria like the perception of order and pattern, form and substance, relevant? Are such personal criteria exclusive? Has a sense of obligation a role? I do not mean obligation with the common meaning of obligation to one's students, one's colleagues, and one's community. I mean, rather, obligation to science itself. And what, indeed, is the content of obligation in the pursuit of science for science? Let me finally turn to a different aspect. G. H.
Hardy concludes his A Mathematician's Apology with the following illuminating statement: The case for my life, then, or for that of anyone else who has been a mathematician in the same sense in which I have been one, is this: that I have added something to knowledge, and helped others to add more; and that these somethings have a value which differs in degree only, and not in kind, from that of the creations of the great mathematicians, or of any of the other artists, great or small, who have left some kind of memorial behind them. Hardy's statement is made relative to mathematicians; but it is equally applicable to all scientists. I want to draw your attention particularly to Hardy's reference to one's wanting to leave behind some kind of a memorial i.e., something that posterity may judge. To what extent, then, is the judgement of posterity (which one can never know) a conscious motivation in the pursuit of science? XI The pursuit of science has often been compared to the scaling of mountains, high and not so high. But who amongst us can hope, even in imagination, to scale the Everest and reach its summit when the sky is blue and the air is still; and in the stillness of the air survey the entire Himalayan range in the dazzling white of the snow stretching to infinity. None of us can hope for a comparable vision of nature and of the universe around us. But there is nothing mean or lowly in standing in the valley below and awaiting the sun to rise over the Kunchenjunga.
Festi-Val - Festschrift for Val Telegdi K. Winter, editor © Elsevier Science Publishers B. V., 1988
49
A COMMENTARY ON DIRAC'S VIEWS ON "THE EXCELLENCE OF GENERAL RELATIVITY" S. CHANDRASEKHAR The University of Chicago, Chicago, Illinois 60637, USA
In a lecture given in 1979, Dirac has expressed certain views on the aesthetic basis that underlies Einstein's formulation of the general theory of relativity. A critical assessment of his views is attempted.
A Cat May Look at a King Lewis Carroll, Alice in Wonderland
The present essay was written a year ago with the sole purpose of clarifying to myself my own views on general relativity and with no intention of publication. If nevertheless I am now emboldened to submit it as my contribution to the Festschrift dedicated to Valentine Telegdi, it is because I have learnt from my long and fruitful association with him that one must probe one's own understanding of a subject without the inhibition of higher authority. Indeed, as one well-known critic once suggested, there may be no better way to sharpen one's musical discernment than, for example, to try to detect, if one can, a misplaced note or a wrong phrasing in Rachmaninoff s playing of a Beethoven sonata. This essay is written in that spirit with the hope that Telegdi will approve the motivation of its writing even if he does not countenance its content. It has often been said that the two great revolutions in physical thought during this century derive from the almost simultaneous founding of the quantum theory and the (classical) general theory of relativity. In contrast to the quantum theory which has grown and flowered in abundance, the general theory of relativity has remained static: all that has been derived in general relativity are strict consequences of Einstein's equation as he wrote it in 1915 without change or alteration. Thus, while there is a general consensus that the general theory of relativity is a "beautiful theory", there is no consensus as to wherein its beauty lies or what the aesthetic elements in it are. To illustrate
50
S.
Chandrasekhar
that these differences are not matters of taste only, I shall detail my differences with the views that Dirac has expressed clearly and precisely in a lecture entitled "The Excellence of General Relativity" [1] that he gave on the occasion of Einstein's centennial. I shall quote from Dirac's lecture and follow the quotations with my commentary. [I] I have enumerated the successes of the Einstein theory of gravitation. It is a long list, quite impressive. In every case, the Einstein theory is confirmed, with greater or less accuracy depending on the precision with which the observations can be made and uncertainties that they involve.
It does not seem to me that the successes of Einstein's theory are either long or impressive. It is true that his prediction of the different rates of clocks in locations of differing gravity, his prediction of the deflection of light when traversing a gravitational field and resulting time delay, his prediction regarding the precession of the perihelion of Mercury, and finally, the slowing down of a binary star in an eccentric orbit by virtue of the emission of gravitational radiation, have been confirmed quantitatively. But all these relate to the departures from Newtonian theory by a few parts in a million; and of no more than three or four parameters in a post-Newtonian expansion of Einstein's field equations. And so far, no predictions of general relativity in the limit of strong gravitational fields have received any confirmation; nor are they likely in the foreseeable future. Should one not argue that a theory which generalizes a theory as well confirmed in its domain as the Newtonian theory of gravity should refer to predictions which relate to the major aspects of the theory rather than to small first-order departures from the theory it replaces? Would the status of Dirac's theory of the electron, for example, be what it is today if its only success consisted in accounting for Paschen's 1916 measurements of the fine structure of ionized helium? The real confirmation of Dirac's theory was provided by the exact departures provided by the Klein-Nishina formula replacing Thomson's formula for electron scattering; and more particularly in the confirmation of the predictions of the theory relating to the creation of electron-positron pairs in cosmic-ray showers and of anti-matter. Or, would one's faith in Maxwell's equations of the electromagnetic field be as universal as it is without Hertz's experiments on the existence and propagation of electromagnetic waves with the velocity of light and without Poincare's proof of their invariance to Lorentz transformations? In the same way, a real confirmation of the general theory of relativity will be forthcoming only if a prediction characteristic of that theory and only of that theory is confirmed. The occurrence of black holes in nature as one of the final equilibrium states of massive stars in the natural course of their evolution is not a confirmation of a prediction of general relativity in any real sense. The notion that light cannot escape from a sufficiently strong gravitational field is an obvious inference not
Commentary on Dirac's "The excellence of general relativity" based on any exact prediction of the theory: it depends only on the verified fact that light is affected by gravity. On the other hand, since the general theory of relativity provides an exact description of the space-time around black holes, only a quantitative confirmation of the metric of the space-time around black holes can be considered as 'establishing' the theory. It is well known that the Kerr solution with two parameters provides the unique solution for stationary black holes that occur in nature. But a confirmation of the metric of the Kerr space-time (or, some aspect of it) cannot even be contemplated in the foreseeable future. [II] Suppose a discrepancy appeared, well confirmed and substantiated, between the theory and observations Should one then consider the theory wrong?... I would say that the answer to the last question is emphatically NO. The Einstein theory of gravitation has a character of excellence of its own.
Again, it seems to me that Dirac's question is not well posed and his answer unsatisfactory. First, with regard to the question whether the general theory of relativity is immutable. Clearly, no physical theory is immutable. The general theory of relativity cannot be an exception: every valid physical theory is circumscribed by limitations inherent to it. Consider the Newtonian theory of gravitation, for example. In Newton's time, light was not supposed to be affected by gravity; and it was not realized that the velocity of light provides an upper hmit to the velocity of propagation of signals; and that, at any point in space-time, there is an accessible future and an experienced past. Once one realizes this fact, then one realizes at the same time that the Newtonian theory is limited by the requirement that, for its validity, bodies should be moving with velocities small compared to that of light; and that one should expect 'discrepancies' when the finiteness of the velocity of light is taken into account. This fact, for example, was realized by Poincare in 1904 before Einstein's formulation of the special theory of relativity in 1905. For example, Pais, in his biography of Einstein, has written [2], Poincare reasons in a more general and abstract way that all forces should transform in the same way under Lorentz transformations. He concludes that therefore Newton's laws need modification and that there should exist gravitational waves which propagate with the velocity of light! Finally, he points out that the resulting corrections of Newton's law must be (v2/c2) and that the precision of astronomical data does not seem to rule out effects of this order.
As another example, consider the quantum theory. Maxwell's theory predicts that accelerated electrons must radiate. Indeed, synchrotron radiation arises precisely on this account. However, this prediction is contrary to the existence of stationary states for electrons in atoms. We conclude that classical electrodynamics and the classical equations of motion of charged particles is
51
52
S.
Chandrasekhar
restricted by the requirement that the relevant actions are large compared to Planck's quantum of action. Therefore we expect (and indeed, find) that departures from the Newtonian theory and classical electrodynamics must occur when the finiteness of the velocity of light and of the quantum of action become significant. If Einstein's theory is assumed to be valid in its domain (and no other assumption would be justified in this context), then its inherent limit is again provided by the quantum of action; and we should expect 'discrepancies' when effects of quantum theory become relevant. Barring new fundamental knowledge in physics, one could assume that departures from Einstein's theory will arise when intervals of distance and intervals of time become comparable to the Planck scales. It seems to me idle to ask whether we shall suddenly find some gross departure from Einstein's prediction - for example, when the proposed experiments of Fairbank relating to the Lense-Thirring effect are completed. Fairbank's experiments are more likely to provide a testimony to the virtuoso character of his experiments! And as to the excellence of the theory, what does that excellence consist of? A theory is not excellent by repeating ad nauseam that our confidence in the theory arises from [III] ... the essential beauty of the mathematical description of Nature which inspired Einstein in his quest of a theory of gravitation.
Our confidence in the theory arises rather from its internal consistency (such as with the requirement of causality and the positivity of the energy) and above all from its freedom from contradiction with parts of physics (such as quantum theory and thermodynamics) not contemplated in the formulation of the theory. (These ideas are pursued in greater detail in the author's Karl Schwarzschild Lecture [3].) [IV] When Einstein was working on building up this theory of gravitation he was not trying to account for some results of observations. Far from it. His entire procedure was to search for a beautiful theory, a theory of a type that Nature would choose Somehow he got the idea of connecting gravitation with the curvature of space. He was able to develop a mathematical scheme incorporating this idea. He was guided only by consideration of the beauty of these equations.
It seems to me that the foregoing statement by Dirac expresses a view which has no historical basis. It is wrong to say that Einstein, in building up his theory, was not "trying to account for some results of observations". And that he was guided solely by considerations relating to the "beauty of the equations". Actually, of course, he was trying to account for what is perhaps the most well confirmed and the most accurately established experimental fact: the equality of the inertial and the gravitational mass. Einstein has stated this over and over again. For example, in the transcript of a lecture, How I
Commentary on Dirac s The excellence of general relativity
53
Created the Theory of Relativity [4], which Einstein gave in Kyoto on the 14th of December, 1922, he said, ... I came to realize that all the natural laws except the law of gravity could be discussed within the framework of the special theory of relativity. I wanted to find out the reason for this, but I could not attain this goal easily. The most unsatisfactory point was the following: Although the relationship between inertia and energy was explicitly given by the special theory of relativity, the relationship between inertia and weight, or the energy of the gravitational field, was not clearly elucidated. I felt that this problem could not be resolved within the framework of the special theory of relativity. The breakthrough came suddenly one day. I was sitting on a chair in my patent office in Bern. Suddenly a thought struck me: If a man falls freely, he would not feel his weight. I was taken aback. This simple thought experiment made a deep impression on me. This led me to the theory of gravity. I continued my thought: A falling man is accelerated. Then what he feels and judges is happening in the accelerated frame of reference. I decided to extend the theory of relativity to the reference frame with acceleration. I felt that in doing so I could solve the problem of gravity at the same time. A falling man does not feel his weight because in his reference frame there is a new gravitational field which cancels the gravitational field due to the Earth. In the accelerated frame of reference, we need a new gravitational field.
And again, in an address given in London in the early twenties [5], The general theory of relativity owes its existence in the first place to the empirical fact of the numerical equality of the inertial and gravitational mass of bodies, for which fundamental fact classical mechanics provided no interpretation. Such an interpretation is arrived at by an extension of the principle of relativity to co-ordinate systems accelerated relatively to one another. The introduction of co-ordinate systems accelerated relatively to inertial systems involves the appearance of gravitational fields relative to the latter. As a result of this, the general theory of relativity, which is based on the equality of inertia and weight, provides a theory of the gravitational field. The introduction of co-ordinate systems accelerated relatively to each other as equally legitimate systems, such as they appear conditioned by the identity of inertia and weight, leads, in conjunction with the results of the special theory of relativity, to the conclusion that the laws governing the occupation of space and solid bodies, when gravitational fields are present, do not correspond to the laws of Euclidean geometry That generalization of the metric, which had already been accomphshed in the sphere of pure mathematics through the researches of Gauss and Riemann, is essentially based on the fact that the metric of the special theory of relativity can still claim validity for small areas in the general case as well.
Or again [6], ...A little reflection will show that the law of the equality of the inert and the gravitational mass is equivalent to the assertion that the acceleration imparted to a body by a gravitational field is independent of the nature of the body.
I may paraphrase what Einstein has said on this and on various other occasions. Consider the Newtonian equations of motion in a general curvilinear coordinate system, qa (a = 1, 2, 3), with the three-dimensional metric,
ds^h^dq'dq',
(1)
S. Chandrasekhar
54
of a mass point moving in a gravitational field described by a potential They take the form [7] m
inerth afiqp = -minen
TaySqyqs
- wgravFa,
V(qa).
(2)
where mmeTl and m grav denote, respectively, the inertial and the gravitational mass and T s the Christoffel symbol, ra,y8=
2{hayr$ + haS
y
— hyS a).
(3)
We observe that there are two terms that contribute to the acceleration: a term that depends on the coordinate system chosen and a term that derives from the gravitational field. They occur, respectively, with the factors w inert and m grav representing the inertial and the gravitational masses and since the two masses are known, empirically, to be equal, the distinction between the two contributions appears unnatural: the disappearance of the mass factor in eq. (2) is, as Weyl has described, an element of magic in the Newtonian theory. Einstein argued: why not abolish the distinction altogether? And he concluded that "all acceleration is metrical in origin". Einstein's conclusion that in the context of gravity all accelerations are metrical in origin is as staggering in its own way as Rutherford's conclusion when Geiger and Marsden first showed him the results of their experiments on the large-angle scattering of a-rays: "It was as though you had fired a fifteen-inch shell at a piece of tissue paper and it had bounced back and hit you". Of course, the transition from the statement that "all acceleration is metrical in origin" to the equations of the field in terms of the Riemann tensor is a giant leap. And the fact that it took Einstein three or four years to make the transition is understandable. Indeed, it is astonishing that he made the transition at all. One can undoubtedly argue that mathematical insight was involved in translating the notion of the metrical origin of all accelerations, gravitational and inertial, in terms of Riemannian geometry. Indeed, it has sometimes been stated (as for example, by Dirac) that "when Einstein was working on building up his theory of gravitation he was not trying to account for some results of observations. His entire procedure was to search for a beautiful theory, a theory of a type that Nature would choose". It does not seem to me that this is a valid description. Let me go back to an earlier time when, after Einstein's formulation of the special theory of relativity, Minkowski provided its basis in geometric terms. You will recall that Einstein's formulation of special relativity was based on the Galilean principle that the velocity of light is the same in two frames of
Commentary on Dirac's "The excellence of general relativity"
55
reference in uniform relative motion. Minkowski's formulation essentially consists in restating this principle as equivalent to the invariance of physical laws to rotations in four-dimensional space-time with the metric ds2 = c2dt2-dx2-dy2-dz2.
(4)
But Einstein's first reaction to Minkowski's formulation was to describe it as " uberflussige Gelehrsamkeit". However, without Minkowski's formulation of special relativity in geometrical form, the transition to Riemannian geometry would not have been made. Indeed, Pais [8] has written in his biography on Einstein, ...in 1912 he adopted tensor methods and in 1916 acknowledged his indebtedness to Minkowski for having greatly facilitated the transition from special to general relativity.
It seems to me then, that Einstein was not looking for mathematical beauty any more than Rutherford was seeking for his simple law of scattering. Indeed, it was Einstein's incredible physical insight that all acceleration is metrical in origin that led him to undertake a study of Riemannian geometry with the triumphant conclusion we all know. It was not the beauty of Riemannian geometry that led him to the theory. It was rather the simplicity of his physical ideas that led him to the beauty of the mathematical theory. [V] The results of such a procedure is a theory of great simplicity and elegance in its basic ideas. One has an overpowering belief that its foundations must be correct quite independent of its agreement with observation.
I have commented on this already. However, it is well to remember that a theory that is derived from simple and elegant basic ideas need not be a "correct' one. In the case of general relativity, the 'simple and the elegant' ideas gave rise to a mathematical structure so consistent and rich in content that one's confidence in the theory results from these latter facts. Perhaps I could explain my meaning as follows: One commonly says that non-linear equations are difficult to solve. Indeed, Einstein remarks in one of his papers, "If only it were not so damnably difficult to find rigorous solutions". This remark is consistent with his earlier reaction to Schwarzschild's discovery of the exact solution for the space-time about a point mass, that he was "astonished that the problem could be solved exactly". The experience during the past twenty-five years has been exactly the opposite: so many problems of physical significance have been solved exactly that one is almost tempted to say that the test whether a question one asks of general relativity is physically significant or not is the solvability of the problem when it is properly formulated.
56
S.
Chandrasekhar
References [1] P.A.M. Dirac, in: Einstein: the first hundred years, eds. M. Goldsmith, A. Mackay and J. Woudhuysen (Pergamon Press, Oxford, 1980) p. 41. [2] A. Pais, 'Subtle is the Lord': the science and the Ufe of Albert Einstein (Oxford University Press, New York, 1982) p. 129. [3] S. Chandrasekhar, Mitteilungen der Astronomischen Gesellschaft Nr. 67, Hamburg (1986) p. 19. [4] A. Einstein, in: History of physics, eds. S. Weart and M. Philhps (American Institute of Physics, New York, 1985) p. 243. [5] A. Einstein, Essays in science (Wisdom Library: a division of the Philosophical Library, New York, 1934) p. 50. [6] A. Einstein, The meaning of relativity (Princeton University Press, Princeton, NJ, 1937) p. 57. [7] E.T. Whittaker, Analytical dynamics (University Press, Cambridge, 1937) p. 39. [8] A. Pais, 'Subtle is the Lord': the science and the Ufe of Albert Einstein (Oxford University Press, New York, 1982) p. 152.
T H E INCREASING ROLE OF GENERAL RELATIVITY IN ASTRONOMY By S. Chandrasekhar Halley Lecture for igj2. Delivered in Oxford, May 2 M R . VICE-CHANCELLOR, LADIES AND GENTLEMEN,—
In a memorable essay, Maynard Keynes wrote Newton was not the first of the age of reason. He was the last of the magicians, the last of the Babylonians and Sumerians, the last great mind that looked out on the visible and the intellectual world with the same eyes as those who began to build our intellectual inheritance rather less than 10,000 years ago. If Newton was not the first of the age of reason, that place could be fairly claimed for Halley. Halley's attitude and approach to the physical sciences was in no way different from ours. In particular, he sought to find, in Nature, manifestations of the basic physical laws. Thus, having convinced himself that the existence of periodic comets is consistent with Newton's laws of gravitation, he set about to discover if there was evidence for them, with results that are well known. In view of Halley's enthusiasm for the Newtonian theory of gravitation, it is, in the first instance, somewhat surprising that none of the earlier Halley lectures should have been devoted to the role which the general theory of relativity may be expected to play in astronomy. This theory which has been described by Hermann Weyl as "one of the greatest examples of the power of speculative thought" was founded on the recognition that the Newtonian theory requires modifications if it is to be compatible with other parts of physics such as electro-dynamics and optics. And in the twenties, the general theory of relativity was an intoxicating subject. Thus, Eddington described his part in the verification of Einstein's prediction of the deflection of light at the solar eclipse of 1919 as "the most exciting event" in his connection with astronomy. And the meeting of the Royal Society in London on
October 1972
The Increasing R6le of General Relativity in Astronomy
161
November 6, 1919 at which the results of this expedition were reported, has been described by Whitehead in the following terms: T h e whole atmosphere of tense interest was exactly that of the Greek drama: we were the Chorus commenting on the decree of destiny as disclosed in the development of a supreme incident. There was dramatic quality in the very setting—the traditional ceremonial and in the background the picture of Newton to remind us that the greatest scientific generalization was now, after more than two centuries, to receive its first modification. In spite of the excitement of the early twenties and in spite also of the fact that the author of general relativity was, in due time, to become the most celebrated representative of science in the twentieth century, the theory itself was not pursued with intensity, in the framework either of physics or of astronomy, during the decades that followed. I am concerned here only with astronomy. And from one point of view, the reasons for the gradual abandonment of the study of the general theory of relativity in astronomy are not far to seek. The effects of general relativity were identified with Einstein's three tests; and their manifestations in the motions of the planets and the Moon were admittedly very small. As such, general relativity did not appear too relevant in the broader contexts of astronomy. Cosmology was, of course, an exception. And in cosmology, general relativity did indeed play a vital role: it directed the course of the observational researches in "the realm of the nebulae", once the expansion of the Universe had been established by Hubble. These cosmological aspects have been referred to in the earlier Halley lectures by Hubble, Sandage, and Schmidt. But cosmology, in spite of its fundamental interest for the physical sciences, is not, if I may say so, the staple of astronomy. But what is the staple of astronomy ? I am afraid that I may be starting a controversy, wholly foreign to my intentions, if I were to be dogmatic about this matter in any way. Certainly, I do not wish to arrogate to myself the wisdom or the prerogative to define the purposes or the elements of a science. So let me be pragmatic and state only that in the past questions pertaining to the continuing outpouring of energy by the stars and the other bodies constituting the astronomical Universe have always occupied a central place: they have stimulated and directed the course of astronomical development on a wide front. And it is in this sense that I envisage an increasing role for general relativity in astronomy. But first, I should like to describe briefly the background against which this increasing role for general relativity may be projected. When one thinks of stellar energy, the question that occurs to one, almost by reflex reaction, concerns the source of the continuing luminosity of the Sun. T h e Sun is constantly radiating energy to the space outside; and so far as one can tell, it has done so at its present rate for at least a few thousand million years. And the principal question concerning this continual outpouring of energy is not so much its intensity as its duration. Let me be more specific. In the nineteenth century, the only known physical process that could release energy from a self-gravitating mass such as a star is by a slow secular
16a
The Increasing R6le of General Relativity in Astronomy
Vol. 92
contraction. By such contraction, gravitational potential energy is released; and while a fraction of this released energy goes towards raising the average temperature of the star, the remaining fraction is available for radiation to space outside. It was in terms of such a contraction hypothesis that Kelvin and Helmholtz sought to account for the radiation of the Sun and the stars. As a physical process that could play a role in astronomy, it is an eminently reasonable one. Indeed, as we now know, it does play a part in current schemes of stellar evolution. But when applied to the Sun to account for its radiation, the hypothesis failed because it provided a continuing source of energy for a period of only a few million years, contrary to the evidence from many directions that the age of Earth's crust must already be measured in thousands of million years. Even though the contraction hypothesis of Kelvin and Helmholtz failed with reference to the Sun, it played an important role in many important astronomical and related developments: it provided, for example, the principal impetus for calibrating and in some cases replacing the qualitative methods of geology for estimating the ages of rocks by the quantitative methods of dating by the content of radioactive minerals. And as I said, it continues to play a part in current schemes of stellar evolution. As I stated, the contraction hypothesis failed the test of duration. One was, therefore, forced to seek sources of energy that could last longer: in particular, provide for the Sun a life of at least io 1 0 years. And the possibility of such a source became evident when it was noted that the mass of the helium nucleus was less than four times the mass of the proton by o-8 per cent. Consequently, if some way could be found for synthesizing a helium nucleus out of four protons, then the energy equivalent of o-8 per cent of the mass of the proton would be freed as available energy. And it was apparent, already in the twenties, that if this energy derived from the binding of the helium nucleus could be released, then we would be assured for the Sun an energy source that could pass the test of duration. Whether the contemplated nuclear transformation could take place under the conditions of temperature and pressure that had been deduced by Eddington and others during the twenties, for the interiors of the stars, could not be confirmed before the advent of quantum mechanics and nuclear physics. T h e understanding of how protons of even relatively low energy can combine to form deuterium nuclei, and also penetrate the nuclei of carbon and nitrogen with sufficient probabilities to result in the effective burning of hydrogen into helium, came only during the late thirties. And much needed information on the crosssections for the various nuclear reactions was obtained during the fifties. In any event, it is a fact that the detailed understanding of the source of the energy of the stars provided the central inspiration for much of the astronomy of the fifties and the sixties. It made possible the development of a detailed theory of stellar evolution; and these theoretical developments in turn stimulated most of the observational photometric studies of star clusters and galaxies; and by directing the studies towards the determination of their ages, they naturally led to the revised cosmological distance-scale and a re-evaluation of the Hubble constant. It appears that the increasing role of general relativity in astronomy that one may witness during the seventies and the eighties will not be unlike the role of nuclear physics during the fifties and the sixties. To state this same view somewhat differently, it appears that physical processes, understandable
October 1972
The Increasing Role of General Relativity in Astronomy
163
only in terms of the general theory of relativity, can satisfy the astronomical requirements that seem to be beyond the scope of nuclear physics. T o mention quasi-stellar sources, radio galaxies, the violent events that seem to occur in the centres of galaxies, and the events reported by Weber and attributed by him to bursts of energy in the form of gravitational waves coming from the centre of our Galaxy, to mention all these, is to conjure up a list that demands processes that will release much larger fractions of the rest mass as energy than the paltry one per cent provided by the binding energies of nuclei. Such processes do seem possible in the framework of general relativity; and the kind of phenomena that must be antecedent to such processes also seem necessary in the larger astronomical contexts. I shall consider first these larger astronomical contexts and then return to the particular processes of energy release that general relativity suggests. In considering the problem of solar energy, I emphasized that in seeking sources of energy for astronomical bodies duration is an essential desideratum. In the case of the Sun, nuclear processes involving light nuclei meet the test: they provide for the Sun a life of some ten thousand million years at its present rate of radiation. By the same token, they cannot provide comparable lives for stars that are substantially more massive than the Sun; for massive stars are proportionately far more luminous. T h u s a star that is ten times more massive than the Sun has a luminosity that is ten thousand times greater. These stars will accordingly burn up their available nuclear fuel in a relatively much shorter time: indeed, they cannot endure for more than ten to twenty million years. Since our Galaxy, in somewhat its present form, must have lasted for a period at least a thousand times longer, the conclusion is inescapable that these stars are young and that they must have been formed within the last ten to twenty million years. An immediate corollary that follows from this last conclusion is that the process of star formation is a continuing process in the Galaxy. On this account the question of the eventual fate of these short-lived massive stars becomes one of central importance for astronomy. And the relevance of this question in the more general context of stellar evolution was recognized long before the problem of stellar energy was clarified. Indeed, the question occurs to one almost inevitably: for no matter what the source of energy is, it must be exhausted sooner or later; and sooner or later the question must be confronted. T h e question was in fact formulated by Eddington in one of his famous aphorisms: "a star will need energy in order to cool". And in the spirit of this aphorism we can ask: will the massive stars have the necessary energy to cool ? Let me rephrase the statement and the question in a less oracular fashion. T h e stellar material in the interiors of the normal stars is mainly in the form of atomic nuclei and electrons: and except in a purely ionized hydrogen-gas, the electrons outnumber the nuclei by a factor exceeding two; and in general they contribute by far the larger share to the total gas pressure. T h e question we ask is: can a star of assigned mass, composed of such matter, attain a state of zero-point energy at a high density? Or, to state the question as R. H. Fowler framed it, can the star attain a state in which it can be described "as a gigantic molecule in its lowest quantum state?" A pioneering investigation by R. H. Fowler in 1926 seemed to suggest that such a state was possible in terms of the equation of state that must govern the electron gas as its concentration is increased at a fixed temperature. This
164
The Increasing Role of General Relativity in Astronomy
Vol. 92
limiting form of the equation of state of an electron gas can be derived from the following picture. We describe the states of an electron gas by quantum numbers, even as we describe the electrons in an atom by quantum numbers. In the limit of high enough concentrations, all the states for the electron with momenta less than a certain threshold value p0 are occupied consistently with Pauli's principle, namely, that no more than one electron can occupy a state of assigned quantum numbers. While states below p0 are all occupied, the states above pQ are all empty. This is the completely degenerate state for an electron gas. Under these conditions it can be shown that the relation between the pressure (p) and the electron concentration (ne) is of the form p — kx (ne)5l3, where k^ is an atomic constant.
FIG.
1
T h e full-line curve represents the exact (mass—radius) relationship for completely degenerate configurations. T h e mass, along the abscissa, is measured in units of the limiting mass (denoted by Mj) and the radius, along the ordinate, is measured in the unit l^=7"rz /ie~ 1 Xio 8 cm. T h e dashed curve represents the relation M = constant X R~* that follows from the equation of state p = kx (« e ) 6 / 3 ; at the point B along this curve, the threshold momentum p0 of the electrons at the centre of the configuration is exactly equal to mc. Along the exact curve, at the point where a full circle (with no shaded part) is drawn, p0 (at the centre) is again equal to mc; the shaded parts of the other circles represent the regions in these configurations where the electrons may be considered to be relativistic (p0 >mc). (This illustration is reproduced from S. Chandrasekhar, M.N., 95, p . 219, 1935.)
October 1972
The Increasing Role of General Relativity in Astronomy
165
On the basis of this equation of state, one can readily determine the structure which a configuration of an assigned mass M will assume when in equilibrium under its own gravity. One finds that equilibrium states are possible for any assigned mass: one finds in fact a mass-radius relation of the form M = constant X R~3. Accordingly, the larger the mass, the smaller is its radius. Also the mean densities of these configurations are found to be in the range of I O 6 - I O 6 grams per c.c. when the mass is of solar magnitude. These masses and densities are of the order one meets in the so-called white-dwarf stars. And it seemed for a time that the white-dwarf stage (or rather the "black-dwarf" stage as Fowler described it) represented the last stage of stellar evolution for all stars. Since a finite state seemed possible for any assigned mass, one could rest with the comfortable assurance that all stars will have the "necessary energy to cool". But this assurance was soon broken when it was realized that the electrons in the centres of degenerate masses begin to have momenta comparable to mc where m is the mass of the electron. Accordingly, one must allow for the effects of special relativity. These effects can be readily allowed for and look harmless enough in the first instance: the correct equation of state, while it approximates to that given before for low enough electron concentrations, tends to p = k2 (« e ) 4 ' 3 as the electron concentration increases indefinitely. (k« is another atomic constant.) This limiting form of the equation of state has a dramatic effect on the predicted mass-radius relation: instead of predicting a finite radius for all masses, the theory now predicts that the radius must tend to zero as a certain limiting mass is reached. T h e value of this limiting mass is 5-76 / V 2 solar masses where fie denotes the mean molecular weight per electron. For the expected value /xe = 2, the limit is 1-44 solar masses. The existence of this limiting mass means that a white-dwarf state does not exist for stars that are more massive. In other words "the massive stars do not have sufficient energy to cool". Fig. 1 exhibits the mass-radius relation that was deduced in 1935 on the basis of the exact equation of state (of which the equations given before are the appropriate limiting forms). The conclusion that was reached at that time was stated in the following terms: The life-history of a star of small mass must be essentially different from the life-history of a star of large mass. For a star of small mass the natural white-dwarf stage is an initial step towards complete extinction. A star of large mass cannot pass into the white-dwarf stage and one is left speculating on other possibilities. Statements very similar to the one I have just quoted from a paper written 38 years ago frequently occur in current literature. But why, it may be asked, were these conclusions not accepted forty years ago ? The answer is that they did not meet with the approval of the stalwarts of the day. T h u s Eddington, commenting on the foregoing conclusion, stated: Chandrasekhar shows that a star of mass greater than a certain limit remains a perfect gas and can never cool down. T h e star has to go on radiating and radiating and contracting and contracting until, I suppose, it gets down to a few kilometres' radius when gravity becomes strong enough to hold the radiation and the star can at last find peace. If Eddington had stopped at that point, we should now be giving him credit
166
The Increasing Rdle of General Relativity in Astronomy
Vol. 92
for having been the first to predict the occurrence of black holes—a topic to which I shall return presently. But alas! he continued to say: I felt driven to the conclusion that this was almost a reductio ad absurdum of the relativistic degeneracy formula. Various accidents may intervene to save the star, but I want more protection than that. I think that there should be a law of Nature to prevent the star from behaving in this absurd way. And similarly E. A. Milne (who was a professor here in Oxford and was a great personal friend of mine) wrote: T o me it is clear that matter cannot behave the way you predict. In spite of the then prevalent opposition, it seemed to me likely that a massive star, once it had exhausted its nuclear sources of energy, will contract and in the process eject a large fraction of its mass; and further that if by this process, it reduced its mass sufficiently, it could find a state in which to settle. A theoretical advance in a different direction suggested another possibility. It is that as we approach the limiting mass along the white-dwarf sequence, we must reach a point where the protons at the centre of the configuration become unstable with respect to electron capture. T h e situation is this. Under normal conditions, the neutron is ^-active and unstable while the proton is a stable nucleon. But if in the environment in which the neutron finds itself (as it will in the centre of degenerate configurations near the limiting mass), all the electron states with energies less than or equal to the maximum energy of the jS-ray spectrum of the neutron are occupied, then Pauli's principle will prevent the decay of the neutron. In these circumstances the proton will be unstable and the neutron will be stable. At these high densities, the equilibrium that will obtain will be one in which, consistent with charge neutrality, there will be just exactly the right number of electrons, protons, and neutrons with appropriate threshold energies that none of the existing protons or neutrons decays. At these densities the neutrons will begin to outnumber the protons and electrons by large factors. In any event it is clear that once neutrons begin to form, the configuration essentially collapses to such small dimensions that the mean density will approach that of nuclear matter and in the range I O 1 3 - I O 1 5 grams per c.c. These are the neutron stars that were first studied by Oppenheimer and Volkoff in 1939, though their possible occurrence had been suggested by Zwicky some five years earlier. From the work of Oppenheimer and Volkoff it appeared likely that a massive star, during the course of its evolution, could collapse to form a neutron star if during the process of contracting it had reduced its mass sufficiently. T h e process would clearly be cataclysmic, and it seemed likely that the result would be a supernova phenomenon. But the formation of a neutron star, as the result of the collapse, will depend on whether a star, initially more massive than the limiting mass for the white-dwarf stars, ejects just the right amount of mass in order that what remains is in the permissible range of masses for stable neutron stars. While the question of the ultimate fate of massive stars with all its implications was not faced till recently, the theory of the white-dwarf stars, based on the relativistic equation of state for degenerate matter, gained gradual acceptance during the forties and fifties. T h e principal astronomical reasons
October 1972
The Increasing Role of General Relativity in Astronomy
167
for this acceptance were twofold. First, the number of the known whitedwarf stars had, in the meantime, increased very substantially, largely through the efforts of Luyten; and the study of their spectra, particularly by Greenstein, confirmed the adequacy and in some cases even the necessity of the theoretically deduced mass-radius relation exhibited in Fig. 1. Secondly, since a time-scale of the order of ten million years for the exhaustion of the nuclear sources of energy of the massive stars requires the continual formation of these stars, one should be able to distinguish a population of young stars from a population of old stars. Spectroscopic studies provided evidence that the chemical composition of the young stars differs systematically from the chemical composition of the old stars; and in fact the difference is in the sense that the young stars appear to have been formed from matter that has been cycled through nuclear reactions. This last fact is consistent with the picture that during the course of the evolution of the massive stars a large fraction of their masses is returned to interstellar space. It also seemed likely that this returning of processed matter to the interstellar space was via the supernova phenomenon. While all these ideas became a part of common belief, it remained only as belief. Their full implications were not seriously explored before the discovery of the pulsars. T h e discovery, in particular, of a pulsar (with the shortest known period) at the centre of the Crab nebula added much credence to the views that I have described, since the Crab nebula is itself the remnant of a supernova explosion that was observed by the Chinese and the Japanese astronomers in the year A.D. 1054 The discovery of the association of further pulsars (of longer periods) with what are believed to be the remnants of more ancient supernova explosions strengthens one's conviction. The story of the pulsars and their identification with neutron stars are matters of such common knowledge that I shall not spend any more time on them. T h e principal conclusions that follow from these theoretical and observational studies can be summarized very simply. Massive stars in the course of their evolution must collapse to dimensions of the order of ten to twenty kilometres once they have exhausted their nuclear source of energy. In this process of collapse, a substantial fraction of the mass will be returned (as processed matter) to the interstellar space. If the mass ejected is such that what remains is in the permissible range of masses for stable neutron stars, then a pulsar will be formed. The exact specification of the permissible range of masses for stable neutron stars is subject to uncertainties because of uncertainties in the equation of state for neutron matter; but it is definite that the range is narrow: the current estimate is between 0-3 to i-o solar mass. While the formation of a stable neutron star could be expected in some cases, it is clear that their formation is subject to vicissitudes. It is not in fact an a priori likely event that a star initially having a mass of, say, ten solar masses ejects, during an explosion, subject to violent fluctuations, an amount of mass just sufficient to leave behind a residue in a specified narrow range of masses. It is more likely that the star ejects an amount of mass that is either too large or too little. In such cases the residue will not be able to settle into a finite state; and the process of collapse must continue indefinitely till the gravitational force becomes so strong that what Eddington concluded is a reductio ad absurdum must in fact happen: "the gravity becomes strong enough to hold the radiation". In other words, a black-hole must form; and it is to the subject of black-holes that I now turn.
168
The Increasing Role of General Relativity in Astronomy
Vol. 92
Let me be more precise as to what one means by a black-hole. One says that a black-hole is formed when the gravitational forces on the surface become so strong that light cannot escape from it. That such a contingency can arise was surmised already by Laplace in 1798. Laplace argued as follows. For a particle to escape from the surface of a spherical body of mass M and radius R, it must be projected with a velocity v such that %v2 > GMjR; and it cannot escape if o a < iGM\R. On the basis of this last inequality, Laplace concluded that if R < 2GM/C 2 = Rs (say) where c denotes the velocity of light, then light will not be able to escape from such a body and we should not be able to see it! By a curious coincidence, the limit Rs discovered by Laplace is exactly the same that general relativity gives for the occurrence of a trapped surface around a spherical mass. (A trapped surface is one from which light cannot escape to infinity.) While the formula for R looks the same, the radial coordinate r (in general relativity) is so defined that 47tr2 is the area of the 3-surface of constant r; it is not the proper radial distance from the centre. That, for a radial coordinate r = Rs, the character of space-time changes is manifest from the standard form Schwarzschild's metric that describes the geometry of space-time external to a spherical distribution of mass located at the centre. For a mass equal to the solar mass, the Schwarzschild radius i?s has the value 2-5 kilometres. At one time, the thought that a mass as large as that of the Sun could be compressed to a radius as small as 2-5 km would havs seemed absurd. One no longer thinks so: neutron stars have comparable masses and radii. T h e problem we now consider is that of the gravitational collapse of a body to a volume so small that a trapped surface forms around it; as we have stated, from such a surface no light can emerge. Let us first consider in the framework of the Newtonian theory what would happen to an incoherent mass, with no internal pressure, distributed with exact spherical symmetry about a centre. It would simply collapse to the centre in a finite time since in the absence of pressure there is nothing to restrain the action of gravity. In the Newtonian theory this result of matter collapsing to an infinite density may be considered as a reductio ad absurdum of the initial premises: a distribution of matter that is exactly spherically symmetric and the absence of any sustaining pressure are both untenable in practice. If either of these two premises is not exactly fulfilled, then the collapse to an infinite density will not happen. Now if the same problem of pressure-free collapse is considered in the framework of general relativity—as it first was by Oppenheimer and Snyder— one finds, as in the Newtonian theory, that the matter collapses to the centre in a finite time (as measured by a co-moving clock). But in contrast to the Newtonian theory, the inclusion of pressure or the allowance of departures from exact spherical symmetry do not seem to make any difference to the final result. The reasons are the following. In general relativity, pressure contributes to the inertial mass; and this contribution becomes comparable to the contribution by the mass density when the radius of the object approaches the Schwarzschild limit. On this account, after a certain stage, the allowance for pressure actually facilitates, rather than hinders, the collapse. It is also clear that small departures from spherical symmetry cannot matter. For unlike in the Newtonian theory, the aim to the centre need not be perfect: it will suffice to aim within the Schwarzschild radius. More generally,
October 1972
The Increasing Role of General Relativity in Astronomy
169
theorems by Penrose and Hawking show that, in the framework of general relativity, allowing for factors that derive from pressure and lack of spherical or other symmetry will not prevent the matter from collapsing to a singularity in space-time if a certain well-defined point of no-return—an event horizon— is passed. If we consider then the gravitational collapse of a massive star and allow for all factors that derive from pressure but permit only small departures from spherical symmetry, the end result is the same as in a strictly sphericallysymmetric pressure-free collapse. There is no alternative to the matter collapsing to infinite density at a singularity once a point of no-return is passed. The reason is that once the event horizon is passed, all time-like trajectories must necessarily get to the singularity: "all the King's horses and all the King's men" cannot prevent it. And as far as the external observer is concerned the energy associated with the departures from spherical symmetry will be radiated away as gravitational waves; and the event horizon will eventually settle down to a smooth spherical surface with an exterior Schwarzschild metric for a certain M—the mass of the black-hole. It is important to notice that the phenomenon of spherical collapse will be described differently by an observer moving with the surface of the collapsing star and by an observer stationed at infinity. This difference is illustrated in Fig. 2. Imagine that the observer on the surface of the collapsing star transmits time signals at equal intervals (by his clock) at some prescribed wavelength (by his standard). So long as the surface of the collapsing star has a radius that is large compared to the Schwarzschild radius, these signals will be received by the distant observer at intervals that he will judge as (very nearly) equally spaced. But as the collapse proceeds, the distant observer will judge that the signals are arriving at intervals that are gradually lengthening and that the wavelength of his reception is also lengthening. As the stellar surface approaches the Schwarzschild limit the lengthening of the intervals as well as the lengthening of the wavelength of his reception will become exponential by his time. T h e distant observer will receive no signal after the collapsing surface has crossed the Schwarzschild surface; and there is no way for him to learn what happens to the collapsing star after it has receded inside the Schwarzschild surface. For the distant observer, the collapse to the Schwarzschild radius takes, strictly, an infinite time (by his clock) though the time scale in which he loses contact in the end is of the order of milliseconds. The story is quite different for the observer on the surface of the collapsing star. For him nothing unusual happens as he crosses the Schwarzschild surface: he will cross it smoothly and at a finite time by his clock. But once he is inside the Schwarzschild surface, he will be propelled inexorably towards the singularity: there is no way in which he can avoid being crushed to zero volume at the singularity and no way at all to retrace his steps. From what I have said, collapse of the kind I have described must be of frequent occurrence in the Galaxy; and black-holes must be present in numbers comparable to, if not exceeding, those of the pulsars. While the black-holes will not be visible to external observers, they can nevertheless interact with one another and with the outside world through their external fields. But one important generalization is necessary and essential.
170
The Increasing Rdle of General Relativity in Astronomy
Vol. 92
It is known that most stars rotate. And during the collapse of such rotating stars, we may expect the angular momentum to be retained except for that part which may be radiated away in gravitational waves. T h e question now arises as to the end result of the collapse of such rotating stars. One might have thought that the inclusion of angular momentum would
FIG. 2 Illustrating spherically symmetric collapse. At each point the future and the past light-cones are drawn; all time-like trajectories must lie within these cones. T h e reception by an observer, orbiting in a circular orbit at a large distance from the centre, of a light signal sent by an observer on the collapsing stellar surface is shown; it makes it clear why no signal sent after the surface passes inside the Schwarzschild surface at r = 2 m can be received by the orbiting observer. Also, notice how all future-directed time-like paths from any point inside r=zm must necessarily intersect the singularity at r=o. (This illustration is reproduced from R. Penrose, Nuovo Cimento, Serie I, Vol. I, p . 252, 1969.)
October 1972
The Increasing Role of General Relativity in Astronomy
171
make the problem excessively complicated. But if the current ideas are confirmed, the expected end result is not only simple in all essentials, it also provides the principal justification for anticipating an increased role for general relativity in astronomy. In 1963, Kerr discovered the following solution of Einstein's equations for the vacuum which has two parameters M and a and which is also asymptotically flat: As2 = -
- 2 [At-a sin 2 6 A<j>}2 + ~
[(r2+a2) A
+ £ Ar2 + p2 Ad2, where p2 = r2 + a2 cos 2 6 and A = r2 + a2—zMr. (The solution is written in units in which c = G = 1; and in a system of coordinates introduced by Boyer and Lundquist.) Kerr's solution has rotational symmetry about the axis 6=0: none of the metric coefficients depends on the cyclic coordinate >. It is, moreover, stationary: none of the metric coefficients depends on the coordinate t which is time for an observer at infinity. Kerr's solution reduces to Schwarzschild's solution when a = o . A test particle describing a geodesic in Kerr's metric at a large distance from the centre will describe its motion as in the gravitational field of a body having a mass M and an angular momentum J = aM (as deduced from the Lens-Thirring effect). It is now believed that the end result of the collapse of a massive rotating star is a black-hole with an external metric that will eventually be Kerr's, all the asymmetries having been radiated away. I shall not attempt to explain the reasons for this belief except to say that they derive, principally, from a theorem of Carter which essentially states that sequences of axisymmetric metrics, external to black holes, must be disjoint, i.e. have no members in common. T h e Kerr metric, like Schwarzschild's, has an event horizon; it occurs at Q r = ., [M+(M2—a2)*]. In writing this formula, I have assumed that a < M; if this should not be the case, there will be no event horizon and we shall have a "naked singularity", i.e. a singularity that will be visible and communicable to the outside world. For the present, I shall restrict myself to the case a < M. Trajectories, time-like or null, can cross the event horizon from the outside; but they cannot emerge from the inside. In this respect also the Kerr black-hole is like the Schwarzschild black-hole. But unlike the Schwarzschild metric, the Kerr metric defines another surface (the stationary limit), external G to the event horizon, whose equation is r = ~2 [M+(M2—a2 cos 2 0)*]. This surface touches the event horizon at the poles; and it intersects the equator (0 = n/z) on a circle whose radius ( = 2 GMjc2) is larger than that of the horizon. On this surface, an observer who considers himself as staying in the same place must travel with the local velocity of light: like Alice, he must run as fast as he can to stay exactly where he is! Light emitted by such an observer must accordingly appear as infinitely red-shifted to one stationed at infinity.
172
The Increasing Role of General Relativity in Astronomy
Vol. 92
The occurrence of the two separate surfaces in the Kerr geometry gives rise to unexpected possibilities. These possibilities derive from the fact that in the space between the two surfaces—termed the ergo-sphere by Wheeler and Ruffini—the coordinate t, which is time-like external to the stationary limit, becomes space-like. Therefore, the component of the four-momentum in the ^-direction, which is the conserved energy for an observer at infinity, becomes space-like in the ergo-sphere; it can accordingly assume here negative values. In view of these circumstances, we can contemplate a process in which an element of matter enters the ergo-sphere from infinity and splits here (in the ergo-sphere) into two parts in such a way that one part, as judged by an observer at infinity, has a negative energy. Conservation of energy requires that the other part acquire an energy that is in excess of that of the original element. If the part with the excess energy escapes along a geodesic to infinity while the other part crosses the event horizon and is swallowed up by the black hole, then we should have extracted some of the rotational energy of the black hole by reducing its angular momentum. T h e possibility that such processes can be realized was first pointed out by Penrose. In considering the energy that could be released by interactions with black holes, a theorem of Hawking is useful. Hawking's theorem states that in the interactions involving black holes, the total surface area of the boundaries of the black holes can never decrease; it can at best remain unchanged (if the conditions are stationary). Now, the surface area of a Kerr black-hole is given by 5 = ^ G2 M [M+(M2-a2)i]. cr By Hawking's theorem, in a process in which energy is extracted from a Kerr black-hole, M and a must both change in such a way that S increases. By writing
M*=M2ir+rkMir2,
where M„=\ { [ M + ( M 2 - a 2 ) i ] 2 + « 2 }* and J ( = a M ) is the angular momentum, Christodoulou has shown that Hawking's condition, 8S > o, is equivalent to SM,> > o. Accordingly, we may consider M{r as the irreducible mass of the Kerr black-hole in the sense that by no interaction with the black-hole, effected by the injection of small amounts matter into it, can we reduce the value of M,>. T h e contribution to M 2 by the term J 2 /4 M, r 2 represents therefore the maximum rotational energy that can be extracted. Another example illustrating Hawking's theorem (and considered by him) is the following. Imagine two spherical (Schwarzschild) black holes, each of mass \ M, coalescing to form a single black hole; and let the black hole that is eventually left be, again, spherical and have a mass $01. Then Hawking's theorem requires that 16 7c 9 K 2 > i 6 71 [2 ( J M ) 2 ] = 8 TTM 2 , or Hence the ma&imum amount of energy that can be released in such a coalescence is
October 1972
The Increasing Role of General Relativity in Astronomy
173
M (1 - 1/V2) t 2 = o - 2 9 j Mc\ In practice, the actual amount may be much less; but it is clear that the processes of the type considered have the potentialities for releasing far larger fractions of the rest mass as energy than nuclear processes. In connection with Weber's observations to which we have referred earlier, the possibility of a large black hole at the centre of the Galaxy and with a mass in the range I O 4 - I O 8 solar masses has often been suggested, for instance by Lynden-Bell and by Bardeen. It may be supposed that such a black hole would be continuously swallowing up stars and accreting matter. As each star is swallowed, or when matter is accreted, we may expect that a certain fraction of the mass energy is radiated as gravitational waves. And strong tidal forces that would also be operative under these circumstances may produce considerable attendant effects such as electromagnetic radiation. Various proposals in these directions are currently actively being pursued. Even if all these attempts to account for Weber's events—their frequency and their energy content—fail, there still remains the question whether naked singularities may not appear under certain circumstances with undreamt-of possibilities though the present view is that "singularities will, forever, remain concealed." It is clear that none of the processes for energy release that I have described is anything more than a mere suggestion. T h e present situation is not unlike that in the twenties when the conversion of hydrogen into helium was contemplated as a source of stellar energy with no sure knowledge that it could be accomplished; only years later were well-defined chains of nuclear reactions that could accomplish it formulated. We may similarly have to wait for some years now. In discussing the various possibilities that may arise as the result of interactions with black holes and among black holes, we are today considering seriously situations that were brushed aside as reductio ad absurdum not so very long ago. For my part, while considering the phenomena associated with event horizons and the impossibility of communication across them, I have often recalled a parable from Nature that I learnt in India fifty years ago. T h e parable, entitled "Not lost but gone before", is about larvae of dragonflies deposited at the bottom of a pond. A constant source of mystery for these larvae was what happens to them, when on reaching the stage of chrysalis, they pass through the surface of the pond never to return. And each larva, as it approaches the chrysalis stage and feels compelled to rise to the surface of the pond, promises to return and tell those that remain behind what really happens, and confirm or deny a rumour attributed to a frog that when a larva emerges on the other side of their world it becomes a marvellous creature with a long slender body and iridescent wings. But on emerging from the surface of the pond as a fully-formed dragonfly, it is unable to penetrate the surface no matter how much it tries and how long it hovers. And the history books of the larvae do not record any instance of one of them returning to tell them what happens to it when it crosses the dome of their world. And the parable ends with the cry . . . Will none of you in pity, T o those you left behind, disclose the secret ?
174
The Increasing R6le of General Relativity in Astronomy
Vol. 92
References Lord Keynes, Newton Tercentenary Celebrations (Royal Society 1947) p. 27. A. S. Eddington, Background to Modern Science (Edited by J. Needham and W. Pagel; Cambridge University Press 1938), p. 140. A. N. Whitehead, Science and the Modern World (Macmillan, New York; 1926) P- 43A. S. Eddington, Internal Constitution of the Stars (Cambridge University Press; 1926) p. 173. R. H. Fowler, M.N., 87, 114, 1926. S. Chandrasekhar, Phil. Mag., 11, 592, 1931. S. Chandrasekhar, Ap. J., 74, 81, 1931. S. Chandrasekhar, M.N., 9 1 , 456 1931; see also, Z. f. Ap., 5, 321, 1932. S. Chandrasekhar, M.N., 95, 207, 1935. S. Chandrasekhar, The Observatory, 57, 373, 377, 1934. A. S. Eddington, The Observatory, 58, 38, 1935. S. Chandrasekhar, M.N., 95, 226, 1935; see p. 237. S. Chandrasekhar, Amer.J. Phys., 37, 577, 1969. J. R. Oppenheimer and G. M. Volkoff, Phys. Rev., 55, 374, 1939. B. K. Harrison, K. S. Thorne, M. Wakano, J. A. Wheeler, Gravitation Theory and Gravitational Collapse (Univ. Chicago Press; 1965). S. Chandrasekhar, in Colloque Internationel d'Astrophysique X I I I ; A. J. Shaler, Novae and White Dwarfs (Vol. I l l , White Dwarfs) (Hermann & Cie, Editeurs; Paris 1941) p. 245. P. S. Laplace, Systeme du Monde, Book 5, Chap. VI (as quoted by A. S. Eddington in Internal Constitution of the Stars (Cambridge U. Press; 1926) p. 6). J. R. Oppenheimer and H. Snyder, Phys. Rev., 56, 455, 1939. S. W. Hawking and R. Penrose, Proc. Roy. Soc, A, 314, 529, 1970. R. H. Price, Phys. Rev., 1972 (in press.) R. P. Kerr, Phys. Rev. Letters, 11, 237, 1963. R. H. Boyer and R. W. Lundquist, J. Math. Phys., 8, 265, 1967. B. Carter, Phys. Rev. L,etters, 26, 331 1971. S. Chandrasekhar and J. Friedman, Ap. J. (1972, in press). R. Ruflini and J. A. Wheeler, The Significance of Space Research for Fundamental Physics (ESRO, Paris, 1071). R. Penrose, Nuovo Cimento, Serie I, 1, 252, 1969. S. W. Hawking, Comm. Math. Phys., 25, 152, 1972. D. Christodoulou, Phys. Rev. Letters, 25, 1596 1970. D. Lynden-Bell, Nature, 223, 690, 1969. J. Bardeen, Nature, 226, 64, 1970. C. W. Misner, Phys. Rev. Letters, 28, 994, 1972. R. Penrose, Nature, 236, 377, 1972.
Printed by Sumfield & Day Ltd. Eastbourne, Sussex, England
Why Are the Stars as They Are? S.
CHANDKASEKHAB
University of Chicago - Chicago, 111.
This lecture recalls some considerations of forty and more years ago which may have some relevance to the current topics which are the concern of this session of the International School of Physics.
1. — Eddington's parable. Domains of natural phenomena are often circumscribed by well-defined scales, and theories concerning them are successful only to the extent that these scales emerge naturally in them. Thus, to the question « Why are the atoms as they are? », the answer «Because the Bohr radius—7&2/(47t2TOee2)~ •~0.5-10~ 8 cm—provides a correct measure of their dimensions* is apposite. In a similar vein, we may ask «Why are the stars as they are? », intending by such a question to seek the basic reason why modern theories of stellar structure and stellar evolution prevail. EDDINGTON [1] effectively posed this question to himself and answered it in his parable of a physicist on a cloudbound planet. His parable as he told it is the following. « The outward flowing radiation may be compared to a wind blowing through the star and helping to distend it against gravity. The formulae to be developed later (eq. (42) in sect. 5 below) enable us to calculate what proportion of the weight of the material is borne by this wind, the remainder being supported by the gas pressure. To a first approximation the proportion is the same at all parts of the star. I t does not depend on the density nor on the opacity of the star. I t depends only on the mass and molecular weight. Moreover, the physical constants employed in the calculation have all been measured in the laboratory, and no astronomical data are required. We can imagine a physicist on a cloud-bound planet who has never heard tell of the stars calculating the ratio of radiation pressure to gas pressure for a series of globes of gas of various sizes, starting, say, with a globe of mass 10 g, then 100 g, 1000 g and so on, so that his w-th globe contains 10" g. Table I shows the more interesting part of his results ». 1 - Rendiconti S.I.F. - LXV
1
2 TABLE
I.
No. of globe
Radiation pressure
Gas pressure
No. of globe
Radiation pressure
Gas pressure
32
0.0016
0.9984
36
0.951
0.049
33
0.106
0.894
37
0.984
0.016
34
0.570
0.430
38
0.9951
0.0049
35
0.850
0.150
39
0.9984
0.0016
« The rest of the table would consist mainly of long strings of 9's and O's. Just for the particular range of mass about the 33rd to 35th globes the table becomes interesting, and then lapses back into 9's and O's again. Eegarded as a tussle between matter and aether (gas pressure and radiation pressure) the contest is overwhelmingly one-sided except between numbers 33-35, where we may expect something to happen ». « What " h a p p e n s " is the stars. » « We draw aside the veil of cloud beneath which our physicist has been working and let him look up at the sky. There he will find a thousand million globes of gas nearly all of mass between his 33rd and 35th globes—that is to say, between | and 50 times the Sun's mass. The lightest known star is about 3-10 3 2 g and the heaviest about 2.10 35 g. The majority are between 10 33 and 1034 g, where the serious challenge of radiation pressure to compete with gas pressure is beginning. »
2. — The (1—/?*)-theorem and the combination
(hc/G^H-2.
There are two curious aspects to Eddington's parable. The first is that, while a combination of natural constants of the dimensions of a mass and of stellar magnitude is clearly implied by the calculation, it is not explicitly isolated—a surprising omission in view of Eddington's later propensities. The second is the logical lacuna in the argument: why is the relative extent to which radiation pressure provides support against gravity a relevant factor to the «happening)) of stars (*)? The omission is easily rectified and the argument at least ameliorated. Consider an enclosure containing matter and radiation. Let the matter be in a state of a perfect gas (in the classical Maxwellian sense) so that the (*) Besides, the values listed in table I are based on the unlikely value of a mean molecular weight ,u = 4. On this least point, see Eddington's [1] own discussion.
WHY ARE THE STABS AS THEY ARE?
3
pressure due to it is given by (1)
*~ =
^
T
'
where k denotes the Boltzmann constant, H the mass of the hydrogen atom, fx the mean molecular weight, Q the density and T the temperature. The pressure due to radiation, in the same enclosure, is given by
(2)
P^W,
where a denotes Stefan's constant. Consequently, if radiation pressure contributes a fraction 1 — /? to the total pressure P, then
{3)
p
=^\aIi=mQT-
We may eliminate T from these relations and express P in terms of Q and /? instead of g and T. Thus
and (5)
'-mv^*-™*
<«"•
Now there is a general theorem (CHANDRASEKHAR [2]) which states that the pressure Pc at the center of a star of mass M, in hydrostatic equilibrium and in which the density g(r) at any point r does not exceed the mean density §(r) interior to that point r, must satisfy the inequality (6)
i(?(|^^if*
where Q denotes the mean density of the star and QC its density at the center. The content of this theorem is no more than the assertion that the actual pressure at the center of a star must be intermediate between those at the centers of two configurations of uniform density, one at a density equal to the mean density Q of the star and the other at a density equal to the density QC at the center. If the inequality (6) should be violated, then there must be some regions in the star in which adverse density gradients prevail; and the occurrence of such adverse density gradients will lead to instabilities. In other
S. CHANDEASEKHAB
4
words, we may consider conformity with the inequality (6) as a necessary condition for the stable existence of a star. The right-hand part of the inequality (6) together with Pc given by eq. (5) requires, as a condition for the existence of a stable star,
or, equivalently, «
/ 6 \ * | 7 * \ 4 3 1— j8, i 1 01'
where in the foregoing inequalities p\, is the value of /? at the center of the star. Now, Stefan's constant (by virtue of Planck's law) has the value (9)
a =
8?r5ft4 15h^-
Inserting this value of a in the inequality (8), we obtain
First we observe that the inequality (10) has isolated the following combination of natural constants of the dimensions of a mass and of stellar magnitude (cf. CHANDEASEKHAB [3]) :
Eeturning to the inequality (8), we can express it as providing an upper bound to 1 — /?„; thus (12)
l-0e
where 1 — /?* is uniquely determined by the mass M and the mean molecular weight [i by the quartic equation
We may suppose that the following tabulation of 1 — /?* was made by Eddington's physicist on the cloud-bound planet, who, after isolating the combination of natural constants (11), had concluded, from its existence, a reason for the « happening » of stars.
WHY ABE THE STABS AS THEY ARE? TABLE
5
II.
WO)/*2
(M/O)/**
1-/3*
0.56
0.01
15.49
0.50
1.01
0.03
26.52
0.60
2.14
0.10
50.92
0.70
3.83
0.20
122.5
0.80
6.12
0.30
224.4
0.85
9.62
0.40
519.6
0.90
1-/3*
3. — «Have the stars enough energy to cooI?»: Eddington's paradox and Fowler's resolution. The same combination of natural constants (11) emerged soon afterwards in the context of resolving a paradox EDDINGTON [4] had formulated in the form of one of his famous aphorisms: «a star will need energy to cool». The paradox arose while considering the ultimate fate of gaseous stars in the light of the then new knowledge that stars, such as the companion of Sirius, exist which have mean densities in the range (10 5 -M0 6 ) g cm~3. As EDDINGTON stated, « I do not see how a star which has once got into this compressed condition is ever going to get out of it. ... I t would seem that the star will be in an awkward predicament when its supply of subatomic energy fails. ». Or, as P O W L E E [5] stated: «The stellar material will have radiated so much energy that it has less energy than the same matter in normal atoms expanded at the absolute zero of temperature. If part of it were removed from the star and the pressure taken off, what could it do? ». Quantitatively, the question arises in the following way. An estimate of the electrostatic energy Ev per unit volume of an assembly of atoms of atomic number Z ionized down to their bare nuclei is given by (11)
JS F =1.32-10 11 Z 2 g-3-,
whereas the kinetic energy EUn per unit volume of the free particles, under the assumption that they are free as in a perfect gas at a density Q and a temperature T, is given by
Now, if such matter were released of the pressure to which it is subject, it can resume the state of ordinary normal atoms only if (16)
A^>Ey,
6
S. CHANDRASEKHAK
or, according to eqs. (14) and (15), only if (17)
p<(0.94-10-3T/^2
This inequality will be clearly violated if the density is sufficiently high. This is the essence of Eddington's paradox (though at a later time he disclaimed to this formulation). FOWLEK [5] resolved this paradox in 1926 in a paper entitled « Dense matter»—one of the great landmark papers in the entire subject of stellar structure and stellar evolution: for, in it the notions of Fermi statistics and electron degeneracy are introduced into astrophysics for the first time. In a completely degenerate electron gas, all the available parts of the phase space, with momenta less than a certain threshold value p0, are occupied consistently with Pauli's exclusion principle. If n{p)dp denotes the number of electrons, per unit volume, with momenta between p and p + dp, then the assumption of complete degeneracy is equivalent to the assertion 9,71 £2
(18)
(P
n(p) = (P>Po) •
The threshold momentum pQ is determined by the normalizing condition
(19)
n = in(p)dp = ^-sp3g,
where n denotes the number of electrons per unit volume. For the distribution given in eq. (18), the pressure P and the kinetic energy _Ekin of the electrons (per unit volume) are given by
(20)
r,
8^ f ,
,
0
and 3>0
(21)
2?kin = ^
jp*T \p*T,dp x<
where vv and Tv are the velocity and the kinetic energy of an electron having
WHY ARE THE STABS AS T H E T ABE?
7
a momentum p. If we set (22)
v„ = plm
and
T„ = p2l2m
(appropriate for nonrelativistic velocities) in eqs. (20) and (21), we find
and
(24)
^ =10h ^:rf 3 ro m
=40
s
gs^
Fowler's resolution of Eddington's paradox consists in this. At the temperature and densities that may be expected to prevail in the interiors of the white-dwarf stars, the electrons will be highly degenerate and 2?tln should be evaluated in accordance with eq. (24), not in accordance with eq. (15); and eq. (24) gives
JUT>
A comparison of eqs. (14) and (25) now shows that for sufficiently high densities _Bkln > Er; and Eddington's paradox does not arise. F O W L E B concluded his paper with the following perceptive statement: « The black-dwarf material is best likened to a single gigantic molecule in its lowest quantum state. On the Fermi-Dirac statistics, its high density can be achieved in one and only one way, in virtue of a correspondingly great energy content. But this energy can no more be expended in radiation than the energy of a normal atom or molecule. The only difference between blackdwarf matter and a normal molecule is that the molecule can exist in a free state while the black-dwarf matter can only so exist under very high external pressure. »
4. — The theory of degenerate configurations; the limiting mass. The internal energy ( = 3P/2) of a degenerate electron gas that is associated with a pressure P is zero-point energy; and the essential content of Fowler's paper is that this zero-point energy is so great that we may expect a star to eventually settle down to a state in which all of its energy is of this form. Fowler's argument can be formulated more explicitly in the following manner.
S. CHANDRASEKHAK
8
According to the expression for the pressure given in eq. (23), we have the relation (26)
P = Kl6i,
r1 - 1 (3Y hTTT^f' *
where
# i - ^ , -
f
20 W {H.H'.
and fj,e is the mean molecular weight per electron. Equilibrium configurations in which the pressure is a monomial of the density are the polytropes of Emden and an equilibrium configuration in which the relation (26) obtains is a polytrope of index f. And the theory of polytropes directly leads to the relation (CHANDEASEKHAE [6])
(27)
l?i = 0.424
GM3B,
or, numerically, (28)
log (B/Be) = - i log (Jf/©) - 1 log p . - 1.40 .
For a mass equal to the solar mass and ^ — 2, the relation (28) predicts a radius R — 1.3-10 _2 J2 o and a mean density of 7-10 5 gcm- 3 . These values are precisely of the order of the radii and mean densities which are encountered in white-dwarf stars. Moreover, according to eqs. (27) and (28) finite equilibrium configurations are predicted for all masses. On these accounts, it came to be accepted that the white dwarfs, rather the « black dwarfs » in Fowler's terminology, represent the last stages in the evolution of all stars. And it seemed for a time that one could rest with the comfortable assurance that «all stars have the necessary energy to cool». But this assurance was shattered when it was realized soon afterwards (CHANDEASEKHAE [7, 8]) that the electrons in the centers of degenerate configurations following the mass-radius relation (27) begin to have momenta comparable t o mc; and that therefore the equation of state for degenerate matter must be modified to take into account effects arising from special relativity. The required modifications of eqs. (23) and (24) allowing for the effects of special relativity are readily made. We have only to insert in eqs. (20) and (21) the relations (29)
v,=
. P2I , ... m(l -{- p2jm2c2)* n
and
T, = mc^l+p 2 /^ 3 )*-!]
in place of the nonrelativistic relations (22). We find that the resulting equation of state can be expressed, parametrically, in the form (OHANDEASEKHAE [9,10]) (30)
P = Af(x),
Q = BX3 ,
W H T AKE THE STABS AS THET ARE?
9
where 7imAcs (31)
A =
87imac3fj,eH B
^h^'
=
^
and (32)
f(x) = x{x- +lf
(2x2 - 3) + 3 sinh" 1 x .
According to eqs. (31) and (32), the pressure approximates the relation (23) for low enough electron concentrations, but for increasing electron concentrations it tends to (cf. CHANDEASEKHAE [7])
(33)
*=*(!!)**»*•
This limiting form of the relation can be obtained by simply setting v„ = c in eq. (20); then ,o,*
-r,
34
F
8JTC f 3
3h r =^JP^P
( )
,
2>JIC 3
.
= Sh^PO;
and the elimination of p0 with the aid of eq. (19) directly leads to the relation (33). While the modification of the equation of state required by special relativity appears harmless enough, it has a dramatic effect on the predicted mass-radius relation for degenerate configurations. The relation between P and Q corresponding to the limiting form (33) is (35)
P = K2oi,
where
Kt = 8 \7t/
U-V-^ (fit H)i'
In this limit the configuration is therefore a polytrope of index 3. I t is well known that, when the index is 3, the mass is uniquely determined by the constant of proportionality in the pressure-density relation. The theory gives (36)
M^
= 4 , ( § ) * 6 . 8 9 = 0.197
ffl^
= 5.76*-© .
(In eq. (36), 6.89 is a numerical constant derived from the explicit solution of the appropriate Emden's equation.) We observe that eq. (36) isolates once again the same combination of natural constants that we encountered earlier in sect. 2. A detailed consideration of equilibrium configurations (CHANDEASEKHAE [8-10]) built on the equation of state given by eqs. (30) and (32) shows that the mass-radius relation for these objects, while it approximates eq. (27)
S. CHANDKASEKHAR
10
for J f - ^ 0 , predicts t h a t the radius tends to zero for M-> MUmlt (see fig. 1). Therefore, finite degenerate equilibrium configurations exist only for M < Miimit; and stars more massive than MUmit «do not have the necessary energy to cool».
5.0
4.0
3.0
20
1.0
0
0.2
0.4
0.6
0.8
1.0
M/M3
Fig. 1. — The full-line curve represents the exact (mass-radius) relation for completely degenerate configurations. The mass, along the abscissa, is measured in units of the limiting mass (denoted by Ms) and the radius, along the ordinate, is measured in the unit l1= I.IZ/J,'1 -10 s cm. The dashed curve represents the relation that follows from the equation of state (26); at the point B along this curve, the threshold momentum p0 of the electrons at the center of the configuration is exactly equal to mo. Along the exact curve, at the point where a full circle (with no shaded part) is drawn, p0 (at the center) is again equal to mc; the shaded parts of the other circles represent the regions in these configurations where the electrons may be considered to be relativistic (p 0 >™c). (This illustration is reproduced from CHANDKASBKHAR [10].)
5. — A criterion when stars can develop degenerate cores. Once t h e existence of t h e limiting mass w a s established, t h e question t h a t required resolution w a s h o w o n e w a s t o r e l a t e i t s existence t o t h e evolution of stars from t h e i r gaseous s t a t e . If a s t a r h a s a mass less t h a n Mlimit, t h e a s s u m p t i o n t h a t i t will eventually evolve t o w a r d s t h e completely degenerate s t a t e a n d become a black dwarf a p p e a r s e m i n e n t l y reasonable. B u t , w h a t if its mass is larger t h a n JJfllmit? Clues as t o w h a t m i g h t t h e n h a p p e n were sought in t e r m s of t h e e q u a t i o n s a n d inequalities of sect. 2 a n d 4 (OHANDRAS B K H A E [11]).
T h e first question t h a t h a d t o b e resolved w a s t h e conditions u n d e r which a star, initially gaseous, c a n develop a d e g e n e r a t e core. A k e y t o t h e solution of this question was provided b y a comparison of e q . (35) w i t h a n expression for t h e electron pressure (as given b y t h e classical perfect-gas e q u a t i o n of s t a t e )
11
WHY ABE THE STARS AS THEY A R E ?
expressed in terms of a parameter /?e and Q as in eq. (5). Thus, with the definition
(where pe now denotes the electron pressure), we can write
Comparing this equation with eq. (35), we conclude that, if
(39)
[fctf) a -jr]
>K
>=8 0 ^m -
the pressure pe given by the classical perfect-gas equation is greater than that given by the equation if degeneracy prevailed—not only for the prescribed Q and T, but also for all g and T having the same /?„. Inserting for a its value (9), we find that the inequality (39) reduces to (CHANDRASEKHAR [11])
or (41)
l - f t > 0.0921 = 1 - 0 *
(say).
For our present purposes, the principal content of the inequality (41) is the criterion that, for a star to develop degeneracy, it is necessary that the radiation pressure be less than 9.2 percent of the total pressure. As we shall presently show, this requirement effectively excludes massive stars. This last inference is so central to all current schemes of stellar evolution that the directness and the simplicity of the early arguments is perhaps worth repeating. The two principal elements of the early arguments were these: first, that radiation pressure becomes increasingly dominant as the mass of a star increases; and, second, that electron degeneracy is possible only so long as radiation pressure is not a significant factor—indeed, as we have- seen, it must not exceed 9.2 percent of the total pressure. The second of these elements is a direct and an elementary consequence of the physics of degeneracy; but the first requires some amplification. That radiation pressure must play an increasingly dominant role as the mass of a star increases is one of the earliest results established in the study of stellar structure; it is, in fact, implicit in the calculation of Eddington's cloud-bound physicist. A quantitative expression to this fact was given by Eddington's standard model, which lay at the base of many of the early studies
S. CHANDRASEKHAR
12
in stellar structure during the twenties and thirties. In this model, the fraction /? ( = gas pressure/total pressure) introduced in sect. 2 is assumed to be a constant through the star. I t follows from eq. (5) that, under this assumption, stars are polytropes of index 3 and that we have the relation (cf. eq. (36)) M = in((^\
(42)
6.89,
where C(/3) has the value given in eq. (5). Equation (42) provides a quartic equation for 0 similar to eq. (13) for /?*. (The values of 1 — /S and /? given in table I are in fact derived from eq. (42) with the unlikely value fx — 4.) Equation (42) gives the mass (43)
M = 0.197ft. ( ! ) * ^ = 0 . 2 2 8 ( ! ) f ^
= 6.65„-0
for a star in which /3 = /3ffl = 0.908. I n the standard model, then, stars with masses exceeding 6.65^-2 O will have radiation pressures that will exceed 9.2 percent of the total pressures. Consequently, stars with masses greater than 6.6^- 2 cannot, during the course of their evolution, develop degeneracy in their interiors; and, accordingly, an eventual white-dwarf state is impossible for them without a substantial ejection of mass. The standard model is, of course, only a model. Nevertheless, all of our experience with stellar models for normal stars has confirmed the general (qualitative) correctness of the inferences that were drawn on the basis of the standard model. In particular, the principal conclusion that radiation pressure plays an increasingly dominant role as the mass of the star increases has stood the test of time; and the corollary that stars with masses exceeding 7 or 8 © cannot develop degeneracy in their interiors has been sustained. These conclusions to which one arrived forty and more years ago appeared so convincing that confident assertions such as the following were made (CHANDEASEKHAE [9,11]):
Given an enclosure containing electrons and atomic nuclei (total charge zero), what happens if we go on compressing the material indefinitely? (1932) « The life history of a star of small mass must be essentially different from the life history of a star of large mass. For a star of small mass the natural white-dwarf stage is an initial step towards complete extinction. A star of large mass cannot pass into the white-dwarf stage and one is left speculating on other possibilities. » (1934) While EDDINGTON [12] would not accede to these conclusions, he, nevertheless, very clearly recognized that the existence of an upper limit to the mass of completely degenerate configurations, if accepted—he, himself, denied
WHY ARE THE STABS AS T H E T ARE?
13
its existence (*)—implied, inevitably, the occurrence of black holes as the end products of the evolution of massive stars. He thus stated: « The star apparently has to go on radiating and radiating and contracting and contracting until, I suppose, it gets down to a few kilometers radius when gravity becomes strong enough to hold the radiation and the star can at last find peace. I felt driven to the conclusion that this was almost a reductio ad absurdum of the relativistic degeneracy formula. Various accidents may intervene to save the star, but I want more protection than that. I think that there should be a law of Nature to prevent the star from behaving in this absurd way.» (1935) The principal conclusion of the foregoing discussion, namely that stars with masses exceeding a certain lower limit cannot develop degenerate cores during their evolution, implies, conversely, that there exist stars with masses in a range in which when they do develop degeneracy, they do so under conditions in which the degeneracy is extremely relativistic. Under these same conditions, the masses of the degenerate cores cannot exceed 5.76 -(0.5) 2 ©~ 1.4O; and, moreover, the density and the radius of the core will be extremely sensitive to the precise conditions—e.g. the exact density—when the degeneracy becomes incipient. I t is clear from these general considerations that the evolutionary sequences that stars, with masses in this critical range, will follow will manifest a great variety of possibilities. The detailed and specific calculations on stellar evolution and nucleosynthesis summarized by ARNETT in this volume (see p. 356) seem to confirm these early expectations—though the concrete calculations and results are, of course, quite beyond anything that could have been anticipated forty years ago.
6. — The minimum mass for gravitational collapse to be possible. Finally, there is one further conclusion, of a converse character, that can be drawn from the inequality (6) and the limiting form (33) of the degenerate equation of state. Combining them, we have the inequality
(44)
!G(f7t)*ef If* >P c = .£2ef .
(*) As did MILNE [13]; he wrote: « If the consequences of quantum mechanics contradict very obvious, much more immediate, considerations, then something must be wrong with the principles underlying the equation-of-state derivation. Kelvin's gravitational age-of-the-Sun calculation was perfectly sound; but it contradicted other considerations which had not then been realized. To me it is clear that matter cannot behave as you predict. ... A theory must not be used to compel belief. ... Eddington is nearly always wrong in his work in the long run, and I am quite prepared to believe that he is wrong here, in his details. But I hold by my general consideration » (1935).
S. CHANDBASEKHAR
14
This inequality, on simplification, gives ( C H A N D E A S E K H A E [14]) Q
/Jir\ f
1
T h e m e a n i n g of this i n e q u a l i t y is t h i s : for a s t a r w i t h a mass less t h a n 1.74/a~2 O , t h e r i g h t - h a n d side of t h e i n e q u a l i t y (6) t o g e t h e r with t h e exact e q u a t i o n of s t a t e (30) will enable us t o set a n upper limit to its c e n t r a l d e n s i t y ; a n d this u p p e r limit t o t h e c e n t r a l density, as it is necessarily derived from less t h a n e x t r e m e relativistic degeneracy, will b e in t h e r a n g e (10 5 -^10 9 ) g c m - 3 (except w h e n t h e m a s s is exactly e q u a l t o 1.74iu^"2 © ) . I n o t h e r words, stars w i t h masses less t h a n 1.74/^T 2 © c a n n o t b e e x p e c t e d t o gravitationally collapse to form n e u t r o n stars or black holes: t h e secular contraction t h e y will undergo d u r i n g their last phases of evolution will b e a r r e s t e d w h e n t h e y reach white-dwarf densities. I t would follow, t h e n , t h a t we can set a lower limit of 0 . 4 3 © for n e u t r o n stars a n d black holes t h a t can form u n d e r t h e p r e s e n t conditions in t h e astronomical universe. T h e various facets of stellar s t r u c t u r e a n d stellar evolution we h a v e considered do suggest t h a t «the stars are as they are, because {hclQf fl-2 ( ~ 2 9 . 2 © ) provides a correct measure for their masses ».
REFERENCES
Since this lecture is strictly an account of « considerations of forty and more years ago », there are no references to papers published after the middle thirties. And except in one or two minor instances, no account is taken of ideas developed since those years. [I] [2] [3] [4] [5] [6] [7] [8] [9] [10] [II] [12] [13] [14]
A. S. 59, S. A. R. S. S. S. S. S. S. A. E. 58, S.
S. EDDINGTON: The Internal Constitution of the Stars (Cambridge, 1926), p. 15. CHANDBASEKHAR: Mont. Not. Boy. Astr. Soc., 96, 644 (1936); also Observatory, 47 (1936). CHANDBASEKHAR: Nature, 139, 757 (1937). S. EDDINGTON: The Internal Constitution of the Stars (Cambridge, 1926), p. 172. H. FOWLEB: Mont. Not. Boy. Astr. Soc, 87, 114 (1926). CHANDBASEKHAR: Phil. Mag., 11, 592 (1931). CHANDEASEKHAE: Astrophys. Journ., 74, 81 (1931). CHANDBASEKHAR: Mont. Not. Boy. Astr. Soc, 91, 456 (1931). CHANDEASEKHAE: Observatory, 57, 373 (1934). CHANDEASEKHAE: Mont. Not. Boy. Astr. Soc, 95, 207 (1935). CHANDEASEKHAE: Zeits. Astrophys., 5, 321 (1932). S. EDDINGTON: Observatory, 58, 38 (1935). A. MILNE: from a personal letter dated February 24, 1935; also Observatory, 52 (1935). CHANDEASEKHAE: Observatory, 57, 93 (1934).
Black holes: the why and the wherefore Subrahmanyan Chandrasekhar
The theme of black holes has received wide publicity in the media and in popular scientific writings; one is left with the general impression that a black hole represents an object beyond imagination and whose place in the scheme of things passes belief. Even in serious scientific writings, descriptive words such as exotic are not u n c o m m o n ; and it is argued that belief in its occurrence in nature requires caution and scepticism. I do not share in these views. On the contrary, I should like to explain why, neither in its conception nor in its role in stellar evolution, the black hole requires any special or sophisticated reasoning. I Let m e start with the extreme simplicity of the reasoning that leads us to the notion of black holes. When a stone is thrown upwards, it will normally ascend to a certain height and will then fall back. The height to which the stone ascends before it falls back, will be higher, the higher the initial speed with which it is projected. And if it is projected with a sufficient speed, it will escape from the earth's gravity. Quite generally, the stronger the initial gravitational pull of the parent body, the faster a stone on it must be projected if it is to escape from the gravity of the larger body. It is one of the consequences of the general theory of relativity that light will also be similarly influenced by gravity. Normally, however, the effect is very small. But it is measurable: for example, the bending of light from a distant star, as it grazes the sun during an eclipse, has been measured. For this reason, one can ask whether a body like the sun could become sufficiently compressed that the force of gravity on its surface would become so strong that even light could not escape from it. If such circumstances could be realized, then the sun would not be visible from the outside
26
New horizons of human knowledge
since, for visibility, an object should be either self-luminescent (like the sun) or able to reflect or scatter light incident on it (like the moon). Under such compression, the sun would have become a 'black hole'. It can be calculated that if the sun, with its present mass, could be shrunk from its present radius of 700,000 km to a radius of 3.75 km, then light emitted tangentially from its surface would go around and around the sun. If the sun were to shrink still further, to a radius of, say, 2.5 km, then n o light it emitted could emerge. As a black hole, it would capture everything that might be thrown into it; and it would not allow anything to escape. But this does not mean that a black hole can have no influence on external objects: it can gravitationally attract external bodies like any other massive object. The radius that a star of mass M must have, in order that it may not be visible from the outside, is called its gravitational or Schwarzschild radius; it is given by 2GM/e2 where G is the constant of gravitation and c is the velocity of light. For a mass equal to that of the sun, this radius is 2.5 km. The argument I have set out for the conditions under which a black hole may occur are by no means novel: it was argued in exactly the same way by Laplace as long ago as 1798. The notion of a black hole depends only on the present observationally verified fact that the propagation of light is influenced by gravity. But it should be emphasized that the notion that light may be affected by gravity is beyond the realm of Newtonian ideas. The phenomenon of light deflection finds a natural explanation only in the framework of the general theory of relativity. And it is to the general theory of relativity that we must turn for a scientifically precise description of black holes. I shall return to this matter later. First I want to consider the question whether objects as massive as the sun can ever shrink to the required dimensions. Can they, in fact, become so small that they will disappear from view and become 'black holes' in the sky ? Current developments relating to the evolution of stars answer both these questions in the affirmative; and it appears that black holes of the kind predicted by theory do in fact occur in nature. The theoretical and observational grounds for these expectations are derived from considerations relating to the final stages in the evolution of stars.
II The radiation of a star is derived from nuclear processes taking place in its interior. In the case of the sun, the process consists in the burning of hydrogen into helium; and at its present rate of radiation, the consumption of hydrogen is so small that the sun can continue to radiate for a time in the order of 10 u years. On the other hand, a star which is ten times as massive as the s u n — a n d there are many such stars—is known to radiate energy at a rate which is about
Black holes: the why and the wherefore
27
10,000 times greater. Consequently, it will exhaust its internal sources of energy in a relatively short time, in the order of 10' years. It is estimated that the galaxy has existed for a period of something like 20 billion years. Therefore, the question as to what happens to a star in the galaxy which exhausts its energy sources in 10 million years becomes an important and a relevant one. As a star exhausts its source of energy it will begin to contract. The density increases as contraction proceeds. When the star has reached a density of about 106 grams per cc, there is a well-defined state into which it can settle down if its mass is less than 1.4 times that of the sun. This is the state which one observes in the so-called white dwarfs (the companion of Sirius is an example). White dwarfs are stars which are very faint and not very massive but which do have densities of the order of 106 grams per cm 3 . The important proviso here is that in order for a star to have the possibility of settling down into such a state, its mass must be less than the limit stated above. Since the existence of this limiting mass is central to much of the subsequent considerations, let m e explain its origin. Ill It is natural that the physical conditions, such as density and temperature, in the interiors of stars should vitally depend on the equation of state which relates the pressure P, the density p, and the temperature T. In normal stars, such as the sun, the relation is the familiar one: P= Rp T
{R = gas constant),
i.e. the pressure increases linearly with the density, a constant of proportionality which is determined, apart from the factor R, by the temperature. Now, suppose that we increase the density indefinitely maintaining the same temperature. At first, it will follow the foregoing linear relation. But gradually a departure will set in and the pressure will begin to increase more rapidly following a p 3/3 -relation. And as the density increases still further, the effects of special relativity intervene and the relation 'flattens' to a p 4/3 -relation. The remarkable fact is that these subsequent parts of the pressure-density relation are independent of the temperature: the relation is the same for all initial temperatures (see Figure 1). This limiting form of the equation of state is described as the fully degenerate state. In this state all the available energy levels for the electrons, up to a certain threshold, are occupied while all the levels above the threshold are left unoccupied. A natural question, now, concerns the structure of a massive configuration, in equilibrium under its own gravitation, when the equation of state is that of fully degenerate matter. It is found that for such configurations, the radius initially
New horizons of human knowledge
FIG. 1. The variation of pressure with density for various initially assigned temperatures.
decreases inversely to its 1 /3-power as the mass increases: MCCR ~1/3. But very soon, by virtue of the later p 4/3 -variation, the radius decreases very rapidly as the mass continues to increase. In fact the radius tends to zero when a certain limiting mass is reached (see Figure 2). It is found that these theoretically constructed configurations have densities in the range 106 to 10s grams per cm 3 . We conclude that these configurations account for the white dwarf stars; and further, that such finite states are not possible when the mass of the star is in excess of the limiting mass, namely 1.4 O. In other words, when a star with a mass in excess of this limit contracts as its source of energy is exhausted, its contraction cannot be arrested at the white dwarf stage, i.e. when its radius is a few thousand kilometres and its density is in the range of 106 to 109 g/cm 3 . Accordingly the more massive stars must contract further.
IV The next stage where the contraction could be arrested would be when the atomic nuclei are so tightly packed that the density of matter becomes comparable to
Black holes: the why and the wherefore
29
\H
*v
^
v"
o>
o* o> o> ©*> ©•* ^
o» o^ ^
# <
FIG. 2. T h e full-line curve represents the exact (mass—radius) relation for completely degenerate configurations. T h e mass, along the abscissa, is measured in units of the limiting mass (denoted by M3) and the radius, along the ordinate, is measured in the unit lx = 7.72fi-' X 108 cm. T h e dashed curve represents the relation that follows from the equations of state; at the point B along this curve, the threshold m o m e n t u m p0 of the electrons at the centre of the configuration is exactly equal to mc. Along the exact curve, at the point where a full circle (with no shaded part) is drawn, p0 (at the centre) is again equal to mc; the shaded parts of the other circles represent the regions in these configurations where the electrons may be considered to be relativistic (p0 ^> mc). (Reproduced from S. Chandrasekhar, Mon. Not. Roy. Astron. Soc, vol. 95, no. 219, 1935.)
that which exists inside the atomic nucleus. That density is not 10 6 g/cm 3 , but a million million g/cm s . The question arises: can a star settle down when it reaches nuclear density of 1012 to 1013 g / c m 3 — t h e density represented in the so-called neutron stars? That depends on the star's mass; stable neutron stars can exist only for a restricted range of masses. But why neutron stars ? And why should their masses be restricted to a limited range? Let m e answer these questions.
30
New horizons of human knowledge
Under normal conditions, the neutron is ^-active and unstable while the proton is a stable nucleon. But if in the environment in which the neutron finds itself (as it will in the centre of degenerate configurations near the limiting mass), all the electron states with energies less than or equal to the maximum energy of the y-ray spectrum of the neutron are occupied, then Pauli's principle will prevent the decay of the neutron. In these circumstances the proton will be unstable and the neutron will be stable. At these high densities, the equilibrium that will obtain will be one in which, consistent with charge neutrality, there will be just exactly the right number of electrons, protons, and neutrons with appropriate threshold energies so that none of the existing protons or neutrons decay. At these densities the neutrons will begin to outnumber the protons and electrons by large factors. In any event it is clear that once neutrons begin to form, the configuration will essentially collapse to such small dimensions that the mean density will approach that of nuclear matter and in the range 1013 to 1015 g/cm 3 . These are the neutron stars that were first studied by Oppenheimer and Volkoff in 1939, though their possible occurrence had been suggested by Zwicky some five years earlier. The second question relates to the statement I made that stable neutron stars can only exist for a restricted range of masses. The reason lies in certain fundamental requirements of the general theory of relativity, which is essential for determining the structure of stars whose radii are in the order of a few kilometres. There is also a fundamental theorem which goes back to Karl Schwarzschild who showed, in a paper published in 1917, that configurations in stable hydrostatic equilibrium in general relativity must have radii in excess of 9/8 of the Schwarzschild radius R,= 2 GMIci which, as I have already stated, determines the maximum radius for a mass to become a black hole. And the argument can be stated as follows. It is well known that, in the framework of the Newtonian theory, the condition for the dynamic instability of a star, derived from radial perturbations, is that the effective ratio y of the specific heats, or, more precisely some average of it, is less than 4 / 3 ; and dynamic stability is guaranteed if y, or, some average of it, is in excess of 4 / 3 . But this result is changed in the framework of general relativity. A star with a ratio of specific heats y, n o matter how high, will become unstable if its radius falls below a certain determinate multiple of the Schwarzschild radius. It is this fact which accounts for the existence of a maximum mass for a neutron star, to which I referred earlier; and, indeed, for the instability of all equilibrium configurations as they approach the value 9/8 of the Schwarzschild radius. Calculations show that, while the permissible range of masses for stable neutron stars is subject to uncertainties, the range is narrow: the current estimate is between 0.3 to 1.5 times the solar mass.
Black holes: the why and the wherefore
31
The principal conclusions that follow from the foregoing considerations can be summarized as follows. Massive stars in the course of their evolution must collapse to dimensions in the order of 10 to 20 km once they have exhausted their nuclear source of energy. In this process of collapse, a substantial fraction of the mass may be ejected. If the mass ejected is such that what remains is in the permissible range of masses for stable neutron stars, then a neutron star will be formed. While the formation of a stable neutron star could be expected in some cases, it is clear that the formation of such a star is subject to vicissitudes. It is not in fact an a priori likely event that a star initially having a mass of, say, ten solar masses, ejects, during an explosion, subject to violent fluctuations, an amount of mass just sufficient to leave behind a residue in a specified narrow range of masses. It is more likely that the star ejects an amount of mass that is either too large or too small. In such cases the residue will not be able to settle into a finite state; and the process of collapse must continue indefinitely until the gravitational force becomes strong enough to hold the radiation. In other words, a black hole must form. V The problem we now consider is that of the gravitational collapse of a body to a volume so small that a trapped surface forms around it; as we have stated, n o light can emerge from such a surface. Let us first consider, in the framework of the Newtonian theory, what would happen to an incoherent mass, with no internal pressure, distributed with exact spherical symmetry about a centre. It would simply collapse to the centre in a finite time since in the absence of pressure there is nothing to restrain the action of gravity. In the Newtonian theory this result of matter collapsing to an infinite density may be considered as a reductio ad absurdum of the initial premises: a distribution of matter that is exactly spherically symmetric and the absence of any sustaining pressure are both untenable in practice. If either of these two premises is not exactly fulfilled, then the collapse to an infinite density will not happen. Now if the same problem of pressure-free collapse is considered in the framework of general relativity—as it first was by Oppenheimer and Snyder—one finds, as in the Newtonian theory, that the matter collapses to the centre in a finite time (as measured by a co-moving clock). But in contrast to the Newtonian theory, the inclusion of pressure or the allowance of departures from exact spherical symmetry do not seem to make any difference to the final result. The reasons are the following. In general relativity, pressure contributes to the inertial mass; and this contribution becomes comparable to the contribution by the mass density when the radius of the object approaches the Schwarzschild limit. On this account, after a certain stage, the allowance for pressure actually facilitates, rather than hinders, the collapse. It is also
32
New horizons of human knowledge
clear that small departures from spherical symmetry cannot matter. For unlike in the Newtonian theory, the aim to the centre need not be perfect: it will suffice to aim within the Schwarzschild radius. More generally, theorems by Penrose and Hawking show that, in the framework of general relativity, allowing for factors that derive from pressure and lack of spherical or other symmetry will not prevent the matter from collapsing to a singularity in space-time if a certain well-defined point of no r e t u r n — a n event horizon—is passed. If we consider then the gravitational collapse of a massive star and allow for all factors that derive from pressure but permit only small departures from spherical symmetry, the end result is the same as in a strictly spherically symmetric pressure-free collapse. There is no alternative to the matter collapsing to infinite density at a singularity once a point of no-return is passed. The reason is that once the event horizon is passed, all time-like trajectories must necessarily get to the singularity: 'All the King's horses and all the King's men' cannot prevent it. As far as the external observer is concerned, the energy associated with the departures from spherical symmetry will be radiated away as gravitational waves; and the event horizon will eventually settle down to a smooth spherical surface with an exterior Schwarzschild metric for a certain M — t h e mass of a black hole.
VI It is important to observe that the p h e n o m e n o n of spherical collapse will be described differently by an observer moving with the surface of the collapsing star and by an observer stationed at infinity. This difference is illustrated in Figure 3. Imagine that the observer on the surface of the collapsing star transmits time signals at equal intervals (by his clock) at some prescribed wavelength (by his standard). So long as the surface of the collapsing star has a radius that is large compared to the Schwarzschild radius, these signals will be received by the distant observer at intervals that he will judge as (very nearly) equally spaced. But as the collapse proceeds, the distant observer will judge that the signals are arriving at gradually lengthening intervals, as is the wavelength of his reception. As the stellar surface approaches the Schwarzschild limit the lengthening of the intervals, as well as the lengthening of the wavelength of his reception, will become exponential by his time. The distant observer will receive no signal after the collapsing surface has crossed the Schwarzschild surface; and there is no way for him to learn what happens to the collapsing star after it has receded inside the Schwarzschild surface. For the distant observer, the collapse to the Schwarzschild radius takes, strictly, an infinite time (by his clock) though the time scale in which he loses contact in the end is of the order of milliseconds.
singula
FIG. 3. Spherically symmetric collapse. At each point the future and the past light-cones are drawn; all time-like trajectories must lie within these cones. The reception by an observer, orbiting in a circular orbit at a large distance from the centre, of a light signal sent by an observer on the collapsing stellar surface is shown; it makes it clear why no signal sent after the surface passes inside the Schwarzschild surface at r = 2m can be received by the orbiting observer. Also, notice how all future directed time-like paths from any point inside r = 2m must necessarily intersect the singularity at r = 0. (Reproduced from R. Penrose, Nuovo Cimento, ser. I, vol. I, p. 252, 1969.)
34
New horizons of human knowledge
The story is quite different for the observer on the surface of the collapsing star. For him nothing unusual happens as he crosses the Schwarzschild surface: he will cross it smoothly and at a finite time by his clock. But once he is inside the Schwarzschild surface, he will be propelled inexorably towards the singularity: there is no way at all to retrace his steps. VII It is currently believed that the X r a y star Cyg X I is a binary system in which one of the components (the X-ray component) is a black hole. The reason for this belief is derived from the following two circumstances. First, the star exhibits a rapidly fluctuating intensity on a time scale of 50 milliseconds or less. From this fact we conclude that the X-ray-emitting regions must be 'compact', with a linear dimension less than 104 km. Second, from the observations relating to the binary nature of the system it has been deduced that the minimum mass of the compact star is six solar masses. Since this minimum of six solar masses is already considerably in excess both of the limiting mass for white dwarfs, and of all estimates of the upper limit for stable neutron stars, the conclusion is inevitable that we are dealing here with a black hole. VIII I now turn to the solutions for black holes which this general theory of velocity provides. A solution describing a stationary black hole must have the following properties. It must partition space into two regions: an inner region bounded by a smooth surface which is the envelope of null geodesies; and an outer region which becomes asymptotically flat, i.e. it becomes the familiar space-time of the special theory of relativity. The boundary surface separating the two regions defines the horizon of the black hole; and it is a necessary consequence of the definition that the space interior to the horizon is incommunicable to the space outside. It is a startling fact that with these simple and necessary restrictions on a solution to describe a black hole, the general theory of relativity allows only a single unique two-parameter family of solutions. This is the Kerr family of solutions in which the two parameters are the mass and the angular m o m e n t u m of the black hole. It includes Schwarzschild's solution as a limiting case appropriate for zero angular m o m e n t u m . Karl Schwarzschild derived his solution in December 1915 within a month of the publication of Einstein's series of four short papers outlining his theory. Schwarzschild sent his paper to Einstein for communication to the Berlin Academy.
Black holes: the why and the wherefore
35
In acknowledging the manuscript, Einstein wrote, 'I had not expected that the exact solution to the problem could be formulated. Your analytical treatment of the problem appears to me splendid.' Roy Kerr derived his solution in 1962. I would include this discovery of Kerr as among the most important astronomical discoveries of our time. It is, in my judgement, the only discovery in astronomy comparable to the discovery of an elementary particle in physics. IX I shall now briefly consider the nature of the space-times around black holes described by the Schwarzschild and the Kerr solutions. The best way to visualize them is to exhibit the 'light-cone structure' in the m a n n e r of Roger Penrose. Imagine that, at a point in space, a flash of light is emitted. Consider the position of the wave front of the emitted flash of light at a fixed short interval of time later. In field-free space, the wave front will be a sphere about the point of emission. But in a strong gravitational field, this will not be the case. The sphere will be distorted by the curvature of space-time about the point of emission. Figure 4 displays these wave fronts at various distances from the centre of symmetry of the Schwarzschild black hole. The section of the wave fronts is illustrated by a plane through the centre of symmetry. One observes that the sections of the wave fronts are circles far from the centre, as one would expect; they are, however, progressively displaced asymmetrically inward as one approaches the centre. And on the horizon the wave front is directed entirely inward towards the centre with the point of emission on the wave front—the wave front has become tangential to the horizon. This is clearly the reason why light emitted from the horizon of a black hole does not escape to infinity. The situation in the interior of the horizon is even more remarkable. The wave front does not include the point of emission: the wave front has detached itself. And since n o observer can travel with a speed faster than that of light, it follows that there can be no stationary observers within the horizon—the inexorable propulsion of every material particle towards the singularity at the centre cannot be avoided. Turning next to the geometry of the space-time in Kerr geometry, we illustrate in Figure 5 sections of the wave fronts of light emitted at various points on the equatorial plane of the Kerr black hole. The singularity in this case is a ring around the centre in the equatorial plane. In contrast to the Schwarzschild geometry, we have to distinguish, besides the horizon—where the wave front is entirely inside the horizon—a second surface where the wave front just manages to be attached to the source of emission. This second surface describes what has been called the
New horizons of human knowledge
36
o o FIG. 4. Effect of the curvature of space-time on the propagation of light from points in the neighbourhood of a Schwarzschild (non-rotating) black hole.
ergosphere. In the region between the ergosphere and the horizon, while the wave front has detached itself from the point of emission, it is still possible for a particle, with a sufficient velocity suitably directed, to escape to infinity. The importance of this intermediate region is that it is possible for a particle entering this region from infinity to break up in two in such a way that one of the fragments is absorbed by the black hole, while the other escapes to infinity with an energy which is in excess of that of the incident particle. This is the so-called Penrose process for extracting the rotational energy of the Kerr black hole. An analogous phenomenon occurs when electromagnetic or gravitational waves of sufficiently small frequencies are incident on the black hole in suitable directions. In these cases, the reflection coefficient for such incident waves exceeds unity and is called super-radiance.
Black holes: the why and the wherefore
FIG. 5. Equatorial cross-section of a Kerr (rotating) black hole. The positions of the wave fronts of light signals emitted at various points should be contrasted with those shown for the Schwarzschild black hole in Figure 4. The rotational energy of the Kerr black hole can be extracted by a particle (P0) that crosses the stationary limit from outside: the particle divides into two particles, one of which (P2) falls into the black hole while the other (P,) escapes from the ergosphere with more mass energy than the original particle (P0). X The possibility of extracting the rotational energy of a Kerr black hole, to which I have referred, leads to some further general considerations. By a careful analysis of the manner in which energy can be extracted by incident test particles, it has been shown that the energy extracted must satisfy an equality of the form: Change in mass = surface gravity X change in surface area + angular velocity of motion x change in the angular momentum.
38
New horizons of human knowledge
A more general result is that the surface area of the black hole must always increase. While this restriction on the change in the surface area consequent to the extraction of energy was first established with respect to the Kerr black hole, Hawking was able to prove that the same restriction must apply quite generally. In other words, the surface area of the event horizon (that is, the boundary of a black hole) has the property that it always increases when matter or radiation falls into the black hole. Moreover, if two black holes collide and coalesce to form a single black hole, then the area of the event horizon around the resulting black hole must also be greater than the sum of the areas of the event horizons around the original black holes. These properties suggest an analogy between the area of the event horizon of a black hole and entropy in thermodynamics. (Entropy can be regarded as a measure of a lack of detailed information resulting from the averaging over the inherent coarse-grained character of a macroscopic system.) The analogy between the laws governing black holes and the laws of thermodynamics has been extended by Bardeen, Carter and Hawking. They show that a small change in the surface area of a black hole is accompanied by a small change in its surface gravity even as in a thermodynamic system a small change in the entropy is accompanied by a small change in the temperature. This analogy would suggest that we may formally identify the surface gravity with temperature. The analogy is strengthened by the fact that the surface gravity is the same at all points of the event horizon just as the temperature is the same at all points in a body in thermal equilibrium. Although there is a similarity between entropy of a thermodynamic system and the surface area of the event horizon of a black hole and between the surface gravity on the event horizon and the temperature, it is not obvious what meaning, if any, one has to give to this analogy. These analogies remained a paradox until 1974 when Hawking was able to show, by applying the method of the quantum theory to the formation of black holes by gravitational collapse, that there is a steady emission of particles from the horizon with a pure thermal spectrum; and further, that the temperature of the thermal radiation so emitted increases rapidly as the mass of the black hole decreases. For a black hole with a mass equal to that of the sun, the temperature is only about a tenth of a millionth of a degree above absolute zero—the consequent rate of radiation is so insignificant that it can have no conceivable physical consequence. On the other hand, a black hole with a mass of a billion tons (which would have the size of a proton) would radiate at a temperature of some 120 billion degrees. At this temperature, the black hole would profusely emit pairs of electrons and positrons; and this emission would increase explosively as the mass of the black hole diminished. The discovery of this purely quantal production of particles at the horizons of black holes has initiated what appear to be the first steps in the unification of
Black holes: the why and the wherefore
39
relativity and quantum theory, a unification that has been sought after for nearly a half century. XI Again in connection with the extraction of energy from a Kerr black hole, a question of considerable significance but of a different sort arises. The Kerr solution represents a black hole with a smooth event horizon only so long as its angular m o m e n t u m is less than an amount determined by its mass. If the angular m o m e n t u m should exceed this limit then we should no longer have an event horizon which conceals the singularity in the interior. We would then have, as one says, a naked singularity, i.e. a singularity with which one can communicate from the outside. Under these circumstances, the world around us would be quite unlike what we have always thought it to be: predictable with respect to the future in terms of the conditions which obtain at the present. For these reasons, Roger Penrose has conjectured that: A system which evolves according to classical general relativity, with reasonable equations of state from generic non-singular initial data, does not develop any space-time singularity which is visible from infinity. This conjecture of Penrose is often referred to as the hypothesis of cosmic censorship. The question arises whether this hypothesis can be proved, or counterexamples to it be given. While no definite counter-examples have so far been given, a rigorous proof of its validity also remains to be given. XII I may, in concluding, summarize what I have tried to convey: that the concepts underlying the notion of black holes are simple and straightforward; that there are ample grounds for believing that they occur aplenty in the astronomical universe; that the general theory of relativity provides a remarkably simple description of the nature of the space-time around black holes; and, finally, that the study of black holes promises the basis for a unification of the general theory of relativity with the two other main disciplines of physics, namely, thermodynamics and quantum theory. I hope I have succeeded in conveying at least some of these impressions.
Proceedings of The Gibbs Symposium Yale University. May 15-17. 1989
How One May Explore the Physical Content of the General Theory of Relativity S. CHANDRASEKHAR 1. Introduction. The general theory of relativity, in its exact nonlinear form, must predict phenomena which have no counterparts in the weak Newtonian limit and are qualitatively different. As is well known, in the weak limit, the predictions relating to the deflection of light when traversing a gravitational field and to the consequent time delay, to the precession of the Kepler orbit as manifested by the orbit of Mercury, and to the changing period of a binary star in an eccentric orbit due to the emission of gravitational radiation, have been quantitatively confirmed. But all these effects relate to departures from the Newtonian theory by at most a few parts in a million and of no more than three or four parameters in a post-Newtonian expansion of the exact equations of the theory. My concern is different: it relates to the exact theory. An example will clarify my meaning. Consider Dirac's relativistic theory of the electron. In the first instance, one naturally asked if his equation will lead to the correct formula of Sommerfeld for the fine-structure separation of the spectral lines of one-electron atoms. It did. But the novel content of this theory is the prediction of the positron and how electron-positron pairs will be created by y-rays of sufficient energy interacting with matter. It is this prediction of anti-matter that is the central feature of the Dirac theory. In the same way, the general theory of relativity must predict phenomena that have no parallels in the Newtonian theory. And we ask what they may be. The very formulation of this equation requires, perhaps, some explanation. One did not, for example, ask how one might explore the physical content of quantum electrodynamics or any similar innovative theory in physics. Why is the general theory of relativity different? The reason is this: it has generally been the case, that when fundamentally new ideas are formulated in a physical context, the predictions following from those ideas are on the border of experimental possibilities, and often are well inside the scope of one's physical intuition. But this is not the case with general relativity. The possibility ©1990 American Mathematical Society 0-8218-0157-0 $1.00+ $.25 per page
227
228
S. CHANDRASEKHAR
of observing the strong effects of gravity are far beyond the scope of even one's imagination; they are a million or more times what we can experience or contemplate. Besides, we have no base for physical intuition to play its part since one cannot even guess the sort of circumstances one should envisage. For these reasons, an entirely different manner of exploration is called for. 2. The singularity theorem. The essential feature of the general theory of relativity that distinguishes it from the Newtonian theory is provided by the singularity theorems of Penrose and Hawking. Very roughly stated, what the theorems assert is the following: these exists a large class of physically realizable initial conditions which, when evolved according to the equations of general relativity, will necessarily lead to the development of singularities in space-time. Stated differently, unlike in the Newtonian theory, the development of singularities is generic to general relativity. This is a qualitative difference. It is illustrated by the following example. Consider an initial distribution of matter, devoid of any internal pressure (i.e., dust), that is perfectly spherically symmetric (see Figure 1). In the Newtonian theory, such an initial distribution of matter will gravitationally collapse to the center in a finite time—the time of free fall—and a singularity of infinite density will form at the center. But this development of a singularity in the Newtonian theory is not generic to the theory in the sense that the slightest departure from spherical symmetry or the slightest amount of internal pressure or rotation will prevent its occurrence. In other words, the minutest departure from the initial conditions stated will efface the singularity. But this is not the case with general relativity. While the initial conditions stated will lead to a singularity, departures from spherical symmetry and allowance for internal pressure or rotation, even of finite amounts, will not prevent the development of the singularity if, during the process of gravitational collapse, a point of no return is transgressed. (At the point of no return, a trapped surface—eventually to become the event horizon of a black hole—forms.) It is this consequence of the singularity theorems that assures us that black holes will form in nature as the result of the gravitational collapse of a star of mass greater than say five or six solar masses during the last stages of their evolution when they have exhausted their source of energy. There remains of course the question, whether during the gravitational collapse of a star, the point of no return will be transgressed. That is an astrophysical question with which I am not presently concerned; it suffices to say that an instability of general relativistic origin is ultimately the cause. The necessary formation of black holes, starting from a wide class of initial conditions, is an example of a phenomenon predicted by general relativity that has no parallel in the Newtonian theory. The question arises: are there other examples? Before I consider this question, it is necessary that I say a few things about the mathematical theory of black holes.
1341 PHYSICAL CONTENT OF RELATIVITY
229
1. A perfectly spherically symmetric distribution of matter, devoid of internal pressure, will gravitationally collapse to produce a singularity at the center in the frameworks both of Newtonian theory and of general relativity. FIGURE
3. On the black-hole solutions of general relativity. I shall restrict myself in the first instance to isolated black holes. A black hole isolated in space represents a space-time with the following properties: 1. It is an exact solution of the Einstein-vacuum or the Einstein-Maxwell equations. (I shall explain later why it is important to include the Einstein-Maxwell equations along with the Einstein-vacuum equations.) 2. The three-dimensional space is divided into two non-overlapping regions—an internal region and an external region—separated by a smooth convex two-dimensional null surface. 3. The null surface is an event horizon, meaning that the internal region is incommunicable to the external region. That is, no signal originating in the internal region can cross the boundary (the surface of separation) and emerge into the external region; and conversely, no time-like or null trajectory which originates in the external region and enters the internal region by crossing the boundary separating them can ever recross the boundary to emerge into the external region: it is lost forever. 4. The space-time is asymptotically flat. It is a remarkable fact that, consistent with the foregoing requirements, there are exactly four black-hole solutions. The black-hole solutions are of two kinds: static and stationary. The distinction is one of whether the black hole is nonrotating or rotating. Consider first the vacuum solutions. A solution that represents a static
230
S. CHANDRASEK.HAR
black hole is necessarily spherically symmetric; and it is given by the Schwarzschild solution. This is Israel's theorem. The Schwarzschild solution provides a one-parameter family of solutions. The parameter is the inertial mass, M, of the black hole. Israel's theorem is to the effect that given the mass of an isolated static black hole, the space-time is uniquely specified by the Schwarzschild solution. Similarly, a solution that represents a stationary black hole is necessarily axisymmetric; it is unique and it is given by the Kerr solution. This is Robinson's theorem. The Kerr solution provides a two-parameter family of solutions. The parameters are the mass, M, and the angular momentum per unit mass a(= J/M). Robinson's theorem is to the effect that given the mass, M, and the angular momentum, J, of the black hole, the space-time is uniquely specified by the Kerr solution. Turning next to the black-hole solutions of the Einstein-Maxwell equations, we have simple generalizations of the Schwarzschild and the Kerr solutions: the Reissner-Nordstrom and the Kerr-Newman solutions. The Reissner-Nordstrom solution describes a spherically symmetric, charged, static black hole. It provides a two-parameter family of solutions; the parameters are the mass, M, and the charge, Q(\Q\ < M) of the black hole. (The mass and the charge are measured in the units c = G = 1 .) The Reissner-Nordstrom black hole has an event horizon exactly as the Schwarzschild black hole. But unlike the Schwarzschild black hole, the Reissner-Nordstrom black hole has a second horizon inside the event horizon. It is called the Cauchy horizon since a time-like or a null trajectory can cross this horizon and escape into a world that is outside the domain of dependence of the external world. An observer who crosses the Cauchy horizon emancipates himself from his past. The two horizons coalesce when |(2| = M. And for \Q\ > M, the space-time has no horizon but the singularity at the center remains. Solutions with \Q\ > M are unphysical since the space-times exhibit naked singularities; and these are forbidden by Penrose's cosmic censorship hypothesis. It may be noted here that the Newtonian attraction and the Coulomb repulsion between two extreme ReissnerNordstrom black holes with charges of the same sign exactly balance. The Kerr-Newman solution represents a rotating charged black hole. It provides a three-parameter family; the parameters are the mass, M, the charge, Q, and the angular momentum per unit mass, a . Absence of naked 2
2
2
singularities require that M > Q + a . The mathematical theory of the Schwarzschild, the Reissner-Nordstrom, and the Kerr black holes is exceptionally complete and exceptionally rich. Thus, contrary to every prior expectation, all the standard equations of mathematical physics can be separated and solved in Kerr geometry. The equations include: Hamilton-Jacobi equations governing the geodesic motion of particles and polarized photons, Maxwell's equations, Dirac's relativistic equation of the electron, and the (linearized) equations governing the propagation and scattering of gravitational waves. The manner of separation
PHYSICAL CONTENT OF RELATIVITY
231
of Dirac's equation in particular has led to a fruitful re-examination of the century-old problem of the separability of the partial differential equations of mathematical physics. Returning to the black-hole solutions of general relativity, one can ask whether there are other solutions besides those we have described if we relax the requirement that they are isolated. For example, are there other static or stationary solutions that describe assemblages of the black holes? It has recently been shown by Rubak that, consistent with the other requirements of smoothness and asymptotic flatness, the only multiple black-hole solution that is allowed is that of Majumdar and Papapetrou that describes assemblages of extreme Reissner-Nordstrom black holes. I shall return to a more detailed consideration of this solution in §5. 4. The theory of black holes and the theory of colliding plane-waves patterned after it. In the general theory of relativity, one can construct planefronted gravitational waves confined between two parallel planes with a finite energy per unit area; and therefore, one can, in the limit, construct impulsive gravitational waves with a J-function energy profile. Parenthetically, it may be noted that one cannot, similarly, construct such impulsive waves in electromagnetic theory. For a ^-function profile of the energy will imply a square root of a J-function profile for the field variables; and the square root of a <5-function is simply not permissible—it is not an allowed mathematical or physical concept. In 1971, Khan and Penrose considered the collision of two impulsive plane-fronted gravitational waves with parallel polarizations. And they showed that the result of the collision is the development of a space-like singularity not unlike the singularity in the interior of black holes. This phenomenon, illustrated in Figures 2 and 3 (see p. 232), is not manifested in any linearized version of the theory: the occurrence of the singularity by a focussing of the colliding waves, in no way depends on their amplitudes. Clearly in this context (and in others, quite generally) nothing short of an exact solution of the problem will suffice to disclose a new phenomenon predicted by general relativity. In any event, the occurrence of a singularity in this example suggested to Penrose that here is a new realm in the physics of general relativity for exploration. However, there was no substantial progress in this area before one realized that the mathematical theory of black holes is, structurally, very closely related to the mathematical theory of colliding waves. This fact is, in itself, a matter for some surprise; one should not have thought that two theories dealing with such disparate circumstances will be as closely related as they are. Indeed, by developing the mathematical theory of colliding waves with a view to constructing a mathematical structure architecturally similar to the mathematical theory of black holes, one discovered a variety of new implications of the theory that simply could not have been anticipated.
232
S. CHANDRASEKHAR tt2+V2 = l
2. To illustrate the space-time resulting from the collision of impulsive gravitational waves. FIGURE
FIGURE 3. Illustrating the development of a singularity when two plane impulsive gravitational waves collide head on: following the collision the two waves scatter off each other and focus to form a singularity.
To describe in detail how the mathematical theory of colliding waves was fashioned in the manner described will require a descent into technical matters that is not appropriate to this occasion. I must restrict myself to an impressionistic description that is still precise enough not to obscure the solid mathematical base.
233
PHYSICAL CONTENT OF RELATIVITY
First, consider the base of the mathematical theory of black holes. The entire set of field equations that govern stationary space-times (and static space-times as special cases) can be reduced to the complex equations, (I -\E\2){[(1
- r , 2 ) E , n ) , ^ - [(I -
^)E,M],M}
=-2E\(\-t12){E,nf-{\-n1)(E,ii)2],
(1)
for the vacuum and for a special class of Einstein-Maxwell space-times, and (l-4<22-|£|2){[(l->72)£,„],„-[(!-H2)^],,} (2)
= -2E*[(\ - r]2)(E, ^f - (\ - M2)(E , ^f],
valid only for a particular class of Einstein-Maxwell space-times. These are the Ernst equations. (I postpone for the present, the consideration of an equivalent pair of X- and 7-equations.) In equations (1) and (2), E is a complex function of the two variables r\ and /i. In a gauge and a coordinate system adapted to black-hole solutions (but not restricted to them) r\ is a radial coordinate and p. — cos 6 , where 6 is a polar angle. Besides, rj = 1 defines the event horizon. In equation (2), Q is a real constant and (3)
\Q\ < \.
It can be shown that if Evac is a solution of equation (1) then (4) ^E.M.^vacVO-4^) is a solution of equation (2). Finally, it is important to observe that more than one combination of the metric functions and the associated electromagnetic potentials will satisfy the same Ernst equation. A particular context may require one choice while another context may require a different choice. And further, once the choice of the combination of the metric functions (and the associated electromagnetic potentials) and of a particular solution of the Ernst equation have been made, the complete specification of the space-time is relatively straightforward and requires no more than elementary quadratures. It is a remarkable fact that all the known black-hole solutions follow from the simplest solutions of equations (1) and (2), namely (5)
E = pr} + iqp,
of equation (1) and (6)
E = (pr, +
iqp)^J(\-4Q2)
of equation (2) where p and q are two real constants subject to the condition (7)
p1 +
q2=\.
234
S. CHANDRASEKHAR
The Schwarzschild and the Reissner-Nordstrdm solutions follow, for example, when (8)
E = t1 and
E = t]\J{l -4Q2)
(<7 = 0).
The charge of the black hole for the Reissner-Nordstrom solution is 2Q. Similarly the Kerr and the Kerr-Newman solutions follow from the general solutions (5) and (6). It is of course necessary that in deriving the blackhole solutions, we make the "right" choice of the combination of the metric functions and electromagnetic potentials that satisfies the Ernst equation. I now turn to the mathematical theory of colliding waves. Before I proceed further, I should make it clear that most of what I have to say from now on derives from work done in close collaboration with Valeria Ferrari of Rome and Basilis Xanthopoulos of Crete. I have been exceptionally fortunate in my collaboration with them. Space-times representing the collision of two plane-fronted waves are characterized by two space-like Killing vectors (dxi and dxi, that span the wave fronts), in contrast to stationary axisymmetric space-times which are characterized by one space-like (d ) and one time-like {dt) Killing vector. Nevertheless the equations governing the space-times of colliding gravitational and electromagnetic waves can be reduced to the same Ernst equations (1) and (2). But the coordinates r\ and // have different meanings: 0 < r\ < 1 is time-like and measures time (in an appropriate unit) from the instant of collision at r\ = 0, and - 1 < n < +1 is space-like and measures distance (again in an appropriate unit) normal to the wave fronts. It should be noted that the choice of the coordinates allows, a posteriori, for the fact that as a result of the collision, a curvature or a coordinate singularity will develop when t] — 1 and n = ± 1 . The origin of the structural similarity of the mathematical theory of black holes and of colliding waves stems from the circumstance that in both cases the Einstein and the Einstein-Maxwell equations are reducible to the same basic equations and, as we shall see, even to the same solution. The richness and the diversity of the physical situations that are described, in spite of this identity, results from the different combinations of the metric functions and the electromagnetic potentials which can be associated with the same solution of the Ernst equation. Thus, the fundamental solution of Khan and Penrose which describes the collision of two impulsive gravitational waves with parallel polarizations, is, in a well-defined mathematical sense, the analogue of the Schwarzschild solution: The Khan-Penrose and the Schwarzschild solutions follow from the solution E = rj of the Ernst equation but, of course, for different combinations of the metric functions. Similarly, the solution E = prj + iqfi leads to the Nutku-Halil solution which describes the more general case when the colliding impulsive waves have nonparallel polarizations (Chandrasekhar and Ferrari 1984). Thus, the Khan-Penrose and the Nutku-Halil solutions play
PHYSICAL CONTENT OF RELATIVITY
235
the same role in the theory of colliding waves as the Schwarzschild and the Kerr solutions play in the theory of black holes. Turning next to the collision of plane impulsive gravitational waves together with electromagnetic waves in the framework of the Einstein-Maxwell equations, one had to confront some conceptual difficulties. Penrose had raised the question: would an impulsive gravitational wave with its associated ^-function profile in the Weyl-tensor, imply a similar (5-function profile for the energy-momentum tensor of the Maxwell field? If that should happen, then the expression for the Maxwell-tensor would involve the square root of the (5-function; and "one would be at a loss to know how to interpret such a function." Besides, there was a formidable problem of satisfying the many conditions at the various null boundaries. On these accounts, all efforts to obtain a solution compatible with carefully formulated initial conditions failed. However, when it was realized that the Khan-Penrose and the Nutku-Halil solutions follow from the simplest solution of the Ernst equation it was natural to seek a solution of the Einstein-Maxwell equations which will reduce to the Nutku-Halil solution when the Maxwell field is switched off (the analogue of the Kerr-Newman solution). The problem is not a straightforward one since in the framework of the Einstein-Maxwell equations, we do not have an Ernst equation which reduces to the Ernst equation for the particular combination of the metric functions that is appropriate for the Nutku-Halil vacuum solution. The technical problems that are presented can be successfully overcome and a solution can be obtained which satisfies all the necessary boundary conditions and physical requirements (Chandrasekhar and Xanthopoulos, 1985; see Figure 4, p. 236). That we can obtain a physically consistent solution by this "inverted procedure" is a manifestation of the firm aesthetic base of the general theory of relativity. With the derivation of solutions for colliding waves that are the analogues of the known black-hole solutions, one might have thought that the theory of colliding waves, patterned after the theory of black holes, was completed. But that is not the case. By asking for solutions that follow from the same simplest solution of the Ernst equation (for space-times with two space-like Killing fields) for other allowed combinations of the metric functions and electromagnetic potentials, one was led to solutions describing space-times with entirely unexpected features. Thus, solutions for both the Einsteinvacuum and the Einstein-Maxwell equations were found (Chandrasekhar and Xanthopoulos, 1987) which violated the then commonly held belief that colliding waves must invariably lead to the development of curvature singularities. One found instead that event horizons formed; and further that the space-time, (extended beyond the horizons) included a domain which is a mirror image of the one that was left behind, and a further domain which included hyperbolic arc-like curvature singularities reminiscent of the Kerr and the Kerr-Newman black holes (see Figure 5, p. 237). It is remarkable that a space-time resulting from the collision of waves should bear such a close
236
S. CHANDRASEKHAR after the collision %
•
after the collision %~8(v) %~HM. <02~//(v) %, 4>0 continuous
-5(«)
-«(«). *o~W(«) %
4*! = *P3 = <»| = 0 everywhere
4. The space-time diagram for two colliding plane impulsive gravitational waves. The colliding waves are propagated along the null directions, u and v . The time coordinate is along the vertical, and the spatial direction of propagation (in space) is along the horizontal. The plane of the wavefronts (on which the geometry is invariant) is orthogonal to the plane of the diagram. The instant of the collision is at the origin of the (u, ?;)-coordinates. The flat portion of the space-time, prior to the arrival of either wave, is region IV. These waves produce a spray of gravitational and electromagnetic radiation which fills regions II and III; and the region of the space-time in which the waves scatter off each other and focus is region I. The result of the collision is the development of a curvature space-time singularity on u + v" = 1 ; and v — 1 or u = 1 are space-time singularities for observers who do not observe the collision. Singular behaviours in the Weyl scalars, *¥4, *¥2, and 4^ (as Dirac ^-functions or Heaviside (//) step-functions, or both, expressing the shock-wave character of the colliding electromagnetic waves) occur, as indicated, along the null boundaries separating the different regions. FIGURE
resemblance to Alice's anticipations with respect to the world Through the Looking Glass. "It [the passage in the Looking-Glass House] is very like our
237
PHYSICAL CONTENT OF RELATIVITY
5. An alternative representation of the space-time diagram, which achieves manifest symmetry between the region of the space-time before and after the formation of the null-surface, is provided by the transformation to the coordinates x = 1 - 2u and y = 1 - 2v . The regions marked I 0 , II, III, IV in this figure and in Figure 1 correspond. A permissible extension beyond the domain of dependence, is by simply attaching to region \e , regions II', III', and IV' isometric to regions II, III, and IV. By this C extension the two worlds joined together are mirror images of one another. FIGURE
passage as far as you can see, only you know it may be quite different on beyond" (see Figure 6, p. 238). In Table 1 (see pp. 239-241), we describe more fully the various solutions that have been derived for black holes and for colliding waves. The visual pattern manifested by this table is a reflection of the structural unity of the subject. The inner relationships between the theory of the black holes and the theory of colliding waves is equally manifest in the simpler context when the metric is diagonal. In this case the basic equations reduce to the twodimensional Laplacean equation, (9)
[(J - WgV),,,],^
[(I - H2)(IgV),lt],M
= 0.
This equation can of course be solved exactly; and the solutions that are relevant in the two theories are listed in Table 2 (see p. 242). In developing the theory of colliding waves in parallel with the theory of black holes, we have in effect systematically examined the consequences of
FIGURE
6. Alice crossing the 'event horizon' in Through the Looking G
TABLE 1. Solution for Ernst equation for Killing vectors
Field equations
E
" , • <>o
Einstein-vacuum
does not exist
i)t . i)ei
Einstein-vacuum
does not exist
£f
P
Solution
1
Schwarzschild
Kerr
PV + iQH: P2 + Q1 = 1
•>, - <>„
",
• >>o
Einstein-Maxwell
does not exist
Einstein-Maxwell
does not exist
>/\/i - 4<22
p>) + iq/t ; p2 + q2 = 1 -
0 ,.0 ,
Einstein-vacuum
'/
Reissner-Nordstrom
Kerr-Newman
4Q2
interchanges A and A
Khan and Penrose
TABLE 1. (continued)
Solution for Ernst equation for Killing vectors
'-*,' • 'V
Field equations Einstein-vacuum
Einstein-Maxwell
2
1
,
x
=\
does not exist
£ t
^
£
vac = (PI + >Q^ x\/l-4(2
' V • °x2
n
Einstein-vacuum
2
Einstein-vacuum
0 i,0
Einstein-Maxwell
2
does not exist
A
Nutku-Halil
2
and x
interchanges x and x
Chandrasekhar and Xanthopoulos
interchanges
Chandrasekhar and Xanthopoulos
1
x
0 i,0 :
Solution
interchanges
P1 + IQM: 2
p +q
<->.i . <) 1
E]
E]
£
j
2
and .V
pr\ + iqfi\ 2 2 P +Q = 1
interchanges A- and x
Chandrasekhar and Xanthopoulos
f\Ai-4<2 2 )
interchanges x and x~
Chandrasekhar and Xanthopoulos
TABLE 1. (continued)
Solution for Ernst equatio n for Killing vectors
Field equations
E
'V • 'V
Einstein-Maxwell
does not exist
0 i . <> :
Einstein-Maxwell
does not exist
''.1 • ' ' . :
Einstein-Maxwell
does not exist
<" = ^ac>
does not exist
P>l + iQM-
p2 + q2 = 1 - 4(2 2
Einstein-Maxwell
P<} + iqu: P' + 1 = 1 Ehlers transform of pi] + iqp
E'
Solution
interchanges [ T .v and x~
Chandrasekhar and Xanthopoulos
interchanges A and x"
Bell-Szekeres
interchanges A and x interchanges x and A
Bell-Szekeres Chandrasekhar and Xanthopoulos
interchanges A and x'
Chandrasekhar and Xanthopoulos
i " = *\ac>
,) , . i)
:
Einstein-hyrodynamics (« = p)
pi] + iqu
TABLE
2. Basic equation: [(1 - r]2)(\gV),,; ], (/ - [(1 - /T)(lg¥
Killing vectors
Field equations
Solution
«, • °o
Einstein vacuum
Ig*+ = l8^f
o,. yrt
Einstein vacuum
y
Einstein vacuum
X
v'yv:
.Y~
<J i . 0 i X
Einstein-Maxwell
X~
0 ,.0 , X
Einstein vacuum
Einstein-Maxwell
X~
(" =
£
vac>
lg*+ =
te^rr+VWW" igx = ig£*
Ig* =
igj^+VW^CO/) 2'
B
1-H
'<* 1 - 7
i l „B i ± / / = lgj^+X,A/>„(/;)/>„(>/) 2 ' l-H
Remarks Schwarzschild solution: sphe Distorted black holes (when Khan-Penrose solution for c polarizations Collision of impulsive gravit gravitational shock waves; pa Collision of impulsive gravit electromagnetic shock waves parallel polarizations; Bell-S Collision of impulsive gravit gravitational and electromag
243
PHYSICAL CONTENT OF RELATIVITY
adopting for the Ernst equation, in its various contexts, its simplest solution. While this approach may appear as an exceedingly formal one, it has nevertheless disclosed possibilities that one could not have, in any way, foreseen: the development of horizons and subsequent time-like singularities or the transformation of null dust into a perfect fluid. 5. Binary black-hole solutions. We have seen how the simplest solution of the complex Ernst equation provides almost all the fundamental solutions describing black holes and colliding plane impulsive gravitational waves. But there exists an alternative pair of real equations—the X- and F-equations— to which also the stationary axisymmetric vacuum solutions can be reduced. These equations are (10)
^(X + Y)[(AX,2),2+(SX,i),i]
A(X,2f+d(X,i)2
=
and (11)
l
-(X
+ Y)[(AY , 2 ) , 2 + (SY , 3 ) , 3 ] = A(Y , 2 ) 2 + 3(Y , 3 ) 2 ,
where A = t]2 - 1 and 3 = 1 -p.2.
(12) By the substitutions,
(13)
x =
—TandY
=
Tzrc'
equations (10) and (11) become (14)
(l-FG)[(AF,2),2+
(SF yi),3]
= -2G[A(F ,2f
+ S(F
^f]
and (15)
(l-FG)[(AG,2),2+(6G,i),i]
=
-2F[A(G,2)2+S(G^)2}.
These equations are real counterparts of the complex Ernst equation (1). And like the Ernst equation, equations (14) and (15) allow the simple solution, (16)
F = -pt) - qp
and
G = -pr\ + qp,
where p and q are two real positive constants subject to the condition, (17)
p2-q2=\.
It has always been my belief that the solutions for X and Y derived from the simplest solution (16) of equations (14) and (15) must provide, in a suitable context, a space-time of some real physical significance. But the X- and 7-equations have remained, like Cinderella, the ignored and neglected stepsister of the Ernst equation (see Figure 7, p. 244). And like Cinderella, the Xand F-equations have been rescued and restored. I should like to tell this tale
1356 244
S. CHANDRASEKHAR
FIGURE
7. X- and ^-functions as Cinderella being rescued.
PHYSICAL CONTENT OF RELATIVITY
245
since it illustrates how one may explore the physical content of the general theory of relativity. Like all tales that are told, I must start in a seemingly remote territory—in this instance the two-center problem. As E. T. Whittaker has written: "the most famous of the known soluble problems" in analytical dynamics, "other than of central motion, is the problem of two centres of gravitation i.e., the problem of determining the motion of a free particle in a plane, attracted by two fixed centres of Newtonian centres of force" (Analytical dynamics, Cambridge Univ. Press, 1937, p. 97). The integrability of the relevant equations of motion, in a plane containing the two centers, was discovered by Euler in 1760. The solution for the general three-dimensional motion was given by Jacobi in his 1842 lectures on dynamics by showing the separability of the Hamilton-Jacobi equation in prolate spheroidal coordinates. The underlying reason for the integrability of this problem is that any ellipse, with the two centers as foci, is an exact trajectory of a particle in the field of force of either of them in the absence of the other and is therefore by Bonnet's theorem—a theorem that could well have been known to Newton—an exact trajectory in the field of force of both of them. It has generally been true that problems that are central to the Newtonian theory play an equally central role in general relativity. On this account, it would appear worthwhile to consider the two-center problem in general relativity. But first we must resolve a conceptual difficulty in considering the problem consistently in general relativity and, strictly, in the Newtonian theory as well: the consideration of two fixed centers of attraction, without allowing for their mutual attraction, requires (to quote E. A. Milne in a different context) the "services of an archangel." However, in this instance, the services of an archangel can be dispensed with by attributing to each of the mass centers electric charges of the same sign so that the Coulomb repulsion between them exactly balances the Newtonian attraction. More generally, in the Newtonian theory we can envisage a static arrangement of any number of mass points Mx, M2, ... , MN (say), at arbitrary locations, with charges Q\, Qi > • • • > QN ° f t n e s a m e sign such that Mj\fG = Qt (i = I , ... , N) (where G denotes the constant of gravitation) and the gravitational attraction between any pair of them is balanced by their Coulomb repulsion. It is a remarkable fact that this same static arrangement is allowed in general relativity in conformity with the Einstein-Maxwell equations. This is the solution of Majumdar and Papapetrou. The solution was correctly interpreted by Hartle and Hawking as representing an assemblage of extreme ReissnerNordstrom black holes. As we have stated earlier, the Majumdar-Papapetrou solution is the sole static (or stationary) multiple black-hole solution that is compatible with smoothness of the space-time external to the event horizons and asymptotic flatness; there is none other.
246
S. CHANDRASEKHAR
Since we are presently interested in the two-center problem, we shall consider the special case of the Majumdar-Papapetrou solution when only two extreme Reissner-Nordstrom black holes are present and the space-time is axisymmetric about the line joining the centers of the two black holes. With the two-center problem in view, we shall, following Jacobi, write the metric in prolate spheroidal coordinates: x — sinh if/ sin 6 cos
x = fi = cos 6 ,
x =
Fn-, = B -, and
F
03 =
B
,r
And it can be shown that the metric of a static Einstein-Maxwell space-time in the coordinates chosen can be written in the form (21)
ds2 = e2'\dt)2
-2v
{dtlf e [r\ - f t ) n2-i f,
2
2.
, (dn)2 i-f
+(>?2-i)(i-/i2)(^)2 where / and e" are the sole metric functions to be determined. A specially simple class of solutions is obtained when (22)
f = 0,
B = ±e",
and e~" is a solution of the three-dimensional Laplace's equation, (23)
[(r12-\)(e-'/),2],2+[(\-fi2)(e-"),,],i=0.
The Majumdar-Papapetrou solution of two extreme Reissner-Nordstrom black holes located on the z-axis follows from the solution A/, M. (24) *1 '+ ' ' +• ' rj-n tf + n where Mx and M2 denote the masses of the two black holes. Their charges, both of the same sign, are Qx = ±MX and (7, = ±M2. The two black holes are located at the coordinate singularities rj = 0 and fi = ±1 . These are not points. They represent smooth two-dimensional surfaces with areas 4nMl
247
PHYSICAL CONTENT OF RELATIVITY
and 4nM2 consistently with the fact that t] — 0 and fi = ±1 represent the event horizons of the two extreme Reissner-Nordstrom black holes of masses Ml and M2 and charges Q{ and Q2 (of the same sign) equal to M[ and M2 respectively. The fact that, for the basic solution, representing the static placement of two extreme Reissner-Nordstrom black holes on the axis of symmetry, / = 0 confirms the fundamental correctness of writing the metric of static axisymmetric Einstein-Maxwell space-times in the form (21). And in view of the special solutions (21) which the Einstein-Maxwell equations allow, the substitutions, (25)
X = e" + B
and
Y = e" - 5,
would appear apposite. By these substitutions, we find that X and Y, in fact, satisfy equations (10) and (11). In other words, the three-dimensional Laplacean equation, (23), which underlies the solution representing the static placement of two extreme Reissner-Nordstrom black holes, is none other than what the X- and Y-equations become when one or the other is zero. As we have stated, the X- and 7-equations were first derived in the context of stationary axisymmetric vacuum space-times. The metric of such spacetimes, in a gauge and in a coordinate system adapted to black hole solutions (but not restricted to them) can be written in the form, ds2
X(dtf
--(dtp-codt)2 -e v2+f
(26)
vW-i)(i-A2)]
(dri)2 , {duf
n2-\
+
\-f
\f(ri2-\),
where x , w, and /i2 + n3 are functions n and n . With the definitions (27)
X = x + o) and
Y = * - co,
we find that X and Y satisfy the same equations (10) and (11)! By virtue of this relation between static Einstein-Maxwell and stationary Einstein vacuum space-times, one establishes the following one-to-one correspondence between them. The correspondence (28)
ev^X,B++a>,
and / = e^^Wd/ir,2
- n2)4,
enables one to pass freely from a metric of the form (21) and an electrostatic potential B, appropriate as a solution of the Einstein-Maxwell equations to a metric of the form (26) appropriate as a solution of the stationary Einstein vacuum equations, and conversely. This one-to-one correspondence is a manifestation of a natural and a harmonious blending of Einstein's relativity and Maxwell's electrodynamics in a single unified structure. It reminds one of the magnificent hall built for
248
S. CHANDRASEKHAR
the five princes in the Indian epic, the Mahabharatta. Here is van Buitenen's translation of the description of the great hall in the Mahabharatta: The hall, which had solid golden pillars, great king, measured ten thousand cubits in circumference. Radiant and divine, it had a superb color like the fire, or the sun, or the moon. Made with the best materials, garlanded with gem-encrusted walls, gilded with precious stones and treasures, it was build well ... and possessed the matchless beauty that Maya imparted to it. Inside the hall Maya built a peerless lotus pond, covered with beryl leaves and lotuses with gem-studded stalks, filled with lilies and water plants and inhabited by many flocks of fowl. Blossoming lotuses embellished it, and turtles and fishes adorned it. Steps descended gently into it; the water was not muddy and it was plentiful in all seasons; and the pearl-drop flowers that covered it were stirred by a breeze ... and it was thick with precious stones and gems. One time the princely Dhartarastra came, in the middle hall, upon a crystal slab, and thinking it was water, the flustered prince raised his robe; ... Again, seeing a pond with crystalline water adorned with crystalline lotuses, he thought it was land and fell into the water . . . . He once tried a door which appeared to be open, and hurt his forehead; another time, thinking the door was closed, he shrank from the doorway.
And as the prince in the Mahabharatta, one realizes, when wandering through the great hall of general relativity, that what one had believed to be Einstein's hall, is in fact a corridor leading to Maxwell's hall; and when one is certain that one is examining the gems in Maxwell's hall, one has inadvertently slipped into Einstein's hall. "So matchless is the beauty" that Einstein has "imparted to it." To return to the X- and 7-equations. Since X or Y equal to zero provides the unique binary black-hole solution with a static space-time that is entirely smooth exterior to the horizons and asymptotically flat, it behooves us to consider the space-time that follows from the (next!) simplest solution of the X- and y-equations, namely, that derived from the solutions of F and G given in equation (16). We find that the metric of the space-time is given by (Chandrasekhar and Xanthopoulos, 1989) (29) 2 ds
=
(P2& + Q2S)2 [(l+/"7)2-<7V]2 ,,.a __ [(1 +pr])2 - < ? V ] 4 al
\ > r/i
,
S, 2 \2
[(1 +ptj)
2
-q
2X3
(dr,)2 2
r, -\
+
{duf
(1-V)
2,2
fi ]
(p2A + q2S)2
,w, 2 ri22-l)(l-n')(d
where (30)
A = r]2-l,
5=
\-n2,
p and q are real positive constants subject to the condition, (31)
p - q = 1 and a < q
PHYSICAL CONTENT OF RELATIVITY
249
is a real positive constant. The electrostatic potential is given by
,32)
2
»--
y
,,.
(1 +ptj) = q n An examination of the metric (29) shows that it represents two charged black holes located on the z-axis at z = +1 and z = - 1 and a string (in general) stretched along the entire z-axis. A string is a line in space-time along which a conical singularity (with no associated curvature singularity) occurs characterized by a deficit defined by the difference between In and the limiting ratio of the circumference to the proper radius of a small circle described normal to the line. The two black holes are of equal mass, (33)
M=l/p,
and opposite charge, (34)
Q = ±(p+l)/pq.
The conical singularity along the axis is characterized by the deficits (in units of In) (35)
d |r|<1 = 1 - a 4 / /
and <5|z|>1 = 1 - a
/p\
where, by the assumption, a < q , both are deficits (not excesses). We observe that when a = q , (36) *ixi<, = °; and the condition for local flatness is met for |z| < 1 , i.e., for the part of the z-axis joining the two black holes; but strings stretch to ±oo from the north pole of the one and the south pole of the other. The surface area, S , of the horizons of the two black holes (in units of An) is given by (37)
S=(\+pfla.
Except for the conical singularity on the axis, the space external to the horizons is smooth. In the manifold extended into the interior of the horizons, time-like singularities with two spatial dimensions occur. But the space-time exhibits no naked singularities and is asymptotically flat. An important feature of these black holes is provided by the inequality, (38)
\Q\-M
=
(\+p-q)/pq>0.
The restriction \Q\ < M for the Reissner-Nordstrom black holes is not applicable to these charged black holes. And finally, it can be shown that the surface gravity on the horizons of the two black holes vanishes identically. This fact suggests that these two black holes are in some sense generically related to the extreme ReissnerNordstrom black holes on the horizons of which the surface gravity also vanishes.
250
S. CHANDRASEKHAR
A comparison with the analogous equilibrium configurations of magnetic monopoles. The multiple black-hole solutions that have been found so far are of two kinds: first, we have the Majumdar-Papapetrou solution which allows a static assemblage of extreme Reissner-Nordstrom black holes, with charges of the same sign, at arbitrary locations, in which (speaking in Newtonian terms) the gravitational attraction between any pair of them is exactly balanced by the Coulomb repulsion between them; and second, we have the solution, that we have described, of two black holes of equal mass (M) and opposite charge (±Q, \Q\ > M) with strings attached to them. A special case of this second class of solution is when there is no string connecting the two black holes but strings stretching to +00 and -00 from the north pole of the one and the south pole of the other. The two classes of black-hole solutions that we have described are analogous to static configurations of magnetic monopoles that have been found. First, we have the Bogomil'nyi-Prasad-Sommerfeld monopoles (with twice the charge of the Dirac monopole) of which we can contemplate a static assemblage (with charges of the same sign) in which the magnetic (Coulomb) repulsion is balanced by the attraction derived from a scalar field. Second, we have the possibility of two Dirac monopoles, of opposite charge, held in place by a connecting string. These two classes of solutions are manifestly similar to the black-hole solutions we have described. It is a remarkable fact that possibilities, contemplated only in recent years at the quantal level, are inherent, already at the classical level, in general relativity. A question that occurs is: Could one conclude from the very natural way in which strings emerge in the binary black-hole solution, that strings are indeed predicted by the general theory of relativity? Concluding remark. The physical insights that we seem to have achieved by developing the mathematical theory of colliding waves deliberately patterned after the mathematical theory of black holes, and by seeking ad hoc the simplest solutions of the equations which simultaneously describe the static Einstein-Maxwell and the stationary Einstein-vacuum equations, suggest that one of the ways in which one may explore the physical content of the general theory of relativity is to allow one's sensibility to its aesthetic base guide in the formulation of problems with conviction in the harmonious coherence of its mathematical structure. REFERENCES
The nature of the lecture precludes giving a list of references in the conventional style. In the context of the different sections the reader may wish to consult the following references: §3 For the theory of black holes: S. Chandrasekhar, The mathematical theory of black holes. Clarendon Press, Oxford. 1983.
251
PHYSICAL CONTENT OF RELATIVITY
§4 The principal references for this section are: K. Khan and R. Penrose, Nature, London 229 (1971). 185. Y. Nutku and M. Halil, Phys. Rev. Lett. 39 (1977), 1379. S. Chandrasekhar and V. Ferrari, Proc. Roy. Soc. London Ser. A 396 (1984), 55-74. S. Chandrasekhar and B. C. Xanthopoulos, Proc. Roy. Soc. London Ser. A 398 (1985), 223-259. Proc. Roy. Soc. London Ser. A 408 (1986), 175-208.
For a more complete discussion along the same lines, see: S. Chandrasekhar, Truth and beauty : aesthetic motivations Chicago Press, Chicago, 1987, pp. 144-169.
in science.
University of
§5 The discussion in this section is based on: S. Chandrasekhar, Proc. Roy. Soc. London Ser. A 358 (1978), 405-420. , Proc. Roy. Soc. London Ser. A 423 (1989), 379-386. S. Chandrasekhar and B. C. Xanthopoulos, Proc. Roy. Soc. London Ser. A 423 (1989), 387-400. T H E UNIVERSITY OF CHICAGO T H E ENRICO FERMI INSTITUTE CHICAGO, ILLINOIS 60637
On Ramanujan
I cannot clearly say anything that will relate to Ramanujan as a mathematician, particularly in this company which includes, among others, Professors Richard Askey, Bruce Berndt, and George Andrews, who have devoted years to exploring and following his many trails. But I do share with Ramanujan the same cultural background in our early formative years: both of us originate in a common social background—he from Kumbakonam and I from Tanjore, both ancient centers of Tamil culture and not very far apart. Besides, Ramanujan's parents and my own grandparents lived in very similar social and financial circumstances. On this account I can probably visualize Ramanujan's background better than even my younger Indian colleagues of later generations. With this common background, I can perhaps throw some light on some conflicting statements that have been made about Ramanujan and 'God' by some of his Indian contemporaries. I refer here particularly to the colorful stories concerning Ramanujan's devotion to the Namakkal Goddess. Quite generally, it may be stated that among those who were brought up in South India during the first two decades of this century, there was (and probably still is) very little correlation between observance and belief. In particular, I can vouch from my own personal experience that some of the 'observances' that one followed were largely for the purposes of not offending the sensibilities of one's parents, relations, and friends. I can say a good deal on these matters, but I shall only state that I do not accept what has commonly been said and written about Ramanujan's religious beliefs. I corresponded with Hardy on this matter while he was preparing for his Harvard Lectures; and I am personally much more inclined to accept his view as expressed in a letter to me dated February 19, 1936. . . . And my own view is that, at bottom and to a first approximation, R. was (intellectually) as sound an infidel as Bertrand Russell or Littlewood.... One thing I am sure. R. was not in the least the 'inspired idiot' that some people seem to have thought him. On the contrary, he was (except for a
period when his mental equilibrium was definitely upset by illness) a very shrewd and sensible person: very individual, of course, and with a reasonable allowance of the minor eccentricities of genius, but fundamentally normal and sane. And this view of Hardy's is corroborated by K. Ananda Rao, himself a mathematician of distinction, who had been Hardy's student and Ramanujan's contemporary in Cambridge. Ananda Rao is well known a n d remembered for his contributions to the theory of Tauberian theorems, function-theory and the theory of Dirichlet series. He has written: In his nature he was simple, entirely free from affectation, with no trace whatever of his being self-conscious of his abilities. He was quite sociable, very polite and considerate to others. He was a man full of h u m o u r and a good conversationalist, and it was always interesting to listen to him. On occasions when I met him, we used to talk in homely Tamil. He could talk on many things besides m a t h e m a t i c s . . . . This view of Ananda Rao is not surprisingly the same as Hardy's. He has written, . . . the picture which I want to present to you is that of a man who had his peculiarities like other distinguished men, but a man in whose society one could take pleasure, with whom one could drink tea and discuss politics or mathematics. Let me now t u r n to the role of Ramanujan in the development of science in India during the early years of this century. Perhaps the best way I can give you a feeling for what Ramanujan meant to the young men going to schools and colleges during the period 1915-1930 is to recall for you the way in which I first learned of Ramanujan's name. It must have been a day in April 1920, when I was not quite ten years old, when my mother told me of an item in the newspaper of the day that a famous Indian mathematician, Ramanujan by name, had died the preceding day; and she told me further that Ramanujan had gone to England some years earlier, had collaborated with some famous English mathematicians, and that he had returned only very recently, and was well known internationally for what he had achieved. Though I had no idea at that time of what kind of a mathematician Ramanujan was, or indeed what scientific achievement meant, I can still recall the gladness I felt at the assurance that one brought u p under circumstances similar to my own, could have achieved what I could not grasp. I am sure that others were equally gladdened. I hope that it is not hard for you to imagine what the example of Ramanujan could have provided for young men
and women of those times, beginning to look at the world with increasingly different perceptions. The fact that Ramanujan's early years were spent in a scientifically sterile atmosphere, that his life in India was not without hardships, that under circumstances that appeared to most Indians as nothing short of miraculous, he had gone to Cambridge, supported by eminent mathematicians, and had returned to India with every assurance that he would be considered, in time, as one of the most original mathematicians of the century—these facts were enough— more than enough—for aspiring young Indian students to break their bonds of intellectual confinement and perhaps soar the way that Ramanujan had. It may be argued, perhaps with some justice, that this was a sentimental attitude: Ramanujan represents so extreme a fluctuation from the n o r m that his being born an Indian must be considered to a large extent as accidental. But to the Indians of the time, Ramanujan was not unique in the way we think of him today. He was one of others who had, during that same period, achieved, in their judgement, comparably in science and in other areas of h u m a n activity. Gandhi, Motilal and Jawaharlal Nehru, Rabindranath Tagore, J. C. Bose, C. V. Raman, M. N. Saha, S. N. Bose, and a host of others, were in the forefront of the then fermenting Indian scene. The twenties and the thirties were a period when young Indians were inspired for achievement and accomplishment by these men w h o m they saw among them. I do not wish to leave the impression that Ramanujan's influence was only in this very generalized sense. I think it is fair to say that almost all the mathematicians who reached distinction during the three or four decades following Ramanujan were directly or indirectly inspired by his example. But Ramanujan's n a m e inspired not only ambitious young men planning scientific careers; it also stimulated to action those with public concern. Let me give one example. When I was a student in Madras one of my classmates (who came from a very wealthy family) was one Alagappa Chettiar. We became good friends; but our lives diverged along different paths after 1930. In the years before and during the second world war, Alagappa Chettiar prospered as an entrepreneur and became a noted philanthropist. He was in fact knighted by the British government. During the late forties after the war, Sir Alagappa Chettiar (as he was then) wrote to me inquiring if it might be useful for him to found a mathematical institute in Madras named after Ramanujan. I enthusiastically supported the idea; and when I returned to India briefly in 1951, the Ramanujan Institute had been founded a few months earlier. Its first director, T. Vijayaraghavan, was one of the most talented among Hardy's former students; he died at a comparatively early age in 1955. C. T. Rajagopal, a student of Ananda Rao, took over the di-
rectorship from him. Already at that time the financial status of the Institute seemed shaky, since Alagappa Chettiar's fortune was melting away. In April 1957, when Alagappa Chettiar died, the fate of the Institute hung in the balance; Rajagopal wrote to me that the Institute 'will cease to exist on the first of next month,' whereupon I wrote to the Prime Minister (Jawaharlal Nehru), explaining the origin of the Institute and the seriousness of its condition. Nehru's prompt answer was refreshing: 'Even if you had not put in your strong recommendation in favour of the Ramanujan Institute of Mathematics, I would not have liked anything to happen which put an end to it. Now that you have also written to me on this subject, I shall keep in touch with this matter and I think I can assure you that the Institute will be carried on.' And it was; but haltingly and precariously for the next twelve years. It is at this Institute in Madras that Ramanujan's Centennial will be celebrated by an International Conference in December. There is very little more I can say. My own view, sixty-six years after my first knowing of his name, is that India and the Indian scientific community were exceptionally fortunate in having before them the example of Ramanujan. It is hopeless to try to emulate him. But he was there even as the Everest is there.
REMINISCENCES AND DISCOVERIES ON RAMANUJAN'S BUST*
by S. CHANDRASEKHAR, F.R.S. University of Chicago, Chicago, Illinois 60637, U.S.A. Ramanujan was elected a Fellow of the Society in 1918 and he died in 1920, slightly more than a year after his return to India in 1919. The story of how a bust of one who had died in 1920 came to be made 60 years later is of some interest; and I hope that it is proper to tell that story on this occasion. In his biographical notice for Ramanujan (Proceedings of the Royal Society A 99, xiii-xxix) G.H. Hardy wrote: It was his insight into algebraical formulae, transformation of infinite series, and so forth, that was most amazing. On this side most certainly I have never met his equal, and I can compare him only with Euler or Jacobi. He worked, far more than the majority of modern mathematicians, by induction from numerical examples: all his congruence properties of partitions for example were discovered in this way. But with his memory, his patience, and his power of calculation he combined a power of generalisation, a feeling for form, and a capacity for rapid modification of his hypotheses, that were often really startling, and made him, in his own peculiar field, without a rival in his day. In his lectures on 'Ramanujan' at the Harvard Tercentenary Conference of Arts and Sciences in 1936, Hardy reassessed what he had written in 1921 as follows: I do not think now that this extremely strong language is extravagant. It is possible that the great days of formulae are finished, and that Ramanujan ought to have been born 100 years ago; but he was by far the greatest formalist of his time. There have been a good many more important, and I suppose one must say greater, mathematicians than Ramanujan during the last fifty years, but no one who could stand up to him on his own ground. Playing the game of which he knew the rules, he could give any mathematician in the world fifteen. *
Remarks made by the author on the occasion of presenting the bust of Ramanujan to the Royal Society, on 11 May 1994.
[153]
In more recent times Ramanujan's reputation has been enhanced considerably. For example, the slight overtone of criticism implied in Hardy's remark 'the great days of formulae are finished' has been countered by 'the days of great formulae are not finished'. With respect to the fall and rise of Ramanujan's reputation, Bruce Berndt (the author of five volumes of 'Ramanujan's Notebooks') has compared him to Johann Sebastian Bach, who remained largely unknown for years after his death in 1750. For Bach, the big turnaround came on 11 March 1829 with Felix Mendelssohn's performance of the St Matthew Passion. For Ramanujan, the roughly analogous event was George Andrews's discovery of the Lost Notebook in 1976. And what is this 'Lost Notebook'? G.N. Watson had worked on Ramanujan's papers for many years before World War II. When Watson died in 1965, The Royal Society asked J.M. Whittaker (the son of E.T. Whittaker) to write Watson's Biographical Memoir. For that purpose, Whittaker had asked Mrs Watson whether he could examine the papers that Watson might have left in his study. There, in Watson's study, as Whittaker recalled, papers covered the floor of a fair sized room to a depth of about a foot, all jumbled together, and were to be incinerated in a few days. One could only make lucky dips [into the rubble] and, as Watson never threw away anything, the result might be a sheet of mathematics but more probably a receipted bill or a draft of his income tax return for 1923. By an extraordinary stroke of luck one of my dips brought up the Ramanujan material.
This 'material', of some 87 loose sheets, was part of a batch of papers Dewsbury (the Registrar at the University of Madras during the years 1909-1924) had sent to Hardy in 1923 and that had, somehow, wound up with Watson. After his 'lucky dip', Whittaker passed them on to Robert Rankin (Watson's successor in Birmingham), who in 1968 handed the 87 loose sheets mentioned, together with other unpublished material, to Trinity College, Cambridge. And there it lay in the Trinity archives without anyone's knowledge till George Andrews rescued it. George Andrews, who had worked on problems related to 'mock theta functions' the subject of Ramanujan's last letter to Hardy a few months before he died, came to Cambridge to explore if any related material was available in the Trinity archives. He was thrilled and excited by what he discovered in the 87 loose sheets deposited in the Archives by Rankin. Andrews told me, at a later time, that when he presented a paper on the 'Lost Notebook' at a meeting of the American Mathematical Society, Dr Olga Taussky-Todd, who was chairing the session, said 'The discovery of the 'Lost Notebook' is as sensational a discovery for the mathematicians as a complete draft of a tenth symphony of Beethoven would have been to the musicians'.
[154]
Let me continue with what Richard Askey has written in this connection: These pages [of the Lost Notebook] are not dated, but from internal evidence they were written late in Ramanujan's life, much of it in his last year. Two thirds of the pages deal with basic hypergeometric series and most of this work is significantly deeper than Ramanujan's earlier work on the same subject. Try to imagine the quality of Ramanujan's mind, one which drove him to work unceasingly while deathly ill, and one great enough to grow deeper while his body became weaker. I stand in awe of his accomplishments; understanding is beyond me. We would admire any mathematician whose life's work was half of what Ramanujan found in the last year of his life while he was dying. Some of Ramanujan's work has one quality which is shared by very little other work. Most mathematics, including some very good work, is predictable. Much of the rest seems inevitable after it is understood, and it would eventually be discovered by someone else. Little of Ramanujan's work seems predictable at first glance and after we understand it there is still a fairly large body of work about which it would be safe to predict that it would not be rediscovered by anyone who has lived in this century. Then there are some of the formulas Ramanujan found that no one understands or can prove. We will probably never understand how Ramanujan found them.
Let me conclude this part of my story with a quotation from Freeman Dyson: The wonderful thing about Ramanujan is that he discovered so much, and yet he left so much more in his garden for other people to discover. For forty-four years I have intermittently come back to Ramanujan's garden; and every time when I come back, I find fresh flowers blooming.
I now turn to the other strand of the story: how the bust came to be made. It begins with a letter of Hardy's to me. Hardy was to give a series of 12 lectures on subjects suggested by Ramanujan's life and work at the Harvard Tercentenary Conference of Arts and Sciences in the autumn of 1936. In the spring of that year, Hardy told me that the only photograph of Ramanujan available at that time was the one of him in cap and gown, 'which makes him look ridiculous.' And he asked me whether I would try to secure, on my next visit to India, a better photograph which he might include with the published version of his lectures. It happened that I was in India that same year from July to October. I knew that Mrs Ramanujan was living somewhere in South India, and I tried to find where, at first without success. On the day before my departure for England in October 1936, I traced Mrs Ramanujan to a house in Triplicane, Madras. I went to her house and found her living under extremely modest circumstances. I asked her if she had any photograph of Ramanujan which I might give to Hardy. She told me that the only one she had was the one in the passport which he had secured in London early in 1919. I asked her for the passport and found that the photograph was sufficiently good (even after 17 years) that one could make a good
[155]
negative and copies. It is this photograph that appears in Hardy s book, Ramanujan, Twelve Lectures on Subjects Suggested by His Life and Work (Cambridge University Press, 1940). It is of interest to recall Hardy's reaction to the photograph: 'He looks rather ill (and no doubt was very ill): but he looks all over the genius that he was.' The rest of the story as told by Richard Askey: The story of the thread from the 'Lost Notebook' to the bust is simple. Andrews has done a lot of very deep work trying to understand what Ramanujan discovered. Eventually the New York Times heard about it and interviewed him. The Hindu followed with a more extensive interview, and also published an interview with Ramanujan's widow, Janaki Ammal. She lamented the fact that a statue of Ramanujan had never been made, although one had been promised. Andrews sent me copies of these interviews, and after a couple of months my subconscious finally got through to my conscious mind and it was clear that a bust should be made. Since Janaki Ammal was 80, time was important,* so it was up to individuals rather than governments or societies, since institutions move slowly. My first reason for wanting a bust was simple: if Ramanujan's widow wanted one she should have it. That was the least we could do to show our appreciation of Ramanujan to someone who had been a great help to him. Later I realized there was a second reason, which Janaki Ammal must have realized all along. She knew Ramanujan, and while she did not understand his mathematics, she knew that he was one of the few whose work will last. As long as people do mathematics, some of Ramanujan's work will be appreciated. In Ramanujan's case a more permanent memorial is appropriate; one which can be appreciated by those who do not understand his mathematics should be added to the memorial Ramanujan made for himself with his work. F r o m the desire to the a c c o m p l i s h m e n t is a long way. (1) A likeness of Ramanujan had to be found. It was provided by the original of the photograph that is the frontispiece in Hardy's book on Ramanujan. (2) A sculptor who would accept the challenge of making a three-dimensional bust from a two-dimensional photograph. Askey found such a sculptor of distinction in Paul T. Granlund of Gustavus College, Saint Peter, Minnesota, U.S.A. (3) A minimum of four busts had to be commissioned. Askey obtained enough funds by donations from the international community of mathematicians for one bust to be presented to Mrs Ramanujan (which it was in the autumn of 1983). Askey acquired one; and my wife and I two. + That was how the bust that you see came to be made. And as Askey has said:
* t
Mrs Ramanujan died on 13 April 1994. She was 94. Since the original four an additional five have been acquired by other institutions, including the one in the Cambridge Mathematical Faculty Library.
[156]
While Granlund does not appreciate Ramanujan's mathematics as those of us who have studied it do, he studied Ramanujan's passport photo deeply, and the results show in the bust. He probably understands some things about Ramanujan that we do not. May I now present the bust of Srinivasa Ramanujan to President Sir Michael Atiyah for his gracious acceptance on behalf of the Royal Society.
[157]
On Reading Newton's Principia at Age Past Eighty
I first met Daulat Singh Kothari in the company of Meghnad Saha at the 1929 December Meeting of the Indian Science Congress in Allahabad. Later, we were Research students together in Cambridge (England) for two years (1930-32). Thereafter, our lives took different courses. But we maintained our friendship, meeting once in several years at his home or at the University in Delhi. I had wanted to see him during my last visit to Delhi in December 1992. Even though Kothari was at that time in retirement in Jaipur, he had agreed to come to Delhi. But illness frustrated his intention and my hopes; and we did not realize at that time that it was to be forever. What follows could very well have been the subject of our conversation, had we met. And I dedicate this essay to his memory. I My interest in the Principia was stimulated by an invitation in 1986 to contribute a paper at one of the many symposia that were being planned to celebrate, in the following year, the tercentenary of the publication of Newton's Philosophic Naturalis Principia Mathematica in 1687. In 1986, I had not more than turned the pages of Cajori's edition of Andrew Motte's English translation of the third 1722-edition of the Principia in 1729:1 found that studying the Principia required considerable concentration to unravel the continuous prose in which Newton had written the geometrical and the analytical equalities and in fact the entire mathematical treatment. Confronted, then, with the prospect of having to prepare a paper with some pretense to substance, I had two choices: either to select and abstract from the extant voluminous Newtoniana or confine myself solely to the Principia and personally assess the intellectual achievement that it is. I chose the latter course, recalling the motto that 'to read for oneself a play of Shakespeare is worth a heap of commentaries.' My study of the Principia, so begun, continued to smoulder till it became earnest enough to embark, a year ago on 'Newton's Principia for the Common Reader' (1995, Clarendon Press, Oxford). And I may add that during the course of my study, I did not find
a single Proposition from which I did not learn something new, some things that I did not know, or some things I should have known. II It is common knowledge that Newton proved Kepler's laws of planetary and satellite motions that bodies (idealized as point masses) describe an ellipse, a hyperbola, or a parabola about the focus on the assumption that they are subject to centripetal attraction inversely as the square of the distance. And further, that Newton derived his universal law of gravitation on that basis. That in truth is a caricature of how Newton did, in fact, arrive at his universal law. Instead of elaborating on this theme, I shall give a brief account of the manner in which Newton proved the second law of Kepler and in what mathematical context. Newton proved in Proposition X that under a law of centripetal attraction proportional to the distance, a particle will describe an ellipse about its centre; and in Proposition XI he proved that under a law of attraction inversely proportional to the square of the distance the particle will also describe an ellipse but about its focus. And he shows how Proposition XI can be deduced from Proposition X; and conversely. In my experience of giving lectures on this topic and talking to students and colleagues, I have rarely found one who had even thought that the two Propositions could be related. But 'Cut the cackle and come to the hosses!' Ill The theorem that Newton proves can be stated as follows: Under the action of a central force inversely as the square of the distance, a particle will describe an ellipse about its focus; and a particle with an equal constant of areas (i.e. angular momentum) will describe the same ellipse about its centre if the law of force is as the distance; and conversely. The theorem is proved in two steps (i) We start with the elementary kinematical fact that the normal acceleration (or C.F. for the component of centripetal force in that direction) that a particle describing a curvilinear orbit experiences is C.F. (in the direction PO) = v2/p,
(1)
where v denotes the velocity, p the radius of curvature, and PO the direction of the inward normal at P (see Figure 1). If this orbit should be described under a centripetal attraction directed towards S, then
Circle of contact.
Figl
C.F. (towards S) = C.F. (in the direction PO) x cosec e (2) p sin e where e is the angle of inclination of the direction of motion, PQ, to PS. On the other hand by the law of areas ('Kepler's first law' proved for the first time in Proposition I) in a short interval of time At, (vAt) X SQ = j A At,
(3)
where A is the constant of areas. Therefore
= _A_ -
A
(4)
2SQ ~ 2SP sin e' a relation proved in Corollary I of Proposition I. By combining equations (2) and (4) we obtain, C.F. (towards S ) =
4(SP) 2 p sin 3 e
(5)
So far no assumption has been made about the nature of the orbit. If we now assume that it is an ellipse, then by the geometry of the ellipse p sin 3 e = semi-latus rectum !&2 =
2
a
ij
lL>
(6)
1376
Fig 2
where a and b are the semi-axes of the ellipse. (If the reader does not know this relation, it will be useful for him to refresh his m e m o r y of what he must have learnt in school!) Therefore, in this case A 2
C.F. (towards S) =
2L(SPy
(7)
= Constant X
2
(SP)~ .
(ii) Now suppose that quite generally the same curve is described by particles with the same angular m o m e n t u m under two different laws of centripetal attraction towards two different centres S and C. Then it follows from equation (5) that (see Figure 2) (C.F.).owardsS _
(SUl
€c\
(C.F.)towardsc
\ s i n ej
(CP\
\SP
In Figure 2, RZ is the tangent and PO is the inward normal to the curve at P and e s and e c are the respective angles of inclination of PS and PC to the direction of motion ZPR. Now, consider the case when the orbit is an ellipse with centre C and focus S. Draw CE parallel to ZPR. Then e s = l-RPS = APEC, and €c = LRPC = ir - Z.ECP. Hence sin e c _ sin AECP _ £ P and equation (8) gives
(9)
1377
(C.F.) towards S (C.F.)towardsC
V EP\ ^P/
(Epy
\SPCP^/
(\\)
l
(cp)(spy
Again from the geometry of the ellipse (as the reader should be able to verify if he does not know it already), PE = a = the semi major axis of the ellipse.
(12)
Hence, l^'rJlowardsS
,
1
(C.F.)
(13)
towards C
From this relation it follows that if the centripetal attraction towards the focus is proportional to (SP)~2, then the centripetal attraction towards the centre is proportional to CP: and conversely Q.E.D.O) Could anything be simpler? I should not be surprised if none of my readers have known of the foregoing demonstration. Yet, they are included between pages 46 and 51 in the 1722 third edition of the Principia.
TV In continuation of Proposition X and XI Newton considers the more general question of when two different laws of centripetal attraction (and/or repulsion), particles with equal angular m o m e n t a will describe the same orbit. If we call two such laws as dual to one another, Newton established the following laws of force, with integral exponents, as dual (Table 1). Newton missed only one other pair (r ~4 and r ~7) which has been shown to be dual only recently. TABLE 1. Dual Laws of Centripetal Attraction attractive, r~2 and attractive, r attractive, r~2 and repulsive, r repulsive, r and repulsive, r repulsive, r'1 and attractive, r~z attractive, r~5 and attractive, r - 5 (The last two are self-dual).
(elliptical orbits); (hyperbolic orbits); (the conjugate branches of a hyperbola); (hyperbolic orbits with the centre of force at the focus of the conjugate hyperbola); (a circle with the centre of attraction on any point of the circumference).
V What I have described in §111 is only one example from Book I out of many. I could equally well have chosen: (1) the geometrical and analytical solutions of Kepler's equation (Lemma XXVIII and Proposition XXXI); (2) the formulation and solution of initial-value problems (Propositions XXXIX, XLI and LXIII); (3) the variation of the elements of a Kepler orbit caused by an external perturbation and the foundations of lunar theory; (4) a spherical distribution of matter attracts an external corpuscle as if its entire mass is concentrated at the center (Proposition LXXVI which has been described as a 'superb theorem' by J. L. Glaisher); (5) the discovery of the method of images, normally credited to Lord Kelvin (Proposition LXXXII); and many more. I am convinced that one's knowledge of the Physical Sciences is incomplete without a study of the Principia in the same way that one's knowledge of Literature is incomplete without a study of Shakespeare.
Newton and Michelangelo
I wish to compare Newton and Michelangelo in the larger context of whether there is any similarity in the motivations of scientists and artists in their respective creative quests. There are many pitfalls in addressing this question: the motivations of individual scientists and artists are diverse; they are strongly dependent on personal tastes and temperaments; and a consideration of this subject in the abstract and in general terms will rapidly degenerate into dilettantism. On this account, I shall restrict my consideration to an example from the most rarefied level of creativity: Newton in writing the Principia and Michelangelo in painting the ceiling of the Sistine Chapel. The Principia and the Frescoes, both in their realms are the supreme, unsurpassed expressions of human creativity. That their origins should be as similar as they are is an astonishing and a revealing fact. Neither Newton nor Michelangelo began their masterpieces with alacrity or enthusiasm: Newton had to be persuaded by Halley and Michelangelo was forced by the imperious insistence of Pope Julius II. But once they started, their visions enlarged; and they both completed their great works in a time—about two years—that is hard even to imagine. Let me first describe the story of the Principia: the origin of his ideas on the laws of gravitation, how they matured and how he came to write the Principia. Newton's first thoughts on gravitation came to him in 1666 while he was sojourning in his manor in Woolsthorpe during the plague years. He deduced how the inverse square law of gravitational attraction between the Sun and the planets from Kepler's third law that the periods are proportional to the 3/2 power of the radii on the assumption of circular orbits. The argument was very simple: Assume that the orbit is a circle of radius r described with a constant velocity v. Then 2T7T
v
„
= T,
where T is the period. By Kepler's third law T oc r3'2.
Therefore 4
T
4ir2 oc
and v2 1 the centripetal force = — a —. r r2 Second, again on the assumption of circular orbits, the attraction between the earth and m o o n and the value of gravity (= g) on the earth, i.e. the attraction of the earth on bodies on the surface of the earth, was compatible with the inverse square law of gravitation. And here again the arguments are simple: In the following diagram the moon located at A at some given instant of time, will continue in the rectilinear path AB but for the attraction of the earth at C. It is the attraction of the earth that is responsible for the moon describing the circular orbit ADEA. Therefore BD represents the effect of the attraction. Thus,
BD
=2*~y = i AD2 ~ 2acc{2nR)2
T
=
r
,
2a\^) 1 BD-BE ~ 2acc(2nR)2
=
E
T
_ ~
BD A7r2R
Ucc
Therefore, = &CC
ATT2R rpy
39.48 X 3.815 X IP 10 ~ (29.74 X 24 X 3600) 2 ~ =
'
But a body (apple!) on the earth, attracted by gravity g ( = 978), accelerates it by the amount acc (of apple) = 978 = 3614 X acc (Moon) = (60) 2 X acc (Moon). In other words, the attraction of the earth falls off as the inverse square of the distance, since the ratio of radius of the moon's orbit and that of the earth is -60. Even though the results were promising, Newton did not pursue the subject further for another 13 years. In 1679, stimulated by some correspondence with Hooke, Newton showed that Kepler's law of areas was not specific to the inverse-square law of attraction but that it was valid for any law of centripetal attraction; and further that for a body revolving in an ellipse, the law of attraction directed to its focus is inversely as the square of the distance. Nevertheless, he was reluctant to publish his results because of his uneasiness with the assumption, q ce (t) _ / Radius of the earth \ g \ Radius of the moon's orbit/ which implies that the earth attracts objects on its surface as if its entire mass is concentrated at the centre—an assumption most emphatically against 'common sense' (unless one had known of its truth already). Newton was to prove the theorem in question in 1685 which he had not suspected before the demonstration. The real story of Newton's development of his theory of gravitation with the concomitant writing of the Principia begins with Halley s visit to Newton in Cambridge in August, 1684 to inquire of Newton the character of an orbit that a particle would describe in a central inverse-square law of attraction. Newton's immediate response was that the orbit will be an ellipse and that he had established that result some years earlier. Halley, astonished at this prompt response, asked for a demonstration. Newton could not find his demonstration among his papers; and he promised to rework his proof and send it to Halley in due course.
In trying to re-derive the result, Newton became sufficiently involved in the subject to give a course of lectures during the Michelmas term of 1684. He forwarded the substance of these lectures to Halley in November of the same year. Halley was so excited with what he received that he visited Newton again in November, and tried to persuade him that he should write out his lectures for publication. After his visit to Newton in November, Halley reported to the Royal Society on December 10, 1684, that he had . . . lately seen Mr. Newton at Cambridge, who had shown him a curious treatise De Motu, which was promised to be sent to the Society to be entered upon their Register. Halley had, thus, at long last persuaded Newton to write out the results of his investigations for publication; and the writing of what was to become the Principia began in earnest. Indeed, Halley was described by Conduit as 'the Ulysses who produced this Achilles.' In the spring of 1685, Newton's attitude to writing the Principia changed when he proved, apparently after some initial difficulty, that the gravitational attraction of a spherical body on an external particle is the same as if its entire mass is concentrated at its center. Newton had not expected this result; and it is not difficult to imagine that it must have goaded Newton to further effort. As Newton wrote to Halley on June 26, 1686, I never extended the duplicate proportion lower than to the superfices of the earth, and before a certain demonstration I found last year, have suspected it did not reach accurately enough down so low; and therefore in the doctrines of projectiles never used it nor considered the motions of heavens. In an address given on the bicentenary of the publication of the Principia, J. W. Glaisher said, Newton proved this superb theorem—and we know from his own words that he had no expectation of so beautiful a result till it emerged from his mathematical investigation—that all the mechanism of the universe at once lay spread before him. Newton's initial expectation of quickly completing the task of writing was thus frustrated. But under Halley's discreet but constant persuasion and quiet encouragement, De Motu Corporum evolved and broadened, swelling first in the early summer of 1685 to a pair of related books and then to a third book on
the 'System of the World' until in its final version was ready for the printer in April 1687. As Newton has recorded, I wrote in 17 or 18 months beginning in the end of December 1684 and sending it to the Royal Society in May, 1686, excepting that about 10 or 12 of the propositions were composed before. As Rouse Ball has written, . . . the first two books were really written in about six months, and the period of eighteen months which the whole composition is said to have occupied includes the time in which copies of them for the press were prepared, and much of the material for the third book collected. The manuscript for the first book was sent to press before June 7, 1686. In it the three laws of motion (which every student of science can recite) are formulated with clear distinction between the notions of the inertial and gravitational mass and their equality; a logical and a coherent account of the dynamics of the motion of particles in general and under the influence of centripetal forces, in particular; the laws of gravitation; and the gravitational attraction of spherical and slightly oblate bodies and of spherical shells. In the third Book, the propositions of Book I are applied to the principal phenomena of the solar system, the determination of the masses and distances of the planets and their satellites and an extended discussion of various perturbations affecting the orbit of the moon. The theory of tides is worked out in detail. The oblateness of the earth is shown to be caused by the earth's rotation; and it is quantitatively related to the difference in the polar and the equatorial values of gravity. Newton also investigates the motion of comets and shows that they belong to the solar system and explains how from three observations the orbit of a comet can be deduced. This extraordinary range and variety of problems, almost all treated here for the first time, and written in magisterial style in less than two years, raises Principia to a level of intellectual achievement unparalleled in the history of science. * ** I t u r n now to Michelangelo. In March 1508, Pope Julius II called on Michelangelo to paint the ceiling of the Sistine Chapel. Michelangelo first stoutly declined on the grounds that he was a sculptor and not a painter and that he had no experience whatever in
painting frescoes. But the Pope would have none of it. And before the imperious insistence of the Pope, Michelangelo had no choice. Even after his acceptance of the Pope's command, he continued to be reserved and bitter; but his attitude changed when the Pope allowed him to change the theme. Originally the Pope had wanted him to paint the 12 Apostles. Michelangelo was able to convince him that it was in his words a 'poor theme.' The Pope commissioned him anew to choose his own theme. The result is what we see today: Not the Apostles, but the entire story of the Genesis: The Creation of the Universe, of Man, and of Evil. The preparation for the frescoes of the Sistine Chapel began in May, 1508. The erection of the scaffolding was finished by July. The actual work of painting, carried out by Michelangelo himself, was started not later than January 1509. He began near the entrance, progressing towards the altar. The three episodes, the drunkenness of Noah, the Deluge, and the sacrifice of Noah were painted first; and they were completed by September 1509. A year later two more and again a year later all 9 histories, together with the Sibyls and the Prophets were completed. On August 14, 1511, on the eve of the feast of the Assumption of Virgin Mary, the frescoes of the ceiling were unveiled; and on October 31,1512, the first mass was celebrated. * * * There is another aspect of great works such as the Principia and the Frescoes of the Sistine Chapel that is characteristic of them; and that is the evolution of creative power which they manifest. Let me consider first the frescoes of the Sistine Chapel. A trained art critic who is also an artist, will be able to illustrate the evolution of Michelangelo's creative power as his painting proceeds from the Drunkenness of Noah to the Separation of Light from Darkness. Being neither an artist nor an art critic, I shall select for comment one feature of the central panels: the face of the Creator in the last five panels in the order in which Michelangelo painted them. (1) In the Creation of Eve: The Creator's face is benign and compassionate—very human; (2) In the Creation of Adam: The stupendous head is God-like but still human; (3) In the Separation of Land and Water: The face presents a powerful image—the fore-shortening of the figure throws into relief the wonderful, creative blessing hands; (4) In the Creation of Sun, Moon and Planets: The Creator breaks forward with a face expressing the stupendous force needed for the creation of the abode of all living things; and finally
1385 (5) In the Separation of Light from Darkness: Never has the ineffable been expressed in art with such intensity. Here, the supreme act of creation attains an almost dehumanized abstraction—in magnificent contrast with the benign features in the Creation of Eve. Let me now turn to the Principia. Unfortunately, the occasion does not allow me to describe the growing intellectual power and physical insight of Newton as we proceed from Book I to Book III of the Principia in the manner I have described the evolution of Michelangelo's artistic power. I must be content with quoting some marvelous statements that Newton makes through the course of the Principia; and these may suffice to illustrate the same things. (1) After formulating the Laws of Motion, the concept of force and the differing notions of the inertial and gravitational mass and the underpinning notions of space and time, Newton realizing the inherent difficulties of resolving the web of intertwining concepts, cuts the Gordian knot with the simple statement: How we are to obtain the true motions from their causes, effects, and apparent differences, and the converse, shall be explained more at large in the following. For to this end it was I composed it. (2) After completing the dynamics of point particles, Newton is ready to prove his superb theorems on the attraction by spherical and non-spherical bodies. And here is his opening statement: In mathematics we are to investigate the quantities of forces with their proportions consequent upon any condition supposed; then, when we enter upon physics, we compare those proportions with the phenomena of Nature, that we may know what conditions of those forces answer to the several kinds of attractive bodies. And this preparation being made, we argue more safely concerning the physical species, causes, and proportions of the forces. (3) In Book III, before stating and proving the marvelously coherent set of 14 Propositions establishing his universal law of gravitation, he prefaces by formulating 4 rules of Reasoning in Philosophy. Here is his Rule I, unsurpassed in clarity and style of expression: We are to admit no more causes of natural things than such as are both true and sufficient to explain their appearances. To this purpose the philosophers say that Nature does nothing in vain and more is in vain when less will serve; for Nature is pleased with simplicity, and affects not the pomp of superfluous causes. (4) And finally the poignant note of the concluding sentence:
And to us it is enough that gravity does really exist, and act according to the laws which we have explained, and abundantly serves to account for all the motions of the celestial bodies, and of our sea. Let me conclude with this note. What I have tried to say during the past half an hour is simply this: When a supremely great creative mind is kindled, it leaves a blazing trail that remains a beacon for centuries.
ON STARS, THEIR EVOLUTION AND THEIR STABILITY Nobel lecture, 8 December, 1983 by SUBRAHMANYAN CHANDRASEKHAR The University of Chicago, Chicago, Illinois 60637, USA
1. Introduction When we think of atoms, we have a clear picture in our minds: a central nucleus and a swarm of electrons surrounding it. We conceive them as small objects of sizes measured in Angstroms (~10~ 8 cm); and we know that some hundred different species of them exist. This picture is, of course, quantified and made precise in modern q u a n t u m theory. And the success of the entire theory may be traced to two basic facte: first, the Bohr radius of the ground state of the hydrogen atom, namely, ^~0.5xl0-*c
m
,
(1)
where h is Planck's constant, m is the mass of the electron and e is its charge, provides a correct measure of atomic dimensions; and second, the reciprocal of Sommerfeld's fine-structure constant,
gives the maximum positive charge of the central nucleus that will allow a stable electron-orbit around it. This maximum charge for the central nucleus arises from the effects of special relativity on the motions of the orbiting electrons. We now ask: can we understand the basic facts concerning stars as simply as we understand atoms in terms of the two combinations of natural constants (1) and (2). In this lecture, I shall attempt to show that in a limited sense we can. T h e most important fact concerning a star is its mass. It is measured in units of the mass of the sun, Q , which is 2 X 10 gm: stars with masses very much less than, or very much more than, the mass of the sun are relatively infrequent. T h e current theories of stellar structure and stellar evolution derive their successes largely from the fact that the following combination of the dimensions of a mass provides a correct measure of stellar masses:
where G is the constant of gravitation and H is the mass of the hydrogen atom. In the first half of the lecture, I shall essentially be concerned with the question: how does this come about? 58
2. The role of radiation pressure A central fact concerning normal stars is the role which radiation pressure plays as a factor in their hydrostatic equilibrium. Precisely the equation governing the hydrostatic equilibrium of a star is dP
Jr
GM(r) =
—TP>
(4)
where P denotes the total pressure, p the density, and M (r) is the mass interior to a sphere of radius r. There are two contributions to the total pressure P : that due to the material and that due to the radiation. O n the assumption that the matter is in the state of a perfect gas in the classical Maxwellian sense, the material or the gas pressure is given by
^ gas
=
~uHpT'
^
where T is the absolute temperature, k is the Boltzmann constant, and /I is the mean molecular weight (which under normal stellar conditions is ~ 1.0). The pressure due to radiation is given by Pr,A = \aT\
(6)
where a denotes Stefan's radiation-constant. Consequently, if radiation contributes a fraction (1 — /}) to the total pressure, we may write
To bring out explicitly the role of the radiation pressure in the equilibrium of a star, we may eliminate the temperature, T, from the foregoing equations and express P in terms of p and p1 instead of in terms of p and T. We find:
and I"/ J- \4 3 i_tf"li/i /3
iz£l'
[W «
J
„,„
„4/s _ 4/3 r,a\n*n p 4 / 3 = C(/3)p (say).
(9)
T h e importance of this ratio, (1— p1), for the theory of stellar structure was first emphasized by Eddington. Indeed, he related it, in a famous passage in his book on The Internal Constitution of the .Stars, to the 'happening of the stars'. A more rational version of Eddington's argument which, at the same time, isolates the combination (3) of the natural constants is the following: There is a general theorem 2 which states that the pressure, Pc, at the centre of a star of a mass M in hydrostatic equilibrium in which the density, p (r), at a point at a radial distance, r, from the centre does not exceed the mean density, p (r), interior to the same point r, must satisfy the inequality, 59
(')
(»)
Fig. 1. A comparison of an inhomogeneous distribution of density in a star (b) with the two homogeneous configurations with the constant density equal to the mean density (a) and equal to the density at the centre (c).
1/3
(8
4/3
p
2/3
M
^Pc
HI")'
4/3
M
2/3
(10)
where p denotes the mean density of the star and pc its density at the centre. The content of the theorem is no more than the assertion that the actual pressure at the centre of a star must be intermediate between those at the centres of the two configurations of uniform density, one at a density equal to the mean density of the star, and the other at a density equal to the density pc at the centre (see Fig. 1). If the inequality (10) should be violated then there must, in general, be some regions in which adverse density gradients must prevail; and this implies instability. In other words, we may consider conformity with the inequality (10) as equivalent to the condition for the stable existence of stars. The right-hand side of the inequality (10) together with P given by equation (9), yields, for the stable existence of stars, the condition,
mv-fT
1/3
2/3
GM
,
(HI
or, equivalently,
M>fi\m\lk
V 3 iz&! /2 i
(12)
where in the foregoing inequalities, /?,. is a value of/3 at the centre of the star. Now Stefan's constant, a, by virtue of Planck's law, has the value a =
for5*4 15/tV'
(13)
Inserting this value a in the inequality (12) we obtain
We observe that the inequality (14) has isolated the combination (3) of
60
natural constants of the dimensions of a mass; by inserting its numerical value given in equation (3), we obtain the inequality,
^ 2 ^(r^;) 1 / 2 ^ 5 - 4 8 0-
(i5)
This inequality provides an upper limit to (1 —f}c) for a star of a given mass. Thus, l-ft=Sl-/3*,
(16)
where (1— /?*) is uniquely determined by the mass M of the star and the mean molecular weight, ft, by the quartic equation,
n
'»-™ffl o.
(17)
In Table 1, we list the values of 1—/3* for several values of ,M 2 M. From this table it follows in particular, that for a star of solar mass with a mean molecular weight equal to 1, the radiation pressure at the centre cannot exceed 3 percent of the total pressure.
Table 1 T h e maximum radiation pressure, (1 —/3*). at the centre of a star of a given mass, M.
1-/3* 0.01 .03 .10 .20 .30 0.40
A////© 0.56 1.01 2.14 3.83 6.12 9.62
1-/3* 0.50 .60 .70 .80 .85 0.90
M/t2/Q 15.49 26.52 50.92 122.5 224.4 519.6
What do we conclude from the foregoing calculation? We conclude that to the extent equation (17) is at the base of the equilibrium of actual stars, to that extent the combination of natural constants (3), providing a mass of proper magnitude for the measurement of stellar masses, is at the base of a physical theory of stellar structure. 3. Do stars have enough energy to cool? T h e same combination of natural constants (3) emerged soon afterward in a much more fundamental context of resolving a paradox Eddington had formulated in the form of an aphorism: 'a star will need energy to cool.' T h e paradox arose while considering the ultimate fate of a gaseous star in the light of the then new knowledge that white-dwarf stars, such as the companion of Sirius, exist, which have mean densities in the range 105—10 gm cm . As Eddington stated 3 61
I do not see how a star which has once got into this compressed state is ever going to get out of i t — It would seem that the star will be in an awkward predicament when its supply of subatomic energy fails. T h e paradox posed by Eddington was reformulated in clearer physical terms by R. H. Fowler. 4 His formulation was the following: T h e stellar material, in the white-dwarf state, will have radiated so much energy that it has less energy than the same matter in normal atoms expanded at the absolute zero of temperature. If part of it were removed from the star and the pressure taken off, what could it do? Quantitatively, Fowler's question arises in this way. An estimate of the electrostatic energy, Ev, per unit volume of an assembly of atoms, of atomic number Z, ionized down to bare nuclei, is given by Ev=
1.32X 10" Z V / 3 ,
(18)
while the kinetic energy of thermal motions, £ k i n , per unit volume of free particles in the form of a perfect gas of density, p, and temperature, T, is given by 3 k
1.24X10 8
Now if such matter were released of the pressure to which it is subject, it can resume a state of ordinary normal atoms only if £ ki „ > Ev,
(20)
or, according to equations (18) and (19), only if p<(o.94xlO-
3
^.
(21)
This inequality will be clearly violated if the density is sufficiently high. This is the essence of Eddington's paradox as formulated by Fowler. And Fowler resolved this paradox in 1926 in a paper entitled 'Dense Matter' — one of the great landmark papers in the realm of stellar structure: in it the notions of Fermi statistics and of electron degeneracy are introduced for the first time. 4. Fowler's resolution of Eddington's paradox; the degeneracy of the electrons in whitedwarf stars In a completely degenerate electron gas all available parts of the phase space, with momenta less than a certain 'threshold' valuep 0 — the Fermi 'threshold' — are occupied consistently with the Pauli exclusion-principle i.e., with two electrons per 'cell' of volume h3 of the six-dimensional phase space. Therefore,
62
ifn(p) dp denotes the number of electrons, per unit volume, between p a.ndp+dp, then the assumption of complete degeneracy is equivalent to the assertion,
-o
(22)
U»M. I
The value of the threshold momentum pm is determined by the normalization condition n = \°n(P)dp
= | | # ,
(23)
where n denotes the total number of electrons per unit volume. For the distribution given by (22), the pressure/? and the kinetic energy £ kjn of the electrons (per unit volume), are given by
P
=%fj\^P
(24)
Ekm = ~\op2T,,dP,
(25)
and
where vp and Tp are the velocity and the kinetic energy of an electron having a momentum/?. If we set vp = plm and Tp = p1'/2m,
(26)
appropriate for non-relativistic mechanics, in equations (24) and (25), we find F
\5h3mP"
20 W
mn
(2?)
and £kin
- lOA'm^ - 40 U
«"
'
(28)
Fowler's resolution of Eddington's paradox consists in this: at the temperatures and densities that may be expected to prevail in the interiors of the white-dwarf stars, the electrons will be highly degenerate and Ekm must be evaluated in accordance with equation (28) and not in accordance with equation (19); and equation (28) gives, £ k i n = 1.39Xl0 1 3 (p/^) 5 / 3 .
(29)
Comparing now the two estimates (18) and (29), we see that, for matter of the density occurring in the white dwarfs, namely p ^ l O gm cm" , the total kinetic energy is about two to four times the negative potential-energy; and Eddington's 63
paradox does not arise. Fowler concluded his paper with the following highly perceptive statement: The black-dwarf material is best likened to a single gigantic molecule in its lowest q u a n t u m state. O n the Fermi-Dirac statistics, its high density can be achieved in one and only one way, in virtue of a correspondingly great energy content. But this energy can no more be expended in radiation than the energy of a normal atom or molecule. The only difference between black-dwarf matter and a normal molecule is that the molecule can exist in a free state while the black-dwarf matter can only so exist under very high external pressure.
5. The theory of the white-dwarf stars; the limiting mass T h e internal energy ( = 3 P/2) of a degenerate electron gas that is associated with a pressure P is zero-point energy; and the essential content of Fowler's paper is that this zero-point energy is so great that we may expect a star to eventually settle down to a state in which all of its energy is of this kind. Fowler's argument can be more explicitly formulated in the following manner. According to the expression for the pressure given by equation (27), we have the relation, P=K^
where * , = - ( - )
^ ^ p ,
(30)
where fic is the mean molecular weight per electron. An equilibrium configuration in which the pressure, P, and the density p , are related in the manner, P=Kpl + Un,
(31)
is an Emden polytrope of index n. T h e degenerate configurations built on the equation of state (30) are therefore poly tropes of index 3/2; and the theory of polytropes immediately provides the relation, Ky = 0.4242 (GMUi R)
(32)
or, numerically, for K\ given by equation (30), log 10 (/WJb) = - | l o g l ° ( M O - f l o g i o ^ - 1 - 3 9 7 .
(33)
For a mass equal to the solar mass and fie = 2, the relation (33) predicts R = 1.26X10~ 2 % and a mean density of 7.0 X 105 gm/cm 3 . These values are precisely of the order of the radii and mean densities encountered in whitedwarf stars. Moreover, according to equations (32) and (33), the radius of the white-dwarf configuration is inversely proportional to the cube root of the mass. O n this account, finite equilibrium configurations are predicted for all masses. And it came to be accepted that the white-dwarfs represent the last stages in the evolution of all stars. 64
But it soon became clear that the foregoing simple theory based on Fowler's premises required modifications. For, the electrons at their threshold energies, at the centres of the degenerate stars, begin to have velocities comparable to that of light as the mass increases. T h u s , already for a degenerate star of solar mass (with fi,, = 2) the central density (which is about six times the mean density) is 4.19 X 10 gm/cm ; and this density corresponds to a threshold momentum p0 = 1.29 mc and a velocity which is 0.63 c. Consequently, the equation of state must be modified to take into account the effects of special relativity. And this is easily done by inserting in equations (24) and (25) the relations, V
"
=
m{\+p2lmV)m
and T
"
=
"J[(\+P2l™2t)m-A>
(34)
in place of the non-relativistic relations (26). We find that the resulting equation of state can be expressed, parametrically, in the form P = Af{x)andp
= Bx\
(35)
where jtm\5 A =
_ 8jimV/ic
~TfT>B-
H
u3
(36)
and J{x) = x ( x 2 + l ) 1 / 2 ( 2 x 2 - 3 ) + 3 s i n t r ' x.
(37)
£ki„ = Ag(x),
(38)
And similarly
where g(x) = 8x3[(x2+\)m-\]-f(x).
(39)
According to equations (35) and (36), the pressure approximates the relation (30) for low enough electron concentrations (x
I-$)"*.".
(«)
This limiting form of relation can be obtained very simply by setting vp = c in equation (24); then 8TIC
p=
ft0
dp
3
_ 2nc
w\J -Ji?
K
4
(41)
and the elimination of/? 0 with the aid of equation (23) directly leads to equation (40). While the modification of the equation of state required by the special
65
theory of relativity appears harmless enough, it has, as we shall presently show, a dramatic effect on the predicted mass-radius relation for degenerate configurations. T h e relation between P and p corresponding to the limiting form (41) is 4/3 u 1/3 V / 3 he v /> = tf2 p 4 / 3 where *2 = - ( - J ^ p .
(42)
In this limit, the configuration is an Emden poly trope of index 3. And it is well known that when the polytropic index is 3, the mass of the resulting equilibrium configuration is uniquely determined by the constant of proportionality, K2, in the pressure-density relation. We have accordingly, t K \3/2 M m
' "
= 4;r
UG)
(hr\m
(2-018) =0.197 ^ J
1
-^5=5.76^-20-
(43)
(In equation (43), 2.018 is a numerical constant derived from the explicit solution of the Lane-Emden equation for n = 3.) It is clear from general considerations 7 that the exact mass-radius relation for the degenerate configurations must provide an upper limit to the mass of such configurations given by equation (43); and further, that the mean density of the configuration must tend to infinity, while the radius tends to zero, and M—>Miimh. These conditions, straightforward as they are, can be established directly by considering the equilibrium of configurations built on the exact equation of state given by equations (35) —(37). It is found that the equation governing the equilibrium of such configurations can be reduced to the form '
{
LA. rf dr\ where
Jo = 4 + 1 ,
(45)
and mexQ denotes the threshold momentum of the electrons at the centre of the configuration and f] measures the radial distance in the unit
By integrating equation (44), with suitable boundary conditions and for various initially prescribed values of j 0 , we can derive the exact mass-radius relation, as well as the other equilibrium properties, of the degenerate configurations. T h e principal results of such calculations are illustrated in Figures 2 and 3. T h e important conclusions which follow from the foregoing considerations are: first, there is an upper limit, Mimm to the mass of stars which can become degenerate configurations, as the last stage in their evolution; and second, that stars with M > M\imn must have end states which cannot be predicted from the considerations we have presented so far. And finally, we observe that the 66
0
0-1
0-2
0-3
0-4
0-5
0-6
0-7
0-8
0-9
1-0
MJMf Fig. 2. The full-line curve represents the exact (mass-radius)-relation {l\ is defined in equation (46) and M3 denotes the limiting mass). This curve tends asymptotically to the curve appropriate to the low-mass degenerate configurations, approximated by polytropes of index 3/2. The regions of the configurations which may be considered as relativistic (g > (Ki/K2)3) are shown shaded. (From Chandrasekhar, S., Mon. Not. Roy. Astr. Soc, 95, 207 (1935).)
combination of the natural constant (3) now emerges in the fundamental context of Miimlt given by equation (43): its significance for the theory of stellar structure and stellar evolution can no longer be doubted. 6. Under what conditions can normal stars develop degenerate cores?
Once the upper limit to the mass of completely degenerate configurations had been established, the question that required to be resolved was how to relate its existence to the evolution of stars from their gaseous state. If a star has a mass less than Afiimi,, the assumption that it will eventually evolve, towards the completely degenerate state appears reasonable. But what if its mass is greater than M iimlt? 67
Fig. 3. The full-line curve represents the exact (mass-density)-relation for the highly collapsed configurations. This curve tends asymptotically to the dotted curve as M—>0. (From Chandrasekhar, S., Mm. Not. Roy. Astr. Soc, 95, 207 (1935).)
Clues as to what might ensue were sought in terms of the equations and inequalities of §§2 and 3 . 1 0 1 1 T h e first question that had to be resolved concerns the circumstances under which a star, initially gaseous, will develop degenerate cores. From the physical side, the question, when departures from the perfect-gas equation of state (5) will set in and the effects of electron degeneracy will be manifested, can be readily answered. Suppose, for example, that we continually and steadily increase the density, at constant temperature, of an assembly of free electrons and atomic nuclei, in a highly ionized state and initially in the form of a perfect gas governed by the equation of state (5). At first the electron pressure will increase linearly with p ; but soon departures will set in and eventually the density will increase in accordance with the equation of state that describes the fully degenerate electron-gas (see Fig. 4). T h e remarkable fact is that this limiting form of the equation of state is independent of temperature. However, to examine the circumstances when, during the course of evolution, a star will develop degenerate cores, it is more convenient to express the electron pressure (as given by the classical perfect-gas equation of state) in terms ofp and /^defined in the manner (cf. equation (7)).
„ -
k
X
nT-J^
^
(47)
where/),, now denotes the electron pressure. Then, analogous to equation (9), we can write
-KaMv 68
(48)
1
10 27 F
1 1 11 III
1
1 1 1 1 MM
1
1 1 1 1 MM
1 1
T = 1.48xl010 ,
!
-
6.64xl0 9 I
=-
r =-
y
'T = 9.21x10°
i i mill
MIIII i i i
/p5/3
SJ
11 nvj
_ 10 23 —y
/
MIIII i i
Pi
i
/ / / P * / Z / vM1 = 2.74xl0 9
CO
Q?
=
/ / / ^
i mill
10 25
10 21 : 10
)/-
IIIIII i
—
x
i
i 1111 ni
I
i i i im l
10'
i
i i i i m l
10'
10*
'
i
10s
p (grn cm ) Fig. 4. Illustrating how by increasing the density at constant temperature degeneracy always sets
Comparing this with equation (42), we conclude that if
LW/) a &J
>K
*~ s U )
he (MM)
4/3'
(49)
the pressure pc given by the classical perfect-gas equation of state will be greater than that given by the equation if degeneracy were to prevail, not only for the prescribed p and T, but for allp and Thaving the same /Je. Inserting for a its value given in equation (13), we find that the inequality (49) reduces to
or equivalently,
960 I - & w* a n pc
. >
l
l-/? e > 0.0921 = l-/8„(say).
(50)
(51)
(See Fig. 5) 69
CO CO
t3 c£>
p (gm
cm-d)
Fig. 5. Illustrating the onset of degeneracy for increasing density at constant p. Notice that there are no intersections for P> 0.09212. In the figure, 1-P is converted into mass of a star built on the standard model.
For our present purposes, the principal content of the inequality (51) is the criterion that for a star to develop degeneracy, it is necessary that the radiation pressure be less than 9.2 percent of (/>e+/>rad)- This last inference is so central to all current schemes of stellar evolution that the directness and the simplicity of the early arguments are worth repeating. T h e two principal elements of the early arguments were these: first, that radiation pressure becomes increasingly dominant as the mass of the star increases; and second, that the degeneracy of electrons is possible only so long as the radiation pressure is not a significant fraction of the total pressure — indeed, as we have seen, it must not exceed 9.2 percent of (/>e+/>rad)- T h e second of these elements in the arguments is a direct and an elementary consequence of the physics of degeneracy; but the first requires some amplification. T h a t radiation pressure must play an increasingly dominant role as the mass of the star increases is one of the earliest results in the study of stellar structure that was established by Eddington. A quantitative expression for this fact is 70
given by Eddington s standard model which lay at the base of early studies summarized in his The Internal Constitution of the Stars. O n the standard model, the fraction /? ( = gas pressure/total pressure) is a constant through a star. O n this assumption, the star is a polytrope of index 3 as is apparent from equation (9); and, in consequence, we have the relation (cf. equation (43))
-«[%}
M=±n\^-
3/2
(2.018)
(52)
where C (fi) is defined in equation (9). Equation (52) provides a quartic equation for/? analogous to equation (17) for/3 . Equation (52) for/? = /Jm gives M = 0.197fti 3 / 2 ( | ) 3 / 2 ( ~ ) 2 = 6 . 6 5 ^ " 2 0 = M (say).
(53)
O n the standard model, then stars with masses exceeding M will have radiation pressures which exceed 9.2 percent of the total pressure. Consequently stars with M > M cannot, at any stage during the course of their evolution, develop degeneracy in their interiors. Therefore, for such stars an eventual white-dwarf state is not possible unless they are able to eject a substantial fraction of their mass. T h e standard model is, of course, only a model. Nevertheless, except under special circumstances, briefly noted below, experience has confirmed the standard model, namely that the evolution of stars of masses exceeding 7—8 Q must proceed along lines very different from those of less massive stars. These conclusions, which were arrived at some fifty years ago, appeared then so convincing that assertions such as these were made with confidence: Given an enclosure containing electrons and atomic nuclei (total charge zero) what happens if we go on compressing the material indefinitely? (1932) 10 The life history of a star of small mass must be essentially different from the life history of a star of large mass. For a star of small mass the natural white-dwarf stage is an initial step towards complete extinction. A star of large mass cannot pass into the white-dwarf stage and one is left speculating on other possibilities. (1934) 8 And these statements have retained their validity. While the evolution of the massive stars was thus left uncertain, there was no such uncertainty regarding the final states of stars of sufficiently low mass. T h e reason is that by virtue, again, of the inequality (10), the maximum central pressure attainable in a star must be less than that provided by the degenerate equation of state, so long as
or, equivalently
*
(55) 71
We conclude that there can be no surprises in the evolution of stars of mass less than 0.43 0 (if ft,. = 2). T h e end stage in the evolution of such stars can only be that of the white dwarfs. (Parenthetically, we may note here that the inequality (55) implies that the so-called 'mini' black-holes of mass ~ 1015 gm cannot naturally be formed in the present astronomical universe.) 7. Some brief remarks on recent progress in the evolution of massive stars and the onset of gravitational collapse It became clear, already from the early considerations, that the inability of the massive stars to became white dwarfs must result in the development of much more extreme conditions in their interiors and, eventually, in the onset of gravitational collapse attended by the super-nova phenomenon. But the precise manner in which all this will happen has been difficult to ascertain in spite of great effort by several competent groups of investigators. T h e facts which must be taken into account appear to be the following. In the first instance, the density and the temperature will steadily increase without the inhibiting effect of degeneracy since for the massive stars considered 1 — /?,, > 1 — /3W. O n this account, 'nuclear ignition' of carbon, say, will take place which will be attended by the emission of neutrinos. This emission of neutrinos will effect a cooling and a lowering of (1— /? e ); but it will still be in excess of 1 — /} w . T h e important point here is that the emission of neutrinos acts selectively in the central regions and is the cause of the lowering of (1— /Je) in these regions. T h e density and the temperature will continue to increase till the next ignition of neon takes place followed by further emission of neutrinos and a further lowering of (1— /? e ). This succession of nuclear ignitions and lowering of (1 — f}t) will continue till 1 — /Je < 1 — /?„ and a relativistically degenerate core with a mass approximately that of the limiting mass ( = 1 . 4 0 for [it = 2) forms at the centre. By this stage, or soon afterwards, instability of some sort is expected to set in (see following §8) followed by gravitational collapse and the phenomenon of the super-nova (of type I I ) . In some instances, what was originally the highly relativistic degenerate core of approximately 1.4 0 , will be left behind as a neutron star. T h a t this happens sometimes is confirmed by the fact that in those cases for which reliable estimates of the masses of pulsars exist, they are consistently close to 1.4 0 . However, in other instances — perhaps, in the majority of the instances — what is left behind, after all 'the dust has settled', will have masses in excess of that allowed for stable neutron stars; and in these instances black holes will form. In the case of less massive stars (M ~ 6—8©) the degenerate cores, which are initially formed, are not highly relativistic. But the mass of core increases with the further burning of the nuclear fuel at the interface of the core and the mantle; and when the core reaches the limiting mass, an explosion occurs following instability; and it is believed that this is the cause underlying super-nova phenomenon of type I. * I am grateful to Professor D. Arnett for guiding me through the recent literature and giving me advice in the writing of this section.
72
From the foregoing brief description of what may happen during the late stages in the evolution of massive stars, it is clear that the problems one encounters are of exceptional complexity, in which a great variety of physical factors compete. This is clearly not the occasion for me to enter into a detailed discussion of these various questions. Besides, Professor Fowler may address himself to some of these matters in his lecture that is to follow. 8. Instabilities of relativistic origin: (I) The vibrational instability oj spherical stars I now turn to the consideration of certain types of stellar instabilities which are derived from the effects of general relativity and which have no counterparts in the Newtonian framework. It will appear that these new types of instabilities of relativistic origin may have essential roles to play in discussions pertaining to gravitational collapse and the late stages in the evolution of massive stars. We shall consider first the stability of spherical stars for purely radial perturbations. T h e criterion for such stability follows directly from the linearized equations governing the spherically symmetric radial oscillations of stars. In the framework of the Newtonian theory of gravitation, the stability for radial perturbations depends only on an average value of the adiabatic exponent, F\, which is the ratio of the fractional Lagrangian changes in the pressure and in the density experienced by a fluid element following the motion; thus, A P/P = T, A p / p .
(56)
And the Newtonian criterion for stability is _
T, =
(M 3
CM
Ti(r)P(r)dM{r)+\ o
J
4
P (r) dM (r) > o
(57)
J
If Ti < 4/3, dynamical instability of a global character will ensue with an ^-folding time measured by the time taken by a sound wave to travel from the centre to the surface. When one examines the same problem in the framework of the general theory of relativity, one finds 12 that, again, the stability depends on an average value of Ti; but contrary to the Newtonian result, the stability now depends on the radius of the star as well. Thus, one finds that no matter how high Y\ may be, instability will set in provided the radius is less than a certain determinate multiple of the Schwarzschild radius, Rs = 2 GM/c2. (58) Thus, if for the sake of simplicity, we assume that Ti is a constant through the star and equal to 5/3, then the star will become dynamically unstable for radial perturbations, if/fj < 2.4 Rs. And further, if Ti —* °°, instability will set in for all R < (9/8) Rs. The radius (9/8) ./?s defines, in fact, the minimum radius which any gravitating mass, in hydrostatic equilibrium, can have in the framework of general relativity. This important result is implicit in a fundamental paper by Karl Schwarzschild published in 1916. (Schwarzschild actually proved that for a star in which the energy density is uniform, R > (9/8)Rs. 73
In one sense, the most important consequence of this instability of relativistic origin is that if T{ (again assumed to be a constant for the sake of simplicity) differs from and greater than 4/3 only by a small positive constant, then the instability will set in for a radius R which is a large multiple of ^?s; and, therefore, under circumstances when the effects of general relativity, on the structure of the equilibrium configuation itself, are hardly relevant. Indeed, it follows from the equations governing radial oscillations of a star, in a first post-Newtonian approximation to the general theory of relativity, that instability for radial perturbations will set in for all K 2GM * < ^ - T 7 ; - i r 2- ,
(59)
r,-4/3 c
where K is a constant which depends on the entire march of density and pressure in the equilibrium configuration in the Newtonian frame-work. T h u s , for a poly trope of index n, the value of the constant is given by
where 6 is the Lane-Emden function in its standard normalization (6 = 1 at § = 0), £ is the dimensionless radial coordinate, §i defines the boundary of the polytrope (where 6=0) and 6\ is the derivative of 6 at £i.
n 0 1.0 1.5 2.0 2.5 3.0
Table 2 Values of the constant K in the inequality (59) for various polytropic indices, n. K n 0.452381 3.25 565382 3.5 645063 4.0 751296 4.5 900302 4.9 1.12447 4.95
K 1.28503 1.49953 2.25338 4.5303 22.906 45.94
In Table 2, we list the values of K for different polytropic indices. It should be particularly noted that K increases without limit for n —* 5 and the configuration becomes increasingly centrally condensed. T h u s , already for n = 4.95 (for which polytropic index p c = 8.09 X 106 /5), X ~ 4 6 . In other words, for thehighly centrally condensed massive stars (for which I^may differ from 4/3 by as little It is for this reason that we describe the instability as global. " Since this was written, it has been possible to show (Chandrasekhar and Lebovitz 13a) that for it—» 5, the asymptotic behaviour of K is given by K-> 2.3056/(5-n). and, further, that along the polytropic sequence, the criterion for instability (59) ,can be expressed alternatively in the form R<
74
0.2264
2W£_J_^,
AP)
(Pt>io'p)
as 0.01), the instability of relativistic origin will set in, already, when its radius falls below 5 X 1 0 Rs. Clearly this relativistic instability must be considered in the contexts of these problems. A further application of the result described in the preceding paragraph is to degenerate configurations near the limiting mass 14 . Since the electrons in these highly relativistic configurations have velocities close to the velocity of light, the effective value of Ti will be very close to 4/3 and the post-Newtonian relativistic instability will set in for a mass slightly less than that of the limiting mass. O n account of the instability for radial oscillations setting in for a mass less than Miimi„ the period of oscillation, along the sequence of the degenerate configurations, must have a minimum. This minimum can be estimated to be about two seconds (see Fig. 6). Since pulsars, when they were discovered, were known to have periods much less than this minimum value, the possibility of their being degenerate configurations near the limiting mass was ruled out; and this was one of the deciding factors in favour of the pulsars being neutron stars. (But by a strange irony, for reasons we have briefly explained in § 7, pulsars which have resulted from super-nova explosions have masses close to 1.4 G>!) Finally, we may note that the radial instability of relativistic origin is the underlying cause for the existence of a maximum mass for stability: it is a direct consequence of the equations governing hydrostatic equilibrium in general relativity. (For a complete investigation on the periods of radial oscillation of neutron stars for various admissible equations of state, see a recent paper by Detweiler and Lindblom 1 5 .) 9. Instabilities of relativistic origin: (2) The secular instability of rotating stars derived from the emission of gravitational radiation by non-axisymmetric modes of oscillation I now turn to a different type of instability which the general theory of relativity predicts for rotating configurations. This new type of instability 16 has its origin in the fact that the general theory of relativity builds into rotating masses a dissipative mechanism derived from the possibility of the emission of gravitational radiation by non-axisymmetric modes of oscillation. It appears that this instability limits the periods of rotation of pulsars. But first, I shall explain the nature and the origin of this type of instability. It is well known that a possible sequence of equilibrium figures of rotating homogeneous masses is the Maclaurin sequence of oblate spheroids 1 7 . When one examines the second harmonic oscillations of the Maclaurin spheroid, in a frame of reference rotating with its angular velocity, one finds that for two of these modes, whose dependence on the azimuthal angle is given by e2"p, the characteristic frequencies of oscillation, o, depend on the eccentricity e in the manner illustrated in Figure 7. It will be observed that one of these modes becomes neutral (i.e., 0 = 0) when e = 0.813 and that the two modes coalesce when e = 0.953 and become complex conjugates of one another beyond this * By reason of the dominance of the radiation pressure in these massive stars and of P being very close to zero.
75
30
3
W
I0 10 I012 I0 14 pz (kg nrf3)
Fig. 6. The variation of the period of radial oscillation along the completely degenerate configurations. Notice that the period tends to infinity for a mass close to the limiting mass. There is consequently a minimum period of oscillation along these configurations; and the minimum period approximately 2 seconds. (From J. Skilling, Pulsating Stars (Plenum Press, New York, 1968), p. 59.)
point. Accordingly, the Maclaurin spheroid becomes dynamically unstable at the latter point (first isolated by Riemann). O n the other hand, the origin of the neutral mode at e = 0.813 is that at this point a new equilibrium sequence of triaxial ellipsoids—the ellipsoids of Jacobi—bifurcate. O n this latter account, Lord Kelvin conjectured in 1883 that if there be any viscosity, however slight . . . the equilibrium beyond e = 0.81 cannot be secularly stable. Kelvin's reasoning was this: viscosity dissipates energy but not angular momentum. And since for equal angular momenta, the Jacobi ellipsoid has a lower energy content than the Maclaurin spheroid, one may expect that the action of viscosity will be to dissipate the excess energy of the Maclaurin spheroid and transform it into the Jacobi ellipsoid with the lower energy. A detailed calculation 1 8 of the effect of viscous dissipation on the two modes of oscillation, illustrated in Figure 7, does confirm Lord Kelvin's conjecture. It is found that viscous dissipation makes the mode, which becomes neutral at e = 0.813, unstable beyond this point with an e-folding time which depends inversely on the magnitude of the kinematic viscosity and which further decreases monotonically to zero at the point, e = 0.953 where the dynamical instability sets in. Since the emission of gravitational radiation dissipates both energy and angular momentum, it does not induce instability in the Jacobi mode; instead it
76
Fig. 7. The characteristic frequencies (in the unit (JtGp) 2) of the two even modes of secondharmonic oscillation of the Maclaurin spheriod. The Jacobi sequence bifurcates from the Maclaurin sequence by the mode that is neutral (a = 0) at e = 0.813; and the Dedekind sequence bifurcates by the alternative mode at D. At 0 2 (e = 0.9529) the Maclaurin spheroid becomes dynamically unstable. The real and the imaginary parts of the frequency, beyond 0 2 , are shown by the full line and the dashed curves, respectively. Viscous dissipation induces instability in the branch of the Jacobi mode; and radiation-reaction induces instability in the branch D 0 2 of the Dedekind mode.
induces instability in the alternative mode at the same eccentricity. In the first instance this may appear surprising; but the situation we encounter here clarifies some important issues. If instead of analyzing the normal modes in the rotating frame, we had analyzed them in the inertial frame, we should have found that the mode which becomes unstable by radiation reaction at e = 0.813, is in fact neutral at this point. And the neutrality of this mode in the inertial frame corresponds to the fact that the neutral deformation at this point is associated with the bifurcation (at this point) of a new triaxial sequence—the sequence of the Dedekind ellipsoids. These Dedekind ellipsoids, while they are congruent to the Jacobi ellipsoids, they differ from them in that they are at rest in the inertial frame and owe their triaxial figures to internal vortical motions. An important conclusion that would appear to follow from these facts is that in the framework of general relativity we can expect secular instability, derived from radiation reaction, to arise from a Dedekind mode of deformation (which is quasi-stationary in the inertial frame) rather than the Jacobi mode (which is quasi-stationary in the rotating frame). A further fact concerning the secular instability induced by radiation reaction, discovered subsequently by Friedman 1 9 and by Comins 2 0 , is that the 77
modes belonging to higher values of m ( = 3, 4, , ,) become unstable at smaller eccentricities though the ^-folding times for the instability becomes rapidly longer. Nevertheless it appears from some preliminary calculations of Friedman 2 1 that it is the secular instability derived from modes belonging to m= 3 (or 4) that limit the periods of rotation of the pulsars. It is clear from the foregoing discussions that the two types of instabilities of relativistic origin we have considered are destined to play significant roles in the contexts we have considered. 10. The mathematical theory of black holes So far, I have considered only the restrictions on the last stages of stellar evolution that follow from the existence of an upper limit to the mass of completely degenerate configurations and from the instabilities of relativistic origin. From these and related considerations, the conclusion is inescapable that black holes will form as one of the natural end products of stellar evolution of massive stars; and further that they must exist in large numbers in the present astronomical universe. In this last section I want to consider very briefly what the general theory of relativity has to say about them. But first, I must define precisely what a black hole is. A black hole partitions the three-dimensional space into two regions: an inner region which is bounded by a smooth two-dimensional surface called the event horizon; and an outer region, external to the event horizon, which is asymptotically flat; and it is required (as a part of the definition) that no point in the inner region can communicate with any point of the outer region. This incommunicability is guaranteed by the impossibility of any light signal, originating in the inner region, crossing the event horizon. T h e requirement of asymptotic flatness of the outer region is equivalent to the requirement that the black hole is isolated in space and that far from the event horizon the spacetime approaches the customary space-time of terrestrial physics. In the general theory of relativity, we must seek solutions of Einstein's vacuum equations compatible with the two requirements I have stated. It is a startling fact that compatible with these very simple and necessary requirements, the general theory of relativity allows for stationary (i.e., time-independent) black-holes exactly a single, unique, two-parameter family of solutions. This is the Kerr family, in which the two parameters are the mass of the black hole and the angular momentum of the black hole. W h a t is even more remarkable, the metric describing these solutions is simple and can be explicitly written down. I do not know if the full import of what I have said is clear. Let me explain. Black holes are macroscopic objects with masses varying from a few solar masses to millions of solar masses. T o the extent they may be considered as stationary and isolated, to that extent, they are all, every single one of them, described exactly by the Kerr solution. This is the only instance we have of an exact description of a macroscopic object. Macroscopic objects, as we see them all around us, are governed by a variety of forces, derived from a variety of approximations to a variety of physical theories. In contrast, the only elements 78
in the construction of black holes are our basic concepts of space and time. They are, thus, almost by definition, the most perfect macroscopic objects there are in the universe. And since the general theory of relativity provides a single unique two-parameter family of solutions for their descriptions, they are the simplest objects as well. Turning to the physical properties of the black holes, we can study them best by examining their reaction to external perturbations such as the incidence of waves of different sorts. Such studies reveal an analytic richness of the Kerr space-time which one could hardly have expected. This is not the occasion to elaborate on these technical matters 2 2 . Let it suffice to say that contrary to every prior expectation, all the standard equations of mathematical physics can be solved exactly in the Kerr space-time. And the solutions predict a variety and range of physical phenomena which black holes must exhibit in their interaction with the world outside. T h e mathematical theory of black holes is a subject of immense complexity; but its study has convinced me of the basic truth of the ancient mottoes, T h e simple is the seal of the true and Beauty is the splendour of truth.
79
1.
Eddington, A. S., The internal Constitution of the Stars (Cambridge University Press, England, 1926), p. 16. 2. Chandrasekhar, S., Mon. Not. Roy. Astr. Soc, 96, 644 (1936). 3. Eddington, A. S., The Internal Constitution of the Stars (Cambridge University Press, England, 1926), p. 172. 4. Fowler, R. H., Mon. Not. Roy. Astr. Soc, ST, 114 (1926). 5. Chandrasekhar, S., Phil. Mag., 11, 592 (1931). 6. Chandrasekhar, S., Astrophys. J., 74, 81 (1931). 7. Chandrasekhar, S., Mon. Not. Roy. Astr. Soc, 91, 456 (1931). 8. Chandrasekhar, S., Observatory, 57, 373 (1934). 9. Chandrasekhar, S., Mon. Not. Roy. Astr. Soc, 95, 207 (1935). 10. Chandrasekhar, S., Z.f. Astrophysik, 5, 321 (1932). 11. Chandrasekhar, S., Observatory, 57, 93 (1934). 12. Chandrasekhar, S., Astrophys. J., 140, 417 (1964); see also Phys. Rev. Lett., 12, 114 and 437 (1964). 13. Chandrasekhar, S., Astrophys. J., 142, 1519 (1965). 13 a. Chandrasekhar, S. and Lebovitz, N. R., Mon. Not. Roy. Astr. Soc, 207, 13 P (1984). 14. Chandrasekhar, S. and Tooper, R. F., Astrophys. J., 139, 1396 (1964). 15. Detweiler, S. and Lindblom, L., Astrophys. J. Supp., 53, 93 (1983). 16. Chandrasekhar, S., Astrophys. J., 161, 561 (1970); see also Phys. Rev. Lett., 24, 611 and 762 (1970). 17. For an account of these matters pertaining to the classical ellipsoids see Chandrasekhar, S., Ellipsoidal Figures of Equilibrium (Yale University Press, New Haven, 1968). 18. Chandrasekhar, S., Ellipsoidal Figures of Equilibrium (Yale University Press, New Haven, 1968), Chap. 5, § 37. 19. Friedman, J. L., Comm. Math. Phys., 62, 247 (1978); see also Friedman, J. L. and Schutz, B. ¥., Astrophys. J., 222, 281 (1977). 20. Comins, N., Mon. Not. Roy. Astr. Soc, 189, 233 and 255 (1979). 21. Friedman, J. L., Phys. Rev. Lett., 51, 11 (1983). 22. The author's investigations on the mathematical theory of black holes, continued over the years 1974—1983, are summarized in his last book The Mathematical Theory of Black Holes (Clarendon Press, Oxford, 1983). The reader may wish to consult the following additional references: 1. Chandrasekhar, S., 'Edward Arthur Milne: his part in the development of modern astrophysics,' Quart. J. Roy. Astr. Soc, 21, 93-107 (1980). 2. Chandrasekhar, S., Eddington: The Most Distinguished Astrophysicist of His Time (Cambridge University Press, 1983).
Publications by S. Chandrasekhar
P A P E R S 1928 Thermodynamics of the Compton Effect with Reference to the Interior of the Stars. Indian Journal of Physics 3, 2 4 1 - 5 0 1
1929 The Compton Scattering and the New Statistics. Proceedings of the Royal Society, A, 125, 231-37
1930 'The Ionization-Formula and the New Statistics. Philosophical Magazine 9, 2 9 2 - 9 9 . On the Probability Method in the New Statistics. Philosophical Magazine 9, 6 2 1 - 2 4 1931 'The Dissociation Formula according to the Relativistic Statistics. Monthly Notices of the Royal Astronomical Society 91, 446—55 'The Highly Collapsed Configurations of a Stellar Mass. Monthly Notices of the Royal Astronomical Society 91, 4 5 6 - 6 6 ' The Maximum Mass of Ideal White Dwarfs. The Astrophysical Journal, 74, 81 - 8 2 ' The Stellar Coefficients of Absorption and Opacity. Proceedings of the Roval Society, A, 133,241-54 'The Density of White Dwarf Stars. Philosophical Magazine 11 (suppl.), 5 9 2 - 9 6 02 Eridani B. Zeitschrift fur Astrophysik 3, 3 0 2 - 5 1932 Ionization in Stellar Atmospheres, III Astronomical Society 92, 150-86 'Model Stellar Photospheres. Monthly 186-95 The Stellar Coefficients of Absorption A, 135,472-90 ' Some Remarks on the State of Matter 5,321-27
(with E. A. Milne). Monthly Notices of the Royal Notices
of the Roval Astronomical
Society
and Opacity, II. Proceedings of the Roval in the Interior of Stars. Zeitschrift fur
92,
Society,
Astrophysik
1933 'The Equilibrium of Distorted Polytropes. I, The Rotational Problem. Monthly Notices of the Royal Astronomical Society 93, 3 9 0 - 4 0 5 'The Equilibrium of Distorted Polytropes. II, The Tidal Problem. Monthly Notices of the Royal Astronomical Society, 93, 449—61 Superior numbers identify titles included in these Selected Papers by volume.
' The Equilibrium of Distorted Polytropes. Ill, The Double-Star Problem. Monthly Notices of the Royal Astronomical Society 93, 462-71 1 The Equilibrium of Distorted Polytropes. IV, The Rotational and the Tidal Distortions as Functions of the Density Distribution. Monthly Notices of the Royal Astronomical Society 93, 5 3 9 - 7 4 ' The Solar Chromosphere. Monthly Notices of the Royal Astronomical Society 94, 14-35 1934 •The Radiative Equilibrium of Extended Stellar Atmospheres. Monthly Notices of the Royal Astronomical Society 94, 444-58 On the Hypothesis of the Radial Ejection of High-Speed Atoms for the Wolf-Rayet Stars and the Novae. Monthly Notices of the Royal Astronomical Society 94, 5 2 2 - 3 8 The Solar Chromosphere (Second Paper). Monthly Notices of the Royal Astronomical Society 94, 7 2 6 - 3 7 The Relation between the Chromosphere and the Prominences. The Observatory 57, 65-68 1 The Physical State of Matter in the Interior of Stars. The Observatory 57, 93—99 The Radiative Equilibrium of Extended Stellar Atmospheres. The Observatory 57, 225-27 1 Stellar Configurations with Degenerate Cores. The Observatory 57, 3 7 3 - 7 7 An Analysis of the Problems of the Stellar Atmospheres. Astronomical Journal of the Soviet Union 11, 5 5 0 - 9 6 The Stellar Coefficient of Absorption. Zehschrift fur Astrophysik 8, 167 1935 'The Highly Collapsed Configurations of a Stellar Mass (Second Paper). Monthly Notices of the Royal Astronomical Society 95, 207-25 1 Stellar Configurations with Degenerate Cores. Monthly Notices of the Royal Astronomical Society 9 5 , 2 2 6 - 6 0 1 Relativistic Degeneracy (with C. Moller). Monthly Notices of the Royal Astronomical Society 95, 6 7 3 - 7 6 'Stellar Configurations with Degenerate Cores (Second Paper). Monthly Notices of the Royal Astronomical Society 95, 676—93 'The Radiative Equilibrium of the Outer Layers of a Star with Special Reference to the Blanketing Effect of the Reversing Layer. Monthly Notices of the Royal Astronomical Society 9 6 , 2 1 - 4 2 'The Radiative Equilibrium of a Planetary Nebula. Zeitschrift fur Astrophysik 9, 266-89 'The Nebulium Emission in Planetary Nebulae. Zeitschrift filr Astrophysik 10, 36—39 Etude des Atmospheres Stellaires. Memoirs de la Societe Royale des Sciences de Liege 20, 3-89 Stjernernes Struktur. Nordisk Astronomisk Tidsskrift 16, 3 7 - 4 4 'Production of Electron Pairs and the Theory of Stellar Structure (with L. Rosenfeld). Nature 135,999 'On the Effective Temperatures of Extended Photospheres. Proceedings of the Cambridge Philosophical Society 31, 390-93 1936 ' The Pressure in the Interior of a Star. Monthly Notices of the Royal Astronomical Society 96, 6 4 4 - 4 7 The Equilibrium of Stellar Envelopes and the Central Condensations of Stars. Monthly Notices of the Royal Astronomical Society 96, 6 4 7 - 6 0
The Profile of the Absorption Lines in Rotating Stars, Taking into Account the Variation of Ionization Due to Centrifugal Force (with P. Swings). Monthly Notices of the Royal Astronomical Society 96, 883-89 On the Distribution of the Absorbing Atoms in the Reversing Layers of Stars and the Formation of Blended Absorption Lines (with P. Swings). Monthly Notices of the Royal Astronomical Society 97, 2 4 - 3 7 1 On the Maximum Possible Central Radiation Pressure in a Star of a Given Mass. The Observatory 59, 4 7 - 4 8 On Trumpler's Stars (with A. Beer). The Observatory 59, 168-70 1937 The Pressure in the Interior of a Star. The Astrophysical Journal 85, 3 7 2 - 7 9 1 The Opacity in the Interior of a Star. The Astrophysical Journal 86, 7 8 - 8 3 1 Partially Degenerate Stellar Configurations. The Astrophysical Journal 86, 6 2 3 - 2 5 On a Class of Stellar Models. Zeitschrift fur Astrophysik 14, 164-88 'The Cosmological Constants. Nature 139, 7 5 7 - 5 8 1
1938 A Method of Deriving the Constants of the Velocity Ellipsoid from the Observed Radial Speeds of the Stars (with W.M. Smart). Monthly Notices of the Royal Astronomical Society 98, 6 5 8 - 6 3 On a Generalization of Lindblad's Theory of Star-Streaming. Monthly Notices of the Royal Astronomical Society 98, 710-26 1 Ionization and Recombination in the Theory of Stellar Absorption Lines and Nebular Luminosity. The Astrophysical Journal 87, 4 7 6 - 9 5 1 An Integral Theorem on the Equilibrium of a Star. The Astrophysical Journal 87, 535-52 1939 'The Minimum Central Temperature of a Gaseous Star. Monthly Notices of the Royal Astronomical Society 99, 673-85 The Lane-Emden Function 9325. The Astrophysical Journal 89, 116—18 Review of Stellar Dynamics, by W. M. Smart. The Astrophysical Journal 89, 679-87 The Dynamics of Stellar Systems, I-VIII. The Astrophysical Journal 90, 1-154 1 The Internal Constitution of the Stars. Proceedings of the American Philosophical Society 81, 153-87 1940 The Dynamics of Stellar Systems, IX-XIV. The Astrophysical Journal 92, 441-642 1941 The Time of Relaxation of Stellar Systems, I. The Astrophysical Journal 93, 285-304 The Time of Relaxation of Stellar Systems, II (with R. E. Williamson). The Astrophysical Journal 93, 305-22 The Time of Relaxation of Stellar Systems, III. The Astrophysical Journal 93, 323-36 'The White Dwarfs and Their Importance for Theories of Stellar Evolution. Conferences du College de France, Colloque International d'Astrophysique, III, White Dwarfs. 17-23 Juillet 1939 (Paris: Hermann & Company), 4 1 - 5 0 3 A Statistical Theory of Stellar Encounters. The Astrophysical Journal 94, 5 1 1 - 2 5 1 Stellar Models with Isothermal Cores (with L. R. Henrich). The Astrophysical Journal 94, 525-36 1942 'An Attempt to Interpret the Relative Abundances of the Elements and Their Isotopes (with L. R. Henrich. The Astrophysical Journal 95, 2 8 8 - 9 8
3
The Statistics of the Gravitational Field Arising from a R a n d o m Distribution of Stars. I, The Speed of Fluctuations (with J. von Neumann). The Astrophysical Journal 95, 489-531 'A Note on the Perturbation Theory for Distorted Stellar Configurations (with W. Krogdahl). The Astrophysical Journal 96, 151-54 •On the Evolution of the Main-Sequence Stars (with M. Schonberg). The Astrophysical Journal 96, 161-72
1943 The Statistics of the Gravitational Field Arising from a Random Distribution of Stars. II, The Speed of Fluctuations; Dynamical Friction; Spatial Co-Relations (with J. von Neumann). The Astrophysical Journal 97, 1-27 'Dynamical Friction. I, General Considerations: The Coefficient of Dynamical Friction. The Astrophysical Journal 97, 255-62 'Dynamical Friction. II, The Rate of Escape of Stars from Clusters and the Evidence for the Operation of Dynamical Friction. The Astrophysical Journal 97, 263—73 'Dynamical Friction. Ill, A More Exact Theory of the Rate of Escape of Stars from Clusters. The Astrophysical Journal 98, 5 4 - 6 0 On the Negative Hydrogen Ion and Its Absorption Coefficient (with M. K. Krogdahl). The Astrophysical Journal 98, 2 0 5 - 8 'Stochastic Problems in Physics and Astronomy. Reviews of Modern Physics 15, 1-89 'New Methods in Stellar Dynamics. Annals of the New York Academy of Sciences 45, 131-62 3
1944 'The Statistics of the Gravitational Field Arising from a Random Distribution of Stars. III, The Correlations in the Forces Acting at Two Points Separated by a Finite Distance. The Astrophysical Journal 99, 2 5 - 4 6 'The Statistics of the Gravitational Field Arising from a Random Distribution of Stars. IV, The Stochastic Variation of the Force Acting on a Star. The Astrophysical Journal 99,47-53 ' O n the Stability of Binary Systems. The Astrophysical Journal 99, 5 4 - 5 8 On the Radiative Equilibrium of a Stellar Atmosphere. The Astrophysical Journal 99, 180-90 'Galactic Evidences for the Time-Scale of the Universe. Science 99, 133—36 2 0 n the Radiative Equilibrium of a Stellar Atmosphere, II. The Astrophysical Journal 100, 76-86 On the Absorption Continuum of the Negative Oxygen Ion (with R. Wildt). The Astrophysical Journal 100, 8 7 - 9 3 On the Radiative Equilibrium of a Stellar Atmosphere, III. The Astrophysical Journal 100, 117-27 2 Some Remarks on the Negative Hydrogen Ion and Its Absorption Coefficient. The Astrophysical Journal 100,176-80 On the Radiative Equilibrium of a Stellar Atmosphere, IV (with C.U. Cesco and J. Sahade). The Astrophysical Journal 100, 355-59 The Negative Ions of Hydrogen and Oxygen in Stellar Atmospheres. Reviews of Modern Physics 1 6 , 3 0 1 - 6 1945 'Ralph Howard Fowler, 1889-1944. The Astrophysical Journal 101, 1-5 2 On the Radiative Equilibrium of a Stellar Atmosphere, V. The Astrophysical 101,95-107
Journal
On the Radiative Equilibrium of a Stellar Atmosphere, VI (with C. U. Cesco and J. Sahade). The Astrophysical Journal 101, 320-27 2 On the Radiative Equilibrium of a Stellar Atmosphere, VII. The Astrophysical Journal 101,328-47 2 On the Radiative Equilibrium of a Stellar Atmosphere, VIII. The Astrophysical Journal 101,348-55 Photographs of the Corona Taken during the Total Eclipse of the Sun on July 9, 1945, at Pine River, Manitoba, Canada (with W.A. Hiltner). The Astrophysical Journal 102, 135-36 2 On the Continuous Absorption Coefficient of the Negative Hydrogen Ion. The Astrophysical Journal 102, 2 2 3 - 3 1 2 On the Continuous Absorption Coefficient of the Negative Hydrogen Ion, II. The Astrophysical Journal 102, 395-401 2 The Radiative Equilibrium of an Expanding Planetary Nebula. I, Radiation Pressure in Lyman_„. The Astrophysical Journal 102, 4 0 2 - 2 8 2 The Formation of Absorption Lines in a Moving Atmosphere. Reviews of Modern Physics 17, 138-56 Reports on the Progress of Astronomy: Stellar Dynamics. Monthly Notices of the Royal Astronomical Society 105, 124-34 1946 The Motion of an Electron in the Hartree Field of a Hydrogen Atom (with F. H. Breen). The Astrophysical Journal 103, 4 1 - 7 0 On the Radiative Equilibrium of a Stellar Atmosphere, IX. The Astrophysical Journal 103, 165-92 2 On the Radiative Equilibrium of a Stellar Atmosphere, X. The Astrophysical Journal 103,351-70 2 On the Radiative Equilibrium of a Stellar Atmosphere, XI. The Astrophysical Journal 104, 110-32 On a New Theory of Weizsacker on the Origin of the Solar System. Reviews of Modern Physics 18,94-102 2 On the Radiative Equilibrium of a Stellar Atmosphere, XII. The Astrophysical Journal 104, 191-202 2 On the Continuous Absorption Coefficient of the Negative Hydrogen Ion, III (with F. H. Breen). The Astrophysical Journal 104, 4 3 0 - 4 5 2 The Continuous Spectrum of the Sun and the Stars (with G. Munch). The Astrophysical Journal 104, 4 4 6 - 5 7 2 A New Type of Boundary-Value Problem in Hyperbolic Equations. Proceedings of the Cambridge Philosophical Society 42, 250-60 1947 On the Radiative Equilibrium of a Stellar Atmosphere, XIII. The Astrophysical Journal 105, 151-63 2 On the Radiative Equilibrium of a Stellar Atmosphere, XIV. The Astrophysical Journal 105,164-203 2 On the Radiative Equilibrium of a Stellar Atmosphere, XV. The Astrophysical Journal 105,424-34 On the Radiative Equilibrium of a Stellar Atmosphere, XVI (with F. H. Breen). The Astrophysical Journal 105, 4 3 5 - 4 0 2 On the Radiative Equilibrium of a Stellar Atmosphere, XVII. The Astrophysical Journal 105,441-60 2
On the Radiative Equilibrium of a Stellar Atmosphere, XVIII (with F. H. Breen). The Astrophysical Journal 105, 4 6 1 - 7 0 On the Radiative Equilibrium of a Stellar Atmosphere, XIX (with F. H. Breen). The Astrophysical Journal 106, 143-44 2 On the Radiative Equilibrium of a Stellar Atmosphere, XX. The Astrophysical Journal 106,145-51 2 0 n the Radiative Equilibrium of a Stellar Atmosphere, XXI. The Astrophysical Journal 106, 152-216 The Scientist. Pp. 159-79 in The Works of the Mind, ed. Robert B. Heywood. Chicago: University of Chicago Press 2 The Transfer of Radiation in Stellar Atmospheres. Bulletin of the American Mathematical Society 53,641 - 7 1 1 2 The Story of Two Atoms. The Scientific Monthly 64, 313-21 James Hopwood Jeans. Science 105, 2 2 4 - 2 6 Solar Research and Theoretical Astrophysics. Science 106, 2 1 3 - 1 4 1948 On the Radiative Equilibrium of a Stellar Atmosphere, XXII. The Astrophysical Journal 107,48-72 2 On the Radiative Equilibrium of a Stellar Atmosphere, XXII (concluded). The Astrophysical Journal 107, 188-215 On the Radiative Equilibrium of a Stellar Atmosphere, XXIII (with F. H. Breen), The Astrophysical Journal 107, 216—19 On the Radiative Equilibrium of a Stellar Atmosphere, XXIV. The Astrophysical Journal 108,92-111 2 The Softening of Radiation by Multiple Compton Scattering. Proceedings of the Royal Society, A, 192,508-18 3 On a Class of Probability Distributions. Proceedings of the Cambridge Philosophical Society 45, 2 1 9 - 2 4 2
1949 The Theory of Statistical and Isotropic Turbulence. Physical Review 75, 8 9 6 - 9 7 On the Decay of Isotropic Turbulence. Physical Review 75, 1454-55; erratum, Physical Review, 76, 158 3 Brownian Motion, Dynamical Friction and Stellar Dvnamics. Reviews of Modern Physics 21,383-88 The Isothermal Function (with G. W. Wares). The Astrophysical Journal 109, 551—54 The Functions G n m (T) and G' nm (T) of Order 6 (m = 6 and m 3= n). The Astrophysical Journal 109, 555 3 On Heisenberg's Elementary Theory of Turbulence. Proceedings of the Royal Society, A, 200, 2 0 - 3 3 3 Turbulence—A Physical Theory of Astrophysical Interest (Henry Norris Russell Lecture). The Astrophysical Journal 110, 3 2 9 - 3 9 1950 On the Integral Equation Governing the Distribution of the True and the Apparent Rotational Velocities of Stars (with G. Munch). The Astrophysical Journal 111, 142-56 3 The Theory of Axisymmetric Turbulence. Philosophical Transactions of the Royal Society 242, 557-77 3 The Decay of Axisymmetric Turbulence. Proceedings of the Royal Society, A, 203, 358—64 The Invariant Theory of Isotropic Turbulence in Magneto-Hydrodynamics. Proceedings of the Royal Society, A, 204, 4 3 5 - 4 9 3
3
The Theory of the Fluctuations in Brightness of the Milky Way, I (with G. Munch). The Astrophysical Journal 112, 380—92 ' T h e Theory of Fluctuations in Brightness of the Milky Way, II (with G. Munch). The Astrophysical Journal 112, 3 9 3 - 9 8 i951 •The Angular Distribution of the Radiation at the Interface of Two Adjoining Media. Canadian Journal of Physics 29, 14-20 Polarization of the Sunlit Sky (with D. D. Elbert). Nature 167, 5 1 - 5 5 3 0 n Stellar Statistics (with G. Miinch). The Astrophysical Journal 113, 150-65 3 The Invariant Theory of Isotropic Turbulence in Magneto-Hydrodynamics. Proceedings of the Royal Society, A, 204, 4 3 5 - 3 9 ' T h e Invariant Theory of Isotropic Turbulence in Magneto-Hydrodynamics, II. Proceedings of the Royal Society, A, 207, 3 0 1 - 6 ' T h e Fluctuations of Density in Isotropic Turbulence. Proceedings of the Royal Society, A, 210,18-25 ' T h e Gravitational Instability of an Infinite Homogeneous Turbulent Medium. Proceedings of the Royal Society, A, 210, 2 6 - 2 9 ' T h e Theory of the Fluctuations in Brightness of the Milky Way, III (with G. Miinch). The Astrophysical Journal 114, 110-22 The Structure, the Composition, and the Source of Energy of the Stars. Pp. 598-674 in Astrophysics—A Topical Symposium. New York: McGraw-Hill. ' S o m e Aspects of the Statistical Theory of Turbulence. Proceedings of the 4th Symposium of the American Mathematical Society: Fluid Dynamics, 1-17. University of Maryland. 1952 ' T h e Theory of the Fluctuations in Brightness of the Milky Way, IV (with G. Miinch). The Astrophysical Journal 115, 94-102 ' T h e Theory of the Fluctuations in Brightness of the Milky Way, V (with G. Miinch). The Astrophysical Journal 115, 103—23 The X- and Y-Functions for Isotropic Scattering, I (with D. D. Elbert and A. Franklin). The Astrophysical Journal 115, 2 4 4 - 6 8 ' O n Turbulence Caused by Thermal Instability. Philosophical Transactions of the Royal Society 244, 357-84 On the Inhibition of Convection by a Magnetic Field. Philosophical Magazine 43, 501-32 Convection under Terrestrial and Astrophysical Conditions. Publications of the Astronomical Society of the Pacific 64, 9 8 - 1 0 4 The Thermal Instability of a Fluid Sphere Heated Within. Philosophical Magazine 43, 1317-29 'A Statistical Basis for the Theory of Stellar Scintillation. Monthly Notices of the Royal Astronomical Society 112, 4 7 5 - 8 3 1953 The Onset of Convection by Thermal Instability in Spherical Shells. Philosophical Magazine 44, 2 3 3 - 4 1 ; correction, Philosophical Magazine 44, 1129-30 The Stability of Viscous Flow between Rotating Cylinders in the Presence of a Magnetic Field. Proceedings of the Royal Society, A, 216, 2 9 3 - 3 0 9 The Instability of a Layer of Fluid Heated Below and Subject to Coriolis Forces. Proceedings of the Royal Society, A, 217, 3 0 6 - 2 7 Some Aspects of the Statistical Theory of Turbulence. Proceedings of the Symposium on Applied Mathematics, The American Mathematical Society, 4, 1-17
'Magnetic Fields in Spiral Arms (with E. Fermi). The Astrophysical Journal 118, 113-15 3 Problems of Gravitational Stability in the Presence of a Magnetic Field (with E. Fermi). The Astrophysical Journal 118, 116-41 The Roots of J_(1 + i^^X-rj)J, + w (\) - J, + 1 / 2 (\T7)J_ U + 1/2) (A.) = 0 (with D. Elbert). Proceedings of the Cambridge Philosophical Society 49, 4 4 6 - 4 8 2 Shift of the l'S State of Helium (with D. Elbert and G. Herzberg). The Physical Review 91, 1172-73 Problems of Stability in Hydrodynamics and Hydromagnetics (George Darwin Lecture). Monthly Notices of the Royal Astronomical Society, 113, 6 6 7 - 7 8 1954 The Gravitational Instability of an Infinite Homogeneous Medium when Coriolis Force is Acting and a Magnetic Field is Present. The Astrophysical Journal 119, 7 - 9 3 On the Pulsation of a Star in which there is a Prevalent Magnetic Field (with D.N. Limber). The Astrophysical Journal 119, 1 0 - 1 3 "The Stability of Viscous Flow between Rotating Cylinders in the Presence of a Radial Temperature Gradient. Journal of Rational Mechanics and Analysis 3, 181-207 The Roots of Yn(\7j)Jn(X) - J n (\7))Y n (\) = 0 (with D. Elbert). Proceedings of the Cambridge Philosophical Society 50, 2 6 6 - 6 8 The Instability of a Layer of Fluid Heated Below and Subject to the Simultaneous Action of a Magnetic Field and Rotation. Proceedings of the Royal Society, A, 225, 173-84 4 On the Characteristic Value Problems in High Order Differential Equations which Arise in Studies on Hydrodynamic and Hydromagnetic Stability. American Mathematical Monthly 61 (Suppl.), 3 2 - 4 5 The Stability of Viscous Flow between Rotating Cylinders. Mathematika 1, 5 - 1 3 On the Inhibition of Convection bv a Magnetic Field, II. Philosophical Magazine 45, 1177-91 Examples of the Instability of Fluid Motion in the Presence of a Magnetic Field. Proceedings of the 5th Symposium on Applied Mathematics, 19-27 2 The Illumination and Polarization of the Sunlit Sky on Rayleigh Scattering (with D. Elbert). Transactions of the American Philosophical Society 44, 6 4 3 - 7 2 8 1955 The Character of the Equilibrium of an Incompressible Heavy Viscous Fluid of Variable Density. Proceedings of the Cambridge Philosophical Society 51, 162-78 4 The Character of the Equilibrium of an Incompressible Fluid Sphere of Variable Density and Viscosity Subject to Radial Acceleration. Quarterly Journal of Mechanics and Applied Mathematics 8, 1-21 The Instability of a Layer of Fluid Heated Below and Subject to Coriolis Forces, II (with D. Elbert). Proceedings of the Royal Society, A, 231, 198-210 2 Energies of the Ground States of He, Li + , and 0 6 + (with G. Herzberg). Physical Review 98,1050-54 The Gravitational Instability of an Infinite Homogeneous Medium when a Coriolis Acceleration is Acting. Pp. 3 4 4 - 4 7 in Vistas in Astronomy, vol. I. New York: Pergamon Press. 1956 On Force-Free Magnetic Fields. Proceedings of the National Academy of Sciences 42, 1-5 3 The Equilibrium of Magnetic Stars (with K. H. Prendergast). Proceedings of the National Academy of Sciences 42, 5 - 9 3 On Cowling's Theorem on the Impossibility of Self-Maintained Axisymmetric Homoge3
nous Dynamos (with G.E. Backus). Proceedings of the National Academy of Sciences 42, 105-09 On the Stability of the Simplest Solution of the Equations of Hydromagnetics. Proceedings of the National Academy of Sciences 42, 273—76 3 Axisymmetric Magnetic Fields and Fluid Motions. The Astrophysical Journal \24, 2 3 2 - 4 3 3 Effect of Internal Motions on the Decay of a Magnetic Field in a Fluid Conductor. The Astrophysical Journal 124, 2 4 4 - 6 5 The Instability of a Layer of Fluid Heated Below and Subject to the Simultaneous Action of a Magnetic Field and Rotation, II. Proceedings of the Royal Society, A, 237, 4 7 6 - 8 4 3 Hydromagnetic Oscillations of a Fluid Sphere with Internal Motions. The Astrophysical Journal 124, 571-79 1957 On Cosmic Magnetic Fields. Proceedings of the National Academy of Sciences 43, 24—27 "Thermal Convection (Rumford Medal Lecture). Proceedings of the Academy of Arts and Sciences 86, 323-39 On the Expansion of Functions which Satisfy Four Boundary Conditions (with W. H. Reid). Proceedings of the National Academy of Sciences 43, 5 2 1 - 2 7 3 On Force-Free Magnetic Fields (with P. C. Kendall). The Astrophysical Journal 126,457-60 "Properties of an Ionized Gas of Low Density in a Magnetic Field, III (with A. N. Kaufman and K. M. Watson). Annals of Physics 2, 4 3 5 - 7 0 On the Expansion of Functions Satisfying Four Boundary Conditions. Mathematika 4,140-45 The Thermal Instability of a Rotating Fluid Sphere Heated Within. Philosophical Magazine 2, 8 4 5 - 5 8 The Thermal Instability of a Rotating Fluid Sphere Heated Within, II. Philosophical Magazine!, 1282-84 3
1958 On Force-Free Magnetic Fields (with L. Woltjer). Proceedings of the National Academy of Sciences 44, 285-89 "The Stability of the Pinch (with A. N. Kaufman and K.M. Watson. Proceedings of the Royal Society, A, 245, 435-55 The Stability of Viscous Flow Between Rotating Cylinders. Proceedings of the Royal Society, A, 246,301-11 2 On the Continuous Absorption Coefficient of the Negative Hydrogen Ion, IV. The Astrophysical Journal 128,114-23 3 On the Equilibrium Configurations of an Incompressible Fluid with Axisymmetric Motions and Magnetic Fields. Proceedings of the National Academy of Sciences 44, 8 4 2 - 4 7 2 On the Diffuse Reflection of a Pencil of Radiation by a Plane-Parallel Atmosphere. Proceedings of the National Academy of Sciences 44, 9 3 3 - 4 0 "Adiabatic Invariants in the Motions of Charged Particles. Pp. 3 - 2 2 in The Plasma in a Magnetic Field: A Symposium on Magnetohydrodyanmics. Stanford: Stanford University Press. "Properties of an Ionized Gas of Low Density in a Magnetic Field, IV (with A. N. Kaufman and K. M. Watson). Annals of Physics 5, 1 —25 On Orthogonal Functions which Satisfy Four Boundary Conditions. Ill, Tables for Use in Fourier-Bessel-Type Expansions (with D. Elbert). The Astrophysical Journal Supplement 3 , 4 5 3 - 5 8 3
4
Variational Methods in Hydrodynamics. Proceedings of the Eighth Symposium in Applied Mathematics 8, 139-41 2 On the Continuous Absorption Coefficient of the Negative Hydrogen Ion, V (with D. Elbert). The Astrophysical Journal 128, 6 3 3 - 3 5
1959 The Oscillations of a Viscous Liquid Globe. Proceedings of the London Mathematical Society 9, 141-49 "The Thermodynamics of Thermal Instability in Liquids. Pp. 103-14 in Max-PlanckFestschrift 1958. Berlin: Veb Deutscher Verlag der Wissenschaften. 1960 The Hydrodynamic Stability of Inviscid Flow between Coaxial Cylinders. Proceedings of the National Academy of Sciences 46, 137-41 The Hydrodynamic Stability of Viscid Flow between Coaxial Cylinders. Proceedings of the National Academy of Sciences 46, 141—43 The Stability of Non-Dissipative Couette Flow in Hydromagnetics. Proceedings of the National Academy of Sciences 46, 253—57 4 The Virial Theorem in Hydromagnetics. Journal of Mathematical Analysis and Applications 1,240-52 The Stability of Inviscid Flow between Rotating Cylinders. Journal of the Indian Mathematical Society 24, 211 - 2 1 1961 The Geodesies in Godel's Universe (with J. P. Wright). Proceedings of the National Academy of Sciences 48, 341-47 2 Diffuse Reflection by a Semi-Infinite Atmosphere (with H. G. Horak). The Astrophysical Journal 134,45-56 4 Adjoint Differential Systems in the Theory of Hydrodynamic Stability. Journal of Mathematics and Mechanics 10, 683-90 The Stability of Viscous Flow between Rotating Cylinders in the Presence of a Magnetic Field, II (with D. Elbert). Proceedings of the Royal Society, A, 262, 4 4 3 - 5 4 The Stability of Viscous Flow in a Curved Channel in the Presence of a Magnetic Field (with D. Elbert and N. Lebovitz). Proceedings of the Royal Society, A, 264, 155-64 4 A Theorem on Rotating Polytropes. The Astrophysical Journal 134, 6 6 2 - 6 4 5
1962 "On Super-Potentials in the Theory of Newtonian Gravitation (with N. Lebovitz). The Astrophysical Journal 135, 238—47 "On the Oscillations and the Stability of Rotating Gaseous Masses (with N. Lebovitz). The Astrophysical Journal 135, 248-60 An Interpretation of Double Periods in j8 Canis Majoris Stars (with N. Lebovitz). The Astrophysical Journal 135, 3 0 5 - 6 The Stability of Spiral Flow between Rotating Cylinders. Proceedings of the Royal Society, A, 265, 188-95. Appendix (with L. Lee), 196-97 "The Stability of Viscous Flow between Rotating Cylinders, II (with D. Elbert). Proceedings of the Royal Society, A, 268, 145-52 4 On Superpotentials in the Theory of Newtonian Gravitation. II, Tensors of Higher Rank (with N. Lebovitz). The Astrophysical Journal 136, 1032-36 "The Potentials and the Superpotentials of Homogeneous Ellipsoids (with N. Lebovitz). The Astrophysical Journal 136, 1037-47 On the Point of Bifurcation along the Sequence of the Jacobi Ellipsoids. The Astrophysical Journal 136,1048-68
4
On the Oscillations and the Stability of Rotating Gaseous Masses. II, The Homogeneous, Compressible Model (with N. Lebovitz). The Astrophysical Journal 136, 1069-81 "On the Oscillations and the Stability of Rotating Gaseous Masses. Ill, The Distorted Polytropes (with N. Lebovitz). The Astrophysical Journal 136, 1082-1104 "On the Occurence of Multiple Frequencies and Beats in the /3 Canis Majoris Stars (with N. Lebovitz). The Astrophysical Journal 136, 1105-7 An Approach to the Theory of Equilibrium and the Stability of Rotating Masses via the Virial Theorem and its Extensions. Pp. 9 - 1 4 in Proceedings of the Fourth U. S. National Congress on Applied Mechanics. 1963 On the Stability of the Jacobi Ellipsoids (with N. Lebovitz). The Astrophysical Journal 137,1142-61 On the Oscillations of the Maclaurin Spheroid Belonging to the Third Harmonics (with N. Lebovitz). The Astrophysical Journal 137, 1162—71 The Equilibrium and the Stability of the Jeans Spheroids (with N. Lebovitz). The Astrophysical Journal 137,1172-84 "The Points of Bifurcation along the Maclaurin, the Jacobi, and the Jeans Sequences. The Astrophysical Journal 137, 1185-1202 4 Non-Radial Oscillations and Convective Instability of Gaseous Masses (with N. Lebovitz). The Astrophysical Journal 138, 185-99 5 The Virial Theorem in General Relativity in the Post-Newtonian Approximation (with G. Contopoulos). Proceedings of the National Academy of Sciences 49, c 0 8 - 1 3 "The Ellipticity of a Slowly Rotating Configuration (with P. H. Roberts). The Astrophysical Journal 138, 8 0 1 - 8 A General Variational Principle Governing the Radial and the Non-Radial Oscillations of Gaseous Masses. The Astrophysical Journal 138, 8 9 6 - 9 7 4 The Equilibrium and the Stability of the Roche Ellipsoids. The Astrophysical Journal 138, 1182-1213 1964 A General Variational Principle Governing the Radial and the Non-Radial Oscillations of Gaseous Masses. The Astrophysical Journal 139, 6 6 4 - 7 4 Otto Struve, 1897-1963. The Astrophysical Journal 139, 423 The Case for Astronomy. Proceedings of the American Philosophical Society 108, 1—6 5 Dynamical Instability of Gaseous Masses Approaching the Schwarzschild Limit in General Relativity. Physical Review Letters 12, 114-16; e r r a t u m , Physical Review Letters
4
12,437-8 5
The Dynamical Instability of the White-Dwarf Configurations Approaching the Limiting Mass (with R.F. Tooper). The Astrophysical Journal 139, 1396-98 5 The Dynamical Instability of Gaseous Masses Approaching the Schwarzschild Limit in General Relativity. The Astrophysical Journal 140, 4 1 7 - 3 3 The Equilibrium and the Stability of the Darwin Ellipsoids. The Astrophysical Journal 140,599-620 "Non-Radial Oscillations of Gaseous Masses (with N. Lebovitz). The Astrophysical Journal 140,1517-28 4 The Virial Equations of the Various Orders (chapter 1 in The Higher Order Virial Equations and Their Applications to the Equilibrium and Stability of Rotating Configurations). Lectures in Theoretical Physics, vol. 6. Boulder: University of Colorado Press On the Ellipsoidal Figures of Equilibrium of Homogeneous Masses (with N. Lebovitz). Astrophysica Norvegica 9, 323-32
1965 The Equilibrium and the Stability of the Dedekind Ellipsoids. The Astrophysical Journal 141,1043-55 Post-Newtonian Equations of Hydrodynamics and the Stability of Gaseous Masses in General Relativity. Physical Review Letters 14, 2 4 1 - 4 4 "The Stability of a Rotating Liquid Drop. Proceedings of the Royal Society, A, 286, 1-26 "The Equilibrium and the Stability of the Riemann Ellipsoids, I. The Astrophysical Journal 142,890-921 5 The Post-Newtonian Equations of Hydrodynamics in General Relativity. The Astrophysical Journal 142,1488-1512 5 The Post-Newtonian Effects of General Relativity on the Equilibrium of Uniformly Rotating Bodies. I, The Maclaurin Spheroids a n d the Virial Theorem. The Astrophysical Journal 142,1513-18 5 The Stability of Gaseous Masses for Radial and Non-Radial Oscillations in the PostNewtonian Approximation of General Relativity. The Astrophysical Journal 142, 1519-40 1966 "The Equilibrium and the Stability of the Riemann Ellipsoids. II. The Journal 145, 8 4 2 - 7 7
Astrophysical
1967 The Post-Newtonian Effects of General Relativity on the Equilibrium of Uniformly Rotating Bodies. II, The Deformed Figures of the Maclaurin Spheroids. The Astrophysical Journal 147,334-52 5 Virial Relations for Uniformly Rotating Fluid Masses in General Relativity. The Astrophysical Journal 147,383-84 5 On a Post-Galilean Transformation Appropriate to the Post-Newtonian Theory of Einstein, Infeld, and Hoffman (with G. Contopoulos). Proceedings of the Royal Society, A, 298, 123-41 5 The Post-Newtonian Effects of General Relativity on the Equilibrium of Uniformly Rotating Bodies. Ill, The Deformed Figures of the Jacobi Ellipsoids. The Astrophysical Journal 148,621-44 5 The Post-Newtonian Effects of General Relativity on the Equilibrium of Uniformly Rotating Bodies. IV, The Roche Model. The Astrophysical Journal 148, 6 4 5 - 4 9 Ellipsoidal Figures of Equilibrium—An Historical Account. Communications on Pure and Applied Mathematics 20, 251 - 6 5 5
1968 "The Pulsations and the Dynamical Stability of Gaseous Masses in Uniform Rotation (with N. R. Lebovitz). The Astrophysical Journal 152, 2 6 7 - 9 1 "The Virial Equations of the Fourth Order. The Astrophysical Journal 152, 293-304 4 A Tensor Virial-Equation for Stellar Dynamics (with E.P. Lee). Monthly Notices of the Royal Astronomical Society 139, 135-39 Astronomy in Science and Human Culture (Jawaharlal Nehru Memorial Lecture). New Delhi: Indraprastha Press, 22 pages 1969 The Richtmyer Memorial Lecture—Some Historical Notes. American Journal of Physics 37, 5 7 7 - 8 4 C o n s e r v a t i o n Laws in General Relativity and in the Post-Newtonian Approximations. The Astrophysical Journal 158, 4 5 - 5 4
5
The Second Post-Newtonian Equations of Hydrodynamics in General Relativity (with Y. Nutku). The Astrophysical Journal 158, 5 5 - 7 9 The Effect of Viscous Dissipation on the Stability of the Roche Ellipsoids. Publications of the Ramanujan Institute 1, 2 1 3 - 2 2
1970 Solutions of Two Problems in the Theory of Gravitational Radiation. Physical Review Letters 24, 6 1 1 - 1 5 ; erratum, Physical Review Letters 24, 762 5 The 2 | Post-Newtonian Equations of Hydrodynamics and Radiation Reaction in General Relativity (with F. P. Esposito). The Astrophysical Journal 160, 153-79 5 Post-Newtonian Methods and Conservation Laws. Pp. 8 1 - 1 0 8 in Relativity, ed. M. Carmeli, S. I. Fickler, and L. Witten. New York: Plenum Press The Oscillations of a Rotating Gaseous Mass in the Post-Newtonian Approximation to General Relativity. Pp. 182-95 in Quanta, ed. P.G.O. Freund, C.J. Goebel, and Y. Nambu. Chicago: University of Chicago Press The Instability of Congruent Darwin Ellipsoids, II. The Astrophysical Journal 160, 1043-48 5 The Effect of Gravitational Radiation on the Secular Stability of the Maclaurin Spheroid. The Astrophysical Journal 161, 561-69 The Evolution of the Jacobi Ellipsoid by Gravitational Radiation. The Astrophysical Journal 161, 5 7 1 - 7 8 5
1971 Criterion for the Instability of a Uniformly Rotating Configuration in General Relativity (with J. L. Friedman). Physical Review Letters 26, 1047-50 5 The Post-Newtonian Effects of General Relativity on the Equilibrium of Uniformly Rotating Bodies. V, the Deformed Figures of the Maclaurin Spheroids (Continued). The Astrophysical Journal 167, 4 4 7 - 5 3 5 The Post-Newtonian Effects of General Relativity on the Equilibrium of Uniformly Rotating Bodies. VI, The Deformed Figures of the Jacobi Ellipsoids (Continued). The Astrophysical Journal 167, 4 5 5 - 6 3 "Some Elementary Applications of the Virial Theorem to Stellar Dynamics (with D. Elbert). Monthly Notices of the Royal Astronomical Society 155, 4 3 5 - 4 7 5
1972 On the "Derivation" of Einstein's Field Equations. American Journal of Physics 40, 224-34 5 A Limiting Case of Relativistic Equilibrium (in honor of J . L . Synge). Pp. 185-99 in General Relativity, ed. L. O'Raifeartaigh. Oxford: Clarendon Press 5 On the Stability of Axisymmetric Systems to Axisymmetric Perturbations in General Relativity. I, The Equations Governing Nonstationary, Stationary, and Perturbed Systems (with J. L. Friedman). The Astrophysical Journal 175, 379-405 5 On the Stability of Axisymmetric Systems to Axisymmetric Perturbations in General Relativity. II, A Criterion for the Onset of Instability in Uniformly Rotating Configurations and the Frequency of the Fundamental Mode in Case of Slow Rotation (with J. L. Friedman). The Astrophysical Journal 176, 745-68 5 On the Stability of Axisymmetric Systems to Axisymmetric Perturbations in General Relativity. HI, Vacuum Metrics a n d Carter's Theorem (with J. L. Friedman). The Astrophysical Journal 177, 7 4 5 - 5 6 5 [Stability of Stellar Configurations in General Relativity] Proceedings at Meeting of the Royal Astronomical Society. The Observatory 92, 116-20 s
s
The Increasing Role of General Relativity in Astronomy (Halley Lecture). The Observatory 92, 160-7'4
1973 On the Stability of Axisymmetric Systems to Axisymmetric Perturbations in General Relativity. IV, Allowance for Gravitational Radiation in an Odd-Parity Mode (with J.L. Friedman). The Astrophysical Journal 181, 4 8 1 - 9 5 5 On a Criterion for the Occurrence of a Dedekind-like Point of Bifurcation along a Sequence of Axisymmetric Systems. I, Relativistic Theory of Uniformly Rotating Configurations (with J . L . Friedman). The Astrophysical Journal 185, 1-18 s On a Criterion for the Occurrence of a Dedekind-like Point of Bifurcation along a Sequence of Axisymmetric Systems. II, Newtonian Theory for Differentially Rotating Configurations (with N. R. Lebovitz). The Astrophysical Journal 183, 19-30 P. A. M. Dirac on His Seventieth Birthday. Contemporary Physics 13, 3 8 9 - 9 4 A Chapter in the Astrophysicist's View of the Universe. Pp. 3 4 - 4 4 in The Physicist's Conception of Nature, ed. J. Mehra. Dordrecht, Holland: D. Reidel 5
1974 On a Criterion for the Onset of Dynamical Instability by a Non-Axisymmetric Mode of Oscillation along a Sequence of Differentially Rotating Configurations. The Astrophysical Journal 187,169-74 The Black Hole in Astrophysics: The Origin of the Concept and Its Role. Contemporary Physics 14, 1-24 5 On the Slowly Rotating Homogeneous Masses in General Relativity (with J . C . Miller). Monthly Notices of the Royal Astronomical Society 167, 6 3 - 7 9 5 The Stability of Relativistic Systems. Pp. 6 3 - 8 1 in Gravitational Radiation and Gravitational Collapse, ed. C. DeWitt-Morette. Dordrecht, Holland: D. Reidel 5 The Stability of Stellar Masses in General Relativity. Pp. 162-65 in Proceedings of the First European Astronomical Meeting, Athens, vol. 3. New York: Springer-Verlag s The Deformed Figures of the Dedekind Ellipsoids in the Post-Newtonian Approximation to General Relativity (with D. Elbert). The Astrophysical Journal 192, 7 3 1 - 4 6 . Corrections and Amplifications to this paper in The Astrophysical Journal 220 (1978): 3 0 3 - 1 1 , are incorporated in the present version Development of General Relativity. Nature 252, 15-17 5
1975 On the Equations Governing the Perturbations of the Schwarzschild Black Hole. Proceedings of the Royal Society, A, 343, 2 8 9 - 9 8 6 The Quasi-Normal Modes of the Schwarzschild Black Hole (with S. Detweiler). Proceedings of the Royal Society, A, 344, 4 4 1 - 5 2 6 On the Equations Governing the Axisymmetric Perturbations of the Kerr Black Hole (with S. Detweiler). Proceedings of the Royal Society, A, 345, 145-67 On Coupled Second-Harmonic Oscillations of the Congruent Darwin Ellipsoids. The Astrophysical Journal 2 0 2 , 8 0 9 - 1 4 6
1976 Verifying the Theory of Relativity. Notes and Records of the Royal Society 30, 2 4 9 - 6 0 6 On a Transformation of Teukolsky's Equation and the Electromagnetic Perturbations of the Kerr Black Hole. Proceedings of the Royal Society, A, 348, 3 9 - 5 5 6 The Solution of Maxwell's Equations in Kerr Geometry. Proceedings of the Royal Society, A, 349, 1-8
6
The Solution of Dirac's Equation in Kerr Geometry. Proceedings of the Royal Society, A, 349,571-75 On the Equations Governing the Gravitational Perturbations of the Kerr Black Hole (with S. Detweiler). Proceedings of the Royal Society, A, 350, 165-74
1977 On the Reflexion and Transmission of Neutrino Waves by a Kerr Black Hole (with S. Detweiler). Proceedings of the Royal Society, A, 352, 325—38 1978 The Kerr Metric and Stationary Axisymmetric Gravitational Fields. Proceedings of the Royal Society, A, 358, 4 0 5 - 2 0 6 The Gravitational Perturbations of the Kerr Black Hole. I, The Perturbations in the Quantities which Vanish in the Stationary State. Proceedings of the Royal Society, A, 358, 4 2 1 - 3 9 6 The Gravitational Perturbations of the Kerr Black Hole. II, The Perturbations in the Quantities which are Finite in the Stationary State. Proceedings of the Royal Society, A, 3 5 8 , 4 4 1 - 6 5 5 The Deformed Figures of the Dedekind Ellipsoids in the Post-Newtonian Approximation to General Relativity: Corrections and Amplifications (with D. Elbert). The Astrophysical Journal 220, 303-13 Why Are the Stars As They Are? Pp. 1-14 in Physics and Astrophysics of Neutron Stars and Black Holes, ed. R. Giacconi and R. Ruffini. Holland: North-Holland Publishing Co. On the Linear Perturbations of the Schwarzschild and the Kerr Geometries. Pp. 5 2 8 - 3 8 in Physics and Astrophysics of Neutron Stars and Black Holes, ed. R. Giacconi and R. Ruffini. Holland: North-Holland Publishing Co. 6
1979 The Gravitational Perturbations of the Kerr Black Hole. Ill, Further Amplifications. Proceedings of the Royal Society, A, 365, 4 2 5 - 5 1 6 On the Equations Governing the Perturbations of the Reissner-Nordstrom Black Hole. Proceedings of the Royal Society, A, 365, 4 5 3 - 6 5 Einstein and General Relativity: Historical Perspectives (1978 Oppenheimer Memorial Lecture). American Journal of Physics 47, 2 1 2 - 1 7 6 On the Metric Perturbations of the Reissner-Nordstrom Black Hole (with B.C. Xanthopoulos). Proceedings of the Royal Society, A, 367, 1 — 14 An Introduction to the Theory of the Kerr Metric and Its Perturbations. Pp. 3 7 0 - 4 5 3 in General Relativity—An Einstein Centenary Survey, ed. W. Israel and S. Hawking. Cambridge: Cambridge University Press Beauty and the Quest for Beauty in Science. Physics Today 32, 2 5 - 3 0 'Einstein's General Theory of Relativity and Cosmology. In The Great Ideas Today, 9 0 - 1 3 8 . Encyclopaedia Britannica C. T. Rajagopal (1903-78) (with A. Weil). Nature 279, 358 6
6
1980 One One-Dimensional Potential Barriers Having Equal Reflexion and Transmission Coefficients. Proceedings of the Royal Society, A, 369, 4 2 5 - 3 3 Edward Arthur Milne: His Part in the Development of Modern Astrophysics (The 1979 Milne Lecture). The Quarterly Journal of the Royal Astronomical Society 21, 9 3 - 1 0 7 The Role of General Relativity in Astronomy: Retrospect and Prospect. Pp. 4 5 - 6 1 in Highlights of Astronomy, vol. 5, ed. P. A. Wayman. Dordrecht, Holland: D. Reidel
5
The General Theory of Relativity: The First Thirty Years. Contemporary Physics 2 1 , 429-49 6 The Gravitational Perturbations of the Kerr Black Hole. IV, The Completion of the Solution. Proceedings of the Royal Society, A, 372, 4 7 5 - 8 4 1981 Review: Oort and the Universe. Science 211, 273 1982 On Crossing the Cauchy Horizon of a Reissner-Nordstrom Black Hole (with J. B. Hartle). Proceedings of the Royal Society, A, 384, 301 - 1 5 6 On the Potential Barriers Surrounding the Schwarzschild Black-Hole. Pp. 120-46 in Spacetime and Geometry: The Alfred Schild Lectures, ed. R.A. Matzner and L. C. Shepley. Austin: University of Texas Press 6
1984 On the Onset of Relativistic Instability in Highly Centrally Condensed Stars (with N. R. Lebovitz). Monthly Notices of the Royal Astronomical Society 207, 1 3 P - 16P 6 On Algebraically Special Perturbations of Black Holes. Proceedings of the Royal Society, A, 392, 1-13 6 On Stars, Their Evolution and Their Stability (Nobel Lecture). Stockholm: The Nobel Foundation, 1984, 5 8 - 8 0 The General Theory of Relativity: Why "It is Probably the Most Beautiful of all Existing Theories." Journal of Astrophysics and Astronomy 5,3—11 6 On the Nutku-Halil Solution for Colliding Impulsive Gravitational Waves (with V. Ferrari), Proceedings of the Royal Society, A, 396, 5 5 - 7 4 5
1985 On Colliding Waves in the Einstein-Maxwell Theory (with B.C. Xanthopoulos). Proceedings of the Royal Society, A, 398, 2 2 3 - 5 9 6 On the Collision of Impulsive Gravitational Waves when Coupled with Fluid Motions (with B. C. Xanthopoulos). Proceedings of the Royal Society, A, 402, 3 7 - 6 5 The Pursuit of Science: Its Motivations. Current Science (India) 54, 161-69 6 Some Exact Solutions of Gravitational Waves when Coupled with Fluid Motions (with B.C. Xanthopoulos). Proceedings of the Royal Society, A, 402, 2 0 5 - 2 4 6
1986 On the Collision of Impulsive Gravitational Waves when Coupled with Null Dust (with B. C. Xanthopoulos). Proceedings of the Royal Society, A, 403, 189-98 6 A New Type of Singularity Created by Colliding Gravitational Waves (with B. C. Xanthopoulos). Proceedings of the Royal Society, A, 408, 175-208 C y l i n d r i c a l Waves in General Relativity. Proceedings of the Royal Society, A, 408, 2 0 9 - 3 2 The Aesthetic Base of the General Theory of Relativity (Karl Schwarzschild Lecture). Mitteilungen der Astronomischen Gesellschaft 67, Hamburg, 19-49 3 Marian Smoluchowski as the Founder of the Physics of Stochastic Phenomena. Pp. 2 1 28 in Polish Men of Science: Marian Smoluchowski: His Life and Scientific Work, ed. R. S. Ingarden. Warsaw: Polish Scientific Publishers 6
1987 On Colliding Waves that Develop Time-like Singularities: A New Class of Solutions of the Einstein-Maxwell Equations (with B.C. Xanthopoulos). Proceedings of the Royal Society, A, 410, 311-36 6 On the Dispersion of Cylindrical Impulsive Gravitational Waves (with V. Ferrari). Proceedings of the Royal Society, A, 412, 75—91 6
6
The Effect of Sources on Horizons that May Develop when Plane Gravitational Waves Collide (with B. C. Xanthopoulos). Proceedings of the Royal Society, A, 414, 1-30
1988 On Weyl's Solution for Space-Times with Two Commuting Killing Fields. Proceedings of the Royal Society, A, 415, 329-45 6 A Perturbation Analysis of the Bell-Szekeres Space-Time (with B.C. Xanthopoulos). Proceedings of the Royal Society, A, 420, 93-123 Massless Particles from a Perfect Fluid. Nature 333, 506 5 A Commentary on Dirac's Views on "The Excellence of General Relativity." Pp. 49-56 in Festi-Val: Festschrift for Val Telegdi, ed. K. Winter. North-Holland: Elsevier To Victor Ambartsumian on His 80th Birthday. Astrofisika 29 (Russian: in English), 7-8 6
1989 The Two-Centre Problem in General Relativity: The Scattering of Radiation by Two Extreme Reissner-Nordstrom Black-Holes. Proceedings of the Royal Society, A, 421, 227-58 6 A one-to-one Correspondence between the Static Einstein-Maxwell and Stationary Einstein-Vacuum Space-Times. Proceedings of the Royal Society, A, 423, 379-86 6 Two Black Holes Attached to Strings (with B.C. Xanthopoulos). Proceedings of the Royal Society, A, 423, 387-400 The Perception of Beauty and the Pursuit of Science. Bulletin of the American Academy of Arts and Sciences 43, 14-29 6
1990 The Flux Integral for Axisymmetric Perturbations of Static Space-Times (with V. Ferrari). Proceedings of the Royal Society, A, 428, 325-49 Science and Scientific Attitudes. Nature 344, 285—86 6 The Teukolsky-Starobinsky Constant for Arbitrary Spin. Proceedings of the Royal Society, A,430, 433-38 6 How One May Explore the Physical Content of the General Theory of Relativity. Proceedings of the Yale Symposium in Honor of the 150th anniversary of the birth of J. Willard Gibbs, The American Mathematical Society and the American Physical Society, 227-51
6
BOOKS An Introduction to the Study of Stellar Structure. Chicago: University of Chicago Press, 1939. Repr. New York: Dover, 1958, 1967. Translations in Japanese (Tokyo: Kodansha Press, Ltd., 1972) and Russian have appeared Principles of Stellar Dynamics. Chicago: University of Chicago Press, 1942. Repr. New York: Dover, 1960 Radiative Transfer. Oxford: Clarendon Press, 1950. Repr. New York: Dover, 1960. A Russian translation exists Plasma Physics: Notes Compiled by S. K. Trehan from a Course Given by S. Chandrasekhar at the University of Chicago. Chicago: The University of Chicago Press, 1960. Repr. 1962,1975 Hydrodynamic and Hydromagnetic Stability. Oxford: Clarendon Press, 1961. Repr. New York: Dover, 1970, 1981. A Russian translation exists
Ellipsoidal Figures of Equilibrium. New Haven: Yale University Press, 1969. Repr. New York: Dover, 1987. A Russian translation exists The Mathematical Theory of Black Holes. Oxford: Clarendon Press, 1983. Repr. in Russian (Moscow: Mer Press), 1986 Eddington: The Most Distinguished Astrophysicist of His Time. Cambridge: Cambridge University Press, 1983 Truth and Beauty: Aesthetics and Motivations in Science. Chicago: University of Chicago Press, 1987 Selected Papers (six volumes). Chicago: University of Chicago Press, 1989-91
A Quest for Perspectives Selected Works of S. Chandrasekhar With
Commentary I his invaluable book presents selected papers of S. Chandrasekhar, co-winner of the Nobel Prize for Physics in 1983 and a scientific giant well known for his prolific and monumental contributions to astrophysics, physics and applied mathematics. The reader will find here most of Chandrasekhar's articles that led to major developments in various areas of physics and astrophysics. There are also articles of a popular and historical nature, as well as some hitherto unpublished material based on Chandrasekhar's talks at conferences. Each section of the book contains annotations by the editor.
SBN 1-86094-285-71
P175 he ISBN 1-86094-201 -6(set)l
Imperial College Press www.icpress.co.uk
'781860"942855
781860"942013"