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!
i — 1,2,... n where
A\
(7.30) (7.30)
In this variables our equations (7.21) have the form 11
dd
((
■ aa AS A 1t ■
s
r sm 9 \o9 rsm9 \d9 — ^
= =
ddAA s
A \ dip op>) )
ll
rlrl
sm9~(rA,): *4(r^),
(7.31) (7-31)
8AT 8 -W = d-r{rAe)The most simple partial solutions of (7.31) are A. A.^ — 0, U, Arr -— A v =
Ag^Q^ 0 A«
(7.32)
A Ar r ==i 4A*9 = 0, 0J
A 4 v,^0. ^0.
(7.33)
and
In case (7.32) we have
J_dAg_ _ 1 sm9 sin# dip
(7.34)
r
l(rAg) = 0. J:(r4*)=0.
(7.35)
These equations yield resp. Ag =c-sin =c-sm9 + .F<,(r,0), Fg(r,9), + r A -Ge{tp,B). ^«g = -(?»(¥>, r Relations (7.36) and (7.37) entail
(7.36) fl).
(7.37)
ap sin 9 + rFg = Gg or Fg{r,9) -Hg{9) = ^{Gg{9,ip)-apsm9). -(Gg(9,ip) - c^sinfl). Fe(r,9) = \Hg{9) r r
(7.38)
Construction
of an autonomous
one-particle Lagrange function
...
61
Finally the partial solution reads Ag = c-sin6,
AT = A{p = 0,
(7.39)
where we dropped the solution of the homogeneous equation *) -He{9). T
In case (7.33) we have J^(sin0i4 v ) = - £ s i n 0 ,
(7.40)
j^(rAv)=0.
(7.41)
The solutions of (7.40) and (7.41) are 4 , = -cot<9, Ar=Ae=0, (7.42) r where again the homogeneous part was left out. " Let us return now to the Cartesian coordinates. In case (7.32) we have, according to (7.29), A\ = cos 6 cos ipA$ , A2 = cos 6 sin tpAg, Ax =
(7-43)
-sinOAe.
Taking into account that x2 arctan — =
sin ip —
x2
j- ,
cos
(Xl+Xl)2 X3
„
— = cos 6 , r
Xi
.
X2
— = sin 6 cos tp , r (x\ +x\)i = rsinO
*) If we replace in (7.34) and (7.35) Ag by ^He{8) c = 0.
.
xi
j- ,
{X\+Xl)2 - .
— = sin 6 sin w , r
these equations will be satisfied for
r) The case Av = Ag = 0, y4r ^ 0 leads to solutions of the homogeneous equations only and so we discard them.
62
The Inverse Variational Problem in Classical
Mechanics.
we conclude that (7.39) in Cartesian coordinates takes the shape (without the homogeneous part) a
!
A
a
(
4.
Z
2N\
X1Z3
A\ = c-u>cosu}sm(fcosO = c arctan— —— = r V X i J r3 = —c I arctan — I —— ( — ) \ %\) ax\ \T' or, generally containing the solution of the homogeneous equation,
* —(—S)S:(T) + ^
<7'">
The solution (7.39) of (7.44) is singular for r = 0 (not interesting) as well as then when ip jumps from the value 2ir to 0. Notice that this singular behaviour cannot be removed by the grad $ term. Solution (7.44) can be represented as depending on two fixed non-collinear vectors t / 1 ' and t/ 2 ) viz. Ai
=-
c arctan
/ ( b « xb<2))x\ d ( bWx \ ( |b(i)|(b(»)x) J ^ {WW\)
+
a$(x) , „ _ " f a T (7-45)
where b^ 1 ' = (0,0, u),
b ( 2 ' = (0, v, 0),
u,« - arbitrary numbers.
Let us evaluate B appearing in (7.10). According to (7.14) and (7.17) * * = -xbir) = dXl
^M dXi
(7.46)
dU
- —rb{r). ui \ —— =
dr thus B(x) = U(r)
(7.47)
where we have incorporated the constant into U(r). To summarize this case ((7.10), (7.44), (7.47)) we have L - J * "
+
c ( « - S ) | ( * ) - * < r >
+
^
.
(,48,
Construction
of an autonomous
one-particle Lagrange function
.... ..
63 63
Let us now discuss the other case, (7.33) and (7.42). Taking into account (7.42), (7.29) and the relations among x and w ip and 6 we get 2~,—2 ( — " + - # ¥ > ) > xi + + x\ xi \ \ r r 1 x\ )
M
= —
A2
— —~——2 (
1" Hf I '
xi + xi V r
/
^3 = 0 ,
(7.49)
H =H
* « (S) •
We deliberately kept here the solution —^~Z^f
r sin 0 rsmv of the homogeneous equation to reduce the singular behaviour of A. Let us us take take
H = ±c. H v=±c.
(7.50) (7.50)
v
Then Then Al
= -
C
- ^ 1 - .
rT £3 £3 = -FF rr
A2=°-^-. = °-^-.
rr X3 X3 =F =FTT '
(7.51) (7-51)
ASs = 0. A\ and A2 are singular for r = 0 as well as along the half-straight line £3 = r or X3 = —r. This solution depends on one fixed vector b . The most general form of (7.22) reads c
(x x b ) ' r |b|r Iblr ± bx '
A-
(7.52)
It is singular for r = 0 and for |b|r = =Fbx > 0. The singularity along half-straight line reminds on Dirac's monopole theory. According to (7.10), (7.47) and (7.52) the admissible Lagrange functions are L =
11 9 2
2
2X
L Lij = y j
+
c
bL . . .. . d$(x) d$(x) kL U{r) + r rr( r| bb| rr ++ bx) ^T' bx) ~ tit 1 '
y / k
c
CijkXjXk ■
(7.53) (7.53)
64
The Inverse Variational Problem in Classical
Mechanics.
Let us return to the solution (7.33) and (7.42) and discard the gauge term. We get c b x ( b x x)j X b 2 r 2 — (bx) 2
A Cr
(bx) 2 ==(b(bxxx)2 x)2 bb 2Vx 2 - (bx)2
^ r^- (bir
•.
-
(7(7.54) 54)
'
The solution (7.54) has a stronger singularity t h a n (7.52); it is singular along the whole straight line when b is parallel to x. The Lagrange function reads x 2
L
_ cc£x)bL ^x)bL_
dd ^
r (b x x ) 2
1
dt
(7.55)
The solution (7.54) was given and examined by Houard. *> The derivations of this Subsection are in many respects standard. Notice that the Euler-Lagrange Equations do not depend on b ' 1 ) or b ' 2 ) in (7.45) as well as on b of (7.53). Hence a different choice of these vectors leads to Lagrange functions, s-equivalent to the former ones and to L°. Let us return to the Equations of Motion (7.12)
=V (— - ^k) x- ^-' \ dxj dxi J 3
9U dxi
^
(7.56)
Since, according to (7.21), dAi dAj _ v—< Xk dAj _ dAj v-* £fc g^-^ dx~ dxi = -+C+cE^> 2^eijkr3 ' J
(7.57) (7-57)
k
we we insert insert (7.57) (7.57) into into (7.56) (7.56) and and get get Li _ •• - _ hi r
6
dU dU(r) ^ dxi
(7.58)
where L is the angular momentum, viz. (7.59) (7.59)
Li = ^2 X ! eHkXjXk ■ i4
kk
Notice the close relation of (7.56) and (7.58) to the expression of the Lorentz force in Classical Electrodynamics. In Electrodynamics one has the Lorentz Equation 3
mXi = Y^Fu,j",
i=l,2,3,
i/=0 i/=0
*) Mat/». PAys. *) C.S. C.S. Houard, Houard, J. J. Math. Phys. 18 18 (1977) (1977) 502. 502.
f^ = 00,,11,,22,,33,,
(7.60)
Construction
of an autonomous
one-particle Lagrange function
. .....
65
where m stands for the mass of the charged particle, exposed to exectromagnetic forces, and jo = j° -= PP JO
ji = = ~f ji ~jl
= = pii, PXi
= 1,2,3 i=
,
is the electric charge and electric current resp.. The (4 x 4) antisymmetric tensor F is F F o =i
dA (dAi _ ^l) dA0\_ dA0 _^ E-. E= l _- (fM* - ^ — - —l ) =- - — °,
\dx°
dx )
dxi
_ - y^dxi _ - dxJ _ Jj dx _ . _ - _ dx,.> ijkHk fa- - - ^e 2-e«*fl fc FlJ
(7.61)
(7.62)
where E and H are the electric end magnetic fields resp. and (Ao,A) is the the electromagnetic 4-vector potential. If we put the mass m and the electric charge p equal to one we get from (7.60)-(7.62) Fikik Fik±k
Xi = Fio-^2 Xi Fio-^2
k
= Ei { +
~®+X S3 X 5Z! eeHkXjHk Hk±iHk j
(7.63) (7.63)
k
or
*~ dxi+^\dxj
J
3 J
da)
3
'
(7.64)
To get (7.56) from (7.64) one has to put
dAp dA0
=
-T.— = dxi axi
dUQx\l dU(M) T. dxi axi
=
Ei
= Ei.
.
(7.65)
Taking into account (7.65) and defining the magnetic field by H ==- c-«£ ^f ^ k
(7.66)
we get from (7.63) and (7.59) the equation (7.58). The equations (7.66) and (7.58) show that our model has many features in common with classical model of Dirac's monopole; then (—c) would play the role of the magnetic charge (see also the relation (7.54)).
66
7.2
77»e The Inverse Variational Problem in Classical
Canonical variables. grange functions.
Mechanics.
Equivalence problem of the La-
Let us consider the classical Hamilton formalism. We have the canonical momenta (see (7.10)) _ dL _
<9$ - Ai Ai + +— axi oxi Then the Hamilton function reads
(7.67)
Pi — = TTTPi a- = — xXi i
H = J^P^i ^ p , * , -- LL == i* §x22 ++u[/(r) H= (r) ==
V
d$v 2+t/(r)
=K^-S)
The gauging of A K = A
Z
-
- -
(7(7.68) 68)
-
(7.69)
is physically irrelevant (as in Electrodynamics). The Hamilton function is conserved as dH dt
= Z.«(i< + lsr) =
=£*(*+£^)-.
(7.70)
where (7.53) was used; notice that
= e /] *»i< XiU = z2z2z2 } , / , } , HjkXjikXi yj vkXjxkxi —=0.o. i
i
j
(7.71)
k
Let us look to the Lagrange brackets of x and x. Taking into account (7.57) and (7.67) we have , -,, iXitXj) -
dAj _ dAj dAi _ ^ J
^-^ xnk \-^ c2_^eijk—,
(7.72)
k
i{**,%} - j T^EL-sx. i.x = 7=5y,
(7.73)
{«<,ij}
(7.74)
■^Xj j JJj j
=
—
0.
U-
These commutation relations are not standard.
Canonical
variables.
Equivalence
problem
of the Lagrange
67
functions. functions.
Let us compute Lagrange brackets for a rotated system. We have Z/ = i x
2
- £ ^ - £ / + !$',
(7.75)
i
where (see (7.6)) x\ = £ f e Rikxk, We have
x\ = Z ^kk Rikik, A' = A(x'), *' = $(x').
6 <"(7.76) >
^-SJ-i:(^-4 f|)^*-§£-£(4-4++*£)*■ Then, taking into account (7.13), we obtain consecutively dp dXkdp k {(xi x■}' y = YV ((^*M ' -_ ^ iMM^^ lI - dp'j'i Xi x Z^ \dx{ dxj dxj dxi J dxj Z^ y ^ i i 9XJ dx ; dxi J dxj
dA
(
' i ^
Wfl
dp dv
'i'i = = dxi dxi
= [7.77)
V ^m dx\dx>m) = 2 J 2 J a'mlRmiRlj = I
h
ml
\
TTl
i "= 2EJ E2 J( af'ml - ^RV ( * ^ ^ u - ay - ^ miRlj = I
TTl
1
1
^ E ^ -^ «*•*>' = ^? £) &
= ^ = tof ~ dx~'
= a^-Eafj^-E^^-^J
I
J
(7.78)
I
(7.79)
|{**, Xi, * Xj,* j} ' = — <), u ,
which coincides with (7.72)-(7.74). We conclude from that the Lagrange as well as the Poisson brackets stay invariant under rotations. According to (5.60) a = l(L1=L, L2 = V). The Euler-Lagrange Equations of the rotated Lagrange function L' read
^f
d2L> .
d2L' „ \
dV _
W / * * , 3+j + * * Xj^3 - g ^ O . ^£*\diidxi I a±i ax,- * dxtdxj dxr dij )) dxi dxt • j
K(7.80)
'
68
77ie The Inverse Inverse Variational Variational Problem Problem in inClassical Classical Mechanics. Mechanics.
From (7.75), taking into account (7.6), we have
*JL - T(i>-A>+^\^
=
= EN-4^U.. d2L' dii dx3
/ 0Ak &># \ ; V dx\+ dx'k dx[)
v
v
V
-
p Ul3Kki
'
(7.81)
8 L1
dx~d±~ = E ^ ^ = ^> dU d2$ . \ + Xk Hji dx dx\ 'i + dxdx) 'idxdx' '*k /) '■
d_V_ y ^ ^ Y 6A'k., I dx ^ V { ^ dx'j Xk 9a;,t ~ ~ ^, 4" * If we insert (7.81) into (7.80) we get 2
dA[ a $' \ „ „ EEE(-^ + a^J^^+^+ + E E E {--^ dx^J R^R^+ Xi+ v-v-v- f
dA v- v- v- dA 'i\ nry n n ■■ vv-- 9U 9U „ „ + E E E QJ-RvK-kiXj + Y, dZTRJi+ + Ei Ek EI Q^-^iK-kiXj +i J2 i^ r % + ° k X
'
* '
OX
5
"$, T
7RljRkiXj
=
&i't\
_
+ 1^2^1^ dx Jx i
k
I
ax
J
kaxl
V-V-V-/9A;
'
(7.82) (7.82)
= = EEE^-4J^^i+^ EEE^-4J^^i+^ + +E E ^ ^ = =o o-s^9U„
By By virtue virtue of of (7.13) (7.13) and and (7.14) (7.14) aa fci (Rx) = fci (Rx) = _ a 4 j _ 9 ^ _
v v
(dAn
dAm\
-~dx' dx'kk &:{ dx\~^ - 2 - Z . ««*»*» * » « «^- ^rma 9g^^JJ '. (Rx) =
^
- ^= - E ^ ^
(7 83) (7.83)
All Lagrange functions...
69
Thus (7.82) can be written, with help of (7.1),
E E E E E ***. ( ^ - ^ ) **,*,*,+ j
k
+
^
l
+
m
n
R k
\
R
E E ^ ~d^ ^
m
n /
=
(7.84)
which coincides with (7.12) (B(x) = £/(r)). Thus the second condition of the Theorem of Henneaux is satisfied (see Subsection 5.2 with a = 1) and we conclude that
ff-l+*a.
(7-85,
where 1/ is the rotated and L the original Lagrange function. In other words L' is equivalent to L. Notice that the Lagrange brackets (7.72)-(7.74) are independent of the constant vectors b ' 1 ' and b ' 2 ' or b. Therefore the Lagrange functions obtained by changing the values of b' 1 ), b^2) or b in the original Lagrange function are not only s-equivalent but they are just equivalent. In other words for V = L(u',v') and L = L(u,v), u' j£ u, v' ^ v, we have (7.85) too. Several authors devoted their papers to similar problems earlier then J. Cislo, P.C. Stichel and me.*' We did not here undertake the task to find all s-equivalent Lagrange functions to the one chosen at the start by us, having the structure (7.48) or (7.55). In the next Section we shall show how to accomplish this program but for a much simpler case, where the magnetic charge c = 0.
*) See e.g. E.L. Hill, Revs. Mod. Phys. 23 (1951) 253, J. - M. Levy-Leblond, Comm. Math. Phys. 12 (1969) 64, J. - M. Levy-Leblond, Am. J. Phys. 39/5 (1971) 502, M. Henneaux, Ann. Phys. N.Y. 140 (1982) 45.
70
8.1
The Inverse Variational Problem in Classical
Mechanics.
All Lagrange functions, s-equivalent to the Lagrange function L — | x 2 — C/(|x|), in (3+1) space-time dimen sions.
In this Section we are going to investigate a somewhat different problem from that in Section 7, although having many common features with it. To start with let us choose an autonomous Lagrange function whose rotationally invariant Euler-Lagrange Equations have a normal form (co incide with the Newton Equations). The most general Lagrange function of this type was given in (7.48), (7.53) or (7.55) and the corresponding Equations of Motion - in (7.63) and (7.64). In this section, however, we restrict ourselves to a special case when c = 0. Then (7.53) or (7.55) read, discarding the gauge terms, L = |x2 -
tf(|x|)
(8.1)
and the Equations of Motions (in the normal form) are Zi+Q*
=»<-/<(*) = 0 ,
i = l,2,3.
(8.2)
Notice that the Lagrange function (8.1) is now rotationally invariant. Our task*) will be to find all Lagrange functions whose Euler-Lagrange Equations are equivalent to (8.2) i.e. they give the same set of solutions as (8.2).t) We shall also investigate the problem of equivalence of the new Lagrange functions, found by us, among themselves and with respect to L, given by (8.1). The set of Equations of Motion, equivalent to (8.1) have the form
J2 ^u(x, x, t) Uj + g^'j
=0.
(8.3)
Here the matrix er has to be found. We require that dettr 7^ 0,oo
a.e.
(8.4)
for obvious reasons. *' J. Cislo,J. Lopuszariski, P.C. Stichel, Fortschr. Phys. 4 3 (1995) 733, M. Henneaux k. L.C. Shepley, J. Math Phys. 23 (1982) 2101, M. Ranada, J. Math. Phys. 32 (1991) 2764. t> N.B.: the Equations of Motion do not need to be any longer forminvariant with respect to rotations.
All Lagrange functions... J4I/ functions...
71 71
Our results in case of (3+l)-dimensional space-time for C / ( | x | )^# a xx22 ++/ 3P,, U(\x\)
a,/3 a, P - constants,
which we are going to prove below, can be summarized as follows. We show that &ij = g$ij 9$ij + G{L)LiLj = gS^ gSij + a°-, a°j,
(8.5) (8.5)
where g ^ 0 is a constant, L is defined by (7.53), viz. i-Ji — jT 7 Li = 3i
CijfcXjXk J7 j eijk XjXk
(8.6) (8.6)
k
and G is an arbitrary function, satisfying the relation G(AL) = A~ A- 33G(L),
(8.7) (8.7)
where A ^ 0 is a arbitrary constant. We have also 2 2 det<7 det
ff
2
(l+G|L|2).
(8.8) (8.8)
An important issue is the existence of a Lagrange function, corresponding to a prescribed a. So we prove that for any x )i , 4 = 4°(' jxi ,( x 4 (<xx' *, x) ) = ' )'
(8.9) (8.9)
satisfying the relations
5>°.±,= £ 4 ^ = 0 , 3
(8.10) (8.10)
3
(8.(8.11) n)
^4 = ^ i == ^ ^ >) S ijt dxk
dxj dxj
dxi dxi
dik Qifc
dij
d&{ dii
(8.12)
there exists a function L° for which d2L°
= CT0
(8.13)
The most general form of the Lagrange function L leading to (8.3) and satisfying PL 82L W7(|x|) dU{\x\) r^ *—• dii cte, dxj ^-> 9dxi i i <9i, ij
JJ
JJ
, gd22L £ . _ dl ^ =_ Q0 dij J 3 9a; dx ^*— r ' 9dii i j 9i,J
| ry
J3
J
(8.14)
72 72
The Inverse Variational Problem in Classical Mechanics.
as well as
d2lL
=
d22l
^
d&i dij dii
lJ
'
3d2L l
=_
dii dxj dij
(8.15)
dxi dxj dij
is, up to the gauge terms,
2 ^^(|x -r/(|xj))E(f-fUK + E(g-gLJ^, ^<w,)
i=9(
+
(8.16) (8-16)
where v is an arbitrary constant vector. To prove the results collected in formula (8.5)-(8.16) we start with the investigation of the Lagrange function L(x, x, t) such t h a t its Euler-Lagrange variation is of the form (8.3). In other words we demand t h a t
d dl dl v-^ , . J.. == aiAxlAX x t) X X3 dtdT -dV ^ ' ' i l dtdX--d^ Y {' 't){Xj
+ +
dU\ dx-) dx-J
is an identity. This implies the identities
d2l _ =
^y , 1
^—>diid±idxj dxj t— j
3 3
((8.17) 8J7)
d2L l
dl _y,y,
diid±idtdt
dxi dx{
d2l
dU
2 2
—'diidii dijdij dxj dxi —' o
(8.18)
Prom (8.17) and (8.4) follows From
det
(d^k)*°
ae
(8.19)
-
Consequently I is regular. We drop the ,,tilde" over L in the sequel for convenience of the reader. Equations (8.18) can be written
d dXi dii
\r^f_dILdU_ 2f1 V { dij dxjddxj ^r Xj
+
BL . \ Xj j ddxj Xj ) J
+
dL _ dL dt ~2dx~dx{ '
((8.20) 820)
Notice t h a t the ralation (8.20) remains true for every Newton's Equation Xi — JiyX.j Z) ,
All Lagrange functions. functions... ..
73
in our case
h Ji —
(8 21)
ggflxj)
= —tr
-
(8.21)
o OXi
Let us differentiate both sides of (8.20) with respect to x.Then the l.h.s. becomes symmetric with respect to the interchange of the indices k and i. This entails*' d2L
d82L
di dikk dxi
(8.22)
dxk dxi dii
From (8.22) follows e.g. d3L diidij
d3L dxk
d±k
diidxj
or, generally, d<J
dijjk jk ___ datj daij _ dxi dxk dxk
daki daki dxj
(8.23)
From (8.17) follows immediately d(7
ddjk doij _ daki_ daki ik _ doij_ d±k dii dij
(8.24)
By By virtue virtue of of (8.23) (8.23) equations equations (8.20) (8.20) can can be be also also written written dd
(^ (x-^
dL dL ,.
_ \r | _ vv -^
Xj
L
dxA^d±j
dU{\x\) dU(\x\)
a
J-z^ ^
dx.
2 L d82L
d±idt-
(8.25)
Let us differentiate both sides of (8.25) with respect to xk- The l.h.s. becomes symmetric under the interchange of the indices i and k, consequently, due to (8.22) and (8.23) we have dd22UU ^v-^ aij ^ °« dxTdx~
^
3 3
=
d2u vV-" ^ d2 U aGh3 j 4 - " dx~dx-
dxk dxj ~ ^ J J
3 3
dxi dx, J
t2^~dx7 - ~dx7 3j
(8.26) -
J
Let us once more differentiate (8.25) with respect to xk. account (8.22), we get Y ^ dakj Xj. _ v ^ ®a*3 ^ Xj
fona. (8 26)
daki
-^dx ~ 2? hdxf~dx~kd~x~i ~ot~W
Then taking into
( (8.27) Z{)
*-
J3 *) Notice that in case the force / does not depend on x relation (8.22) is always valid. This follows from (5.10), (5.43) differentiated with respect to x as well as (5.3) and (5.4).
74
The Inverse Variational Problem in Classical
Mechanics.
Notice*' that (8.17), (8.24) and (8.26) are Helmholtz's Conditions (5.3), (5.4) and (5.51). The missing piece is the equation (5.41)
££f E
&aij t
k
dx^k
, V ^ ®aV
, daV
■
n
(8.28) (8-28)
+ V5 dx: k ^ + ^9t = 0. k
expressing the fact that each entry of a is a Constant of Motion (not only the trace of a). We have only to replace the matrix er by the matrix G. The relation (8.23) follows from (8.28); we have to differentiate it with respect to xi, interchange the indices i and I and subtract so obtained expression from the former one (where i and I were not interchanged). So far every thing was true for every f (x, t), i.e. when we replace in the formulae above (—gradt/) by f. In the sequel, however, we are going to make use of the specific form of the second term in (8.2). We have «
OXJ
dxj
= ^ r r
, dr dr
/ J
W
, x
. '
3
x
(8.29) '
Hence d2U(\x\) oxk OXJ
ldf(r) r dr
(8.30)
In the next Subsection we shall investigate the case when £*0. dr The case
(8.31)
f = 0 dr leads to the equation for U(r), namely
(8.32)
d fld(/{r)\ ndU(r)\ dr \r
„ =Q
dr J
fliUir)
. \
{r^r = ~f{r))
(8.33)
or or 2 U(\x\) = ax2ax+ +f3, p,
where a and 0 a 3-dimensional a freely moving discuss the case
(8.34)
are constants. Relation (8.34) encompasses the case of harmonic oscillator" (notice that a can be < 0) and of particle in the 3-dimensional Euclidean space. We shall (8.34) in Subsection 8.3.
*' I am indebted to Dr. J. Cislo Oslo for a hint.
The case U(\x\) £/(|x|) ^ £ ax2 + fi, where a, 0 - constants. constants.
8.2
75 75
The case U(\x\) U(\x\) ^ a x 2 + 0, where a, 0 - constants.
According to (8.29) and (8.30) we have (8.35)
r x fi = f( f(.r)xj fj ) j >,
dfi 8fj dfk dfk 1 ldf(r) df(r) x x a— = OXj a— = ~~i— J + finhj = oxu r dr k j W+/W
'W"
^ =4 r *
J
(8.36) (8 36)
•
"
From (8.26) and and (8.36) follows (8.37) (8.37)
a j GjjXjXk /_j®kjXjXi . VijXjXk — = ^2 kjXjXi ■
^/ 3i
3j
us multiply both sides of (8.37) by Xk and and sum sum over over k. k. We We get Let us by Xk r22
x / ], ao~ijXj — 9gx%, X *ixi = i>
r 99 = =^2 ^2^2^jkXjXk. ^2 ajkXjXk . 33
3
(8.38) (8.38)
k k
We are going to show that g in (8.38) is a constant; *' this would mean that x is an eigenvector of a corresponding to the eigenvalue g. From (8.38) and (8.24) follows
ydaijx £— d±k
jx k]x =ydak J
= 3
^— d±i
3
8g_ 8g dii
3
or
(8.39) (8.39)
dg _ dg_ P-H = -J-xk .
dxk oxi oxk axi From (8.38) and (8.23) we obtain da daij tj
dg
E E -dx^ i
Xj+aik X +aik
or
=
dx-kXi+9Sik
3
dg a
_ dg_ X
i — Q
X
k •
& & » • dxk* = oxi Let us us differentiate (8.39) and and (8.40) with respect to xi and and xi x\ resp. 2 d g d*g „. x Xi + dxk dxi dxk oxi
dg d2g dg d'g li a~^° ~ dxi »■ axi a Xk dxk oxk oxi axi
+
dg dg ~57~°kl > axi axi
(8.40) (s 40) -
8 41 ((8.41) -
*' The proof of this conjecture was presented to me in this form by Dr. J. Cislo.
76
77ie The Inverse Variational Problem in Classical
d2g
Mechanics.
d2g
(8.42)
Relation (8.42) can be exploited in (8.41), viz. dg . _ jfo dg^_ . (/ d d22g9 _ d d22gg \\ ^ = d^^ \dxidx \dx {dxk
dxkdxij dxkdxij
d±i dxt
dxk
(8.43)
Let us multiply both sides of (8.43) by n and sum over I. We get ( &2g _ d2g \ x 2 = dg_ _ 8g_^ (_** *L.) x 2 = 0LXk - *LXi. \dxidx dx dii dx \dxidxkk dxkkdxi) dxi) dii dxkk Taking into account (8.39) we conclude from (8.44) that d2g9 dxi dxk
d2g dxk dii dxi
=
(8.44) (8.44)
(8.45)
This relation inserted in (8.43) yields d
d
9 x dxt d±6u
9 xx dx dx~kk6ti ■ If If nn > > 22 we we are are able able to to choose choose k k= — II ^ £ i. i. Then Then from from (8.46) (8.46) follows follows dg _ do ^d±i = 0, =
/Q Aa, (8.46) (8 46)
-
;8.47) (8.47)
i.e. g does not depend on x. Notice that Y ^d i
E dcnj k
\
J
k
d
(9Xi)
n
(8.48)
J
Due to (8.48) and (8.35) condition (8.28) reduces to
*-? dxk k
(8.49)
dt
k
Let us further consider the relation x"^ Y ^ d{xiOijXj)
_
•4-* ■A" dxk a xXi = EE J2J2 IiXia^ ~d^Ix3i ++ H 12^a*i HCTi ** i ~ i ^%i ++ E^ =
=
=
i*
j J
ii
jj
"
jj
x Y,YliXij-^ kEE ^xJ++229»-
i
(8.50) (8.50)
The case E/(|x|) C/(|x|) 7^ ^ a x 2 +fi, + / ? , where a, p - constants. constants.
77 77
On the other hand sr-yr^ sr-S^d(xiad[xiOijXj) JJxj)
2 d(gx 2 )
dg
22
2 . 2 . ^ ^ == ^- r^ ==^^x* + 2 ^^' 2-2tet 1
J
(8.51) 8 51
(- )
Combining (8.50) and (8.51) we get (8.52) 1
1
Let us multiply both sides of (8.49) by :Ej and Xj and sum over z and j . By virtue of (8.52) and (8.38) we get
< 8 ->
*(?&«. +£)-*?fr-30--
(8.53)
Since g does does not not depend depend on on xx we weconclude conclude from from the the identity identity (8.53) (8.53) that that
§H
OXk
(8.54)
Formulae (8.47) and (8.54) prove our assertion that g is a number, an eigenvalue of a for the eigenvector x. We are going now to show that g is also an eigenvalue of a for the eigenvector x. To this aim let us differentiate (8.38) with respect to Xk- We get
E xi jr~+ i
°lk
8 55 ^(8.55) ' ^
= gSik
'
i
We have also We have also Qg^ Q_ x 2-*<*«*'■> 2 , -dt ^x3~ = 2^ dt^3 3) 3
d(gXl)
= - ma r =U°--
(8 56) (8.56)
-
3
Let us multiply both sides of (8.49) by Xj, sum over j and use (8.56). We obtain (8.57) 33
k*
Let us further multiply (8.55) by ±k, sum over k and make use of (8.57). The result is ^"Y^OikXh V j f c Z f c -= k
g±i. fix,.
(8.58)
78
The Inverse Variational Problem in Classical
Mechanics.
This proves our conjecture. Moreover, the relation
E §■* ** = £ §^** = £ £(****) - 47^*° = ° (8'59) k
k
3
k
3
3
and (8.49) entail that ^ j T = 0-
(8-60)
In other words cr is a function of x and x only. Notice that the considerations of this Section are valid for any Euclidean space of dimension n > 2. Because of (8.38), (8.58), (8.47) and (8.54) it is convenient to introduce the notation *ij = gSij + °% •
(8.61)
Then by virtue of (8.17), (8.38) and (8.58) we have a% = ^ ,
(8.62)
E^=°-
(8-63)
E4^=°-
(8-64)
i Notice that because of (8.23) and (8.24) we have
E^»E^=0. 3
v ^ dan .
E ^=EftJ»i=°-
(8-65) (8-66)
Our next task will be to find the structure of
ij
=
G L
ii
The case fU(\x\) ( | x | ) jt ^ ao x 2 + /?, fi, w/tere where a, 0 - constants. constants.
79 79
where L is given by (7.53). Taking into account (8.62) we have G= = G(x, G ( x ,xx)).. G
a% ij = ~ GLiLj GLiLj,,
a
(8.67)
If we insert (8.67) into (8.65) and (8.66) we get dG
LkLiXj E d±~ d±~LkLiXj 3i
+GLi +GLi
e
x xx x kijXiXj kijXiXj ++GL GLkk^2 ^2 E E etij *#i i ji i ==0°
J2Ys J2Ys
3 3
i
li
i
li
or 01 v - 9G
(8.68) 8 6
E -Q-.xi = ° and and
3
v-
&G ■
v^dx- X3=0 9G . j
( - 8)
3
J
22 g^x> = °
resp..
(8.69)
res
( 8 - 69 )
P-
3 3 Let us apply the method of chracteristics to both equations, (8.68) and Let us The applyrelations the method of chracteristics to both equations, (8.68) and (8.69). (8.69). The relations —- = —^ and and -7^ = -r^-, i, 1,2,3 i,j3 ==1,2,3 (8.70) (!••/■■ ;
(JLJb I
Xi
Xj
entail the partial solutions XiXj X}Xj
XjXi — 6jj'jti-/Jt .
Hence (8.71)
G = G{L). G(L). We know already that, due to (8.67),
4 = G{L)LiLi. G{L)LiLi.
(8.72)
We recall that by virtue of (8.61) the relations (8.23) and (8.24) imply
0f& *£«,. ^ = && M = =^ = dxi
dxk axk
(8.73) (8.73)
dxj OXJ
da°i a8a°^k = da% 54 = ao| 1 Td±i F * = IT ^ dxk = d±j ftr* 9xfc dxj
, -
,
8(8.74) 74
80
The Inverse
Variational
Problem
in Classical
Mechanics.
Then from (8.72) and (8.73) follows /
/ /I
o r tlimXmLjLik 07-
+
GLj€kimxxmm + y7 ^ GLj€kim,
y ^ GCjimXfnLk
m 771
m 771
— /
/ /
nj
£lkmxm,LiLj mLiLj
—=
771 m
+ y ^ GCikm LrCikmxxm-L>j m-L>j
771
+ /
771
/
GLiCjkmxm
771
or 7/
J
/
/7
€lirn ^livaxm*rn^JJj^jlJJk
oj
1
->r *^ /7 ,. ^J^-kimxm
~r T /7 , GtjiTnXTnLi}z
771
T7l
771
m
™ =
— I^i I1 7/ ,.. / \
/Z
or
e
x
+ 7y
lkm mLj m^j
771
y
xx
— ((8.75) 8 75)
m
II
GCjkm GCjkm m m
777 771
/
To put (8.75) in a more perspicuous form let us choose kfc==2,2,
j; == l,i ,
ii == 33;;
then d8G G _ _ .
^ — I2LiXi ^ l ^ i ■^—L O JL2 OL2
SG _ ,_
.
_ . .
~ —M -L2Z2 + - ^-^—L + £L332Z: 33)H + 1{L-2X2 COL\ IJI
dG + -^-j—L f -^j-L 3Liii 3Li±i OL Oh33
x2 - 2Li±i + L3x3) - G{L2±2
= 0
or
L =0 ([12^: E | |Li^ ++ 3G3)G ^^ 1 , 1 = 0 as L x = 0. Relation (8.76) for L\i.\
(8.76) (8.76)
^7^ 0 implies
dG
(8.77)
i
This relation indicates t h a t G is a homogeneous expression with respect to L of (—3) degree, viz G(AL) = AA~33G(L) G(L)
(8.78)
for each constant A. To summarize: according to (8.61) (Jij = o"ij = g&ij gSn + G(L)LiLj, G(L)LiLj,
(8.79)
The case U{\x\) V{\x\) ^± a x 2 + /3, j3, where a, ft0 - constants. constants.
81
where g is a constant and G(L) is a homegeneous function of (-3) degree in L. What is the set of Lagrange functions, belonging to a given by (8.79)? We are going to show that for each a, which conforms to the constraits (8.61)-(8.64) and (8.73)-(8.74) there exists such a Lagrange function, satisfying equations (8.18), for which 8d22LL
_
, v (8.80)
The most general solution of (8.80) reads L =
2YIYI
cFijdiidxj + y^/9 ]P#fc(x,£)a; o-ijdiidxj k(x,t)xk fc + $(x,i) =
1
k > = |ffx \g± + + L° + ^Y. **(x, **(*, t)x t)ikk + *(x, $(x, t),
2
(8.81)
k
where L° is a certain particular solution of the equation
d'2L° l3
dii dij
(8.82)
'
where er° is given by (8.72) and (8.78). In (8.81) we made use of (8.79). We are going to show that the equation (8.82) has a solution. To this end let us introduce ad hoc three auxiliary functions L°, i = 1, 2,3, defined as follows by the equations | ^ = 4a ,
Si,
«'
3j == 1,2,3,
(8.83)
for t arbitrary but fixed. Due to (8.82) we have
|4-§&=o. ,,»_,,„ j,k =
OXk
l,2,Z.
OXj
Therefore, for a fixed i, there must exist such a function L°, which fulfills (8.83). Here we are exploiting the Converse of Poincare Lemma. Having the solution L° of (8.83) for each i we are looking for the solution of the equations fir0
^ - = L°. axi
(8.84) (8.84)
82
The Inverse Variational Problem in Classical
Mechanics.
Again it follows from (8.82) and (8.83) that dij dii and again using the converse of the Poincare Lemma we come to the conclusion that there must exist such a function L° for which (8.84) and, as a consequence of it, (8.82) are both satisfied. This accomplishes the proof of our conjecture. Next step is to examine the structure of L, given by (8.81), in more detail. It follows from (8.73) and (8.82) that 2 L° d2L° \\ = Q dd /f dW W (8.85) dii V dij dxk dxk dik dik dxj) dxj) \ dij Consequently the expression d2L° d2L° dij dxk dik dxj does not depend on x. Thus we have d2L° d2L° _ d2L° d2L° (8.86) dij dx dxk dxj dij dik dxj i = v dijdx dikdxj dijdxdx dikdxj k=v k k k k i=v where v is an arbitrary constant vector. Let us introduce the function L1 defined as follows ^
\dii
dii
(8.87)
i=vy
By virtue of (8.64) and (8.82) we have dV_ dij
_
= d22Lll
^y . d2L° . l t-r1 dij dii % dL° dl? ir~dij dxj d22L° d L° dij dij dii dii
8L° dL" ITdij dxj
dL° 8L° d±j dij
dL° 8L0 d±j dij
i=v
, % 8 88 (8.88) i=v 0
-
i=v
dL 0 lJ lJ ' dij dij dii dii ' i.e. L 1 satisfied relation (8.82). Moreover, L1 satisfies also the symmetry requirement (8.22), namely taking into account (8.88) and (8.86) we have d2Ll d2L° d82L° diidxj diidxj diidxj ■ _ (8.89) 2 (8 89) d L° d2L° dd2L2Ll l " dij dxi dij dxi . _ dij dxi
The case f/(|x|) ^ a x 2 + / 3 , where a, /3 - constants.
83
Notice that L° does not, in general, fulfill the relation (8.22) although it solves (8.82). As L 1 and L° are both particular solutions of (8.82), we may choose L1 instead of L° in (8.81). We have then L = \gy? + L 1 ( x , x ) + ^ * f e ( x , i ) i * + $ ( x , i ) .
(8.90)
Let us insert (8.90) into (8.18). We have then the identity
Z-* \ dxj
dxi)
2
^
a L°
v^
3
dt
dxi
2
d L°
= E te^-'i - E di-^
x i=v
_V-^-i.+V— +
^
a^i dfj
^ ^
V&E,
J
4 - 9 i j a±j
/a*i _ OVA . 5xi) Xj
-9t{r)xt + 2^yd
+
dxj
...
>
+5/(r)
^(891)
x+ J
i=v
9*, _ s$ _ dt dxi ~ J+
dt
U
dxi
'
This identity is linear in x therefore 89j _ d^j dxj dxi implies *< = ^ : * ( x , t )
(8.92)
and, because of (8.92),
^
= 5/(1x1)^ + ^
=^-^(1x1) +
^
implies $ = - 3 (7(|x|) + ^ + c
(8.93)
84
The Inverse Variational Problem in Classical
Mechanics.
where c is some constant. By virtue of (8.92) the term in (8.90) ^Yd x ~dx~k ~ k X kXk + ~dt ~di ~~~dt ~dt is a gauge term and can be discarded. Thus L = 59 ( | x 2 - t / ( | x | ) ) + L 11(x,x) ))
(8.94)
where we made use of (8.93). L1 is given by (8.87) where L° is an arbitrary solution of (8.82). According to (8.38) and (8.58) the constant g is an eigenvalue of a belonging at least to two eigenvectors, x and x. In view of (8.4) we must have g^O. 9^0-
(8.95)
The matrix er has three eigenvalues and eigenvectors. Two of them we know already. To find the third eigenfunction let us make the ansatz that it is proportional to L. With help of (8.79) we get 2 2 + GCL)^] GCL)^] )^. )^.
^aijL^ig
(8.96)
j3
Hence the eigenvalues of a are g and
g,
g + G(L)\L{ G(L)\L\2 (8.97)
2 2 deto- = <; <;2( (5 + G ( L ) | L | ) ^ 0
as indicated in (8.8). Let us derive from L (8.94) the Equations of Motion. We have dlL_ dj . dL° 8L° ^dii = gxt' + — d±i - — dii oxi
8d2L dii dxj
dxi
82L° dii dxj
dXi
(8.98) 8.98
k=v k=v
d2L° dx^ dxj .■ _ dxi
(8.99)
Q2 L
' a-3" = ^ *+ 4 o di~dx- = 9 ^ + a*i
dL_ _ _ dU{\x\) dUJM) dxi dxi
(8.100) yy^ , I d2L° ^— \dxjdxi £-* \dijdxi
d2L° dijdxi
.
\ . J J
(8.101) ' '
The case U(\x.\) The U(\x\) ^ a x 2 + fi, where a, fi - constants. constants.
85
Hence, taking into account (8.99)-(8.101), the Euler-Lagrange Equations read ^
d2L°
/
d2L°
\ .
„
at/(|x|) _ ^ / / d2L° L° 9dx{ ii
ad2L° L°
dij dxi ^£-? \\dxjdxi
d±j dijdxidxi
v-
o ■•
° \\ . .■_ /
3J
(8.102)
=^ + « + E4^« Since
y„.a*a ^ o gj
,103) ((8.103)
v ^ o gt/(x|) JJ
jj
3j
we may add (8.103) to (8.102) and using (8.61) obtain eventually
8 1M (8.104)
12'* i > 4(*><++*£?)-<> ^H
<- >
J
j
i.e. equations (8.3), equivalent to (8.2). Let us check equations (8.28) telling us t h a t every entry of a is a Con stant of Motion. We have
^r = = EElf^r ^r JrL§^ ^ ++GG^^ as as
£—>
dt
dLkk dt —» dL
dt at
+ GL
+ GL^^ == °o
8 105 (8.105)
(- )
dt at
k
\~~^ cv~^ ^-^ ^
dLi
i
L-.
^ ^ v -^^ ^^. 4
h.
' v ^'v ^ 1 1IdLT dU kX x = 0 = -2^2^^ i K~dV j k j
oU vU
'
8 106 (8.106)
(-
)
k
Obviously we have also d m ^-Trcr —Tra dt
= = 0
(8.107)
as well as -~Tra T r < raa=0. =0. dt at
(8.108)
86
8.3
The Inverse Variational Problem in Classical
Mechanics.
T h e case U(\x\) = a x 2 + 0.
For the potential U(\x\) = ax 2 +P,
a,/3 - constants
(8.109)
the Equations of Motion are Xi + 2axi = 0,
i = 1,2,3.
(8.110)
To find the set of s-equivalent Lagrange functions L we have to make use of the identity 92L
^
..
-^
J
£-*• dii l d£jJ J
d2L
^— dii l dxj3 3
. 3
d2L
dL _
dii dt
dxi (8.1H)
= y^gjj(x,x,f)(ij- + 2axj) j
where deter ^ 0 , o o
a.e..
(8.112)
Relation (8.111) we may also write d2L di^i-= d
(
ai
>
= a
(8
»>
„ v-^ dL
^ dL . + X3+
8L\
2
dL
-U3)
,„
^[-^w,*' ^dx- m)= 0x--
<8-114>
We have, as before, d2L dxi dij
d2L dxj dii '
dajk _ doij _ duki axi oxk oxj da
Jk d±i
=daii
dxk
=
dak1 dij
' ,.„., '
v
*'
'
From (8.26), (8.27) and (8.109) follows 2a^2dtjSjk i
=2a^2^kjSji, i
(8.118)
The case U(\x\) C/(|x|) = = a x 2 + /3. 0.
4j - dxk
3
3
dt ~ U
^i dxiJ X} J
3
87
reSp
(8.119)
-
From (8.118) follows for a ^7^ 0 that (8.120)
ffijfc = 0~ik — ^
The same follows for arbitrary a from (8.113). From (8.119) by virtue of (8.117) and (8.120), we get „2 a v— ^» du v sr-^ ^ du . du / ~K^XJJ - / 7;—XJ: - —- = 0, 4 ^ dxjJ *— dii J dt 33
J
, (8.121)
J
3
where u stands for arbitrary entry of er. Equation (8.121) we may solve applying the method of characteristics. We have dx\ dxy _ dx2 2ax\ 2ax2 2ax2
dx\ dx22 _ _dx^ dx3 _ dxi _ _dx x,\ x2 ±3 X\
dx3 dx3 2ax3
,
,„ •.„„.. (8.122)
The partial solutions are the Constants of Motion \x2 + ax2 = Hj Hj, ,
ji = 1,2,3, (8.123)
x
XkiXk ~ XiX XiXk k == Lik Lik ~= ~L>ki -Lki — 22 e-ikiLi,, kixk ~ / i t-iklLl I1
%i%k 1 ZOiXiX AOtXiXk XiXk -+k
= Ai Aikk =
— -A-ki = A-ki.■
There are 9 of them but not all of them are independent from each other, e.g. we have*) AHiHj = 2*L% AHiHj 2aL% + A% or
(8.124) 4HX1H \H H22H H33 + A12A23A31
=E HXXA\ A\33 + +
H2A231
+
H3A A\22.
Nevertheless we may write u = u(Hi,H2,H ,L,L31, ,H3,Li2,L 3,Lu,L2323 3i,
A12, A), A12, A23, A31 3i),
(8.125)
where u stands for any 0{k . We are not going to give here a general solution of the stated problem, as the variety of solutions is very large. We shall inspect two special cases where (1) u , , £ 3L3li)) U ( 1 = u(Hi, U ( # i , tHf 22, , tHf 33, , £L12, l 2 , £L23 23
*) Relations (8.124) were given by Dr. J. Cislo.
(8.126) (8.126)
88
The Inverse Variational Problem in Classical
Mechanics.
and
(8.127)
u^=u{H H2,H ).,A23,A31). u^ = uu(H Hi, l2,AH 23,A 3l12 u 3,A 3,A
The functions (8.126) and (8.127) have to satisfy the conditions (8.116) and (8.117). Let us look at
dan _ <9<7i2 da\2 dx dx\ 8x22 dxi in case u ' 1 ' we get dau .. dan
dau dan
2a— r — Z 2 - » i - f dL l7 dH 2
dH2 =
dax2 i2 2a Xi ~xir a Hi dHi
12
dL12
, .
dan
^ 3 - 5dL 7— = 23
dL23 .. dai2 . da\ dai2 +X2 dLi2 ^?—^3^7—• dLzi <9Li2 dL3i
( 8 i 2 g )
(8.128)
In a similar way from
dan dx2
_
dai dai22 d±i
follows dan ■ dan X2 T - Xl -WTT + FdL dH2 0L12 x2 adai2 . = Xl ~aIT ~ ##! dHi
dan ~ AdL T - ^ 3 == oL 23 A AdaX2 dai2 x 7^7— * + -^7—^3 -Kj— 5L dL x3 ■■ dL dL ni 2
(8.129) (8-129)
3X 3X
It turns out t h a t the relations derived from relation (8.117), like (8.129), do not carry, in the case considered by us now, any new information as compared with the relation (8.116) and we shall ignore them.
The case £/(|x|) = a x 2 + 0.
89
R e l a t i o n s (8.116) yield i n a d d i t i o n t o (8.128) in c a s e « W t h e e q u a t i o n s „
dan ~aITXi oH3 „ dai2 Za ~xTrX2 alii „ dan dH3
. dan ~X2~XT dL23 . da\2 ~ x*~x? oL\i . do\i dL23
2a
da\3
.
da\3
9#2 9(713
.
2 a ^ - r — x 3 -x2 9.H3 _
9(722
.
dH3 „
, . dau dai3 . da\3 = 2a +X2 ~AT~ ^TTXl ~a? 0L31 dH\ oL\i , . <9cri2 „ da22 . daii ^ S ^ T — = 2 a — — X i + x2 — 0L23 oHi dLi2 , . <9ci2 „ da\3 . dan 0L31 dH2 dL\i 9(713
9(723
9Li2
.
0L23
oi/i
9(713 , .
9(713
2a
9(722
»
, .
9(722
9Z/23
9(733
„
— l-EioT— = 9L23 9L31
. 9(733
201—77-2:2 - x i — 9.H2 9Ll2
. dan ~a?— > dLi3 . da22 x3 —— , dL\3 . da\3 dL23
VXI
9(733
^7TXl 9i/i
, .
, .
+
X2
9(723
9L3i
9//2
. 9(733
9(723
.
X3
9(723
. 3(723
9Li2
dL3x
9(733
.
o7 9Li2
X3
9(723
, .
9Li2 9(723
l-X3^7— = 2 a — — x 3 -x2 91*23 9ii3
— 9Z/23
/o T on\
9(733
^7— > 9L3i 9(723
9Z-23 ,
9(723
t-^iF?91/31
N o t i c e t h a t for a 7^ 0 x a n d x c a n n o t b e e x p r e s s e d i n t e r m s of t h e p a r t i a l s o l u t i o n s ( 8 . 1 2 3 ) , i n p a r t i c u l a r in t e r m s of Hi a n d L ^ = —L^i, i, k — 1 , 2 , 3 ; if t h a t w o u l d b e t h e c a s e x a n d x s h o u l d solve t h e e q u a t i o n ( 8 . 1 2 1 ) , b u t this does not occur. T h e a n s a t z t o solve e q u t i o n s (8.128) a n d (8.130) is
1
2
3
(
8
m
)
F r o m (8.113) w e g e t t h e L a g r a n g e f u n c t i o n , c o r r e s p o n d i n g t o ( 8 . 1 3 1 ) , v i z
a
ii(Hi)\±i=Ui
'
-' k
(Xi -Ui)dlLi
+
' j^k
(8.132) j
w h e r e c is c h o s e n s o t h a t t h e i n t e g r a l in (8.132) m a k e s s e n s e . T h e f u n c t i o n s / a n d g h a v e t o b e e v a l u a t e d from t h e i d e n t i t i e s ( 8 . 1 1 4 ) . W e h a v e •sr^fdfi
9fj\
■
22 (3*7 " tel 3
Xj +
. d f i d g ^ r
"* " ax7 - 2 " I > ^ + J
, +2axj /
7777-^1 dui = 0 .
(8.133)
90
The 77se Inverse Variational Problem in Classical
Mechanics.
Since dHi we have
~n OXi
= X
i
fXi dan fXi "^— Ui du dUi = / U / ~E7T i i - I Je, J '« dHi dHi ii ii == UU ii Jcc.. I
— an <Tii| i ._ C i . = an — <7ii\ii=c^ .
dan —— dUi — du ~5 i ~ dui dui ii=Ui ii=u<
If we insert (8.134) into (8.133) we o b t a i n If we insert (8.134) into (8.133) we o b t a i n y(dU_dJl\i+d_h_ £—• \dxj \dxj dxij dxij J3 dt Z-" dt —2ay^ajjXj —2g } ^crjfXj - 2ax 2axia . ian\ =Ci ii\ i.=c ._,.
'
dg
- -— = 0.
(8.134) (8.134)
(8.135) (8.135)
OXi
Since the t h e l.h.s. of the t h e identity (8.135) is linear in x we we have
dfi_ = dfl dxj
dxi
or
A - M fdxi eS
(8.136) (8,36,
and dg d2$ dg n n v~— = -^— =—2a — 2a>> anXi anXi——2axian\laXiGiA- „„++——————75«
or or
9= =~ ~aa ^2 Y ^2PijXjXi ZJ ^ i * * -2a^, ~2aJ2 i
j^i
x
xi <Jii\±±i=Ci .=c. dxi (8.137) i <7ii\ dxi++— — .. (8.137)
i
In we insert (8.136) and (8.137) into (8.132) and discard the gauge terms we get IXi;
a
'
/ /
-
;
ii(Hi)\Xi=Ui
(*« _ Ui)dui + | J]]^Vj f c *•» ; ' ; * _ i Wi
a
~ YJ Yl a3kXjXk ~2aY.J *j£ /■Xi i
Jc
'
a
*i(Hi)\ii=c, dxi = (8.138) (8-138)
' i
+ |\ YJ 2_J ajkXjik UjkXjXk - a 2 J ajkXjXk (TjkXjXk++ const. const. +
The case C/(|x|) = a x 2 + p. p.
91
where Hll = ax} H ax} + + \x\ \x\ , (see (8.123)) and a is given by (8.131). We used also the formulae
^
= -«
and - — = 22axi. aii. OXi OXi
This is the most general form of the Lagrange function, under the proviso a conforms to (8.131), which function yields as its Euler-Lagrange Equations the Equations of Motion equivalent to (8.110). This Lagrange function satisfies the condition (8.115). The Hamilton function reads H = ]T ^ZuiHi) + Y, ^2^2a k B= Eii(tfi) + S jkff{±XjX i*(i*i**
+ M+axjX i**)k).•
8 139 ((8.139) - )
In case u' 2 ' we proceed in a similar way as in case of u^1)*' . We are not going to write out all the 8 equations for a. We again make an ansatz, namely C7ij—=aijctijAij (7jj Ji-ij ,
,
aija.ij==ajictji- -constants, constants,forfori i^ ^j ,j ,
(8.140) (8.140)
while no assumption is made as far as an, i — 1,2,3, are concerned. The relevant relations, obtained after performing some simple calculations, are dan dan ~dih da33 dHi « ,
_ ~ _ ~ =
9(7 2222 _ 9(7l 9(7122 _ <9<7 9 F 7 ~ O~AT2 ~ ai2 dan _ dai3 _ dH ~ Q* l »3 « 3, -~ «dA * l3» ■
5(722 _ 9(733 _
(8.141)
9(723 _
~dlh dH3 ~ dA^ ~~ ~~Q23 ° 23 Ms" ~ ~dih ggg _ „ dan _ dan _ dan 0A ~ 9 A " a A i 3 ~ '' 8~M2 dA23 12 2 3 ~ dAl3
. _ iz =i 1,2,3. - '/'"i-
*) The case u ' 2 ' , formula (8.127), is a slightly generalized version of an example given by S. Hojman and H. Harleston, J. Math. Phys. 22 (1981), 1414.
92
The Inverse Variational Problem in Classical
Mechanics.
Hence
+ f2{Hf2(H + 2), 2),
033 = aQ3i-#i + a ot3232H H22 3iHi
+
(8.142)
f3{H3).
The Lagrange function corresponding to a given by (8.140) and (8.142) can be computed in similar way as in case u^1'. The most general Lagrange function for u^2\ when fi(Hi) = 0, i = 1,2,3, is L = - jT >J &ij(tftf 3
i^j
L = - jT >J &ij(tftf
+ iaxjx? +
8aXi±iXjXj)-
+ iaxjx? +
8aXi±iXjXj)-
(8.143)
} ij*3
The case (3) (3) The ucase = u (L 1 2 , L23, L3U
A12,A23,A31)
U(3) = W ( 3 ) ( ^ 1 2 , ^ 2 3 , i 3 1 , A 1 2 , > l 2 3 , ^ 3 l )
(8.144) (8.144)
does not give any new results. It is a special case of (8.131) where an — const. % = 1,2,3. It has to be emphasized that ansatzes (8.131) and (8.140)-(8.142) do not exhaust all possible solutions. Notice, however, that our equations are linear, so the superposition of solutions yields again a solution*' 8.4
The Hamilton formalism for the model investigated in Subsections 8.1 — 8.3. Equivalence sets of Lagrange functions.
We shall first discuss the case considered in Subsection 8.2 where tC/ /((| |xx| |))^#aaxx22++ / 3 .
(8.145)
In this case the most general form of the s-equivalent Lagrange functions is given by (8.94) 2 - f/(|x|)) ++ Li L1 ++ ^ d*J^±, M , L = gi ^-U(W) 3 (|x
*' I profited from a discussion with Dr. J. Cislo on this problem.
. 146) (8 (8.146)
93
The Hamilton formalism for the model... model...
where (see (8.87))
^{dij
dxJkJX>
(8.147)
and the last term in (8.146) is the gauge term. The canonical momentum reads
dl 8L11 5$ 5$ Pi = «r= gxi + ~^7Pi = « r - = gxi + ~^7- + + ^— -77- = axi axi oxi = OXi OXi OXi .
dLo_dlS dLo _ d!P_ d±i dii dii i = v
9a$ $ dxi '
(8.148)
where we used (8.64). Thus p is fixed up to an arbitrary gradient of a function of x. By virtue of (8.82) we may, in principle, express x as a function of x and p . The form of this interrelation depends on the shape of L1 The Hamilton function is defined as usual (we skip in the following the gauge term) (8.149)
H = = 22 ^pi±i-L. H Pi^i ~ L ■ i
To compute (8.149) we first inspect the term
Ef-sy-i1.
(8.150) (8-150)
From (8.147) and (8.64) follows
£ | ^ - L l =1-00-Ear*'-* 33
(8(8.151) 151)
"
3 J
Thus the contribution coming from L1 just vanishes. Let us go back to (8.148). From (8.146) and (8.151) we conclude that 2 F = = g(^ + H + t/(|x|)). U(\x\)). ff(ix
(8.152)
Notice that in terms of x and x H looks like a Hamilton function of a particle in an external field; but in the variables p and x its form can become complex. The Hamilton function (8.152) is conserved as
— =9l2*i{xi+tei i
J=0.
(8.153)
94
The Inverse Variational Problem in Classical
Mechanics.
It is interesting to find out whether all Lagrange functions given by (8.146) are equivalent among themselves and equivalent to (8.1). To an swer this question we have to investigate the Lagrange brackets (or Poisson brackets). We have for the Lagrange brackets Y^ fdxr dpr *-— \dxi dxj
\Xi, Xj3 J-
d2L diidxj dii dxj
dxr dxj
dpr\ dxij
d2L dijdxi dij dx{
(8.154)
'
■j^Xj, Xj y \Xi) Xj j
— —
ij — ji — jJ C?ij — O'ji — l"*is\Xi)Xjj
\Xi) Xj j \Xi) Xj j
— —
U , U ,
where we used (8.15). Since according to (8.79) Oij = gS^ gSij + G(L)LiLj, G(L)LiLj,
(8.155)
the Lagrange functions belonging to various functions G are not equivalent to each other; they are only s-equivalent among themselves. The set, equiv alent to (8.1) consists of Lagrange functions for which G = 0. In general, the equivalence sets for G ^ 0 consist of only one single Lagrange function, characterized by fixed G. Let us now investigate the case treated in Subsection 8.3. We have U(\x\) = aaxx 22 ++ / ?3 ,,
(8.156)
where a and ft are constants. The most general form of the s-equivalent Lagrange function for the case u^ (see the formalae (8.126) and (8.138)) reads Xi
/
E H °ii(H du -Y Zu(Hi)+ °ii(Hii)\ )\ii=Ui ii=Uiduii-Y,/ n( i)
-'
+
i
+
x x z J z2 aVjkijXk - az2z2 a 2 J ^2 aaJk jkXjX + 22z2z2 ikxixk ~ j k k + const.
k k
jyik jyik
k k
jjtk jjtk
where where 2 Hi = \x —-— i-2 + ax]2 , dS(z) = a(z) , . and dz Hi = |x^ + aXi , — — = a(z) and
(8.157)
The Hamilton formalism for the model...
95
We may compute the canonical momenta Pi=^-=
/ "
OXi
0u{Hi)\ii=tt.
J
dm + Y^OijXj t&
(8.158)
Then the Hamilton function reads
H = Y, Xu(Hi) + Y, 5Z tr«t(k±*if + aXiX^ i
k
( 8 - 159 )
■
i^k
For Cij = Oij
relation (8.159) reduces to (N.B
dT,
"}tHi)
= 1 => XW = Hi)
H1 = | x 2 + a x 2 The Hamilton function in (8.159) is conserved viz. dll
\—*
-rr = 2_j
dl±i
Uii
~Af
+
V - ^ ^--V
1^,1^
i
i
- / .-
a
^Xi{xk
~
\
~
+ 2axk) = 0
k^i
as —— = Xi(Xi + 2axi). We are going now to show that Lagrange functions (8.157) belonging to different <x are not equivalent to each other. We have \Xi, Xj j — U ,
{xi,Xj} = aij ,
(8.160)
\Xi, Xj j — U .
The set of Lagrange functions marked by certain choice of er, where at least one au(Hi) is not a constant, consists only of one single Lagrange function and is inequivalent to other Lagrange function of this type which belong to different a. If all an, i = 1,2,3, are constants then all Lagrange functions belonging to a^ = c ,i,j — 1,2,3, c - arbitrary constant, form an equivalent set, not equivalent to the former.
96
Tfte The Inverse
Variational
Problem
in Classical
Mechanics.
For the case u' 2 ' the Lagrange function corresponding to cr given by (8.140) and (8.142) with /;(#*) = 0, i = 1,2,3, reads (see (8.143)) L = | 22_, J z Z aa
The canonical momenta are Pk Pfc = = iy~r- = 2^aifc(|a;itx "^aiki^ikif 2 + + aaxi k^x\x 2 + 2axA 2aa;itZtZi) :a;ia;i).•
(8.162)
i^k
Then the Hamilton function reads H = 2 j 2 j Q y ( | i ; 2 a ; 2 + \ax\x2
* ?vv
+aXi±iXji (8.163) ( 8 - 163 )
22
This Hamilton function is conserved. For a free particle (a = 0) we have for the cases it' 1 ' and u' 2 ' a 1 L(2)
ii{\u\){%i
--'
'
Y. aijilx2
= I £
- "i)dUi + 2 5 Z 5Z 3 «#i i **i resp.
°iiiiiJ (8.164)
and w H # ( 1 ) =Y1 = X!
CTi ii:i; , )uidui + /^2i(TijXiXj / aii{\u "sd™?)"**! + \'Y l5ZzC > J' 2
1 JciCi
H{2) H{2)
i
IJ2J2ai^2ii2j = IJlH^i^i i^j
i
res reg
pP-
The Equations of Motion are in both cases ^CTijXj
=
0.
j
For a*11' =
/ 0 0 1\ 0 1 0 \\ 11 00 00 //
Ji
&J m
(8.165)
Examples.
97
we have, discarding the gauge terms, L{1) = \x\ + xxxz
(8.166)
and the Equations of Motion are *3=0,
aJ2=0,
3-1=0.
(8.167)
1
Notice that Z/ ) in (8.166) is not equivalent to L = \-k2 .
(8.168)
Notice also that /0 1 0\ cr = 0 0 1 , \i 0 0/ corresponding to x2 = 0,
det«r = 1
x3=0,
(8.169)
ii=0
(8.170)
(in this order) is not admissible, as cr is here not a symmetric matrix.*' In this case there does not exists a Lagrange function yielding (8.170).
8.5
Examples.
According to (8.94) the most general Lagrange function for J7(|x|) / ax 2 +/3 reads
i-rt^-tfMHE^-sLK
(817I)
Here L° is a certain particular solution of the system of equations (8.82),
c%
(8172)
Hkr where 0% = G(L)LiLj,
(8.173)
G(XL) = A" 3 G(L)
(8.174)
are given by (8.72) and (8.78). *' see R. M. Santilli, Foundations York, 1978.
of Theoretical Mechanics.
I., Springer Verlag, New
98
8.5.1
The Inverse Variational Problem in Classical
Mechanics.
Example of Henneaux and Shepley.
Let us present here the nice example of Henneaux and Shepley*' , the only example preserving rotational symmetry of the Euler-Lagrange Equations. The model is as follows 0
d2L>
_
e
[L '_J e Li = y^ / , ijkXjik
3
]
(8.175) £ - a constant.
k
Then the equations of motion should be
^
+ H)+E^W,=°
(8-176)
which are evidently forminvariant under proper rotations. To find L1 let us make the ansatz for L° namely a = x2,
L° = L°(a,b,c),
b = xx,
c = x2
(8.177)
L° is then invariant with respect to rotations. Then d2L° diidij'
_ d2L° dV*XiXi+
o9 2
2
L° dbdc{XiXj+XiXj}
+
J
(8.177)
2
Ad
+4
L° . .
Qc2
5L°
XiXj +
^
d{j .
On the other hand ^
L i L j
=
e — / T ox 3, L (L2)2
2
"ijf
' XX^XjXj H- XiXj)
(8-178) X X{Xj
X X{Xj 1
If we compare (8.177) and (8.178) we get d2L° = db2 ~
ex 2 ILI3 '
*tf
£*1
flLP
|L|3 '
2
4
5c
2
a 2 L 0 _ exx 965c ~ ILI3 ' 9c
e_ |L| '
*) M. Henneaux and L.C. Shepley, J. Math. Phys. 2 3 (1982) 2101.
(8 179)
-
99 99
Examples. From (8.179) we conclude that that L = l\££~ L°0 =
fdz
_
^
+fiab) bh f % + / ( a , b) = = -(ac -b2)l*i{ac +h2)i+f f(a, b), aa JJ zi a Z2 a
where
2 L 2 = ac a c - 6b2 = zz,,
dc efc=-d,z. = -dz. aa From (8.180) and (8.179) follows d22L° £ b22 £ L° _ 1 E 2 2 <% ~ a ( a c - 6 ) i a (ac-62)f or
(8.180)
(8.181) 2 d 9 2 // _ <%22 ~ db ~
d2f
£C £C ( aacc--6622))§I (8.182) (8.182)
^db-2 = 0 . This implies
(8.183) (8.183)
f = r{a)b s{a). r(a)b + s{a). If we insert (8.183) into (8.180) we obtain 2 2 L°=-{ac-b L°= -(ac-b )5+r{a)b )i +r(a)b + s(a). sla). a According to (8.147)
(8.184)
(8.185) We do not need to perform this construction as for v = 0 1 L £ ( a c _2 6 2 ) i . eJH = L1 = =e^ z = -{ac-b )2
(8.186) x a As direct computations show function L\ given by (8.186) satisfies indeed the condition (8.22). The Lagrange function then reads L ==g(\x2-U{x))+e^. L 9(^-U{x))+e§.
(8.187)
Thus we conclude that (8.187) is the most general rotationally invariant form of the Lagrange function. The canonical momenta are dL
(
£ x2±i — (xx)a;j\ (xx)ij\
dxi dX
V
9
/ \ ° (. e(Lxx)\ ■ , e(Lxx)\
x
lLl
X | L |
/
'
(8.188) (8-188)
100
The Inverse Variationat Variational Problem in Classical
Mechanics.
To invert (8.188) with respect to x and p let us introduce an auxiliary vector (8.189)
Ji = i- ( x xXp p) « ) j.. Ji = 99
Let us multiply both sides of (8.188) by eijiXj. We get *' ee ((LL xx xx)) xx xx // e el1 \\ Ji = x x x i - - ^ j i f : = Li ( 1 + - - r r . Hence
L
x2 L
l l j t = ( x x x ) l - -5 ^r=u
(V i + - g\MJ mj.
j m
Hence L
or
(8.190) (8.191)
L
and and
(8.190)
U = Ji j ) T |
(8.191)
<-"-'("m)'-("+$ |J| = - ( £ <7
+ 55|L|)
(8.192) (8.192)
Taking into account (8.192) and (8.191) we have L
(8.193) (8-193)
i = jM-
From (8.188) we get, exploiting (8.193), 1 / e(L x x ) A; 1 / e(J x x ) A Pw fr^-Ua | , ) = ( aiTi J) •■ T ^ -9 ^ x2|L| J " g{ > x2|J| g V x 2 |L| / 5 V x 2 |J| J Formula (8.194) presents x expressed in terms of p and x. The Hamilton function in terms of p nad x is
H = |x2 2 ++C/(|x|) c/(|x|)= =i(H) § (H)-a g_l£ +g l+ilLi L +y(W) tf=ix +£,(,*,)
(8.194) 8.194
(8.195) (8.195)
aass xx
22 __ =
1 („2 f„2 3o((JJXx X (( JJXX XX )) V 1 x )) PP ££ V \\ 2 V X 22|J|| J | ++ ((XX22)2J2 5ff2 V X ) 2 J 2 )J ~i f/ 1
gg22 \\
2
2 ITI \ 2 g g jJ| | e 2x 2 2 2 2
x x x x yJ
If If ee = = 00 we we get get the the Hamilton Hamilton function function for for the the model model (8.1). (8.1). 2 *) *) (a (a xx b) b) xx cc = = (ac)b (ac)b - (bc)a; (bc)a; (L (L xx x) x) xx xx = = -x -x2h h
as as Lx Lx = = 00
(8.196) (8-196)
Examples.
8.5.2
Example
of
101
Stichel.
Let us present another interesting example given by P.C. Stichel *' . The model is 3 4a% == £ (r(L))" ( r ( L ) ) " 3LiLj,
T(L) = [ L 2 - ( b L ) 2 ] 2 = | L x b | ,
(8.197)
where b ^ 0 is an arbitrary fixed vector and e a constant. We are going to show that
i1 dd2Tr (bx) ( b x3 Fdiia ^dij-
L = r _ 3 L i LJ
*
-
((8.198) 8198)
The prove (8.198) it is sufficient to consider the particular case b = (0,0,1).
(8.199)
Then from (8.197) follows T(L) = (L 2 + L 2 )i = [x22i:i 2 + x22a;^2 - 2(xx)i3a;3]5 2(xx)i;3a;3]5 .
(8.200)
We have dT 1 2 ^r- = -^~ = ^[Si3(^ ^[^3(x2xl i3 - (xx)i 3 ) + Xix\ ±ix\ - XiX XiX3x3xz}. 3].
(8.201)
dT Xi°> 0 ^dx-Xi=
= '
(8-202)
5 2 —1 Xi = T,
(8.203)
OXi OX{
11
Hence
J2d~.
(8.202)
i
2-dii i
'
2
and
0 £aS^ ~ * <9£* 9a:-, = ' i
J
" i 9ij d iJ, %
J
*) Private *> Privatecommunication communicationfrom fromDr. Dr.P.C. P.C.Stichel. Stichel.
(8 204 (8.204)
- )
(8.205)
102
The Inverse Variational Problem in Classical
Consequently the shape of g?,g± d2T g-gdi^x-
= =
Mechanics.
must be (8.206) (8-206)
H^)^-f{X'iQ)L^-
Let us differentiate (8.201) taken for i = 1 with respect to X\. We get V *■ =
W
1 / •
X^
2
•
~T ~ f ? ( i i a ; 3 ~ x^^3?
\2
"^
= ^L\
r2
.
(8.207) (8.207)
If we compare (8.206) with (8.207) we conclude that
/(*.*) = f| or
„
,x
(bx), 2
/(x,x) = ^
(8.208)
which accomplishes the proof of relation (8.198) We have (8.209)
L° = ^ .2 (bx) ' Since T T is homogeneous of first degree in x we have
£ a^7^ = L■ ■
8 21 (8.210)
( - °)
^-dx~Xi=L
Choosing v\ ui = — uv22 = — 0 we get, making use of (8.185) rl _ rO L L _— e^ X (Xl±l — 3
-
+ X2X2)
- X3(XJ + X%)
c^p (z? + Z2) o
o—1
(8(8.211) 211)
-
5
Notice that the second term in (8.211) satisfies (8.22). It remaines to check
a2
_ ^ _ diidxj diidxj
2
/ £ r \ == _ d* _ (JL.\ \V(bx) ( b x ) 22 // dijdxi dijdxi
((J*L) £r \
\(bx)2/ ' V(bx)V
,(8.212) 8 2U) ° l >
{
To do that it suffices to show for b = (0,0,1) v
^
d2 ((L2 + L2)i\ _ d_ ((LJ+LiyA k diidxk \ x3 ) dxi y x3 )
(8.213)
Examples.
103
We have
^ifca^{^(x2^ _ h,3(-x2xj or
+ i2a;
3- 2 ( xi ) i 3^3)H =
+ 3(xx)± 3 x 3 - 2x2x3i) x33il
xkx\ - xkx3x3 xzT
E^P^-t^'
(8.214)
Let us differentiate both sides of (8.214) by ii*. j . We have
y ±i t—1 k
=
d2 d2
LirA ((Lj + LiyA
dxk dii \
\
x3
J
+ +
/
2 2d(i33(L 2d(x (Lj1+Ll)i) + L22)i)
({Ll + Lp'A _d_ ({L* Lpi\ dx% \
\
d = 2_d_ =2
x3
l{Lj f(LJ+Ll)i\ + Ll)i\
J /
= = /Q (j-i r\ (8.215)
which accomplishes the proof. So we may instead of (8.211) use £T 1 -L U = U -= p g j Ll Lo ~ ~ (bx^ The 1 he Lagrange Lagrange function function of or the the model model is is
L = 5 5( (I Ii i2 2- -Ct 77 ((| |xx| |) )) )++££II»»
(8.216) (8.216)
(8.217)
and the the Euler-Lagrange Euler-Lagrange Equations Equations are are and
99 (xi U ++^ ^J M p ) ) ++e Y. e j n\M)~ : r O3L LiL^ D - ^ i ;= =0. 0.
(8.218)
3
It is obvious that neither the Lagrange function nor the Equations of Motion, are rotationally invariant or covariant resp. We have
99 lx ++ g (bx) 2 dx t ~ V K ~=dxdxr [ * -9w^) Pi
J
(8.219) (8 219)
-
The r.h.s. of (8.219) is not linear in x. Nevertheless, relation (8.219) can be inverted i.e. x can be expressed in terms of x and p , but the procedure is tedious and the formula complicated.
104
The Inverse Variational Problem in Classical
8.5.3
Example
Mechanics.
of Ranada.
Next two examples for the case f/(|x|) ^ a x 2 + (3 are due to Ranada** . We choose first G = ( L{L\L i ^ L2Ls3) - 1
(8.220) (8.220)
According to (8.82)
L -P^ =L\L-rrr ^diidxj L
8 221 (8.221)
(- )
1
2 3
To solve this system of equations we start with d2L° _ £1 W dx\ = TT L2L3 '■ dx{ L 2L 3 We we are going to use the tne tormuia formula
(8.222) 8222
(
)
dZ dz _ J (b2 2 - ac) 2 - b - az 1 ft = *1 l n I {b - ac)i - b - az ) (8.223) n 2 2 22 2 ./ + 2bz 26z + + c ~ 2 ( 6 - a c ) 5 \(b - ax)i J az + \ {b ax)? + b +az + az )J
for ac-b2
< 0. In our case for
d (9L° _ J±--r dii 9±i 2
ac-b
= -\x\L\
dfei ff d±1
1
*J y L L22LL3 3 ' '
< 0 and according to (8.223) we get
\ „ dL° 1 / x2(x3xx1 -x3xxi) x) ln ■5^: T ) 3) = F " = -—7—: ? : = \ ) +gi{x,x + 5i(x,a: = 2,x23,a; Dxi Xi {X1xX2 - Xix XXX2) J (8.224) /onnAs da;i ^i \\ xX33(x x 2 2) J ( 3 l ,lrJ / ^~ l \ +gi{x,x , 2,x3^ = ). = ln<^ — ) +gi{x,x2,x3). Notice that it does not matter whether we choose +(x\L\)* +(x\L\)i or — (i (a:22L!,22)5 )2 in (8.223). From (8.224) follows L° = X\X xix33
L 2 (lnL 2 - 1)
+gi±i + +P1X1 +
— L33{\nL (lnL3 3 - 1) +
X\X xxx2 2
(8.225) (8.225)
hi(-x,x ). h1(-x.,x2,x 2,x33).
From From (8.224) (8.224) and and (8.221) (8.221) follows follows 2 2 L° d 1 = | ggi d L° 1 dgx= 11 9a ± <9±2 112:2 — :i i a : 2 + ^~ 9 i 2= T L3 ■ ' . i fl . =— ax\ ax2J. Math. i 1 i 2Phys. — xix L3 ♦' M.F. Ranada, 32 2(1991)ox 2764. 2 *' M.F. Ranada, J. Math. Phys. 32 (1991) 2764.
(8.226)
8.226 '
105
Examples.
From (8.226) we infer that 01 = = 0i(x,x3). a2r0
Examining Examining aa ?? ^^
in in aa similar similar way way we we find find that that
00il ==00ii((xx))..
(8.227)
Using (8.225) and (8.221) let us inspect d2L° _ 0X2 dx\
X! 1 x2x\x Xxx —x\x — 2 2 2 xxx2
d2h _ dx\ 0X2
L2 L\L LiL33
(8.228)
Hence dh-t . dki 1 . . . , . xi -- xx33xi)-—- = — ln(xix 2 - xXix (x33xi Xi)xx2)2 ) - (x ax xX22 0x22 jf
OXj2 d±2 J x\X xix33x\ x\ + + xx22({xx22xix i + + xx33xi)x xi)x x\iix3 2 2 + +x\iix
(8.229)
+/z (x,x = +h22(x,x 3) 3 ) = 1
1 L T == — In 33 x2
1l
L s 1In — jJ.+.
x2
LL\x
g 21((x, (x, xx■33))\ ==
= — I n L i + 3+g , x 33)) {x,x 2 (2x x2 or Li(lnLi -- 1) l) + + g32(x,x + h/i2(x,x (8.230) LiflnLr 3 )x 2(x,x3)x 2 2 + 2(-K,x3)3 ).■ x2x3 In a similar way we get from (8.225), where gi and h\ are replaced by (8.227) and (8.230) resp., and from (8.221) the relation hftix = =
d2L _ J_ dx2 dx3 Lx
dg2 _ 1 dx3 Lx Lx
or or (8.231) (8.231)
02 = 02 ( x ) .
g2 = g2(x). We have We have — L 2 (lnL 2 - 1) - — L° = X3X\ X3X1
-\
X2X3 X 2X3
XX\X i X 22
L3(\nL3
- 1)+
Li(lnLi - l) + 3i(x)£i c/i(x)ii +g +322{x)± (x)x2 2 + /i 2 (x,x 3 ).
(8.232)
106
The Inverse Variational Problem in Classical
Mechanics.
f\2 rr 0 p\2
From ~=r- in a similar way as before we find a x
3
(8.233) (8.233)
h2 =33(x) + h(x). h2 = g3(x) + h(x). Since Since XiJbi — U ,
/
y XiJbi ~ u, i i
we get eventually from (8.232) and (8.233) 11
L —
11
L\ InLi
X3X1
X2X3
1
L2lnL2
X\X2 X1X2
L3\nL3+
(8.234) (8.234)
+ ^3i(x)xi + 2__/gi{x)xi +/i(x). +h(x). i
We need, however, the function L1, given by (8.185). Prom (8.234) we obtain i i
r r L L = =
i±1i , L22 I I nnL- 3-
Xi Xi
x22 . L3z I I nnL\ --
x2 x2
L3
x33 x3 x3
L\
Ii Li In-—h In-—h L2
(8.235) 8 235
L2
/• T Li\ \ (x\ LT2 x■2 LT3 x3 +, I Xi — In — L2 + x—2 In — L3 + — x3 In —Li\ + \[— — + x— In Li — + x— In L—J x i In L \xi L33 x22 Li x33 L22J Let us check relation (8.22). We have e.g.
d2L} dxidx2
i ij 1 xi xi L3 Xi ii 1 xi L L33 X!
x\x xfx22 1 X\X ±1x22 xi L\ L\ V\ vi 1 x j I<3|x=v L3\ic=v %\
d2L} dx2dxi dx^~
x\x2 L\
x2 1 x 2L 3
( "
)
x=v x=v
1 vi x xix L 3L|3x\ii=v =v
and
-_
i
V2 _*2. x2 — 1 +. v2 x2 L 3 x2 x2 L3 x2
If (8.22) is satisfied we should have If (8.22) is satisfied we should have
( ** | i2\ V
xi
1
12/^3
-(v*
\x2
V xi x2J L3 \x2 Relation (8.237) is, in deed, true. Relation (8.237) is, in deed, true.
i
x\x\ 1 1 v2 x x22 L\ L% x2 L33\\k=v
1
1 L3\k=v L3\k=v V
A) I
a;i/l3|
xi)
x = v
I/ 3 |x= v
"
""
(8.236) (8.23b)
(8.237) (8237)
Examples.
107
The Lagrange function is 2 L = g(±x g(^2-U(\x\)) -U(\x\))
+ L\L\
(8.238)
and the Euler-Lagrange Equations are
9
(.. /..
3E/(|x|)\ 3*7(1x1) \
^v ^
1
+
{** - -Or) £
r, r ,
.
L
L^LZ ^=°
(8.239) 8
^ - 239 )
■
Both, the Lagrange function as well as the Equations of Motion are not rotationally forminvariant. The canonical momentum is dL /( . Pi = 7^ rF-- == f 5f [Ui ii Pi = dxi V dxi \ 8.5.4
1 11 /7 7I 22 ., L 7,22 In - = 4 - IInn -— gxi \ I n L L43 3 g xi V L3 L3
Second example
of
\\ ) /) ,, /
i = v / i=v/
etc.. (8.240) etc.. 8.240
Randda. Ranada.
The other example of Ranada*' Ranada*) .
°~k
(8.241) (8-241)
G = j , or
IS" = kL^ d2L°
JL_
diidij
L\
*
^
(8.242)
3
Let us start with d 2 L° _ 1 3if Li
(8.243)
This entails ix2 7° +Bi(x,x2,x L° = 1ly-2^- + AAi(x, i ( x , i±2,2:3 2 , i 3 ) i)xi ; i 4- J3i(x,ia,i • 3 )3).
(8.244)
■ki
To compute Ai A\ and 7?i i?i we proceed in a standard way. We compute e.g. d2L° _ iixg d±i dx2 L\
dAi __ L2 d±2 L\
or Ai = Ax =
x3xx fdM f dLi ^ xsxt _x xx3xil 1 / -JY+M+is)A2{-x.,x3) == 1 1 — x3 J L\ x3 Lx
*) M.F. Ranada, J. Math. Phys. 32 (1991) 2764.
+
+A2(x,x3),
(8.245)
108
The Inverse Variational Problem in Classical
Mechanics.
where we use dL\ — —Xzdx2 ■ After performing these not very exciting computations we end up with L° = | ^ ^
+ y]Fl(x)±i + F(x).
(8.246)
Xo -L/l
To get L1 we choose vi = v$ = 0 , vi ^ 0. We have V —■ ^ - " 9ij i
-i—M J
2
a;? L i °
and
Hence L ^ i ^ f
(8.247)
The Lagrange function reads L = 5(Ix2-[/(|x|)) + | ^ M
(g.248)
and the Euler-Lagrange Equations are (
9 Xi
dtf(|x|)\ +
v-, 1
r
r
LiL
[ ~ ~0T) ^
Lf ^ = ° ■
(8-249)
Both, the Lagrange function and the Equations of Motion, are no longer forminvariant with respect to rotations. The canonical momentum is Pi=xi
1 L2 + ——, x3 Li
etc..
(8.250) '
In both cases III and IV, the procedure to express x in terms of x and p is theoretically possible but of no practical use.
109 109
Examples.
8.5.5
Example
of Cislo.
We wind up with an example given by J. Cislo*) concerning the case of an oscillator in (2 + 1) dimensional space-time (a membrane). For simplicity reasons we put in (8.110) 2Q = 1. 2a =
Then the Euler-Lagrange Equations read Xi + Xi=0,
Xi ~\~ %i — U ,
i£ = = 1,2. 1,2.
(8.251)
The time independent Constants of Motion are
Hi =
\{x\+xi),
H22 =
\{x \(x22+x +x22), (8.252)
L = = X\X X\X22 — — xXxXxX 22 ,
A= = x\i,2 + X\x xxx2 2■. These Constants of Motion are not independent from each other e.g. we have L2 + A2 = 4H AHXXH H22 .■
(8.253)
The symmetric (2 x 2) matrix a consists of entries which all are Constants of Motion, consequently actij ti ^
(8.254)
=aij(Hi,H aij{H uH2,A). 2,A).
The conditions (8.116) and (8.117) read dann . da 2+ dH2 M? da22 . da22 Xl . or or
m mXl
+ +
dan . dai2 daX2 . x2 . Xl = Xl + oA oHi oA ~dA m ~dAX2' a fl a da . da\ . da X2 . da22 daX22X2. + daX2 . 22X2. Xl
~dA ~dAX2 = = dlf dif22X2
[dan fdffu
dax2\ da \
. .
I'da22 (da22
daX2\ daX2 X2\
. fda22 . +, (da22
+
fdan fda
8 255 ((8.255) - )
~dA -dAXl
dax2\ da \
. .
daX2\ daX2x\
. . = ()
= x2 X2 + n x2Xl X2 + {-jA-m) {d^-^A) UHT^A {-JA-m Xl = 0^'
{-WM-X2 ) \^-^2) 2
{dH;-^A) ' + \jmr^A-)Xl = *-
*' *' Private Private communication communication from from Dr. Dr. J. J. Cislo Cislo
(8.256)
(8.256)
110
The Inverse Variational Problem in Classical
Mechanics.
The relations (8.255) and (8.256) have the structure Y±i + Z±2=0,
(8.257)
where Y and Z are Constants of Motion. Relation (8.257) entails that for nonvanishing Y and Z ±2
should also be a Constant of Motion, which is not true. We conclude that d a n _ da12
.
-dir2-^A'
(8 258)
-
dan _ da12
.
.
(8 259)
-jA-m' gq
.
'
~22 _ da12
,
.
(8 260)
■dA-dlh'
-
da22 _ da12 -dih'^A
(8 261)
-
From (8.258) and (8.261) we get, using the Converse of the Poincare Lem ma, d\ * n = g]^ .
ff22
(8-262)
<9A = ^ ,
where A = \{HUH2,A).
(8.263) From (8.260) we infer that
d\ al2 = ^
(8.264)
Equation (8.259) does not supply any new information. But relation (8.258) yields still d2X _ d2X dHi 8H2 ~ 8A2 '
(8 265
"
)
Let us introduce the euxiliary quantities B = HX+H2,
C = H1-H2.
(8.266)
IIllll
Examples.
Then (8.265) can be written as a wave equation (membrane) dd22\x dd22\\ 5d22A \ _ 2 2 22 dB ~~dJP~dC dB dA ~-°' dC2 ~~ '
(8.267) (8 267) -
where B plays the role of the time variable and A and C stand for two independent space coordinates. Every X(A,B,C) satisfying (8.267) gives rise to a matrix cr, given by (8.262) (8.264). From relations (8.262) (8.264) one can construct the Lagrange function, Euler-Lagrange Equations as well as the Hamilton function. The variety is enormous. Let us inspect a very special and simple case \A = = a{\A a{\A22 + +H HiXHHa). 2).
(8.268) (8.268)
According to (8.262) - (8.264) we have a n = aH2 -= \a(x\ \a{x\ + x\), (8.269)
o-22 = aH cr aHix = - \a{x{ \a{x\ + x\), x\), <Ti2 — (T12 — aA aA == aa(xi±2 ( i i i 2 ++ x\X x\X2)2).■ Since
w = an
(8(8.270) 270)
d2L
-
-dxj=ai1 we get we get
L = -a{x\ +xl)±l L = -a{x\ +x\)x{
+ F(xi,x ,x )±i +F(xi,x22,x22)xi
+G(xi,x ,x ). +G(xi,x2,x2).2 2
(8.271) (8.271)
From (8.271) and (8.269) we obtain From (8.271) and (8.269) we obtain d32L dF OF ..,. , „. = aa; ax22xi a;i + + -sv-^- =CTi2 at2 == a{xix22 + x^x xxx22) n. ox\ ax2 ox2 or or (8.272) (8.272)
F = axix x2 + Fi(xi,x2). F = axtx22X2 + Fi(xi,x%). The The equation equation U
XJ
i
.9
O \J
l / . O
r
)\
112
77ie The Inverse Variational Problem in Classical
Mechanics.
yields + F2{xux,x x2%)x)x2 2
G = -ax\x\
(8.273) (8.273)
+ J(xi,x J{x1,x2).2 ) • +
If we insert (8.272) and (8.273) into (8.271) we get 2 1 1 L + ^^ Fiii Fiii + + J. J. (8.274) (8.274) L = = -a(± -a{x\ 2 + + 12)2:1 x\)x\ + + axix^xxj) axix2xix2 + + jax\x\ ~ax\x\ +
i=l
The Lagrange function L should satisfy
82L
A
y, J}
£-r" 9±i A* dii <9x, dij J
3
82L
3
dL _ Q
±
J -^-r* £—•3 idii i <9x,dx«
J
J
dxidxi
(8.275)
which entails Fi
=
—
dxi
(gauge (gauge term) term)
(8.276)
and J7 = x ? x2Xnf2.. — -ja ~CLX\ 4 l z
(8.277)
J
Hence
±J — r f l X i Xo 1 7 8 * 1 X 5 " I - i flX^Xi T
+ax\X2±\X +axiX X\X22 — \ax\x\
— \ax\x\
.
(8.278)
The Equations of Motion are 2 2
^2
3
= a(xix 2 + XiX2)(xi + Xi) 4- 2 a(x, + Xi)(x 2 + x 2 ) = 0.
These equations are2 not forminvariant under = a(xix + xix \a(x\rotations. + x2)(x2 The + x2Hamilton ) = 0. func2 )(xi + Xi) + tion Thesereads equations are not forminvariant under rotations. The Hamilton function reads H = \ax\x\ + \ax\x\ + jax 2 Xi + H = \ax\x\ + \ax\x\ + jax 2 Xi + (8.280) .2.2 . (8.280) ■\-ax\X2X\X2 + 3- a x2 1 xa 2 +1-ax\x\ +axix 2 xix 2 + -ax1x2 + -ax^x^ ■
The model of Subsections 8.1 - 8.4 for n ^ 3.
113
The canonical momenta are Pi = \ax\x2
\ax\x\2 + ^aiix
+ ax\x ax\x 2x22x2,
p2 = \ax\xi \ax\x2
+ \ax\x \ax\ 2x2 + ax\X2X\
(8.281) The r.h.s. are not linear in x. 9
The model of Subsections 8.1 - 8.4 for nn ^^ 3 .
In Subsections 8.1 - 8.4 we dealt mostly with the Euclidean space n = 3 and time as an independent variable. Except for some derivations (following e.g. the formula (8.66) down to (8.78)) our considerations apply to n > 2. In particular in case U ^£ a x 2 + P 0 for n = 2 all Lagrange functions are equivalent. This can be shown as follows. Relations (8.38) and (8.58) imply that x as well as x are eigenvectors of er, belonging to the same eigenvalue g. Therefore in case n = 2 av = gl. gl.
(9.1)
This makes the ansatz (8.3) trivial. A nontrivial issue is the case n = 1. The fundamental identity for an autonomous Lagrange function, which we are going to call £/, is , d2V dU(\x\) d2V dV _ , « . « . 2« , . dxdx dx dx dx ox ox ax ox'' ox If we differentiate (9.2) with respect to x and use the notation d2 V _
((9.2) 9.2)
(9.3)
we get .da _ dU(\x\) da _ ^ - « = 0 . dx dx ox ax dx ox The method of characteristics leads to the equation dx dx dx dx
-- + + dxm m= = °° dx
(9.4)
The Inverse Variational Problem in Classical
114
Mechanics.
valid for x / 0 and ^ / 0, which integrated yields \x2 + U(\x\) = H - conts..
(9.5)
a = a{H).
(9.6)
Thus
The general solution of (9.3) reads L' = [ cr{\u2 + U(\x\)){± - u)du + # ( x ) i + $ ( x ) ,
(9.7)
where $(x) and ty(x) are functions to be found. We may write
•« = £ =
then the term
is a gauge term and can be omitted. To find $(a;) we insert (9.7) into (9.2). We get r <9cr d£/ , d® / —-—udu— ./c 3ff i = „ dx dx or, taking into account that
dU —a = 0 dx
<m dx we get dU,
.
^(---l
. i = c
5*
dt/
) - ^ - ^ = o.
(9.8)
Relation (9.8) entails a$ _ dU dx ~ ~ ff'*=e dx~
where
^lx=c = ^(|c2 + c/(N)).
(9.9)
Hence ,V(«)
CT(V)
■/V(aso)
dF + const.,
(9.10)
All autonomous
s-equivalent one-particle Lagrange functions...
115
where V = \ + U(\x\). If we insert (9.10) in to (9.7) and drop the gauge terms we get ri
rV(x)
a{\u2+
V =
U(\x\))(x-u)du-
Jc
a(V)dV.
(9.11)
JV{xo)
The Lagrange functions (9.11) for various a's are not equivalent; they are only s-equivalent. We have *
=
%
■
*
-
*
=
ox = jXa{\u2
+ U{\x\))udu + J%{\t?
+ U(\s\))^^-ds
= (9.12)
rH(x,x)
= I
a{H)dH,
J H(x0,±o)
where rccb &o a r e some constants. The derivations indicate that the one-dimensional case differs essentially from the model with n > 2.
10.1
All autonomous s-equivalent one-particle Lagrange func tions for (1+1) space-time dimensions.
In this Section we undertake the important task to find the largest set of s-equivalent Lagrange functions to a given one (which we do not, as a mat ter of fact, specify) in case of one point particle in a (l+l)-dimensional space-time. This set we are going to decompose into subsets of Lagrange functions, inequivalent to each other. The problem we intend here to tack le is important as the results obtained in this way allow us to apply the inequivalent classical Lagrange functions in the quantization procedure e.g. in Feynman's method of ,,integral over all paths". We have here an essential extension of our considerations of Section 9, concerned with the case n = 1. In Section 9 the form of the initial Lagrange function, say L°, was specified viz. L° = \x2 -
U(x).
116
The Inverse Variational Problem in Classical
Mechanics.
In this Section we do not specify the shape of the initial autonomous *' Lagrange function L. We do not need to bother whether it exists; it exists always*) If we have an equation in the normal form (10.1)
x = f(x,X,t) f(x,x,t), , X
this equation will be equivalent to the Euler-Lagrange Equation for a Lagrange function L if the identity d82L d2L . &L_d2_dL_ IL_dL cPL a^/o -e. 2 dx dxdx dxdt dx dx ' dx2 is fulfilled. Let us defferentiate (10.2) with respect to i . We get
(10.2)
dG . dG. &G dG . dG dG_dG_ dG dj_ Bf „ „ „, (10.3) dx dx dt dt dx In (10.3) the substantial time derivative of G was taken along the trajectory x = g(t,g(t,a,0), a, /?),
a, a,/3 0 - some constant parameters,
(10-4) (10.4)
which solves the equation (10.1). In (10.3) (10.5) G = 0 —. 0= (10.5) dx2 Equation (10.3) coincides with Helmholtz condition (5.41). Conditions (5.3) and (5.4) are trivially satisfied and (5.51) just vanishes. In contradistinction to n > 1 equation (10.3) has always a solution. From (10.3) we get the solution r(t,a,/3) = exp j - j ^ r (g(t,a,0), ^(t,a,0),t) r(t,a,0)=exp{-j^(g(t,a,0),^(t,a,0),tj\dt.
J dt.
(10.6)
Let us invert (10.4) as well as ±=—g(t,a,P) & = -jj.9(t,a,0)
(10.7)
with respect to a, (3 and x, x. We get the Constants of Motion a —
$(x,x,t),
0/3 == * ( !V(x,x,t). ,£,*)•
(10.8)
*' In Section 10.2 the case of Lagrange functions depending explicitly on t will be considered. t> I am obliged to Dr. J. Cislo for calling this to my attention. See e.g. F. Bolza, Lectures on the Calculus of Variations, New York, 1931.
All autonomous
s-equivalent one-particle Lagrange functions... functions...
117
Hence = a($(x,x,t),y(x,x,t))r($(x,±,t)y{x,x,t),t), a($(x,x,t),,i&(x,x,t))r($(x>±,t)iJ!{x,x,t),t),
G(x,x,t) G{x,x,t)
(10.9)
where a is an arbitrary function and $ and \1> are Constants of Motion. Integrating G in (10.9) twice with respect to x we get the desired result, i.e. the Lagrange function. We conclude that for n = 1 equation (10.1) is always equivalent to an Euler-Lagrange Equation for a certain Lagrange function. Let us return to the autonomous Lagrange function L and its EulerLagrange Equation d32L.. L ox21 dx
d2L . ox dx ox dx
d2L dx ox2z
BL dL _ „ ox dx
.n
„„ v (10.10)
a.e.
If there exist other Lagrange functions, say L', which give rise to EulerLagrange Equations equivalent to (10.10) (the same set of solutions) then we should have the identity d2V.. dx2
d2L' . dx dx dxdx
dL' _ dV dx
(d2L.. \\dx dx2 2
d2L . dxdx
dL\ dxdx) J
(10.11)
where a = a(x,x') ^ 0,oo a{x,x') T^O.OO
a.e.
(10.12)
or d2V _ dx2
d2L dx2
dx \ dx
(10.13)
J
dx \dx
J
(10.14)
Our goal is to find a as well as V Let us differentiate (10.13) with respect to x and (10.14) with respect to x. We get dad2L d3L 2 2+a2 +<J 2 d~x~dx dx dx dxd± dx dx~ da d 77— dxdx dx dx
—
d33L' V 2 2 " dxdx dx'dx'
(dL. V (ot d33L x. = 8d33L' T\\ xx. l—x-L *=— -r— x — L )+a—5 ) + a—2 2 —5 22 ■ \dx dx 8xdx dx dxdx \dx )) dx dx
{(10.15) }
(10.16); V
The Inverse Variational Problem in Classical
118
Mechanics.
If we subtract (10.16) from (10.15) the latter multiplied on both sides by x we obtain*' dadU _da_dH_ _ 8adH_dadH =0j dx dx
dx dx
(1017) (10.17)
H = H(x,x)
(10.18) = —x-L. -pr-x — L. dx ox The quantity (10.18) is the Hamilton function linked to L and is conserved by virtue of (10.10). The mothod of characteristics applied to (10.17) yields dH ., dH,. ..„_ „n -^—dx + -—- ax = dH an = 0 ox dx dx ox or a = a(H). a(H).
(10.19)
Thus, on the trajectory, a is a Constant of Motion. Let us denote d2L *—p = G(x,x).
(10.20)
Then from (10.13) we get L'V = = //
G(x,u){±-u)du+^(x), G(x,u)(x-u)du+%(x),
(10.21) (10.21)
where the constant*' c is so chosen that the integral on the r.h.s of (10.21) does not diverge. In (10.21) we discarded the gauge term. L' given by (10.21) should still satisfy (10.14); this allows us to fix the arbitrary function *$>{x). We have
^(S^ L ) = £(f G ( a : ' u ) t i d u - $ )Because of (10.18) and since dH _ d2L . 2 dx dx *' Notice that (10.17) is one of the Helmholtz conditions. t) We could assume that c = c(x), but we discard this possibility.
(10.22)
All autonomous
s-equivalent one-particle Lagrange functions.. functions....
119
we get from (10.22) taking into account (10.20) dH
8 ( /■* ,TT,dH
\
*& = *(/. ^ftrL*-*;
, (10.23) (io 23)
-
To make thing easier let us introduce ad hoc the function To make thing easier let us introduce ad hoc the function (10.24)
•o-(z) « = . -±1 ^ - .
(10 . 24)
Then (10.23) reads Then (10.23) reads
^S ->£= < U=Lt -. *- 1 |(= E << *f l>, --SS( fH ) f)U or or (10.25) (10.25)
**(«) (i) = const.. = -- S S (( fHf )) || ii ==cc + + const.. If If we we insert insert (10.25) (10.25) into into (10.21) (10.21) and and use use (10.20) (10.20) we we get get
L
'=f f _>-«>*-=
.
x u
~
(10.26)
f G(*,u)«ki-E(Jff). G(x,u)du-T,(H). = i /" This is the most general form of an autonomous Lagrange function, sequivalent to L. It is interesting that the Lagrange functions V are not equivalent to each other and to L. To see that we have to check the Lagrange brackets of x and x. We have dL g , =dx->
(10.27) (10-27)
, S P
p'== / G(x,u)du + xG{x,x) - —Tx — - i = p' xG{x,x)-—-— = / G(x,u)du (10.28) 7C off Ci-* "i-* yc and consequently <92L
(10.29)
{X,±}L = ^ ,
{x,x}LL'< = G{x,x) G(x,x) = a{H){x,x} a(H){x,x}LL {x,x}L[ — {x,x} L> = —{x,x} LI L> cr cr22
,
(10.30) (10.31)
120
The Inverse Variational Problem in Classical
Mechanics.
Since a is not, in general, a constant our conjecture is correct. For the example discussed in Section 9 we have {x,x}LLoo {x,±}
= =
1,
(10.32)
{x,x}LLI,
= =
a{H). . a{H)
(10.33)
Notice that the only quantity which chracterizes the various Lagrange functions (10.26) is cr(H). Thus each equivalence set consists of one single Lagrange function, provided a is not a constant. The Hamilton function is rH{x,x) rH{x,x)
H' = - p'x - V = = H(ff) T,(H) + const. = = /
JJH(xo,±o) H(xo,xo)
a(z)dz ,
(10.34)
where xo, in are some constants. Here we used (10.28) and (10.26). Our L' of (10.26) coincides with the Lagrange function obtained by the method of Engels*) using (10.13), (10.14), (10.18) and (10.19). The inverse problem for the case of dimension (1 + 1) was treated to large extent by many scientists, to mention only Darboux, Jacobi, Hirsch, Hamel or Kiirschak.t) 10.2
All s-equivalent one-particle Lagrange functions for (1+1) space-time dimensions.
As in Subsection 10.1 we do not specify the shape of the initial Lagrange function L, we drop only the restriction of autonomy. Thus we take for granted the existence of*' , L = L(x, L(x,x,t). i , t).
(10.35)
If there exist other Lagrange functions, say L', which give rise to EulerLagrange Equations equivalent to those of L, then we should have d2L' d2L „. -dx^^a-dx^^G{x^t)
. .,
d2L , ^ 2# T0 dx
a.e.
(10.36)
*) E. Engels, II Nuovo Cimento 2 6 B (1975) 481. *' " G. Darboux, Lecons sur la Thiorie Generate de Surface. Paris, 1894, C.G.J Jacobi, Zur Theorie der Variationsrechung und der Differentialgleichungen, Works vol 4, A. Hirsch. Math. Ann. 49 (1897) 49, G. Hamel, Math. Ann. 57 (1903) 231, J. Kiirschak, Math. Ann. 60 (1905) 157. t ' Of course L does not need in general to depend explicitly on t.
All s-eguivalent s-equivalent
one-particle Lagrange functions for (1+1) space-time dimensions.
d2V . dHJ__8L[_ 9HJ__dL[_ ft d2L d2 L x + dxdxX+x+ dxdt dx ~aa\dxdxX+dxdt
dxdx didt
dx ~ \d±dx
dxdt
8L\ dL\ dx J '
dx) '
121
(10.37) (10 37)
'
where where a(x,x,t) a{x,x,t)
/£ 00
a.e.. a.e..
(10.38) (10.38)
We showed in Subsection 5.3 that a is a Constant of Motion*' From (10.36) we get assuming that we already evaluated cr L' — u)du ++ ip(x,t)x ip(x,t)x ++ ip{x,t), ip(x,t), V = = // G(x,u,t)(x G(x,u,t)(x-u)du
(10.39) (10.39)
where c is a constant and tp and ip are here still arbitrary functions. Let us insert (10.39) into (10.37). Then I* (dG dG\ dip dip r fdG dG\ , (hp_dip__ u + Jc \dt +Udx) dt ds~ 22 22 d LL dL\ dL\ = ^ f dd LL . | d \dxdx dxdt dx J
[(10.40)
'
Since dp dip dt
dip dx
do not depend on x , we may put in (10.40) x = c. Then (10.40) reduces to dip dt
dtp dx
=
L (
d2 L \dxdx
|
d2L dxdt
dL\] dx J
i i=c =c
(10.41)
The most general solution to (10.41) reads dR(x,t) V^^x—' dRjx,t) dR(x,t) dt
frxx\ I" /f d22L Jc, [ \dxdx
2 ddH LL_^A,] dL\} ,, dxdt dx J *==
,
(10.42) (10 42) '
where is aa constant constant and and R R is is an an arbitrary arbitrary function. function. If If we we insert (10.42) where c' c is insert (10.42) into (10.39) and discard the gauge term we get into f 10.39) and discard the gauge term we get rX rX
V =
Jc Jc
p-x px
G{x,u,t){x-u)du+ G(x,u,t){x-u)du
+
Jc1
Gf\*=
*> D.G. Currie & E.J. Saleton, J. Math. Phys. 7 (1966) 967.
(10.43) (10.43)
122
The Inverse Variational Problem in Classical
Mechanics.
where x = f{x,x,t)
(10.44)
is the Newton Equation, obtained from the Euler-Lagrange Equations of L. The Hamilton function H reads H=
I G{x,u,t)uduf G({,c,t)f(£,c,t)d£. (10.45) Jc J c' Notice that in case L depends explicitly on t the Hamilton function H is no longer a Constant of Motion. So a is no longer a function of H, although it is still conserved. The commutation relations are similar to those in Subsection 10.1, so the Lagrange functions V are not equivalent to each other and to L. 11.1
Construction of the most general autonomous oneparticle Lagrange function in (3+1) space-time dimen sions giving rise to rotationally covariant Euler-Lagrange Equations.*^
In Subsections 7.1, 7.2 we investigated a similar problem: it was a special case, when the Euler-Lagrange Equations coincided with the Newton Equa tions. The study of the general case turns out to be much more complicated. It is simple for one particle in (1+1) dimensional space-time. Notice that even more complicated is the problem to find all Lagrange functions, s-equivalent to one chosen by us and having the structures of Lagrange function studied in this Section. The latter problem was solved in Sections 8.1 8,5, alas, for a very special case of a Lagrange function of a simple structure L = | x 2 - £/(|x|).
(11.1)
Let us formulate first more precisely our problem. We assume the ex istence of an autonomous Lagrange function*) for one point particle in a *) The considerations of Section 11 were presented to me by Dr. J. Cisto. Unfortunatelly, the main result, the formulae (11.8) and (11.9), are already contained in, J.C. Houard, J. Math. Phys. 18 (1977) 502 although without clear derivation. t> We could start with a Lagrange function depending explicitly on time; this would not cause much trouble. For clarity reasons, however, we refrain from doing this.
Construction
one-particle Lagrange function... function... 123
of the most general autonomous
(3 + l)-dimensional space-time. Its Euler-Lagrange Equations should be rotationally forminvariant. These equations read _ y^ ^ d2L ., ~2^ ~ 2-* dii dx, dxj XjXj
n Tl
+ +
J
3
d2 L , 8L dL Xj d±i dx* dx3 Xj~8x~' ~dx~'
^ ^
((11.2) 1L2)
J
3
where L(x, x) is the Lagrange function we are searching for. The forminvariance of (11.2) under proper rotations S € 50(3) means that ri(Sx,Sx,Sx) ri(Sx,Sx, Sx) = = ^ 5 i f c r^2S x,x). f c (ikxrk, (x,x,x).
(11-3) (11.3)
k
As x appears on the r.h.s of (11.2) linearly, condition (11.3) is equivalent to the following two conditions
Gi3(Sx,Sx) = J2J2hi7nbjn( mn\X->' -) J
m m
x
j
(11.4) (11.4)
nn
Gi(Sx, Sx) = ^2 J2 SimGm{x, (x,
xx), ),
(11.5)
m
where Ci 13 3
_ d2L ~ d±i dii dxj
(11.6) d2L
dL
y>
dxi
z—1 diidxj 33
, 3
(11.7)
JJ
Our procedure in solving the stated problem is as follows. We are going first to find all functions which satisfy condition (11-4), and consequently choose out of this set of functions those which conform to the condition (11.5). Let us summarize the results we were able to derive. The most general autononomous Lagrange function giving rise to rotationally forminvariant Euler-Lagrange Equations reads up to to gauge terms L= K i r ( x 2 , x x , x 2 ) + ^ i i ^ -[ [W W((xx22))PP((xx|| xx ,, b ) ] ,
(11.8)
i
where K and W are arbitrary functions, b ^ 0 is an arbitrary constant vector and 1 |x| EYlAb x)iXi , f W , b xxx) n , ,. ,> {Xi P (11.9) P ((xxx|, xb ), b=) —rarctan ,\x\R a r c t a n [E ^,i xXYx\sbbMj ,xXJ X) ; X.,4 jJ .■ (11.9)
124
The Inverse Variational Problem in Classical
Mechanics.
The Lagrange function (11.8) yields the following Euler-Lagrange Equations ^ d2K .. *-? d±i d±j J
^ d2K . _ dK ^—> dii dxj 3 dxi
-(2x2w - w) y A [ £ - ^ U + v
+
' *->dxj [ | x x x | 2 ] 2^W'
+
4(TX3X)J
W"
J
(11.10)
^ | | ,
where da The Lagrange function, given by (11.8), depends, besides x and x, also on one additional constant vector b. Of course, the Euler-Lagrange Equa tions do not depend on this additional parameter as this would break the rotational symmetry of these equations. We shall show that the Lagrange functions are weakly invariant with respect to a change of b. Thus all the Lagrange functions for different b's are equivalent. It is easy to see from (11.8) that the Lagrange functions are not, in general, rotationally, invariant. We show, however, that they are weakly invariant under rotational transformations. In other words they are all equivalent. 11.2
Evaluation of the function Gij.
First we are going to investigate the structure of G^. As we required con dition (11.4) to be fulfilled, it can be shown that he most general conjecture for Gij symmetrical in the indices i and j (due to (11.6)) is G^ = K^Sij
+ K{2)XiXj + K^iiXj
+ Kw(xiXj
+ xiXj)+ (11.11)
+K{-h){xiLj
+ XjLi) + K^\±iLj
+ XjLi),
where Li = ^ T ]T] UjkXjXk = (x x x)j i k
(11-12)
Evaluation of the function G ; J .
125
and the functions K^\ ... K^ depend only on the variables
p = '£iXiii=xx,
(11.13)
i.e. JifW = lf<*)(a,)9,7)
A = l,...6.
(11.14)
We are going to prove the conjecture (11.11). Let us notice first that the vectors x, x and L = x x x are linearly independent,*' unless x is parallel to x. They span a 3-dimensional space. Hence Lj
(11.15)
form a basis for the space of (3 x 3) tensors. If we take into account that ^LiLi=L2,
^XiL,=0,
i
y^ZjZ/j = 0,
i
i 2
^ii(xxL)i=0,
^]ii(xxl)i=L ,
i
i
^L,(Lxx)i=0,
^ii(Lxx),=0,
i
i
^i;(xxL)i=0,
(11.16)
i
^ii(Lxx)i=L2 i
we conclude that for the matrix Gji expanded in terms of (11.15) the co efficients in front of these expansions are invariant with respect to proper rotations, under the proviso that (11.4) is satisfied. For example the ex pansion - L ^ ^ G ^ x x L L f L x x ) , *
i
(11.17)
j
renders the coefficient in front of Xiij which is, in deed, invariant with respect to proper rotations.^ In other words, it is a function of x 2 , xx and *' Linear independence means that the relations Ax+jix + ^L = 0 implies A = fj. = v = 0, for x }f x. f)
1
—L - t f <4>[x(x x L)[x(L x x)] = KW , (L2)2
1126 26
The Inverse Variational Problem in Classical
Mechanics.
x 2 Moreover, it is uniquely determined by Gy. The same applies to other coefficients of the expansion. Notice, however, that in (11.11) we used a basis different from that in (11.15). But one can check that*' LiLj L/i±jj
2
=L + {xx)(xi±j — J J Sij 0{j -p (XX)(2^Xj + -f" XjXi) XjXiJ
2 - xX2±i±j X{Xj - iX XiXj X{Xj ..
(11.18)
So we may replace the entry LtLj in (11.15) by <%. This new basis was used in (11.11). This accomplishes the proof of our conjecture. The necessary and sufficient condition for the local existence of L, satisfying the conditions (11.6), is the symmetry of r)C
rG - dG* (11.19) (11.19) l]k ijk S f ^ - d±k with respect to permutations of the indices i, j , k. Hence we are looking for such K(x\ A = 1, 2 , . . . 6, for which this symmetry occurs. Let us introduce the auxiliary quantities L±(x,x) = l [ L ( x , x ) ± L ( - x , - x ) ] , (11.20) G*(x,x) = l[G | [ Gi3v(x,x) ( x , x ) ± Gy(-x, Gy(-X, --x)] x)]. Then, taking into account (11.6), 3 d& L± L± dik dii dij
_ dG% i ~ dik
G±± l k
(11.21)
is symmetrical with respect to interchange of the indices. According to (11.11) and (11.20) G± = = K^Sij K^dij G+-
+ K^XiXj
K^Xiij + K^x iXj
2
^ ^ ( tfx ( <x >(x(x — x L ) x[ xL)[x(L ( L x xx )x)] ] = 0
•) O
LiLj
}/ , . // , d/? , ,. /} i^iab^jcdXaX eiab^jcdXaXccXXbb±Xdd a b e d
=
z2zZz-,z2 a
=
e
d
zj2jzj2j a
=
b
b
2 2
e
5 i
S
*c "re
**J
Sac
6bj
Sbc
J'('
J a c 5 M
= S,d Old
xbbxxdd = *«* xaxc± 6bd
- ibcSad)XaX ad)xaXccXXbbXXbb
d
2
* i i i ) , (11.22) + ZjXi),
etc..
— =
S V °ij
K^{Xl±j + K^Hnij
2
ify 6ij(x (3C x 2 * 2 - ((xx) x x ) 2 ) + . .. .. = SijL SuL2
+ ...
... =
Evaluation
G% = K^ixiLj
of the function
Gij. Gij.
+ XjLi) XjU) + K^iiiLj
127 127
xjU). + XjL,).
(11.23)
Let us first concentrate upon Gf,k. We have the following Statement Statement 1 1 We We assert assert that that 22 „ ,+ , _ <9 AT(x22,xx,x ,xx,x2 ) a AT(x IJ dii dij
(11.24)
or, in accordance with (11.6) and (11.20), + x L+ = X ( x 2 , xxx, x , xx22) + Y X(x Y, ^F?(x)xi ( ) ^ + ^^++((*x)) ,,
(11.25)
i
where F + and J+ are arbitrary functions. Proof of the Statement 1: According to (11.14) dK^ dii
dK^ l dpXi+2
-~di-=-w
nBKW
(11.26) (1L26)
^rXidj-
Using (11.26) and (11.24) we get G
U =^
= ^TS^Xk + KWS^
+2 ——5ijik
{z) +K 5ikXj K^Sikij
+
+ K^Sjkxi+
K{3)Sjk±i+
PEW dKW_. . . H—^— l i i j i t + 2 — - — 2 ^ 2 ; * + "T^ T'&
no
XiXjXfc XiXjXfc -f-f-
Q _ XjXj'Xfc XiXjXf- "r -\-
oj
dp
r\st Z a/i X{XjX %i%j%k "V & k ~r
r\
(11.27) U 1 "*') _T
XiXjXk~T XiXjX^-j-
dp
X{XjXk -\- L
XiXjX XiXjXkk ■.
Let us make use of the symmetry of Gfj Gj,kk with respect to the interchange of the indices i, j , k and compare the coefficients in front of the tensors OijX OijXk,k ,
OijXk , OijXk,
XiXjXfc XiXjXk, ,
X%XjX XiXjXk k
e tc.. etc..
(11.28) (11.28)
These tensors are linearly independent (see the proof of conjecture (11.11)). As the consequence the following relations should be true dp '
(11.29)
128
The Inverse Variational Problem in Classical
Mechanics.
tf(3)=2^l,
(11.30)
2 ^ — = ^— 07 op
(11.31
Formulae (11.29)
(11.31) entail
-w=2jhr
(n 32)
-
It follows from (11.31) that there exists a function if (°) for which (locally) 2ffCfl = ^
,
JfW = * ? .
(11.33) (11.34)
07
From (11.29) and (11.34) we infer that there is a function if for which (locally) K{0) = 2 ^ ,
(11.35)
K W = 2~
(11.36)
07
Moreover, by virtue of relations (11.30), (11.33), (11.34), (11.35) and (11.36), we get eventually
^
=0 -
if(3)=4—,
^
( 4 )
=
2
|^-
(11-37) (11.38)
(1^)
Relations (11.21), (11.22) and (11.36)-(11.39) imply (11.24) and (11.25). This accomplishes the proof. Let us now investigate L~ and G~. We have the
129 129
Evaluation of the function Gij. Gij.
Statement 2
If G G~- defined denned by (11.23) satisfies the condition
dGj dG±L dGjjL =_ dG~ kj dxk d±i dii
(11.40)
G~ then G {{ has the form
^^kr'^M-^M
(11.41)
N.B.: Of course, if G~ has the form (11.41), it must satisfy (11.40). Proof of the Statement 2: To this aim we introduce two auxiliary quantities, namely A0
=
B0 Bo
2 i5) 2 ( 6 ){6) i f ( 5 ) xK + JxC+K x xxx, ,
(11.42)
2 2
= K^xx
(11.43)
+ K^x K^x
Then, due to (11.23),
J2 G-jXj = A0Li, j1
L Y, G&i = =B°BL°> YjGiJ^ i-■
(11.44) ( 1L44 )
j3
We are going to show that the functions B = BB0L0L2 2
A = A0L2 ,,
(11.45)
do not depend on x. The function AQ and Bo depend only a, (1 and 7. Taking this into account we have ^^
T i
'
dA 8Ao0 flB
^^
T
i
'
E4t = °i
i
fdA fdAo0
^odA dA00. .
\\ (11.46) (1146)
130
The Inverse Variational Problem in Classical
Mechanics.
Using (11.40), (11.44), (11.45) and (11.46) we get
- ^(EE<^.) = fdG-j i *->^
j
^
dLi\
Xk
OXk
\°
\dii
J
J
lJ
*-f \ oxi
axi
(11.47)
:
dxk) axk J
In a similar way we get £?-=0. dxk provided we exploit the relation
(11.48)
arj
«
(n 49
- >
J
Prom (11.47) and (11.48) as well as from (11.42), (11.43) and (11.45) we conclude that A = A(x.2), B = B(x 2 ). Taking once more into account (11.42), (11.43) and (11.45) we may express K^ and K^ in terms of A and B, viz. K^
=
-^[A(ah-B(a)p],
K^
=
-^[B(a)a-A(a)p}.
(11.50)
Thus the necessary form of K^ and K^ is given by (11.50). We are going now to show that Gy defined by (11.23) and (11.50) satisfies condition
Evaluation of the function Gij. Gij.
131 131
(11.40). We have G
ij H
= =
771^2 TT/T^ {Ma)[x {Mrfi*22(xi (xiLLjj + XjLi) - xx(xiLj + XjLi))+ L v > +B(a)[-xk(xiLj + XjLi) + x22{xiLj + ijLi)}} . +B(a)[-xk(xiLj + XjLi) + x {xiLj + ijLi)}} .
Notice that Notice that dd AA (f W\
dxi dxi U UV V ""
Hk tjikiiX tjikiiX11 Hk
(11.51) (11.51)
2Lj J^k J^k Hi Hi Lktkiiii Lktkiixi -- 2Lj L L2 2
" "
( () )
~'
Therefore
[A (h\ + JL(h\ A fh\ _i2i \JL(h\ 2 [dii V W o»i \ L2V . " [dii V W dxj U A = (T2\2 /L2^2 X 2_/) ^{LjLktku z2iLjLkekii =
+ LiLktkii)xi = =
(11.52)
= , _ 2 , 2 [Lj(xi± [L,(x*x2 2 - iixx) + Li(xjX. LiixjX.2 - XJXX.)] XjXk)] . A similar computation yields
i \JL(*±\ + -!L(h.X 2 2
2
2 1[d±i x\L )
[dii I W'
d±j \L2 )\ dij3 \\L /-i )_
"
(11.53)
2
2
= , 2 . 2 [Lj(x 4 x - ijxx) Xjxx) + Li{xjX. - XjX±)] Xjxx)] Notice that the replacement of x by x and x by x on the r.h.s. of (11.53) produces the r.h.s. of (11.52) with a minus sign. If we insert the l.h.s of (11.52) and of (11.53) into (11.51) we get
r-ij - -A\JL(i±\2 + -!L(k.\ _ " 2 [dxi \L J dxj \L2J u « " 2 [ai^W fej UvJ _B _d_ (LA 2 2 [d±i [dii \L )
_d_(LA dij \L2)
'
Since V^v^
we conclude that
JL(hb\ dij \L2J
d
(Lk\
v->. d
(Xiij - Xj±i\
. f(11.54)
132
The Inverse Variational Problem in Classical
Mechanics.
is symmetric under the interchange of j with k viz.,
A (M = A A (M (M 2 d±j U d±j \L2vJ
(1L55) (11.55)
d±kk \L*J U / d±
As the interchange of x with x in
m
(11.56)
causes only a change of sign of (11.56) we have also
A(M 2
(u(11.57) 57)
A(M
£(&)-£ @) dxj U /
dxk
-
W)
The formula (11.55) and (11.57) enables us to express G~- in (11.54) in a more compact form, namely
G" - G7l = -A— (Aj - B— (A j
(11.58)
This coincides with formula (11.41) of our assertion. Moreover, from relation (11.55) follows
A [A A [A (Ml (Ml == A A [A [A (Ml (M 2 dx [dxj v \ Lw / lJ a**k [a^-
dii L 22 /J / a±i [dxj ^ \U
«« m (11.59)
Due to the symmetry displayed in (11.55) and (11.57) as well as taking into account (11.59) we get by using (11.58) the relation (11.40). This accomplishes the proof. Our main task is still to find L~. We have the Statement 3
We assert that
IT = -\A{a)-
+ B(o)]P(x|x t b) + ^J ]pF-r(( x ))±i ii + - ( x ) , (11.60) + JJ~(x), i
where „r.r, , i. •, i, % 11 |x|(b l x l(b xx x)x x)x P , ' \ n; P (x x | x,b) X b ) = — a r Carctan t a n ' ' R \x\ ( x(xx bx )b)(x ( x xxx x) ) and F F~ and JJ~ are arbitrary functions.
(11.61) (1L61)
Evaluation of the function Gij. Gij.
133
Proof of the Statement 3: To show (11.60) and (11.61) we have to look for a function P for which dP _ Lj dij ~ L 2
(11.62) 1L62
W, = v
(
If P satisfies (11.62) then
_ ^ _
I
Txx
d_ (LA
- ^ [i^-H - £ (EO
(IL63 (11.63)
'
To see that (11.63) is really true one should carry out differentiations, indicated on both sides of (11.63), making use of (11.62) and L 2 = x 2 x 2 - (xx) 2 . The following expression should be then an identity XjLi + XiLj ^Lj + —j—g I ^Ci/yifcL ^2eikjXkL2 2 - 2LiX. 2LiX22±j Xj + 2Li(xx)xj x 2 (^ \ 2 2 = = -- 2LjXiX = - = -- jj ^eyfcifcL ^CijkXkli1 2LjXiX2 + + 2Lj{xx)±i 2Lj(xx)xi J1 ..
J = (11.64) (U 64) -
To check whether (11.64) ir really an identity it is enough to multiply both sides first by (5y and sum over i and j , then multiply both sides of (11.64), say, by Xi or ±i and sum over i ox j . The local existence of P follows from equations (11.55) by virtue of the Converse of the Poincare Lemma. It is obvious, due to (11.62), that we should have (11.65) j
J
3
J
and
Y.L*wrl-
(11.66) (1L66)
3
To proceed further we have to enter into some geometrical considerations. By virtue of (11.65) we have P(x,px P(x,x) P(x, px + crx) ax) = P(x, x)
(11.67)
>
134
The Inverse Variational Problem in Classical
Mechanics.
for L ^ 0 and for arbitrary numbers p and a, a 7^ 0. We infer from (11.67), that P is constant on the plane, spanned on the straight lines x and x. Thus P = P&),
(11.68)
where tp is the angle, measured between a fixed plane passing through the x-line and a plane, encompassing the x-line as well as the x-line. If we denote the rotation around the x-axis through the angle tp by Sv we have d ,„ .. x xx L ^(5,x),=0 = - ^ j - = R
/,,-„» (11.69)
Then due to (11.66), /dP\
x~,dPL]
1
_ (1I 70)
U L = ?^M=H
.
PM = ^
(11.71)
-
or
We are going now to express (p in terms of three independent vectors x, x and b. The x-line and the b-line mark the plane from which we are going to measure the angles. Vectors (x x b) and (x x x) are perpendicular to x; the angle between these two vectors is just ip. We have |x|(b x x)x S m ^|xxb||xxx|' (1L72) C
(x x b)(x x x) °S^=|xxb||xxx|-
(1L73)
|x|(b x x)x ^ (xxb)(xxx)'
(1L74)
t a
=
Prom (11.71) and (11-74) we conclude that n/ ^ n/ 1 • 1 \ 1 lxl(t> x x)x P(
Evaluation of the function Gij. Gij.
135
We shall need the formulae (a x b)c = (b x c)a = (c x a)b
(11.77)
(a x b) x c = (ac)b - (bc)a. (bc)a.
(11.78)
and
As a matter of fact, we used already (11.77) in getting (11.76). To compute the denominator on the r.h.s. of (11.76) we have to take into account that on one hand [(x x b) x L] 2 = (bL) 2 x 2 ,
(11.79)
on the other hand [(x x b) x L] 2 = (x x b) 2 L 2 - [(x x b)L] 22.
(11.80)
Hence, comparing (11.79) with (11.80), we conclude that the denominator [(x x b)L] 2 + x 2 (bL) 2 = (x x b ) 22L 22.
(11.81)
To compute the numerator notice first that according to (11.77) (x x x)(x x b) = [(x x b) x x]x and we have [(x x b) x x]x(x x b)i - [(x x b)x][(x x b) x x]j = [(x x b) x x]x(x x b)i - [(x x b)x][(x x b) x x]j = = [{[(x x b) x x] x (x x b)} x x] { = = [{[(x x b) x x] x (x x b)} x x]; = = (x x b) 22 (x x x)j = (x x = (x x b) (x x x)j = (x x
(11.82) (11.82)
b)22Li. b) Li.
Here we used twice the formula (11.78).*' Thus (11.82) yields the numerator of (11.76). Taking into account (11.81) and (11.82) we get
dP(x\x,b) _ U dii ~ L2
(11.83)
*' To get in (11.82) the second expression from the first one we have to identify in (11.78) x xxbb) ) xxx x, , aa ~~ ( (x
b ~ xx xx bb, ,
cc~~xx;;
to in the the second second expression expression to get get the the third third expression expression we we have have to to identify identity in a ~~ xx xxbb, ,
b~x,
cc~~ xx x bb .
136
The Inverse Variational Problem in Classical
Mechanics.
This accomplishes the proof. Combining our previous results of Statements 1 and 3, i.e. formulae (11.25) and (11.60), we finally get for the Lagrange functions satisfying condition (11.4) L = K(x2, xx,x 2 ) -
fe.4(x2)
LX
+ B(x 2 )l P(x|x, b)+ '
(11.84)
+ Y,Fi(x)xi + J(x), i
where F = F+ + F " ,
.7 = .7+ + J~ .
(11.85)
In the next Section we are going to choose from the variety of functions (11.84) those functions which, in addition, are compatible with condition (11.5). Before doing that, however,we shall write the Lagrange function L in a more shapely form. We write A as 2 ^ ( a ) - ^ =- M , a a
W
r(d,s^W. da
(11.86)
Then L = K + hW - ^pj xxP - BP + ]P Fi±i + J.
(11.87)
i
Let us prove an auxiliary identity, namely
E *,£o^-(nr-5)-r +
. r J § ^
(11.88,
The l.h.s of (11.88) reads
2W'xxP + wY,ii^i
(11-89)
Symmetry
of Gij and evaluation of the Lagrange function.
137
So we have to compute ^ p . dP _ x x ^-r' ldxi x2
p
[(b x x)x]{[(x x b)(x x x)]xx - x 2 [(x x b)(x x x)]} x 2 (x x b) 2 (x x x) 2 xx x2 _xx x2 xx ^ x2
[(b x x)x][b(x x x)]j[(xx)3:, - x 2 ^ ] x2(xxb)2(xxx)2
niQuI '
[
[(b x x)x][b x (x x x)]j[(x x x) x x], _ x 2 (x x b) 2 (x x x) 2 [(b x x)x](bx) x 2 (x x b) 2
Here we made use of formulae (11.77) and (11.78) as well as of the relation (a x b)(c x d) = (ac)(bd) - (ad)(be).
(11.91)
Relations (11.89) and (11.90) prove the identity (11.88). If we insert (11.88) into (11.87) we get
L = K + Y/ii-faT.(wp)
~ BP + J2iiNiW
+ J
'
( 1L92 )
where Ni^=Fi-W^)^^
(11.93)
Formula (11.92) does not yet present the final shape of the Lagrange func tion we are looking for. We postpone it to the next Subsection.
11.3
Symmetry of Gij and evaluation of the Lagrange func tion.
The condition (11.5) imposes some restrictions upon S, N and J in formula (11.92). We are going to prove the following. Theorem 1 We assert that the Lagrange function L, satisfying condi tions (11.4) and (H-5)> has the form (up to gauge terms). L = K(x2,xx,x2) + ^ i i ^ - [ P y ( x 2 ) P ( x | x , b ) ] . i
(11.94)
138
27ie The Inverse Variational Problem in Classical
Mechanics.
Proof of the Theorem: By virtue of (11.92) we have*) 2 y , d y, d2L L ,,
dL dL dxi
*—•d±i d±idxj dxj 33 *-? i ' 2 - dK if . _ dK V ^ dd2K J X *-? dii dxj = Saij dx--\dx-dx^ j
(11.95) (1L95)
ft
where where Qi = — Qi ^ - ( (J J + 5BPP) ) ++ i xx (V« x NN))..
(11.96)
A look at (11.95) and (11.96) makes it clear that only Q can spoil the forminvariance of G under rotations. Let us compute first Vx {BP). One *' In (11.92) the term
E x 1 £-(^).z is contained. The contribution coming from Z in (11.95) is dZ dxi
d2Z
^
>
^
d±i dxj
d a
v-v--
x, = - > J
> x,
*-" 4^ *-? *-£
J
dxj
AJ i tdxILM = _YEVI i ] dx * VLaw 9X
y v
■ •
d
xfc
d
,„ „„™ r™ (WP)
dxk d±iK
dx
J *v
/
by virtue of (11.62); notice that W does not depend on x. Since
dx
by virtue V of (11.62); « notice that W does not depend on x. Since the r.h.s. of (A) reduces to
-EE(*£<«')&
the r.h.s. of (A) reduces to
-EE (*£*£")&
'
=
{A) {A)
Symmetry
of Gi} and evaluation of the Lagrange function. function.
139 139
can show in a similar way as in (11.90) that*) d
x 9 p__ _pp3H i _ ((xx)(x x x ) ( x xxxx), )j dx{ dxi x2 x2(xxx)2
|
((xb)(x xb)(xx x bb)j ), x 2 (x x x) 2
(11.97)
Then -^-(BP) -£-(BP) axi
(-2x2B'-B)P^xz x b ) ( x x t ) ji ££ , ( x x ) ( x x x ) ii jBg ((xb)(xxfe) 2 2 22 22 x (x x x) x 2 (x x x) 2
=
(11.98)
Notice that x Q should not change under rotational transformations as required by (11.5). The change caused by rotation through an angle tp around x is - ]^^[[22xx22BB' '((xx22))--B B (( xx 22 ) ] \x\ \x\ Hence we infer that
(11.99)
(11.100)
2aB'(a)-B{a)=0 2aB'(a) - B{a) = 0 *' To get (11.97) we proceed as follows. We know the formula for ^
%
dx%
given by (11.90). We conclude from (11.75) that
and
P(x|x,b) = = - P- P( (xx| b| b, ,xx))
(A) (A)
P(Ax|x,b) = A A-"11PP((xx||xx,,bb))..
(B) (B)
From (A) and (11.90) we infer that
[{h
x)
„, i. , , „, ,- , * [(b x x)xl(xx) 5 >, Ap x|x,b)=-p xix,b)t+ :(V^ E ( ( -P(x|x,b) = -P(x|x,b)L J - ^ i d
*-r"
x b
x2 +
b ! dxi
2
x (x x
b ) r2
^ (C)
and from and from (B) (B) we we get get >Xi
4-
= - P
dxi
(D)
Assuming that x, x and b are not complanary vectors we make the ansatz (1) ) 3 3) — P ( x | x , b ) = uv^Xi ii + w i/ 2( 2>(x (x x X x)t x)i + W ?/ >(x ( x xx b)j b ) ; .. dxi dxi
(E) (E)
2) (2) To compute o' tA2) , uv^ and v3~> we we use use (D),(D), (11.90) (11.90) andand (C).(C). Eventually Eventually we we get get (11.97). (11.97).
140
The Inverse Variational Problem in Classical
Mechanics.
or B{x22) = C\x\, B(x C|x|,
C --a caoconstant. nstant.
(11.101)
We see that Q is forminvariant with respect to rotations if and only if axi
(11.102)
|x|(x x (x x by
and V t, x N = A(x22 )x.
(11.103)
If we take the rotation on both sides of (11.102) we get ( (xb)(x x b ) \ „ x c ( v , x ^ ^ ) 1= C ^ = J 0 . C \V |x|(x x b) = j C |x| = 0
( ^1XKW)
WF
-
Here we exploited (7.21) and (7.49). (11.104) implies Here we exploited (7.21) and (7.49). (11.104) implies C=0 C =0 and consequently (see (11.101)) and consequently (see (11.101)) B(x 2 ) = 0. B(x 2 ) = 0. Prom (11.102) follows then Prom (11.102) follows then 2 — == / / ( x2)X )^ dx~ ^ >
,
(11.104) (1L104)
(11.105) (11.105) (11.106)
or J(x) = J ( x 22 ) .
(11.107) 2
Thus we may include J(x ) into the term K in (11.92) . Taking all this into account we get from (11.92) L = K + YJ±i-fa(WP) i
+
YjiiN^-
(11.108) (11.108)
i
Now let us consider (11.103). The divergence of (11.103), taken on both sides, yields 0 = 2a\'(a) 2aA'(a)+2A(a) + 2\(a)
(11.109)
A
(11.110) (".HO)
or =j^3
Symmetry
properties of the Lagrange function. function.
141
Then the equation (11.103) reads V N= = c C ' ^ 33 v , xxN |x|
*
(11.111) (1L111)
'^F
From the first equality in (11.104) we guess the partial solution of (11.111), viz. , N N (( Pp)) = =c c
(, x( xbb))( (xxxx bb )) |x|(x x b) 2
(11.112)
The general solution consists of the r.h.s. of (11.112) and a gradient of any function of x, viz. c f rxb)(xxb) * ; ) ( x * g 2+ V * ( x ) . (11.113) |x|(x |x|(x xx b) b) 2 Let us recall (11.90) and compute with help of it and (11.112) N N ==
^ i ^ ( | x | p ) = x x | x r l p + | x | ^ i i l£ = i i
i i
b x b _ [( [(b x x)x](bx) _ ^Y ^ .■ (xb)(x ( x b ) ( x xx b)j )t _ 22 X1 l l | x | ( x x bb)) ^ | i | ( x x xx)) 22 |x|(x t-r \x\(x
(11.114) (11.114)
i
The relation (11.114) inserted in (11.108) yields i
* + + EY^/i^ [ (il(W ^ ++C''|x|)P] £^xi. L =K C'\x\)P} + Y,^^ii+ i
(11.115)(H-115
i
We can include C"|x| into W. Thus we get (11.94), where the gauge term has been omitted. 11.4
Symmetry S y m m e t r y properties of the Lagrange function.
In this Subsection we are going to prove the following Statement 4 The Lagrange function L is weakly invariant under transfor mations of rotation; this means that for any proper rotation S G S(Os) L(Sx,Sx)-L(x,x)=d/(x^;S) for some function / .
(11.116)
142
The Inverse Variational Problem in Classical Mechanics. P r o o f of t h e S t a t e m e n t 4: Let us consider Let us consider L(x, x) = L(Sx, 5 x ) ,
(11.117)
L(x,x) = Z ( 5 x , S x ) ,
(11.117)
where S € 5 0 ( 3 ) . With the notation
where 5 € 50(3). With the notation x' = 5 x ,
x' = 5 x ,
x' = 5 x ,
(11.118)
x' = 5 x ,
(11.118)
we have, according to (11.4) and (11.5),
we have, according to (11.4) and (11.5), g 2 L(x',x') _ S ^ x ^
- 2 ^ 2 ^ 2 ^ 2 ^ SkiSljSkmSln it
!
m
'
ramra
n
_ d 2 L(x,x)
-
(11.119)
"
and in in aa similar similar way and d2l
y.
,
^^— 9 i j <9±, <9i,
di_ _ ^ dL__Y* 3J
JJ
3 3
dxi eta;
da 2L£
.
9L dL
^—' ^ - ' dii S i ; dij d±, 3'
dxi dxi
JJ
33
(11.120) '
Thus we get from (11.119)
L-L
= Y/Dk(x;S)x Y/kD+^( k(x;S)x x-S),k+^(x-S),
(11.121)
kk
where D and $ are some functions. If we insert (11.121) into (11.120) we obtain
yfdDjjx-S) ^— V 9XJ
G>£,(x;5)\ dxi J
d*(x;S) dxi
j
_Q
(11.122)
As x appears linearly in (11.22) we have
or
dDi _ dDj_ dDj dlh 1 -7Tdxj = -Kdxi, OXj OXi
Di
_d$(x,5) ^dx~' OXi
=
$$ == const. const.
(11.123) (11.123)
v
$(5) == Cconst. HS) nSt °
-
'
((11.124) 1L124
)
Symmetry
properties of the Lagrange function.
143
Consequently we have, taking into account (11.121) L-L=^[¥(x;S)+t$(S)]
(11.125)
This accomplishes the proof It can be shown generally, applying the procedure above, for any quan tity which appears in the Lagrange function but does not appear in the Euler-Lagrange Equations that the Lagrange function is then weakly in variant under the change of this parameter. Let us, nevertheless, give another proof for the case of the vector b which method displays explicitly the form of the gauge term. To this end notice that P(x|x, v) + P(x|v, w) = P(x|x, w ) .
(11.126)
This follows as a result of adding together two angles. Let us denote by L the Lagrange function L = J Fs:(x 2 ,xx,x 2 ) + 5 3 i i ^ [ w ( x 2 ) P ( x | x , b ) ] .
(11.127)
i
From (11.94) and (11.127) we infer that L-L
= £ > A { W r ( x 2 ) [ p ( x | x , b ) ) - P(x|x,b))]} = i
= 5>^{W(x2)P(x|b,b)}=
(11.128)
= |[W(x2)P(x|b,b)]. We did not so far investigate other symmetries of the Equations of Motion given by (11.10), barring the rotational covariance. Now we shall examine the case when the Equations of Motion are invariant under time reflection x-»x,
x->--x,
x-»x.
(11.129)
The best way to discuss this stringent symmetry restriction is to inspect the behavior of the Lagrange function given by (11.94) under the time reflection. We get L = tf(x2, - x x , x 2 ) - 5 > ^ [ W ( x 2 ) P ( x | x , b ) ] . i
(11.130)
144
The Inverse Variational Problem in Classical
Mechanics.
If the second term in (11.130) does not vanish and we want to keep the absolute value of the Lagrange function unchanged we have to demand if(x2,-xx,x2) = -A-(x2,xx,x2);
(11.131)
the function K has to be an odd function with respect to xx. In this case the Equations of Motion remain unchanged. The consequence of (11.131) is that a K which only depends on x 2 and x 2 (e.g. K = | x 2 — V(|x|)) has to vanish. In case the second term in (11.94) equals zero (W = 0) the requirement of symmetry with respect to the time inversion reduces to K(x2,-xx,x2)
= ±K(x2,xx,x2).
(11.132)
Also in this case the Equations of Moton remain unchanged. We wind up our considerations by giving two special examples, which seem to be of some interest. If W = C\x\, where C is a constant then, according to (11.94) and (11.114)
r = *<*-.*) + 0Wj£$3.
(u.133)
Formula (11.133) coincides with (7.50) when K = | x 2 - U(\x\).
(11.134)
Thus (7.50) is a special case of (11.133). For W = x 2 we have the Lagrange function which depends on two constant vectors b and c / b, viz. L = K{x2,xx,x2)
12
+ (xx)P(x|x,b) + (xb)P(b|x,c),.
(11.135)
The largest set of Lagrange functions of one-particle sys tem in a (3+1) dimensional space-time, s-equivalent to a given Lagrange function yielding rotationally forminvariant Equations of Motion (formulation of the prob lem).
In this Section we are concerned with the following very important problem: to find the largest set of Lagrange functions of one particle in a (3 + 1)
145
The largest set of Lagrange functions. . .
dimensional space-time, s-equivalent to a given Lagrange function of the form presented by (11.94). A similar problem for one particle in a (1 + 1) dimensional space-time was solved in Section 10. The actual issue for (3 + 1) dimensions is much more complicated. The solution of it would be of paramount importance in Physics, as it could be used as a starting point for quantization procedure in, say, Feynman's approach to Quantum Mechanics. Unfortunately, we are not able to offer the reader a solution of this problem. It is not yet solved; it is, really, a difficult task; the reader will realize, why it is so, when he continues reading this Section. So we present here only a convenient formulation of the posed question. In Subsection 11.1 the most general autonomous Lagrange function, yielding rotationally symmetric Euler-Lagrange Equations was given (see formulae (11.8) and (11.9)), namely 2 L L= = K(x\xx,x K(x2,xx,x2))
+ J2^[W(x2)P(x\x,b)}, + Y/^[W(x2)P(x\x,b)},
(12.1)
i
where , , „, 1 (|x|[(b x x l x l l P x x, b = — arctan ' LV ' J .
|x|
(12.2)
[ (x x b)L J
The Equations of Motion (see (11.10) are ^ 82K .. 2— dxi dij j ditidxj j
y ^ d2K . *-? diidxj dii dxj j
J
dK dxi
3
-<**.-*>££(§)*,+
(12.3)
+ ( 2 x W "{2xW'+4{xx) + 4 ( x x ) 2 W2W")^, ")^, where „ NN W'(a)=
dW(a) dW(a)
L L= = xx xx xx ..
/
da
;
,
2
a = x2 , a = x2,
a
-
8= xk, P = xx,
-2 7 = x2 2 ,
7
= x ,
(12.4) (12.4)
146
The Inverse Variational Problem in Classical
Mechanics.
The equations (12.3) can be also written as an identity, in extenso, ^
d2K
„ . d2K 8K „. d2K n. 2 2x 2x + 28ii + 2:ClXi ^ + ***'1w8j + ^ + ^d^
/
?22 ([ft^ 3^ r + ^Wd~i 3
aT
+4x UiXjix^(x f^) J^)(x Jf^+x Jf^+ +L3Jf^) + JfV+i JfV .
+2
2
d d2K^
22 tI.I2 l289 KK + Xi|x|
W
++
+ A..
+4
.
2 d2K d K
™57^ s^ ™ ^ -2a,f-(2x2^-^)^.
2 2 , „. „. .„ .„ a a/s: ^ + 2a; x -
'
w
(12.5)
1 ] ( X^ e «« u * L 2 _ Li[2xjX2 - 2(xx)ii] 2(xx)i i ] 1 • .(si/W+i;./W+z./W)+ + (2x 2 ^' + 4 ( x x ) 2 W " ) ^ = 0 ,
where z< = /i(x,x) = Z i / W + i j W + LJ< S ), / « = /«(a>(9,7)
(12.6)
are the Newton Equations and we assumed that x | f i i.e. x, x and L are linearly independent and form a coordinate system in the 3-dimensional Euclidean space. The orthogonal set to x, x and L is
X x
— =
2X — p X £jx L I 7 r x' ,
,. j_i xX±
XX X xx x 2 .. - — X + — X , =_ -—x+—x,
L L
~" L *2
2
, (12.7) (12.7)
The largest set of Lagrange functions...
147
Let use rewrite (12.5) in a different manner, namely Y d2K
no
d2K
dK\
,
m
^w+2^)fi2)+{2awW'-w)if{3)+ n„
"/
d2K
d2K
d2K
2
AOd
dK]
K\
.,.,
LV
(12.8)
- ( 2 ^ -
W
) A
/
( 3 )
+ 4
^
+ 2 7
| | . ]
+
+Lj(2aH/'-W)^/(1»+2|^/(3) + + (27W" + 4/? 2 W")^ = 0 , where L 2 = a 7 - /3 2 .
(12.9)
If L 2 tends to 0 and all expressions remain regular during this limiting procedure, then (2aW - W)jf{3)
=0,
(12.10)
=0,
(12.11)
L 2 =0
{2aW - W)/3f{3) L 2 =0
and (2aW - W)f{1)
2_
+21W' + 4p2W" = 0.
(12.12)
We do not make any specific assumptions which would restrict the structure of the equations (e.g. of a sort of 2aW — W = 0 which implies W = cot? , c ^ O a constant). Then from (12.10) and (12.11) follows /<3)
=0 L 2 =0
(12.13)
148
T/ie The Inverse Variational Problem in Classical
Mechanics.
and from (12.12) ,(D| (1)|
13
IL2=O
+ *pw 2^W" 7W' + =_2iw 2aW'-W
((12.14)
'
X
'
We conclude from (12.13) and (12.14) that >( Q ,/3, 7 ), /< 3 >=L 2 $( 3 >(a,/?,7), j ^ + 2 / ^" ^T 2aW'-W S ^ ;
(1)
(12.15)
^
+ ^~WV*>P,1), ,,,,
(12.16) (me)
(1)
+
where $'3^ and i'1' have weaker singularities (poles) than 1 for L 2 = 0. Let us examine the coefficient of L, in (12.8). Using (12.15) and (12.16) we get (2aW - W)^ W)* ( 1 ) + + 2^L2$<3> = 0 or £(i) SW
= =
3 2 (1 — $< — $ ( 3>) = = LL2 $$( 1 > ) d-y 2aW -W
_22— _ —
(12171 (12.17)
Consequently 7
V
2aW"-W
(12.18)
;
and the force / is for 2aW / W tj fj
Xj
~
2 yW W" ■yW - 2[3 2/3'W" 2aW'-W 2aW> - W
(12.19)
-^W5Sp^-^)
rt,S, +
The Equations of Motion (12.8) can now be written L 9K 2 2(L L ((aa*Z *£ 2 ++2WB ^^ __ ++ 22M\^ dK ( L 2 )22)2
V a/3
d 7 ) d72aW'-W
^dpdj
^
W
|+
+ (2aW> + (2aW - W)j] Wft $< $<33>> ++ (/}!** ( 0 ^ 4+ + 2277 d2R ^ —)\ f(2)_ fW-
~22
' + 2/m ? 2" ^ / dc?22tftf , f^7 W"
- { 22aW>-W2 J [ W +
d922KK
8K\ 8K\
oa J [aaW ++ 22P/ W ++ 22 *7 J ++
OQ 2
a2/v: ^
+
a2x
^
-
2
air ax
^ = 0'
Wd~l
*7 J
('12 20') (12.20) ' ^
V
The largest set of Lagrange functions.. junctions... .
"_ / d2K _ (2 a 92R V dpdj
+
149
2 2 2 dK (L igd K\ dK ( L 2 )) 2 AB — \ P d~i 3 7 22 ) fry d 7 2aW - W
$<3> +
-(2aW'-W)p -(2aW'-W)p
+(^+<£^0)/ m + 2p2W"\
c,(lW'
(n
' H 2aW-W d2K da da 93 7
d2K 2a
(12.21)
AOd
+0
2
K\
){ Wih W)"
d2K dp dp dj dj
7
The The equations equations (12.20) (12.20) and and (12.21) (12.21) have have the the structure structure (3) A$ A$ ( 3 )
{2) + + Bf Bf{2)
(3) {2) D D $$ ( 3 ) + + Ef £/(2) = = F,F,
= = C, C,
(12.22) (12.22)
where where A- A
2R
(J 92K 22 (n V dp a/3
flP Ii 228
^d ' K dpdj
II 22 d ^K \ 1 d^ K d1 ) )
L 22 )) 22 22 (( L
d12aW'-W
II
+ (2aW -W)^, +(2aW'-W)j, „ „
2 nd d2K K n
„„
da 22Kx
„__nloW' ^ + 2p 2 / 2W W" (/ d2K K 7W' + 2aW"-WT r 3 / ? 2 °~ 2aW'-W 2
{
+
d2 K K fliTX dK\ + + d$d1 dpd fry) dp d 1 1J
a P 2+P
2
dd'K K dK dK frK_ frK 11 2 + + 2 ~ 8(3 dada'' ~ dadp ~ ~ dp 22 22 2 2 K ,d K\ (L if,»d ff dd K K\ dK dK (L2))2 a + 2 \^ Qdpdj d1 2aW 2aW -- W <9/3<9 7 + 5dj7 2 J) <9 w 7 -(2aW -W)p, -{2aW -W)p, P Pdadp
d2K K-2 — 28 4 — *3d%K dj dpdj dj22 dj <9/?<9 <9 7 7 2 fiW' <922if /7W + + 2P 2/32W"\ Ty"\ /
F
= 2 ( 2aW>-W ) {2adm _ 4p / ? - ^ - 2 7rr ^ dad~< dad~< dpdj dpdj
+
2 K\ jntad&K\
^W)
, (12.23) (12 23)
~
150
The Inverse Variational Problem in Classical
Mechanics.
For AE - DB ± 0 (3)
a.e.
CE-FB AE-BD'
(12.24) AF-DC AE-BD
(2) J
{
'
If we change the Lagrange function changing K and W the structure of the Lagrange function should remain the same; this is our requirement. The Newton Equations should also remain unchanged. Taking this into account we should have for K ->■ K' and W -> W w 31
C'E'-F'B1 A'E' - B'D' '
,
m
J
A'F'-D'C A'E1 - B'D' '
,„„„,., v
;
Where A',B', C", D', E' and F' correspond to expressions (12.23) where K and W were replaced by K' and W resp.. From (12.25) and (12.26) follows C'E' - F'B' = X(CE -
FB),
A'F' - D'C = \{AF - DC), A'E' - B'D' = \(AE -
(12.27)
BD),
where X = X(a,/3,j).
(12.28)
The relations (12.27) form the starting point for finding the answer to the question stated at the beginning of this Section. As an illustration of the method presented above let us inspect a model, investigated by us earlier, in Section 8, but from a slightly different point of view. Our starting point will be the Lagrange function (8.1) L=\1-U{a)=K{a,1)
(12.29)
and the Equations of Motion dU .. BU Xi + — = Xi + 2—Xi = 0.
(12.30)
Hence, according to (12.6), /(1) = - 2 ^ ,
/( 2 ) = /(3> = 0.
(12.31)
The largest set of Lagrange functions...
151
Our goal is to find the largest set of Lagrange functions V of the struc ture given by (12.1), s-equivalent to the Lagrange function (12.29). Notice that all those Lagrange functions have to have the same Newton's Equa tions (12.31) (normal form). To start with let us write (12.8) in the form of a system of three equations, taking into account (12.31), viz. „ / d2K
d2K
nn
, „„ d2K /
2a
d2K
-{ W^ -(2aW
, d2K 2
ind
+ 4p
K\dU
w)^
-W)^-
+ (fW
8K\
8U (12-32)
dK 2
nnd
+2
K
^
+1
d2K
=0
mh >
n
(12 33)
-
+ 2{32W") = 0,
(12.34) oa as |L| / 0. Here U as well as W are functions of a only. If we differentiate (12.34) with respect to 7 we get W = c = const.
(12.35)
Then (12.34) reduces to
C =
(12 36)
S °
-
which entails W =0
(12.37)
as
— ^ ±0
a.e.
oa by assumption. We conclude that the set of Lagrange functions, s-equivalent to (12.29) does not encompass Lagrange functions containing the term
appearing in (12.1). We infer also that the equations (12.32) and (12.33) remain unchanged and L = K(a,P,rr)-
(12-38)
152
The Inverse Variational Problem in Classical
Mechanics.
Let us differentiate (12.32) with respect to 7 and (12.33) with respect to j3 and subtract the latter from the former. We obtain d2K dp2
t
4
d2K = 0. da 97
(12.39)
If we introduce the new variables "f 7= =
a = z2-zi, - zi ,
z2 Z+2+Zi z\ ,
(12.40)
then the equation (12.39) becomes d2K 8d22KK n d22K K 2+ + ~dp dz\ dz2 =0 ~ w ^[-^[
,^nA% (12.41) (12 41)
-
a 2-dimensional wave equation (equation of the membrane) where P and z\ play the role of the space coordinates and z2 - of the time coordinate. Hence we conclude that every solution of equations (12.32) and (12.33) has to satisfy the wave eqaution (12.41) (or (12.39)). Notice, however, that not every solution of the wave equation satisfies the equations (12.32) and (12.33). One can easily check that K== - ( aa77- -/ 3P22))>i ,, K
(12.42)
a
discussed in the example 8.5 I., satisfies (12.41). Let us focus on the equation (12.33). If we denote (12.43) (12-43)
■g^ = Q, 07 equation (12.33) can be written in the form
(■>-*-m-<%+»%-°
(12.44)
or, using the theory of characteristics, d/3 dp
^ ^M 7 _- 22Q a 1 ' oa
_
dj
=
da =
"" ~Tp^M ~ ~7p™M Ww ""
(12.45) (12 45)
'
da oa
The partial solutions, obtained from (12.45), read
^ + U(a) == ci, %y+U(at) Cl,
aj - 2=c P2 2.= c2 . aj~P
(12.46)
The largest set of Lagrange functions... functions...
153 153
Hence the general solution of (12.44) is 2 ^=QC — = llQ(§7 ++ t / ( aU(a), ) , a 7 -ai/ -p 3 2 )),,
(12.47)
where Q is an arbitrary function. From (12.47) follows II Q{^u + U(a),au-/32)du
K=
+ F{a,P). F{a,P).
(12.48)
On the other hand let us insert (12.47) into (12.39). We get
=4 =4 +4
(1,(12.49) 2 49)
BQ A.dUdQ dUdQ ,. dQ d2K 8Q dQ = 4 = 4 1- 47 d(32 da da dc\ dc2
w £ ^ ^>
-
where notation of (12.46) was used. The solution of (12.49) reads fp Jb
(dUdQ fdUdQ
dQ\ dQ\
\dadcL
dc2J
+F1(a,j)l3
+ F2 (a, 7 ) .
(12.50) (12.50)
0=v
Let us compare (12.48) and (12.50). We have
f Q\1=udu + F(a,p) = fjdUdQ [\(dUdQ
8Q\ dQ\
(a
(12.51)
NJ
+F + F22 (a, +Fi(a,7)j8 (a,7). 1(a,j)(3 + Let us further differentiate both sides of (12.51) twice with respect to 0 and then once with respect to 7. We get 2dUd^
da dcf oc{
d2Q | 1 \dUa ' 0C2 dc2 [[da oa
1 d2Q \\ dci oci dc ac22
QdQ =
dc oc22
0.(12.52)
To get a solution of Q which depends only on c\ and c2 but not on a we conclude that d2Q dc\
=
d2Q _ Q dci dc2
or, after solving the equation (12.52), _ = a + —Z-T e +gI , Q Cl + r r - T T + Q= aci 2(c2)2 I9 1
(12.53) (12.53)
154
The Inverse Variational Problem in Classical
Mechanics.
where a, g and e are some constants. From (12.53) follows £ {
K = a(l+u),+
^-/
) k +
f7
+
F(a,fl.
(12.54)
If we insert (12.54) into (12.32) we realize that a =0
and
F(a,/3) = gU(a).
(12.55)
This leads us to the familiar result (see (8.187)) 6
K = g(^ + U(a)) +
-^.
(12.56)
As another example let us inspect the case when L = ^ii-^-[WP)
(12.57)
i.e. K = 0. Then the Equations of Motion reduce to >y(2aW' - W)f{3) = 0 , P(2aW - W)f{3) = 0 , (2aW - W)fw
(12.58)
+ (2jW + 4(32W") = 0 .
The only nontrivial case is when Xi =
2W'-f + W'P2 2aW'-W
Xi + f
m ±l
'
(12 59
- ^
2
where /( > is an arbitrary function of a, (3 and 7. The Equations of Motion read
} v
j 2
/
(12.60)
2
+(2k W - 4(xx) W")S- = 0. In other cases W is a constant or just zero and the equations (12.58) or (12.60) do not make sense.
Construction
of the most general two-particle Lagrange function..
.
155
Can we expect that equation (12.58) admits other Lagrange functions belonging to W ^ W ? To answer this question consider W(a) = W(a)\{a).
(12.61)
Then, according to (12.59), we should have W _ W'X + WX' 2aW< - W ~ 2a(W'X + WX')-WX' f\n\
/-in
W" 2aW -W~
W"X + 2W'\' + WX" 2a{W'X + WX1) - WX
'
If we rewrite these equations as differential equations for A we obtain X'W2 = 0 , (12.63) X"(W2 - 2aWW') + X'(2aWW"
- 4aW'2 + 2WW') = 0
or A = const.
(12.64)
The set of Lagrange equations, s-equaivalent to (12.57) is L' -XL,
A = const.
(12.65)
i.e. all are equivalent. 13.1
Construction of the most general two-particle Lagrange function in (1+1) space-time dimensions giving rise to Euler-Lagrange Equations covariant under Galilei transformation.*)
In the former Sections we investigated several aspects of the problem of Lagrange functions and Euler-Lagrange Equations for one particle in (1 +1) and (3 + l)-dimensional space. In the present Section we are going to consider the non-relativistic case of two point particles moving according to the rules of the Lagrange formalism in (1 + l)-dimensional space-time. The case of (3 -I- l)-dimensional space-time will be discussed in Section 14. We are not going to specify the form of these equations; the only restrictions imposed upon them are the following •) J. Lopuszariski & P.C. Stichel, Fortschr. Phys. 45 (1997) 1.
156
The Inverse Variational Problem in Classical
Mechanics.
i) these equations should be forminvariant with respect to the Galilei transformation, defined below, and ii) should be derivable from a Lagrange function as its Euler-Lagrange Equations. The unknown Lagrange function, we are starting with, (13.1)
L = L(x,x,R,R,t), L(x,x,R,R,t), depends on x H R M
xX
= = = =
x^-x™, x^-xM, (1)
TOia; m
1 == 1,2,
(2)
^+«*&2 , M ' mi +m mi^O, + m22^0,^ 0 , mtpQ, dx{t) Rk =_ dR(t) dt ~dT'
= -5-'
=
+ m z
x^ = x®(t), =x{i\t), x®
(13.2) (13.2) (13.3) (13.3)
m22 ^^ 00 ,,
^r
(13.4)
(13 5) (13.5)
-
The positions of the particles at the time t are given by x^^t) and x^(t). The real quantities mi and 7712 are arbitrary constants satisfying (13.4). The masses of the particles can, in general, not be evaluated unless the 3-rd Newton's Law is observed. Thus rrii i= 1,2 should not be necessarily identified with masses of the particles. The Galilei transformation for the variables x^\ x^ and t read (x^,x{2\t)
-> g(a,v,s)(xW,xW,t) ^(o.u.sJC^ 1 ),^ 2 ),*) =
(2) = (z (1) + a + vt, x{2) +a -)- a + vt,t vt, t + a). s).
(13.6)
In (13.6) a is the spatial translation, v - the velocity (special Galilei transformation, boost) and s - the time translation. The forminvariance of the Equations of Motion under the Galilei tranformation (13.6) means that the Euler-Lagrange variations of L should not depend on R, R and t. Thus we have --x d dL dL ,. == fix / (> : CX> a ;X> a ;R)j> R) dtdx--dxd t 5 ± - ^ ' ' ' '
(13.7) ^
d dL dL 9L „ , . .. »~. ;— —7- = r KF { (x,x,x,R). m R ) Jtd-R-dR dtdR OR = ' ' ' ' -
(13.8)
The question we are going to answer is: what is the most general form of L, when (13.7) and (13.8) are observed.
^
Construction
of the most general two-particle Lagrange function... function...
157
If follows from (13.1) and (13.7) that f(x, x, x, R) = (x, x)x + f2(x,x)R, (x, x)R, f(x,x,x,R) = ffo(x,x) fi(x,x)x 0{x, x) + /i
(13.9)
where /h! = £dx2> '
(13.10) (13-10)
h = - ^ dxdR dxdR
(13.11)
Since dh 8h ^ l= = dJl == QQ dR dx dR dx we conclude that we conclude that
(13.12)
hf2 ==Mx). h{x). In a similar way In a similar way F(x,x,x,R) F(x,x,x,R) where where
(13.13) (13.13) = F {x,x) + Fi(x,x)x = F00(x,i) + Fi(x,x)x
+ F {x,x)R, + Fa{x2ti)R,
(13.14) (13.14)
fF1i == ^^ = = ^/ 2 ( x- ) , dxdR 9a; dR 2 _ 9rP-T L F2 = ^-=-
(13 15) (13.15) -
^ = ^ = 0 , dx dR
(13.17)
(13.16) (13.16)
Since
we conclude that F2 = F2F(x). 2(x).
(13.18)
Now we are in position to integrate (13.10), (13.11) and (13.16), taking into account (13.13) and (13.18). Prom (13.16) and (13.18) follows L = g{x)R g(x)R22 + gi{x,x,R,t)R ffi(x, x, R, t)R + g2(x,x, g2(x,x,R,t), R, t),
(13.19)
2g = FF2.2.
(13.20)
where
158
The Inverse Variational Problem in Classical
Mechanics.
From (13.15) and (13.19) we get dgi(x,x,R,t) = dx
h{x)
or 9\ -9n{x)±
+ gio(x,R,t),
(13.21)
where gu(x) = f2(x).
(13.22)
Finally from (13.10) and (13.19), taking into account (13.21), follows
or 92 = g22(x,x) + g21(x,R,t)x
+ g20(x,R,t),
(13.23)
where g22{x,x)=
fi(x,u)(x-u)du
(13.24)
and the constant c is so chosen that the integral on the r.h.s. of (13.24) makes sense. If we insert (13.21) and (13.23) into (13.19) we obtain L = 922(x,x) + g2o{x,R,t) +
g2i(x,R,t)x+ (13.25) + g(x)R2 .
+g10(x,R, t)R + gn(x)xR
Let us once more use (13.7) and (13.9). We have d2L d2L • X+ R+ did-X dIdR didl-^
d2L
dL
r
,
=M
x
, ( 13 " 26 )
s
^ -
If we insert (13.25) in here we get, after some cancellations, d2g22{x,x) dxdx
. X
dg2i{x,R,t) OR
dg22{x,x)
dg20{x,R,t)
5
o
ox d
■
- -^R
2
=
ox fo(x,x).
dg2i{x,R,t) dt dg10(x,R,t) 3
ox
■
nio^
~
\\.a.£t)
H
Construction
of the most general two-particle Lagrange function...
159
The coefficients of R and R2 in the identity (13.27) have to vanish, which yields dg21{x,R,t) dR
dg10{x,R,t) dx
_n - °'
,.,„., (1328)
g(x) = g = const..
(13.29)
Taking into account (13.28) and (13.29) the identity (13.27) reduces to 82g22{x,x)
.
dg22{x,x) /o(l x) = Yx ' X ° ° (13.30) _ dg20(x,R,t) _ dg2i(x,R,t) dx dt The l.h.s. of (13.30) does not depend on R and t, the r.h.s., on the other hand, does not depent on x. Consequently x
didx aX X
d g22 . dg22 --^r-^-x+—— +f0 = g0{x) ox ox ox
, „ „ , (13.31)
and dg2o . dg2i + = 9o{x)
-^x-
^r
,,,,„! (13 32)
-
-
Due to (13.8) and (13.14) the identity X^x + ^RR + -5^- - | | = F0{x,i) (13.33) 8Rdx 8R dRdt dR is still left for inspection. With help of (13.25), taking into account (13.28) and (13.29) we get ■ 2dgu(x) ox
, dg10{x,R,t) at
dg20{x,R,t) dR
_
This can be also written as cw •-, -2dgii{x) F0(x,x) -x'—z = °X dg10(x,R,t) dg20(x,R,t) _ = dt dR— =
(13.34) h{x)
-
Let us differentiate (13.32) with respect to R and exploit (13.28). We get d (dgw(x,R,t)
dx-{—m
dg20(x,R,t)\ =0
m—J
-
,.„,, (13 35)
-
160
The Inverse Variational Problem in Classical
Mechanics.
This means that the differentiated function (13.35) depends at most on R and t. On the other hand we learn from (13.34) that it depends at most on x. Hence dgw — at ~W
dg dg20 20 —— = K — COnSt const.
K= an = - -3R
. . (13.36a)
'
To recapitulate: we have OR —M
dg2i(x,R,t)
_ dgio(x,R,t) dg10(x,R,t) dx dx
dg2i{x,R,t) dj^R^
_ dgdg 20{x,R,t) 20(xji,t)
—m
OR—
_
=
°'
= gQ{x)
=
9o{x)
/ t a a«u\
(13.36b) (13 36b)
-
(13 36c)
(13.36c)
and (13.36a). It is convenient to use the following notation gio(x, R, t) = gio(x, gio{x, R, t) - at, nt,
(13.37a)
g21 (x, (*, R, t) = g21 (x, R, t) - g0g{x)t. 0(x)t.
(13.37b)
Then (13.36a)-(13.36c) read dg20 _ dgw _ „
^dt = ~ 0,'
(13.38a)
dg i _ dg _ ff-^=0, dt dx
(13.38b) 13 8b
^OR2
20
dffio _ dg21 _
-dx--^R~°dx dR ~
( , , 13 38c (13.38c)
( -
)
The l.h.s.'s of (13.38a)-(13.38c) look as a 3-dimensional rotation of a vector g9 = (521,310,520) (921,910,920)
with respect to the variables Y =
(x,R,t).
Hence, by the Converse of the Poincare Lemma, 92i 921 ^
910 9io
= ~
d${x,R,t) d$(x,R,t) ^ dx
, '
(13.39a)
-=
d$(x,R,t) d${x,R,t) Q£ dR ',
(13.39b)
Construction
920 920
of the most general two-particle Lagrange function... function...
d$(x,R,t) = — gt
(13.39c) (13.39c)
'
at—
~
161
Let us once more recall the Lagrange function (13.25) and insert in there (13.37a)-(13.37b) and (13.39a)-(13.39c). If we discard the gauge term 4* we obtain L = g22{x,x) 922(x,x) + go(x)tx g0(x)tx + gn(x)xR
+ KtR gR22 ntR + gR
(13.40)
We may introduce _ dG(x) --90{X) * ( * ) == % * •■ -Hx-
(13.41) (13-4D
g0(x)tx + KtR = --~-[t{G{x) — [t(G{x) - KR)] + G(x) G{x) - KR
(13.42)
Then
and the modified Lagrange function (13.40) reads 2 + + gR gR2 , ,
L= L = U(x,x)-KR+^^-R U(x, X)-KR+ —f^R at at
(13.43)
where where (13.44) (13.44)
U(x, x) = g22(x, x) + G(x) U(x, x) = g22(x, x) + G(x) and and
dG(x) _ (13.45) ^ s f t =l ( flu(a:)« ) . —j— ax ax Thus (13.43) is the final form of the Lagrange function we were looking for. The Galilei forminvariant Equations of Motion are d22U . dxdx
d22U.. U dx2
^ _ ^dGf,_dU_ _ ^ L _ Q
dx
dx
(1346) (13.46)
d2G .o dG., n ?, d (dG . n ■ \ n 0—xi2 2 + + 2gR 2gR + = j—t (^x = 0 . (13.47) + ^—xx + + KK= l —x + + 2gR 2gR + + « t )Ktj=0. Let us write (13.46) and (13.47) in the form of linear algebraic equations with respect to x nad R, viz. d2U.. dx2
dG -~__ _ dU dx dx
d2U . dxdx
(13.48)
162
The Tfte Inverse /nuerse Variational Problem in Classical
dG.. ,t
Mechanics.
92G.2
-J
(1349 (13.49)
E * -0*'-'
»
We denote
*£-(£)'=»*<>-
(13.50)
a.e..
Taking into account (13.2)-(13.4) as well as (13.50) we get + + " P ff 5MM VV 35 ^^ dxdx ) M dx m2dGd2G.2 2 _dH]_ m2dG_d^G_. d2 U ( d2G . 22 \
+ +
21 M M 3^x ^a i ^
9G 9G 9a: dx
2 + 2 + " ^9i( ' V"s + a 9a;? a *! JJ +
2 U_., _ //_9a 2 t/ _ 9X/V dU\ \ dx dx dx JJ \dxdx dx
, „ „ , (13.51) (13 51)
"
m ■■(2) __ 11 \r, i fdU dd22 U rn^dG [Q "H fgg ^ -\ \ mi. 9(7 ff 9 7 7 + ~ " P -~v[ [ Mw{~dx~~dx~d± 1^ 9a ~ 9a T 9± a 7x)++W~dx~ 2 t f & KK+
1x
dGd2G ., m^dG_d^G_. mi 2
92£/ / G .,2\\ dHJ_( d92G
+ M dx dx2%1 + 2 X 9± 2 K \K++ adx^ Mto^ + M )J-~ +
2 _dG U . _dU\ _dUY _ 9 G /I d92E7 dx \dxdx dx J
(13.52) (13 52)
'
If the 3-rd Newton's Law is observed i.e. there exists such fi\ > 0 and fi2 > 0 that 1) {2) ^/ii£ i i ((1) +n +/J.2x2X^
== 00 ,,
(13.53)
then from (13.51)-(??) follows (\x //iim xm22 9 \ M + +
fj, M2mi\ /dU /9E7 2mi\ M ))\dx \dx
d2U \ dxdxX)
2 (firms Mgron gg dGd G .2\ /fiim2 H2fni\ /(dG dGd'G K + K+ 2X VV M M ~ M M / \ f t s )\dx~dx~~dx ~dx~~dx2X)~ j J.\d52U2 c / /( 9d22GG . 2,\\ + K+ X
-^ ^W\ M )-
__dG 9 G /t d922U [/ , _ 9 C E /A\ 1 _ 9a; \dxdx dx J
~
(13.54) (13 54)
-
Galilei forminvariance
163
of the Euler-Lagrange Equations... Equations...
If we choose mi mi = =m pi, ,
(13.55)
m22 = = H2, m H2 ,
then (13.54) reduces to d2UK+^. f ±2)_dG(dHl._dU\ d2G . 2 \ 8Go ( d2U . ^£( dx22 \ dx22 ) dx \dxdx \ dx dx
dU\ dx) dx)
(13.56)
Since we do not intend to restrict the freedom of choice of U(x,x) conclude from (13.56) that G = const, const. or just
K 0, « == 0,
G (2 = 0.
we
(13.57)
These relations will have some significance in the discussion of the next Subsection. 13.2
Galilei forminvariance of the Euler-Lagrange Equations for two particles in (1 + 1) space-time dimensions.
Let us start with equation (13.47) which entails that
is is If If
22 ^ ^ - x + 2gR + KKt=^-(r t= ^ ( r ++ I«i =S zUt ) ) = dx dt z ox at a Constant of Motion, where a Constant of Motion, where T = G{x) + 2gR. V = G{x) + 2gR. (13.57) is satisfed then G = 0 and (13.57) is satisfed then G = 0 and 1 2 r = g ( MMli * ( 1 ) + M 2 z ( 2 ) ) .
r = g( o;( ) + ^ ) ) .
The canonical momenta dx^ dxW
y
computed from (13.43) yield m P
,(2)p
8U dU dx dU dx
dG ■ dGdx
dTmi dTmi dt M '
dGdx
dTm22 dt M
(13.58)
(13.59) (13.59) (13.60)
The Inverse Variational Problem in Classical
164
Mechanics.
and W+pW ppW+pW
^ ==P. (13.61) (13.61) at In case the 3-rd Newton's Law is observed relations (13.60) and (13.61) imply = =
M = 2g. 2g.
(13.62)
Notice that p is not conserved as long as the 3-rd Noewton's Law does not hold. From (13.58) follows namely dp dt dt
^ == -—K. —K.
(13.63) ' The Lagrange function (13.43) is weakly invariant under the Galilei transformation 3,(0' xW
= xx(i) (i)
=
+t
+vvt
a , + a,
+
ii ==l , 1,2. 2.
(13.64) (13.64)
we have SL —Ka —— ——(—Kta), SLaa = = —no, (—Kta), at
(13.65a) (13.65a)
".-sEan* 1 " "!*">■
(13 65b) (13.65b)
5LV - -Kvt + —-V —v + 2gRv = — [(-±Kt2 + I > ] ,
(13.66a)
SL -r.{pvt). 5LVv = —(pvt). dt
(13.66b)
i.
i
-
The „Noether currents"*"> or generators of translation and the „boost" resp. are ja = pP + + lit Kt = = S, S , ja =
(13.65c)
*' When the Lagrange function L is weakly invariant under a symmetry transformation then „ _ dip{xM,xW,t) (A) oLi — dt On the other hand 6L = 4E" \dxM (4r-M%) + did) #TT^(')Nj))
Galilei forminvariance
jv = =pt+\nt pt+\nt22
165
of the Euler-Lagrange Equations... Equations...
(13.66c)
-v. -V.
Notice that due to (13.58)
■Jr =0 dt
(13.65d)
Jdjxv,
, 13.66d (13.66d)
and dp dT dT —— r ==-ft—t ++Pp+ + KtKt —-—-==-nt — K,t + p + Kt-p +p+Ht-p=0. = 0. at dt at dt at dt
The Hamilton function or the generator of the time translation reads Z^ L^,
gx(i) gx(i)
i = = ^-x ^-x dx ox
+ + 2^-xR 2^-xR dx ox
2 + + 2gR 2gR2 -- U U+ + KR KR -- ^-xR ^-xR dx ox
T—XR + gR2 = -=— x + -TT-XR
ox dx
ox dx
-U
+
We have 2 dH _ d2U , 2 do^U U ... dU.. dt dxdx dx2 dx dGdG..dGdG~ ■ ~ n + -^xR — x R + -^r-xR + 2gRR ox dx dx ox _ / d2U . \ dx dx +
d22U U± . dx2
KR.
d2G .2 ■ dx2 dU . dU.. ■ ~-x - -^—x -^rx + KR nR = ox dx dx ox
dG~_dU\. dGz_9U\ dx dx J
(d (d22GG « 9G.. 9G.. „ ==
\ -•
4£ -- dxto dxto
//
{!^x
+
2 -- gR gR2 = = (13.67a) (13.67a)
Hx-X + 29R + K ) R
(13.67b)
„n
= 0
>
or
di(') dxto dt dt v V
I
(B)
^£f V V dxto dam
dt dt did)) did))
++
dt dt [[^^
Bxto Bxto
J)
Taking and (B) (B) the the Taking into into account account the the Euler-Lagrange Euler-Lagrange Equations Equations we we get get from from (A) (A) and „Noether current"
i=Eii l i ( i ) -* w . i , J ! ' { )'
which is a conserved quantity as dj _ d I sr-^ 8L
\ _
Jt~ dt \^> dxM ~dt~ dW> ~~^J ^J ~~
166
The Inverse Variational Problem in Classical
Mechanics.
where we exploited (13.46) and (13.47). The Lie-Cartan relations for Galilei group expressed in terms of the Poisson brackets, using the formula*' r„
m
d A
dA
[*.*! + £ - £
(13.68) (13-68)
are d
r•
y . , f ilnl = - ^3a = - «
(13.69)
(see (13.65d)), [jv,H] - ^ == ~P~ -p-Kt Kt ==~i°-ja b'«, -H"] == ~~Qf
(13.70)
(see (13.66c)-(13.66d)) [ja,ja]=0, ba,ia] = 0 ,
(13.71)
[jvja] = =\ptT,p] = = --\T,p] = -2 -2g9 = = --A [7»,i«] b* - r,p] rr)P] = A
(13.72)
as \p,p] [p, p] = 0 and .„ , r _ „ _ , dG n mi dG m2 [T,p] = [r,p] = [G [G ++ 22gR,p) ^ p ] == -— + +2 2g-jr _ ++ 22 5^- | = 22g, 5 - -- -=5l ox M Ox M
(13 731 (13.73)
[jv,jv] == 0. [jv,jv] 0. If the 3-rd Newton's Law is valid and K — G = 0 we have instead of (13.43) the Lagrange function L = = U{x,x) [/(j;,i) + + \MR | M f2,l 2 ,
(13.74)
*' We derive here the formula (13.68). If we have a quantity A yl(x,p,t) "' ( x , p , t ) we have — A(x,p,t) v
dt
= >
—-j-x (l > H
^vai(')
TTTP 1 1 ' I H
a P (>r
;
a<
l(/i) J
Since 5p(0 '
ax(;)
expression (A), using (B), takes the form d dt dt
dA dt dt
which coincides with (13.68).
^p( sr-f dA 9H ^2f V 5 i ( 0 dpd) XaxM dpi*)
dA dH \ dpM dlw) dpWdxWj '
(*)
Construction
of the most general two-particle Lagrange function...
167
where we used also (13.62). Moreover (13.69), (13.70) and (13.72) are replaced by [ja,H] = \p,H] = 0, [jv,H] = \pt = nix{1) -n2x{2\H]
(13.2.69') = -p,
[jv,p] = -M,
(13.2.70') (13.2.72')
resp.. The quantities K and 1g in the algebra (13.69)-(13.73) are considered to be central generators, which commute with all other. In case the 3-rd Newton's Law is valid there is only one central generator M. 14.1
Construction of the most general two-particle Lagrange function in (3+1) space-time dimensions giving rise to Euler-Lagrange Equations covariant under Galilei transformations.*^
The assumptions made by us here are similar to those made in (1 + 1) dimensional case mutatis mutandis. We are looking for the most general form of a Lagrange function L = L(x,x,R,R,£) which we assume to exist and to give rise to Euler-Lagrange Equations, forminvariant with respect to the Galilei transformation (x(M^5(S,a,v,S))(x(M = (14.1a) = (5x ( i ) + a + vt, t + s)
i = l,2,
where 5 is an element of the 50(3) group, a 3 x 3 real orthogonal matrix, a is the vector of spatial translation and s - the time shift. In this Section we restrict ourselves mainly to the transformations (x(i),t)->g(l,a,v,S)(XW,t);
(14.1b)
at the very end we shall make use of the required rotational covariance. *) J. Lopuszariski & P.C. Stichel, Fortschr. Phys. 45 (1997) 1.
168
The Inverse Variational Problem in Classical
Mechanics.
We have the identities V d2£ . yV ■ V *-* dij dxk ^—" *-* dij dxk ~* + y . _ ^ L _ „ ^— dij dxj dik
d2L ■ R dij dRk 5±j 5 f t
J
k
d2L _ dL_ - — &&j dxj 9 i j dt 5t 5XJ
y, 92 L ^ *—' dii dij dRk k
(14.2)
J
x x ) i , + ^ / j2)2 ) (x,x xx ) f t = (x, x) + Y fffk = zj ff (x, i)+52 $ ((x, -*)** + E 4 ( ' )^
and and
fc
k
k
k
^ d2L . y, d2L X. + V 0 f t fix* Xk+ * Y 5 f t dxk z ^ dL + > — ^dRjdxk ^dRjdik
y > d2L ■ d2L y > d2L ■ + d2L V 5ft 5ft * 5 f t St ^ dRj 8Rk 2 * + dRj dt .. ^ 5 L xk + > Zz ^ft = r'd± ?d±]dRk
dL dL + 5ft + dRj (14.3)
x xx ) f t . = (x, x) *) + i f > (x, *)Su x)x* + + E £ F< i f$ >( (x, = ii ff > > (x, +£ £ FltHx, - )^ • k
k
From 5 ft dRi
5ifc dxk
we infer that 2, 2) 44 2 ) ==/< / ] , 2 ) ((*). x).
(14.4)
Prom From
djJl 5ff
= =
dRi 5 ft
djf_ dFJl^=Q dxk 5ijt
follows
*f = *f{x). if^^fCx).
(14.5) (14.5)
Let us integrate the relations d2L dijdxk ^ ^ dxjdRk
^tx'x>' = / ( 2 , (x)
J j k [ h
(14.6) (14.7)
Construction
of the most general two-particle Lagrange function... function...
d2L (2) : r— = F}, (x) . Jk dRjdRk The result is
169
(14.8)
L = 9gW W (x, x) + g<0101> (x, R, i) t) ++ £J2 9f 9?2)2) (x, R, t)±, tfy + 3
)1 0 )
+ 5^>9fj > ( x(,xR,1Rt ), ^^ 3++£^ ]^T< ^, S1 )1 () (x x) )i i; ^^ i + +
7
(M9) (14.9)
*
where ^ - = t f ( x , x ) ,
(14.9a)
si^W = /]"'«,
^M-Jff'W,
(14.9b)
2 f^f f( (xx)) = F^ ^( (xx ) ;
(14.9c)
the remaining terms are not fixed; they are at most linear in x and R. Let us preform a gauge transformation on (14.9) namely L-¥L L -> L = L L- - — ~ *$((xx ,, R R ,, i ) , at at where "3/ is defined as a certain solution of the equation 0 11 R ,I tt )) .. ^ff(° ))(x,R,t) ( x , R , t ) = -- ** (( xx ,I R
(14.10)
(14.11a)
Hence L = 5 ( 0 0 )(x,x)
+
^g(.02)(x,R,i)ii
+
i 1 ) s
+EE?fw^> where where
i (02) _
.7
fc
(14U) (14.11)
k (02) _
5 *
(14.11b)
170
The T/ie Inverse Variational Problem in Classical -(10) £(10)
=
yj
Mechanics.
(10)_ 5 5* * (10)
yj
(14.11c)i
.
d R
*
Keep in mind that the Equations of Motion are invariant with respect to the gauge transformations. Hence from (14.2) we get the identity d2L
^
82L
^ ++
d2 L +
■
dl
=/
, ( 0 )).. . „ , „ . _ , (x x)(14 12)
E^7^^ EaM^^ E^7^^ ^a-^^+a^-^ 5^-^ > '
"
or taking into account (14.11)
V *y» ay* ±±nk ++ y y ddHl y JlXk ± k++ y L- d±j dxkk *-*
k
2-J dx QXk ^ k
dd9
j£l^RRlXk+ tiKl
2_, dx gXkk *-?
k
y9jrRk+ °jr_*r_ydjrikik__ Rk+djryjT_ydjr ++yvjr £— 2— dR 8Rk
dt
K
l-
dxj
Q-I10)
k ft
/I
kk
Ii
J
L.
O (11)
K K
At
*— dxi dxj
J
(14.13)
3J
J J
Relation (14.13) is an identity. As the coefficients in front of Rj, Rjik RjRk do not depend either on x or R we have (2)
(14.14a)
9ki fffci == Cu Cki == Qk Cik = = const. const. oxk
and
(14.14b)
oxj
02) agj (x,R,t) ag<10>(x,R,t) =_Q0. gg^M,gT^M
.14c) (14 (14.14c)
Taking (14.14a)-(14.14c) into account, we get from (14.13)
ay»)(x,x) v ay°°)(x,x)
JOj(X,Xj - > > Joj(x,xj =
— —
22
9 ag(°°)(x,x)
XXfcfc HH
--
^^9ff(° y /a f f (° )(x,R,t) )(x,R,t) _ affff (° >(x,R,t)\ ^ .
V ~^M
= =
2
&i
dxk
02) ag((02) gg (x,R,t) at
^
dxj
)JXk
+
h
( w1 5 )
(14.15)
Construction
of the most general two-particle Lagrange function...
171
The l.h.s. of (14.15) does not depend on R and t, the r.h.s. is at most linear in x. Hence dg[02)(x,K,t) dt
,_, =gj W ,
c^
"
(14.16a)
dx~3
- l ^ ^ m
J
(x).
(14.16b)
i
Notice that from (14.14b) follows (by using the Converse of the Poincare Lemma)
•Tw - ^fer1
' I416c »
We still have to investigate the identity v ^ d2L > ■ ^dRjdxk
v^ d2L i ^dRjdRk
.
Xk + >
■
Hk H
d2L : dRjdt
dl dRj
=
(14.17)
= F o i (x,x) derived from (14.3). Taking again into account (14.11) we obtain 3$< 1 0 ) (x,R,t).
n )
¥ ^ ** ? ¥ ~^^*+ +
+^ ^ ^ ,
dg< 10) (x,R,*) ^3iJ t
^5ff(02)(x,R^). Xk \ dRj =
*+
( x ) .
3gj. 10) (x,R,t) 3t
^d$0)(x,R,t)h \ dRj
(14.18) _ ~
Rk
F0j(x,±)
or, because of (14.14c) /agj10)(x,R,Q _ c^10>(x,R,t)\ ^ ^- ^ aiifc dRj J k
+ E E
^
i A + fCl|M^ j(l[ , i ,
(1419>
172
The Inverse Variational Problem in Classical
Mechanics.
The r.h.s. of (14.19) does not depend on R therefore 9g( 1 0 )(x,R,i) dRk
agj 1 0 ) (x,R,t) dRj
which implies by using the Converse of the Poincare Lemma Pi ' ( x , R , t ) =
^
(14.20)
We have still
The l.h.s does not depend on R and t, the r.h.s — on x. Consequently
W^-Xll-tor*** =
at
=
(i4.2i)
= 5 ( 2 ) (x). ' Notice that (14.16a) entails ^
0 2
>(x,R,*)=0.
(14.22)
By virtue of (14.14c) and (14.22) d 2 g (° 2 >(x,R,t) dRk dt
2 { l0) g k (x,R,t)
=d
= Q
dxj dt
hence
agj10)(x,R,t) dt does not depend on x. Comparing this with (14.21) we conclude that -g{10) (x, R, t) = g{k2) (x) =
Kfc
= const.
(14.23)
We introduce a new notation which turns out to be handier in solving our problem. We define gf0) (x, R, t) = gf0) (x, R, t) -
Kjt,
(14.24)
Construction
of the most general two-particle Lagrange function... function...
~gf (x,>R R, gf2) (x, R, - g\0) (x)t. gSM2)J(x 1t)t)== jJ">(x > Rt) I *)-^(x)t.
173 173
(14.25) (14.25)
Instead of (14.14c) we have
agj02)(x,R,t) ^j 10) (x,R,Q
o
(14.26)
instead of (14.16a) 02 (02) S< / ( ° 2 )>(x,R,*)=S ( x , R , 0 = 5 ( O 2 ) ((x,R) x,R)
(14.27)
and instead of (14.23) ~9W=gW(x,R).
(14.28)
* r = s r (*.*)■ We may introduce
14 29 ((14.29) - )
(10) 10) G (10) (x, R) = G G((10) (x, (x,R) ( xR, , Rt), i-) -E^ K "J ^i ?i^* +
By virtue of (14.20), (14.24) and (14.28) the r.h.s of definition (14.29) can be justified as follows d(g(10> _ d(?(10> ~dR~~ dRj dRj
K] Kjt
~
(14.30) (14.30)
1 0 ) (x,R). = 3 J 11 00 )) ( x , R , t ) - ^ t = g(. 10) (x ) R).
Taking into account (14.27) and (14.30) the equation (14.26) can be written 02) 10 ffj 4 ° 2 ) (x,R) ( x , R ) = ^-G< >(x,R) + + ajQ(x). i(x).
(14.31) (14-31]
Let us collect (14.14a), (14.16c), (14.24), (14.25), (14.30) and insert it in (14.11). We have
L=/>°>(x,x)+Y, ^ L = fl<°°>(x,x) + £ v^ , , x + J2 aja (X)XJ
3
(1
2 i+E»rw«i+
^ 130
v^9G( 1 0 >(x,R)£ • R
+ J2 J ( )*i++J2 E j3
d&1 x,R)±
° ) j X ' R ) ^ - + £ e , f (x)tx,+
J3
i
^
•
K R t+ ^Q^. H* ~ i++EE *i^*+ i l J3
3
+E E ea \\. .( x( )x^)4^- + E E c«M c*M ■ +Y,E
174
The Inverse Variational Problem in Classical
Mechanics.
If we discard the gauge term rfG<10>(x,R) dt
we obtain L = 00>(x,x) + £ 5 j.°>(x)**i + X > ( x ) ^ + I > ^ * __«#'>(*)
+E E
J ag<
.
'
+
( 14 - 32 )
. .
c R
*«** + E E ^ i ■
* i i 3 The Lagrange function (14.32) encompasses two terms which depend ex plicitly on t. To get rid of them let us examine the equation (14.16b), viz. gggfrjM ggj 0a) (x,R,0_ v m
^
^
-Li'Ml i
W-
According to (14.25) and (14.31) we have §j M ) = J - G ( 1 0 ' ( x , R ) + f l i ( x )
+ 3 (°»(x)i.
(14.33)
If we insert (14.33) into (14.16b) and differentiate both sides with respect to t we get
fri0)w..0
<w 5a;
fc
9a; j
or ffj0)W
= ^J»l0)(x).
(14.34)
Thus exploiting (14.34) we get for one of the terms of L in (14.32)
5 > f ( * ) t i i = | [ 3 ( 0 ) ( x ) t ] - 5 (°>(x).
(14.35)
j
Moreover
E ^J* = 1 E «>*** - E **^ •
(14-36)
Construction
of the most general two-particle Lagrange function.. function... .
175 175
Thus, using (14.35) and (14.36) and discarding the gauge terms,we arrive at the autonomous Lagrange function 8G BC
L KjRj+Y. Y -g-r^i+Y. L= = K(x,x)~Y^ A-(x,x)-]T KjRj+Y. Y -g^^+Y jj
ii
i 3
'
ii
Y CCijikR) ,(14-37) Y HRiRi >(14-37) ii
where a K(x,x) aj(x)x K(x, x) = g^(x,x) g(°°> (x, x)++yY, i ( xj ) ^ -- g^(x) 9{0) (*)
(14.38) (14-38)
.7 j
and Gf1}f > =SEGG G 3. 3.
(14.39)
Formula (14.37) is not yet the final form of the Lagrange function, we are looking for, as we did not yet take into account the consequences of the rotational part of the Galilei transformations. The Euler-Lagrange Equations derived from L of (14.37) read ^ d2K . £f dxs dxt '
d2K .. *-f dxs dxt f'
y , dCU- _ 8K_ ^ *-f *-f dx dxss ll dx dxss ''
T-,
+E -^-x +y+ 2C y y* -5—-^—Xiit+y 2c RA = -«*. E E s&** a r * S *-* *-?• oxi dx *-* ox ^r - —"■ t
tf
ir
t
tt.
t
1
st t
(14.40)
<14 41) (14.41)
-
tt
Notice that the equations (14.41) can be also written ^ Gj\G + 2j2C 2YJCjkjR kR ^L( (x)j{x) + = --««,,.. k) k\ =
(14.42) (14.42)
The equations (14.40) and (14.41) should be rotationally forminvariant under transformation x -> x x' — = Sx, xx --> > xx'' = = SS xx ,,
(14.43) (14.43)
x x— —► T x'x'==SSx x. .
This imposes some restriction on the construction of L. We are not going to enter into the lengthy calculations concerning the structure of K(x,x),
176
The Inverse Variational Problem in Classical
Mechanics.
we are going however, to straighten up other problem. It follows imme diately from the forminvariance of the equations (14.41) under rotation transformation that Kj = 0
(14.44)
and dj = —Sij ,
A - a constant.
(14.45)
Although G itself does not need to be a vector, | 2 i and af, ^ should be tensors with respect to th indices i, j , k. We are going, however, to show that*' Q = X2,
Gi(x) = G(a)zJ,
(14.46)
up to an inessential constant. The proof runs as follows. We start with the ansatz (see (11.11) and the considerations following it) U
^^5jkG^{a)+xjxkG^{a)+Y,t^iG{z){a).
Xk
(14.47)
i
By differentiating (14.47) with respect to Xi and making proper choice of indices i, j , k we get for i = 3, j = 1, Ar = 2 0 2 Gi ——— dx2ax3
„ d2G^ dG^ = 2xlx2x3—^r 2 + G{3> + 2x2s —— , da da
for i — 2, j = 1, k = 3
or QQ(3)
Gi3) + (x2+x2)-^-
= 0.
(14.48)
From (14.48) follows for x2 = x3 = 0 G(3)(Q)=0.
(14.49)
Thus we have -^-
= SjkG^(a)+xjXkG^(a).
*' the idea of the proof was given by Dr. J. Cislo.
(14.50)
Galilei forminvariance
177
of the Euler-Lagrange Equations... Equations...
Let us examine 8Q (9Gj\ Ad_ ffdG±\ ;A _. _d_ _d_ (dGj\
dxi \ dxk )
dxk \ dxi )
After performing the differentiation and comparing results with each other we get QQ(I)
2S 2SjkjkXi-^Xi-^--
dG^ + + 8ikXiSjkGxwlG{2).. ijXk-^+ SijXkGp) 8ijXkG(2) = = 25 iSijXk-Q— +
(14.51)
For j — k ^ i we get 2 3GW =
G(2)
(14.52)
oa da If we insert (14.52) into (14.50) we get dGj3 dG dx & *k
. MU „ .m . .
„ J
= -—(z,G oxk
(1)
dG^(a) 9a &*
(1453) (14.53)
( a ) + const).
If we put G'1) = G the proof of relation (14.46) is accomplished. Thus the final shape of the autonomous Lagrange function is L = K(x,x) + (xR)G(x 2 ) + v 8r( + 2 ( x x ) ( x R— ) ^U M +2(xx)(xR)
. ++ ^- ( R R ) ;
(14.54)
N.B. We did not discussed the structure of K(x, x ) . 14.2
Galilei forminvariance of the Euler-Lagrange Equations for two particles in (3 + 1) space-time dimensions.
The Euler-Lagrange Equations corresponding to the Lagrange function (14.54) read ^ d22K . ^ 832lK K .. l *-* ^ dx <9is s dx Sitt * L* ^ - dx dxss dx dxtt „< vft „ dG(a)^ +G(a)R. + Zxs—±-l-Y,xtRi
-
dtf = }£- >
(14.55) (1455)
178
The Inverse Variational Problem in Classical
^(xJG(a)
Mechanics.
(14.56)
+ XRXR J)=0. J)=0.
The latter equation implies that — (xjG{a) + +XRj) jt(xjG(a) XRj) = =pj, Pi ,
j3== 1,2,3 1,2,3
(14.57) (14.57)
are Constants of Motion. For an arbitrary function G(a) this conservation law differs from Newton's Third Law. In such a situation we are not able to evaluate the proper mass values of the particles under consideration. The constants mi and m^ (mi + va2 = M ^ 0) were chosen arbitrarily and do not need to correspond to real masses. Only in case when G = = /i n = const.
(14.58)
we may establish the true masses of the particles, viz. A
Hi mmii +M, +(i, Hi = -TT
(14.59a)
A ^2 V2 = = j^m j^rn22
(14.59b)
-- fi, H,
/Ui n% — A = total mass = M fit + fi%
(14.59c) (14.59c)
Then ■(1) Pi = m&fl +, itoif- (32 ) (14.60) Pj = p.\x) + V2Xj is the total momentum. The Lagrange function (14.54) is not Galilei invariant. It is, however, weakly invariant under the transformation (14.1). The ,,Noether currents" for translations and „boosts" can be computed as follows. We have on one hand
SL = J2 JttTT3J •■vivi ++°°( ( vv)), > SL^Yl
(14.61) (14-61)
j
Tj ^XJG Tj^XjG
+ XRJ + XRj
(14.62)
On the other hand
" - i E E j W - 1 (§)<*+%>•
(14.63)
Galilei forminvariance
of the Euler-Lagrange Equations. . ..
179 179
Hence the Noether currents" for translations are j(a)_lr._ ji a) = T =Pj i It >
((14.64) 14 64)
-
(see (14.57)), and for the boost — (14.65)
jf^Pit-Tj.
Both „currents" are Constants of Motion as a consequence of p in (14.57) being a Constant of Motion (see (14.56)). Both are also generators of the Galilei group. To see that we are going to use as auxiliary tool the relation . . _ . dA dA (14.66) where H is the time translation generator of the Galilei group and [ •, ■ ] denotes the classical Poisson brackets. The Lie-Cartan commutation relations for J* a ' = p, H and J^v> are as follows. In case of J^a\ p is conserved and does not depend explicitly on t; therefore \pk,H]=0. \Pk,H]=0.
(14.67)
In case of Ji"' this vector is also conserved but depends explicitly on t (see (14.65)). Taking into account that | £ = 0 (see (14.62)), we get v) d Vl{:\H] ,H} = = -p-kPk [J =- °^ ^ rr = ,,
(14.68)
{ V) == -\6 [{J4 i ,Pk} ,Pk] == -[ru -\Ti,Pk] -A<5jjfe, Pk} ik,
(14.69)
[j^^j^] [j^^^]
= -[t -[tpPii-r -rii,tp ,tpkk-T -Tkk]]
=
= -t{pi, = -t\pi, Trk] -- t[Ti, t\Ti,pPkk]] ++ [Ti, [Ti,rTkk] == t\6 t\sik -- t\S txsikik == 0, o, \pi,Pk]=0. \Pi,Pk] =0.
(14.70) (14.71)
To obtain the relation (14.69) one has to use the formula ^Pk]
=d^k+
X
W6ik
- d^k
+A
M
5ik = Xkk
(14.72)
180
The Inverse Variational Problem i n Classical Mechanics.
and to get (14.70) ane has to use
which is trivial as I7 depends on x(') and x(') only. Relation (14.71) follows also trivially. We do not present here the commutation relations of generators of the rotations S (see (14.1)) of the group SO(3) with themselves and other generators as they are standard. Relation (14.69) yields the extension of the group by the mass generator A. The Hamilton function is
This Hamilton function is a conserved quantity, due t o the Equations of Motion (14.54) and (14.55).
15.1
All two-particle Lagrange functions s-equivalent to a given autonomous Lagrange function yielding Galilei forminvariant Equations of Motion in (1+1) spacetime dimensions.
In Section 13 we presented the construction of the most general two-particle Lagrange functions in (1 1) dimensions, a Lagrange function giving rise to Euler-Lagrange Equations covariant under the Galilei transformations.*) The goal of this Section is to find all autonomous Lagrange functions, s-equivalent to the original Lagrange functi0n.t) Unfortunatelly, we are not able to acomplish our task and to construct the whole set of such Lagrange functions. Notice that under the s-equivalent Lagrange functions can be and are also such whose Euler-Lagrange Equations although equivalent t o the original ones are no longer Galilei covariant. In Subsection 15.2 we give a full account on the set of s-equivalent Lagrange
+
*)
J . Lopuszariski & P.C. Stichel, Fortschr. Phys. 45 (1997) 1.
t ) J . Cislo, J . Lopuszariski, P.C. Stichel, Fortschr. Phys. 46 (1998) 45
All Euler-Lagrange
Equations, forminvariant
under the Galilei transformations.
181
functions, all of which have Euler-Lagrange Equations forminvariant under the Galilei transformations. In Subsection 15.3 we give some examples which — we hope — will clarify our derivations. In Subsection 15.4 we give a natural extension of the set obtained in 15.2 to a partial subset of EulerLagrange Equations equivalent to the former which do not display Galilei covariance. Finally in Subsection 15.5 we touch upon another important problem where the equations in normal form (Newton's Equations) but not necessarily Euler-Lagrange Equations are forminvariant under the Galilei transformations. *) 15.2
All Euler-Lagrange Equations, forminvariant under the Galilei transformations.
We start with a Lagrange function given by (13.43), viz. L = U{x,x)-KR
+ G'{x)xR + gR2-
G'(x) = ^ ^ - , (15.1) ax where K and g are constants, U and G — arbitrary functions. Other quan tities are defined by (13.2)-(13.5). The Galilei forminvariant Equations of Motions are w t t .L:I s 0f +
var fl L:II
C 5 + ^ i - g = O1
= G'x + 2gR + G"x2 +K = 0.
(15.2) (15.3)
We assume that (13.50) holds true. Let us resolve the equations (15.2) and (15.3) with respect to x and R. We have
2
=M- 2fl (^*"^) +o,(tf, * ,+ ' B) ] s/ - (ir '* ) ' (is - 4) V =
dx2
\dxdx
dx)
(15.5)
IR{X,X).
Notice that the equations (15.4)-(15.5) remain true for each element of the set of s-equivalent Lagrange functions. *' An illuminating discussion with Dr. J. Cislo is acknowledged.
182
The Inverse Variational Problem in Classical
Mechanics.
We are going to show that the most general Lagrange functions, fulfilling the requirements presented above for g ^ 0 kept constant, are of the form
- S ( i ? i ) + i - ( 5 ' ( a ; ) ) 2 i 2 - ~G{x)+G'(x)xR+
(15 6)
'
+#.R2 - R.R, where cr is an arbitrary function and Hl(Xii)
= x9-^f±
- g (G'(z)) 2 - C7(M) - ^ G ( x ) ■ (15.7)
As far as the quantities k and G(x) are concerned the explanation is as follows. If the first Newton Equation*) (15.4) does not have the form f, = foix) + f2(x)x2
(15.8)
we have G
=
G,
(15.9)
h
=
K.
(15.10)
This is the most frequently encountered case. If, however, (15.4) has the from (15.8), i.e. we have a restriction imposed upon U in (15.1), there exists an additional set of solutions, namely a) for h=h{x)x2
(15.11)
we may have G' = G' + A e x p j - I f2{x)dx\
,
(15.12)
where A is an arbitrary constant; then it follows that & = «, *' First equation in normal form.
(15.13)
All Euler-Lagrange
Equations, forminvariant
under the Galilei transformations.
183
0) for fx=fo(x)+±2—\n\f0(x)\
(15.14)
we may have
G' = G'--P--
(15.15)
where 7 is an arbitrary constant, and K = K + J.
(15.16)
The latter set of solutions is not empty as we shall give examples where (15.8) accurs (Subsection 15.3). The Lagrange functions L, given by (15.6), give rise to Galilei covariant Euler-Lagrange Equations, which are linear combinations of the original equations (15.2) and (15.3). For g ^ 0 kept constant we have anI
+ a12II = ^ x ox1
+ G'R+P^x-^, ox ox
(15.17) ox
a2il + (T22II = G'x + 2gR + G"x2 + k ,
(15.18)
where a=(an
ai2
) =
_ U(Hl} 1 \
+
l°±poL
_ ^ l
2g(G'-G')
, 1
L
V
deta = a{Hl) ^ 0
G
+
| _ i ^ ^ : s (G'-G')G'
a.e.
V
(15.19)
j
' /
(15.20)
and V and Hi are given by (13.50) and (15.7) resp.. Since the trace of the cr-matrix should be a Constant of Motion (see Section 5.3) we infer from (7ll+(T22=a(ff1)
+ l+ §;
P = G'-G
(15.21)
that
P2 j - = Constant of Motion. (15.22) The case g J^ 0, g = 0 is easy to handle as any linear combination of elements of L with constant coefficients is again an element of L. It suffices
184
The Inverse Variational Problem in Classical
Mechanics.
to choose two proper elements of L belonging to the same g / 0 and subtract one from the other; then the g-term of the so obtained Lagrange function will vanish. The case g = g = 0 is rather simple to handle too. The most difficult case is that of g = 0 but g / 0; the model has to satisfy then certain restrictions: fr, given by (15.5), has to be of the form fR = iRfi(x) + IR,2{x)x2,
(15.23)
fRfl
=
~(G'fmfi+K),
(15.24)
fRfl
=
~(G'f,j
fx,o
=
- ~ ,
(15-26)
ha
=
~^G"
(15.27)
where
+ G"),
(15.25)
and the relation ^
= /»,o/fl,2
(15.28)
has to be satisfied. 15.2.1
Case
g^O.
According to (15.2)-(15.5) the functions fx and / # satisfy the identities d2U , d2U . -^Tfx+G'f R+-—xz ox ox ox
dU n — =Q, ox
G'h + ZgfR + G"x2 +K = 0.
,
15.29
(15.30)
Our goal is to find all possible Lagrange functions L = U{x, i)-kR
+ G'{x)xR + gR2 = 0
(15.31)
leading to Galilei covariant Euler-Lagrange Equations, having the same set of solutions as (15.2) and (15.3), and consequently to the some Newton's Equations (15.4) and (15.5). Taking this into account we get the following conditions to be fulfilled, viz. d2U, ~,„ d2U dU -^jfx+G'fR + ^—-x--—=Q, ox* ox ox ox
v(15.32)
'
All Euler-Lagrange Equations, forminvariant
under the Galilei transformations.
G'U + WR + G"x2 +k = 0.
185
(15.33)
Let us consider first the case when g ^ 0 and g ^ 0 too. Since g and g are constant quantities we may assume 9 = 97^0,
(15.34)
which does not restrict the generality of our investigation (scaling). Our first task is to solve (15.33). Let us subtract both sides of (15.30) from both sides of (15.33) resp.. We obtain P{x)fx + ^ ±
2
+ v = 0,
(15.35)
where P{x) = G'-G',
V = K-K.
(15.36)
If P = 0, (15.35) entails v = 0. Thus G' = G',
H = K.
(15.37)
When can we expect that /3 ^ 0 ? Should this be the case then we infer from the identity (15.35) that fx = fo(x) + f2(x)x2
,
(15.38)
where fo(x) = - ^ ,
(15.39)
hi*)^-\fx
(15-40)
We conclude that if fx, given by (15.4), has a different form from that given by (15.38)-(15.40) the quantity /3 and consequently v have to vanish. This is the most common case to be encountered. If, however, fx has the form given by (15.38)-(15.40) we may expect additional set of Lagrange functions corresponding to /3 ^ 0. Two separate cases have to be considered a) /o = 0: (15.39) entail v = 0 and (15.40) implies P = Aexp
{-Jf2{x)dx^,
where A / 0 is some constant,
(15.41)
186
The Inverse Variational Problem in Classical
Mechanics.
(3) /o ? 0: then 0 =~ , Jo
v^O
(15.42)
and /0/2 = ^
•
(15.43)
To summarize: we have usually the case (15.37); this can also happen when the Newtonian Equation (15.4) has the form (15.38), e.g. in case a) for A = 0 and in case fi) when v = 0. However, if the Newtonian Equation (15.4) has the shape (15.38) a new set of Lagrange functions is admissible, for which the Newtonian force has the shape a) fx = i2f2(x),
K = K,
G'=G
0) fx=fo(x)+x2-?-ln\fo(x)\, ox
+ \expt.-f
R = V + K,
f2(x)dx\
,(15.44)
G' = G - ■£ , (15.45) /o
where A nad v are arbitrary functions. We shall see scrutinizing some examples in Section 15.3 that the sets given in a) and fi) are not empty. So far we exploited only the equations (15.30) and (15.33). Now we are going to make use of (15.29) and (15.32) keeping in mind (15.34). We may write the Lagrange function (15.1) in the form L = L!+L2,
(15.46)
where Ia(x,x)
= U(x,x) - ^-(G')2x2
L2{x, x,R, R) = ^-(G')2x2 4ff
+ £-G,
- ^G 2g
+ G'xR + gR2 - KR.
(15.47) (15.48)
Notice that we may write L2 in the form L2 = gq2 - isq ,
(15.49)
q = R+—
(15.50)
G.
All Euler-Lagrange
Equations, forminvariant
under the Galilei transformations. transformations. 187
This simplifying substitution is, however, only valid for g ^ 0. We shall come back to that in Subsection 15.4. The Euler-Lagrange Equations for (15.46) are 2 var va.rflR L L :: G'x G'x + + 2gR 2gR + + G"x G"x2 + +KK = = 00, ,
(15.51) (15.51)
0| ( G ' i292 i0i 12 . i : | ( ^ ) - ^ » . ) 0=' 15 0.(15.52) '"•^slSrj-^+l' '^ ^ " » ^'' - 52 ) v a
t
Hence Hence according according to to d Ld_ ((dLA -^L\ dt\dx)
+
+
9
+G
i
+
(15.47) (15.47) and and (13.50) (13.50) dLl dLj -_ dx
-5-Ms=*-s
+ff
<
0V+
*>-» -
(15.53)
= 0.
If we compare (15.53) with (15.4) we see that L\ is a Lagrange function yielding (15.4). Conversely, let us assume that we have a Lagrange function L1=L Li =l(x,x), Li(x,x),
(15.54)
which gives rise to the Newton Equation (15.4) and a function G'{x) G'(x)
(15.55)
satisfying equations (15.33), keeping the restriction (15.34). Let us introduce the quantity L Z, = L Lix + LL22,
(15.56)
where x2 --^G -^G + + G'xR G'xR ++ gR gR22-kR.- h R . (15.57) \g \ ) 2g \g \ ) 2g Then L yields equations (15.4) and (15.5). To prove our conjecture let us use Euler-Lagrange Equations for L, viz. 2 2 L22 == ^(G'(X)) L — (G'(X)) X
G'x + 2gR + G"x2 + k = 00,,
5 i ++ 2 + 5< 5 1 ,i+! + s))= 0 !(§)-§ ++ | (<** ' **^ "" * =^-
(15.58)
<(15.59) 15 59)
-
188
The Inverse Variational Problem in Classical
Mechanics.
We conclude from (15.58) and (15.59) that d_ fdLi\ ggi !(§)-§-■ dt \ dx J dx = Q
(15.60)
According to our assumption L\ is a Lagrange function linked to (15.4). Consequently (15.60) yields (15.4) too. If we replace in (15.58) x by fx and confront so obtained relation with (15.33), we obtain (15.5), preserving condition (15.34). Formula (15.56) gives rise to all L of the form (15.31), leading to Equations of Motion, whose set of solutions coincides with the set of solutions of equations arising from (15.1), provided (15.34) is observed. In other words, all L and L lead to the same Newton's Equations (equations in the normal form). In Section 10 we learnt how to find the whole set of Lagrange functions V, s-equivalent to the given L. In the present case we have to replace L by Li and V by h\. Then using (10.26) and (15.47) we have Lx == x j"a{Ht(x,n))^*'"* f a{Ht(x,u))d2L^2'u) In
du - S(Jfft) £(#0 (15.61)
= £x- / r cr(ir 1 (x,«))D(ar,u)Ai-E(iri) J = 2— / aiHx (i, u))27(ar, u) du - E(if x ), where where
9 Jc
2 .dU - x— Ex _= a:.din ^ 7 - - TLi = x— (G 9a; ox 4g
2
- {/ - K— G , 2p 0
* - - 5 ~ ^ = -5-5""->'-w -
(15.62)
£&>-.<.,.
Finally by virtue of (15.61) and (15.57), we have X 2 L l = x* [[ a(tfi(*,u)) v(Hi{x,u)) ((~ 0 - i^-((G ' ) ) du - E(ffi)+ 3 7 \ (15.63) (15.63) 1 22 2 2 22 + - -(G'(z)) ^ GG + Gifl + gR RR +—(G'(
U(x,x) = x^ x f
l U U) U)
-^HlHl) ) -^
2 22 2 + j-g(G'(x)) ^(G'(x)) x -^ x -^G. gG.
9
duJ
(15.64) (15.64)
All Euler-Lagrange
Equations, forminvariant
under the Galilei transformations. transformations. 189
We are going to show how to get from (15.63) the a-matrix. start with the most frequent case 13 = 0, 0,
G = G,G,
K = k K = K
= g^0. 9=~9^0. g
Let us
(15.65)
T h e Euler-Lagrange Equations for (15.63) are II = G'x + 2gR + G"x2 +K + K = 0, 0, 11
v a rf fi i L: Z: var
~
[ ,„J&U
var.L:
[ ^ l ) ( ^ -
(G')2\
i
(15.66)
22 (G') (G') L 1].. -
2 r j + ^
j - +
G G + d + Gi+Gik
44g9 G' '"G"^+^-tdY-t9 + YgG=' + Glk = -%" + «*[{%-«£)» +,
=| " where
agi . fd22uU dHt _ .(d X \ dx2 ~ X
2
._&U_ ■ d U _ PGx^ ,GlG, Gl, dxdx 2g +
_dU_ K_G,dx 2g
2 {G') (G')2\ \
n
2g2 ) '
dm = .ffni_ Jk = ±^L
&GG„. __ 9L II±22 _ !E_du_ - ±G±G' ,
dx dxdx 2g dx dxdx 2g Consequently Consequently var x L = <j\il + <J\2lI = 0, var x L = o\\I + <J\2lI = 0 ,
where where
va,iR L = o\iI varj? L = a\il
+ (JiiII + o-nll
^(HO ff=M#i)
=
^)(i-|/r)=o,
-3x- = {^--2«7)> dx
™
(15.67)
dx 2g dx 2g
= 0, —0,
%{\-a{E ))\ ^ i(i-«m)y x
ddeta etu = = a(H cr(Hi). 1). It is obvious t h a t the trace of (15.71) is conserved.
, , ^
(15.68) (15 68)
'
(15.69) (15.69)
(15.70) (15.70)
(15.71) (1571) (15.72)
190
The Inverse Variational Problem in Classical
Mechanics.
In case a) (0 ^ 0), given by (15.44), we have fx=i2f2(x),
K = K,
r
(1573)
)
r
1
G = G' + Aexp I - / f2(x) dx \ = 0 + G'. The Euler-Lagrange Equations of L, given by (15.63), read v a r ^ Z : G'x + 2gR + G"x2 + K = 0
(15.74)
or, due to (15.73), varfi L: II + 0(x - fx) = 0.
(15.75)
By virtue of (15.4) we get from (15.75) var„i:
7/+ |
+
[ ( 2 ^ - (G'f)
2+
ff ov
*(ss*-f)- < H
=
i576
( »
We have further
var.L: ^ ) [^
- ^L j + i ^ - j
X+
= — ( G ' Z + 2ffjR + G"i 2 + K ) +
(15.77)
(G')2\
.^.r/^c/
+xpL_o_GI,i2_eu.G^Q ox ox
2g
dx
2g
Taking into account (15.74) and (15.76) we get from (15.77) ~
G' \2g0T
/
G'0\ ,
+a(Hl) I-^-Il] 2 9
„] (15.78)
=0.
All Euler-Lagrange Equations, forminvariant under the Galilei transformations. 191 191
Finally var.L:
(^.
+
,~ (G' G1
a(H1)\l+
~ ' G'G'0 G'G'p
(15.79) (15-79)
s
,,TT„,G'\ ,G'\
TT TT
n n
From (15.79) and (15.76) we infer that the
w
\ , G'0
„(II\G' „(ll\G'
,G' li x
\ M. V v \ p deta = a(H deter o-(Hi). 1).
G'G'0 \
_ 2L0 v v
/)
(15.80)
I'
(15.81)
The trace of (15.80)
02 1+ + ^^ )) + + ^ ^ is conserved (see Subsection 5.3). It is the right place to give an independent proof of that fe- is a conserved quantity. To show that notice that in cases a) and 0) there exists such a /? ^ 0 which fulfills equations (15.35). It is easy to check that L= L = j—x - i2 2—- ^— fpdx, 3dx , g^O, 9*0, 4ff 2gJ 4g 2g J is a Lagrange function leading to equation (15.35) and consequently to equation (15.4). From Section 10, using also the formula (15.47) for L\ as well as for L we get
d2L, _dU
(G1)2 _V _
. dL a{H)
-W=^-^lg--29-
W
_.p.P ~ a{H)Yg ■■
where a is a Constant of Motion as
dt
dt y ^ dx &c
J y
2g 25 V \
dx 3a;
J /
'
due to (15.35). Notice that from (15.82) follows immediately
(15.82) (15 82) '
192
The Inverse Variational Problem in Classical
Mechanics.
which accomplishes the proof. A similar consideration applied in case (3) (/3 / 0), given by (15.45), mutatis mutantis yields for the a-matrix the same expression, (15.80). To complete the case g / 0 let us inspect the case when g = 0. To this aim we use two Lagrange functions of the set L both with g ^ 0, say ZM = U(1){x,x)-h^R+{&l))'xR 2
+ gR2,
{2)
Z< > = U (x,x) - K&R + (G^)'xR
+ git ,
such that not all U^\ «W and ( G « ) ' are equal to U^\ resp.. If we subtract (15.83) from (15.84) we get 1(3) = fj{i)fa
±)
_ k(3)R
+
(15.83)
2
(15.84) h™ and (G<2))'
(G^yxR,
(15.85)
with the term „gR2" missing. A linear combination with constant coef ficients of two elements of L yields again an element of L and a linear combination of equations (15.32) and (15.33) as its Euler-Lagrange Equa tions. This means also that they yield the Newton Equations, (15.4) and (15.5). Hence H3' belongs to the set of L with g — 0 and is the desired Lagrange function with the missing term „gR2", unless I/' 3 ' = 3j|. 15.2.2
Case g = 0.
Let us first inspect the case 9 = 9 = 0.
(15.86)
In this case we have to have mandatorily G'?0,
G'?0
(15.87)
to conform to condition (13.50). Now the equations (15.30) and (15.33) have the structure of equation (15.35), viz. G'U + G"x2 + K = 0 ,
(15.88)
2
G'U + G"x + k = 0.
(15.89)
Then either a)
fx=i2f2(x),
« = 0,
G' = A e x p | - / f 2 d x \
,
(15.90)
K = 0,
G' = A e x p l - f f2dx\
,
(15.91)
All Buler-Lagrange
Equations, forminvariant
under the Galilei transformations. transformations. 193
or P)
=i2-^-ln\fo(x)\+fo(x), fx=x — \n\fo(x)\+f0(x), ox
K^O, n^O,
G' = = ~ , ~, (15.92) G' /o Jo
K ^ O , G' MO, G' ==-^-.- f . Jo
(15.93)
Hence G'=nG', G' = fiG' ,
(15.94)
K k == fiK, jUK ,
ft being a number which we may put equal to 1 without resctricting the gererality of our derivations (scaling). We have according to (15.1) and (15.31) L x) - KR++G'(X)±R, G'(x)xR, L = = U{%, U(X,X)-KR
(15.95)
L L= = U(x,x) U(X,X)-KR- KR + G'(X)XR. G'{X)XR.
(15.96)
The Euler-Lagrange Equations read , d2U . dxdx
8U .. dx2 9CJ„+G, dx2
|
d2U ± dx dx
_w__ dx dU dx
= Q
(15.97) (15.98)
and K= 0 G'x + G"x2 + K
(15.99)
as (15.88) and (15.89) coincide. For L(x, x) x)=L-L = L — L == U(x, U(x, x)x) -- U(x, U(x, x)x)
(15.100)
we get, by substracting (15.97) from (15.98), d2L.. dx2
d2L . dxdx
8L _ dx
(15.101)
To find L we use (15.101) and (15.99). We should have d2L „,, .,dG(x) ,8G(x) — =aM^L>.
(15.102) (15.102)
194
The Inverse Variational Problem in Classical
Mechanics.
From (15.102) we get
r
(15.103) l-G' f a(x,u)(x-u)du + h1{x)x + h2(x). L — G'\ a(x,u)(x — u)du + hi(x)x + h2(x). In the following we shall discard the term hi(x)x as a gauge term. If we insert L, given by (15.103), into (15.101) we get the identity rx
g
rx
G'—- I auduoudu-h'2 h2 -= bG"x aG"x22 + +&K, G" I/ audu + G'-—OK ,
(15.104)
where we made use of (15.99). Let us differentiate both sides of (15.104) with respect to x, We obtain -G"ax + G'^-x%(G"x2 + an) +&K)=0. = 0. (15.105) ox ox ox ox We are going to solve (15.105) by using the method of characteristics. If we introduce an auxiliary function $(x,x,a) $(x,x,a)=0,= 0,
(15.106)
relation (15.105) takes the shape - G'x^pG'x^- + (G"x2 + + &K)~
(15.107)
(15.108) (15108)
T G dx T=G
we we obtain obtain a = cG', cG',
(15.109)
where c can be a function of another partial solution of (15.108). For our purpose it is, however, enough to know only one single solution for &, as we learnt in Section 10 how to construct the whole set of s-equivalent Lagrange functions, starting from a given one in (1 + 1) dimensional space-time. So we restrict ourselves to c = 1. We are going to compute h2 using (15.104) and (15.109). We obtain ti2 = = -G'G'c -G'G'c22
-G'K
All Euler-Lagrange
Equations, forminvariant
under the Galilei transformations. transformations. 195
or h2 = -\{G') -l(G')22c22
-GK+
const.
(15.110)
Let us insert (15.109) and (15.110) into (15.103). We get I = \{G') L \{G'fx2x2 2
- {G'fcx (G'fcx - GK GK
(15.111)
or for c = 0 2 2 L == HG') \{G')2Xx2-GK. -GK.
(15.112)
The Lagrange functions (15.111) and (15.112) lead to (15.4). To get other solutions one has to use (10.13)
(T( = )(G )2 ^ ==.(H)g=^)( G')*,
S"
^S ^ ' '
(15U3)
(15.113)
where, according to (10.18) and (10.10), = zUG')2X + GK GK (15.114) ox ox is a Constant of Motion. To get the a-matrix we start with (10.26), (15.100) and (15.95). We get H=^X-L H = ^ X - L
L = x(G')2 I
a(H{x,u))du-Z(H)
+
U(x,x)-KR+G'xR(15.U5)
and varx L: L : 1I + a{H)G'II v a r TL L :: 11 77 = = 0. var^ 0.
= 0,
(15.116) (15.117)
Consequently (I
a(H)G'\
(15.118)
There is still pending the problem to be solved; what conditions have to be imposed upon the system when g = 0 but g / 0. If g = 0 we have to have obligatorily G' / 0. The equations (15.30) and (15.33) read 2 2 K = 0, G'fx+G"x + G"x +K 0,
(15.119)
196
The Inverse Variational Problem in Classical
G'fx + 2gfR + G"x2+K = 0.
Mechanics.
(15.120)
From (15.119) follows /* = /*,o(z) + /*,2(x)± 2 ,
(15.121)
f*fi = - § ,
(15-122)
ha = ~
(15-123)
where
Then from (15.120) we may derive / H , namely / * = /fi,o(z)+/*,2(*)± 2 ,
(15.124)
fRfi = -^(G'fx,o
(15.125)
where + ii),
fR,2 = -^z(G'fx,2 + G"). (15.126) 25 Keep in mind that g is supposed to be different from zero. Let us insert (15.122) and (15.123) into (15.125) and (15.126). We get G' 2gfR,o = ^ - K ,
(15.127)
2 ^ , = ~G>1 ( ^ ) .
(15.128)
If we differentiate (15.127) with respect to x and compare with (15.127) we get G ' ^ £ = -«/*,2
(15.129)
or, taking into account (15.122), ^ ~
= /x.o/ii.2 ■
(15.130)
This is the necessary condition for the existence of a solution involving g 7^ 0. We are going now to show that this is also a sufficient condition.
All Euler-Lagrange
Equations, forminvariant
under the Galilei transformations. transformations. 197 197
To start with, let us assume that (15.130) is, in deed, satisfied. Prom (15.127) and (15.128) we obtain G' «K = -2gf -2gfRRfi ,o + K— ,,
(15.131)
& = & = -2gG' -2gG'I I2& J^-dx. dx .
(15.132)
We have to check whether k is really a number and that (15.131) and (15.132) solve the equations (15.125) and (15.126). The equation (15.125) is trivially satisfied, taking into account (15.131) and (15.122). By virtue od (15.131), (15.132) and (15.130) we have dk
~dfR,o 0r.-dfufi
Tx = -29-dx~+
,
d (G'\
.dfRfi 00-<9//?,o
,2 „-/ H SR,2
G7 = -29~dx--2^-GT
s - * - s - « E ^ J = - ' r s ~ * « - s r ==( u(15.133) -ufl +R
Tx
= - t f ( ^ - / . ^ ) = 0 . Since, according to (15.131), k does not depend on x, it must be a constant. Let us now check on the equation (15.126). Using (15.132), (15.131) and (15.123) we get f & = G" = -2gG" -2gG" jI —"dx~^dx- 2gf 2~gf = ^G"G' ^G"G' - 2~gf 2gfR,2Ra Rft = R>2 G G (15.134) J (15.134) = —fxfiG —fxfiG - 2gf 2/fi, R^22 ,
which proves our assertion. For the case j = 0we have L leading to (15.119) or (15.121), viz. (15.135)
L = & X 22-KG. - K G . L=^Y-X
2
Consequently the most general solution for L is a(H(x,v))dv-■£(&)+ L = x{G')2 f[ a(H(x,v))dv - £(ff)+ Jo0 2 2
+ ^-(G') x 43 4c;
- ^G + G'xR + gR gk - kR, --^G 2g
where we used (15.56), (10.26), (15.57) and
g-«*>•*
22
(15.136)
198
The Inverse Variational Problem in Classical
Mechanics.
H is given by (15.114). To summarize: dealing with a Lagrange function L for which g = 0 we are able to construct a Lagrange function L involving g ^ 0, if and only if the function / R is of the form (15.123)-(15.125) and the condition (15.130) is satisfied. It is worthwhile to notice that the case of two particles in a (1 + 1) dimen sional space-time as well as the case of a time dependent point particle in a multidimensional Euclidean space was treated extensively by Douglas*' . His approach differs from ours. 15.3
Examples.
In this Subsection we are going to present some examples concering former Subsections and to convince the reader that the sets given in a) and /3) (formulae (15.11)-(15.16) are not empty. Example
of type a) (15.11)
(15.13).
L = \x2 + \R2
(15.137)
i.e. U=\x2,
K = 0,
G'=0,
g = \,
V= 1.(15.138)
The Equations of Motion are var^L : l = x = 0,
(15.139)
var f l L : II = R = 0.
(15.140)
Since
U = \x2 then Hi = \x2 is a conserved quantity. *' J. Douglas, Trans. Am. Math. Soc. 50 (1941) 71.
(15.141)
Examples.
199 199
Because of (15.139) we have the case a), /3 ^ 0 2 x = x2f2x(x), h{x),
/2(z) / 2 ( z ) = 0, 0,
K =00, K = , (15.142)
G' = /3 = A A ## 00 ,,
2
^ =A.
v
First we consider the case g = g = | . According to (15.63) f*
.
.
2 2 2 22i 2 + MR + gR22 L a{\v22))dv-Y,(\x dv - E(ix ) + + i\\A L= =x x I a{\v ) x + XxR + gR Jo
(15.143) (15.143)
The Euler-Lagrange Equations of (15.143) read
Hence
[a{\x22)) + \2]x + \R = 0, [a(±x
(15.144)
Xx + R = 0. 0.
(15.145)
.= (^ + A 2 J), J).
(15.146) (15.46,
deter let a = cr(|i; a(\x22))
(15.147)
-(^
2
A
) +A 2
in accordance with (15.80) and (15.81). To get the case where the term g = 0 we choose in (15.143) two functions i
~ = \x r odv-\Y,+ A—x2+\xR+\R2, L^ L(1) = \x / a d w - | E + — i 2 + A i ^ + | i ? 2 , Jo 2 7o 2
2 2 2 L< = -\x - § i [/" adv+^+^R adw+§£+|i? L{2)) = , , Jo Jo
(15.148) (15.149)
as a and A are arbitrary. We subtract (15.149) from (15.148) and we get L= =x x L
fx 2 a{\v )dv-Y,{\x22)) a{\v2)dv-Y,{\x Jo Jo
A2 22 + —x —x + + \xR. \xR. + 2 2
The Equations of Motion are [o{\x2)2) + \A22]x [cr(|a; ]a; + \R A/i = Q, 0,
(15.150)
Ax = 0.
(15.151)
200
The Inverse Variational Problem in Classical
Mechanics.
Consequently
a=(^2A)+A2 J). For special choice u{\x2)
(15.152)
= —A2 = —1, A = 1 we have
L = xR
(15.153)
and for 5=|,
ff=
(!
(15154)
i)
as well as for
5 = 0,
a=
(°i
< 15 - 155 )
I)
The latter means that the equations (15.139) and (15.140) are interchanged. Notice that (15.137) and (15.153) are not equivalent. Example
of type a)
L = ±ax2e-2ax
(15.11)-(15.13). - be-ax + cRxe-ax
+ \dk2 + —R,
ad - c2 = Ve2ax + 0.
(15.156) (15.157)
Hence U = \ax2e~2ax 9-
id, z
- be'ax ,
K=
G = --e~ax
aM c
, (15 158)
'
-
The Equations of Motion are varx L : I = a{x + ax2)e~ax vaiR L:II=
+cR-ab
-c(x + ax2)e~ax
= 0,
(15.159)
+d{R-—)=0.
(15.160)
The normal form of (15.159) and (15.160) is x = ax2 ,
(15.161)
Examples.
R=—
201 201
(15.162)
c
c resp.. Thus we have the case a) with 0^0. Hence resp.. Thus we have the case a) with 0 ^ 0. Hence __ abd 9 = 9. abd ~ c +a A t_t X„ x c , K=K= G= e , 9 = 3. c a P ~ adA-2 c2 ' £!
(15.163) , 15.163
2 ' We haveValso~ the ad-c Constant of Motion We have also the Constant of Motion
^i = ^
= |
5
(--^e-a-=flx.
We construct L according to (15.6) or (15.63), viz. 2 nrl - r 2
ff*x
L = ^±-e-2"* d
(Hl (x, / aa'Hx (x, v))(x v))(x -- v) v) dv+ dv+ Jo J0 (15.164) (15.164)
2 2 ax ax e- 2 ^ --b-{c + ^-Acc ++ A) \)2x2±e-2ax -{c++\)e\)e~+ + + i_( 9/7 r 2d c a a * * ++ ^R R . + (c ++ \)xRe\)xRe±R22 ++ ^^R.
For a = 1, c2 = a — 1, a > 1, d = 1, A = —c, a = 0 we have L=\x2 + \R2
(15.165)
as in example 1. The (T-matrix for equations (15.159) and (15.160) reads /
^-<*) l )
- < j ( # i ) $ e - Q I + £±Ae"ai
( c ++ A* )-e *—+A *<*
Ac(c+A;
" —
aZ> a
\
\
dXe- * ^ ^
\
V
/a(ffO
1 . +
ad-c2{
e
— 2 1
, Ace" " 1i + ^ ^ 1
[-a(i?x)^ + ^ ] e — \
"*"
)+
d(,ax
\
^
+
c
J ■
X)
/
(15.166) (15166)
202
The Inverse Variational Problem in Classical
Mechanics.
For A = 0 (15.166) reduces to a=('T{f)
$(l-*™«-~)
(15.167)
and for a{H\) = 1 we get an identity matrix as it should be. Example
of type /3)
(15.14)-(15.16).
x2 L = — + ax + \x2x2 + xxR + \R2 .
(15.168)
x2 U = — + ax + \x2x2 , K-0, 2 a » ■
,„ (15.169)
Hence
G=\x2,
g=\,
23 = 1.
The Equations of Motion are v a r x L : I = (1 + x2)x + xR + xx - a = 0 ,
(15.170)
va,TR L: II = xx + R + x2 = 0
(15.171)
or in normal form x = a,
(15.172)
R= -x2 -xa.
(15.173)
Because of (15.172) we have the case (5) with /3 =fi 0. We have x = /o + i2 — In |/o|, X
1 9
fo = a,
K
G=ia:2--i,
,
« = ./,
3 =3 = | ,
tf2
i/2
^7 = — ■
(15.174)
L1=U-±(G')2x2 + ^G=^+ax.
(15.175)
i2 Hi = —— ax
(15.176)
is conserved.
Examples.
203
Let us construct L using (15.63) rX
L=x Jo
n
a{Hl{x,v))dv-Y,{Hl{x,x)) 2
-v Ux
x2-
+ \(x--) K 2
- -x) + (x - -\ xR + \R
aJ
(15.177)
- vR.
The cx-matrix reads ( <*{Hi)-i{x-i)
-a(H1)x
+ x-^
+ ^x(x-^a)
a=\
\ .(15.1'
For 1/ = 0 and u(77i) = l w e get the identity matrix. The standard
model. 2
L = \x
- V{x) + \R2
(15.179)
or U = \x2 - V(x),
K
= G' = 0,
g = \.
(15.179)
The Equations of Motion are vaixL:
dV I = x+-— = 0 ,
(15.180)
v a r f l L : 77 = R = 0 .
(15.181)
Li = | x 2 - V(sc),
(15.182)
Hi = \x2 + V{x),
(15.183)
Since
then
which is conserved. Because of (15.180) we have the case (3 = 0,
G'-G'=0,
K = K = 0,
g = g=\.
(15.184)
L = x f a(Hi {x, v))dv - £(i?i (x, i)) + \R2 . 7o
(15.185)
Thus
204
The Inverse Variational Problem in Classical
Mechanics.
The <7-matrix is a
={
i)
o
For a(Hi) = Hi we have e.g. i=
Yi
+
L = J2^-
15.4
Yv
+
^2'^2-
arctan [ J ^-x
(15J86)
) - In (§x 2 + V) + | i ? 2
(15.187)
Generalization of the set of Lagrange functions L (ad mission of Euler-Lagrange Equations not covariant un der Galilei transformations).
In the previous Subsections we were interested in such Lagrange functions L, whose Euler-Lagrange Equations were Galilei covariant and led to the same set of solutions as the original equations, we started with. In this Subsection we are going to construct a set of Lagrange functions, say Lex, whose Euler-Lagrange Equations are no longer Galilei covariant, but still lead to the same set of solutions as those Lagrange functions be longing to the set L. To start with, let us turn back to formulae (15.46)-(15.50). For g ^ 0 we have L = Li + L 2 ,
(15.188)
where Li(x,x)
= U(x,x) - ~(G')2x2
L2(x,x, R, R) = j-{G')2x2
--^G
+ £-Q, + G'xR + gR2 - KR.
(15.189) (15.190)
If we define the new variables q = R+—G, "9
(15.191)
Generalization
of the set of Lagrange functions
L...
q = R+~G',
205
(15.192)
we may write L2 in a concise form L2{q,q) = gq2-Kq.
(15.193)
With this notation the Lagrange function L, (15.188), can be written as a sum of two Lagrange functions, Li(x,x) and L2(q,q), depending each on distinct variables, the interaction term missing. Thus we may apply to both terms of L the procedure established for one-partice case in Section 10. For L2 we have H2(q,q) =gq2 + nq = ,
(15.194)
= gR2 + G'xR + — {G')2x2 +
KR+-^G.
H2{q,q) is conserved as var, L2 : 2gq + K = 0.
(15.195)
The extended set Lex reads Lex = LX+L2,
(15.196)
where L1=x
rai{Hx(x,v)) Jo
f 5 r \dv2
L2 = 2gq [ a2(H2(q,v))dv Jo
{
^P-) 2g )
dv
~X2(H2(q,q)).
(15.197)
(15.198)
We may replace in L2, given by (15.190), the original quantities G and K by G and k resp., keeping g / 0 constant. Thus we may introduce new variables q and q instead of q and q, viz. q = R+^-G 2^
= q+^f0dx, 2g 7
^=i?+^G'=g+-^/?i, 2ff 2g
(15.199) (15.200)
206
The Inverse Variational Problem in Classical Mechanics.
H2=H2(q,q)=gqi+£q.
(15-201)
H2 is also conserved. The Lagrange function Lex = LX+L2
(15.202)
consists still of a sum of two Lagrange functions depending on different set of variables. While L\ does not change we have L 2 = 2gq~ f Jo
a2(H2(q,v)) dv - X2(H2(q,q~)).
(15.203)
To get the Euler-Lagrange Equations we may treat L\ and Lx separately. We have v a r x Z i : a1(H1)I-a1(H1)^-II
= 0,
(15.204)
where / and / / are given by (15.2) and (15.3) resp., va,rx L2 : a(H2(q, q))j-
[G'x + G"x2 +k + 2gRj = 0 .
Now compare relations (15.74)-(15.76) (similar considerations applied in case 0) yield the same result), viz. G'x + G"x2 + K + 2gR= ^-1
+ (l - ^ - \ 11 = 0.
(15.205)
Hence ~ vaxx L2 :
■ G18 CT2{H2(q,q))—I+ G> f
+a2(H2(q,$)-\l-^)lI
(15206)
G'B\
= 0.
As L\ does not depend on R and R it gives no contribution to the variation with respect to R. We just have va,iR L2 : cr2(H2{q, q)) (G'X + G"x2 +K + 2gR^j = 0
Generalization
of the set of Lagrange functions
L... L...
207
or, taking into account (15.74)-(15.76), v a r f l L 22:
aa2{H (q,q))~-I+ 2(H22{q,q))-^-I+
+a (q^)(^-^jn +
(15.207)
= 0.
Finally we get ~ G18 oi{Hi) + a2(H {H22)—— ) —
~ varxLex:
v
1+ JJ
\ L
C" -a\{H{)-—h -ax(Hi)-—h 29
L
5
(15.208) (15.208)
♦*<*>£ (i-^)]"-o. varflL Z e x : a2(H2)^I
+ a2(H2) (l - ^ - \ II = 0 ,
(15.209)
where H2=H2(q,^). The most general
-ax{H,)%+d2{H2)%
(l - ^ ) \
a= \^
2
a2(H2))if-f
a2{H2){i-^-) ){l-^)
deta = <jlcr {Hl(H (H 1)d12)d 2(H 2).2).
, (15.210) )) ' (15.211)
It is obvious that the cr-matrix (15.210) depends explicitly on R and R and so the Euler-Lagrange Equations, originating from Lex, break the Galilei covariance of the original equations. IfCT2= 1 we get the c-matrix (15.80). For /? = 0 a=\
f/ aMHi) i(ffi) V V
00
((ffff33{{H Haa))--ooii((H Hii))))gg ff
where H2 is given by (15.194).
(H2) 2(aff2»)
\\
, /) '
(15.212)
208
The Inverse Variational Problem in Classical
Mechanics.
It is easy to see that the matrices (15.210) and (15.212) are not the most general tr-matrices which would correspond to the Euler-Lagrange Equations yielding the same set of solutions as the original equations but not conforming to the requirement of Galilei covariance. To get complete set of such Lagrange functions one has to solve the Helmholtz Equations (Subsection 5.1). 15.5
Case of Galilei forminvariant Newton's Equations cor responding to Euler-Lagrange Equations which are not Galilei covariant.*) (formulation of the problem)
In Subsections 15.2 and 15.3 the starting point for searching for s-equivalent Lagrange functions was a certain Lagrange function which led us to EulerLagrange equations forminvariant under the Galilei transformation. The set of equivalent Euler-Lagrange Equations encompasses also equations which are not forminvariant under the Galilei transformation. It is obvious that the Galilei covariant Euler-Lagrange Equations put in the normal form (Newton's Equations) must be also forminvariant. No tice that these equations in their normal form remain unchanged for each element of the set of all Euler-Lagrange Equations of the s-equivalent La grange functions, we derived in Subsection 15.2 and 15.3. These Newton's Equations inserted into any system of Euler-Lagrange Equations yield a system of identities. This implies that in case we are given a system of Galilei covariant normal equations and this system, inserted into an arbi trarily chosen system of Euler-Lagrange Equations, discussed in Subsec tions 15.2 and 15.3 (see e.g (13.46) and (13.47) or (13.43)), does not satisfy these identities, then no one of Lagrange functions and no one system of Euler-Lagrange Equations, corresponding to these Lagrange functions, is sequivalent or equivalent to a Lagrange function of type (13.43) or Equations of Motion of type (13.46) and (13.47). This in turn implies that no one of these new set of Euler-Lagrange Equations can be forminvariant under the Galilei transformation. To make this plausible let us investigate the following example. Assume we are given the Newton's Equations x = (1 + x)2 = fx(x),
(15.213)
*) The idea comes from Dr. J. Cislo. 1
Case of Galilei forminvariant
Newton's Equations... Equations...
R= =x= = fR(x). fR(x).
209
(15.214)
Obviously these equations are forminvariant under the Galilei transforma tion. The most general Lagrange function leading to Euler-Lagrange Equa tions forminvariant under the Galilei transformation is given by formula (13.43), viz. L = U(x, U(x,x)±)-KR - KR + + G'XR G'XR + gR gR22
(15.215)
and the Equations of Motion are d2U d2U.. _,> _,——x+—jx ——x+ + G'R1—x 1 ox ox ox
dU nn — =0, =0, ox
, 15.216 (15.216)
G"x2 + + G'x + 2gR + K = 0.
(15.217)
If we insert (15.213) and (15.214) into (15.217) we get (G" + G')x2 + 2G'x + G' + 2qx 2gx + K = 0
(15.218)
ga == KK
(15.219) (15.219)
or = G' G' = = 0. 0. =
Let us insert (15.219) into (15.215). We get L = U(x,x). U(x,x).
(15.220)
Equation (15.216) reduces to d2U dx2
2
d2U dxdx
±
9V_^Q dx
(15.221)
which can be viewed as an equation for U. The solution of (15.221) inserted in (15.220) would reproduce (15.213) but not (15.214). Thus (15.215)(15.217) are incompatible with (15.213) and (15.214). We may, however, find a Lagrange function which will be in agreement with (15.213) and (15.214). This Lagrange function L will be not of the type (15.215). The same concerns (15.216) and (15.217). The Lagrange function we are talking about is for 1 + x / 0 L =
~^TTi
(1
+x
~ \iTi)
+R
+ R( (]nl n 1 1
15 222 ^ - I I + * ! - * ) ■ ( 15 - 222)
210
The Inverse Variational Problem in Classical
Mechanics.
We have var f l : —*— x - x - 1 = 0 , 1+ x 1
var^ : - —
^r
(15.223)
/• l +X j R + r - k T,
1 ^
\ .. , x+
1 + x) 2 I 1 + x 2(l + x)2J „ + 71 +^ x + 25{l7 7+^x)22 + 5 72l+x ^ - + * = °"
(15.224)
Equation (15.223) is Galilei covariant, (15.224), however, is not. (15.223) follows
From
X
(1 + x)2
= 1 ,
hence (15.224) reduces to - l - z + 4r+ii+77-^ + I =0 2 2 2 1+x l+x or
If we compare the structure of the Lagrange function (15.222) with (15.215) we see that the structure is essentially different. The same applies to (15.216)-(15.217) and (15.223)-(15.224). Another example similar to the former one is given by the Euler-Lagrange Equations x + (dx + C0)2 = 0,
(15.225)
R-^r-. -^--CIR\ Cix + Co
v
.„ , ^ + l l =0, \(Cia; + Co)2 J
(15.226) '
belonging to the Lagrange function L = — ln(Ci± + Co) - R{dx
+ Co),
(15.227)
where C\ and Co are some constants. Although equation (15.225) is Galilei covariant, equation (15.229) is not. Of course the Newton's Equations are both Galilei covariant. The structure of the Lagrange function (15.227) differs essentially form that of (15.215).
An Outlook. Application in the Feynman Approach to Quantum Mechanics. Mechanics.
211
We wind up this Section with a statement of J. Cislo.*) We quote his result without proof. Assume that we have a system of equations x = f(x,x), f{x,x),
(15.228)
R = — h{x,x) h(x,x)
(15.229)
in their normal form. They are obviously invariant under the Galilei transformations. Cislo asserts that this system of equations is equivalent to a certain system of Euler-Lagrange Equations of some Lagrange function iff det det
fa fa \ \\ \
0P
da da dt dt
d§_ __ i u dh , d0 dt _ i2 u 9dhi ~, dt 2 9i ~
\\ lf>d£ j lf>d£ 2^ dx I 2^ dx I
= = 00,,
(15.230)
where a
=
d±_ldLdj_ +
=dx--idtdi 2dtdx ~dx
13
dh 9h 1x d dh dx ~2 dtdx = ~dx~ 'Jtdi
\_(djy
4{di) 4\dx)
'
\df\dj_dh dh 4dx dx 4didi
+
((15.231) 15 231)
-
(15.232) (15 232) '
and
— = tJL.
-JL 1L
dt dx dx dt is the substantial time derivative. 16
(15.233)
An Outlook. Application in the Feynman Approach to Quantum Mechanics.
The approach to Quantum Mechanics proposed by Feynman is known as the ,,integral over all paths" method.t) It is based mainly on the following observation. Let us consider a classical system, the development in time of which is described by a Markovian process. The time dependent probability distribution function reads P(h,t P(b,t2\a,h). *' Private communication. t) R.P. Feynman, Revs. Mod. Phys. 20 (1948) 367.
(16.1)
212
The Inverse Variational Problem in Classical
Mechanics.
Formula (16.1) is the conditional probability to find the system under con sideration in the state b at the time t2 under the proviso that it was in state a at the time t\. This probability depends neither on the history of the system prior to t\ nor on the history of the time interval (t2 — *i). For such a probability distribution function the Smoluchowski Equation holds, viz. P{b, t2\a,h) = J2P(b>*2|c, t')P(c, t'\a,h),
h < t' < t2 .(16.2)
c
Let us now consider a quantum system described by the transition am plitude ip(b,t2\a,ti).
(16.3)
We have P(b,t2\a,t1)
= \
(16.4)
But instead of (16.2) we have 2
p
Qu
(M2|Mi)= £ > ( M 2 M > M ' M i )
,
(16.5)
c
or ip{b, t2\a, tx) = J2
(16.6)
c
The reason for the discrepancy between the classical and quantum descrip tion is that in classical picture of Nature the system must pass on his way from a at time t\ to b at time t2 through a certain classical state c at time t', {ti < t' < t2) if its space-time trajectory is continuous. This kind of reasoning can not be accepted in the quantum picture of Nature, as we did perform measurements of the state of the system at time ti and t2 but not at time t'. If we want (16.2) to hold true also in quantum picture of Nature we have to perform a measurement at t' too. For a particle moving in a three dimensional Euclidean space from (16.6) follows by iteration
(16.7)
An Outlook. Application
in the Feynman Approach to Quantum Mechanics.
213
Feynman postulates for infinitesimal At (we put h = 1) = A~lexpliL
ip{x',t + At\x,t)
(x, * ~ * j At\
(16.8)
So the contribution in (16.7) comes from xs close to xs+i, s = 1 , . . . n — 1, because of the strong oscillations of (16.8). Here L = L(x,x) is the classical autonomous Lagrange function. For a nonrelativistic particle (of mass=l) in the potential field V(x) we should have according to Feynman
(sy^M*^-™)*}
(169)
In the last formula we could also write V(x') as well as \{V{x!) + V(x)) instead of V(x). To justify Feynman's idea let us take the simplest model considered above for which H = | p 2 + V(x),
(16.10)
where p is still a classical momentum and start directly from the first principle of Quantum Mechanics, namely p(x',t2\x,t2)
= (x'\exp{-iH(t2-t1}\x.)
,
(16.11)
where (• | denotes the bra-state and | •) the ket-state of Dirac and H = H(x,p),
pi = - i - ^ ,
i = 1,2,3
(16.12)
is the Hamilton operator. We have (x'\exp{-iHt}
\x) = lim / d 3 x n _i . . . / d 3 x x • At ^°J J
(16.13)
■(x'te^^IXn-O-.-lxiXxile-^Mx). Let us use in (16.13) the formula e(a+b)t
_
Um
f e 4 e *n ] .
n—>oo \
/
( 1614 )
214
The Inverse Variational Problem in Classical
Mechanics.
Then, taking into account (16.10) we get (x'\exp{-iHt}\x)
=
/ d 3 x n _ i . . . / d3xi
= Jim Atn=l J
J
A
.{x'|e-^ *e-
iVAt
|xn_1)...|x1)(16.15)
•(x1|e-l^Aie-iVAt|x) = = lim
/ d3xn_i ... / d3xi •
n—►oo J
J
■ (x'| e~'^At
| x n _ ! ) . . . |xi) ( Xl | e - J ^ A t |x) ■ e~iVt
Let us concentrate on (x'|e-^At|x)= /d3p(x'|e-^At|p)(p|x) = ,
(16-16)
2
= yd3pe-l^At(x'|p)(p|x). As 3
(P|x)=(^)
2
(16.17)
we may write 2
(x'|exp{-iyAO|x) = =
{h)
/ d 3 p e x p{- ! ^p 2 - i p( x '- x )} =
= {h)'j *>-{-%{* + &<*■» , (x'-x)2 (Ai) 2
+
, i(x'-x)2\l 2 A< ;j
-(s)7"-{-"(-"^)>
(16-18)
An Outlook. Application in the Feynman Approach to Quantum Mechanics. Mechanics.
215
To evaluate the integral / ^ e x p | - ^ (
P
+
^ )
2
|
(16.19)
we use the saddle point method and get eventually f . A t .2- l , . ,„ . . v a ( i (x'-x)2l (( xx '' || ee xx pp || -- iz— ^ pp 2J| || xx )) = = (( 22^ ^A A 0 t e ^ ) - 2 e x p x| - vp | ^—^ '^ | j
(16.20)
Finally (x',i
-(dsy-K^-^M" ■(ski),-'K»fi^-vw)AI}=
(16 21 (16.21)
-»
= (ds)'-{«(-^)MThe formulae (16.9) and (16.21) coincide. The derivation of (16.21) from (16.11) holds for all Lagrange functions, equivalent to the Lagrange function L = | x 2 - V(x). If we namely multiply I, by a number a then
t
also gets an a-factor; this amounts, roughly speaking, to a change of the value of the Planck constant from h = 1 to h = a. Moreover, adding to L a term dt (we stick to the autonomous Lagrange functions) causes only a change in the phase of the transition amplitude.*' t) *' N.B.: it can happen that this change of the phase can have sometimes some physical quantum effects. t) I enjoyed a discussion with Dr. Z. Haba on these topics.
216
The Inverse Variational Problem in Classical
Mechanics.
Let us assume that the relation (16.8) postulated by Feynman holds true irrespectively of the structure of the classical Lagrange function. Then the following problem arises. The classical Equations of Motion have closer links to Physics than the Lagrange function, as the dynamical behavior of the model is determined by the Equations of Motion. But, as we learnt already, there are often many Lagrange functions yielding the same equivalent Equations of Motion. Which Lagrange function should one then choose? For the simplest (1+1) space-time dimensional case we have L = \x2 - V(x). V(x).
(16.22)
The Lagrange functions s-equivalent to (16.22) are 382L
* /
a(H)\i=u—
du-11(H)
(16.23) (16.23)
=
-C*&L*-™ where = a(z) S4-.W
H = \x2 + V(x), V(x),
^
(16.24)
and S is an arbitrary function. From (16.23) follows H
~Xdx
L
X
dH(H) dH .
JC
dH
±=u
, /"* dE(ff)
dU
+ X
,
dH , ,
or H' = EE((H #).
(16.25) 2
l
For Y,W(H) YS \H) = \H L (D =
i^±* ±
+
we have ±_ | i 2±y2(V{X) v{x? __ a ; ) _ ii_ v{xf
(16 26)
(16.26)
2
for E< Jp){H) >(if) = lntf we have for V(x) V(ar) > >0 Li2) L (2>
= = i ^ ^a ar rcctt aa n {j§^j ( ^ j )
- In (§± 2 + V(x)) V(a:)) .
(16.27)
An Outlook. Application
in the Feynman Approach to Quantum Mechanics. Mechanics.
217
The question arises: are the expressions
A
?"{'{i!lsf-rM)"}
and
A exp ^"■4 «» {' (s (aww1 1+ 5+ ^w s ^w1! v' w-"^
At\x,t)
is concerned? Let us have a closer look to the case (16.26). We have for the canonical momentum w ppW
- = = \x3 +\±^iV{x). = ^ — xV{x).
(16.28)
Equation (16.28) is an algebraic equation of 3-rd degree with respect to x. For V(x) > 0 we may introduce new variables _ gW 1 ± = v^IvT*. r'=§rjgft-vm. " \v\JWV
(16.29) <#>*>
Then (16.28) reads t3+t-r + t - r = 00. .
(16.30)
We are only interested in t real as a function of r. We have dt_ _ _ 1
?dr = T-K^ > 0 1 + Zt2 >
dr l + 3t2 for t real. For r = 0, t (1) = 0, *<2'3) = ±i and dtW dt' 1 )
(16.31) (16-31) '
v
= 1
So t^(r) is a monotonously growing function crossing the point t = 0, r = 0 at angle 45° and for large r behaving like i(1» = (r)3 t(D (r)i
(16.32a)
218
The Inverse Variational Problem in Classical
Mechanics.
or 1) x = (6p( )')s x= (6p (1)
For large p^ H^
(16.32b)
the Hamilton function (V ~ 0) is = \H2~\tfp^ ~ \tf p{l)*
f ~1.36p^ .
For V(x) = 0 the dependence of x on p ' 1 ' is given by (16.32a)-(16.32b). For V(x) < 0 we have = 00, , t33 -- t -- rr =
(16.33)
where t and r are defined as in (16.29) . In this case A _ 1 dr ~ ~ 1 - U2
(16.34)
Forr = Owe 0 w e hhave a v eti(( 11 ) = 0 ;;ii(<2 )2 = ) =l ,li,(i3()3 )== - l1, , *™ &£■ dr
|i . In In the the interval interval
= = -1, ^ r=0
dr
= r=0
2 / T < r << 22 // T 7 3 3V3 V3 the behaviour of t(r) is singular in the sense that to each r belong 3 real values of t. For the domain r > | v / | and r < — \\J\, there is a one to one correspondence between iP-' real and r. Two roots of (16.33) t^ and t (3 ' are in this domain complex conjugate. The shape of i' 1 ' is similar to an stretched capital letter "S". The asymptotic behaviour for large r is given by (16.32a)-(16.32b). If V(x) is bounded from below and V(x) < 0 we may add to V(x) a constant A > \V(x)min\. Then we have the first case considered by us, V(x) = V(x) + A > 0, and the relation between real p^1' and x is one to one. Another example which can be pursued in more detail than the former example is (3+1) dimensional model. We have 2 L L= = §x2 - U(\x\) U(\x\)
(16.35a)
7 ( 2|-U(\x\))+s^. L' = g{\x2 _ J g(^ x|))+eU.
(16.35b)
and
An Outlook. Application
in the Feynman Approach to Quantum Mechanics.
219
Here
F =
' K?) 2 -^ + ^ + f / ( | x | ) -
(i6 36)
-
where Ji = - ( x x p ) i ,
» = 1,2,3
(16.37)
and e is a parameter. To get H' one has to replace p by p and J by Ji = | - [ ( x x p); + (p x x)i]. Both expressions (16.35a) and (16.35b) inserted into Feynman's formula should yield the same result, which is not at all obvious. We leave these questions open for further investigation. If it should happen that for two inequivalent but s-equivalent Lagrange functions the expressions for the transition amplitude should be essentially different i.e. the quantization procedure would be different in both cases, that would indicate that the Equations of Motion do not determine the quantization scheme; the decisive factor would be then the Hamilton functions and the connected with it canonical formalism. A possible way out, which would make sense in Physics, would be to prove that two Feynman transition amplitudes, belonging to two inequiva lent Lagrange functions, give in the limit of time tending towards (+) or (-) infinity resp. the same results (up to an inessential phase factor). In other words the use of two different Lagrange functions would not affect the Smatrix. To check whether this can really happen needs a lot of work. The hope that this is feasible comes from Relativistic Local Quantum Field The ory; there Wess and Zumino made some calculations based on perturbation theory and the theory of Borchers classes (based on Axiomatic Quantum Field Theory) and got reasonable, promissing partial results.*' I cannot say anything more ... ,,So steh' ich hier, ich kann nicht anders, helfe mir Gott. Amen," 7 '
*' I am thankfull to Dr. J. Wess for a discussion. t) Martin Luter in Worms (1521).
Index
Euler, Theorem of — 2-3 Euler-Lagrange Equations (Equations of Motion) 13, 22-23
Acceleration 4,9 Boost 4 Boost, Infinitesimal — 6 Bracket, Lagrange — 35-36, 43, 67 Bracket, Poisson — 33-36, 43
Feynman's Method of ,,Integral over All Paths" 115, 145, 211-215 Force 8-9, 12-13 Forminvariant Vector 54,57 Forminvariant Tensor 57
Cartesian Coordinate System 1-2, 9, 14 Central Charges 7, 167, 182? Characterises, Method of — 19-22, 47-48 Constants of Motion 14, 191-192 Constraints, Holonomic —, Anholonomic 10
Galilei Group 4, 156, 164, 167, 179-180 Galilei Group, Infinitesimal — 5-7 Gauge Term 39 Hamilton Formalism 66 Helmholtz's Conditions 23-29, 30-32, 208 Helmholtz's Conditions, Douglas Method of Solving — 30-32 Henneaux, Theorem of — 33-39
Degrees of Freedom 10 Dirac Monopole 63, 65 Douglas' Method of Solving Helmholtz's Conditions 32 Dynamics, Analytical — 7 Energy, Kinetic — 12 Energy, Potential — 13 Equations, Equivalent System of — 29, 33 Equations, Linear Partial Differential — 19-22 Equations of Motion 13, 22-23 Euclidean Space 1,4
Kinematics 1-7 Lagrange Lagrange Lagrange 14 Lagrange 33, 39 220
Equations 10-13 Function 13-14 Function, Autonomous — Functions, Equivalent —
Index
Lagrange Functions, s-Equivalent — 33, 39 Lagrange Function, Weakly Symmetrical — 53 Lorentz Force 64-65 Mass of a Particle 9 Matrix, a — (and Its Trace) 40-43 Newton's Laws (Equations) 8-10, 15 ,,Noether Current" 164, 178 Particle, Material Point 8 Poincare's Lemma 16 Poincare's Lemma, Converse of — 16-18 Rotations (SO(3), Group of Proper —) 1-5, 53 Rotations, Infinitesimal — 4-6 Translations, Group of— 1, 4 Translations, Infinitesimal — 6 Translations, Time — 6 Variational Problem, Inverse — 22 Velocity 4, 8-9 Work 12
221