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!
0) we have
)(E) = y({E, 1)) + . . . + cp((£', rc)). Now 7>p<£ is a single-valued function on the £-plane. In particular, if ( PW(x, y% P) G(x,y% E) = \ 0 then G(x, y, E) is the Green's matrix of LA . Let g(x, P) = V(x, x, P). Then g(x, P) has the following important properties. a) The group of periods of g(x, P) is the same as that of the potential V. b) It is algebraic on F. c) It gives the "spectral decomposition" for the matrix A(x, E) that is, g2 =ft g(x, (E, k))-g(x, (E, /)) = 0 for k * I (where k, I label the sheets), Trpg(x, P) = 1, TrPW-g(x, P) = A(£, JC). d) =r-^- = [g, Q^ ] for any B E A. e) The variational derivative of the functional p(E) {v} defined in (3.2.8) (in the periodic case) has the form -=-(at-aj)g{
(see (2.1.17)).
f) As P -• {k}, g(x, P) has an expansion of the form where
1 Here the dash denotes differentiation with respect to x = xA. From c) and e) it follows immediately that the systems (3.2.9) are Hamiltonian. Let us find the zeros and poles of the matrix elements gj(x, P), which are algebraic on F. From c) it follows that the poles of g(x, P) are precisely the branch-points of F, that is, the points at which the different branches of the algebraic function W(E) merge. We denote the set of branch-points by the symbol 3)w. There are Nn{n - 1) of them. From the expansion f) it follows that at the infinitely distant part of F
114
B. A. Dubrovin, V. B. Matveev and S. P. Novikov
all the gj have zeros and their disposition is as follows: for / ¥= j, gj(x, P) has double zeros at all the points P = {k} (k =£ i, j) and simple zeros at the points P = {/}, P = {/}; gf(x, P) has double zeros at all the points P = {k}, (k ¥* 0, and gf(x, P) = 1 at P = {/}. There is a convenient notation to describe the distribution of zeros and poles of a function defined on F. Let
- - - > %n) of L can be extended to a canonical basis (£i> * ? ! , . . . , $ „ , %)The proof is by induction. For n = 1 the assertion is obvious. Let n > 1. We consider V. = n Ker 7(£;). Now K , D l and dim V\ = « + 1. Therefore, n(T}15 ^ ) * 0 for any />1 T?! e V\ L, otherwise, l(L, rij would be an (« + l)-dimensional Lagrangian plane. We normalize r\x so that n(£,, TJ I ) = 1. It is not hard to see that / (| x, r\ x) and V = Ker 7 (£ x) n Ker 7 (TJ ,) give a decomposition o_f Kinto a skew-orthogonal sum: V- /(f t , 17,) © K, and the restriction of n to Fis non-degenerate* Now Kcontains (with respect to si\y) the Lagrangian plane L = l($2,..., $n) with the basis (f 2 , . . . , $ n ), and by the inductive hypothesis it can be extended to a canonical one (with respect to 12| p^) by vectors r}2,... , t\n e V. Clearly, (17,,. . . , Tjn) are the required vectors." . . . , / # ) 0* = 1, ...,&)> so that they are essentially independent, for which it is enough to insist that the determinant (X, Y) = X(OJ(Y)) - Y(a>(X)) - u>([X, Y]), applied to co = dip, shows that X(d ) is symmetric and well-defined, that is, does not depend on the extensions A' and Y of the vectors £ and 17. Next, we suppose that Hess H and Hess h are positive-definite, hence equivalent. Then the level surfaces of H and h near x are closed curves. We construct the action-angle variables &', (£ for H in a neighbourhood of x. By § §6.4 and 6.5, the function «^° is uniquely determined up to a constant by the level lines of//, which we normalize by setting &>(x) — 0, and the cyclic coordinate (Si is uniquely determined up to a constant. In the same way, we construct action-angle coordinates ^ ' , (&' for the Hamiltonian h. The required canonical transformation (if it exists) must take the level lines of H to level lines of h, consequently, the normalized action & to the normalized action &'\ the cyclic coordinates & and 6Lf must go into each other, up to a constant. Therefore, there is a canonical transformation carrying H to h if and only if H and h, as functions of their normalized actions, coincide. What we have said shows that knowledge of H together with all its derivatives at a singular point does not allow us to recover the Hamiltonian itself up to a canonical transformation; for example, we can add to H any function of & that is flat at x. We should mention that under the above hypotheses (m = n - 1, positive-definiteness of the Hessians), the algebras of local symmetries are nonetheless unique, that is, are carried into each other by a canonical transformation. For these algebras, which consist only of functions of the actions, are identified under identification of the coordinates ^ ° , (§, and «^% @L'. Ut such that F * ^ ^ ) = Slo\Uo and F| fix(A) = id. We try to construct a family of diffeomorphisms Ft such that the form F*(cor) does not depend on t, that is j ^ F*(cot) = 0, 0 < t < 1, and Fo is the identity map. Except for the condition F| fix(A) = id, this would give us the result needed, because £l0 = co0 = F5(co0) = Ff(coi) = Ff(£lx) and f/x such that F*(£11\U) = £lo\u, F\L = id. Then the composition c = b~lo F is such that c*(£lL) = S2|p- and c(L) = L, so that we reach the following result. THEOREM 7.2 (Weinstein [12]). Let L C K be some Lagrangian manifold. Then there are neighbourhoods K D V D L and T*(L) DWDL and a diffeomorphism c: V -> W such that c*(£2L) = £l\v and c(L) = L (that is, the neighbourhood V of L in K has the same canonical structure as the neighbourhood W of L = so(L) in T*(L)). COROLLARY 7.2. Every canonical diffeomorphism f:K-*K sufficiently close to the identity can be included in a family of canonical diffeomorphisms /,: K^K,fo= id, / , = /. We do not make meaning of the words "sufficiently close" precise, and leave it to the reader to do this on the basis of the proof presented below. Let nf K X K -> K be the projection onto the z-th factor (/ = 1, 2) and co = 7rf(£2) - 7rf(£2). Then, clearly, dco = 0 and co is non-degenerate, that is, co is a canonical structure on I X K. Let /: K -• K be a canonical diffeomorphism. Then for the "graph" /: K -• K X K of /: fix) = (x, fix)), we have /*(co) = /•irf(S2) - / M ( 1 2 ) =_ = dr 1 o/)*(S2)-(7r 2 o/)*(n) = 1*(«) - / • ( « ) = 0, that is, f(K) is a Lagrangian manifold in K X I . If / = id = 1, then \(K) = A is the 0i*(co) = 0, hence, co = dip, where y = - V a regular map onto, where ^ - 1 (u) = Lv for all v G V is a Lagrangian manifold and a solution of the Hamilton-Jacobi problem, that is, lies on any one of the level surfaces of H. We denote by Qf: V -» R (/ = 1, . . . , « ) the coordinate functions on V. If V} v = (c l 5 .. . , cn)9 then Lv = A {&=,}, where Qt = ^*(Q/) are the parameters of the Lagrangian manifolds Lv, regarded as functions on K. The following lemma shows that the Q{ form a family of n commuting functions on U whose differentials are linearly independent owing to the regularity of \p. LEMMA 8.1. Let Lv be an n-parameter family of Lagrangian manifolds lying on level surfaces of both fx and f2, where flt f2 G $F(K). Then {fu h) = o. In other words, ff = ip*(hf), where ht G &(V) (/ = 1, 2). Clearly, { (L"), W ) = X%) are defined. We compute Tv The functions 7rf = /(n f ), where nz- = p(Xf), form a coordinate system on I * . The structure formula [Xt, Xj] = Xc^Xk n = of the ring of symbols S is canonical if and only if the diagram S 4 , D(S) S 4> D (S) commutes, that is, if and only if H(qs) = qjo^oq)- 1 for all s G S. The proof is obvious. Of course, the concept of a canonical transformation introduced above corresponds in the classical case K = C°°(M, C) or K = C°°(M, C) to the "usual" canonical transformation in T*(M). 6. Infinitesimal canonical transformations. The analogue of an infinitesimal canonical transformation, that is, a vector field that respects the Hamiltonian structure, must, in general, be taken to be a derivation of the ring S that respects the Hamiltonian structure. We give a precise definition. DEFINITION 2. A derivation \|j £ D(S) of the ring of symbols S is called canonical if it satisfies the relation Lhpoh+hoAty = 0, where, by definition, Dty(d) = id, ip], and AiKs ds') = (\\)s)dsr + sd(tys'), s, s' e S. S2)i — i ^ i ' (Psa}2> where {,}* are the Poisson brackets of Sf. If sx 6 Ker cp then 0 = {q>slf cps3}2 = cpfo, s2}lt that is, {sx, s2} 6 Ker cp. Thus, Hs^Ker cp) cz Ker cp, sx 6 Si. Ideals that are invariant under the action of any Hamiltonian derivations are called stable. In other words, an ideal of a canonical ring S is stable if it is also an ideal for the Lie structure of S determined by the Poisson brackets. PROPOSITION 5. Let q be a stable ideal of a canonical ring S. Then by the equality {cpsj, cps2}' = cpfo, s2}, where cp: S -+Slq is the natural projection, a canonical structure on S/q is well defined; and cp is then a canonical homomorphism. The proof is obvious. In view of this proposition it is natural to regard the canonical factor ring S/q as a localization of the canonical ring S. The geometrical picture corresponding to a stable ideal q C S is that the relevant "submanifold" Spec S/q of the "manifold" Spec S is tangent to all the Hamiltonian vector fields. It would be natural to define the Hamiltonian orbit of a subset A C Spec S as the smallest set AH D A, AH C Spec 5, that touches all Hamiltonian vector fields. Intuitively, one can imagine AH to be the set obtained from A by means of all Hamiltonian transformations, that is, shift operations along the trajectories of Hamiltonian vector fields. Of course, these intuitive geometrical concepts cannot be given, in general, a precise meaning. However, the algebraic version of the set AH exists and is given by the following definition. DEFINITION 3. The algebraic orbit of a point p G Spec S is the "submanifold" Spec SlO(p) -> Spec S, where O(p) is the maximal stable 1
We now consider the question in how many ways a basis (£ 1 ? . . . , %n) of a Lagrangian plane Z, can be complemented to a canonical basis (£i> ••• > £w, *h, . . . , ! ? „ ) of F. Every such complementation gives rise to a decomposition of F into the direct sum of Lagrangian planes L and L' = /(7h, . . . , rfo), F = Z, © L \ We claim that the converse is also true. LEMMA 1.3. For given L, £ 1? . . . , £„ and l! satisfying the above conditions, the complementary vectors r\u . .. , rin can be uniquely recovered. The required vector 17. must belong to Vi = n Ker7(£,) and V.- n L'\ now Vf D L and dim V* = « + 1. Since L and Z/ form a direct sum, dim (Fy. n L') = 1. If 0 * rjj e F;. n L', then n(£;., rjj) ¥= 0, otherwise
The structure of Hamiltonian mechanics
181
ft| y, - 0, contradicting the fact that dim Vj = n + 1. Setting r\. = rjj./ft(£;., TJJ), we obtain the required basis. Its uniqueness follows from the fact that ft(£;-, TJ) = 1 for only one vector TJ on the line V.- D Z,'."
LEMMA 1.4. The space of Lagrangian planes complementary to a given Lagrangian plane L is ho mo topically trivial (= contractible). (For facts about homotopy theory, see [5]). Let L' be a Lagrangian plane complementing L, {£t,... , f n ) a basis of L, (TJX , . . . , Tjn) a basis of L' such that ft(£ j., TJ;) = 6^- (Lemma 1.3), and (al,..., an, / ^ , . . . , 0 n ) coordinates with respect to the basis ( f c j , . . . , £nt i?i»• • •» Tjn). Then every complementary plane L" is given by a system of equations at; =
n
2 flj. /3fe (i = 1 , . . . , « ) , where the matrix || tfj || is uniquely determined by L". A straightforward
verification shows that L" is Lagrangian if and only if || a f || is symmetric. Thus, the space of complementing Lagrangian planes is homeomorphic to the space of symmetric matrices of order n, which is contractible."
§2. Fields and 1-forms on
2.1. Having studied the geometry of the localization of dp at some point v G <$>, we now consider the situation globally. A global analogue to 7 is the map V: 3)($) -* AJ(4>) defined by the formula T(X) = dp( , X) = -X J dp. This T is clearly J2r($)-linear, and its localization is (-7), since \YiX)iY)\ iv) = dp(Y, X)iv) = dpviYvi Xv) = -yp(Xv)(Yv), v G $, that is, T(X)V = -yviXv). Since 7 is an isomorphism at every point v G
F (a — + P—) —adp — fidq.
2.2. Canonical manifolds and canonical coordinates. We have seen that there is a canonical structure on the manifold $ = J*(M), that is, a non-degenerate closed 2-form dp. Similarly, there is such a structure on any domain U C
182
A . M. Vinogradov and B. A. Kupershmidt
By canonical coordinates on K we mean (local) coordinates (p,q) in which £2 can be written in the form £1 = dp A dq. As we shall prove in §7, such coordinates exist in a neighbourhood of any point a G K. An example of canonical coordinates are special coordinates on $>. Just as in the case of a phase space, we can introduce the concept of a Lagrangian manifold L on an arbitrary canonical manifold K. Namely, L is called Lagrangian if £L\L = 0 and L has maximal dimension, namely, n. We can also define a map F by the formula T(X) = £2( , X). From the nondegeneracy of £2 it follows, as before, that F is an isomorphism. We note that F is also given in canonical coordinates by (2.1). We shall make use of the isomorphism F and call a vector field X Hamiltonian if the 1-form T(X) is the differential of some function H E Jf(K). In classical mechanics, every function on the phase space is called Hamiltonian. We extend this term to apply also in the case of arbitrary canonical manifolds. We write the fact that H is a Hamiltonian of a field X in this way: T(X) = dH, or H ««* X, or X = XH. If a field X is Hamiltonian and in the canonical coordinate system (p, q) we have X = a ^— + j8 ^—, then, as follows from (2.1), j3z- = —H q., oij = Hp,. The differential equations corresponding to the Hamiltonian field X are (2.2)
q = Hpi
p = _#,.
These then are the famous Hamilton equations corresponding to a Hamiltonian function H. 2.3. By a canonical transformation in ^ we mean a diffeomorphism >4: K -•AT preserving 12, that is, ,4*(£2) = 12. This also means that 12(Z, Y) = G4*(12))(X 7) = 0{A(X)9 A(Y)), VX, 7 G 2)(K). If L C K, then ^*(co)| I = -4*{co|A(L)} for any diffeomorphism A: K -+ K and any form co € AP(AT). If ^4*(co) = co, then co^ = ^*{o)|A(L )}. Hence a canonical transformation carries Lagrangian manifolds into Lagrangian manifolds. If {At} is a one-parameter group of canonical transformations, {At} *-> X, then ^(12) = lim —*— 0
f
= 0. In this connection, a vector
field X for which X($l) - 0 is called an infinitesimal canonical transformation. The one-parameter group {^4*} corresponding to an infinitesimal canonical transformation consists of canonical transformations (Lemma 0.1). 2.4. If the form co G A1 (K) is closed, but not exact, then it is only locally the differential of some function / G jf(K). Therefore, it is natural to call vector fields corresponding to closed forms under the isomorphism F locally Hamiltonian. For such fields the Hamiltonian exists only locally. We remark that if K is simply-connected (in the case <£ = T*(M) this is
The structure of Hamiltonian mechanics
183
equivalent to M being simply-connected), then locally Hamiltonian fields are Hamiltonian. The connection between locally Hamiltonian fields and infinitesimal canonical transformations is described by the so-called fundamental theorem of mechanics. FUNDAMENTAL THEOREM OF MECHANICS. Locally Hamiltonian vector fields are infinitesimal canonical transformations, and vice versa. We have T(ft) = d(X J £2) + X J da = -dT(X). Therefore, if X is locally Hamiltonian, that is, dF(X) = 0, then X(£l) = 0; and if X is an infinitesimal canonical transformation, that is, X(£l) = 0, then dT{X) = 0." Since the form Sla on Ta(K) is non-degenerate, O^ _A_ ^ ^ A £la =£ 0 n times
(this can be checked by writing £la in some canonical basis). Therefore, £2 = S2 A • • • A 12 is a volume form on K. If X is an infinitesimal n times
canonical transformation, {At} *=> X, then X(£2) = 0, so that Af(£l) = 12. In particular, by virtue of the fundamental theorem of mechanics we have the following result. LIOUVILLE'S THEOREM. Shift operators along trajectories of a Hamiltonian vector field preserve the ("phase") volume, that is, the form 12. 2.5. The form p is natural. Let A: Mx -*•M2 be a diffeomorphism. We construct a map A: T*{MX) -* T*(M7) by the rule A\T*(M ^ = (A *) - 1 . Now A is a foliated smooth map, with a smooth inverse A'1, therefore, A is a foliated diffeomorphism: 7r2° A = A © rr1. Let u>2 G A^ATj), w, =.4*(co2), and 5 W# be sections corresponding to the forms ojt, i = 1, 2. Then V ^ ^ Afj, ^ ( 5 W (x)) = - G4*)-1* (w, ijc) = (A*rl C4*(«,| A(X) )) - co 2 | A(X) « 5 Wa U(x)), that isM o 5 W j = 5Wa» ^ . COROLLARY 2.1. Le/ p,- Z?e r/ze universal form on <J>t (/ = 1, 2). 77K?rt i"*(pa) = p,, rtar w, ^ e universal form p w natural. Let « , e A1 (M,), co2 = (^i*)"1 ( W l ); then S^ (A*(p2 )) = (4» 5 W j )* (p2) = (5Wjo ^)*(p 2 ) = = A*(SUJ (pa)) = i4*(wa) = w lf thatis, 5*^ (i4*(p3)) = w,
V ^ t e A1 (/If,). Since the universal form
p t is unique, >i"*(p2) = p, .• COROLLARY 2.2. If A: M-+Mis a diffeomorphism, then A*(p) = p. We now turn to the infinitesimal version of the preceding arguments. Let Z e 2b (^0» {At} *** X {^4^} <=> X. Since ^J'(p) = p, we have ^(p) = 0, and from it° At=At<> nix follows that n(X) = I : f o TT* = = {djdt) A*\^Qo TT* = (d/dt) (A* o rr*) k = 0 = (d/dt)(n* o - 4 P | ^ O = w*W/ir)i4fI^Q = TT* <> JT. Since r ( A = d(p(X)) - X(p) = d(p(X)), X is a Hamiltonian field with the Hamiltonian # = p(Z). Let us compute p(A) at v G
p(X)(v) = v(X n ( v ) ).
We mention without proof a fact that we do not use later:Jtet B: $>-*<& (Ye 25(Q>)) be a diffeomorphism such that B*(p) = p (y(p) = 0; then B(Y) has the form Apt) for some v4CY).
2.6. Transformation properties of canonical manifolds. Let (Kt, £2Z-), i = 1, %
be canonical manifolds, and A: ATt -* K2 a canonical diffeomorphism, that is, ,4*(£22) = « ! . LEMMA 2.1. r 2 °,4 = U " 1 ) * Ti.
184
A. M. Vinogradov and B. A. Kupershmidt
VX, Y e 3) (iQ, rx(X)(Y) = Sl^Y, X) = A*(n2(A{Y), A(X))) = = A*[T2(A(X)), (A(Y))] = [A*(T2(A(X)))] (10, that is, rx(X) = A*[T2(A(X))].m COROLLARY 2.3. Let h G ^(K2) and g = A*(h). Then A(Xg) = Xh. Applying Lemma 2.1 to the field X - Xg, we find that T2(A(Xg)) = W"1)* [T^Xg)] = (A-1)* (dg) = dh, hence, A(Xg) = T-2l(dh) = Xh.m COROLLARY 2.4. Let A: K -+ K be a canonical transformation. Then Let Z G 3)(K). We denote by Z$: 3){K) ->• S)(Z) and Z A : A1 (K) -+ A1 (AT) the corresponding Lie derivatives. LEMMA 2.2. Let Z be an infinitesimal canonical transformation. Then
roz% + zAo r = o.
The transformations {Gt}, {Gt} <=» Z are canonical, and by Corollary 2.4, T(Gt(X)) = (G; 1 )* (F(I)), VZ E S (#). Hence
= (G*t-E)/t(T(X)).
As f -> 0, we obtain r(Z(AT)) = -Z(T(X)).m §3. Poisson brackets
3.1. Let Xx and X2 be two Hamiltonian fields on K with Hamiltonians H1 and ^ 2 ? respectively. Applying Lemma 2.2 for Z = Xx to I 2 , we obtain 0 = T((Xl)(X2)) + Xl(T(X2)) = T([X2, Xt]) + Xx(dH2\ that is, T([Xl, X2 ]) = (/(AT! (H2)). Thus, the commutator of two Hamiltonian fields Xx and X2 is another Hamiltonian field, with Hamiltonian XX(H2) = {Hu H2}, which is called the Poisson bracket of Hx and H2. By definition, {#i, # 2 } = XxC^) is the derivative of H2 along the field Xx = XHr We note that?/?!, H2\ = dH2(Xt) = TiX^iX^ = Sl(Xlt X2). Hence, the* Poisson bracket is bilinear over R and skew-symmetric. Furthermore, the Jacobi identity is satisfied: V#i, H2, H3 G ^(K), (3.1) {Hu {H2, H3}) + {Ht, {/f3, ^i}} + {#3, {#1, # 2 }} = 0, because the left-hand side of (3.1) is {Hu X2(H3)} — {H2, X1(H3)} — -[Xlt X2] (H3) = XX(X2{H2)) -X2(Xl(H3)) ~ [Xlt X2](H3) = 0. Thus, the Poisson bracket makes JF(K) into a Lie algebra, which is denoted by 3)&>. If ZBffi is the Lie algebra of Hamiltonian fields on K, we come to a homomorphism of Lie algebras X = T-1 o d: 3)$> ->• 3$&eThe following formulae are obtained straight from the definition: (3.2)
{#!, H2H3} = {Hu H2}H3 + {Hu H3) H2.
If {H, Ht} = 0 (i = 1, . . . , k\ ^ e JT(Rk), then (3.3)
{#, q , ^ , . . ., Hh)) = 0.
The structure of Hamiltonian mechanics
185
If (p, q) are canonical coordinates, then V
•/
\
t.
2/
^ \
dPi
dqt
dqt
dPi)-
dp IT
~q
df •
3.2. Let A: Kx -> K2 be a canonical diffeomorphism, and let {,}* denote the Poisson bracket with respect to the form Q,f (/ = 1, 2). LEMMA 3.1. Iff, ge^(K2), then A*({/, g}2) = {A*(/), i4* (*)}i. Let f=A*(f), g = A*(g). Then 4(*)0 = Xf, A(X-) = ^ (Corollary 2.3).
Therefore A* ( f^gj) = A*(n2(Xf, X )) = ^ W " 1 ^ / ) , A~l{Xg)) =
= «1(Z/,^-)={/,g}1. • LEMMA 3.2. Let {H, /} = a = const, {At} <^> XH. Then A*(f) = f + at. {#, /} = X H (/) = a=const. For details, see §0." 3.3. Vertical fields on <£. The following useful facts follow from (2.1). LEMMA 3.3. Let co € A 1 ^ ) , X = r" 1 (7r*(co)). T^w X is vertical, that is, is tangent to the fibres of the projection ir: $ -> M, and is constant along these fibres. In particular, if f G F(M), then the field Xn*^ is vertical. COROLLARY 3.1. Iff g G F(M\ then {n*(/), n*(g)} = 0. COROLLARY 3.2. Let co, v G A1 (M), {At} <=• r* 1 (ir*( co)). §4. Symmetries
4.1. Let XH be a Hamiltonian field on K. It is natural to call a canonical transformation A: A" -> A" a symmetry of the Hamiltonian system with Hamiltonian// if it preserves the field XH: A(XH) = XH. By Corollary 2.3, A(XH) *=> C4"1 )*(//). Since the Hamiltonian of a Hamiltonian system is uniquely determined up to a constant, H = ( / l " 1 ) * ^ + const, or A*(H) = H + const. Conversely, if A*(H) = H + const, then ^ ( ^ ) = X#. It is natural to define a symmetry of a Hamiltonian / / as a canonical transformation ^4 such that ^4*(/T) = //. Clearly a symmetry of the Hamiltonian H is also a symmetry of the system Xff, but the converse is false. EXAMPLE 4.1. A particle in a gravitational field. K = T*(R), n = dp,H =-£- + q, where 4 is a coordinate on R. If,40/, p) = (+fl ) p),flgR, then,4*(<jf) = q + a, A*(p) = p, A*(a) = ft,,4*(/0 = H + a. For a ^ 0, A is a symmetry of the system XJJ, but not a symmetry of the Hamiltonian H. Thus, there may be fewer symmetries of the Hamiltonian H than of the system XJJ. However, for "the majority" of Hamiltonians the totality of their symmetries is the same as that of symmetries of the corresponding Hamiltonian system. For if A*(H) -H + a,0^=aGR, then the level surfaces {x I H(x) =* c} and {x I H(x) = c + a) must be "identical", since they are carried into each other by A, that is, the family of level surfaces of the function H is periodic with period a. Consequently, if this family is not periodic for any a e R, then the symmetries of H and XH are the same. EXAMPLE 4.2. The harmonic oscillator. Let K » r*(R), H=(p2 + q2)/2. That the level surfaces of H are non-periodic follows from the fact that the surfaces {H - c } are empty sets if c < 0 and nonempty if c > 0.
186
A M. Vinogradov andB. A Kupershmidt
In connection with this example we remark that if the domain of definition of H is not R, then the family of level surfaces of this function is necessarily non-periodic. In the physical literature it is usual to consider symmetries of H, although from the point of view of the theory of systems of canonical equations it is more natural to consider symmetries of XH. Therefore, henceforth we call symmetries of XH non-standard symmetries of H9 and the usual symmetries of H standard. 4.2. We now consider the infinitesimal version of the preceding arguments. Let {At} be a one-parameter group of symmetries of XH and {At} <=» Y. Then 1) y(«) = 0; 2) Y(XH) = [Y, XH] = 0. It is natural to call a field Y satisfying 1) and 2) an infinitesimal symmetry of X H. In other words, an infinitesimal symmetry is an infinitesimal canonical transformation that commutes with XH. There is a natural connection between infinitesimal and finite symmetries. LEMMA 4.1. Let Y be an infinitesimal symmetry of XH and {At} *==• Y. Then At is a non-standard symmetry of H. Since 7(ft) = 0 and Y(XH) = 0, we have A*(£l) = ft and At(Xjj) = XH. By the fundamental theorem of mechanics, the field Y in question is locally Hamiltonian. Let h be its local Hamiltonian; then 0 = T([XH, Y]) = d{H, h}, hence {h, H) = a = const. Let us explain the meaning of this constant. Since At is a non-standard symmetry, Af(H) - H = c(t) is also constant. Since {At} is a group, c(t + s) + H = A*+S(H) = A*(A*(H)) = A*(c(s) +H) = c(t) + c(s) + H, that is, c(t + s) = c{t) + c(s). Since c(t) is smooth, c(t) = cot. Hence the symmetries occurring in {At} for t =£ 0 are simultaneously either all standard (if c 0 = 0) or all non-standard (if c0 ¥= 0). Further, {h,H} = Y(H) = = (d/dt) G4*(/7))|,=o = c0. Thus, a = c 0 . Therefore, if {h, H} = 0 (for any local Hamiltonian h of 7), then Y should be called a standard infinitesimal symmetry of ff, and non-standard if {h, H} = const ¥= 0. REMARK 4.1. If there is a point x G K where dHx = 0, then H has no non-standard symmetries, since {h, H} (x) = dHx(Yh\x) = 0 VA 6 Jf(K)We consider the infinitesimal analogue to the example with a particle in a gravitational field. Let Y = ^—, then A t(q, p) = (q + t, p), h=p, {h, H} = 1, Af(H) = H + t, that is, Y is a non-standard infinitesimal symmetry. We remark that if there is a non-standard infinitesimal symmetry Y and Y is full, then all the level surfaces of H are diffeomorphic, so that At({H = a}) = {H = a + ct), c ^= 0. But if F is a standard infinitesimal symmetry, then Y is tangent to the level surfaces of H, so that Y(H) = 0. In what follows we shall often omit the word "infinitesimal" in front of "symmetry", and we call a function / G JF(K) with {/, H) = const a symmetry (standard or not) of the Hamiltonian H.
The structure of Hamiltonian mechanics
187
4.3. Having discussed symmetries of Hamiltonian systems, we now consider the general case of a canonical manifold. Here it is appropriate to introduce one particular type of symmetry that arises on the phase space T*(M). Let H G jF(cD), X € &(M), X G 3)(O). Since p(X) is the Hamiltonian of X, we have the following lemma. LEMMA 4.2. If X(H) = 0, then {p(X), H} = 0. It is precisely symmetries of this special type that are, as a rule, utilized by physicists. Their popularity is explained by the fact that they are "visible", since they are directly related to the configuration space M (see §10.7). The other, "hidden", symmetries, are not immediately obvious, and sometimes it is not at all easy to guess whether they exist (see §10.6). 4.4. Every domain U of a canonical manifold K can itself be regarded as a canonical manifold on which the canonical structure is given by the form SI\y. Furthermore, with every Hamiltonian H e ZP{K) we can associate its (/-part HJJ = H\ y, which can be regarded as a Hamiltonian on U. Here the restriction of XH to U is, clearly, a Hamiltonian field on U with Hamiltonian Hv. Thus, we can speak of the {/-part of H or XH and, consequently, of symmetries of all the above types for these parts. We call such symmetries partial. It is clear that the restriction to U of every symmetry of one type or another is a partial symmetry. The converse is false, because the whole may be less symmetrical than a part. In particular, the Hamiltonian may fail to possess non-standard (infinitesimal) symmetries, while some {/-part of it does have them. EXAMPLE 4.3. The harmonic oscillator. As we have seen above, the harmonic oscillator has no nonstandard symmetries. With the same notation, we take for U the phase plane without the origin of coordinates, that is, U~R2\ {0}. On U there is defined the many-valued function Arctan (p/q) whose differential is single-valued, because the separate branches of Arctan (p/q) differ from one another by a constant irk: d(Arctan (p/q)) = P
q q
q2 + q2
~
P
= K The form \is closed, but not exact on U, so that the field
— is locally Hamiltonian. Since X( Bp
P imal non-standard symmetry of the Hamiltonian—
) = i x is an infinites2
1 , 1^.
Apart from partial symmetries, it is sometimes useful to consider local symmetries of a Hamiltonian near some fixed point x e K, that is, germs of partial symmetries at x. The study of local symmetries is significantly easier than that of global ones. Besides, in the most interesting cases, they can be completely computed. In § 7.4 it will be shown that near non-singular points all Hamiltonians have identical structures. The Lie algebras of local symmetries close to the simplest singular points can also be described, but to do this is already rather complicated, and we omit it (we treat a particular case in §6.5). §5. The structure of Hamiltonian systems with a given degree of symmetry
We now study some structural features of a Hamiltonian system that arise as a consequence of its symmetry properties. More precisely, we consider an invariant that characterizes the symmetry of such a system, which we call the "degree of symmetry", and we clarify what structural properties a system with a given degree of symmetry has. We shall show, in particular, that the symmetries generate a double fibering of K. Unfortunately, a theory that could be used to test the presence of one symmetry or another of a given Hamiltonian and, in particular, to calculate its degree of symmetry, is still waiting to grow out of "alchemy" into
188
AM. Vinogradov and B. A. Kupershmidt
"chemistry". The classical many-body problem is a well-known illustration of our helpnessness with similar questions. Therefore, eschewing alchemy, we shall pass this problem by in silence. 5.1. We say that a system of functions fx, . . . , fk G ^F{K) is essentially independent if the set {x G K | rank (dfx \x, . . . , dfk \x) < k} has no interior points. By the degree of symmetry s(H) of a Hamiltonian H we mean the largest number k such that there is an essentially independent system of functions fx, . . . , fk G JF(K) for which {ft, fj} = {ft, H) = 0 (1 < i, / < k). LEMMA 5.1. / / {ft, fj} = 0, 1 < /, / < k, then the fields Xt = Xf, are \ therefore Xt is tangent to the surfaces {fj = Cj}, 1 < j < k, hence also to their intersection P c ." COROLLARY 5.1. 1 < s(H) < n. Let (fly . . . , fk) be the system of functions figuring in the definition of s(H), and x G ^T a point such that k, hence k < «, that is, .?(//) < «. On the other hand, if the system consisting of the function fx = H is essentially independent, that is, the set B = {x G K\dHx = 0} has no interior points (int B = 0 ) , then s(#) = 1. If int B ¥= 0 we can take for /j a function that coincides with H on ^T \ int B and for which int [{x G JTItf/j | x = 0} O int 5] = 0 . We omit the proof that such a function exists, both because it is banal and because Hamiltonians of this sort do not occur "in nature"." 5.2. We start by studying Hamiltonian systems of degree of symmetry > k. Let fi, . . . , fk be some collection of essentially independent symmetries of a Hamiltonian //, X = XH, Xt = Xp and {^4j} <==* Xf. Since {/«, fj} = 0, we have [Xt, Xf] = 0 and A\ * A{ = A\* A\. The first important structural feature of our system consists in the possibility of fibering K into a family of disjoint submanifolds (with singularities) of dimension 2n — k, "woven" out of the trajectories of X. To construct these manifolds, we consider the level surfaces of the functions f{. PZ(X) = {x\ft(x) = X} and their intersection k
Pc = H PjiCj), c = (c l 5 . . . , ck). By Lemma 5.1, the trajectories of X /i
lie on the manifolds Pc. In what follows it is convenient for us to assume that f\ - H (for otherwise we could add H to the functions fl9 . . . , fk and argue as in the proof of Corollary 5.1). Under this assumption, the sets Pc lie on the level surfaces of H. If x G Pc and dfx\x, ••• , dfk \x
The structure of Hamiltonian mechanics
189
are linearly independent, then near x the set Pc is a submanifold of dimension In - k, that is, Pc n U is a submanifold, where £/ is some neighbourhood of x in #. By the essential independence of the system of functions fx, . . . , fk, the points where the differentials dfx \x, . . . , dfk \x are linearly dependent form the "inessential" part of K. Therefore, loosely speaking, Pc is a submanifold near almost all points x G Pc. Henceforth we call the sets Pc manifolds, regardless of the fact that they may have singularities, and we bear in mind that all our arguments for which it in fact matters that the Pc are genuine manifolds (without singularities) are valid for those points x G K near which Pc is a manifold. Clearly, Pc C\ Pc> is non-empty only if c = c\ and U Pc = K. In this sense we say that K is fibred by the ^-parameter family {Pc}, c G R*, of manifolds of dimension 2n-k. Let 7: / -* K be a trajectory of XH, I C R. Since 7(7) C Pc for some c G R*, and since Pc is closed, the closure of 7CO also lies in Pc. This shows that "the more symmetrical the Hamiltonian system, the less intricate are its trajectories in K". Furthermore, we obtain the following necessary geometrical condition for a Hamiltonian system to have degree of symmetry > k: there must be room for the closures of all the trajectories of the Hamiltonian system on (2n — A:)-dimensional submanifolds of K. This circumstance enables us to give fix an upper estimate for the degree of symmetry of a Hamiltonian system if we know some of its trajectories. We must, however, remark that it is very difficult to realize this possibility in practice. 5.3. It is pertinent to ask how far the functions fx, . .. , fk that form a maximal collection of pairwise commuting symmetries of a given Hamiltonian H are uniquely determined. If we form new functions // =
—-1 vanishes on a
set S
(x) = 0> . Clearly, the converse also holds, so that we 0/1
*
have the following result. PROPOSITION 5.1. The symmetries fx, . .. , fk of the Hamiltonian H are uniquely determined up to a transformation ff =
190
A. M. Vinogradov and B. A. Kupershmidt
if and only if the collection of manifolds Pc is uniquely determined by H. An example of a non-unique collection is given by the Hamiltonian H = qx G .F(r*(R 2 )), where (qt) (i = 1, 2) are coordinates in R 2, and (qiy pt) is the corresponding special system of coordinates. The collection f\ = Qi> fi ~ Q2> is clearly not functionally related to the collection 1i.
Pi-
There are, however, Hamiltonians for which Pc is uniquely determined. A relevant example will be given in §6.5. We remark that the reason for uniqueness in that example lies in the fact that the trajectories of X H fill out everywhere densely "almost all" the manifolds Pc, which are, therefore, invariantly defined as the closures of the trajectories of XH. A similar circumstance apparently occurs for "almost all" Hamiltonians whose level surfaces are compact. A priori one can suppose that there are similar but more complicated mechanisms for the behaviour of the trajectories of X H ensuring the uniqueness of the family {Pc} in general. To sum up, we can say that the degree of symmetry characterizes the degree of intricacy of the trajectories of XH, while the character of intricacy of these trajectories is connected with the problem of the uniqueness of the family {Pc}. Arguments to be introduced in §6 below for the case k = n show that the ratio s(H)/n, 0 < s(H)/n < 1, can be interpreted as the degree of integrability of the corresponding Hamiltonian system, so that integrability and symmetry are practically one and the same. The interrelations between properties of symmetry and solubility (integrability) that arise in Hamiltonian mechanics are essentially the same as in the theory of algebraic equations, where they are described by Galois theory. Unfortunately, the lack of methods of finding the algebra Sym H is a serious obstacle to the practical use of a "Hamiltonian Galois theory", even if it were constructed. 5.4. For further investigation of the structure of a Hamiltonian system in a neighbourhood of an arbitrary point x G K we make use of the map o^: (tu
. . . , tk) = t -* At(x)
= A) o . . . oA1^ (x). Generally speaking, ax
is not defined for all t E Rk, because the fields Xx, . . . , Xk need not be full. On the other hand, it is not hard to see that there is in Rk a neighbourhood Ux of zero such that ax is defined on Ux. Decreasing Ux if necessary, we can choose a neighbourhood Vx, 0 G Vx C Ux, such that Alt.oA{,(x) = A[oA\{x) for any / G Vx. Such neighbourhoods Vx we call admissible. We ; denote by Nx the set o^(Kx) = {y G K\y = At(x\ t G Vx). Our next aim is to describe the structure of Nx. LEMMA 5.2. Suppose that rankC*^, . . . , Xk\x) = s. Then rank {Xx \y, . . . , Xk\y) = s for any point y G Nx. Let y = o^(O = At(x), t G Vx. Then Xt\y = (dAt\x)(Xt\x) in view of the fact that At maps Xf to itself. Furthermore, dAt\x is an isomorphism, since At is a diffeomorphism. Therefore, rank (Xt \y, . . . , Xk\y) = X ^ , . . . , Xk\x) = s. •
The structure of Hamiltonian mechanics
191
LEMMA 5.3. If rank (Xx \x, . . . , Xk\x) = s, then Nx is an s-dimensional embedded manifold. 1) Let 5 = k. By Lemma 5.2, rank dax\t = k, t G Vx. Hence Nx is an embedded manifold of dimension k. 2) Let s < k. We choose s linearly independent vectors from among Xl\x, . . . , Xk\x. Let these be, for example, X1\x, . . . , ^ | x • We denote by F^ the intersection of Vx with the subspace {T} = {(tu . . ., ts, 0, . . ., 0)} in R*, and let ax = ax\p. By Lemma 5.2, rank dS^ = s, and from the part of Lemma 5.3 already proved we conclude that Nx = Imc^ is an embedded manifold of dimension 5. Further, A^ = Nx, since Xs+1, . . . , Xk are tangent to Nx: at x this is so because Xs+1 \x, . . . , Xk\x can be expressed linearly in terms of Xx \x, . . . , Xs\x, and since (dAr\x)(Xt\x) = Xt\ ^, the vectors ^5+1 la (T)> • • • ' ^itla (T) c a n i n t n e s a m e w a v ^ e expressed linearly in terms of Xl | ^ ( T ) , . . . , ^ | t t x ( r ) . " REMARK 5.1. The map c^. and the manifold Vx can, clearly, be formed by means of the construction we have described, for an arbitrary system of commuting fields Xx, . . . , Xk on some manifold M. In particular, the fields Xt = T~1(dfi), where {ft, fj} = const, are such a collection. REMARK 5.2. Let rank (Xl \x, . . . , Xk \x) = k. Then A^it)) = ax(t + t') = ax(rt(t')), where tt t' e Vx, rt(t') = t + t'. Since — (rre) = -^—, where et is the /-th basis vector in R^, (tj(et) = 5/7), we have c^ (-^-) = Xt\N . Thus, Q:^ identifies Nx with Fx and the system of fields Xt\N with the system of coordinate fields, that is, ocx is a multidimensional analogue to the concept of a trajectory (see §0). Let y £ A^, .y = At(x), and z E Ny, z = ylf'(y). As we remarked above, the point At'+t(x) may be undefined (because the Xf are not full), whereas the point At>(At(x)) = At>(y) = z may be defined. Therefore, it is natural to consider the "transitive completion" of Nx, that is, the set k N Nxx = {y G K\ 3 tx, . . . , tx e R : y = A^o . . . <>At (x)}. It is clear that if Nx H j ^ is non-empty, then Nx = Ny. By Lemmas 5.2 and 5.3, the set Nx, which consists of the Ny, y G Nx, is an embedded submanifold. Thus, we see that K is fibred into non-disjoint fibres formed by the embedded submanifolds Nx. Also, owing to the essential independence of the functions fx, . . . , fk, the open everywhere dense set Uo in K where the differentials dfx, . . . , dfk are linearly independent is fibred into embedded submanifolds of dimension k. In the complement Vx = K \ Uo we can choose a set Ux, open in Vx, on which rank (dfx,. . . , dfk) = k - 1, and in the same way it is fibred into embedded submanifolds of dimension k — 1, and so on. Thus, the collection of symmetries of a Hamiltonian
192
A. M. Vinogradov andB. A. Kupershmidt
system determines a "hierarchic" division of the canonical manifold K into submanifolds Us( s = 0, . . . , k) at whose points the rank of the system dfx, . . . , dfk is equal to k - s and each of the submanifolds Us is "fibred" by submanifolds of the form Nx of dimension k - s. Thus, the system of symmetries determines two fiberings of K: into the manifolds Pc and into the manifolds A^. Lemma 5.1 shows that every Nx lies in Pc, c = F(x). Thus, the Nx, x G Pc, in their turn fibre the manifolds Pc, so that we get a "double" fibering. It is useful to note that the A^ are pre-Lagrangian, that is, £2|jy = 0- For S^CX/IJC, Xf\x) = {/,-, fj} (x) = 0, and the vectors Xt\x(i = 1, . . . , k) generate TX(NX). Therefore, £lx\T ^ ) = 0. 5.5. We turn now to the case when the fields Xt\^ are full. Here we can completely describe the structure of the manifolds A^.. THEOREM 5.1. / / the fields Xt\jff are full and rank (Xt \x, . . . , Xk\x) = s, then Nx is diffeomorphic to a manifold of the form Tm X Rs~m, where Tm = S1 X . . . X S1 is an m-dimensional m times
torus. Just as in the proof of Lemma 5.3, we can reduce the proof to the case s = k. Owing to the fullness of the fields Xt\^ the map ax is defined on the whole of R*, and by Lemma 5.2 it is a local diffeomorphism. We consider Ker ax = {t G R^lc^O) = x}. Now Ker o^ is a subgroup of R*. For if tx, t2 G Kera^, then o^Oi + t2) — <xa (^)(^) =
= ax(t2)=x and x = ax(0) = ax{tx + {-tx))^ao^{u){-tl)
= ax{-tx\
that
is, ("~ tx) G Ker o^. together with tx. It is not hard to see that Ker ax is discrete. This means that every bounded set in R^ contains only finitely many points of Ker ax or, equivalently, that there is a neighbourhood W of the origin in R* such that W D Ker ax = {0}. This last is true in view of the fact that ax is a local diffeomorphism. Let y G Nx and a ^ ) = 0 ^ 2 ) = y- Then ax(t1 - t2) = a ^ ( r i ) ( - r 2 ) = = 0Cy(—1 2) =A_(tj(y) = x, that is, tx —1 2 G Ker a^.. Conversely, if
t' G Ker 0^, then ax(tx + t') = a^^t^
= ax(tx) = y. Thus,
o^iti) = OLx(t2) if and only if tx~t2 G Ker axi that is, if and only if tx and t2 lie in one and the same coset of Ker ax. For this reason, the formula o^dt]) = 0^(0, where [t] is the coset of t G R* under Ker o^., gives a well-defined map 0^: R*/Ker o^. -> Nx. This is clearly a one-toone correspondence and coincides locally with ax because ax is discrete. Therefore, 0^ is a diffeomorphism. By the theorem on discrete subgroups of R* (see [25]), R*/Ker c^ « Tm X R^""m, where m is the dimension of the linear space spanned by Ker o^., which is equal to the number of integral generators of this subgroup."
The structure of Hamiltonian mechanics
193
§6. Completely integrable systems
We now examine in greater detail for s(H) = n the situation described in the preceding section. In view of Corollary 5.1, we should then call the corresponding Hamiltonian system completely symmetric. But for reasons that will become clear later, such systems are called completely integrable (see also §5.3). 6.1. An important property of completely integrable systems is the following: the manifold Nx is Lagrangian if x is a point in general position, that is, rank (Xx | x , . . . , Xk\x) = n. For in this case dim Nx = n and, as we saw above, tl\^ = 0. In the case in question, the manifolds Pc, <4as a rule", also are of dimension n, so that if Nx C Pci then Nx coincides with one of the non-singular components of Pc. (By a non-singular component we mean a connected component of Pc \ P c sing , where i^ingis the set of singular points of Pc.) We now assume in addition that the fields Xt {i = 1, . . . , n) are full (or full in the relevant domain of K). In this case all the fibres Nx = Nx of general form are of dimension n and, by Theorem 5.1, Nx = Tm X Rn~m. The integer m in this equation depends, of course, on Nx, that is, on x. Thus, we arrive at a function m = m(x) which takes integer values and is defined for all points x E K where dim Nx = n. It is clear that m(x) = m(At(x)) and that m(x) is constant along the fibres Nx. We investigate its structure in more detail. Suppose that an ^-dimensional submanifold W C K satisfies the conditions (T): x G W, W is Lagrangian, W D Nw = w
Vw G W, TW(W) n TW(NW) = 0,
and m is defined on W. If W is a Lagrangian manifold intersecting Nx transversally at x, then a sufficiently small neighbourhood of x in W' satisfies (T) (since Nx is a submanifold). This shows that there are manifolds satisfying (T). We consider the map a= aw: W X Rn ^ K, a(w, t) = aw(t) = At(w), w G W, t G R". LEMMA 6.1. a is a local diffeomorphism. We must show that da\{w§t): T(wt)(W X R") -• Ta(wt)(K) is an isomorphism for V(w, t) e W'x R».'But it follows from (T) that this is so for the points (w, 0). We now introduce r r :B'XR w ->WXR' 2 , rT(w, t) = (wy r + r).
Then a = V
a
° r.T, so that da\^wt) =dAt\w° da\(w 0)o dr_t\(wt) for
r = t, from which it is clear that da\^w^ is also an isomorphism.• It is obvious that the set a{W X R n) = U Nw is open, contains A^, and is invariant under all At. If y E ^(H7 X R"), then there is some w G W such that y G Nw. Hence m(y) = m(w;) and in studying the function m near x it is enough to study it only on W. LEMMA 6.2. There is a neighbourhood V of x in W such that m{x) < m(w) \fw G V.
We note that a - 1 (^) n (w X R") = Kera^. Since a is a local
194
A. M. Vinogradov andB. A. Kupershmidt
diffeomorphism, of1 (W) is an ^-dimensional submanifold of W X R" transversally intersecting the fibres of1 (Nw) = w X Rn. We denote by Wt the connected component of of1 (HO containing t G Ker ocx, and we set Wi = M^f. (/ = 1, . . . , m(x)), where ^ , . . . , rm(jc) is a basis of the discrete . mfr) subgroup Ker o^. Let PJ= {w e W\W{ n (w X R " ) ^ 0 } . T h e n F = n ^ is a non-empty open neighbourhood of x in K. The intersection Wt n (wX R") if it is not empty, consists of a single point, which we denote by lt(w), and we set lt = lv. Since lt(w) depends smoothly on w, the vectors lt(w) (/ = 1, . . . , m(x)) are independent in some neighbourhood V C V of x in W. But l((w) e Ker o^ , so that rank (Ker a j > m(x) if w e V, that is, m(w) > m(x). • COROLLARY 6.1. The function m(x) is locally constant, that is, the union of the open sets on which m(x) is constant is everywhere dense in its domain of definition. This follows from Lemma 6.2 and the fact that m{x) is integral-valued." LEMMA 6.3. Suppose that m(x) is constant near x and that the neighbourhood V constructed in the proof of the preceding lemma is connected. Then a~l(V) = U Wt, where Wt = Wt n aTl(V). fe Ker ax
Let E be some connected component of a~l(V), distinct from all the Wt, t E Ker o^, and WE = {w G V\(w X Rn) D E # 0 } . Clearly, WE is open in V and x $ WE. Let ^(u;) be the unique point of intersection of w X Rn with E, w e WE. Since e(w) G Ker aw, rank (Ker aw) = m(x), and the vectors /x(u;), . . . , lmM(w) are linearly independent for w G K, () we see that e(w) = 2 A ^ w ) / ^ ) , where the //(w) are rational numbers. /= l
The functions X^w) are clearly continuous, hence are constants X(y so that the domain WE, which is diffeomorphic to E, is connected. If m{x)
w0 G WE\ WEi then the point e(w0) = S limit point for E = U wGWE
X^^WQ)
lies in Ker o^ , being a
d » . For the same reason its connected com-
ponent E' in a'1 (E) contains E, and Er \ E $ e(w0). This contradicts the fact that E is a connected component." COROLLARY 6.2. li(w), . . . , lm^X)(w) form a basis of Ker c^ /or M; G V. 6.2. Suppose that PV satisfies (T), vf = «*(/;), rt - 7r*(fz), where n TT(U;, t) = t = (*!, . . . , r w ), (w, t) e W X R . Since a is a local diffeomorphism, the form ^2 = a*(£2) defines a canonical structure on W X Rn. It is clear that a (-^r-)= Xt. Hence the field-^— is Hamiltonian (with
The structure of Hamiltonian mechanics
195
respect to 12), oc*(ff) = vf is its Hamiltonian (Corollary 2.2), and ^ - (12) = 0. We calculate 12 in the coordinates (u, r). If 12=
2
^r/
%<*"/
1< /,/<«
+
(bkidTk f\dri
2 1 <&<
it follows from / - (12) = 0 that pL = ^ i =^L
+c
kidvk
A du7), then
= 0, that is,
atj, bkl, and c^ are functions of the variables v only. Next, Nw, and hence also w X R" = a"1 (#„,) are Lagrangian manifolds. Therefore, 0 = &\WXRn = 2 bkl(v(w))dTk A dr, from which it follows that k,l
bkl = 0. Because W X {0} C W X Rn is Lagrangian, we have 0 = £l\wx {0} = 2 ckl(v)dvk f\ dv^ that is, c^ = 0. Finally, since k,l
n
dVj = — !23/3T/ = 2 a^-^dVf, we have a/;(u) = 5/;-, so that we have proved the following result. LEMMA 6.4. 12 = 2 <2uf A i=i
dT
v
COROLLARY 6.3. (complement theorem). Suppose that in a domain U C K the differentials^ of the functions ^u • • •» ^ n ^ ^ independent and that {$V ^ ; } = 0. 772e« /or VuG [/ r/zer^ is 0 neighbourhood U' C U and functions & , . . . , ( ? „ € #(CO ^wc/z f/z^r r/z^ /
, . . . , nmM), n{ G Z. We define a free action of the group
196
A. M. Vinogradov andB. A. Kupershmidt
on T*(W) by the rule: k'(wx,. = (wly...,wn,
irl + ku
. . , wn, itx,. . . , itn) =
. . . , TTW(X) 4- km{x)i irm(xl+1>..
. , 7rrt),
where
£ = ( & ! , . . . , /:m(x)) £ Z w ( x ) . By construction /f(u;) is the point of the fibre T*(W) with coordinates 7iy = 5^., i = 1, . . . , m(x). Hence, if a(z) = y, y GNW, then (ao0)- 1 (y) = /TH^OO) = / T 1 ^ + K e r a j = = /3"1(z) + j3"1(Kerau;) is the orbit of 0~*OO under the given action of Z/nix) xhus, the map a ° 0 induces a diffeomorphism 7: T*(W)/ZmM -» a(W X /?"), 7(Zm(x)-w) = a(0(w)), w G r*(H/). If ^: T*(PV) -• T*(Pi/)/Zm(x) is the natural projection, then there are manyvalued functions
&t = H & t ,
ifii = 0
(J=1, ...,n).
Therefore, ^t = at = const, and &t = fat + 09, where j3j = Htpfax, . . ., a n ), so that the system has been completely integrated, which explains its name. The numbers j3z- = j3z(x) (/ = 1, . . . , m(x)) can be understood as the frequencies with which a point moving along the trajectory of XH passing through x runs round the torus Tm{x) (by virtue of the diffeomorphism Nx = Tm(x) X RB"WW constructed above). The frequency vector j3 = (jSj, . . . , ]3w(;c)) is not uniquely determined by H and x, since it depends on the choice of the representation of Nx in the form Tm X Rn~m. But in the important special case m = n the frequency vector (ft, . . . , j3w) is uniquely determined by the choice of cyclic coordinates on the torus Tn ^ Nw. The tori TV^ are, generally speaking, in this case completely determined by H itself. For if all the frequencies
The structure of Hamiltonian mechanics
197
0i, • . • , Pn on Nw « Tn are independent over Z, then the trajectories of XH are everywhere dense on Nw, so that Nw is in this case the closure of one of the trajectories. If the "frequency map" Nw -• (3(w) = (jSj (w), . . . , pn (w)) is an epimorphism, then the trajectories of XH are everywhere dense on almost all tori Nw. Therefore, these tori and, as is not hard to see from continuity arguments, all the others are in this case uniquely determined by H. This is certainly so if 0 on U. In this case the Hamiltonian system is called nondegenerate. The cyclic coordinates &t on Nw « Tn are uniquely determined by the choice of basis in Ker aw, that is, up to a integral unimodular transformation. Since such a transformation must depend continuously on w, but is integral, it does not, in fact, depend on w. So we come to the following result. PROPOSITION 6.1. Suppose that m = n and that the Hamiltonian system under discussion is non-degenerate on the domain U of Theorem 6.1. Then the tori Nw, w G U, are uniquely determined by the Hamiltonian, and the frequency function (3(w) = (Pi(w), . . . , $n(w)) is uniquely determined up to a integral unimodular transformation of the image. 6.5. Cyclic "actions". We now indicate a method of finding the coordinates ^u . . . , $>mix) introduced above directly. Let %•(«), a = (<*!, . . . , a n ), at = ff(w), be the cycle on Nw « Tm(x) X ^ defined by the equations &j = 0, / =£ /. If p' = 2 ^id&i, then
Slfr = dp' and
] P' = j ^ i d & i = ^i jj dg f =.#*,, Y. (a)
yt (a)
since &t\Nw is
yt (a)
constant. Let p" be another form such that dp" = Q\u. Then d(p' — p") = 0 and
\ {p — p") does not depend on a, that is, it is yt\a)
a constant ot. This follows from Stokes' formula owing to the fact that the cycles 7z(
(6.2)
j p" + at. And so ()
&t(a)= J
p
Complementing the system obtained in (6.2) by functions &i = &i (fi, • • • > /„) (i = w + 1> • • • > «) in such a way that the system <^n • • •» ^n is independent, we obtain half of the coordinates introduced above, with H = H(^u . . . , $*n). This is enough to study the frequency map and the frequencies of the system in question. To find the corresponding coordinates S* we consider any Lagrangian manifold V intersecting all the fibres Nw, w G U, transversally. Then, if Yi = X$>i
198
A. M. Vinogradov and B. A. Kupershmidt
and {A}} <=» Yi9 the manifolds Vt ... t = (A\ o . . . o 4 " ) (F') are the surfaces {&i=tt | i = l , . . . , n } . Therefore, the question of constructing the functions (Hi is reduced to the integration of the Hamiltonian fields
r l 5 . . . , Yn.
If m = n, then (6.2) defines all the functions #V In the case # = T*(M) we can take for p the universal form p = 2 pzdg/5 whose integrals along one path or another are customarily called the action. Therefore, the ($*, ffit) are in this case called "action-angle variables". 6.6. To conclude this section, we apply the theory we have set out to the problem of the local classification of Hamiltonians at a singular point. Let n = 1, that is, dim <J> = 2, H(x) = h(x) = 0, dH = dHx = 0. We ask, is there a canonical transformation in a sufficiently small neighbourhooa of x that takes H to h and leaves x fixed? For this, first of all, it is necessary that there is an ordinary (not canonical) transformation satisfying the stipulated conditions. By Morse's lemma (see [9]), the equivalence of the Hessians of H and h at x, provided that they are non-degenerate, is necessary and sufficient for this. We recall that the Hessian Hess(<^) of a function
§7. The theorems of Darboux and Weinstein
One of the possible formulations of the classical theorem of Darboux asserts that locally all canonical manifolds of dimension 2n have the same structure. We shall prove this below, together with vastly more general facts discovered recently by A. Weinstein. We recommend the reader also to look at the classical proof of Darboux' theorem (see, for example, [2]). 7.1. We consider the following situation. Let 120 and ^2j be two closed non-degenerate 2-forms on a manifold M, dim M = 2n. By an exact deformation A = (cot, yt) of £20 a n d ^ 1 w e mean two ^-smooth families of forms a>t e A2(M) yt G A1 (Af), 0 < t < 1, such that co0 = ft0, CJOI
=
£2I,
dcot = 0, dyt = — co t and cor is non-degenerate for all /. We
The structure of Hamiltonian mechanics
199
call x G M a fixed point of such a deformation if yt\x = 0, 0 < t < 1. The set fix (A) of fixed points of A is clearly closed. THEOREM 7.1. Suppose that two closed non-degenerate 2-forms £20 and £lt on M are joined by an exact deformation A = (cor,
we can set F = Fx. But 0 = £ F*(co,) = F*(7'(cor) + £- cor) = F*(r'(
200
A. M. Vinogradov and B. A. Kupershmidt
the domain Ux C Kx by taking F = A © F, U1 = A(UX).* 7.2. We now consider the problem of canonical coordinates (see §2.2). From the results of Section 7.1 we deduce the following proposition. PROPOSITION 7.1. Every point x of an arbitrary canonical manifold K has a neighbourhood furnished with a canonical system of coordinates. Let R2" be the space of sequences (ql9 . . . , qn, pX) . . . , pn). We n
introduce a canonical structure on R2" via the form £2t = D dpf A dqt. Let y e R2n. Then by Corollary 7.1 there is a diffeomorphism F: Uo -> Uu x G Uo C K, y G Ux C R2n, such that F*(£lx\Ui) = ft^. Therefore, the functions £f*i = F*(pi)i &t = F*(qf) form a canonical system of coordinates in Uo.* 7.3. Our next aim is to investigate the structure of a canonical manifold K near some Lagrangian manifold L. With the help of Theorem 7.2 we shall show that it is the same as that of T*(L) near I = so(L), OGA^L). In what follows we prepare the ground for the application of this theorem. Let L C K be a Lagrangian manifold, and X{x) the totality of all Lagrangian planes in TX(K), x E L, transversal to the Lagrangian plane TX(L) C TX(K). Then X{x) is a contractible manifold (Lemma 1.4). The set X = U X(x) can be naturally furnished with the structure of a smooth manifold. We have the obvious projection p: X -> L, p(X(x)) = x, which is, clearly, a smooth fibering with fibres %(x). The fact of elementary topology according to which there exist sections of a fibering with contractible fibres (see, for example, [7]), together with a procedure of smooth approximation proves the validity of the following assertion. LEMMA 7.1. There is a family of Lagrangian planes l(x) E X(x), x G L, depending smoothly on x. We remark that the map vx\ lx -* T*(L) defined by the formula (M**), £) = %( a > £)> a e /(JC)' ^ G Tx(L)> i s a n isomorphism (see §1.5). The totality A^ of all vectors OL G l(x), x G L, is, clearly, a submanifold of r(#), and the projection 7%: N -+ L (TT: T(K) -• ^T) turns it into a smooth vector bundle over L with fibres /(x). The map v\ N -> T*(L) defined by the formula v\i(X) = vx is an isomorphism of vector bundles. The vector bundle iV -> L is, clearly, a fibering isomorphic to the normal fibering of L in K. We identify L with the zero section of 7%: TV -> Z and 7^ (/(*)) with /(*), x G Z. This allows us to identify ^(AO with TX{K). The "tubular neighbourhood" theorem (see [11]) then guarantees that there is a diffeomorphism a: U' -* K, where U' is some neighbourhood of L in TV, such that a\L: L -+ L and da x : rx(A^) -> ^ ( ^ 0 , x G I , are the identity maps. Let U = v(U'). Then b = a o v'1: U -> iT is a diffeomorphism, and by construction fc*^^ = (^jr) x Vx G L, where 12L G A 2 (7'*(I)) is a canonical 2-form
The structure of Hamiltonian mechanics
201
on T*(L). We set £20 = &>\HU), Slt = (b*)'1^)Then £20 and nl are closed non-degenerate 2-forms on b(U) D L, £ 2 ^ = Sll\xi Vx E L, and £20|z = Q>l\L, since L is Lagrangian. The form cof = £20 + r ^ j -r2 0 ) is such that cot\x = £lo\x, V x G Z. Hence, there is a tubular neighbourhood Vo, b(U) D Vo D L, such that cjt\v is non-degenerate for all f, 0 < f < 1. Let # : T*(Z) -• r*(Z) be a "homothety" with coefficient ( 1 - 0 , that is, $(\x) = (1 - O X , , Xx e T*(L) C r * ( I ) , j8f = = b © 0/ o O" 1 )!^: F o -> Ko, and let 7 ' G 3J(F0) be the family of vector fields on Vo corresponding to the family ]3r If co = £2j —12 0, then doo = 0 and 7r(co) = d(Yf J co). Therefore, •£- |8*(co) = ft*(rr(co)) = d^fiX*
J co)} and
0f(co) - co = d { f 0 * ( y ' J co) J r } . Since ft = / o ( ^ ) l F o ° ^f"1, where o /: L -* K is the natural embedding, and /*(co) = co|^ = 0, we see that I j8*(7r J co) dt. Noting, finally, that o dcOf -— = co = dip, we conclude that A = (cop <pr =
202
A. M. Vinogradov and B. A. Kupershmidt
diagonal in K X K. By Theorem 7.2, there are a tubular neighbourhood V of A ^ K in K X A: and a diffeomorphism F: V -* W C r*(iQ such that F|^ = id and (F~1)*(co) = ^ ^ | ^ , where £lK is a canonical form on T_*(K). Now let /: K -* K be a. canonical diffeomorphism such that f(K) C V. Then F(/(iO) C W C r*(iO, and if / together with its first derivatives is sufficiently close to the identity map, then since also F(T(K) = so(K) C T*(K), 0 E A1 (JO, it intersects the fibres of the projection irK: T*(K) -* A" transversally, that is, it is of the form sx(K) = F(f(K)l A e A*(iO. But F<7(iO) is Lagrangian, therefore d\ = 0, and the canonical field X = F" 1 (7r^(X)) is such that ^4f-5x = 5 Q _ ^ X , where {-4*} «*» X (Corollary 3.2). We consider the manifolds F~*(stx(K)). If/ together with its first derivative^ is sufficiently close to 1, then F~1(stx(K)) has the form of the "graph" ft(K), where ft: K -> K is some smooth canonical map, 0 < t < 1, owing to the fact that, as before, the Lagrangian manifolds F~1(stx(K)) like A, intersect the fibres of TTJ transversally. Finally, since / i s "sufficiently close" to 1, the ft are also diffeomorphisms. By construction, / 0 = id, fx = / • 7.4. We now show that all Hamiltonians have the same structure in neighbourhoods of their non-singular points. PROPOSITION 7.2. Let H G &{K)and dHx ¥> 0. Then there are canonical coordinates (p, q) near x such that H = qx.
Let (cf>, (g) be some canonical coordinate system near x. By a linear (non-homogeneous) canonical transformation of the variables (<£P, &) we can, clearly, achieve that dHx = d(2ilx» H(x) = &-i(x). Assuming that is so, we consider the function Ht = H + t(&x — H) and we try to find a family of canonical transformations Ft such that F*(Ht) = const. F o = id. If we can do this, then in view of the fact that H = F$(H0) = F*(&i), Ft is the transformation we are looking for. But ±Ft(Ht) =F? IY1 (Ht) 4 - ^ - / 7 ] , where ® (K)^Yl <=> {Ft}, therefore, the condition ^ us to the condition Y\Ht) + &x - H = 0,
or
F*(Ht) = 0 leads
X1 (ht) = fix - H,
where F(X f ) = dHt, and ht is a local Hamiltonian of Yx. Let N be a hypersurface in K containing x and transversal to all the trajectories of all the fields X1, 0 < t < 1. Since by construction dHt\x^=d&x\x=dHx, we have X*\x = —
TTTT
* so that iV can be taken to be a sufficiently
small neighbourhood of x on a hypersurface iV' containing x and not tangent to the vector -^-1 . Now when we give on JV an arbitrary Ou l
X
The structure of Hamiltonian mechanics
203
function «pr as initial condition ht\N = \pt, we can find ht itself by solving the equations Xf(ht) = &t - H. In view of the fact that the trajectories of X* do not touch N, this equation can be solved by the method of characteristics. Thus, the required fields Yl - Y~l(dht) exist and hence so does the required transformation F r " COROLLARY 7.3. Let xt G K((i = 1, 2), where Kt are canonical manifolds of the same dimension. If Hi G &{Kt), # i ( * i ) = H2{x2), and dHi\x. ^ 0 (/ = 1, 2), then there are neighbourhoods Ut of xt and a canonical transformation f: U^ -* U2 such that f*(H2) = Hx. We introduce canonical coordinate systems (p, q) and (IF, g) near xt such that H1 = qlt H2 = fii, and we identify them." §8. Invariant Hamilton—Jacobi theory
8.1. In this section we discuss in an invariant way the theory of the Hamilton-Jacobi (H-J) partial differential equations and the connection between solutions of these and of the corresponding Hamiltonian system. Essentially, the whole H-J theory is covered by the following two almost obvious geometrical facts: 1) a solution of an H-J equation is a Lagrangian manifold lying on a level surface {H = c}; 2) if there is an H-parameter family of solutions of an H—J equation, that is, an ^-parameter family of Lagrangian manifolds lying on level surfaces of H, then the parameters, regarded as functions on <£, form a family of n pairwise commuting symmetries of H. The usual method of developing the H—J theory achieves a certain mystery by the introduction of coordinates. We draw the reader's attention to the fact that all the lengthy passages in the following exposition are concerned only with harmonizing the invariant point of view with the usual coordinate one. 8.2. We recall that an H-J equation is a first order partial differential equation of the special form H(q, ^-) = c, where q = (qx, . . . , qn), bu du ( bu bu \ , ~— = ( r — , . . . , ^— I and c is 0 Q
' OQ i
O £1 „ •
,
, i
a constant.
A special feature of this equation consists in the fact that H does not depend on u. We interpret this invariantly, and to do this we remark that, as a consequence of the specific nature of the H—J equation, together with each function u(q) that is a solution of it, all functions of the form u(q) + const are also solutions. In view of the fact that the totality of such functions is uniquely characterized by the differential of each of them, we can say that the H—J equation is an equation in the differential du of The H-J equations are usually written in the form Uf -
204
A. M. Vinogradov and B. A.
Kupershmidt
an unknown function u(q). We consider next the equation H(q, p) = c in T*(R n ), p e T*(Rn). This is the equation of a level surface F of H on T*(Rn). If u(q) is a solution of the H-J equation, then the graph of its differential L = f(q, p)\ p = sp (#)} lies on F. Clearly, the converse is also true: if L is the graph of the differential of some function u(q) and if L lies on F, then u(q) is a solution of the H—J equation. But, as we already know, the graphs of differentials of functions on R" are Lagrangian manifolds in T*(Rn) that project one-to-one onto Rw, and vice versa. Therefore, the solutions of the H—J equation are Lagrangian manifolds lying on F and projecting one-to-one onto R". On the other hand, even the simplest H—J equation for the harmonic oscillator I ( I T ) + y ^72 = c n a s n o global solution, that is, a solution in singlevalued functions defined on the whole of R". Fig. 1 illustrates the graph of a many-valued function u(q), whose differential has the circle p2 + q2 = 2c as its graph.
Fig. 1.
It is therefore sensible to give up asking for a one-to-one projection and to count as solutions of the H—J equation arbitrary Lagrangian manifolds lying on the surfaces F. These Lagrangian manifolds, which have no oneto-one projections, can be regarded as graphs of the differentials of manyvalued functions (see §1.3). Resuming what we were saying above, we find it natural to define an H—J equation as an arbitrary hypersurface F (possibly with singularities)
The structure of Hamiltonian mechanics
205
in T*(M) or, more generally, on an arbitrary canonical manifold K, and its solution as an arbitrary Lagrangian manifold L lying on F. Next, by the H—J problem we mean the task of finding Lagrangian manifolds lying on F. Then the statement of the H—J problem is indeed invariant, since the terms in which it is expressed do not appeal to coordinates and are invariant under canonical transformations. REMARK 8.1. Arguing in a similar way, one can develop a theory of many-valued solutions of arbitrary non-linear partial differential equations (see [13]). 8.3. We now turn to the H—J theory, which consists in the study of the connection between solutions of the H-J equations and those of the system of Hamilton's ordinary differential equations. This connection is given in an invariant way by the absorption principle. Let F be a hypersurface (that is, dim F = 2n - 1) on a canonical manifold K. As we saw earlier (see §1.4), there is then a field of direction lines on F. Namely, at each point x E F there is defined a line (= one-dimensional subspace) lx(F) C TX(T) C TX(K), the direction line of the hypersurface ^ ( F ) in TX(K) with respect to the skew-bilinear form £lx. The absorption principle. Every Lagrangian manifold L lying wholly on F is tangent to the direction field of F. What we are saying means that lx(T) C TX(L) for every point x € L. But this follows from the local (linear) absorption principle (see § 1.4). For the linear space TX(L) is a Lagrangian plane in TX(K) with respect to £lx. Further, since L C F, we have TX{L) C TX(T). Consequently, the Lagrangian plane TX(L) contains the direction line of ^ ( F ) , that is,/X(F).» Properly speaking, from now on we use the absorption principle in the following form: if a Lagrangian manifold L lies on a non-singular part of a level surface of //, then the corresponding Hamiltonian vector field XH is tangent to L. This formulation of the absorption principle follows straight from the first. For if F = {H=c} and x E L, then *n\x e / ^ H C TX(L). Since XH is tangent to L under the preceding conditions, L is woven out of the trajectories of XH, that is, if x G L, then the trajectory of XH passing through x lies wholly in L, therefore, L is invariant under the shift operators {At} *=> XH. This can be regarded as a third formulation of the absorption principle. 8.4. A method of constructing solutions of an H-J problem follows immediately from the third formulation of the absorption principle. Namely, let L = U At(L*\ where {At} *=* XH, and let I ' be an arbitrary (n - 1)dimensional manifold lying on F, transversal to XH and such that £l\L> = 0 (that is, £l\L> is pre-Lagrangian). For, if x £ At(Lf), then TX(L) = Tx(At(L')) © {\XH\X, X G R} = Tx(At(L')) © /X(T). Hence it follows that TX(L) is a Lagrangian plane in TX(K), that is, L is a Lagrangian
206
A. M. Vinogradov and B. A. Kupershmidt
manifold. In other words, L is the union of the trajectories of XH issuing from the points of the submanifold L\ Hence it is clear that, although the field XH and the operators {At} are connected with a specific choice of the function H having F as a level surface, the manifold L above, which is a solution of the H-J problem, does not depend on this choice, that is, is determined only by F and the pre-Lagrangian (n — 1 )-dimensional manifold L' lying on F. 8.5. From what we have said above it follows that a solution of an H—J problem is everywhere locally of the form L = UAt(L')y where L' is an {n — 1 )-dimensional pre-Lagrangian manifold in F (generally speaking, this is not true globally). Therefore, at least locally, a solution of an H—J problem reduces to 1) finding all possible L' and 2) constructing L from L'. Technically, 2) is connected with the solution of a system of ordinary Hamilton differential equations with initial conditions lying on L', which are in principle always soluble. We now consider the question of constructing {n- 1 )-dimensional pre-Lagrangian manifolds L' C F = {H = c}. We suppose first that K = T*(M) and TT\L> is a diffeomorphism of L' onto TT(L') = N. Then N is an (n - 1 )-dimensional submanifold of M. We note, next, that for any point x £ TV we have the natural embedding o^ = (^(N): TX(N) -* TX(M) and the associated projection /L = QL*: THM) -> T*(N). Let Tfi(M) = U T*(M) C T*(M). Now is clearly a (2n - l)-dimensional submanifold of T*(M) that maps naturally onto 7*(A0 via the map j3 = a* = a*(N): T$(M) -• T*(N), where P\T*(M) ~ Px- Thus, j3 is a linear fibre map, that is, the fibre T*(M) is mapped by j3 onto the fibre T*(N). In particular, all the inverse images Ry = (/TMCy) = O?; 1 )^), y e T*(N), are lines in the fibres T*(M). We call these lines characteristic with respect to N. Now let F^ = F fl 7^(M); then F^ is a (2« - 2>dimensional manifold with singularities. We denote by Fjy the non-singular part of F ^ , that is, F^ with its singular points discarded. Then F^ is a {In - 2)-dimensional manifold in T$(M) C T*(M). We call a point z e T% characteristic if the characteristic line R , y = ]3(z), touches F^ at z, that is, Tz(Ry) C ^ ( F ^ ) . The point of the concept of a characteristic point is made clear by the following proposition. PROPOSITION 8.1. The map j(3|ro is a local diffeomorphism everywhere except at the characteristic points. If a point z € Fjy is not characteristic, then Ker (dfi)z = Tz(Ry) £ TX(T%), therefore, W f l l , ^ ^ ) = rfOJl^k is an isomorphism. These same arguments show that Ker(d(P\ro )\z) = ^ ( R ^ ) , if z is characteristic."
The structure of Hamiltonian mechanics
207
COROLLARY S.I. If P C T*(N) is some m~dimensional submanifold and P O j3(F£) = 0 , where Tfc is the totality of all characteristic points in Fjy, then P = (]3|ro ) - 1 (P) is an m-dimensional submanifold of Fjy. Clearly, 0 has the following natural property. PROPOSITION 8.2. P*(pN) = P\TMM)' wnere PN is tne universal \-form on T*(N) and p that on T*(M). COROLLARY 8.2. The line Tz(Ry) C Tz(Tfi(M)), z G Tfi(M), y = 0(z), is a direction line of the hypersurface T$(M) C T*(M). We must show that O,(£, r?) = 0, £ G Tz(Ry), r? G TZ(T$(M))- But ^PIr^(M) = 0*WPjy)- Therefore, if X is a vector field on such that Zfl G r fl (Rj (a) ) for all a G 7$(M) and £ = * z , and if F G 8(ft(ilf)), F z = T?, then d(P*(j>N))(X, Y) = X(P*(pN))(Y)-Y[p*(pN)(X)]. But X is compatible with 0, and fi(X) = 0. Therefore, P*(PNXX) = pN(RX)) = 0 and X(0*(pN)) = P*(PiX)(pN)) = 0, that is, d(P*Q>N)XX, Y) = 0. In particular, 0 = COROLLARY 8.3. T/ze /orm 12| r o Nr c w non-degenerate, and the map j3|ro xr c /5 locally a canonical diffeomorphism of(T% \ F^, ^2|ro x r c ) Combining Corollaries 8.1 and 8.3, we came to the following result. PROPOSITION 8.3. Let L" C T*(N) be some Lagrangian manifold. Then L' = ( P l r f t u ^ r 1 ^ " ) C ( r ^ \ T^) is a Lagrangian submanifold of the canonical manifold (T^ \ F^, r2|ro x r c ) . In particular, dim L' = n - 1, S2| r = 0, and L' C T. Proposition 8.3 provides us with the method of constructing (n - 1)dimensional pre-Lagrangian manifolds L' C T that we were talking of at the beginning of this section. For, given some function \p on N, we can take for L" the graph of the differential of this function, that is, L" = sdf(N) C T*(N)9 and then find the inverse image (/3|ro X p C )" 1 (I) = L, a task that is simplified by the fact that j£3|ro Nrc is a local diffeomorphism. Since from the local point of view any Lagrangian manifold in T*(N) can be represented by means of a suitable canonical transformation in the form of the graph sdf(N), the above procedure makes it possible to construct all the pre-Lagrangian manifolds L' described in Proposition 8.3. Finally, we remark that the method described above for constructing the manifolds L' works not only when K = T*(Af), but also when K = U C 7*(Af) is some domain of T*(Af). To see this we need only restrict all the above arguments to U. Furthermore, every domain V of an arbitrary canonical manifold K in which canonical coordinates are introduced is canonically isomorphic to some domain U C T*(M), therefore, our method works inside such a domain. Finally, we recall that any point x G K has a canonical neighbourhood V. Thus, this method allows us to
208
A. M. Vinogradov and B. A. Kupershmidt
construct, at least locally, all possible pre-Lagrangian manifolds Z/ C F. REMARK 8.2. Suppose that N = {<7i = 0} in suitable local coordinates (ql9 . . . , qn). Then to specify a function
The structure of Hamiltonian mechanics
209
§9. Hamiltonians of mechanical type
In this section we define and study Hamiltonians of mechanical type. The reason for singling out these Hamiltonians from the rest is the following. Firstly, the kinetic energy of a mechanical system is a positive-definite quadratic form on its configuration space M, in other words, a Riemannian metric on M. Secondly, as is well known, the energy of a mechanical system is made up from its kinetic and its potential energy, which depends only on the position of the system. Therefore, the Hamiltonian (= energy) of a mechanical system has the form: Riemannian metric (regarded as a function on T*(M)) plus TT*(F), F G jf(M). Relativistic mechanics, which is concerned with the Minkowski metric, makes it necessary to consider also pseudo-Riemannian metrics. For this reason it is natural to call the sum of a pseudo-Riemannian metric and a function of position, that is, a function of the form TT*(F), V G JF(M), a Hamiltonian of mechanical type. Of course, not every Hamiltonian of mechanical type is the Hamiltonian of a real mechanical system. However, it is convenient to study the common properties of genuine mechanical Hamiltonians in a wider framework. We remark that the possibility we pointed out above of regarding kinetic energy as a Riemannian metric leads to deep interconnections between geometry and mechanics, many aspects of which are still waiting to be studied. We give some examples of this kind in the next section. 9.1. A metric on M is a non-degenerate symmetric ^(M)-bilinear form G(X, Y) on the ^(MJ-moduIe 3)(M). Its non-degeneracy means that the map g: 3)(M) ->• A1(M), g(X) = G(X, ) is an isomorphism of ^(M)-modules. As well as by G, a metric can be specified by a non-degenerate symmetric
^(M)-bilinear form G on A^M): G(u, v) = G(g"l{oi),
g-l{v)\u9ve\l{M).
The corresponding ^(M)-quadratic form G(CJ, co) defines a function G G F($) by the ^equation G(co, co) = S*(G). Clearly, G is uniquely recoverable from G. We call a function H G F($) o£the form G + TT*(K), V G F(M) a Hamiltonian of mechanical type, G the kinetic and 7r*(K) the potential energy^oi the system. We show below that the representation of H in the form G + 7T*(K) is unique. 9.2. To compare our approach to mechanics with the usual one based on variational principles (that is, with the Lagrangian approach), we need some new concepts. First of all, we have to understand the invariant meaning of a system of second-order ordinary differential equations on a manifold, which we need for an invariant treatment of the Euler-Langrange equations of motion. We recall that such a system in Rn has the form Qi = fM> i ) 0' = 1, . . . , » ) or, equivalents, hi = vu vt =ft(q,
v) (i = 1, . . ., 7i).
210
A.M. VinogradovandB.A.
Kupershmidt
The latter is a system of first-order differential equations, with which, according to §0, there is associated a vector field on a 2«-dimensional manifold with local coordinates (q, v). In view of the geometrical meaning of the equations qt = vf it is natural to assume that this field is given on the tangent space to the configuration manifold M, and the equations qt = Vj express the fact that it is special. More precisely, a field Y e ®(T(M)) is called special if dn(qv)(Y(qv)) = v y/{q, v) E T{M). A more substantial global definition of a special field is based on the concept of a universal derivative. We call a function r{f) G ^(T(M)) a universal derivative of / E &{M) if for all X E @(M) (9.1)
S*x(x(f)) = * ( / ) .
Here Sx: M -* T(M) is the graph of the vector field X, that is, Sx(q) = Xq E Tq(M) C T(M), q E M. Clearly, r defines an R-linear derivation of the ring ^(MJinto the j^(M)-module F(r(M)). Since (9.1) is satisfied for all / G ^ ( M ) , it can be written in operator form: (9.2)
55OT =
I.
Using r, we can define special fields on T(M): a field Y E 2)(T(M)) is special if 7 o TT* = r. This definition is, clearly, equivalent to the first. 9.3. Hamiltonian vector fields on T*(M) are closely connected with special vector fields on T(M). This connection is realized by means of the Legendre map, which we are now going to describe. Let X e 2)(T*(M)). The Legendre map associated with the field X is the map ax\ T*(M) •+ T(M) that is defined by (9.3)
ax{q, p) = (, dniq, p)(X(q, p))) 6 Tq(M)a
T(M).
We note that ax = ax if and only if (Xx ~X2) is 7r-vertical, that is, X —X 2) = 0. If the Hamiltonian field XH is vertical, then X = 7r*(/z) for some h G jf(M)\ hence, we have the following result. LEMMA 9.1. ax = ax if and only if H = H' + ir*(h), h £ f(M). H
H
Being clearly fibred, ax takes sections of the cotangent bundle to sections of the tangent bundle, and we arrive at a map ax: A.X(M) -• 3)(M). It follows from the definition that (9.4) ax(co) = S*OXOJX*. When X = XH, we write aH and o:^ instead of ax and ax, respectively. It then follows from (9.4) that (9.5)
^H(
THEOREM 9.1. If
OLH
= S*>({H, «•(/)}).
is a diffeomorphism, then
&H(XH)
is special
We verify that a^XH) o ?*(/) = afi~l o j ^ a j o v*(f) = T(f), V / E
The structure of Hamiltonian mechanics
211
Since TF° aH =TT, this reduces to the equation oc^~1o XHo TT*(/) = = c$~ 1 ({#, 3t*(/)}) = r(/), which, by (9.1), is equivalent to the fact that Sfoafi-^dH, Jt* (/)}) = Y{f) for all 7 G Z>(M). Setting co = ^ ( D and noting that Y = aH(co), <%*o SY - Sw , we arrive at (9.5)." 9.4. Now we can give an invariant characterization of Hamiltonians of mechanical type. THEOREM 9.2. A Hamiltonian H is of mechanical type if and only if the map aH is an isomorphism of ^{M)-modules or, equivalently, c %lr*(M): T*{M) -+ Tq(M) is an isomorphism of linear spaces for all q <EqM. The equivalence of the two formulations follows from the method of constructing aH from aH. If H is a Hamiltonian of mechanical type, then it follows from the definition that aH is an isomorphism. Let aH be an isomorphism. We define G G F(<S>) by 2TT* ° SfriG) = {{H, n*(/)}, n*(/)}, / G ^(M). It is not hard to see that aH = ag. By Lemma 9.1, H - G'= ir*(V), FG ^(M). • Let F = 0 (a free system) and G = glJ(q)pjPj (locally). A straightforward calculation shows that the equations corresponding to the field OLH(XH) have the form (9.6)
V* + Tkkhqi = 0
(i = 1, . . . , » ) ,
where F ^ are the Christoffel symbols of the Riemannian connection associated with the metric || gi}-\\ = || gl} II"1. This is the equation of a geodesic of the metric || gfj\\ (see [1]). From this and the equation TT#<% = re we obtain the next result. PROPOSITION 9.1. The projection of a trajectory of Xg on M is a geodesic of the metric G. REMARK 9.1. Proposition 9.1 admits the following interpretation: "a ball rolling by inertia" on a Riemannian manifold describes a geodesic. 9.5. The connection between the Lagrangian and the Hamiltonian approaches to mechanics consists in the following. PROPOSITION 9.2. / / aH is a diffeomorphism, then the system of second-order equations corresponding to the special field OLH(XH) is equivalent to the system of Euler-Lagrange equations corresponding to the Lagrangian X = (o^1 )*(p(XH) - H). We cannot give here an invariant proof of this proposition because for this we would need an invariant account of the Lagrangian formalism (see [6], [8]). The reader can verify it by appealing to coordinates. §10. Mechanical systems. Examples
From the mathematical point of view, to study the motion of a given mechanical system whose configuration space is a manifold M means to study
212
A.M. VinogradovandB.A. Kupershmidt
the Hamiltonian system XH on T*(A/), where H is the total energy of the system. This approach to classical mechanics can be called Hamiltonian mechanics. If we call the theory expounded earlier the Hamiltonian formalism, then we can say that the language of Hamiltonian mechanics is the Hamiltonian formalism. However, the Hamiltonian formalism, serves also as a language for a whole range of other theories. One example of this kind is geometrical optics. The equation describing the propagation of rays of light in an inhomogeneous and anisotropic medium is of the form 3
(lo.i) V
'
y £
2 df -J—ilL\ -
Oii(x)
\dxt)
-
where Oj(x) is the speed of propagation of light at the point x along the x r axis. Now (10.1) is the Hamilton-Jacobi equation corresponding to the p\ p\ p\ Hamiltonian — + — + —. The corresponding mechanical system is of mechanical type, and as a result there arises an optical-mechanical analogy which has played an exceptionally important role in the development of a whole range of areas of physics and mathematics. In fact, it was just this analogy that led Hamilton to create his mechanics (see [14]). It also played a decisive role in the research of Schrodinger, leading to the discovery of the fundamental equation of non-relativistic quantum mechanics - the Schrodinger equation (see [15]). Here is another example. The symbol of a linear differential operator SB on a manifold M can be regarded as a function on T*(M), that is, as a Hamiltonian (see [16]), Formally, to do this we must make the substitution -^— => pt in the highest terms of SB. The properties of the dx( Hamiltonian system arising in this way are closely connected, for example, with properties of local solubility of SB (see [17]). Altogether, the Hamiltonian formalism is widely used in questions of various kinds in the theory of linear differential operators (see [18]-[20]). This is the most important domain of its non-mechanical application. In this section we restrict ourselves to some specific mechanical questions connected with the use of the Hamiltonian method, and we discuss a number of examples. 10.1. Thus, to investigate the motion of a mechanical system by the methods of Hamiltonian mechanics, we have to find first the configuration space M of this system, and then to specify on its phase space, that is, on T*(M), the appropriate Hamiltonian function, that is, the energy of the system. The principles that allow us to give an answer to these two questions form the physical basis of mechanics. As soon as these questions are answered, the relevant mechanical problem becomes purely mathematical
The structure ofHamiltonian mechanics
213
and falls within the framework of the Hamiltonian formalism. Thus, Hamiltonian mechanics, as a physical theory, is made up of two parts: the physical, which consists of the principles referred to, and the mathematical, which is the Hamiltonian formalism. To find the configuration space M of a mechanical system S is a purely geometrical question, because M is, by definition, the totality of all possible positions that S can take. In every mechanical system S one can select a number of points, say Ai9. . . , Am, whose position in the physical space R3 determines the position of S (we are not considering systems of the type of a string, which have infinitely many degrees of freedom; for these the Hamiltonian formalism is more complicated, see [59]). The points Al9 . . . , Am cannot, generally, speaking, take arbitrary positions in R3 and in relation to one another, because, one says, of the presence of constraints. Analytically, these constraints are given by relations of the form fj(wl9 . . . , wn) = 0 (/' = 1, • • • > N), where wk = (xk, yk, zk) are the coordinates of Ak in R 3 . Thus, the position of S can be characterized by a point of the space R3m = {(wt1 . . ., wm)} satisfying the system (10.2)
f,(Wl, . . ., wm) = 0
(/ = 1, . . ., N).
In other words, the configuration manifold of S is the submanifold M of R3m given by the system (10.2). Here we assume that the constraints /i > • • • > fa are independent, which means analytically that the conditions of the implicit function theorem are satisfied for (10.2). In practice, systems of mechanical interest, apparently, all satisfy this assumption, from which it follows that the configuration space is a smooth manifold. It is possible to have constraints that operate instantaneously, and these are included when the system arrives at an extreme position. There are all sorts of restrictions acting as constraints of this kind. For example, appropriate restrictions do not permit the pendulum of a clock to deviate from its equilibrium position by more than an angle
214
A. M. Vinogradov andB. A. Kupershmidt
In the last example there must be such a rule for the collision of the particles. A wheel able to roll without friction on a plane is a system with nonholonomic constraints, that is, constraints imposed on the speed (or momentum) of various parts of this system. In this example, the nonholonomic constraint has the form V = cor, where V is the translational and co the angular velocity of the wheel, and r its radius. To describe non-holonomic mechanical systems the Hamiltonian formalism must be enriched with the theory of reaction constraints.1 Thus, within the framework of the Hamiltonian formalism developed above we can study holonomic mechanical systems with configuration spaces without boundary and exceptional points (or their motion in interior domains without exceptional points). We call such systems admissible. We remark that not every smooth manifold M can play the role of the configuration space of some admissible system. The fact that M is given by a system of equations of the form (10.2) shows that the normal bundle of M is trivial (see [11]). This entails, in particular, the orientability of M. There are also other algebraic-topological properties that the configuration space must have. 10.2. Since all possible positions of a mechanical system are represented by points of its configuration space M, the motion of S from the position a0 G M from time t0 up to time tx is represented by a curve !'• [t0, tx] -+ M, y(t0)
= a0, where the point y{t) G M represents the
position of S at time t, t G [t0, tt ] . The tangent vector to 7 at time t is called the state of the system at this moment of time. Mathematically, the fact that the system in question is holonomic means that any vector tangent to M can be realized as the state of S in the process of one of its possible motions. Therefore, T(M) is the state space of S. The calculation of the kinetic energy of a mechanical system in a given state is based on the following two principles: A) The kinetic energy of a material point of mass m moving with velocity V is equal to mV2/2; B) kinetic energy is additive, that is, if a system S is mentally split into two parts Si and S2i then its kinetic energy in some state is equal to the sum of the kinetic energies of S^ and S2 in those states that are defined by their motion as constituent parts of S. If S consists of finitely many material points, then the rules A) and B) allow us to find the kinetic energy by a simple summation process. In the case of a continuous distribution of mass (for example, in a rigid body) summation is obviously replaced by integration. The procedure indicated leads, by A), to some positive-definite quadratic function on Ta(M) for all a G M, that is, we obtain a Riemannian metric G on M. Then for the kinetic part of the Hamiltonian of S we must take the function See Appendix II.
The structure of Hamiltonian mechanics
215
1/4 G e ^(T*(M)) (see §9.1). The appearance of the factor 1/4 in front of G is explained by the fact that the kinetic energy is transferred from T(M) to T*(M) by the Legendre map %* = a ^ , where H^n is the kinetic part of the total mechanical energy H E ^(T*(M)) of S, and the linear Legendre map aH differs by a factor 2 from the usual metric identification of the fibres of T*(M) and Ta(M) coming from the metric
(H^n\ on T*(tf).
The configuration space M and the kinetic energy G are determined solely by the internal structure of S, that is, by the geometry of the system. In contrast, the potential energy is determined from the external conditions (external field of force) in which the motion is proceeding. Since for a system with a given geometry the external conditions can be of a very diverse physical nature, there can be no general principles for finding the potential energy. Nevertheless, in view of the fact that the potential energy depends only on the position of the system, finding the potential energy never presents difficulties in practice if the physics of the interaction of the system with the external force field is known. The remarks made in the preceding paragraphs are, as a rule, sufficient to construct without difficulty the manifold M and the appropriate Hamiltonian H on T*(M) for a given specific system S. If y: I -> T*(M) is a trajectory of XH, then IT ° 7: I -+ M describes one of the possible motions of S and, moreover, every curve in M describing the motion of S has the form shown. It seems to us that this approach to solving specific problems of classical mechanics is most simple and elegant. 10.3. It is natural to call the motion of a mechanical system free when external influences are absent, that is, when the motion is determined just by its geometry; the system itself is als£ called free. The Hamiltonian of a free mechanical system is of the form G/4, where G is the Riemannian metric on M determined by its kinetic energy. As we have seen, the geodesies of G are the trajectories of the motion of the system in question. We show now that very often the study of non-free motion of a mechanical system can be reduced to that of free motion. Let i / = ( j + 7 r * ( F ) b e some Hamiltonian of mechanical type and Pc = {H = c) one of its level surfaces. Let Uc C M be the domain consisting of those x £ M where V(x) =£ c. Then the function TT*(F) - c vanishes nowhere in Uc = Tr'l(Uc)9 and so the hypersurface Pc n Uc in />
"'
Uc can be given by the equation
l
clearly, a metric on Uc and Gc = c_l*(v) p
c
n
=
'"*'
1
*(y\ G = \. Since Gc = _ „,. . G is,
This leads t0 the
G\$c> we see that
Uc $c = !}next lemma. LEMMA 10.1. The hypersurface Pc O Uc in Uc = T*(UC) coincides with the unit level surface of the metric Gc.
216
A. M. Vinogradov andB. A. Kupershmidt
Owing to this lemma, the fields X = XH and Y = YQ are tangent to Pc n Uc, and, by the uniqueness of the field of direction lines of a hypersurface, we have Ya = Aa)Xa> where a £ Pc C\ Uc and / G J F (Pc n Uc). We can find the function / by using the linearity of : (dGc)a - f(a)dHa. But dGc - ^
^
V
- ~ '
S
°
that at
points a G Pc D Uc where Ga = c - 7r*(F)(a), we have
42
()
'that
is /=
'
i < - ; \
e
Let 7: / -> 7*(Af) be a trajectory of 7 lying in Pc O f/c, and 7 ' : / ^ / a trajectory of the field 7*(/~1) -7— on /. Then 7 = 7 o 7' is a trajectory at
of X, that is, 7* ° X = —rr <> 7*, where r' is a coordinate on /'. For ( _o 7 *]=_oy*o 7 * ;
It follows from this fact that im 7 = im7, that is, the trajectories of the Hamiltonian systems with Hamiltonians H and Gc on Pc n ^ coincide if we regard them as submanifolds of T*(M). As / =£ 0 on Pc n Uc, 7' is a diffeomorphism of/' into /, and we can transfer the parameter t' on / ' to / by means of (7'*)" 1 , that is, reparametrize /. Then 7 is a trajectory of X relative to the parameter t' on /. So we can say that the trajectories of the system XH lying on Pc O Uc are obtained from the trajectories of the system Xg lying on Pc O Uc by a reparametrization. To find the parameter t' on / we must solve the equations corresponding to the field 7*(/" 1 ) -T-. Therefore, r', as a function of r, is given by the equation at
t
t' - \ y*(J)dt. Bearing in mind what we have said and the results of §9. we obtain the following theorem. THEOREM 10.1. Every motion of an admissible mechanical system S with energy H = c is described in the domain Uc C M by a geodesic v: I ->• M of the metric 4GC, where G is the kinetic energy of S, reparametrized by —_ *(x/\' c v \V)
wnere
t
*5 tne
arc
length of v(I) in the metric 4GC.
The points of the hypersurface {V = c} - M \ Uc, where the metric Gc is not defined, clearly correspond to those positions where the system moving with energy H = c stops instantaneously on arrival (the momentum
The structure of Haliltonian mechanics
217
(or velocity) in this position is zero). If the function V is bounded, V < cOi then Uc = M for energy values c > cOi hence, the motion of S with energy c is everywhere described by the geodesies of the metric 4GC. If M is compact, this is always so. Theorem 10.1 allows us to use various facts of Riemannian geometry to obtain mechanical results. For example, the theorem of Lyusternik and Fet that on every closed Riemannian manifold (that is, compact without boundary) there is a closed geodesic (see [21]) leads to the following assertion. THEOREM 10.2. Every admissible mechanical system with a compact configuration space M for energy values H > max V has at least one periodic motion. 10.4. In this subsection we present some facts concerning symmetries of systems of mechanical type. When do such systems have non-standard infinitesimal symmetries? To answer this question we find the points y G T*(M) where dHy = d(G + ir*(V))y = 0. Clearly, we have the following result. LEMMA 10.2. dHy = 0 if and only if y G M and dVy = 0. Using now a remark in §4.2, we obtain the next proposition. PROPOSITION 10.1. A system of mechanical type can have a nonstandard infinitesimal symmetry only when the differential of V does not vanish anywhere. We turn now to non-hidden symmetries, that is, symmetries of the form X, where X G 3) (M). Let X <=* {At} and X «=> {At}. Then directly from the definition of At and X we have A*(H) = A*{G + TT*(F)) = = fif(G) + 7T*C4*(F)). Thus, A*{H) is once more a Hamiltonian of mechanical type, and so its decomposition into the components A* (G) and 7T*C4*(F)) is unique. If X is a symmetry of XH, then Af(H) = H + const, hence, A* (G) = G and ir*(A*(V)) = 7r*(F) + const, from which it is obvious that A*(G) = G, A*(V) = V + const or, equivalents, X(G) = 0, X{V) = const. Thus, At is an isometry of the Riemannian manifold (M, G) and X is an infinitesimal symmetry of it, also known as the Killing field of G. This leads to the next result. THEOREM 10.3^ Every non-hidden infinitesimal symmetry of the Hamiltonian H = G + ir*(V) is the Killing field of the metric G for which X(V) = const. To conclude this subsection, we clarify the connection between standard infinitesimal symmetries of the Hamiltonians H = G + TT*(F) and
Gc = ^ ^ 0 . Let h e &(T*(M)). Then {h, Sc} = ^ ^ ^ { h ,
n* (V)}. In view of the fact that
^
% 5} +
218
G\P
A.M. VinogradovandB.A.
Kupershmidt
= (c - 7r*(F))|i> > we have
It is clear from this that a function /iG 5 (T* (M)) that is constant on the trajectories of XH lying in Pc n Uc is constant on the trajectories of XQ lying on {Gc = 1} = Pc C\ Uc, and vice versa. Loosely speaking, the standard infinitesimal symmetries of XH coincide with the symmetries of Xg on {§c=l} . This follows from the fact that the geometrical shapes of the trajectories of XH and Xg in Pc n Uc are the same (see §10.3). In conclusion, we consider three classical examples, which have played a prominent part in the history of mechanics. 10.5. Kepler's problem. This is the problem of the motion of a particle A in the gravitational field of a fixed point mass. The position of this mechanical system is uniquely determined by the position of A, that is, its configuration space is R 3 . Suppose that the fixed attracting mass is situated at O G R 3 . This point is clearly exceptional. Let qlt q2, q3 be Cartesian coordinates in R 3 , q = \Jq\ + q\ + q\, q = (qlt q2, q3)\P\, p2> Pz t>e the respective momenta, p2 = 2 pf and p = (plt p2, p3). For an appropriate /= l
choice of units, the Hamiltonian function of the system in question has the form H = ^r — —, where o: > 0 is some constant. Here G = p2/2 and G is the standard Euclidean metric on R 3 . Rotations about O are isometries of R3 that preserve the potential energy -oc/q. An infinitesimal rotation about the qt axis is of the form Xt = - qi+ x -^
+ qi+ 2 ^
, where the indices
are taken mod 3. Then the Hamiltonian of the field X t is equal to Mt = -qj+xPt+2 + Qi+2Pi+i' By Lemma 4.2, {#, Mt}= 0. We note that {Mu Mi+1} = Mi+2. This shows that the linear span of the functions Mj(i = 1, 2, 3) forms a Lie algebra under Poisson brackets, which is isomorphic to the Lie algebra of the group SO(3) (see [2]). Let 3
M2 = 2 Mf. Then, clearly, {#, M2} = {Mi, M2} = 0. Therefore, z=l
fi = H, fi = M2, f3 = Mx is a complete collection of pairwise commuting symmetries of H, s(H) = 3, and Kepler's problem is completely integrable. REMARK 10.1. We have exploited here the following useful method of finding pairwise commuting symmetries. Suppose that the Hamiltonian system admits a finite-dimensional algebra L of symmetries, generated by, say, the functions hl9 . . . , hk. Then every invariant of L (a polynomial over L* that is invariant under the adjoint action of L on L*), regarded as a polynomial in the hi9 commutes with all the hi9 and, clearly, also with H.
The structure of Hamiltonian mechanics
219
We recommend the reader starting from this to find an exact solution of Kepler's problem, using the methods developed above. We remark that the preceding arguments are valid for any central potential, that is, a potential of the form tp(q). The Newtonian potential — a/q has, over and above this, the following remarkable symmetry property, Let M = (Mx ,M2,M3) and A = MX p - - q (we use the standard vector notation). Then, as can be verified immediately, {H, A i} = 0, where A = (A x, A,,A 3 ) . The functions Nj= 2 | / / p Aj are also, clearly, symmetries of//. Furthermore, a straightforward calculation shows that
X
{Mh ^ } = 0, {NU Ni+i} =
±Mi+2,
where the plus sign corresponds to the case H < 0 and the minus sign to the case H > 0. The formulae (10.3) show that the linear span of the functions Mj, Nj (i = 1, 2, 3) forms a Lie algebra of dimension 6, isomorphic to the Lie algebra of the group 0(4) for H > 0 and of 0(3,1) for H < 0. The triple of pairwise commuting functions//, N% a n d .? Ni f o r m s a collection of commuting symmetries of //that cannot be expressed functionally in terms of//, M, and M2. This gives us yet one more example in connection with the problem discussed in §5.3.
10.6. Geodesies on an ellipsoid. This example demonstrates the fruitfulness of the "mechanical" approach to a geometrical problem. We want to find the geodesies on the triaxial ellipsoid
=:1
} ' a > b > c > °- Let G be the metric onM
induced by the embedding M C R 3 . From the mechanical point of view it is natural to examine the Hamiltonian H = G/2 of the corresponding free mechanical system and investigate it for symmetry. If we could find a symmetry of this Hamiltonian, independent of it, this would mean that it is completely integrable. We note that, as the ellipsoid has no infinitesimal symmetries, the symmetries of H, if there are any at all, must be hidden. The geometrically natural coordinate system on some surface K R 3 is the system of its lines of curvature. A practically useful way of finding the lines of curvature is based on a theorem of Dupin, which consists in the following. Let Ft(x, y, z) (i = 1, 2, 3) be functions on some domain U C R 3 , where the level surfaces of the F,- intersect at right angles at all points of U (a tri-orthogonal system). If P = {Fi = c) and u ~ F2\P, v = F3\P, then (u, v) is a coordinate system on P for which the lines u = const and v = const are lines of curvature of P (see [22] p.334). The functions Xz(x, y, z), where \t(i = 1, 2, 3) are the roots of the equation —^-r+ f^r + - ^c— A= 1, form, as is not hard to check, a tria—A b—A orthogonal system and M = {kx = 0}. The resulting coordinates (w, v) on M are connected with the Cartesian ones by the equations x
2
_ a(u — a) jv—a) ~~ (b-a)(c-a)
'
y
2
_ b (u—b) (v—b) (a — b)(c — b)
'
2 Z
_ c(u—c)(v—c) ~" (a—c)lb — c)
'
In the coordinates (w, v) the metric G can be written in the form i
^
l
]
,
where q>(">)= ( . _ w ) ( 6 I l > ) ( t _ w ) •
220
A. M. Vinogradov and B. A. Kupershmidt
Therefore, (10.4)
H = -±
where (u, v, plf p2) is a special canonical coordinate system on T*(M). We rewrite (10.4) in the form (10.5)
- ^ — #M
=
_ E L — Hv.
Then this equation gives H implicitly. We now make use of the following general lemma, which lies behind the method of separation of variables. LEMMA 10.3. Let o f o , . . . , \pk, H) = Mx, . . . , \pt, H) be an equation giving H implicitly, where
J?}= 1=1
j=l
k
But the expression
S c^.o^. {
symmetric in / and /, is zero." Applying this lemma to (10.5), we find that the function f)2
p2
Q2
h = * — JJU = _Ji_ —1— + -JL- 2 i s a symmetry of the Hamiltonian H and clearly independent of it. The two-parameter family of Lagrangian manifolds Mc c = {H = q} O {/^ = c2} can be put in the following form by expressing px and p2 in terms of H = cx and h = c2, Pi = ± y/^p(u)(ciU + c 2 ), P2 = ± V ^ ( u ) ( c i y + ^2)- From this we find an explicit expression for the two-parameter family of solutions of the corresponding Hamilton—Jacobi problem: (10.6)
S(u, v, eu ca) =
We take H = (Si and /z = fi2 as one half of new canonical variables and complement them by the corresponding variables gPt and e^a. Then in these
The structure of Hamiltonian mechanics
221
coordinates Hamilton's equations take the form $i = (t2 = 0, gPi= —1, ^ 2 = 0, so that &i=Ci (*=1, 2), 0*4 = — t + c3i 0*a = e4,C| GR(i = 1,. . . , 4). But &t =-•!£-
(= - P~)> therefore,
For fixed values of the constants clt c2 and c 4 , (10.8) is a relation between w and v, that is, it gives an implicit equation of the geodesies. A parametrization of the geodesies can be derived from (10.7). 10.7. A rigid body. The Hamiltonian formalism on lie groups. By a rigid body we mean in mechanics a system of material points (finite or infinite), the distance between any pair of which is invariable. An orthonormal frame of reference rigidly connected to a rigid body uniquely defines the position of the points of this body in R 3 . We consider a free rigid body. The configuration space M of this mechanical system clearly coincides with the totality of all orthonormal frames of reference % = {a\ 8i, e2, 83} in R 3 , where a G R3 and the ordered triple of vectors e^ is orthonormal. Therefore, M = R3 X SO(3). We fix one such frame g0. Then every frame % G M uniquely determines a motion g{%)\ R3 -* R3 carrying g0 to %. This lets us identify M with the Lie group S (3) of all motions of R3 . L e t % G M. Then T
is equal to m 2 vf/2, where m is the mass of the body and z=l
v = (i>i, v2, v3), and the second is a quadratic form in the coz-(/ = 1, 2, 3), 3
where co = S cof et. By obvious symmetry arguments, it does not depend /i
We recall that w G R 3 denotes the rotation about the dxis of the vector w in the positive direction with angular velocity I co i.
222
A. M. Vinogradov and B. A. Kupershmidt
on %. Therefore, So can be affixed to the body (at the centre of gravity) 3
in such a way that this quadratic form is diagonal: 2 /fco?/2. The constants /-l
If are called the principal moments of inertia of the body in question. Thus, 3
3
G
n
/= l
/= l
X = 2 atX( (respectively, co = 2 a^f),
where a* 6 *F(^). In particular,
every metric G o n 3 c a n b e written in t h e form G(X, Y) = 2 /*//«/&> where X = 2 a , * , , 7 = 2 ft*,, M / ; = M// e j F ( £ ) . /,/= 1
/= 1
i-1
If G is left-invariant, then the ixi}- are constants. The formula for the kinetic energy of a rigid body is, as is now clear, an expression of a similar kind when 3 = 3(3) = M. Since G is left-invariant, the Hamiltonian H = G/4 of a rigid body is left-invariant. If (pfi vt) are coordinates in T*(M) dual to (vi9 ojj), then
The structure of Hamiltonian mechanics
223
clearly H = 2 ( 2 1 . + J l . ) . Thus, the theory of the motion of a rigid body leads us to the necessity of considering left-invariant Hamiltonian systems on Lie groups. The theory of left-invariant Hamiltonian systems on a Lie group g is simplified considerably if we discuss only the left-invariant part of the Hamiltonian formalism on T*{S). This is justified, because lg(p) = p V g £ G (Corollary 2.2), that is, the fundamental forms p and dp on r*(G), as well as the map F, are left-invariant. This allows us to make full use of the property of being left-invariant, which clearly entails the fact that § is a group of (non-hidden) symmetries of the leftinvariant Hamiltonian H on T* (*§). Thus, we consider the left-invariant Hamiltonian formalism on T*(&), which means that we deal only with functions, forms, fields, etc. on T*{3) that are invariant under the diffeomorphisms lgi g G § , Every leftinvariant function
leads to [£ /f Xf) = 2 equation
^. Applying the form p to both sides of the
dpOi ^(Ily) = II,., n ; =•• dp{Xit Xj) = X((p(Xf)) ~ Xf(p(Xt)) - pdX,., A)]), n ., hence ^ / = 2 ( 2 c^irk ) ^ - . If we obtain Xt
=
k=l
/ ( ^ I > • • • > Kn) is some left-invariant Hamiltonian, then
/(#) = /(TT15 . . . ,
TTW)
and * # = 2 - ^ - Xn. = 2 -^— £ f . Hence
To finish this brief excursion into the left-invariant theory, we remark
224
A. M. Vinogradov and B. A. Kupershmidt
that any function h invariant under the adjoint representation of g on L* is a symmetry of any left-invariant function if we regard it as a function on T*(&) (it is obvious that h is an invariant if and only if Tj(h) = 0). For more details about this, see [23]. Let us return, however, to the rigid body. Above, we took as a basis of Te{3(3))=:T<$o(M) =r f l o (R 3 ) © Te(SO(3)) the vectors elf e2, e3 of unit translations along the inertial axes of the body in question (a basis of Ta (R 3 )) and the vectors e 4 , es, e6 of unit rotations about the axes ex\ e2, e3 (a basis of Te(SO(3))). Then
!
\ei,e3\ = [ei, ei+3] = 0, l < i , / < 3 , [e4, e5] = e69 [e5, ee] = e4, [e6i e4] = e5, [e4, e2] = e3, [ek, e3] = — e 2i [e5, et] = — e3, [e5, e3] = eu [e6, et] = e2, [e6, e2] = — ej. 3
v2- v
p*
When we now apply (10.9) to the Hamiltonian H = 2 ("2^+27:) t and calculate the values of the structure constants, which are given by (10.10), we obtain (10.11)
Thus, the left-invariant Hamiltonian system for a rigid body is:
(10.12)
The Legendre map a^ is clearly left-invariant and gives rise to a map aj^: L* -* L. In the case in question, alH is given by the formulae Vf = Pj/m, coz- = ^//z-. Therefore, the system of equations on L associated with the field &lH(XlH) has the form
(10.13)
i .
j,-l,
'
aa
h h
~
/2
gyp,, I] . ( j ' - / a ^3
m,m..
The first three equations of this system are, as is not hard to see, the laws of conservation of momentum along the coordinate axes, written in the moving frame g, and are therefore of no interest. But the second three equations are the famous Euler equations of motion of a rigid body. For this reason, the field <xlH(XlH) on L in the case of an arbitrary left-
The structure of Hamiltonian mechanics
225
invariant Hamiltonian H on T*(3) can be regarded as a generalization of Euler's equations for a rigid body (provided, of course, that aH is a 3
3
diffeomorphism). The functions p2 = 2 p2 and v2 = S *>? are invariants //= l / l of G(3), so that A^a = X\? = 0; hence they are symmetries of H. The projection of the manifold {H — c v} n {p2 = c2) D {v2 = c3} on the *>-space (= T*(SO(3)) C 7$o(S(3))) is given by the equations (10.14)
S ^ .
= Cl
-.^.f
*
= c ,.
Thus, the motion of a rigid body is described in non-parametric form by the curves (10.11) or, after applying aH, by the curves (10.15)
3
2 /*©?= const, t=i
3
H /?©? = const, i=i
which are, therefore, the solutions of Euler's equations. The following connection between the trajectories of X H and of the "Euler field" <xlH(XlH) is an immediate consequence of the definitions: if oo(/) is a trajectory of the Euler field, then the projection y(t) of the trajectory 7(0 of XH on & is of the form 7(0 = exp co(f), and vice versa. Here exp: L-* & denotes the exponential map (see [2]). The case of a rigid body fixed at one or two points can be treated similarly. The theory expounded above is, as we have seen, the restriction of the Hamiltonian formalism to the subring of JF(T*(S)) consisting of leftinvariant functions. A similar theory can be constructed, if instead of this subring we take jany other subring of JfiT*^)) that is invariant under the operators lgy g G &. The smaller this subring, the simpler the resulting theory. The study of the motion of a rigid body in, say, a gravitational field can most conveniently be brought within the framework of such a theory. APPENDIX I Bifurcations Here we want to describe how the manifolds Pc vary with a change of c G R*. It is not hard to give to the loose arguments presented below a rigorous meaning in the framework of the theory of singularities of smooth maps (see, for example, the survey [6]). In particular, this refers to the concept of "general position", which lies at the basis of our discussion and which nonetheless we leave undefined. We begin with the following general situation: K is some m-dimensional manifold,/!, . . . , / * £ &(%), k < m, is a collection of functions in
226
A. M. Vinogradov and B. A. Kupershmidt
"general position", Pc = {x G K I f^x) = ct, i - 1, . . . , k }, c = (c l 5 . . . , ck) G R*. Then we make modifications, taking into account the fact that {ft, fj) = 0. Let Bif (ft, . . . , fk) = {a; G A' | <2ft \x, . . . , dfk\x are linearly independent }. Generally speaking, Bif (ft, . . . ,fk) is a manifold with singularities, whose points in "general position" are the x for which rank (dfx \x, . . . , dfk\x) = k - 1, dfti, =* 0 (/ = 1, . . . , k). We restrict ourselves to the analysis of these points. Let P'c. = {x GK\ft0c) = cz, i = 1, . . . , k- 1}, c' = ( c 1 , . . . > c ^ 1 ) € R * -1 . For almost all c G R*""1 the manifold P'c> can be assumed to be non-singular, that is, rank (dfx \x, . . . , dfk_x \x) = k - 1. A point x lies in P'c> n Bif CA, . . . , /„) only when ^(4|i?0|x = 0, that is, when x is an extremum point of fk\P>, on P'c>. The extremum points of a function in "general position" form a discrete set (see [6]), so that the set P^> H Bif (/j, . . . , fk) can be regarded as such. When the (k — 1 )-dimensional parameter c varies, each of the x G Pc'» Pi Bif (/i, . . . , fk) also varies and describes a (k— 1 )-dimensional manifold. Thus, Bif (A, . . . , fk) = U (P^ O Bif (/*!, . . . , fk)) is a (k- l)-dimensional manifold with singularities. We now consider the map F: K -> R^, F(x) = (/^x), . . . , /^(JC)). Then the set bif (ft, . . ., fk) = F(Bif(ft, . . . , /^)), is, in general, also a (fc - 1 )-dimensional manifold in R* with singularities and self-intersections. It is called the bifurcation diagram of the system .ft, . . . , fk and it splits F(K) C R* into a number of connected domains, say Bt. These domains have the property that under some natural conditions of completeness all the manifolds Pc, c G Bf, are diffeomorphic, while Pc and P? are not diffeomorphic if c and c belong to different domains Bt. Thus, the manifold Pc changes on passing through the bifurcation diagram from one domain Bf to another. Let us prove this.
Let Bt = F-l{Bt\ x G Bt, c(x) = (ft(x), . . . , fk(x)). Then_
Ker dFx = TX(PC(X))- We introduce a Riemannian metric on Bt (this is always possible (see [1])) and consider an orthogonal complement Ox to Ker dFx in TX(K) in the sense of this metric. Now dFx\0 : Ox -•r ^ ^ R * ) is an isomorphism, since dFx is an epimorphism because x G Bf. Every field Z G 3}(Bt) can be lifted to a field Z G 35(Bt), by setting Zx = (dFx\0 )~l(Znx)). Here, clearly, Zo F* = F*Z. It follows from this equality that F o Cr = Ct<> F, where {C*} <=* Z, {CJ «=> Z at places where Ct is defined. In this context we call F full if Ct is defined on F~l (V) whenever Ct is defined on V C 5 f .
The structure of HamUtonian mechanics
LEMMA. Let F be full
Then all the fibres Pc, c G Bi9 are
227
diffeomorphic.
PROOF. Let c, c G Bt and let Z G 3) (Bt) be_ a field such that one of its trajectories passes through the points c and c, and Ct(c) = c for some f G R. If F is a neighbourhood of c where Ct is defined, then Ct is defined on F~l(V) D Pc. Since F ° C, = Ct<>F, we have Ct(Pc) C P-, and C_r(P-) C P c , from which it is clear that (?t\p : Pc -> P- is a diffeomorphism. We give the following almost obvious proposition without proof. PROPOSITION. 1) A map F is full if VZ G 3) (Bt) and V# G Bt, the trajectory of Z starting at x is defined for those t for which the trajectory of Z starting at F(x) is defined. 2) A map F is full if the preceding condition is satisfied for some system of fields {Za} that generates the &{B^-module S(£*). 3) The property of a map F of being full does not depend on the choice of the Riemannian metric in Bt. 4) / / F is full then F\$;. Bt -• Bt is a smooth fibering with the fibres Pc = F'Hcl cGBf (see [7]). We turn now to the case of interest to us, that of a canonical manifold K and a commuting system of essentially independent functions fl9 . . . , fk. Then a point x lies in P^ Pi Bif(fi9 . . . , fk), as before, if d(fk\p>,)\x = 0. But x cannot be isolated, because of the conditions Xt(fk) = {ft, fk} = 0, Xt = Xf. For it follows from Xf(fk) = 0 that d(fk\p>,)\Ai = 0, where {A]} <=* Xt. More precisely, there is some C
t\X)
_
(k- l>dimensional manifold Nx> C Pc> passing through x and consisting of points where d(fk\P>,) = 0. The manifold Nx is constructed from fu - - > h-\ Just as Nx from fx, .. ., fk. Thus, bif (A, . . . , fk) is a (fc- l>parameter family of {k- l)-dimensional manifolds, that is, a manifold of dimension 2k - 2. Nevertheless, Ker dFx O Tx(Bif(fi9 .. . , fk)) = TX(NX) is o f d i m e n s i o n k-\, s o that Bif(/\, . . . , fk) is, as before, a (k- 1 >dimensional manifold with singularities. For the rest, everything proceeds as before. We consider one simple example that illustrates the theory of bifurcation: a "particle in a central field". In this case Af = R2 \ {0}, T*(M) **MXR2
a n d # = - - ^ — + U(x), where (,, q 2) are Cartesian
coordinates in Ra DM, (qp pfi (i- 1, 2) is the corresponding special coordinate system in T*(tf), x = sj(q\ + q\). The Hamiltonian His, clearly, symmetric with respect to rotations about the origin, and 3
the field X& 35{M) corresponding to the one-parameter group of turns about the origin is q x Therefore, the Hamiltonian/, of X is qtp2 -p%q2, pair/j = / / , / a . Since
d
q%—. dq2 dqt and {/2, H) - 0. We consider the bifurcations of the
t— Pi
dq2,
the set Bif (f^, / 2 ) consists of the points where the rank of the matrix
228
A. M. Vinogradov andB. A. Kupershmidt
/
Pi
\—Q2
Pi
y (x) q\
V {x) q
P2
—Pi
Ql
where V(x) = U'{x)/x, does not exceed 1. An elementary calculation shows that B i f ^ , f2) consists of "pieces of two types: Bifj = {p1=± s/(V(x))qif p a = * s/(V(x))qlt V{x) > 0 } -points in "general position"; Bif, = {JDJ =p2 = 0, V(x) ~ 0 } -points "not in general position". The projections of these pieces on the plane R2 = {/ lt / 2 } are *
TTt
/-A
-.
±
U'(y) > 0,y > 0 is a parameter;
where a^ = £/(*/) and the Xj are roots of the equation U'(x) = 0. Thus, bif(f lf / 2 ) = F(Bif t ) U F(Bif 2 ). For example, let £/(JC) = 1/x ("repelling potential"). Then Im F = {ft > 0 } r bif(/ t ,f2) = Q. But if ^(x) = -l/x("attractingpotential"),thenImF=R 2 ,bif(/ 1 ,/ 2 )= { 2 / ^ 1 = — 1 } . For other examples, see [26].
APPENDIX II The Hamiltonian form of mechanics with friction, non-holonomic mechanics, invariant mechanics, the theory of refraction and impact
A. V. Bocharov and A. M. Vinogradov In this appendix we present the Hamiltonian formulation of the branches of classical mechanics indicated in the heading, about which nothing is said in the main text. Here, in contrast to the traditional approach, we are in no way guided by the Lagrangian point of view. The invariant Hamiltonian formulation of the branches of mechanics discussed in subsections 2, 3, and 5 is apparently new. 1. Mechanics with friction. The action of frictional forces is manifested by a decrease of the momentum of a mechanical system in "the direction opposite to its direction of motion". Therefore, the resistive action of frictional forces can be described by adding to the non-resistive Hamiltonian field of the system a vertical vector field on T*(M), in which all the vectors that are tangent to the fibres T*(M) "face" the point (JC, 0) G T*(M). In particular, Xp is such a field. Hence, any field of frictional forces is of the form \Xp, where X is a non-negative function on T*{M). Thus, if 38 is the usual Hamiltonian of the system in question, then its motion, when friction is taken into account, is described by a vector field on T*(M) of the form X = Y~\dM + Xp). The function X is the coefficient of friction, and its form is determined by the concrete physical conditions. 2. Mechanics of non-holonomic systems.1 The kinematic constraints imposed on a system by their very meaning must be understood as a For a Lagrangian treatment of non-holonomic systems see [28]. See also [29].
The structure of Hamiltonian mechanics
229
certain submanifold IT C T*(M) on which the phase trajectories of the system have to lie. Thus, it is necessary to indicate the principles on which a vector field Z G 3) (II) is constructed for a given Hamiltonian of mechanical type. It is convenient to formulate it, by introducing the concept of a partial symplectic structure. Let Q be a linear subspace of a linear space V, and P = p0 + Q an affine subspace. We say that a bilinear form ft: V X V-*- R is non-degenerate onP if it is non-degenerate on its associated linear subspace Q. If ft is such a form, then P can be uniquely written in the form p l + Q, where p t is a vector skeworthogonal to Q (with respect to ft), the so-called canonical displacement (with respect to ft) of the affine space P. If a skew-symmetric form ft is non-degenerate on P, then the map 7 ^ 'P-*Q*> 7ft(*) = x _in is an "isomorphism". We consider the isomorphism r ^ : Q -*• Q*, r^(y) = y J ft. The maps 7 ^ and r n are connected by the relations
f
in which p t is the canonical displacement of P.
Let L be a smooth manifold. We say that a partial symplectic structure is given on L if there are specified a distribution of affine subspaces C TX(L) VJC G I , and a differential 2-form £1 G A2 (I), with 1) £1 non-degenerate on J j for any x e I , 2) d£2 = 0. Together with {T°x} we consider the distribution of linear spaces T* C TX(L) and the field of canonical displacements & (T°x = c£Px + r j ) . The family of maps 7 ^ } : (7^)* -* T°, ^{Lln y (T*)* -• 7^ obviously defines linear operators By (1), Z(co) = ^ + y(co), where <£P is the field of canonical displacements. DEFINITION. We call the \-form co a generalized Hamiltonian of the vector field Z(co). When co = df we call the function f a Hamiltonian of the field Z(co). d Y The existence of the map C°°(L) -* A 1 ^ ) "^ 35(L) allows us to define on a manifold with a partial symplectic structure a pairing (/, g> = y(jy)(g) satisfying all the axioms of the Poisson bracket (see [16]), except the Jacobi identity. Let M be the configuration manifold of a mechanical system and G its kinetic energy (a Riemannian metric on M). Let Px C TX(M) be a distribution of /^-dimensional affine subspaces, realizing the kinematic constraints imposed on the system. We introduce the following objects: 1°. 0^: TX(M) -> T*(M) - the Legendre map associated with the metric 2°. a A:-dimensional distribution Tlx = OLX(PX) C 3°. a 2^-dimensional distribution r^(II) C 7^(11) on the manifold
230
A. M. Vinogradov andB. A. Kupershmidt
=
U IL :
rg(n) = {ierb(U) \dn u a e i W is the inverse image of Px under the natural projection ir: II -• M. FUNDAMENTAL ASSERTION. The distribution T° and the 2-form (dp)\n give a partial symplectic structure on IT. If we now understand by non-linear kinematic constraints an arbitrary submanifold II C T*(M), fibred over M into smooth surfaces 11^ C TX(M), then the partial symmetric structure on II must be given "backwards". Namely, let 2>_E II, 7r(2>) = x; we replace the surface 11^ C T*(M) by its linearization U(b) at b. We construct the element Px(b) =
= a ; 1 ^ ) ) C TX(M\ We set Tgfjl) = {\e Tb(Tl)\dn\b(O ePx(b)}
etc.
Let F e A1 (M) be the "field" of forceJn which the system is situated, regarded as a differential 1-form, and let G be the quadratic function on T*(M) corresponding to the metric^G. By a generalized Hamiltonian of the system we mean the 1-form E = dG - ir*(F), where ir: T*(M) -* M is the natural projection. DEFINITION. Suppose that a manifold II realizes the kinematic constraints imposed on a mechanical system, co € E\n, £2 = (dp)\n. Then the trajectories of the vector field Z(co) are called the phase trajectories of this mechanical system. We assume that the forces admit a force function F = dU, SS = G - 7T*(£/). Since a free mechanical system would move along the trajectories of the Hamiltonian vector field X^>, the vector R^ = zd&e—X^€ must be taken to be the force of reaction of the nonholonomic constraints. Let us describe the objects introduced above in special coordinates on T(M) and T*(M), assuming that the kinematic constraints are linear, and let us verify that the well-known Chaplygin—Kanjel equations (see [27]) and the formulae for the force of reaction follow from the principle we have formulated. Let dim M = n, n = k + m. 1 . The constraints Px = \ 2) asi(q)qt + as0(q) = 0, s = 1, .. . m J, where Qi > • • • > Qn> #i > • * • > kn
are
special coordinates on T(M).
2°. The Hamiltonian E = 2 -i—- dpt + 2 £,-(#, 3°. The inverse Legendre map q{ = ^— (z = 1, . . . , « ) , where i> • • - > Qn> Pi> - ' ' > Pn
are
special coordinates on
PROPOSITION. For CJ = £ | n tfie vector field Z(co) is o / tfie form
The structure of Hamiltonian mechanics n
m
n
231
^
\ s (s = 1, . .. , m) are functions on II. In particular, when F = d£/, <#?=£ - TT*(C/), we have the classical Chalygin—Kanjel equations m dqt _
2J
b&e
dpj
ffl
Si-^-+«<.o
_
= O (s=l, ...,
««-2 (2**.)-£-• i=i
3=1
If the configuration manifold of the system is a Lie group ^, and the distribution Pg C Tg(S) and the kinetic energy of the system are leftinvariant, then so is the surface II C T*(3). In this case the phase trajectories of the system are projected naturally onto the linear space IIe C I * = TtjiS), where e is the unit element of the group. Here the equations of the projected phase flow can be uniquely recovered from the element of the metric Ge: Te X Te ->• R and the structure constants of the Lie algebra L = Te(§), similar to the way this happens in the holonomic case (see §10). In particular, the classical equations "in quasicoordinates" describing the rocking of a body on a rough plane are equations of such a flow on the appropriate Lie algebra. Bringing mechanics with friction and non-holonomic mechanics together, we come to the following general principle. If a force field F and a coefficient of friction are given on the configuration manifold M of a non-holonomic mechanical system, then the non-holonomic constraints on the system must be interpreted as a submanifold II C T*(M\ and the phase trajectories o£this system are the trajectories of the vector field Z(co), where co = (dG - TT*CF) + Xp)| n . 3. The Hamiltonian formalism with values in a representation of a Lie group.
The formalism developed below is suitable for the description of mechanical systems whose configuration manifold is a Lie group &, whose kinetic energy is left-invariant, and whose potential energy is "only slightly noninvariant". We show below that the study of such a system reduces to that of a system with left-invariant Hamiltonian on a Lie group &u an extension of S. We consider the right representation of £ on C°° ( $ ) : g: f— Rg(f), where Rg is right multiplication by g. Let V c C°°($) be an invariant subspace of this representation, A( V) C C°°{^) the subalgebra generated by V, and ir*(A) = {«•(/) I / 6 A(V)}t (n : T*($) -+ % is the projection). We denote by C™(T*($)) the algebra of left-invariant Hamiltonians (see § 10).
232
A . M. Vinogradov and B. A. Kupershmidt
FUNDAMENTAL ASSERTION. C£(r*).7T*U) is a subalgebra of C°° (T*($)) and is closed under the Poisson bracket. We indicate at once the main source of invariant subspaces of the right representation in C°°(#). Let
EXAMPLE. The potential energy of a rigid body, fastened at one fixed point and situated in a uniform gravitational field, is a function on SO(3), but is not right-invariant. The orbit of this function under the right representation has a three-dimensional linear span. From now on we assume that dim V < °°. Let L = Te(&) be the Lie algebra of S. Regarding V as a commutative Lie algebra, we examine the action of I on V induced by the right representation, and the corresponding semidirect product Lj of the algebras L and V. Now the ground ring K = C£(T*)*ir*(A) can be regarded as a ring of functions on Lf = L* © V* or, equivalently (§10), as a ring of left-invariant Hamiltonians on the corresponding Lie group Si» In particular, the Hamiltonian of the original mechanical system G + 7r*(F) is of this kind ifuGF. It is not hard to check that the canonical structure arising on K from the Lie algebra L\ (see [16]) is the same as the structure induced on it by the Poisson bracket, by the fundamental assertion (see above). EXAMPLE. For a heavy rigid body fastened at a fixed point, §x is the group of motions of three-dimensional Euclidean space. The classical equations of motion of a rigid body are equations on a trajectory of the Hamiltonian vector field X^> if the Hamiltonian 3
3
of the heavy rigid body is regarded as an element of the ring of functions on WSJ. 4. Invariant mechanics on a uniform space. Generalizing the left-invariant theory, we can construct an invariant Hamiltonian formalism on a uniform space. The analogue of the ring CL(T*(&)) in this case is the ring of functions on T£ (M) (x0 is some fixed point) that are invariant under the action of the stabilizer of x0. From results of Gel'fand (see [30]) it follows, for example, that the Poisson bracket of any two O(«)-mvariant functions on r*(G*), where G% is a Grassman manifold, vanishes. 5. The theory of refraction and impact. We suppose that the phase space of a mechanical system is divided by a hypersurface F into two domains and that the behaviour of the system in these domains is described by distinct Hamiltonians. We discuss below the rules of "refraction and reflection" of a phase trajectory of such a system on F.
The structure of Hamiltonian mechanics
233
Let M be a smooth manifold and F C M a hypersurface splitting M into two connected components: M\T = Af. U M+. By a function divided by M we mean a pair of functions fT: MT UT-^R, each of which is infinitely differentiate at interior points and is extended to a smooth function / T : M -> R. (In particular, the restriction of a divided function to F is two-valued.) We denote the ring of divided functions by j?Fi(M). If M is a symplectic manifold, <## G jfT(M), then a Hamiltonian vector field X$Q and the notion of a trajectory of this field are defined on each of the components. A point x G r is called minus-opposing for SB if the vector (X^JX points into Af+, and called plus-opposing for <$? if ( X ^ ):x. points into M_. There is a natural requirement, which we now impose on a trajectory: the trajectory must pass from opposing points into non-opposing points along the characteristics of the hypersurface (see §8). Let e, E G R, E > 0 and let y_: (—8, 0] -> F U M_ be a trajectory of the vector field X^_ lying on the level surface {SB. = E}, where x = 7_(0) £ F n {c$L = £ } . We consider the characteristic / passing through x. A point y G / is called decisive for x if >> lies in / H {o^- = £"} and is not minus-opposing or j> lies in / C\ {$£+ = E) and is not plus-opposing. PRINCIPLE OF TRANSITION. A trajectory y_ can be extended from any decisive point as the trajectory of the corresponding Hamiltonian vector field (that is, the field X$Q in the case of non-plus-opposition and X$$_ in the case of non-minus opposition). Having formulated the principle of transition in this form, we deliberately refrain from quoting the theorem about the uniqueness of the trajectory, which has an obvious physical justification. We indicate (in the spirit of [16]) an algebraic interpretation of the principle of transition. Instead of J^r (M) we consider a fairly arbitrary i^-algebra K with a canonical structure { , }. Then the analogue of F in M is a self-stabilizing ideal J c= K, and the analogue of the field of characteristic directions is a family of derivations Xg- K-+K/J, g G J , xgif) = *({g, /}), where TT: K-^K/J is the natural homomorphism. Let / C R1 be an interval, 0 G /, and let Fo (I) be the ring of functions divided by the point 0. We denote by f o ( / ) c . f o ( f l the subring of functions continuous at zero. Let SB G K. DEFINITION. We call a homomorphism of i?-algebras 7: K-+ JF0(I) a trajectory of the Hamiltonian $8, divided by the ideal J, if 1°. the diagram
234
A. M. Vinogradov and B. A. Kupershmidt
is commutative; 2° ~~ 3°. it follows from x^(/) = 0 Vg£ J , that y(f)£JF0(I)We examine the case Af = T*(L); N C X is a submanifold of codimension 1, T = U r * ( I ) . The axioms 2° and 3° in the definition above have a x<=N
simple mechanical meaning; 2° is the law of conservation of energy on passing through N, and 3° implies the continuity of the projection of the trajectory on L and the law of conservation of "the component of momentum tangential to N". Let U be a domain in L, and ql9 .. ., qn, plt . . . , pn be special coordinates on T*(U) such that N = {qn = 0}, M+ n T*(U) = = fe > 0}, M_ n r * ( ^ = {gn < 0}. Let SS = SS=P(qi> . . . , qn, pl9 ... , pn). Then the set of plus-opposing points of H is described by the inequality
(3)
ig
and the domain of minus-opposing points by the inequality
(4)
i^
EXAMPLE. We suppose that the branches SS^ of $8 are positivedefinite quadratic functions on each fibre T*{L). We consider in T*(L), x G N the pair of surfaces S+ = {$8\ = E} and S_ = {(f^_ = E), Let / be a characteristic passing through a minus-opposing point on S_. It is easy to verify that a) / intersects S_ again at a minusopposing point, therefore, there is always a "reflected" trajectory: b) if / intersects S+ at two points, then exactly one of them is plus-opposing and there is a "refracted" trajectory: c) if / touches S+9 then the point of contact is non-opposing. The cases b) and c) correspond to refraction of a trajectory of the system on N, and the case when / does not intersect S+ at all to a completely internal reflection. EXAMPLE. (Capture of a trajectory). We consider on T*(R 2 ) and r = to = 0}, the Hamiltonian &8? = =Fp1p2. The domains of plus- and minus-opposition of this Hamiltonian coincide: {pj, < 0 , q2 = 0 } . Therefore, there are no non-opposing points of S8 on the characteristic Pi = const < 0, and trajectories falling on such a characteristic cannot be extended. It would be interesting to clarify whether there is any physical
The structure of Hamiltonian mechanics
235
effect corresponding to this mathematical situation. In conclusion we point out that impulsive motion of systems can also be studied by means of the formalism we have constructed. For example, elastic impact of a rigid body at a fixed barrier without friction is described by reflection of the trajectory of the rigid body from a hypersurface T C T*(SO(3) X R 3 ). Here we must postulate that the Hamiltonian 30+ is "identically equal to °°", that is, that {SS+ =E} n r = 0 for any E. References
[1] R. L. Bishop and R. J. Crittenden, Geometry of manifolds, Pure and Appl. Math. 14, Academic Press, New York-London 1964. MR 29 #6401. Translation: Geomeiria mnogoobrazii, Mir, Moscow 1967. MR 35 #4833. [2] S. Sternberg, Lectures on differential geometry, Prentice—Hall, Engjewood Cliffs, N.J., 1964. MR 33 #1797. Translation: Lektsiipo differential'noigeometrii, Mir, Moscow 1970. MR 52 # 15255. [3] R. Narasimhan, Analysis on real and complex manifolds, North-Holland, Amsterdam 1968. MR 40 #4972. Translation: Analiz na deistviteVnykh i kompleksnykh mnogoobraziyakh, Mir, Moscow 1971. [4] M. F. Atiyah and I. G. Macdonald, Introduction to commutative algebra, Addison— Wesley, Reading, Mass. -London 1969. MR 39 #4129. Translation: Vvedenie v kommutativnyu algebru, Mir, Moscow 1972. MR 50 #2138. [5] Sze-tsen Hu, Homotopy theory, Pure and Appl. Math. 18, Academic Press, New YorkLondon 1959. MR 21 # 5186. Translation: Teoria gomotopii, Mir, Moscow 1964. [6] H. Goldschmidt and S. Sternberg, The Hamilton-Cartan formalism in the calculus of variations, Ann. Inst. Fourier (Grenoble) 23 (1973), 203-267. MR 49 # 6279. [7] N. E. Steenrod, The topology of fibre bundles, Princeton Mathematical Series 14, Princeton University Press, Princeton, N.J., 1951. MR 12-522. Translation: Topologia kosykh proizvedenii, Izdat, Inost. lit., Moscow 1953. [8] B. A. Kupershmidt, The Lagrangian formalism in the calculus of variations, Funktsional. Anal, i Prilozhen. 10:2 (1976), 77-78. = Functional Anal. Appl. 10 (1976). [9] J. Milnor, Morse theory, Ann. of Math. Studies 51, Princeton University Press, Princeton, N.J., 1963. MR 29 #634. Translation: Teoriya Morsa, Mir, Moscow 1965. MR 31 #6249. [10] V. V. Lychagin, A local classification of first order non-linear partial differential equations, Dokl. Akad. Nauk. SSSR 210 (1973), 525-528. MR 50 # 1295. = Soviet Math. Dokl. 14 (1973) 761-764. Local classification of non-linear first order partial differential equations, Uspekhi Mat. Nauk 30:1 (1975), 101-171. = Russian Math. Surveys 30:1 (1975), 105-175. [11] R. Thorn, Quelques proprietes globales des varietes differentiables, Comment. Math. Helv. 28 (1954), 17-86. MR 15-890. Translation: in the coll. "Fibre spaces and their applications", Izdat. Inost. lit, Moscow 1958.
236
A. M. Vinogradov and B. A. Kupershmidt
[12] A. Weinstein, Symplectic manifolds and their Lagrangian submanifolds, Advances in Math. 6 (1971) 329-346. MR 44 # 3 3 5 1 . [13] A. M. Vinogradov, Many-valued solutions, and a principle for the classification of nonlinear differential equations, Dokl. Akad. Nauk. SSSR 210 (1973), 11-14. MR 50 # 1 2 9 4 . = Soviet Math. Dokl. 14 (1973), 661-665. [14] F. Klein, Vorlesungen uber die Entwicklung der Mathematik im 19ten Jahrhundert, Springer-Verlag, Berlin 1926. Translation: Lektsii o razvitii matematiki v XIX stoletii, ONTI, Moscow—Leningrad 1937. [15] E. Fermi, Quantum mechanics, Univ. of Chicago Press, Chicago 1961. MR 24 # 1534. Translation: Kvantovaya mekhanika, Mir, Moscow 1968. MR 40 # 3 8 0 3 . [16] A. M. Vinogradov and I. S. Krasil'shchik, What is the Hamiltonian formalism? Uspekhi Math. Nauk. 30:1 (1975), 173-198. = Russian Math. Surveys 30:1 (1975), 177-202. [17] Yu. V. Ergorov, On the solubility of differential equations with simple characteristics, Uspekhi Mat. Nauk. 26:2 (1972), 183-199. MR 45 # 5 5 4 1 . = Russian Math. Surveys 26:2 (1971), 113-130. [18] A. M. Vinogradov, The algebra of logic of the theory of linear differential operators, Dokl. Akad. Nauk. SSSR 205 (1972), 1025-1028. MR 46 # 3498. = Soviet Math. Dokl. 13 (1972), 1057-1062. [19] L. Hormander, linear partial differential operators, Springer—Verlag, Berlin— Heidelberg-New York 1963. MR 28 # 4 2 2 1 . Translation: Lineinye differentsiaVnye operatory s chastnymi proizvodnymi, Mir, Moscow 1965. MR 37 # 6 5 9 5 . [20] V. P. Maslov, Teoriya vozmushchenii i asimptoticheskie metody (Perturbation theory and asymptotic methods), Izdat. Moskovsk. Gos. Univ., Moscow 1965. [21] L. A. Lyusternik and A. I. Fet, Variational problems on closed manifolds, Dokl. Akad. Nauk. SSSR 81 (1951), 17-18. MR 13-474. [22] J. Favard, Cours de geometrie diffe'rentiale locale, Gauthier-Villars, Paris 1957. MR 18-668. Translation: Kurs lokal'noi differentsiaUnoi geometrii, Mir, Moscow 1966. [23] A. A. Kirillov, Elementy teorii predstavlenii, Nauka, Moscow 1972. Translation: Elements de la the'orie des representations, Mir, Moscow 1974. MR 52 #14134. [24] A. Appel', Teoreticheskaya mekhanika (Theoretical mechanics), Izdat, Fizmatgiz, Moscow 1960. [25] L. S. Pontryagin, Nepreryvnye gruppy (third ed.), Nauka, Moscow 1973. MR 50 # 10141. Translation of second ed.: Continuous groups, Gordon and Breach, New York 1966. MR 34 # 1439. [26] S. Smale, Topology and mechanics, I, Invent. Math. 10 (1970), 305-331; II, ibid, 11 (1970), 4 5 - 6 4 . MR 46 # 8 2 6 3 , 4 7 # 9 6 7 1 . = Uspekhi Mat. Nauk 27:2 (1972), 7 7 - 1 3 3 . [27] Yu. I. Neimark and N. A. Fufaev, Dinamiki negolonomnykh sistem (The dynamics of non-holonomic systems), Nauka, Moscow 1967. [28] A. V. Vershik and L. D. Faddeev, Lagranzheva mekhanika v invariantnom izlozhenii (Lagrangian mechanics in an invariant presentation), in the coll. "Teoreticheskaya i matematicheskaya fizika" ("Theoretical and mathematical physics"), vol.11, Izdat. Leningrad. Gos. Univ., Leningrad 1974.
The structure of Hamiltonian mechanics
237
[29] F. A. Berezin, The Hamiltonian formalism in the general Lagrangian problem, Uspekhi Mat. Nauk. 29:3 (1974), 183-184. [30] I. M. GePfand, Spherical functions on symmetric Riemannian spaces, Dokl. Akad. Nauk. SSSR 70 (1950), 5 - 8 . MR 11-498. [31] R. Abraham, Foundations of mechanics, Benjamin, New York-Amsterdam 1967. MR 36 # 3 5 2 7 . [32] V. I. Arnol'd, Matematicheskie metody klassicheskoi mekhaniki (Mathematical methods of classical mechanics), Nauka, Moscow 1974. [33] C. Godbillon, Geometrie differentielle et mecanique analytique, Hermann, Paris 1969. MR 39 # 3 4 1 6 . Translation: Differentsial'naya geometriya i analiticheskaya mekhanika, Mir, Moscow 1973. [34] Dynamical Systems (Proc. Sympos. Univ. of Bahia, San Salvador 1971), Academic Press, New York-London 1973. [35] Dynamical Systems, Theory and Applications, Lecture Notes in Physics 38. MR 52 # 14236. [36] J. K. Moser, Lectures on Hamiltonian systems, Mem. Amer. Math. Soc. 81, Amer. Math. Soc, Providence, R.I., 1968. MR 37 # 6060. Translation: Lektsii o gamil'tonovykh sistemakh, Mir, Moscow 1973. [37] J. M. Souriau, Structure des systemes dynamiques, Dunod, Paris 1970. MR 41 # 4 8 6 6 . [38] Symposia Mathematica vol. XIV (Convegno di Geometria Simplettica), INDAN, Roma, January 1973. [39] A. Avez, A. lichnerowicz, and A. Diaz-Miranda, Sur l'algebre des automorphismes infinitesimaux d'une varie'te symplectique. J. Differential Geometry 9 (1974), 1—40. MR 50 # 8 6 0 2 . [40] W. B. Gordon, A theorem on the existence of periodic solutions to Hamiltonian systems with convex potentials, J. Differential Equations 10 (1971), 324-335. MR 45 # 1 2 0 0 . [41] S. M. Graff, On the conservation of hyperbolic invariantttori for Hamiltonian systems, J. Differential Equations 15 (1974), 1-69. MR 51 # 1 8 7 8 . [42] P. B. Guest, A non-existence theorem for realizations of semisimple Lie algebras, Nuovo Cimento A (10) 61 (1969), 593-604. MR 40 # 2 2 5 4 . [43] R. Hermann, Spectrum generating algebras and symmetries in mechanics, I, II, J. Math. Phys. 13 (1972), 833-842. MR 47 # 2930. [44] A. A. Kirillov, Local lie algebras, Uspekhi Mat. Nauk. 31:4 (1976), 5 7 - 7 6 . = Russian Math. Surveys 31:4 (1976), 5 5 - 7 6 . [45] A. lichnerowicz, Cohomologie 1-differentiate des algebres de lie attachees a une variete' symplectique ou de contact, J. Math. Pures Appl. 53 (1974), 4 5 9 - 4 8 3 . MR 51 # 4 3 1 5 . [46] E. Lacomba Zamora, Mechanical systems with symmetry on homogeneous spaces, Trans. Amer. Math. Soc. 185 (1973), 4 7 7 - 4 9 1 . MR 46 # 9 7 5 9 . [47] L. Losco, Integrabilite'en mecanique ce'leste, J. Mech. 13 (1974), 197-223. MR 51 # 1 4 1 5 4 . [48] S. V. Manakov, A remark on the integration of the Euler equations of the dynamics of a rigid body, Funktsional. Anal, i Prilozhen. 10:4 (1976), 9 3 - 9 4 . = Functional Anal. Appl. 10 (1976).
238
A. M. Vinogradov and B. A. Kupershmidt
[49] A. S. MIshchenko and A. T. Fomenko, On the integration of the Euler equations on semisimple lie algebras, Dokl. Akad. Nauk. SSSR 231 (1976), 536-538. = Soviet Math. Dokl. 17 (1976), 1591-1593. [50] M. A. Ol'shanetskii and A. M. Perelomov, Geodesic flows on symmetric spaces of zero curvature and explicit solutions of the generalized Calogero model for the classical case, Funktsional. Anal, i Prilozhen. 10:3 (1976), 86-87. = Functional Anal. Appl. 10 (1976). [51] R. M. Santilli, Dissipative and lie admissible algebras, Meccanica 4 (1963), 3-11 (Italian). [52] A. Weinstein, Lagrangian submanifolds and Hamiltonian systems, Ann. of Math. (2) 98 (1973), 377-410. MR 48 #9761. [53] M. Verge, La structure de Poisson sur Palgebre symmetrique d'une algebre de Lie nilpotente, Bull. Soc. Math. France 100 (1972), 301-335. MR 52 #657. [54] P. A. M. Dirac, Lectures on quantum mechanics Yeshiva Univ., New York 1964. Translation: Lektsiipo kvantovoi mekhaniki, Mir, Moscow 1973. [55] L. M. C. Coelho de Souza and P. R. Rodrigues, Field theory with higher derivativesHamiltonian structure, J. Phys. A, Ser.2, 2 (1962), 304-310. [56] M. Flato, A. lichnerowicz, and D. Sternheimer, Deformations of Poisson brackets, Dirac brackets and applications, J. Math. Phys. 17 (1976), 1754-1762. [57] P. L. Garcia, Symplectic geometry in the classical theory of fields, Collect. Math. 19 (1968), 73-134 (Spanish). MR 39 # 5089. [58] B. Kostant, Quantization and unitary representations, Lecture Notes in Math. 170. MR 45 #3638. [59] B. A. Kupershmidt, Geometriya mnogoobraziya dzhetov i struktura lagranzheva i gamiVtonova formalizmov (The geometry of a jet manifold and the structure of the Lagrangian and Hamiltonian formalisms), Trudy Moscov. Mat. Obshch., Moscow 1977. [60] L. D. Faddeev, The Feynman integral for singular Lagrangians, Teoret. Mat. Fiz. 1 (1969), 3-18. [61 ] E. Shmutser, Osnovnye printsipy klassicheskoi mekhaniki i klassicheskoi teorii polya (The basic principles of classical mechanics and classical field theory), Mir, Moscow 1976. [62] O. I. Bogoyavlenskii and S. P. Novikov, On a connection between the Hamiltonian formalisms of stationary and non-stationary problems, Funktsional. Anal. i. Prilozhen. 10 (1976), 9-13. = Functional Anal. Appl. 10 (1976). [63] L. J. F. Broer and J. A. Kobussen, Canonical transformations and generating functional, Physica61 (1972), 275-288. [64] A. M. Vinogradov, I. S. KrasiTshchik, and V. V. Lychagin, Primenenie nelineinykh differentsiaVnykh uravneniu I (Applications of non-linear differential equations. I), MIIGA, Moscow 1977. [65] A. M. Vinogradov, I. S. KrasiTshchik, and V. V. Lychagin, Geometriya prostranstv dzhetov i nelineinye differentsial'nye uravneniya (The geometry of jet spaces and nonlinear differential equations), MIEM, Moscow 1977. [66] C. S. Gardner, Korteweg-de Vries equation and generalizations. IV. The Korteweg-de Vries equation as a Hamiltonian system. J. Mathematical Phys. 12 (1971), 1548-1551. MR 44 #3615.
The structure of Hamiltonian mechanics
239
[67] I. M. Gel'fand and L. A. Dikii, Asymptotic behaviour of the resolvent of SturmIiouville equations and the algebra of the Korteweg-de Vries equations, Uspekhi Math. Nauk. 30:5 (1975), 67-100. = Russian Math. Surveys 30:5 (1975), 77-113. [68] I. M. Gel'fand and L A. Dikii, Fractional powers of operators and Hamiltonian systems, Funktsional. Anal, i Prilozhen. 10:4 (1976), 13-29. = Functional Anal. Appl. 10 (1976). [69] I. M. Gel'fand, Yu. I. Manin, and M. A. Shubin, Poisson brackets and the kernel of the variational derivative in the formal calculus of variations, Funktsional. Anal, i Prilozhen 10:4 (1976), 30-34. = Functional Anal. Appl. 10 (1976). [70] B. A. Dubrovin, V. B. Matveev, and S. P. Novikov, Non-linear equations of Korteweg-de Vries type, finite-zone linear operators, and Abelian varieties, Uspekhi Mat. Nauk. 31:1(1976), 55-136. = Russian Math. Surveys 31:1 (1976), 59-146. [71 ] V. E. Zakharov and L. D. Faddeev, The Korteweg-de Vries equation is a fully integrable Hamiltonian system, Funktsional. Anal, i Prilozhen. 5:4 (1971), 18-27. MR 46 #2270. = Functional Anal. Appl. 5 (1971), 280-287. [72] B. A. Kupershmidt and Yu. I. Manin, Equations of long waves with a free surface, conservation laws and solutions, Funktsional. Anal, i Prilozhen. 11:3 (1977), = Functional Anal. Appl. 11 (1977). [73] N. V. Nikolenko, On the complete integrability of the non-linear Schrodinger equation, Funktsional. Anal, i Prilozhen. 10:3 (1976), 55-69. = Functional. Anal. Appl. 10 (1976). [74] Y. Watanabe, Note on gauge in variance and conservation laws for a class of non-linear partial differential equations, J. Mathematical Phys. 15 (1974), 453-457. MR 49 #5588. Received by the Editors 20 October 1975 Translated by A. J. Mclsaac
WHAT IS THE HAMILTONIAN FORMALISM? A. M. Vinogradov and I. S. Krasil'shchik In this paper the basic concepts of the classical Hamiltonian formalism are translated into algebraic language. We treat the Hamiltonian formalism as a constituent part of the general theory of linear differential operators on commutative rings with identity. We take particular care in motivating the concepts we introduce. As an illustration of the theory presented here, we examine the Hamiltonian formalism in Lie algebras. We conclude by presenting a version of the "orbit method" in the theory of representations of Lie groups, which is a natural corollary of our view of the Hamiltonian formalism.
"The present essay does not pretend to treat fully of this extensive subject, - a task which may require the labours of many years and many minds; but only to suggest the thought and propose the path to others." (From the introduction to "On a general method in dynamics" by W. R. Hamilton.)
A curious fact that has come to light in the course of the last ten or fifteen years is the appearance of the Hamiltonian formalism of classical mechanics as an essential component of a number of mathematical theories, far removed from mechanics as such. Examples of this kind are the orbit method in the theory of representations of Lie groups (see [10] and [11]), the theory of local solubility of linear differential operators (see the survey [8]), the theory of a canonical operator (see [13] and [14]), and others. It is of interest that the Hamiltonian formalism plays different functional roles in these theories; this suggests that the Hamiltonian formalism is part of a broader mathematical system. We believe that this is the "general theory of linear differential operators" in the sense of [2], and that this circumstance, although demonstrated in the corresponding specific investigations completely explains the many-sidedness of the Hamiltonian formalism and allows us to hope for a deeper understanding of the possibilities of Hamiltonian method, as well as for new domains of its applicability. Our paper is concerned with the necessary first step in this direction: translating the basic concepts of the Hamiltonian formalism into its natural algebraic language (here we arrive at some new "mechanical" invariants of commutative rings). 241
242
A. M. Vinogradov and I. S. Krasil'shchik
The new facts, if they are new, are simple and illustrative. We present our exposition in two versions, "general" and "Lie", mainly to show the degree of unification our approach yields. Throughout we try to motivate the concepts we introduce. For an understanding of many of them, apart from the "standard" facts, an acquaintance with the elementary parts of the theory of affine schemes (see [12], for example) is needed. The authors are very grateful to V. I. Arnol'd, who read this work in manuscript, and to the referee for many useful remarks. 1. Linear differential operators over a commutative ring K with identity. We introduce here the necessary information from the theory of linear differential operators over a ring K. An outline of this theory, which will help to understand the role of the "Hamiltonian formalism", can be found in [2] and [3]. For surveys of the theory of differential operators from the point of view of the theory of schemes, see [4] and [7]. Let A and B be AT-modules and let A: A -*B be an additive mapping. For each element k G K we define the mappings kA9 Jc+A and 8k (A) by putting (kA)(a) = kA(a), (k+ A)(a) = A(ka), a G A, and 5fc (A) = k+A - kA. We also set 5 ^ , ^ , .. .,k = 8ko o 5 FUNDAMENTAL DEFINITION. An*additive mapping A.A-+B is called a linear differential operator of order < s over K if for each choice of k0, kx, . . ., ks G K we have 6
V V
...,
fc#(A)
=0.
If the ring K has additional structure, we can consider differential operators that respect this structure. If, for example, A' is a C-algebra, we can speak of C-linear differential operators over K. When K - C°°(M) is the ring of all smooth (real or complex) functions on a smooth manifold M, and A and B are A'-modules of smooth sections of vector bundles over M, then our concept of an R- or C-linear differential operator coincides with the classical one (see [15]). The set of differential operators of order < s from A to B is an Abelian group under the natural addition. We can make this Abelian group into a AT-module in two different ways. The first is by the multiplication (k, A) -* kA, and the second by (k, A)-* k+A. We denote the corresponding modules by Diffs(A,B) and Dif£C4,£) and we write Diff(+>04, B) for the bimodule obtained by considering both structures simultaneously. It is clear that
Diff%)(A,B)CDiff(f)(A,B)
when m<s, so that we arrive at the
bimodule Diff <+) (A, B) = U Diff£+)G4, B), filtered by the bimodules s>0 DiffJ+)G4,5), s>0. Let A, B9 and C be ^-modules, Aj G Diff£+)(£, Q and A2 G Diff,(+)C4,£).
What is the Hamiltonian formalism?
243
The obvious relation 5 J t (A l o A 2 ) = ^ ( A j ) ° A2 + Aj o 8k(A2), kGK, shows that composition of differential operators defines a multiplication Diff <+) W, B) ®,Diff i+> (5, O -*Diff <%U, C). In particular, Diff *** (A, A) is a filtered ring. We consider the ring Diff <+> K = Diff<+> (£, JO. The graded ring Smbl, K = S Smbls ff ^ 0
+)
corresponding to it, where Smbls K = Diff< i^/DiffJ^ #, is called the rmg o/ symbols; it is obviously commutative. For each ^"-module A we introduce a module DiffJ>1 and a differential operator R: DiffJ ^4 ->^4of order < s such that for any ^-module B the correspondence / - > # % / £ Hoi%(2?, Diff*^4) is an isomorphism of the modules Horn* (B, Diff J A) and DiffJ (5, ^4). It is easy to see that by setting Diff* A - Diff* (#,,4), and fl(A) = A(l), A 6 DiffJ ^4, 1 6 #> we obtain a realization of the functor A -* Diff*^4. 2. The functors D^- and A*. The reader will see below that, perhaps somewhat unexpectedly, the Hamiltonian formalism cannot, in general (that is, for an arbitrary ring K), be stated in the language of differential forms. In this section we provide enough information on the calculus of differential forms over a ring K to explain this circumstance. When K = C°° (M, R or C), the concepts introduced here coincide with the classical ones. We recall that when K = C°°(M, R), each smooth vector field X on M induces a derivation dx: K-> K, that is, an additive mapping
such that dx (fif2)= f\ dx(f2)+ fi dx(f\)J\Ji
e
K, and vice versa.
Since the concept of a derivation is defined for any ring K, we naturally use the phrase "vector field on A"" to mean a derivation of K. We write D(K) for the AT-module of derivations of K. Now we recall that a derivation of K with values in a ^-module A is an additive mapping A: K-+A such that A(klk2) = k1A(k2) + k2A(kx). We write D(A) for the ^-module of all such derivations. It is obvious that each of these derivations is a differential operator of order < 1 and that there is an embedding D(A) CDiff* A), which is a first order differential operator. Let i? be a subset of a module A. We write D(R CA) for the set of all derivations A: K-+A such that A(K)CR; we define Diff^+)(i?c:^) similarly. In particular D(R C A) = D(R) if R is a submodule of A. We now define a sequence of functors Dt and embeddings Df(A) C (Diff*)'O4) inductively. We put Do = id, Dx = D, Dt(A) = D{DU1(A) c= (Difff)*-1^)}. The composition
244
A. M. Vinogradov and I. S. Krasil'shchik
Dt(A) = D{Di_t(A) cz jr f j of the natural embeddings gives an embedding D;(A) C (Diff j)z(>l), which is a differential operator of order < 1. The embeddings at(A): Dt(A) cz D^CDiffJ" A) are constructed inductively from the natural embedding oti(A): D(A) C Diff j 04) and can be shown to be differential operators of order < 1. Since ^ A)} = = 0{Z>f we see that Dt(A) = D{DUA)
cz (Difff)*
can evidently be identified with D{lm a M U ) cz (Difffl^W)} cz Z>{Z)<_2(Difft4) cz which defines oij(A). We remark that the functor embedding Dt(Dj) C Z)/+/- can be established directly by induction. The module At = At(K) (i = 0, 1, 2, . . .) of /-dimensional differential forms over A^ is defined as a representing object for the functor Dt; that is, so that for every ^"-module A there is an isomorphism Dt(A) = HomK (A£, .4). As we know, A = A1 is the factor module of the free ^-module generated by elements dk (k G k) subject to the relations d(k1k2) = k1d(k2) + k2 dkx,d{kv +k2) = dkx +dk2,kt,k2 EK. It can be shown that the module At exists and is the /th exterior power of A. In this context we mention that the homomorphisms Ar 0 A,- -> Ai+j of representing objects generated by the functor embedding Dt(Dj) C Di+J is nothing but exterior multiplication of differential forms. We note that every differential operator A: A ^ -+At of order < 1 determines for every AT-module A a mapping (PAG4): Homji: (A£, A) -> HomK (A M , Difft ^4) by the formula
Horn (A*, A) — - » Hom x (A M j IMI+A)
IIII
a.(A) l D
||
Dt(A) commutes. In particular, for / = 1 the operator dx is the derivation of K with values in A, and d(k) = dk,kE K, and the isomorphism D(A) = Homk (A, A) associates with each derivation AGD(A) a homomorphism hA : A -+ A so that the diagram
What is the Hamiltonian formalism?
245
commutes. The operator d defined here can be shown to have many of the properties of the classical exterior differentiation of forms. For example, d2 = 0 and d(cox /\ co2) = dcoi f\ co2 + (—I)1©! A d(o2, (Oi 6 Af, co2 £ A7-. The representation of the module A* as the zth exterior power of A allows us to regard the elements w G A j as /-linear skew-symmetric functions on D(K), which corresponds to the classical treatment of a differential form on a smooth manifold. The classical formula for the exterior differential: i+1
(d©) (x l f ..., x <+1 ) = S (-1) 73 " 1 xP«> (Xu ..., x p _ t , x p + 1 , ..., x l+ 1 ) + P=I
+
ZJ ( — 1 ) P + 9 ( 0 ( [ ^ P T
X g ] , Xu
. . . , Xp_1? X p + 1 , . . . , Xg_i, X g + 1 ,
...,Xi+1),
where Xj, . . ., Xi+1 6 D(K), co 6 Af, remains valid. It is important to realize that a form co 6 At(K) need not, in general, be uniquely determined by its values on vector fields; that is, it can happen that u(Xx, . . ., Xt) = 0 for all choices of Xx, . . ., XjGDiK), and yet co =£ 0; there is an example of this on p. 0G0. This cannot happen when the module A = A(K) is projective and locally finite, for example, when K = C°°{M, R or C). In 4. we shall use the fact that if a 3-form is determined by its values on elements of D(K), then to prove that co is closed we need only verify that it vanishes on the expression above. 3. The cotangent (phase) space of a commutative ring K. We recall that the arena of action of the Hamiltonian formalism of classical mechanics, the phase space, is the cotangent space T*(M) the manifold M of configurations. The algebraic theory of differential operators, which forms the foundation of our constructions, is based on the transition from smooth manifolds to commutative rings. Here M is replaced by C°°(M, R) or C*>(M, C). REMARK. The theory of affine schemes (see [12], for example) permits us to recover the manifold M from C°°(M, R) or C°°(M, C). But by now it is clear that the ring view-point is preferable, because two distinct rings correspond to one and the same manifold. We must, therefore, find out, purely in terms of the algebraic structure of K, what this T*(M) should be, more accurately, what ring "corresponds" to T*(M). Answer: the ring "corresponding" to T*(M) is Smbl^AT), the ring of
246
A. M. Vinogradov and I. S. Krasil'shchik
symbols of K. What is the basis for this statement? Let us consider the classical case K = C°°(M, C). Then, as is well known, the ring Smbl* (K) can be identified naturally with the ring of those smooth functions on T*(M) that are polynomial along the fibres of the projection jt: T*(M) ->M. The "manifold" Specm {Smbl^ZlX "bad" points (we do not define "bad" precisely) is then the complexification of the fibering 7r, so that T*(M) a SpecjSmbl^iO. Further, by what we have said above, each element of Smbl^K, regarded as a function on SpeclSmbl^Z}, in accordance with the theory of affine schemes, can be uniquely recovered from its restriction to the subspace T*(M) cz SpecfSmb]^}. Thus, the points of the complement Spec{Smbl # iir}\ T*{M) are inessential from the point of view of the "geometrical realization" of the ring S m b l ^ , that is, its representation as a ring of "functions on something". This shows why T*(M) can be regarded as the geometrical equivalent of Sfnbl^AT. NOTE 1. In quite a number of problems (the complex germs of Maslov [13], for example) it is useful, and even necessary, to consider the "complex" points of SpeclSmbl^Z). NOTE 2. What we have said above can be understood as our justification for taking SpecjSmbl^} as the cotangent space of the "manifold" Spec K. Later we shall write briefly T*(K) instead of Spec {Smbl,,K} (on this point see also [6]). 4. The Hamiltonian mapping. Let A G Diffm K and let us write smblm (A) for the image of A under the quotient homomorphism Diffm K -» Smblm K. If A,- G Diffm (K) (/ = 1, 2), then [Aj, A2 ] = AjA 2 — A 2 Aj G Diffm +m _x K. This is an immediate consequence of the equality 6fe(A1A2) = 6fe(Aj) -A2 + Ax-6fe(A2), which was established in 1. If st = smblm .(Az) (/ = 1, 2), then it follows from what we have said that the element smbl mi + m^ _l[Al,A2], depends only on 5 / (/= 1, 2). We call this element the Poisson bracket of st and s2 and denote it by {sx, s2) 6 Smbl mi+m2 _ 1 K. The following properties (analogous to those of the classical Poisson bracket) can be derived immediately from the definition: 1) Ui + s2, s} = to, s} + {s2, s}, 2) {%, s2} = — {s2, ^ j , 3) {s, s^} = s1-{s, s2} + s2-{s, s x }, 4) {Sj, {S2, S8}} + {52, {53, Si)} + {S8, {«!, S2}} = 0. As 3) shows, the mapping Hs0: s h-+Hso(s) = {s0, s} for fixed s0 ES= Smbl* K is a derivation of S. Since for us the concepts "derivation of 5"' and "vector field on 5 " are synonyms, we can regard Hs0 as a "vector field" on T*(K); more precisely, Hs0 is the "Hamiltonian vector field" on T*(K) corresponding to the "Hamiltonian" sQ. As 3) shows, the mapping H. S -+D(S), s H-> HS is a derivation of S with values in
What is the Hamiltonian formalism ?
247
the S-module D(S). There is, therefore, an S-homomorphism h: A(S) +->/)(£), such that the diagram S —*-J)(S)
XA A(S) commutes. We note that specifying the derivation H is equivalent to specifying the homomorphism h; the Poisson bracket can, therefore, be expressed in terms of h. Indeed, {su s2} = (h dsx, ds2), where the brackets <,> denote the natural pairing of A(*S') and D(S) induced by the isomorphism D(S) — Homs (A(£), S). Instead of h we can consider equivalently the bilinear form Qh(Ki K) = (MM» ^2) o n MS). The form £lh is skew-symmetric because so is the Poisson bracket. When h is an isomorphism, we can consider the skew form £lh(Dx, D2) = ^C/z^COi), h~l(D2)) on D(S). If Qh 6 A2(S), it is natural to interpret it as a differential 2-form on SpecS =T*{K). PROPOSITION 1. / / the form £lh exists and is defined by its values on D(S), then it is closed, that is, d£lh = 0. PROOF. We need only prove that for any three derivations £>;££>(£) 0 = 1 , 2 , 3 ) (1) dQh(D^ D» D3) = D&h{D* D3) - D.Q^D,, D3) + + D3Qh(Du D2) - Qh(lDl9 D2], D3) + + QMD^ D3l D2) - Qh([D2, D8], Dx) = 0 (see 2). Since A(S) as a module over S is generated by the elements of the form ds, and since h in our case is an isomorphism, the module D(S) is generated by derivations of the form Hs with the same relations. Therefore, since (1) is linear in all three arguments, we need verify (1) only for Dt = Hst ( / = 1 , 2, 3). But then dQh(HSl, Hs2, Hs3) = HSli2h(Hs2, + Hs3Qh(Hsv
Hs3) - Hs9Qh(Hst,
Hs3) +
Hs2) - Qh(lHSl, Hs2], Hs3) +
+ Qh(lHSl, Hs3], Hs2) - Qh([Hs2, Hs3], Hs%) =
= 2({*i, fe, s3}} + {53, fo, s2}} + {s2 {s3, sx}}) = 0 by the Jacobi identity for the Poisson brackets. In the classical case K = C°°(M, R) we have llh = 2 dpi A dqt in the i
corresponding coordinates on T*(M). So we see that when h is an isomorphism, £lh is a 'geometrical realization" of h (or of H, or of the Poisson bracket). Thus, the mapping H (or h) is a more fundamental object than £lh. Thus, although according to the conventional point of view the Hamiltonian (= canonical) structure is interpreted as a 2-form on T*(M)
248
A. M. Vinogradov and I. S. Krasil'shchik
(which in this case is equivalent to specifying H or h), in the general situation we must understand by the Hamiltonian (or canonical) structure one of the three objects: the operator //, the homomorphism h, or the form £lh. It stands to reason that each of these three objects must satisfy the relevant conditions. For example, the operator H must be a derivation from S into D(S) for which Hs1(s2) +Hs2(s1) = 0. From now on we mean by a canonical structure on a ring S (which is not, in general, the ring of symbols Smbl+ K for some ring K) a derivation H: S -+ D(S) satisfying the condition just stated. The pair (S,H) is then called a canonical ring. We note that when M is a canonical (= symplectic) manifold, then there is a natural canonical structure on the ring of functions C°°(M, R) or C°°(M, C). Thus, the concept of a canonical ring is a generalization of that of a symplectic manifold. We remark again that the equation Hsi(s2) = {su s2} defines a "Poisson bracket" on the canonical ring S and that this in turn determines the canonical structure H: S-*D{S) uniquely. The concept of a canonical ring (canonical manifold), being structural, taken by itself is void of a "functional" meaning. However, the study of canonical rings is both justified and necessary, because they arise naturally as factor rings and subrings of "functional-valued" rings S of the form Smbl* K. For example, the 2-dimensional sphere S2, equipped with its standard canonical structure, is a factor object of a Hamiltonian formalism in the Lie algebra of the rotations of 3-dimensional space (see 11) and this, in turn, is a "left-invariant subobject" of a Hamiltonian formalism in T*(G), where G is the group of rotations of 3-dimensional space. NOTE. In classical mechanics the fundamental form Q ;= 2 dpi /\ dqt is exact and arises purely geometrically as the differential i
of the canonical 1-form p = 2 Pt dqt on the phase space T*(M); and this i
1-form is determined naturally by the following universal property: s^(p) = w , where GO is an arbitrary 1-form on M and s^ is the section corresponding to co of the cotangent fibration JI: T*(M) -> M. The following arguments show that in the general situation this construction is not possible. Suppose that K is an algebra over the field of rational numbers and let © = dfr £ At(K)t k£ K. Further, let o = smblm A, A G Diffm K. We define
m times
The resulting mapping s^ : Smbl* K-* K is a ring homomorphism and in the case K - C°°(M, R or C) corresponds to the homomorphism
What is the Hamiltonian formalism?
249
si:C00{T:¥{M))^C00{M), generated by the section s w . Therefore, the homomorphism s^ is an algebraic equivalent to the concept of a section of T*(M). However, for an arbitrary ring K the existence of a mapping of the type s^ does not, in general, enable us to construct a form p. The following example illustrates this. Let K be the R-algebra of all continuous functions on the line R 1 that are once differentiable at 0. We denote by jufl C K, a E R, the maximal ideal consisting of the functions that vanish at a. Clearly, A^K) =£ (0), therefore there are derivations of K with values in K/n0. On the other hand, the modules D(K/ixa) for all a¥=0 are trivial (because Ma = M
Aitf) - * Z ) ( S ) Ai(5)
commutes, that is, if h = Dqp.oho Acp. Here we understand by Acp and Dy the ^-module morphisms: Acp(s ds') =
The morphisms Aq> and Dq> are obviously related with respect to the natural pairing of At(S) with D(S), in the following sense: (3)
{d, A
250
AM
Vinogradov and LS. KrasiVshchik
PROPOSITION 2. The diagram (2) commutes if and only if £lh is invariant under cp. PROOF.
Q^Acpftj), Aq>(A,8)) = (fcAq)^), Aq>(Xa)> = = q> (Dyoho Acp(Xa), X2> = cp (M^i), ^2> =
PROPOSITION 4. A derivation i|? is canonical if and only if it satisfies the following two conditions: 2)
The proof is obvious. We introduce some more notation. As above, let K be a commutative ring with identity and S = Smbl* K its ring of symbols. We write Ham S for the set of Hamiltonian vector fields on T*(K) (that is, derivations of S of the form Hs, s G S), and Can S for the set of canonical vector fields on T*(K) (that is, canonical derivations of S). Condition 2) of Proposition 4, applied to an arbitrary element s ' G 5 indicates that if {s> s'} = f\J?s, s'} + {s, tys'}, and is, therefore, satisfied for all derivations i|> = /Zs", s" 6 S (in this case it is simply the Jacobi identity
What is the Hamiltonian formalism?
251
for Poisson brackets). Thus, we have the embedding Ham S C Can S (as sets). Further, the modules Ham S and Can S are Lie rings (and when K is provided with the structure of an A -algebra they are Lie algebras over A). Condition 2) of Proposition 4 shows that Ham S is an ideal in Can S. In the classical case K = C°°(M, R) the "fundamental theorem of mechanics" asserts that Ham S = Can S locally (and globally, if M is simply-connected). This is not so, in general, as is shown by examples below.1 Thus, we must understand by the "fundamental theorem of mechanics" for a ring K a theorem describing the structure of the Lie ring (A -algebra) Can iS/Ham S, that is, describing by how much Can S is "bigger" than Ham S). As in 5., the concepts introduced here correspond in the classical case completely to the usual ones. 7. Localization. Definition 1 of 5. admits of some generalization. We say that a ring homomorphism cp: St -> S2 of canonical rings is a canonical homomorphism if it preserves the Poisson brackets, that is, if
We mention that in the case of an arbitrary canonical manifold M, since the fundamental form is nondegenerate, the set h~x (Can S) coincides with the set of all closed 1-forms on Af, and h'1 (Ham S) with the set of all exact 1-forms. Therefore, Can S/Ham S = H1 (M, R).
252
A.M. Vinogradov and I.S. Krasil'shchik
ideal of S contained in the ideal p. Intuitively, we can interpret SlO(p) as the "ring of functions on pH". PROPOSITION 6. For every prime ideal let p C S the ideal O{p) exists, is uniquely determined, and is prime. PROOF. The first two assertions are obvious. Let us only prove that &{p) is prime. (a) If O(p) — /?, there is nothing to prove. (b) We write q for O(p). Suppose that q is a proper subset of p. First we show that q is its own radical. For the set R(q) of elements s E S with sn E q for some n is a stable ideal that contains q and is contained in p (because p is prime). Thus, by the maximality of q, R(q) = q. (c) We consider an element s E S/q and the set qs of elements s' E S such that s"sr E q for some n. It is clear that qs is an ideal and qs D q. The identity Ha(s'ysn+1 = Ha(s'sn+1)-Ha(sn+lysf = Ha(s'sn+l)-(n+\)sns'Ha(s) n+l shows that Ha(s')s G q if s' G qs, o G S, that is, that qs is stable. Note also that qs C p if s £ p, because p is prime. (d) Let s, sf G S with 5 s' E qr. We assume that one of s and 5', say 5, does not lie in q. We consider the ideal qs. If there is an element s" E qs with 5" $ p, then q C q^> C p and so gs» = #. Since sns" E g, we see that sn E 4S» and so sn G q. As a consequence of (b) we have 5 E q9 which contradicts our assumptions. Thus, q ^ qs C /?, therefore q - qs. And since 5' E gs we have 5' E #. COROLLARY 1.^4 Hamiltonian structure on S defines a mapping O: Spec S ->Spec S under which every ideal p C S is associated with O(p). This mapping O is a projection of the space Spec S onto the subspace Spec# S consisting of all stable prime ideals of S. COROLLARY 2. The set of algebraic orbits in Spec S is in one-to-one correspondence with the set of stable prime ideals of S, that is, Spec# S is the "orbit space". Even when the geometrical concept of the Hamiltonian orbit pH of a point p E Spec S, which we have discussed above, can be given a meaning, it can happen that nevertheless pH ¥= Spec S/Q(p) (obviously, we always have pH C Spec S/0(p)). This has to do with the fact that Spec S/O(p) is only the "algebraic closure" of pH and therefore can be considerably larger than pH. In that case the algebraic orbit Spec S/O{p) contains some set of geometrical orbits of the form qH, q E Spec S. Simple examples illustrating this behaviour are provided by the Hamiltonian formalism in Lie algebras (see 15.). The following definition corresponds intuitively to the operation of excising the geometrical orbits qH of non-maximal dimension from the algebraic orbit Spec SlO(p). DEFINITION 4. Let p G Spec S be a prime ideal of S. The principal part of the algebraic orbit of p is defined as the set
What is the Hamiltonian formalism ?
25 3
o(p) = oPROPOSITION 7. 1) The partition of Spec S into the principal parts of algebraic orbits is a partition into disjoint classes. 2) The algebraic orbit Spec S/q coincides with its principal part if and only if q is a maximal stable ideal. The proof is obvious. The Poisson centre Z(S) of a canonical ring is the set of those s € S for which Hs = 0. Obviously, Z(S) is a subring of S. The following proposition shows that the Poisson centre of S/q is algebraically irreducible when q is a maximal stable ideal. PROPOSITION 8. If q is a maximal stable ideal, then the Poisson centre of S/q (that is, the "ring of functions" on the orbit corresponding to q) is a field. PROOF. The ring S/q is endowed with a Hamiltonian structure and is characterized by the fact that it has no proper stable ideals. Clearly, Z(S/q) ^ = 0 . Let z £ Z(S/q). Then the principal ideal (z) is stable; consequently, it coincides with the whole ring S/q. Thus, z is invertible in S/q; evidently, its inverse z"1 belongs to Z(S/q). Thus, with an algebraic orbit of the indicated type there is connected a field Z(S/q), an extension of the base ring K. It is natural to call this the field of infinitesimal characters of the algebraic orbit Spec S/(3(q). We end this section by remarking that the concepts introduced here can be taken as the basis of a theory of "Hamiltonian affine schemes", an algebraic-geometric analogue to the concept of a canonical manifold. 8. Lagrangian manifolds and the Hamilton-Jacobi equation. It is known
that every Lagrangian manifold L C T*(M) lying on the level surface S£ = 0 of the Hamiltonian function is invariant under the corresponding Hamiltonian vector field. This absorption principle, translated into the algebraic language, leads to the following definition: an ideal / C S is called autostable if it is invariant under all operators Hs, s £ / , and a "maximal autostable ideal" is Lagrangian. The submanifold Lj C T*(K) corresponding to a Lagrangian ideal is also said to be Lagrangian. We mention that in the classical case K = C°°(M, R) the algebraic concept introduced above of a Lagrangian ideal corresponds to the geometric concept of a connected Lagrangian manifold in T*(M). Thus, the fact that some (connected) Lagrangian manifold lies on the level surface 3£ = 0 can be regarded as "geometrical realization" of the situation where the principal ideal of S generated by Si 6 S is contained in the corresponding Lagrangian ideal. Below we use this interpretation to determine the algebraic meaning of the Hamilton-Jacobi partial differential equation. As a preliminary we have to give a geometrical meaning to the Hamilton-Jacobi equation in the classical context:
254
AM. Vinogradov and I.S. Krasil'shchik
Here g l 5 . . . , qn are local coordinates in U = M, and (g, p) are the naturally arising canonical coordinates in 7T"1 (C/) C T*(M), where n: T*(M) -* A/ is the natural projection. With this equation we can associate a function H(q, p) on r*(Af). Then the fact that a function / G C~(M) satisfies (4) means that the manifold L = L(f) C T*(M) defined by the equations Pi = j-m (i = 1, . . ., n)
in ir"l(U) C T*(M)
lies on the level surface
H(q, p) = 0. Submanifolds of the form L(f) are uniquely characterized by the following two properties: 1) L(f) is Lagrangian; 2) 7r| I( ^is a diffeomorphism. Therefore, the differentials of solutions of (4) are in one-to-one correspondence with those Lagrangian manifolds La {H = c} for which ir\L is a diffeomorphism. This geometrical interpretation of the concept of a solution of the Hamilton-Jacobi equation is not invariant under canonical transformations of the natural canonical structure on T*(M). The obstruction is the requirement that TT\L be a diffeomorphism. Discarding it we come to regard as a solution of (4) (more accurately, its differential) any Lagrangian manifold L lying entirely on the level surface H = 0. NOTE. There are a number of other weighty reasons, both of a physical and mathematical character for using our general concept of a solution of the Hamilton-Jacobi equation. This geometrical formulation, in turn, has the following "algebraization" by virtue of the indicated algebraic meaning of the inclusion L c {# = 0}: a solution of the "Hamilton-Jacobi" equation $8 = 0, where 3£ t- S, is a Lagrangian ideal J C S containing the principal ideal {$£) . Let us call the problem of finding all Lagrangian ideals J C S containing a given principal ideal (36) the "Hamilton-Jacobi problem". Then in the classical case K - C°°(M, R) the Hamilton-Jacobi problem is that of finding all (generalized) solutions of the Hamilton-Jacobi equation. In view of this it is natural to regard the Hamilton-Jacobi equation as the "geometrical" way of expressing the Hamilton-Jacobi problem. We end this section by indicating the algebraic meaning of an "ordinary" solution of the Hamilton-Jacobi equation; we restrict ourselves, for simplicity, to rings K whose additive group is divisible. With this aim we draw attention to the following method of constructing a manifold L(f). Let Xf be the Hamiltonian field on T*(M) corresponding to the Hamiltonian ir*(f), where TT: T*(M) -• M is the natural projection. We write At: T*(M) -• T*(M) for the shift diffeomorphism along the trajectories of Xf. Then, in the coordinates considered above, At(q, p) = (q, p — t j - ) . In particular, Ax(L(f)) = L(0),
What is the Hamiltoniam formalism?
255
and L(0) is the zero section of the vector bundle n: T*(M) -* M. Since Xf has global trajectories, we represent the automorphism A*:
C~(T*(M), R) -> C°°(T*(M), R) in the form A? = etXf, where
£tz/ = 2
f~ a n d ^7 is interpreted as the derivation operator
Xf: C°°{T*{M\ R) -> C°°(r*(M), R), or in "algebraic" notation, A* = '#<"*/). Therefore, the operator ^ = £>-#<"*/) carries the Lagrangian ideal / 0 corresponding to Z(0) into the Lagrangian ideal /^ corresponding to /. Thus, the description of Jj reduces to that of Jo. In turn, clearly / = 2 Smblj K. Thus, in general, we have defined Jf as the inverse J
i>0
image of the ideal 2 Smbl^ K czS
under the automorphism e~HJ^, where
i>0
/ G A" and j: K -* Smbl* A^ is the natural embedding. We remark here that e-Hj(f)(s) = S ( ~ 1 } 'fi> { s ) for any 5 G Smblm K since i / j ^ (5) = 0 for i > m, so that the operator e~Hi^ has a meaning for any ring whose additive group is divisible. Thus, an element / E K is an "ordinary" solution of the Hamilton-Jacobi equation if the Lagrangian ideal Jf contains the ideal (3£). NOTE. What we have said in this section demonstrates the purely algebraic nature of the Hamilton-Jacobi problem. However, we have not touched on the vastly more important functorial aspect of this problem, that is, "why" it needs to be solved. To do this we would have to develop the general theory of differential operators on commutative rings in vastly greater volume than we can allow in this paper. A partial functorial meaning of the formulation of the Hamilton-Jacobi problem can be extracted from [2] and [3]. 9. An example. The cotangent spaces of cyclic groups! We consider the cyclic group Zn of order n, a field Fp of characteristic p, p a prime, and the group ring K = Fp[Zn]. The "cotangent space" T*(Zn, Fp) of Zn over Fp is naturally understood to be the cotangent space of K. Below we characterize T*(Zn, Fp) and also obtain the "fundamental theorem of mechanics" for the rings K. In what follows we use the representation of K as the factor ring Fp[x]l(xn - 1). Then the Fp-algebra of differential operators Diff* K is isomorphic to K if n is prime to p, and is generated by the elements, X, bx subject to the relations X *dx — dx*X = —1, dnx = 0 if n is a multiple of p. Therefore, [ K with the trivial Lie structure if n is prime to p, S = Smbl*A = { [ Fp[x, y]/(xn
- 1 , yn),
{x, y}=— 1, if n is a multiple of p.
256
AM
Vinogradov and IS. KrasiVshchik
Thus, T*(Zni Fp) is a "regular n-gon" embedded in the affine space Spec Fp[x] if n is prime to p, and an "w-fold regular «-gon" embedded in Spec Fp [x, y] if n is a multiple of p. An easy calculation shows that the "fundamental theorem of mechanics" (that is, Can S/Ham S) is trivial if n is prime to p, and is isomorphic to the Abelian 2m2-dimensional Lie algebra over Fp if n - mp, m integral. 10. The Hamiltonian formalism in Hamiltonian rings. As we have seen, from the formal-structural point of view, we need only the existence of Poisson brackets to construct a "Hamiltonian formalism". On the other hand, the above method of defining Poisson brackets was based only on the fact that in the filtered ring Diff* K D . . . D Diffs K . . . D K the commutator of elements At(i = 1 , 2 ) of filtration < vt is an element of filtration < vl + v2 — 1. Therefore, we can construct a "Hamiltonian formalism" for any unital associative filtered ring F D . . . D Fs D . . . D Fo in which the commutation relation is linked up with a filtration in the indicated manner. We call such rings Hamiltonian. Then the graded ring Smbl# F = S
FIIFI-I
associated with F under the assumptions made is
commutative and unital. The Poisson bracket {a1? a2} = smblroi+TOa_1[61, 62] of the elements ai = smblm/ 5; (i = 1, 2) is well defined if by the symbol smblm 5 where 6 has filtration
x
In this context, see [ 1 ] .
What is the Hamiltoniam formalism ?
25 7
3
/(**) = bik ft it = 1, 2, . . . , 7i) (that is, 5, = - J r ) and A(5) is also free and is generated by elements of the form d{ - dxi (/ = 1, 2, . . . , « ) . Moreover, (e/,-, 3^ > = 5 ^ , that is, the bases described above are conjugate (we recall that D(S) = Homs (A(S), S)). Also, there are natural isomorphisms D(S)~L*
(*\ / = 1, 2, . . . , n)
be the structure equations of the Lie algebra L. The Poisson brackets are obviously defined by their values on the generators xh and
Thus,
n n
xt=
n
ZJ \ 2J wj^fc/ Oj
(i = 1, 2, . . . , n).
n
The derivation d = ^, atdt
is Hamiltonian if and only if there is an
element 5 G S such that af = {5, xt) (i = 1, 2, . . ., n), that is, n
n
The element 5 is the "Hamiltonian" of 3. n Now we write out the conditions for the derivation d = 2 a^ f to be canonical. As we know (see 4.), 3 is canonical if and only if d(5x> s2} = {dsl9 s2} + {st, ds2} for all su s2 G S. In our situation the verification that 3 is canonical obviously reduces to checking the following equalities: d{xt, Xj} = {dxtJ Xj} + {#*> dzj} (h / = 1, 2, . . ., n) or n
n
n
n
n
h=l
m=l
k—l
m=l
ft=l
(/, 7 = 1 , 2 , . . . , 7i).
We consider the matrices C = \\ cfj- || and ^4 = 11fl/;-||, where n Ctj — 2\
^ijxki
a
ij~^r
(h 7 - ^ l » A, . . . , n).
258
AM
Vinogradov and LS. Krasil'shchik
Note that C is skew-symmetric. We write 3C = || 3c/;- ||. Then the conditions (*) can be rewritten in the compact form dC = C*A{ - A'C, where A1 is the transpose of A. 12. The "fundamental theorem of mechanics" for Lie algebras. Let L be a Lie algebra, U its universal enveloping algebra, and S the corresponding ring of symbols. As already remarked, a "fundamental theorem of mechanics" for the ring U is to be understood as a theorem describing the factor module Can (//Ham (/, equipped with the natural structure of a Lie ring. Let us write y(L) = Can (//Ham U. We remark that when L is a Lie algebra over a ring K, then y(L) can be made into a Lie algebra over K. EXAMPLE. Let L be an Abelian Lie algebra. It is clear that Ham U(L) = 0 and Can U(L) = D(S). Hence y(L) = D(S). EXAMPLE. Let L be the two-dimensional Lie algebra over a ring K with two generators yx and y2 and the Lie bracket [y\, y2] ~ yi\ its ring of symbols is the polynomial ring K[xu x2] with the Poisson bracket {xu x2} = x2. The derivation 3 = c^Bj + OL2d2 is Hamiltonian with Hamiltonian s if ax = -X2I7 » an(* a 2 = x2f^ > while d is canonical if
Therefore, 7(L) = K with the trivial Lie structure. We remark that the natural filtration S = U Sj of the ring of symbols induces a filtration of its module of derivations D(S); namely D(S)=
U Dt (S),
where Dj(S) consists of precisely those derivations that map Sx into Sj. It is not difficult to check that Can U = U Can.- U and z>0
l
Ham U = U Ham,- U where Can,- U = Can U O £>;(£), and Ham,- tf = Ham tf n Dt(S) (/ = 0, 1, 2, . . . , m, . . .). Since Ham,- U C Canz- U, we have, in turn y(L) = U Y*(£), where Yf(L)
= Can, (//Ham, (7,
?f (L)
cz yi+1(L)
(i = 0, 1, 2, . . ., m).
It follows immediately from the definition that Canx U is, modulo elements of Can0 U, the Lie algebra of endomorphisms of L, and Hamj U the Lie algebra of inner endomorphisms of L. Therefore, yi(L)/yo(L) - Hl(L; L), where the space L is endowed with its natural L-module structure by means of the adjoint representation ad: L -> End L.
What is the Hamiltonian formalism?
25 9
Thus, the Lie algebra y(L) is in a certain sense an "enveloping object" for the first cohomology group Hl(L\ L). 13. The "fundamental theorem of mechanics" for semisimple Lie algebras. THEOREM I. If L is a finite-dimensional semisimple Lie algebra over a field of characteristic zero, then y(L) = 0. PROOF. We define the Poisson commutator of S as the subspace {S, S} of S consisting of the finite linear combinations of all elements of the form {sl? s2}, 5 i' S2 £ $ with coefficients in K. We consider the linear representation R of L on Can U: R: L-+- End(Can U), where i?(Z)if = [if, Hj(l)], op 6 Can U, I £ L ; this turns Can U into an Z-module. Further, we consider the filtration of Can U defined in 12. and set Ct = Can,- U. Then Ci C Ci+1 and Can U = .U C r It is easy to see that if £ Ct if and only if if is a canonical derivation and ap/(Z) £ £*, Z £ L. Let us check that the Ct are invariant subspaces of R. For let Zj, l2 G I, ^/(Zj) 6 S f . Then
and since ^ / ( y 6 5«, we have (R(hH)(j(h)) = W(h), j(h)} 6 5 4 . Thus, /^j- = R\c.: L -> End Ct is finite-dimensional and is therefore completely reducible'(see [16], Ch. VI, §3). Hence it follows ([17], Ch. V) that Ct = Ci © C°i, where Ci is the subspace of Ct formed by all vectors that are annihilated by the endomorphisms in Rt(L) and Cf is generated by the vectors of the form R;(l)(c), where / E L and c E Ct. It is clear that C? c Cf+1 and C^ C C?+1. Let Can^ Z7 and Can0 U be the subspaces of R defined similarly. Then f and LEMMA 1. / /
U C\
Can U = Can^ f/ 0 Can0 U. if G Can U, then if{£, S} cz{S, S},
^Z(S) aZ(S).
PROOFS. 1) if{«i, s2} = {ifsl7 s2} + {5l, if52} e {S, S}. 2) The fact that z G Z{S) means that {5, z) = 0 for all 5 E 5, and so 0 = if{2, s) = {ifz, s} + {2, ^fs} = {t z ' s> , that is, if(z) 6 Z(S). LEMMA 2. S = Z(S) 0 {S, S}. ' We consider the representation R': L -+ End 5, where R'(l) = Hj(l), I E L. An argument quite similar to that by means of which we have established above the decomposition Can U = Can0 U 0 Can^U, shows that S = S* 0 S°, where S° = {5 6 £: 5 = 2#j(*i)(*i)} a n d
260
A.M. Vinogradov and IS. KrasiVshchik
S* = {s G S: Hj(l)(s) = 0, I 6 L}. Therefore, it only remains to remark that S*=*{S, S) and S* = Z(S). We return to the proof of the theorem. Let i|) 6 Can*7 L. This means that i|? is a canonical derivation and that for all / G Z and s G 5 0 = (tf(Z)(i|>))M = hp, #/(0K*) = {*;(Q, 5}.
Thus, Can^£7 consists of those i|? £ Can U that are carried by j(L) into Z(5). But since L is semisimple, [Z,, L] = Z,, so that j(L) a {S, S}, and according to the lemmas just proved, Gan^ U = 0. So we arrive at the following fundamental equation: Can0 U = Can U. Now let i|> 6 Gan° 17. Then there exist lu l2, . . . , /^ G L and *!» *2» • • ., t* € Can f/ such that iK*')=2 (^(^Hi)^') = ft
A
i=i
•
<
z
5
2 {*<0W),*'} = ( 2 **( ()» '} for all 5' e S. Consequently i=l
?=i
h
\|5 = Hs £ Ham *7, where ^ — 2 tr/(^)- Therefore, >=i
Ham f/ C Can U C Can0 t/ C Ham J7, from which it follows that Ham U = Can U. The theorem is now proved. 14. Example. The fundamental theorem of mechanics and symmetries of linear vector fields. Let us consider in the ^-dimensional Euclidean space Rn a linear vector field
z =
n
A
2 *wr
where
=
n
^ 2 *'*** ( i s = = 1 ' 2 ' •••' w)»
and the matrix .4 = 11 a^ II is non-zero. Speaking invariantly, we regard the field X as an element of D(K), where K is the ring of smooth functions on Rw, Rrt C Spec K. Throughout this section we agree to understand by symmetries of X those diffeomorphisms Rn ->- Rn that carry trajectories of the field in question into trajectories of the same field. The set of diffeomorphisms of Rn that carry each trajectory of X into itself form in the Lie algebra Sym X an ideal Sym X ("proper" or "inner" symmetries of X) whose elements, evidently, are of the form D = fiX, ju E C°°(RW). In what follows we confine ourselves to the algebraic case, that is, to those symmetries whose coefficients are polynomial functions on Rn (for such symmetries, X and n belpng to R[xj, . . . , xn]). Speaking invariantly: we regard X as ail element of D(K), where A' is a ring of polynomials on R", Rn C Spec K. From now on the notation Sym X and Sym X refers to the sets of precisely these symmetries. Let S(X) = Sym X/Sym X. We associate with X the (n + 1 )-dimensional Lie algebra L{X) with the
What is the Hamiltonian formalism?
261
generators yi, y2, . • . , yn + \ and the relations [*/*, yj] = 0, [yn+1, yt] = 2 aikyki
1 < i, / < H. The algebra L(X) is
obviously solvable and is isomorphic to the subalgebra of S = Smbl* K (relative to the Poisson brackets) spanned by the coordinate functions Xi, x2, • • . ,n x and the Hamiltonian function V, aikpiXk is the natural lifting of X in T*(RW). *•k^ THEOREM 2 The Lie algebra y(L(X)) is the semidirect product of S(X) and a commutative ideal J that can be identified with the space Coker X (the field is regarded as an operator on the ring R[xl9 . . . , xn]). PROOF. Let z 1? z 2 , . . . , zn + 1 be the images of the elements ^IJ y-i-> • •»•yn + i under the canonical embedding /: L(X) -> Smbl* L(X). n+i
We consider the derivation d — >J a '~9i7 °f Smbl*
3 is Hamiltonian
with Hamiltonian 5 G Smbl* L(X) if and only if Bt
ds A=l
(N)
<= 2
( i = l , 2, . . . , » ) .
In the present case the conditions for 3 to be canonical, according to the formula (*) in 11. can be written in the form of the following system of equations: • = 0,
(S) Solving (S) we obtain
/*
\
0
\
ft=i
where px, p2, . . ., Pn, P»
In this context we note that the condition [D, X] = XX for D = V, X* -J— i=t
262
A.M. Vinogradov and I.S. Krasil'shchik
to be a symmetry of X has the following form in coordinates: Ahd^)=XAt
( i = l , 2, . . . , « ) .
If we now associate with every D E Sym X the derivation
of Smbl* L(X), then by the remark just made we get a Lie algebra monomorphism p:Sym X -• Can £/(Z(XA)). Further, it is easy to check that p(Sym X) n Ham U(L(X)) = p(Sym X), that is, p': 5(Z) -> T ( I ( X ) ) is an embedding of Lie algebras. There is also the embedding T:
R [xi9 * a , . . . , xn] -> Can U (L (X));
% (q>) -
Then { T ^ ) , t((p2)} = 0, that is, Imcp is a commutative subalgebra of Can U(L(X)). Furthermore, it is immediately clear from (N) and (S') that: 1) Can U (L(X)) = Ham U (L(X)) + Im p + Im r; 2) Im r n Im p = 0;
3) {P:(Z)),T(CP)}=2 ( ^ g - M - ^ ^ ( 2 ^fi~^) ( t h a t is' {Im p, Im T} CI Im T); and 4) Ham U (L(X)) H Im r consists of the Hamiltonian fields with Hamiltonians of the form ty C R \z\, . . . , Znl, that is, fields of the form —X(^)-^—^, where X (i|>) ,= ^ aikzh^-» If we now write r' for the composition of the embedding r with the projection Can U (L(X)) -> y(L(x)), then from our remarks it follows at once that y(L(X)) = Im p' © Im r', where Imt'^CokerX and Im p' are subalgebras of y(L(X)), and Im r' an Abelian ideal. NOTE 1. We restrict our attention to the set of elements of filtration 1. Then, if A is invertible, yo(L(X)) = 0 and SX(X) = yx(L{X)). Thus, the Lie algebra of linear symmetries of X, factored by the "proper" linear symmetries, is isomorphic to the first cohomology group H1(L(X); L(X)). NOTE 2. This theorem together with the theorem of 12. and the first example of 11. describes completely the invariant y(L) for all 3-dimensional Lie algebras. 15. In this section we examine the connection between the concept of an algebraic Hamiltonian orbit (see 7.) of S(U) and that of an orbit of the ("geometrical") coadjoint operation Ad* of the Lie group G o n l * , the conjugate space of L, U = U(L). These latter orbits are of interest owing to their role in representation theory (the "orbit method", see [9] and [10]). In connection with this we also introduce below a version of the "orbit method", which comes immediately from our view of the Hamiltonian
What is the Hamiltonian formalism?
263
formalism. First of all we recall that S = S(U) can be identified naturally with a ring of polynomial functions on I * ; this gives an embedding of topological spaces L* C Spec S. The following statement describes the geometrical orbits in L* in "Hamiltonian" terms. PROPOSITION 9. If p e L*, then the geometrical orbit G(p) of p is obtained by shifting p along the trajectories of all possible Hamiltonian vector fields (that is, G(p) - pH, see 7). PROOF. Since shifts along trajectories of Hamiltonian vector fields include, in particular, shifts with respect to trajectories of fields of the form Hx, where x E Im /, /: L -> S(U) is the canonical embedding, it is obvious that every manifold of the form pH contains the geometrical orbit G(p) of p E L*. On the other hand, any Hamiltonian field Hs, s G S(U), being an S-linear combination of fields of the form Hx, xG Im / (this follows from the relation H(sls2) = sxHs2 + s2Hsx), touches each of the geometrical orbits in L*. Thus, G(p) D pH. COROLLARY 1. The geometrical orbits G(p) depend only on L, and not on G. COROLLARY 2. The embedding G(p) c 0 ( p ) holds, with equality when G(p) is an algebraic variety. If the Lie group G has the structure of an algebraic group, then its orbit in some vector space V is an algebraic variety with an algebraic hypersurface excised. More precisely, each orbit is given by a system P0(v) ± 0, P((v) = 0 (/ = 1, 2 k)9 where v G V and Po> Pu . . . , Pk are polynomial functions on V. This proves the next proposition. PROPOSITION 10. When G is an algebraic Lie group, each geometrical (= Hamiltonian) orbit G(p), p E L*, coincides with the principal part of the corresponding algebraic orbit O(p). The study of Lie algebras of the form L(X) shows that y(L(X)) can indicate very sensitively whether orbits are algebraic. Without going into details of the calculations, we remark, that when n = 2, the Lie algebra y(L(X)) is commutative if and only if the geometrical Hamiltonian orbits of L(X) are not algebraic varieties. We complete our study of Hamiltonian orbits of Lie algebras with the following remarks. Let Qh: A(U) 0 A(U) ->• S(U) be a canonical form, U = U(L), and A(U) - A(S(U)). This form, in general, is degenerate and has a "kernel" Ker £lh consisting of those elements X £ A(U) for which £lh(k, X') = 0 for all X' G A(U). However, the Hamiltonian map h: A(U) -+ D(S(U)) allows us to consider in D(S(U)) the submodule Im h, which as is easy to see is formed by the linear combinations of the form u 2 s\Hsi, st, si 6 S(U). Thus, £lh determines on Im h a non-degenerate il
264
-AM. Vinogradov and IS. Krasil'shchik
closed bilinear skew-symmetric form £Lh. These arguments, together with Proposition 9 demonstrate the following result. PROPOSITION 11. The orbits G{p) are the maximal submanifolds of L* on which £lh is non-degenerate. COROLLARY: The orbit G(p) is a canonical manifold. In conclusion we consider briefly a general principle of constructing representations of Lie groups, which is a paraphrase of the "orbit method". The classical problem in representation theory is to search for all irreducible representations on an a priori given class of spaces (Hilbert spaces, for example). It seems to us that this contains an element of arbitrariness and can hardly be called natural. We therefore replace the postulation of a "natural" class of spaces by a natural principle suggested by our view of the Hamiltonian formalism. Roughly speaking, the classical quantization procedure is a method of reconstructing a differential operator from its symbol (the classical Hamiltonian). In general, there is no natural map Diff* K -• Smbl* K (or vice versa); to make the reconstruction possible, K must have further structure. In the case K = C°°(M) such a structure is the connectivity of the manifold M (see [15], for example). In particular, a Lie group has a left-invariant connection, which can be used naturally. Thus, let I be a Lie algebra and G its corresponding simply-connected Lie group. If xu . . . , xn is a basis in L, then we call the coordinates of a point x G I with respect to this basis the exponential coordinates of exp (x) G G. Again, let x G L and let lx be the operator of left translation by x in G, that is, lx(g) = exp(x)*g, g G G. Then the left-invariant differential operators on G (that is, the elements of U(L)) are defined by the condition (5)
l*xDol* = Dx,
where Dx is the value of the differential operator D at the point with exponential coordinates x. There is obviously a one-to-one correspondence between the values at 0 of the differential operators on G and the elements of S(U). Thus, (5) establishes1 the required map Q: S(U) -» U. We also set Sm = Q"1. The element Sm(D), D G U, is called the full symbol of D. Let / be an ideal in S(U) that is stable under Hamiltonian derivations. It determines a left ideal QJ C Urn the universal enveloping algebra in the following way: the ideal QJ consists of those elements u G U for which Sm(w'w) G / for all elements u G U. The left ^/-module U/QJ determines a representation of L. It is natural to expect that the so arising "simplest" (possibly irreducible) representations then come from "really simple", namely maximal stable ideals of S(U). We call the method described here of obtaining representations of Lie 1
SeeGel'fand[5].
What is the Hamiltonian formalism?
265
algebras the principle of localization by symbol. We illustrate it by one example. EXAMPLE. Let L be the three-dimensional nilpotent Lie algebra consist/0 /U Xi Xi xxz3\ ing of t h e m a t r i c e s of t h e f o r m [[00 0 0 x, x2\ ( t h e Lie algebra of t h e
Vo o o
Heisenberg group). Then each left-invariant differential operator on the corresponding Lie group is, when expressed in exponential coordinates, a linear combination of operators of the form e^e^e"3* , where ePi
-
d
x
*
d
— &ri"-"2~te3f
It is easy to see that
8
-JL + 3-A-
**~dx2^
min(m,n 2 ) ..
2 dz3'
{
8b3
-J_
""a*3"
{
where yx, y2, y-s generate S(U) and Sm zt — yt (i = 1, 2, 3). There are two types of maximal stable ideals in S(U). Firstly, the principal ideals Jr generated by irreducible polynomials r(y3). The relation Sm(uE3) = Sm(u)y3, u 6 U, shows that QJr is the left ideal of U generated by r(e 3 ). The corresponding representation space R(Jr) is the space of differential operators on the plane with real or complex coefficients, depending on the form of the polynomial r. The image of e 3 in this representation is the operator of multiplication by X, r(X) = 0. If we denote by {enu n2} (nn rc2 = 0, 1, 2, . . .) a basis of R(Jr), then the operators corresponding to e± and ea, act in the following way: R ( 8 l) eni, n2 = e r»i+l, n2>
R (E2) eni, n2 = enu n2+l ~ ^ienu
W
The maximal stable ideals in S(U) of the second type are Jp.q = (p» 9S ys)i where p and q are irreducible polynomials depending only on yx and y2, respectively. Observing that the commutator of any two elements ul9 u2 ^ U lies in the left ideal of U generated by s3, we can verify that QJPiQ is the left ideal of U generated by />(ei), g(e2) and 83. The corresponding representation spaces R(Jp q) are the tensor products B.l^1]/(p) 0 RU2]/(g) on which R(&3) = 0 and the actions of JR(8X) and i?(e2)are multiplication by the cosets of ex and e2 inRle^/ip) and R[s2]/(g) respectively. Thus, we can observe better the correspondence between our principle and known results on representations of the Heisenberg group (see [9]). We shall study this method in greater detail elsewhere. In conclusion, the authors apologise for using the word "natural" excessively often.
266
A.M. Vinogradov and I.S. Krasil'shchik
References
[1] F.A. Berezin, Some remarks on the enveloping algebra of a Lie algebra, Funktsional. Anal, i Prilozhen. 1:2(1967), 1-14. MR 36 #2750. = Functional Anal. Appl. 1 (1967), 91-102. [2] A.M. Vinogradov, The algebra of logic for the theory of linear differential operators, Dokl. Akad. Nauk SSSR 205 (1972), 1025-1028. MR 46 # 3498. = Soviet Math. Dokl. 13 (1972), 1058-1062. [3] A.M. Vinogradov, Many-valued solutions and a classification principle of non-linear differential equations, Dokl. Akad. Nauk SSSR 210 (1973), 11-14. = Soviet Math. Dokl. 14 (1973), 661-665. [4] P. Gabriel, Proprie'tes generates des schemas en groupes, Lecture Notes in Math. 151* Expose VIIA. Springer-Verlag, Berlin-Heidelberg-New York 1970. [5] I.M. Gel'fand, The centre of an infinitesimal group ring, Mat. Sb. 26 (1950), 103-112. MR 11-498. [6] A. Grothendieck, Categories cofibrees additives et complexe cotangent relatif, Lecture Notes in Math. 79, Springer-Verlag, Berlin-Heidelberg-New York 1968. MR 39 # 2835. [7] A. Grothendieck, Ele'ments de geometrie algebrique. IV. Etude locale des schemas et des morphismes schemas, Inst. Hautes Etudes Sci. Publ. Math. 32 (1967). MR 39 # 220. [8] Yu. V. Egorov, On the solubility of differential equations with simple characteristics, Uspekhi Mat. Nauk 26:2 (1971), 183-199. MR 45 # 5541. = Russian Math. Surveys 26:2 (1971), 113-130. [9] A.A. Kirillov, Unitary representations of nilpotent Lie groups, Uspekhi Mat. Nauk 17:4 (1962), 57-101. MR 25 # 5396. = Russian Math. Surveys 17:4 (1962) 53-104. [ 10] A.A. Kirillov, Lektsii po teorii predstavlentigrupp, IV. Predstavleniya i mekhanika (Lectures on the theory of group representations, IV. Representations and mechanics), Izdat. Moskov. Gos. Univ., Moscow 1971. [11] B. Kostant, Quantization and unitary representations, Lectures in modern analysis and applications, II, Springer-Verlag, Berlin-Heidelberg-New York 1970, 87-208. = Uspekhi Mat. Nauk 28:1 (1973), 163-225. [12] Yu. I Manin, Lektsii po algebraicheskoi geometrii, I. Affinnye skhemy (Lectures on algebraic geometry, Part I. Affine schemes), Izdat. Moskov. Gos. Univ., Moscow 1970. [13] V.P. Maslov and B. Yu. Sternin, Kanonicheskii operator!kompleksnyi sluchai) (Canonical operators (The complex case)), in the coll. "Problems of modern mathematics", vol. I, All-Union Inst. Sci. Tech. Information (VINITI) 1973, 169-159. [14] V.P. Maslov and M.V. Fedoryuk, Kanonicheskii operator (veshchestvennyi sluchai) (Canonical operators (The real case)), in the coll. "Problems of modern mathematics", vol. I, All-Union Inst. Sci. Tech. Information (VINITI) 1973, 85-168. [15] R. Palais, Seminar on the Atiyah-Singer index theorem, Annals of Mathematics Studies, No.57, University Press, Princeton, N.J., 1965. MR 33 # 6649. Translation: Seminar po teoreme At'i-Zingera ob indekse, Mir, Moscow 1970. [16] J-P. Serre, Lie algebras and Lie groups, Benjamin, New York 1965. Translation: Gruppy Li i algebry Li, Mir, Moscow 1969. [17] Seminaire "Sophus Lie", Theorie des algebres de Lie. Topologie des groupes de Lie, Ecole Normale Supe'rieure, Paris 1955. MR 17-384. Translation: Teoriya algebr Li. Topologiya grupp Li, Izdat Inost. Lit., Moscow 1962. Translated by Ph. Spain
Received by the Editors, 19 July 1974