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!
X' be a closed subspace of X'. Then i»S is a unique sheaf on X' such that i*S\x = 5,
t , 5 | X ' \ x = 0.
Indeed, i f x ' s X c X', then i*S(U') = S(U' n X) for any open neighborhood U of this point. If x' $ X, there exists its neighborhood U' such that U' H X is empty, i.e., i*S(U') = 0. The sheaf i*S is called the trivial • extension of the sheaf S. By a morphism of ringed spaces (X, 9^) —» (X', 9V) is meant a pair (ip, $) of a continuous map ip : X —> X' and a sheaf morphism $ : 9t' —> 0*9l or, equivalently, a sheaf morphisms ?*9t' —• 9^ [414]. Restricted to each stalk, a sheaf morphism $ is assumed to be a ring homomorphism. A morphism of ringed spaces is said to be: • a monomorphism if
X x X X'
is closed. In this case, one says that (X, JH) is separated over X'. A scheme (X, 9t) is called separated if it is a separated over Spec Z. All afflne schemes are separated.
Chapter 1 Commutative Geometry
73
A morphism of schemes ( X , 9 i ) - » ( * ' = Spec .4,21) is said to be locally of finite type (resp. of finite type) if (X, 9t) has an open affine cover (resp. a finite open affine cover) {Ui = Spec.4j} such that each Ai is a finitely generated ,4-algebra. A general morphism of schemes tp (1.8.4) is said to be locally of finite type (resp. of finite type) if there is an open affine cover {V{\ of X' such that every restriction of
(iR(tf')). Given a local-ringed space (X, 91), a sheaf P on X is called a sheaf o/9lmodules if every stalk Px, x £ X, is an 9tx-module or, equivalently, if P(U) is an 9t(t/)-module for any open subset U C X. A sheaf of 9l-modules P is said to be locally free if there exists an open neighborhood U of every point x G X such that P(U) is a free 9t(f7)-module. If all these free modules are of finite rank (resp. of the same finite rank), one says that P is of finite type (resp. of constant rank). The structure module of a locally free sheaf is called a locally free module.
The following is a generalization of Proposition 1.7.8 [220]. 1.8.1. Let X be a paracompact space which admits a partition of unity by elements of the structure module S(X) of some sheaf S of real functions on X. Let P be a sheaf of 5-modules. Then P is fine and, consequently, acyclic. • PROPOSITION
74
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let (X, SH) be a ringed space. A sheaf of 9^-modules S is said to be quasi-coherent if for each point x of X there exists a neighborhood U of x and an exact sequence M —*N —• S| l / -»o, where M and N are free 9t|(y-modules. Let 5 be a locally free sheaf of 9t-modules of finite type. It is said to be of finite presentation if, locally, there exists an exact sequence 9tm —>9tn —>5->0, where m and n are positive integers (which need not be globally constant). A locally free sheaf of finite type is called coherent if the kernel of any homomorphism 9t|j} —* S\u, where n is an arbitrary positive integer and U is an open set, is of finite type. Obviously, if S is coherent, then it is of finite presentation, and is quasi-coherent. If 91 itself is coherent as a sheaf of 9t-modules, then it is called the coherent sheaf of rings. In this case, every sheaf of Dt-modules of finite presentation is coherent. For instance, the structure sheaf of a locally Noetherian scheme is a coherent sheaf of rings. Let (X = Spec A, 9^ = 21) be an affine scheme. Then every quasicoherent sheaf S of 2l-modules on X is generated by its global sections. The correspondence S —> S(X) defines an equivalence between the category of quasi-coherent sheaves on X and the category of .4-modules. If .4 is Noetherian, then the coherent sheaves and the finite ^.-modules correspond to each other under this equivalence. Let (X,fR) be a separated scheme and il = {[/*} an affine open cover of X. For each quasi-coherent sheaf S of 9^-modules, the cohomology H*(X;S) of X with coefficients in S is canonically isomorphic to the cohomology H*(iX;S(U)). One defines the cohomological dimension cd(X) of a scheme (X,SR.) as the largest integer r such that Hm(X;S) ^ 0 for a quasi-coherent sheaf of fK-modules on X. For instance, if X is an affine scheme, then cd(X) = 0. The converse is true under the assumption that X is Noetherian. The sheaf Cg of smooth real functions on a smooth manifold X provides an important example of a local-ringed spaces. Remark 1.8.3. Throughout the book, smooth manifolds are finitedimensional real manifolds, though infinite-dimensional Banach and Hubert manifolds are also smooth. A smooth real manifold is customarily assumed
Chapter 1 Commutative Geometry
75
to be Hausdorff and second-countable (i.e., it has a countable base for topology) . Consequently, it is a locally compact space which is a union of a countable number of compact subsets, a separable space (i.e., it has a countable dense subset), a paracompact and completely regular space. Being paracompact, a smooth manifold admits a partition of unity by smooth real functions. One also can show that, given two disjoint closed subsets N and N' of a smooth manifold X, there exists a smooth function / on X such that /|AT = 0 a n d /|AT' = 1. Unless otherwise stated, manifolds are assumed to be connected and, consequently, arcwise connected. We follow the notion of a manifold without boundary. D Similarly to the sheaf C^ of continuous functions, the stalk C£° of the sheaf Cj^P at a point x € X has a unique maximal ideal of germs of smooth functions vanishing at x. Though the sheaf Cjj? is defined on a topological space X, it fixes a unique smooth manifold structure on X as follows. 1.8.2. Let X be a paracompact topological space and (X,9\) a local-ringed space. Let X admit an open cover {Ui} such that the sheaf 9K restricted to each Ui is isomorphic to the local-ringed space (R",C^,). Then X is an n-dimensional smooth manifold together with a natural isomorphism of local-ringed spaces (X,9i) and (X,Cj^). • THEOREM
One can think of this result as being an alternative definition of smooth real manifolds in terms of local-ringed spaces. A smooth real manifold X is also algebraically reproduced as a certain subspace of the spectrum of the K-ring C°°(X) of smooth real functions on X as follows [17]. Let A be a commutative R-ring and Specm.4 its maximal spectrum. The real spectrum of A is the subspace SpecR.4 C Specm A of the maximal ideals X such that the quotients A/I are isomorphic to R. It is endowed with the relative Zariski topology. There is the bijection between the set of R-algebra morphisms of A to the field R and the real spectrum of A, namely, HomR(.4,R) 3
Ker > e SpecRA SpecR.4 9 x H-> irx e HomR(.4,R), 0. «p, O\ <8> qj. UK, Ane Cp+1'q+1(Z), L~ : CP'"(Z) 9 and W of <& are equivalent, the representations TT* and Try of C(0,a) are also equivalent. Let us provide the convolution algebra C(<&,a) with a different norm | | / | | = sup ||7r(/)||, where TT runs through all bounded representations of the algebra C(<S,a). The completion C*(<8,cr) of C(<3,cr) with respect to this norm is a C*algebra, called the convolution C* -algebra of the groupoid <8. Let A be a positive measure on the unit space (25° of a locally compact groupoid 0. Then \{f\@o), f £ C(<5,a), is a positive continuous form on C(<3,cr) equipped with the inductive limit topology. If <S is an r-discrete groupoid and A a measure of total mass 1, this form is extended to a state on the C*-algebra C*(<8,a) of 6. 3.5 °-", \ of <j>. ) A a + ( - l ) H I ^ u\(t) e CE ) as graded local-ringed spaces, where $ is an even graded morphism. In particular, every morphism ip : M —> M' yields the smooth supermanifold morphism ( Cfa, where U H, P'. DEFINITION 6.9.11. A quotient of an action of a G-Lie supergroup on a G-submanifold P is a pair (M,TT) of a G-supermanifold M and a Gsupermanifold morphism 7?: P —» M such that: (i) there is the equality
nx : A -> A/x = R.
Any element a € A induces a real function fa : Spec R «4 3 x )—> nx(a)
on the real spectrum SpecRA This function need not be continuous with respect to the Zariski topology, but one can provide SpecR.4 with another
76
Geometric and Algebraic Topological Methods in Quantum Mechanics
topology, called the Gel'fand topology, which is the coarsest topology which makes all such functions continuous. If A = C°°(X), the Zariski and Gel'fand topologies coincide. 1.8.3. Given the R-ring C°°(X) of smooth real functions on a manifold X, let fix denote the maximal ideal of functions vanishing at a point x e X. Then there is a homeomorphism THEOREM
Xx-X3x^
fixe SpecRC°°(X).
(1.8.5)
a Let X and X' be two smooth manifolds. Any smooth map 7 : X —> X' induces the R-ring morphism 7* : C°°{X') -» C°°{X) which associates to a function / on X' the pull-back function 7*/ = / o 7 on X. Conversely, each R-ring morphism C : C°°{X') -» C°°(X) yields the continuous map C*1 (1-8.3) which sends SpecKC°°(X) C SpecC°°(X) to SpecRC°°(X') C SpecC°°{X') so that the induced map Xx'C* o xx •• X -» X' is smooth. Thus, there is one-to-one correspondence between smooth manifold morphisms X -> X' and the R-ring morphisms C°°(X') -> C°°(X). Remark 1.8.4. Let X x X' be a manifold product. The ring C°° (X x X') is constructed from the rings C°°(X) and C°°(X') as follows. Whenever referring to a topology on the ring C°°(X), we will mean the topology of compact convergence for all derivatives [374]. The C°°(X) is a Frechet ring with respect to this topology, i.e., a complete metrizable locally convex topological vector space. There is an isomorphism of Frechet rings
C^POSC^PO *£ C°°(X x X'),
(1.8.6)
where the left-hand side, called the topological tensor product, is the completion of C°°{X) ® C°°(X') with respect to Grothendieck's topology, defined as follows. If Ei and E2 are locally convex topological vector spaces, Grothendieck's topology is the finest locally convex topology on Ei <&EQ. such that the canonical mapping of E\ x E2 to Ei ® E2 is continuous [374]. It is
77
Chapter 1 Commutative Geometry
also called the 7r-topology in contrast with the coarser e-topology on E±®E2 [356; 417]. Furthermore, for any two open subsets U C X and V C X', let us consider the topological tensor product of rings C0O([/)(g)C°o(C/'). These Due to tensor products define a locally ringed space (X x X',Cx®Cx,). the isomorphism (1.8.6) written for all U C X and U' C X', we obtain the sheaf isomorphism /TOO Zi./-«X)
.TOO
/ I n 71
(1.S.7J
OX®OX, -CXxX,.
D Since a smooth manifold admits a partition of unity by smooth functions, it follows from Proposition 1.8.1 that any sheaf of C^-modules on X is fine and, consequently, acyclic. For instance, let Y —> X be a smooth (finite-dimensional) vector bundle. The germs of its sections make up a sheaf of C^-modules, called the structure sheaf Sy of a vector bundle Y —> X. The sheaf Sy is fine. In particular, all sheafs Ox, k G N+, of germs of exterior forms on X is fine. These sheaves constitute the de Rham complex —>K — >u
x
— > Ux
— > • • • Ux
—>••• .
(l.o. a)
The corresponding complex of structure modules of these sheaves is the de Rham complex 0 - • R —* C°°{X) -1+ Ol{X) -iU • • • Ok(X) - ^ • • •
(1.8.9)
of exterior forms on a manifold X. Its cohomology is called the de Rham cohomology H*(X) of X. Due to the Poincare lemma, the complex (1.8.8) is exact and, thereby, is a fine resolution of the constant sheaf R on a manifold. Then a corollary of Theorem 1.7.6 is the classical de Rham theorem. THEOREM
1.8.4. There is the isomorphism Hk(X)=Hk(X;R)
(1.8.10)
of the de Rham cohomology H*(X) of a manifold X to cohomology of X with coefficients in the constant sheaf M.. Q Remark 1.8.5. The de Rham cohomology groups H*(X) of a compact manifold X are finite-dimensional real vector spaces so that dimHh(X) is
78
Geometric and Algebraic Topological Methods in Quantum Mechanics
the k-th Betty number of X, and X(X)=
J2dimHk(X) k
is the Euler characteristic of X. Remark 1.8.6. sheaves
•
Let us consider the short exact sequence of constant 0 - > Z —»R —>I7(1)-»O,
(1.8.11)
where C/(l) = M/Z is the circle group of complex numbers of unit modulus. This exact sequence yields the long exact sequence of the sheaf cohomology groups O ^ Z —>R —>tf(l) -^Hl{X;Z) HP{X;Z)
^HP(X;R)
—> • • •
-^H\X;R)
—*HP(X;U(1))
—>HP+1(X;Z)
—*••-,
where H°(X;Z)=Z,
H°(X;R)=R
and H°(X; U(l)) = U{\). This exact sequence defines the homomorphism H*(X;Z)^H*(X;R)
(1.8.12)
of cohomology with coefficients in the constant sheaf Z to that with coefficients in R. Combining the isomorphism (1.8.10) and the homomorphism (1.8.12) leads to the cohomology homomorphism H*(X\Z)->ir(X).
(1.8.13)
Its kernel contains all cyclic elements of cohomology groups Hk{X\ Z). • Given a vector bundle Y —> X, the structure module of the sheaf Sy coincides with the structure module Y(X) of global sections of Y —» X. The Serve-Swan theorem, shows that these modules exhaust all projective modules of finite rank over C°°(X). As was mentioned in Introduction, this theorem has been proved in the case of a compact manifold X. However, its proof can be generalized to an arbitrary smooth manifold due to Theorem 10.6.2.
79
Chapter 1 Commutative Geometry
1.8.5. Let X be a smooth manifold. A C°°(X)-module P is isomorphic to the structure module of a smooth vector bundle over X if and only if it is a projective module of finite rank. • THEOREM
Outline of proof. Let {(^,V'c)>/9Cc}' £» C = 1> • • • i^> D e a finite atlas of a smooth vector bundle Y —* X of fibre dimension m. Given a smooth partition of unity {/^} subordinate to the cover {U(}, let us set
It is readily observed that {£|} is also a partition of unity subordinate to {U^}. Every element s £ Y(X) defines an element of a free C°°(X)-module of rank m + k as follows. Let us put s% = ip£ o s\u( • It fulfils the relation a
€=Z>«'csC-
(1-8-14)
C
There are a module monomorphism F : Y(X) 3 s » (llSl,...,
lksk) G "©* C°°(X)
and a module epimorphism *: m © f c C°°(Jf) B (tu...
,tk) ^ (su ... ,sk) €Y(X),
3i=
Y^PH^i3
In view of the relation (1.8.14), 9oF
=
IdY{X),
i.e., Y(X) is a projective module of finite rank. The converse assertion is QED proved similarly to that in [430], Theorem 1.8.5 states the above mentioned categorial equivalence between the vector bundles over a smooth manifold X and projective modules of finite rank over the ring C°°{X) of smooth real functions on X. The following are corollaries of this equivalence • The structure module Y*(X) of the dual Y* —> X of a vector bundle Y -» X is the C°°(X)-dual Y(X)* of the structure module Y(X) of Y -> X. • Any exact sequence of vector bundles 0 -> y —>Y' —>Y" -> 0
(1.8.15)
80
Geometric and Algebraic Topological Methods in Quantum Mechanics
over the same base X yields the exact sequence 0 -» Y(X)
—* Y'(X)
—• Y"(X) -> 0
(1.8.16)
of their structure modules, and vice versa. In accordance with Theorem 10.6.5, the exact sequence (1.8.15) is always split. Every its splitting defines that of the exact sequence (1.8.16), and vice versa. • For instance, the derivation module of the M-ring C°°(X) coincides with the C°°(X)-module Ti(X) of vector fields on X, i.e., with the structure module of the tangent bundle TX of X. Hence, it is a projective C°°(X)-module of finite rank. It is the C°°(X)-dual TX(X) = O1{X)* of the structure module O1(X) of the cotangent bundle T*X of X which is the module of differential one-forms on X and, conversely, O1(X) = Tx{X)*. It follows that the Chevalley-Eilenberg differential calculus over the R-ring C°°(X) is exactly the differential graded algebra (O*(X),d) of exterior forms on X, where the Chevalley-Eilenberg coboundary operator d (1.6.6) coincides with the exterior differential. Accordingly, the de Rham complex (1.6.12) of the R-ring C°°(X) is the de Rham complex (1.8.9) of exterior forms on X. Moreover, one can show that (O*(X), d) is a minimal differential calculus, i.e., the C°° (X)-modu\e O1(X) is generated by elements df, f € C°°(X). Indeed, using the notation in the proof of Theorem 1.8.5, one can write
O\X) B <$> = £Z|0 = J^lfacb? = ^2(l^d(l^)
-
l^^dk),
(1.8.17) where (xM) are local coordinates on [/{ and /^xM and 1% are functions on X. Remark 1.8.7. Let us note that the above mentioned ChevalleyEilenberg differential calculus over the M-ring CCO(X) is a subcomplex of the Chevalley-Eilenberg complex of the Lie algebra T\{X) with coefficients in C°°(X). It consists of skew-symmetric morphisms of T\(X) to C°°(X) which are not only R-multilinear, but C°°(X)-multilinear. The ChevalleyEilenberg cohomology of smooth vector fields with coefficients in a trivial representation and in spaces of smooth tensor fields has been studied in detail [160] • • Let Y —> X be a vector bundle and Y(X) its structure module. The r-order jet manifold JrY of Y —> X consists of the equivalence classes j^s, x £ X, of sections s of Y —> X which are identified by the r + 1 terms of their Taylor series at points x £ X (see Section 10.7). Since Y —> X is a
81
Chapter 1 Commutative Geometry
vector bundle, so is the jet bundle JrY —» X. Its structure module JrY(X) of the C°°(X)-module Y(X) is exactly the r-order jet module Jr(Y(X)) in Section 1.2 [261]. As a consequence, the notion of a connection on the structure module Y(X) is equivalent to the standard geometric notion of a connection on a vector bundle Y —» X [296]. Indeed, connection on a fibre bundle Y —> X is defined as a global section F of the affine jet bundle JXY -> y . If V —> X is a vector bundle, there exists the exact sequence 0->T*X®Y—>JlY
—>y->0
(1.8.18)
x over X which is split by F. Conversely, any slitting of this exact sequence yields a connection Y —> X. The exact sequence of vector bundles (1.8.18) induces the exact sequence of their structure modules
o-»o 1 (x)®y(x) -->JlY{x) —>y(x)->o.
(1.8.19)
Then any connection F on a vector bundle Y —> X defines a splitting of the exact sequence (1.8.19) which, by Definition 1.3.1, is a connection on the C°°(X)-module Y(X), and vice versa. Let now P be an arbitrary C°°(X)-module. One can reformulate Definitions 1.3.2 and 1.3.3 of a connection on P as follows. DEFINITION 1.8.6. A connection on a C°°(X)-module P is a C°°(X)module morphism
V:P-*O1{X)<»P,
(1.8.20)
which satisfies the Leibniz rule V ( / p ) = # ® p + /V(p),
f€C°°(X),
pGP. D
DEFINITION 1.8.7. A connection on a C°°(X)-module P associates to any vector field r e 7i{X) on X a first order differential operator V r on P which obeys the Leibniz rule
V T (/p) = (Tj4f)p + / V r p .
(1.8.21)
• Since O1(X) = Ti(X)*, Definitions 1.8.6 and 1.8.7 are equivalent.
82
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us note that a connection on an arbitrary C°° (X)-module need not exist, unless it is a projective or locally free module (see Theorem 1.8.11 below). The curvature of a connection V in Definitions 1.8.6 and 1.8.7 is defined as the zero-order differential operator iJ(T,T') = [V T ,V T .]-V [T , T /]
(1.8.22)
on a module P for all vector fields T,T' e TX{X) on X (cf. (10.6.64)). Let us turn now to the notion of a connection on a local-ringed space. Let (X,d\) be a local-ringed space and ty a sheaf of 9"l-modules on X. For any open subset U C X, let us consider the jet module Jx{ty{U)) of the module ty{U). It consists of the elements of £H(C/) ® ^P(E7) modulo the pointwise relations (1.2.14). Hence, there is the restriction morphism
jlmu)) -
j\y(v))
for any open subsets V C U, and the jet modules ^(^(U)) constitute a presheaf. This presheaf defines the sheaf 3ls$ of jets of ty (or simply the jet sheaf). The jet sheaf O^SH of the sheaf SH of local rings is introduced in a similar way. Since the relations (1.2.14) and (1.3.1) on the ring W.(U) and modules ty{U), J71(?J(C/)), v71(fH([/)) are pointwise relations for any open subset U C X, they commute with the restriction morphisms. Therefore, the direct limits of the quotients modulo these relations exist [303]. Then we have the sheaf O 1 ^ of one-forms over the sheaf Dt, the sheaf isomorphism 0;1(q3) = (fH©C»1fH)®^, and the exact sequences of sheaves 0-*£> 1 9l
(1.8.23) (1.8.24)
They reflect the quotient (1.3.3), the isomorphism (1.3.10) and the exact sequences of modules (1.3.11), (1.3.13), respectively. Remark 1.8.8. It should be emphasized that, because of the inequality (1.7.1), the duality relation (1.3.5) is not extended to the sheaves 59^ and O 1 ^ in general, unless VfR and Ol are locally free sheaves of finite rank. If 93 is a locally free sheaf of finite rank, so is CJ1^D Following Definitions 1.3.1, 1.3.2 of a connection on modules, we come to the following notion of a connection on sheaves.
83
Chapter 1 Commutative Geometry
DEFINITION 1.8.8. Given a local-ringed space {X,9\) and a sheaf *p of Dt-modules on X, a connection on a sheaf ^3 is denned as a splitting of the exact sequence (1.8.23) or, equivalently, the exact sequence (1.8.24). •
Theorem 1.7.3 leads to the following compatibility of the notion of a connection on sheaves with that of a connection on modules. 1.8.9. If there exists a connection on a sheaf ?p in Definition 1.8.8, then there exists a connection on a module *P(t/) for any open subset U C X. Conversely, if for any open subsets V c U
• X be a vector bundle. Every linear connection Example 1.8.9. Let Y —> F on Y —> X defines a connection on the structure module Y(X) such that the restriction T\u is a connection on the module Y(U) for any open subset U C X. Then we have a connection on the structure sheaf Yx- Conversely, a connection on the structure sheaf Yx defines a connection on the module • Y(X) and, consequently, a connection on the vector bundle Y —» X. As an immediate consequence of Proposition 1.8.9, we find that the exact sequence of sheaves (1.8.24) is split if and only if there exists a sheaf morphism V :
(1.8.25)
satisfying the Leibniz rule V(/s) = d / ® s + /V(s),
f€A(U),
s£V(U),
for any open subset U € X. It leads to the following equivalent definition of a connection on sheaves in the spirit of Definition 1.3.2. DEFINITION
sheaf q3.
1.8.10. The sheaf morphism (1.8.25) is a connection on the
•
Similarly to the case of connections on modules, the curvature of the connection (1.8.25) on a sheaf ty is given by the expression R = V 2 :
(1.8.26)
The exact sequence (1.8.24) need not be split. One can obtain the following criteria of the existence of a connection on a sheaf.
84
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let if! be a locally free sheaf of !!H-modules. Then we have the exact sequence of sheaves 0 -» Horn («p, O1^ ® qj) - • Horn («p, (91 © O 1 ^ ) ® q?) - • Horn («JJ, «P) -> 0 and the corresponding exact sequence (1.7.16) of the cohomology groups 0 - • H°(X; Horn («JJ, O 1 ^
ff°(X;Hom(q},q3)).
Its
if1(X;Hom(«P,O19Il(8>q3)) is called the Atiyah class. If this class vanishes, there exists an element of Hom(qj ) (JR©O 1 lR)®qj)) whose image is Id?P, i.e., a splitting of the exact sequence (1.8.24). In particular, let X be a manifold and fH = Cjf the sheaf of smooth functions on X. The sheaf OC;-? of its derivations is isomorphic to the sheaf of vector fields on a manifold X. It follows that: • there is the restriction morphism D(C°°(U)) —> V(C°°(V)) for any open sets V C U, • QCx is a locally free sheaf of CjJ?-modules of finite rank, • the sheaves 3CJJ? and Olx are mutually dual. Let q3 be a locally free sheaf of C^-modules. In this case, Horn (^3, O\ ® <}3) is a locally free sheaf of C^-modules. It is fine and acyclic. Its cohomology group if 1 (X; Horn (qj,OJf®qj)) vanishes, and the exact sequence
O^C3c(g)«p^((7^©C>3f)«g»q3^
(1.8.27)
admits a splitting. This proves the following. PROPOSITION 1.8.11. Any locally free sheaf of C^-modules on a manifold X admits a connection and, in accordance with Proposition 1.8.9, any locally free C°°(X)-module does well. •
Chapter 1 Commutative Geometry
85
In conclusion, let us consider a sheaf S of commutative CjjP-rings on a manifold X. Basing on Definition 1.3.4, we come to the following notion of a connection on a sheaf S of commutative C^-rings. DEFINITION
1.8.12. Any morphism OC^?
3TMVTG5S,
which is a connection on S as a sheaf of CjJ?-modules, is called a connection on the sheaf S of rings. • Its curvature is given by the expression R(T,T')
= [Vr, VT.] - V[TiT.],
(1.8.28)
similar to the expression (1.3.20) for the curvature of a connection on modules. 1.9
Algebraic varieties
In this Section, we briefly sketch the relevant material on affine and algebraic varieties [397]. Throughout, K is an algebraically closedfield,i.e., any polynomial of non-zero degree with coefficients in K. has a root in K,. The reason is that, dealing with non-linear algebraic equations, one can not expect a simple clearcut theory, without assuming that a field is algebraically closed. If K. fails to be algebraically closed, one can extend it in an appropriate way. A subset of the n-dimensional affine space /Cn is called an affine variety if it a set of zeros (common roots) of some set of polynomials of n variables with coefficients in K.. Unless otherwise stated, the dimension n holds fixed. Let K.[x] be the ring of polynomials of n variables with coefficients in a field K. Given an affine variety V, the set I(V) of polynomials in K.[x] which vanish at every point of V is an ideal of JC[x] called the characteristic ideal of V. Herewith, V = V if and only if I(V) = I{V). Therefore, an affine variety V can be given by the generating set of its characteristic ideal I(V), i.e., by a finite system fc = 0 of polynomials /» € K\x\. An affine variety which is a subset of another affine variety is called a subvariety. An affine variety is said to be irreducible if it is not the union of two proper subvarities. A maximal irreducible subvariety of an affine variety is called its irreducible component. Note that any affine variety can
86
Geometric and Algebraic Topological Methods in Quantum Mechanics
be written uniquely as the union of a finite number of irreducible components. An affine variety V is irreducible if and only if /(V) is a prime ideal. Moreover, let I be a prime ideal of fC[x) and V be an affine variety in Kn of zeros of elements of X. Then I(V) = 1. This fact states one-to-one correspondence between the prime ideals of K.[x] and the irreducible affine varieties in K,n. In particular, maximal ideals correspond to points of Kn. The intersection and the union of subvarieties of an affine variety V are also subvarieties. Thus, subvarieties can be taken as the system of closed sets of a topology on V which is called the Zariski topology on the affine variety V. Unless otherwise stated, an affine variety is provided with this topology. Given an affine variety V, the factor ring
Kv = K\x]/I(V) is called the coordinate ring of V. One can think of K.y as being the ring of K,-valued functions on V such that, for any / e ICy, there exists a polynomial on Kn which equals / on V. If an affine variety V is irreducible, the ring /Cy has no divisor of zero. Remark 1.9.1. There is the following correspondence between the affine varieties and the affine schemes. Let a /C-ring A be finitely generated, and let it possess no nilpotent elements. This is the case of an algebraic scheme of finite type over a field K. in Example 1.8.2. One can associate to .4 the following affine variety. Given a set ( a i , . . . , an) of generating elements of A, let us consider the epimorphism <j> : K\x] —> A defined by the equalities 4>(xi) = ai. Zeros of polynomials in KeT(j> make up a variety whose coordinate ring is exactly A. Conversely, every affine algebraic variety V yields the affine scheme Spec/Cy such that there is one-to-one correspondence between the points of Spec/Cy and the irreducible subvarieties of V. • If V is an irreducible affine variety, its coordinate ring /Cy contains no divisor of zero. Then let Ry be the field of quotients of /Cy (see Remark 1.1.1). Its elements can be regarded as rational functions on the affine variety V, and .Ry is called the function field of V. There is the monomorphism /Cy —> Rv (1.1.1). The function field -Ry is finitely generated over /C, and its transcendence degree is called the dimension of the irreducible affine variety V. Let W be an irreducible subvariety of V and I(W) the subset of ICy consisting of elements which vanish on W. Then I(W) is a prime ideal of
Chapter 1 Commutative Geometry
87
/Cy. Let us consider the multiplicative subset K\> \ /(W) and the subring Rw = (Kv \ / ( W ) ) " 1 ^ of i?y (see Remark 1.1.1). It is called a local ring of a subvariety W. Functions / £ Ry\> C i?y are called regular at W. For a given function / € Rv, the set of points of V where / is regular is Zariski open. Given an open subset U C V, let us denote Ru the ring of regular functions on U. Assigning RJJ to each open set U, one can define a sheaf of rings Dty of germs of regular functions on V. Its stalk at a point x G V coincides with the local ring Rx. The sheaf 9lv is called the structure sheaf of the affine variety V. The pair (V,£Hv) is a local-ringed space. Let us consider a pair (X, 9\) of a topological space X and some sheaf Ui of germs of K-valued functions on X. This pair is called a prealgebraic variety if X admits a finite open cover {Ui} such that each Ui is homeomorphic to some affine variety Vi and 5K|[/i is isomorphic to the structure sheaf 91 Vi °f Vt- Let us note that the Cartesian product V x V of affine varieties V G /Cn and V € Km is an affine variety in /C n+m though the Zariski topology on V x V' is finer than the product topology. Since the Cartesian product X x X' of prealgebraic varieties is locally a product of affine varieties, this product is a prealgebraic variety. A prealgebraic variety (X, 0\) is said to be an algebraic variety if the diagonal map X —> X x X is closed in the Zariski topology of the product variety. This condition corresponds to Hausdorff's separation axiom. If W is a locally closed subset (i.e., the intersection of open and closed sets) of an algebraic variety, it becomes an algebraic variety in a natural manner since the germs of regular functions at x £ W are taken to be the germs of functions on W induced by functions in the stalk 9tx. The definitions of irreducibility and local rings of subvarieties for algebraic varieties are given in the same manner as before. From now on, by a variety is meant an algebraic variety. Any variety (X, !Eft), by definition, is a local-ringed space. Given a variety, one says that its point x is simple and that V is nonsingular or smooth at x if the local ring Rx of x is regular. Since the problem is local, one can assume that V is an affine variety in /C". Then the simplicity of x implies that x is contained in only one irreducible component of V and, if this component is r-dimensional, there exists n — r polynomials fi(x) in the characteristic ideal I(V) of V such that the rank of the matrix (dfi/dxj) at x equals n — r. A point of V which is not simple is called a singular point. The set of singular points, called the singular locus of V,
88
Geometric and Algebraic Topological Methods in Quantum Mechanics
is a proper closed subset of V. A variety possessing no singular point is said to be smooth. A point x of a variety V is called normal if the local ring R(x) is normal. A simple point is normal. Normal points make up a non-empty open subset of V. An irreducible variety whose points are all normal is called a normal variety. Example 1.9.2. When K = C, an algebraic variety V, called a complex algebraic variety, has the structure of a complex analytic manifold so that the stalk fHy.i = Rx at x 6 V contained in the stalk C v of germs of holomorphic functions at x and their completion coincide. If x is a simple point, C v x is the ring of converged power series and its completion is the ring of formal power series. • A regular morphism (V, fK) —* (V', dM) of varieties over the same field fC is defined as a morphism of local-ringed spaces
$ :
where ip* is the pull-back onto V of K.-valued functions on V. An isomorphism of varieties is also called a biregular morphism. Let V and W be irreducible varieties. Let a closed subset T C V x W be an irreducible variety such that the closure of the range of the projection T —> V coincides with V. Then the function field Ry of V can be identified with a subfield of RT- If Ry = RT, then T is called a rational morphism of V to W. One can show that, if T : V —> W is a rational morphism and x G V is a normal point of V such that T(x) contains an isolated point, then T is regular at x. Let V be an m-dimensional irreducible affine variety. One can associate to V the two algebras Diff * (/Cy) and Diff«(R\>) of (linear) differential operators on the coordinate ring /Cy and the function field Rv of V, respectively [310]. In this case of the function field Ry, one can choose a separating transcendence {x1,... ,xm} basis for Ry over K. Let us consider the derivation module X)Ry of the /C-field Ry. It is finitely generated by the derivations di of -Ry such that di{x^) = 5\. Moreover, any differential operator A e Diff r{Ry) is uniquely expressed as a polynomial of 9, with coefficients in Ry. Let DRV be the i?y-dual of the derivation module dRy. Given the above mentioned transcendence basis for Ry, it is finitely generated by the elements dxi which are the duals of di. As a consequence, the ChevalleyEilenberg calculus over Ry coincides with the universal differential calculus O*Ry over Ry.
Chapter 1 Commutative Geometry
89
The case of the coordinate ring /Cy is more subtle. In this case, Diff»(/Cy), denoted by Diff,(V), is called the ring of differential operators on V. In general, there are no global coordinates on V, but if fC is of characteristic zero and V is smooth, the structure of Diff »(V) is still well understood. Namely, Diff*(V) is a simple (left and right) Noetherian ring without divisors of zero, and it is generated by finitely many elements of Diffi(V). If V is singular, the construction of Diff «(V) is less clear. One can show that any differential operator on /Cy C fly admits a unique extension to a differential operator of the same order on i?y. Thus, one can regard Diff »(V) as a subalgebra of Diff»(fly). Furthermore, a differential operator A on .Ry which preserves /Cy is a differential operator on /Cy. In particular, it follows that Diff *(V) has no zero divisors. In contrast with the smooth case, Diff *(V) fails to be generated by elements of Diff i(V) in general. One has conjectured that this is true if and only if V is smooth [334]. This conjecture has been proved for algebraic curves [323] and, more generally, for varieties with smooth normalization [416]. Let us note that, if a field K is of positive characteristic, the ring Diff *(V) is not Noetherian, or finitely generated, or without zero divisors [400].
Chapter 2
Classical Hamiltonian systems
This Chapter summarizes the relevant material on geometry of Poisson manifolds and classical Hamiltonian systems quantized in the sequel. 2.1
Geometry and cohomology of Poisson manifolds
Subsections: A. Symplectic manifolds, 91; B. Presymplectic manifolds, 96; C. Poisson manifolds, 97; D. Symplectic and Poisson reductions, 103; E. Koszul-Brylinski-Poisson homology, 108; F. Lichnerowicz-Poisson cohomology, 109. We start with symplectic manifolds. Every symplectic manifold is a regular non-degenerate Poisson manifold, and vice versa. A. Symplectic manifolds Let Z b e a smooth manifold. Any exterior two-form Q on Z yields the linear bundle morphism & :TZ ->T*Z, z
ftb
:v^-v\Sl(z),
v &TZZ,
z G Z.
(2.1.1)
One says that D, is of rank r if the morphism (2.1.1) has the rank r. The kernel Kerfi of fi is defined as the kernel of the morphism (2.1.1). If ft is of constant rank, its kernel is a subbundle of the tangent bundle TZ. In particular, Ker Q contains the canonical zero section 0 of TZ —> Z. If Kerfi = 0 (one customarily writes Ker ft = 0), a two-form Q. is said to be non-degenerate. It is called an almost symplectic form. Equipped with such a form, a manifold Z becomes an almost symplectic manifold. It is never odd-dimensional. Unless otherwise stated, we put dhnZ = 2m. In 91
92
Geometric and Algebraic Topological Methods in Quantum Mechanics
accordance with Theorem 10.10.5, a 2m-dimensional manifold Z admits an almost symplectic form if and only if the structure group GL(2m, R) of the tangent bundle TZ of Z is reducible to the symplectic group Sp(2m, R). A closed almost symplectic form is called symplectic. Accordingly, a manifold equipped with a symplectic form is a symplectic manifold. A symplectic manifold (Z, 0) is orientable. It is usually oriented so that A Q. is a volume form on Z, i.e., it defines a positive measure on Z. Remark 2.1.1. As concerns the existence of a symplectic structure, there is a strong dichotomy between compact and non-compact manifolds. If a non-compact manifold admits an almost symplectic form, it is provided with a symplectic form, too. Namely, every almost symplectic form on a non-compact manifold is homotopic among almost symplectic forms to a symplectic form [200]. At the same time, there is no sufficient conditions of the existence of a symplectic structure on a compact manifold [35]. For example, there exist homeomorphic pairs of smooth four-manifolds with isomorphic tangent bundles such that one admits symplectic structures, while the other does not [413]. In addition to an almost symplectic form, a compact symplectic manifold (Z, O) necessarily possesses the non-zero de Rham classes w[fi] = [Afi] €Hr{Z),
r = l,...,m.
Therefore, the even de Rham cohomology groups of a compact symplectic manifold are non-trivial and, consequently, its even Betty numbers differ from zero. More subtle conditions, involving the Seiberg-Witten invariants, have been discovered in [413]. • A manifold morphism £ of a symplectic manifold (Z, fi) to a symplectic manifold (Z', Q,') is called a symplectic morphism if O = £*Q'. Any symplectic morphism is an immersion. One should distinguish a symplectic morphism from a symplectomorphism in the terminology of [279] which is a symplectic isomorphism. A vector field u on a symplectic manifold (Z, £1) is an infinitesimal generator of a local one-parameter group of symplectic automorphisms if and only if the Lie derivative Lufi vanishes. It is called a canonical vector field. A canonical vector field u on a symplectic manifold (Z, fi) is said to be Hamiltonian if the closed one-form u\Q is exact. Any smooth function / £ C°°(Z) on Z defines a unique Hamiltonian vector field $/ on Z such
Chapter 2 Classical Hamiltonian Systems
93
that tf/jn = -4f,
4f = n*W),
(2.1.2)
where fi" is the inverse isomorphism to & (2.1.1). Remark 2.1.2. There is another convention [l], where a Hamiltonian vector field differs in the minus sign from (2.1.2). D Example 2.1.3. Given a manifold Q coordinated by (q1), let
TT,Q :T*Q^Q be its cotangent bundle equipped with the holonomic coordinates (ql,pi — qi). It is endowed with the canonical Liouville form
6 = pM and the canonical symplectic form n = dd = dpif\dqi.
(2.1.3)
Their coordinate expressions are maintained under holonomic coordinate transformations. The Hamiltonian vector field -df (2.1.2) with respect to the canonical symplectic form (2.1.3) reads 0/ = &fdi - djd\
(2.1.4)
Of course, f2 (2.1.3) is not a unique symplectic form on the cotangent bundle T*Q. Given a closed two-form ^ o n a manifold Q and its pull-back ~K*Q(J> onto T*Q, the form Q.^ = n + nlQ(j) is also a symplectic form on T*Q.
(2.1.5) •
Example 2.1.4. Given a symplectic form Q on a manifold Z, the twoform —ft is also a symplectic form on Z. • The canonical symplectic form (2.1.3) plays a prominent role in view of the classical Darboux theorem [279].
94
Geometric and Algebraic Topological Methods in Quantum Mechanics
THEOREM 2.1.1. Each point of a symplectic manifold (Z, fl) has an open neighborhood equipped with coordinates (qi,Pi), called canonical or Darboux coordinates, such that fl takes the coordinate form (2.1.3). • Similarly to a Riemannian manifold endowed with the Levi-Civita connection, any symplectic manifold can be provided with a symplectic connection. It is a symmetric linear connection preserving a symplectic form. In order to characterize symplectic connections in full, let us start with connections on an almost symplectic manifold (Z,Q.). Since Q is associated to a reduced 5p(2m)-structure, the well-known theorem [250] states the existence of a linear connection K* (10.6.66) on T*Z which preserves Q, i.e., V>frap = dxSla0 + /fA"afiM/3 + Kx»pnQll = 0.
(2.1.6)
We refer the reader to [424] for the complete picture of connections preserving a given almost symplectic form. PROPOSITION 2.1.2. (i) An almost symplectic form, parallel with respect to a symmetric connection, is symplectic. (ii) A symplectic connection exists on any symplectic manifold. •
Outline of proof. Substituting the equality (2.1.6) into the expression for dQ,, one easily justifies that dCl = 0 if a connection is symmetric. Given a symplectic form 0, a symplectic connection preserving fi can be constructed locally by putting its coefficients zero relative to the Darboux coordinates. Then one can glue a global symplectic connection by a partition of unity. QED Given a symplectic connection K, its covariant coefficients kx^ = ^aKxau
(2.1.7)
are introduced. Relative to the local Darboux coordinates, these coefficients are totally symmetric. Let K' be another symplectic connection with covariant coefficients k'x . Then the difference k'x — k\^u is a totally symmetric tensor field, and this property is coordinate-independent. Thus, it has shown the following [57; 165; 424]. THEOREM 2.1.3. Symplectic connections on a symplectic manifold (Z,Q.) are assembled into an affine space modelled over the vector space of sections • of the tensor bundle VT*Z.
Chapter 2 Classical Hamiltonian Systems
95
Given the curvature tensor R of a symplectic connection, one introduces the covariant curvature tensor
which obeys the relations rx^.u/3 = rx^pu,
r\pVp + r^px + rvpx^ + rpx^u = 0.
Furthermore, the covariant curvature tensor rx^vp is split in a unique fashion into two parts r = E + W, irreducible under the action of the group 5p(2m, R), where E = 0 if and only if the Ricci tensor vanishes [57; 424]. Accordingly, one speaks either on an Einstein symplectic manifold if E = 0 or a simple symplectic manifold if W = 0. Let us turn to special submanifolds of a symplectic manifold. These are coisotropic, symplectic, isotropic and Lagrangian submanifolds. Let in : ./V —> Z be a submanifold of a symplectic manifold (Z,il). The subset OrthnTTV := ( J {v € TZZ : v\u\ti = 0, u £ TZN]
(2.1.8)
zeN
of TZ\N is called orthogonal to TN relative to the symplectic form Q. or, simply, the il-orthogonal space to TN. There are the following bijections Orth n (OrthnTA0 = TN C TZ\N, n b (Orth n TA0 = AnnT./V C T*Z\N, Ann (OrthfiTA0 = fib(TiV) C T*Z\N. If Ni and N2 are two submanifolds of Z, then TNi C TA^2 implies OrthnTiV! D OrthnTN2 over A^i n N2, and wee versa. We also have OrthnCTWi n TiV2) = O r t h n T M t a n ^ + Orth n TAr 2 | WinJ v 2 , TA^ n OrthnTAT = Orth n (Orth n TAf + TN). It should be emphasized that TN n OrthnTAT ^ 0,
TZ\N ^TN + OrthnTN,
96
Geometric and Algebraic Topological Methods in Quantum Mechanics
in general. The first set is exactly the kernel of the pull-back £lN = i*NCl of the symplectic form fi onto the submanifold N. One considers the following special types of submanifolds of a symplectic manifold such that this pull-back is of constant rank. A submanifold N of Z is said to be: • coisotropic if OrthnT7V C TN, dim N > m; • symplectic if Cl^ is a symplectic form on JV; • isotropic if TN C OrthnTiV, dim N <m; • Lagrangian if N is both coisotropic and isotropic, i.e., OrthnTV = TN, dim N = m. Clearly, Jl^ = 0 if N is isotropic. A submanifold of dimension 1 is always isotropic, while that of codimension 1 is coisotropic. We refer the reader to [131] for classification of germs of special submanifolds of a symplectic manifold. B. Presymplectic manifolds An exterior two-form w on a manifold Z is said to be presymplectic if it is closed, but not necessarily non-degenerate. A manifold equipped with a presymplectic form is called presymplectic. Example 2.1.5. Let (Z, ft) be a symplectic manifold and %N : N —> Z its coisotropic submanifold. Then i*ND. is presymplectic form on N. D The kernel Ker wofa presymplectic form u> of constant rank is an involutive distribution, called the characteristic distribution [279]. It defines the characteristic foliation of a presymplectic manifold {Z,LJ). The pull-back of the presymplectic form w onto any leaf of this foliation equals zero. The notion of a Hamiltonian vector field on a symplectic manifold is extended in a straightforward manner to a presymplectic manifold. However, a function on a presymplectic manifold need not admit an associated Hamiltonian vector field. Any presymplectic form has a symplectic realization, i.e., can be represented as the pull-back of a symplectic form. Indeed, a presymplectic form u> on a manifold Z is the pull-back
of the symplectic form fiw (2.1.5) on the cotangent bundle T*Z of Z by its zero section 0. It is easily justified that the zero section 0(Z) c T*Z is a coisotropic submanifold with respect to the symplectic form fiw on T*Z.
Chapter S Classical Hamiltonian Systems
97
Therefore, the morphism 0 of the presymplectic manifold (Z, UJ) into the symplectic manifold (T*Z,fiw)exemplifies the coisotropic imbedding. This construction can be refined as follows. If a presymplectic form is of constant rank, it admits the following symplectic realization [187; 188]. 2.1.4. Given a presymplectic manifold (Z, w) where u is of constant rank, there exists a symplectic form on a tubular neighborhood of the zero section 0 of the dual bundle (Kerw)* to the characteristic distribution Kerw —> Z such that (Z, u>) can be coisotropically imbedded onto PROPOSITION
0(Z).
•
If the characteristic foliation of a presymplectic form is simple, there is another important variant of symplectic realization, namely, along the leaves of this foliation [189; 423]. PROPOSITION 2.1.5. Let a presymplectic form w o n a manifold Z be of constant rank, and let its characteristic foliation be simple, i.e., a fibred manifold n : Z —* P. Then the base P of this fibred manifold is equipped with a symplectic form Q, such that u> is the pull-back of Q, by n. •
C. Poisson manifolds A Poisson bracket on the ring C°°{Z) of smooth real functions on a manifold Z (or a Poisson structure on Z) is defined as an R-bilinear map C°°(Z) x C°°(Z) 3 (f,g) ~ {f,g} G C°°(Z) which satisfies the following conditions: • {gJ} = ~{f,g} (skew-symmetry); • {/. (5, h}} + {g, {h, /}} + {h, {f,g}} = 0 (the Jacobi identity); • {h, fg] = {h, f}g + f{h, g] (the Leibniz rule). A manifold Z endowed with a Poisson structure is called a Poisson manifold. A Poisson bracket makes C°°(Z) into a real Lie algebra, called the Poisson algebra. A Poisson structure is characterized by a particular bivector field as follows. THEOREM
2.1.6.
Every Poisson bracket on a manifold Z is uniquely
defined as {/, /'} = w(df, df) = w^drfdyf
(2.1.9)
98
Geometric and Algebraic Topological Methods in Quantum Mechanics
by a bivector field w whose Schouten-Nijenhuis bracket It is called a Poisson bivector field [426].
[W,U;]SN
vanishes. •
Example 2.1.6. Any manifold admits the zero Poisson structure characterized by the zero Poisson bivector field w = 0. •
Example 2.1.7. Given a Poisson bivector field w on & manifold X, the bivector field — w is also a Poisson bivector field on X. • Example 2.1.8.
Let vector fields u and v on a manifold Z mutually
commute. Then u A v is a Poisson bivector
field.
•
Any bivector field w o n a manifold Z yields the linear bundle morphism ->TZ, wi :aH-«,(z)[a, a£T*zZ. (2.1.10) z One says that w is of rank r if the morphism (2.1.10) is of this rank. If a Poisson bivector field is of constant rank, the Poisson structure is called regular. Unless otherwise stated, only regular Poisson structure are considered. A Poisson structure determined by a Poisson bivector field w is said to be non-degenerate if w is of maximal rank. wt :T*Z
Remark 2.1.9. The morphism (2.1.10) is naturally generalized to the homomorphism of graded commutative algebras O*(Z) —> T*(Z) in accordance with the relation w«(0)(<7i,..., ar) = (-l) r 0(«;«(cri),.. .,w*(ar)),
<j) e O r (Z),
o-i G 0\z).
It is an isomorphism if the bivector field w is non-degenerate.
•
There is one-to-one correspondence Qw <-> u>a between the almost symplectic forms and the non-degenerate bivector fields which is given by the equalities wn(
(2.1.11a) (2.1.11b)
where the morphisms w^ (2.1.10) and Q^ (2.1.1) are mutually inverse, i.e.,
Chapter 2 Classical Hamiltonian Systems
99
Furthermore, one can show that there is one-to-one correspondence between the symplectic forms and the non-degenerate Poisson bivector fields. However, this correspondence need not be preserved under morphisms. Namely, let (Zi,wi) and (Z2,W2) be Poisson manifolds. A manifold morphism g : Z\ —> Z2 is said to be a Poisson morphism if {/ OQj'o g}, = {/, f'}2 o g,
/ , / ' G C°°(Z2),
or, equivalently, if w2 = Tg o w\ where TQ is the tangent map to g. Herewith, the rank of w\ is superior or equal to that of W2- Therefore, there are no pull-back and push-forward operations of Poisson structures in general. Nevertheless, let us mention the following construction [426]. THEOREM 2.1.7. Let (Z,w) be a Poisson manifold and 7r : Z —> Y a fibration such that, for every pair of functions (/, g) on Y and for each point y EY, the restriction of the function {n*f, n*g} to the fibre TT~1(y) is constant, i.e., {TT*/, n*g} is the pull-back onto Z of some function on Y. Then there exists a coinduced Poisson structure w' on Y for which IT is a Poisson morphism. •
Example 2.1.10. The direct product Z x Z' of Poisson manifolds (Z, w) and (Z',w') can be endowed with the direct product of Poisson structures, given by the bivector field w + w' such that the surjections prx and pr2 are Poisson morphisms. • Example 2.1.11. Let (Zi.fii) and (Z 2 ,O 2 ) be symplectic manifolds equipped with the associated non-degenerate Poisson structures Wi and w2. If dimZi > dimZ 2 , a Poisson morphism g : Z\ —> Z2 need not be a symplectic one, i.e., w2 = Tg o wi and fli j= g*Q2. D A vector field u on a Poisson manifold (Z, w) is an infinitesimal generator of a local one-parameter group of Poisson automorphisms if and only if the Lie derivative Luw = [u,w]SN
(2.1.12)
100
Geometric and Algebraic Topological Methods in Quantum Mechanics
vanishes. It is called a canonical vector field for the Poisson structure w. In particular, for any real smooth function / on a Poisson manifold (Z,w), let us put 0, = w^df) = -w[df = w^dpfdv.
(2.1.13)
It is a canonical vector field, called the Hamiltonian vector field of a function f with respect to the Poisson structure w. Hamiltonian vector fields fulfil the relations (2-1.14)
{f,9} = #f\d9, [tf/Al = %,<,},
f,geC°°(Z).
(2.1.15)
For instance, the Hamiltonian vector field df (2.1.2) of a function / on a symplectic manifold (Z, O) coincides with that (2.1.13) with respect to the corresponding Poisson structure wu- The Poisson bracket defined by a symplectic form fi reads
{/.0} = iV|0/Jfl. Since a Poisson manifold (Z, w) is assumed to be regular, the range T = w8(T*Z) of the morphism (2.1.10) is a subbundle of TZ called the characteristic distribution on (Z,w). It is spanned by Hamiltonian vector fields, and is involutive by virtue of the relation (2.1.15). It follows that a Poisson manifold Z admits local adapted coordinates in Theorem 10.6.7. Moreover, one can choose particular adapted coordinates which bring the Poisson structure into the following canonical form [426]. THEOREM 2.1.8. For any point z of a Poisson manifold (Z,w), there exists a coordinate system
(z\...,zk-2m,q\...,qm,Pl,...,pm)
(2.1.16)
in a neighborhood of z such that (2.1.17)
• The coordinates (2.1.16) are called the canonical or Darboux coordinates for the Poisson structure w. The Hamiltonian vector field of a function / written in this coordinates is
•81 = &fdi - difd\
Chapter 2 Classical Hamiltonian Systems
101
Of course, the canonical coordinates for a symplectic form Cl in Theorem 2.1.1 are also canonical coordinates in Theorem 2.1.8 for the corresponding non-degenerate Poisson bivector field w, i.e., w = di A d».
Q = dpi A dq\
With respect to these coordinates, the mutually inverse bundle isomorphisms Qb (2.1.1) and wi (2.1.10) read n b : v'di + vidi H-> -Vidj + v^pi, w* : Vidq1 + v*dpi i-> v% - Vid\ Given a Poisson manifold (Z,w) and its characteristic distribution T, the above mentioned notions of coisotropic and Lagrangian submanifolds of a symplectic manifold are generalized to a Poisson manifold as follows. A submanifold N of a Poisson manifold is said to be: • coisotropic if w" (Ann TN) C TN, • Lagrangian if W* (Ann TN) = TN n T. Integral manifolds of the characteristic distribution T of a Poisson manifold {Z,w) constitute a (regular) foliation T of Z whose tangent bundle TT is T. It is called the characteristic foliation of a Poisson manifold. By the very definition of the characteristic distribution T = TT, the Poisson 2
bivector field w is subordinate to ATJ-. Therefore, its restriction w\p to any leaf F of J- is a non-degenerate Poisson bivector field on F. It provides F with a non-degenerate Poisson structure {, } F and, consequently, a symplectic structure. Clearly, the local Darboux coordinates for the Poisson structure w in Theorem 2.1.8 are also the local adapted coordinates I \
\
z
i---iz
k-2m
iz
i_
i
— H iz
m+i _
—Pi)i
\
• _ ,
1 — 1 , . . . , 771,
(10.6.27) for the characteristic foliation T, and the symplectic structures along its leaves read Q.F = dpi A dqx.
Remark 2.1.12. Provided with this symplectic structure, the leaves of the characteristic foliation of a Poisson manifold Z are assembled into a symplectic foliation of Z. Moreover, there is one-to-one correspondence between the symplectic foliations of a manifold Z and the Poisson structures on Z (see next Section). •
102
Geometric and Algebraic Topological Methods in Quantum Mechanics
Since any foliation is locally simple, a local structure of an arbitrary Poisson manifold reduces to the following [426; 436]. 2.1.9. Each point of a Poisson manifold has an open neighborhood which is Poisson equivalent to a product of a manifold with the zero Poisson structure and a symplectic manifold. D
THEOREM
In particular, this local structure of a Poisson manifold is used in order to prove that every Poisson manifold (Z, w) admits a Poisson connection, i.e., a symmetric linear connection which preserves the Poisson bivector field w [426]. Such a connection is locally a product of a symplectic connection and a symmetric connection on a manifold with the zero Poisson structure. Let (Z, w) be a Poisson manifold. By its symplectic realization is meant a symplectic manifold (Z',Q,) together with a Poisson morphism Z' —» Z which is a surjective submersion. 2.1.10. Each point of a Poisson manifold has an open neighborhood which is realizable by a symplectic manifold. • THEOREM
Outline of proof. In local Darboux coordinates, this symplectic realization is described as follows. The Poisson structure given by the Poisson bracket (2.1.17) with respect to the canonical coordinates is coinduced from the symplectic structure given by the symplectic form Cl = dpi A dq% + dz\ A dzx
with respect to the coordinates (z1,.
. . , Zk~2m,ZU.
. . , Zfc- 2 m ,
• • • ,Pm)
by the surjection (zx,lx,q\Pi)
^
{zx,q\pi). QED
Remark 2.1.13. It follows from Theorem 2.1.9 that each point of a Poisson manifold has an open neighborhood which is a presymplectic manifold with respect to the presymplectic form Q = dpi A dq%
Chapter 2 Classical Hamiltonian Systems
103
written relative to the local Darboux coordinates (zx,q\pi). Moreover, let the direct product in Theorem 2.1.9 be global, i.e., a Poisson manifold (Z, w) is the Poisson product Z = P x Y of a symplectic manifold (P, fl) and a manifold Y with the zero Poisson structure. Then Z is provided with the presymplectic form prjfi. Conversely, let the characteristic foliation 7T: Z —> P of a presymplectic form u on a manifold Z in Proposition 2.1.5 be a trivial bundle Z = P xY. Then Z is a Poisson manifold given by the Poisson product of the symplectic manifold (P, fi) and Y" equipped with the zero Poisson structure. • D. Symplectic and Poisson reductions By a symplectic reduction of a symplectic manifold (Z, Q) is meant a fibration vr : N —> P of a submanifold N C Z over a symplectic manifold (P, Op) such that the relation 7T*nP = eNa holds [l; 279]. One says that (P,flp) is the reduced symplectic manifold of the symplectic manifold (Z, Q.) via the submanifold N. The submanifold (TV, i^fi) is presymplectic. Therefore, Proposition 2.1.5 shows that a symplectic reduction via N may take place if the presymplectic form Cl^ = i*N£l on N is of constant rank and its characteristic foliation is simple. These conditions appear to be both necessary and sufficient as follows [279]. 2.1.11. Let (Z,il) be a symplectic manifold and N a submanifold of Z. (i) If TV : N —> P is a symplectic reduction, then: • the pull-back presymplectic form f2jv on N is of constant rank equal to dim P; • KeiflN = VN; • the connected components of fibres of TT are the leaves of the characteristic foliation of OJVFurthermore, if fibres of TT are connected, the characteristic foliation of OJV is simple, and it is precisely the fibration N —> P. (ii) Conversely, if the presymplectic form HM on a submanifold TV is of constant rank and its characteristic foliation is simple, the base P of the corresponding fibration N —> P has a unique symplectic form such that this fibration is a symplectic reduction of (Z, O) via N. This is a unique symplectic reduction via N with connected fibres. Furthermore, if N —> P' THEOREM
104
Geometric and Algebraic Topological Methods in Quantum
Mechanics
is another symplectic reduction, there exists a surjection P —> P' which is a local symplectic isomorphism. • A coisotropic submanifold N provides the most interesting case of a symplectic reduction when the presymplectic form QN is of constant rank which is neither dim N nor 0. Example 2.1.14. Let (N,w) be a presymplectic manifold, where a presymplectic form u> is of constant rank and its characteristic foliation is simple. Let (Z, fi) be its symplectic realization by means of a coisotropic imbedding N —» Z (see Proposition 2.1.4). Then, by virtue of Proposition 2.1.5, there is the symplectic reduction of (Z, 0.) via the coisotropic submanifold N. D The reduction procedure is extended to Poisson manifolds as follows [300]. By a reductive structure on a Poisson manifold (Z, w) is meant a triple (Z, N, E) of a submanifold N of Z, a vector subbundle E C TZ\N, and a fibration ir: N —» P if the following conditions are satisfied: • Ed TN is tangent to the fibres of the submersion n; • if df and dg where f,g£ C°°{Z) belong to AnnE, so does d{f,g}w. A reductive structure is called a Poisson reduction if: (i) P is a Poisson manifold with a Poisson bivector field W, (ii) for any local functions / , g on P and for any local extensions f,]j onto Z of the pull-backs / o TT, g o n such that df, dg C Ann E, the relation {/, 9}w ° IN = {/, g}w ° 7T holds. One says that (P, W) is the reduced Poisson manifold of (Z, w) via (IV, E). The following are the necessary and sufficient conditions of a reductive structure on a Poisson manifold to be a Poisson reduction [300; 426]. PROPOSITION 2.1.12. Let (Z,N,E) be a reductive structure on a Poisson manifold (Z,w). This is a Poisson reduction if and only if
w»(Ann£;) CTN + E.
(2.1.18)
• A Poisson reduction possesses the following functorality property [300].
Chapter 2 Classical Hamiltonian Systems
105
PROPOSITION 2.1.13. Let (Z,N,E) and (Z',N',E') be Poisson reductions, and let $ : Z -> Z' be a Poisson map such that §(N) C iV', T$(E) C £', and $ sends the leaves of N to the leaves of N'. Then $ induces a unique Poisson map $ : P —> P', called the reduction of $, such that TT'O<3> = $o7r. D
Remark 2.1.15. In order to compare symplectic and Poisson reductions, let (Z, fl) be a symplectic manifold and w the corresponding Poisson bivector field on Z. Let iV be a submanifold of Z such that the presymplectic form 0 ^ = *jv^ is of constant rank. Put E = OrthTTV. Then EDTN = KerQ,N is tangent to the characteristic foliation of fi^. If there exists a symplectic reduction 7r : iV —» P of (Z, £l) to (P, Op), one can show that (Z,N,OrthTN) is a reductive structure on the Poisson manifold (Z,w). Indeed, the condition (i) of a reductive structure holds since the integral manifolds of the distribution Ker f2/v are connected components of fibres of N —> P. Furthermore, for a symplectic manifold, df € Ann (OrthTiV) if and only if the Hamiltonian vector field •df for / belongs to w* (Ann (OrthTTV) ) = TN.
(2.1.19)
Then i?/,tfg € TN implies {*f,*g} = *{f,g}ZTN. Hence, the condition (ii) of a reductive structure is also satisfied. Moreover, the injection (2.1.18) obviously takes place due to (2.1.19). It follows that (Z, N, OithTN) is a Poisson reduction and the corresponding reduced structure W on P is associated to the symplectic form fip. On the other hand, if we put E = Ker ClN, the injection (2.1.18) implies TV to be coisotropic, since w^AnnE) = TN + OrthTiV, and one can apply the previous result.
•
The construction in Remark 2.1.15 can be generalized to Poisson manifolds as follows [426].
106
Geometric and Algebraic Topological Methods in Quantum Mechanics
LEMMA
2.1.14. Let N be a. submanifold of a Poisson manifold (Z,w) such
that C(N) :=w*(AnnTN) nTN
(2.1.20)
is a distribution on N. Then C(N) is involutive. Furthermore, if N is transversal to the leaves of the symplectic foliation T of w, the distribution • C(N) defines a transversally symplectic subfoliation of T D N. If the foliation determined by the distribution C(N) (2.1.20) is simple, i.e., is a fibred manifold n : N —> P, then Ff)TN projects onto a symplectic foliation ir(F) of P which provides P with a Poisson structure. This is called the leafwise reduction of (Z,w) via N. It follows in the same way as for the symplectic case in Remark 2.1.15 that, if to"(AnnTN) is of constant dimension, then (Z,N,w$(AnnTN)) is a reductive structure. Due to l o ^ A n n ^ A n n T W ) ) ) C TN and the relation (2.1.18), the reductive triple (Z, JV, io"(AnnTW)) is also a Poisson reduction, while the reduced Poisson manifold P is exactly the one defined by the leafwise reduction. Given a Poisson manifold (Z,w), let (P, W) be a reduced Poisson manifold of (Z, w) via a submanifold N. Then one can say that the Poisson algebra C°°{P) of smooth real functions on P is a reduction of the Poisson algebra C°°(Z) on Z. Therefore, if AT is a closed imbedded submanifold of Z, this reduction can be described in an algebraic way, namely, in terms of ideals of the Poisson algebra C°°(Z) as follows [245]. Given a closed imbedded submanifold N of a Poisson manifold (Z,w), let us consider the set IN = Keri^ C C°°(Z)
(2.1.21)
of functions f on Z which vanish on N, i.e., i*Nf = 0. It is an ideal of the R-ring C°°(Z). Then, since iV is a closed imbedded submanifold of Z, we have the ring isomorphism C°°(Z)/IN S C°°(N).
(2.1.22)
Let us consider the space of all vector fields u on Z restrictable to vector fields on N, i.e., u\N C TN. It is TN = {u e T(Z) : u\df €lN,
/ e IN}.
(2.1.23)
Chapter 2 Classical Hamiltonian Systems
107
Then we obtain at once that the Hamiltonian vector field df of a function / on Z belongs to T/v if and only if $f\dg = {f,g] &IN,
9&IN-
Hence, the functions whose Hamiltonian vector fields are restrictable to vector fields on TV constitute the set I(N) := {/ G C°°(Z) : {f,g} G IN, 9 G IN),
(2.1.24)
called the normalizer of I^. Owing to the Jacobi identity, the normalizer (2.1.24) is a Poisson subalgebra of C°°(Z). Let us put r(N):=I(N)nIff.
(2.1.25)
This is a Poisson subalgebra of I(N) which is non-zero since I2 C I'(N) by virtue of the Leibniz rule. The following Theorem returns us to the Poisson reduction procedure [245]. THEOREM 2.1.15. Let N be a closed imbedded submanifold of a Poisson manifold Z. Let us suppose that W*(AnnTN) and C(N) (2.1.20) are of constant dimension, and that the foliation determined by C(N) is simple, i.e., a fibred manifold N —> P. Then (Z, ./V,iuit(AnnT./V)) is a reductive structure such that there is the ring isomorphism
C°°(P) S I(N)/I'(N).
(2.1.26)
Since the quotient in the right-hand side of this isomorphism is a Poisson algebra, we obtain a Poisson structure on P. • This result is based on the fact that all sections of w* (Ann TN) —+ N are the restriction to N of Hamiltonian vector fields of elements of IN , while all sections of C(N) —* N are the restriction to N of Hamiltonian vector fields of elements of I'(N). In particular, if N is coisotropic, IN C I(N), i.e., IN = I'(N) is a Poisson subalgebra of C°°(Z). Theorem 2.1.15 leads to the following algebraic definition of Poisson reduction [245]. DEFINITION 2.1.16. Let P b e a Poisson algebra on a manifold, J an ideal of V as an associative algebra, J" its normalizer (2.1.24), and J ' = J"C\J'. One says that the Poisson algebra J"/J' is the reduction of the Poisson • algebra V via the ideal J.
108
Geometric and Algebraic Topological Methods in Quantum Mechanics
In particular, an ideal J of a Poisson algebra V is said to be coisotropic if J is a Poisson subalgebra of V. E. Koszul—Brylinski—Poisson homology Given a Poisson manifold (Z,w), let us consider the operator 5: Or{Z) -» Or~\Z),
(2.1.27)
5w:=w\od-dow\, r
S(fodfi A • • • A dfr) = ] T ( - i y + 1 { / o , / i K i A • • • A dfi A • • • A dfr + i=l
53
( - l ^ / o ^ / i , fj} A dA A • • • A df, A • • • A dfj A • • • A dfr,
l
on the differential algebra O*(Z) of exterior forms on Z. It is called the Poisson codifferential [72; 426]. The operator 5 (2.1.27) is nilpotent and obeys the relation d o 5 + 5 o d = Q. We have the chain complex 0 <— C°°(Z) ^-O\Z)
J-
SOP(Z) ^
,
called the canonical complex of a Poisson manifold. The homology H%an(Z) of this complex is said to be the canonical homology or the KoszulBrylinski-Poisson homology of a Poisson manifold. In the case of a symplectic manifold (Z, Q), the canonical homology is related to the de Rham cohomology of Z as follows. With a symplectic form, one can introduce the symplectic Hodge operator *:Or(Z)->O'2m-r{Z), which obeys the relations
*(**) =
*(j)=^-Qi((j))\rAn, m!
*(m*
2m = dimZ,
(-l)m°W(a)\(*
Then the Poisson codifferential 5 (2.1.27) on a symplectic manifold is given by the formula ty = ( - l ) W + 1 * d * &
This relation leads to the isomorphism H^n{Z) = H^-^Z),
(2.1.28)
Chapter 2 Classical Hamiltonian Systems
109
where H*(Z) is the de Rham cohomology of Z [72; 426]. F. Lichnerowicz-Poisson cohomology Given a Poisson manifold (Z,w), let us introduce the operator u?:T r (Z)-»T r + i(Z),
w{0) :=-[w,0],
^T,(Z),
(2.1.29)
on the graded commutative algebra %,(Z) of multivector fields on Z, where [.,.] is the Schouten-Nijenhuis bracket. This operator is nilpotent and obeys the rule w(ti A«) = {w(tf) A u + (-l) |l?l i? A w(v).
(2.1.30)
It is called the contravariant exterior differential [426], and makes %(Z) into a differential algebra. Its de Rham complex, called the LichnerowiczPoisson complex, reads -^•••) (2.1.31) where Sj^(Z) denotes the center of the Poisson algebra C°°(Z). Accordingly, cohomology HlP(Z,w) of this complex is called the LichnerowiczPoisson cohomology (henceforth the LP cohomology) of a Poisson manifold. 0-.SAZ)
-^C°°(Z)
- ^ T i ( Z ) A , . . . T r _ l ( Z ) ±+Tr(Z)
Example 2.1.16. If / G %{Z) = C°°(Z) is a function, -€>(f) =
[w,f]=4f
is its Hamiltonian vector field. Hence, the LP cohomology group H^P(Z, w) coincides with the center S? of the Poisson algebra C°°(Z). The first LP cohomology group HlP(Z, w) is the space of canonical vector fields u for the Poisson bivector field w (i.e., ~Luw = —w(u) = 0) modulo Hamiltonian vector fields -w{f), f S C°°{Z). The second LP cohomology group F£ P (Z, W) contains an element [w] whose representative is the Poisson bivector field w. We have [w] — 0 if there is a vector field u on Z such that w = w(u) = —Luw. If [w] = 0, a Poisson manifold (Z, w) is called exact or homogeneous. D Due to the property (2.1.30), the LP cohomology HlP{Z,w) is a graded commutative algebra with respect to the cup product {•d}
^{v}:={&Av}.
110
Geometric and Algebraic Topological Methods in Quantum Mechanics
In accordance with the relation (10.6.20), it is also provided with the bracket
[W,M]:={[Af]}. The contravariant exterior differential w is related to the exterior differential d and the Poisson codifferential 5W (2.1.27) by the formulae > 6 O*(Z),
w{w\
(2.1.32)
W
|1?| = H - 1 . (2.1.33)
wW\
The first of them shows that 10" is a cochain homomorphism of the de Rham complex (0*{Z),d) of exterior forms on Z to the Lichnerowicz-Poisson complex {%,— w) (2.1.31). It yields the homomorphism [w*]:H*(Z)^HZp(Z,w)
(2.1.34)
of the graded commutative algebra of the de Rham cohomology to that of the LP cohomology. The homomorphism (2.1.34) is an isomorphism if the Poisson bivector field w comes from a symplectic structure on Z [426]. In this case, the isomorphism (2.1.34) and the isomorphism (2.1.28) lead to the isomorphism TTCa.nir7\
-"»
Trim — iirv
K^) — -"LP
n,,\
\Z,W).
In general, there is the following relationship between the LP cohomology and the canonical homology of a Poisson manifold. By virtue of the relation (2.1.33), the interior product of exterior forms and multivector fields induces the corresponding contraction (,) : HlP(Z,w)
x H™(Z) -
tfocan(Z),
<{0},{0}> = W < « } ,
of vector spaces of the LP cohomology and the canonical homology of a Poisson manifold. Unless w is a symplectic structure, a serious drawback of the LP cohomology lies in the fact that it fails to satisfy the Poincare lemma, i.e., it is not trivial on a domain.
2.2
Geometry and cohomology of symplectic foliations
There is above-mentioned one-to-one correspondence between the symplectic foliations of a manifold Z and the Poisson structures on Z. Due to this
Chapter 2 Classical Harniltonian Systems
111
fact, one can provide the leafwise geometric quantization of Poisson manifolds as symplectic foliations (see Section 5.1). We start with some basic facts on geometry and cohomology of foliations. Let T be a (regular) foliation of afc-dimensionalmanifold Z provided with the adapted coordinate atlas (10.6.27). The real Lie algebra TX{T) of global sections of the tangent bundle TT —> Z to T is a C°°(Z)-submodule of the derivation module of the K-ring C°°(Z) of smooth real functions on Z. Its kernel Sj^(Z) C C°°{Z) consists of functions constant on leaves of T. Therefore, T\{T) is the Lie 5jr(Z)-algebra of derivations of C°°(Z), regarded as a 5^(Z)-ring. Then one can introduce the leafwise differential calculus [175; 215] as the Chevalley-Eilenberg differential calculus over the S^(Z)-ring C°°(Z). It is denned as a subcomplex 0->S?(Z)
—^C°°{Z) -1+31(2)... -^->5 d i m j r (Z)-^0
(2.2.1)
of the Chevalley-Eilenberg complex of the Lie 5jr(Z)-algebra T\{T) with coefficients in C°°(Z) which consists of C°° (.^-multilinear skew-symmetric maps xT1(F)-*Coo(Z),
r = l,...,dixnT.
These maps are global sections of exterior products ATT* of the dual TT* —> Z of TT —> Z. They are called the leafwise forms on a foliated manifold (Z,T), and are given by the coordinate expression 4> =-.K-iM1 r\
A---Adzir,
where {dz1} are the duals of the holonomic fibre bases {<%} for TT. Then one can think of the Chevalley-Eilenberg coboundary operator d
A dzil A • • • A dzir
as being the leafwise exterior differential. Accordingly, the complex (2.2.1) is called the leafwise de Rham complex (or the tangential de Rham complex in the terminology of [215]). This is the complex (A°'*,df) in [422]. Its cohomology H^(Z), called the leafwise de Rham cohomology, equals the cohomology H*(Z; S^) of Z with coefficients in the sheaf S?r of germs of elements of Sjr(Z) [143; 322]. We aim to relate the leafwise de Rham cohomology H^(Z) with the de Rham cohomology H*(Z) and the LP cohomology HlP(Z,w) [175].
112
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us consider the exact sequence (10.6.29) of vector bundles over Z. Since it admits a splitting, the epimorphism ij- yields that of the algebra O*(Z) of exterior forms on Z to the algebra 1$*{Z) of leafwise forms. It obeys the condition ij- o d = d o i^, and provides the cochain morphism dzi ^ dz\ (2.2.2) of the de Rham complex of Z to the leafwise de Rham complex (2.2.1) and the corresponding homomorphism i*T:{R,O*{Z),d)->{Sr{Z),T*{Z),d),
dzx^Q,
[*>]• : H*(Z) -> H£{Z)
(2.2.3)
of the de Rham cohomology of Z to the leafwise one. Let us note that [iJ-] r>0 need not be epimorphisms [422]. Given a leaf ip : F —> Z of !F, we have the pull-back homomorphism (R,O'(Z),d)->(R,O*(i?),d)
(2.2.4)
of the de Rham complex of Z to that of F and the corresponding homomorphism of the de Rham cohomology groups H*(Z) -> H*(F).
(2.2.5)
PROPOSITION 2.2.1. The homomorphisms (2.2.4) - (2.2.5) factorize • through the homomorphisms (2.2.2) - (2.2.3). Outline of proof. It is readily observed that the pull-back bundles %*FTT and i*FTT* over F are isomorphic to the tangent and the cotangent bundles of F, respectively. Moreover, a direct computation shows that iF{d4>) = d(i*Fct>)
for any leafwise form
dz* ~ dz\
(2.2.6)
of the leafwise de Rham complex of (Z, T) to the de Rham complex of F. Accordingly, the cohomology morphism (2.2.5) factorizes through the leafwise cohomology H*{Z) [^H}(Z) ! ^ F ( F ) .
(2.2.7) QED
113
Chapter 2 Classical Hamiltonian Systems
Let us turn now to symplectic foliations. Let T be an even dimensional foliation of a manifold Z. A d-closed non-degenerate leafwise two-form Qj? on a foliated manifold (Z, F) is called symplectic. Its pull-back i*FQf onto each leaf F of F is a symplectic form on F. If a symplectic leafwise form fi^- exists, it yields the bundle isomorphism
$r : TT -> TT\ z
Q^ : v h-» -v\Qjr(z),
(2.2.8)
v £ TZT.
The inverse isomorphism fij^ determines the bivector field
wn(a,p)=nr(niAi»>nUi*rP)),
a,peT*zZ, z e Z, (2.2.9)
2
on Z subordinate to ATT. It is a Poisson bivector field (see the relation (2.2.16) below). The corresponding Poisson bracket reads
{f,f}r = '&fW,
4f\Slr = -df,
(2.2.10)
^f = n^(df).
Its kernel is S>(Z). Conversely, let (Z, w) be a (regular) Poisson manifold and T its characteristic foliation. Since krnxTT C T*Z is precisely the kernel of a Poisson bivector field w, the bundle homomorphism w*:T*Z
z
^TZ
factorizes in a unique fashion u,« :T*Z
%
z
^TT* ^TT z
(2.2.11) '
^TZ z
v
through the bundle isomorphism v}T:TT* -+TT,
w\ : a H-> -W{Z) \a,
a€TzF*.
(2.2.12)
The inverse isomorphism w^- yields the symplectic leafwise form JV(v,«') = w(w^(v), w^v')),
v, v' e TZT,
z^Z.
(2.2.13)
The formulae (2.2.9) and (2.2.13) establish the above mentioned equivalence between the Poisson structures on a manifold Z and its symplectic foliations, though this equivalence need not be preserved under morphisms.
114
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us consider the Lichnerowicz-Poisson complex 2.1.31. We have the cochain morphism w* : (R,O*(Z),d) -> (S?(Z),%(Z),-w) t t f ^ X n , . . . ,ar) = (-l)>(t<;»(ai),... V K - ) ) , W o u;' = —w o
(2.2.14) <* G O\Z),
of the de Rham complex to the Lichnerowicz-Poisson one and the corresponding homomorphism (2.1.34) of the de Rham cohomology of Z to the LP cohomology of the complex (2.2.13) [426]. PROPOSITION 2.2.2. The cochain morphism to" (2.2.14) factorizes through the leafwise complex (2.2.1) and, accordingly, the cohomology homomorphism fiu"] (2.1.34) does through the leafwise cohomology
IT(Z)
[
%H*r(Z) -^HtP(Z,w).
(2.2.15)
• Outline of proof. Let %, (.F) C T* (Z) denote the graded commutative subalgebra of multivector fields on Z subordinate to TJ7, where 7o(^r) = C°°(Z). Clearly, (S^(Z),Tif(Jr),w) is a subcomplex of the LichnerowiczPoisson complex (2.1.31). Since w oft^.= -Cl^ o d,
(2.2.16)
the bundle isomorphism w^ = Cl^p (2.2.12) yields the cochain isomorphism
n% : (SAZ),r(Z),d) -» (Sr{Z),Tt(?),-w) of the leafwise de Rham complex (2.2.1) to the subcomplex (T*(!F), w) of the Lichnerowicz-Poisson complex (2.1.31). Then the composition ijroSl^: (Sr(Z),T(Z),d) -> (ST(Z),%(Z),-w)
(2.2.17)
is a cochain monomorphism of the leafwise de Rham complex to the LP one (2.1.31). In view of the factorization (2.2.11), the cochain morphism (2.2.14) factorizes through the cochain morphisms (2.2.2) and (2.2.17). Accordingly, the cohomology homomorphism [«;"] (2.1.34) factorizes through the cohomology homomorphisms [i*^] (2.2.3) and [ i f o 4 ] : H£(Z) -» HlP(Z,w).
(2.2.18) QED
Chapter 2 Classical Hamiltonian Systems
2.3
115
Hamiltonian systems
Subsections: A. Poisson and symplectic Hamiltonian systems, 115; B. Presymplectic Hamiltonian systems, 117; C. Hamiltonian systems with symmetries, 119; D. Partially integrable Hamiltonian systems, 123. This Section addresses some types of autonomous Hamiltonian systems on Poisson, symplectic and presymplectic manifolds whose quantization is studied in the sequel. A. Poisson and symplectic Hamiltonian systems Let us recall that dynamic equations, by definition, are differential equations which can be algebraically solved for the highest order derivatives. An autonomous differential equation on a smooth manifold Z is the particular case of a general notion of a differential equation in Remark 10.7.2. Let us consider the trivial fibre bundle R x Z —> R. Due to an isomorphism J^R x Z)=RxTZ, an autonomous first order differential equation on Z is equivalently defined as a closed submanifold (E of the tangent bundle TZ of Z. Its solution is a (local) vector field on Z which lives in £. An autonomous first order dynamic equation (or, simply, a dynamic equation) on Z is a section of the tangent bundle TZ —* Z, i.e., it is uniquely identified with a vector field on Z. Given a Poisson manifold (Z,w), a Poisson Hamiltonian system (w,7i) on Z for a Hamiltonian H € C°°(Z) with respect to the Poisson structure w is defined as the set
Sn= \J{veTzZ: v-wi(dH)(z)=0}.
(2.3.1)
zez By a solution of this Hamiltonian system is meant a vector field •& on Z, which takes its values into TN D SH- Clearly, the Poisson Hamiltonian system (2.3.1) has a unique solution which is the Hamiltonian vector field •&H = w\dU)
(2.3.2)
116
Geometric and Algebraic Topological Methods in Quantum Mechanics
of H. Hence, S-n (2.3.1) is a first order dynamic equation, called the Hamilton equations for the Hamiltonian H with respect to the Poisson structure w. Relative to local canonical coordinates {zx,ql,Pi) (2.1.16) for the Poisson structure w on Z and the corresponding holonomic coordinates (zx,q\pi,zx,q'l,pi) on TZ, the Hamilton equations (2.3.1) and their solution (2.3.2) take the form
Pi = -diH,
zx = 0,
tiH = dt-Hdi - diHd\ The following theorem points out the possibility of reducing a Poisson Hamiltonian system if a Poisson manifold admits a Poisson reduction [300; 426]. THEOREM 2.3.1. Let (Z,N,E) be a Poisson reduction and P a reduced manifold of Z via (N,E). Let H be a Hamiltonian on Z such that the flow of its Hamiltonian vector field tin preserves the submanifold N and the bundle E, and dH belongs to the annihilator of E. Then there is a Hamiltonian H on P such that H\N = ir*H and
• Let (Z, fi) be a symplectic manifold. The notion of a symplectic Hamiltonian system is a repetition of the Poisson one, but all expressions are rewritten in terms of a symplectic form fi as follows. A symplectic Hamiltonian system (fi, H) on a manifold Z for a Hamiltonian H, with respect to the symplectic structure Cl is the set Sn •.= \J{v e TZZ : -uJQ + dH{z) = 0}.
(2.3.3)
z€Z
As in the general case of Poisson Hamiltonian systems, the symplectic one (fi, H) has a unique solution which is the Hamiltonian vector field fin of
n, i.e.,
#n\n = -dH. With respect to the local canonical coordinates (ql,Pi) for the symplectic structure Cl, the Hamilton equations (2.3.3) read
Pi = -diH.
(2.3.4)
Chapter 2 Classical Hamiltonian Systems
117
Let (Z, w, TV) be a Poisson Hamiltonian system. Given a smooth function on Z, its Lie derivative along the Hamiltonian vector field flu (2.3.2) reads Unf
=ti
(2.3.5)
It is called the evolution equation. If {H, / } = 0, a function / is constant on integral curves of the Hamiltonian vector field fin- It is called the first integral. It is readily observed that the Poisson bracket {/, / ' } of any two first integrals / and / ' is also a first integral. Consequently, the first integrals of a Poisson Hamiltonian system constitute a Lie algebra. Since {H,f} = -0f\dH = -UfH, the Hamiltonian vector field $ / of a first integral / is an infinitesimal symmetry of a Hamiltonian Ti. B. Presymplectic Hamiltonian systems The notion of a Hamiltonian system is naturally extended to presymplectic manifolds [33; 186; 294; 407]. Given a presymplectic manifold (Z, Q), a presymplectic Hamiltonian system for a Hamiltonian Ti £ C°°(Z) is the set SH = \J{veTzZ:
v\£l + dH{z)=0}.
(2.3.6)
z€Z
A solution of this Hamiltonian system is a Hamiltonian vector field $-^ of Ti. The necessary and sufficient conditions of its existence are the following [186; 294]. PROPOSITION 2.3.2. The equation
v\Cl + dfi(z) = 0,
ve TZZ,
(2.3.7)
has a solution only at points of the set N2 = {z € Z : Ker z ft C Kei zdH}.
(2.3.8)
• Outline of proof. It is readily observed that the fibre (2.3.7) of the set Su (2.3.6) over z £ Z is an affine space modelled over the fibre Kerz12 = {«£ TZZ : v\Q = 0}
118
Geometric and Algebraic Topological Methods in Quantum Mechanics
of the kernel of the presymplectic form 0.. Let a vector v G TZZ satisfy the equation (2.3.7). Then the contraction of the right-hand side of this equation with an arbitrary element u £ Ker z fi leads to the equality u\dH{z) = 0. In order to prove the converse, it suffices to show that dTi(z) € ImO1". This inclusion results from the injections dH(z) G Ann(KerdW(z)) C Ann(Ker z 0) = Imfi1". QED
Let us suppose that a presymplectic form O is of constant rank and that N2 (2.3.8) is a submanifold of Z, but not necessarily connected. Then Ker fi is a closed vector subbundle of the tangent bundle TZ, while S-H\N2 is an affine bundle over N2. The latter has a section over N2, but this section need not live in TN2, i.e., it is not necessarily a vector field on the submanifold N2. Then one aims to find a submanifold N C N2 C Z such that Sn\NnTzN^(D,
zGN,
dH(z) G Qb(TN),
zGN.
or, equivalently,
If such a submanifold exists, it may be obtained by means of the following constraint algorithm [33]. Let us consider the overlap S-H\N2 r\TN2 and its projection to Z. We obtain the subset N3 =
nz{Sn\N2nTN2)cZ.
If N3 is a submanifold, let us consider the overlap Sw|jv3 f~l TN3. Its projection to Z gives a subset ./V4 C Z, and so on. Since a manifold Z is finite-dimensional, the procedure is stopped after a finite number of steps by one of the following results. • There is a number i > 2 such that a set Ni is empty. This means that a presymplectic Hamiltonian system has no solution. • A set Ni, % > 2, fails to be a submanifold. It follows that a solution need not exist at each point of A^. • If Ni+i = Ni for some i > 2, this is a desired submanifold N. A local solution of the presymplectic Hamiltonian system (2.3.6) exists around each point of N. If &\TN is °f constant rank, there is a global solution on N. Sections of the vector bundle Ker Q —> Z are sometimes called gauge fields in order to emphasize that, being solutions of the presymplectic
Chapter 2 Classical Hamiltonian Systems
119
Hamiltonian system (Q, 0) for the zero Hamiltonian, they do not contribute to a physical state, and are responsible for a certain gauge freedom [33; 407]. At the same time, there are physically interesting presymplectic Hamiltonian systems, e.g., in relativistic mechanics when a Hamiltonian is equal to zero (see Section 5.6). In this case, KerdW = TZ and the Hamilton equations (2.3.7) have a solution everywhere on a manifold Z. The above mentioned gauge freedom is also related to the pull-back construction in Proposition 2.1.5. Let a presymplectic form flona manifold Z be of constant rank and let its characteristic foliation be simple, i.e., a fibred manifold TT : Z —> P. Then Q is the pull-back 7r*Op of a certain symplectic form fip on P. Let a Hamiltonian fi be also the pull-back TT*HP of a function Tip on P. Then we have Kerft = VN C Ker dH, and the presymplectic Hamiltonian system (Q, 7i) has a solution everywhere on a manifold Z. Any such solution $ w is projected onto a unique solution of the symplectic Hamiltonian system (Clp,Hp) on the manifold P, while gauge fields are vertical vector fields on the fibred manifold Z —> P. One calls (Z, Q,,n*Hp) the gauge-invariant Hamiltonian system, which is equivalent to the reduced Hamiltonian system on a physical phase space P. C. Hamiltonian systems with symmetries Let us consider Hamiltonian systems on symplectic and Poisson manifolds, subject to a Lie group action. By G throughout is meant a real connected Lie group, g is its right Lie algebra, and g* is the Lie coalgebra (see Section 10.6E). We start with the symplectic case [l; 279; 299; 301]. Let a Lie group G act on a symplectic manifold (Z, Q) on the left by symplectomorphisms. Such an action of G is called symplectic. Since G is connected, its action on a manifold Z is symplectic if and only if the homomorphism e H-> £e, e e g, (10.6.34) of the Lie algebra g to the Lie algebra T\(Z) of vector fields on Z is carried out by canonical vector fields for the symplectic form Q on Z. If all these vector fields are Hamiltonian, the action of G on Z is called a Hamiltonian action. One can show that, in this case, fe, e € g, are Hamiltonian vector fields of functions on Z of the following particular type. PROPOSITION
2.3.3. An action of a Lie group G on a symplectic manifold
120
Geometric and Algebraic Topological Methods in Quantum Mechanics
Z is Hamiltonian if and only if there exists a mapping (2.3.9)
J-.Z-+Q*, called the momentum mapping, such that ££Jft = - d J £ ,
Je(z) = (J(z),e),
eSfl.
(2.3.10) D
The momentum mapping (2.3.9) is defined up to a constant map. Indeed, if J and J' are different momentum mappings for the same symplectic action of G on Z, then d((J(z)-J'(z),e))
= 0,
ee3.
A symplectic manifold provided with a Hamiltonian action of a Lie group is called a Hamiltonian manifold. Let H be a Hamiltonian of a Hamiltonian system on a Hamiltonian manifold (Z, Cl,G). If H is G-invariant, then L€.W = &JdW = {W,J e } = 0,
SGQ,
and all the functions Je (2.3.10) are first integrals of motion of the Hamiltonian system in question. The Poisson brackets of these functions are also first integrals of motion. Let us obtain their Poisson brackets. Given g e G, let us us consider the difference a(g) = J(gz)-Ad*g(J(z)),
(2.3.11)
where Ad*g is the coadjoint representation (10.6.38) on T*. One can show (see, e.g., [l]) that the difference (2.3.11) is constant on a symplectic manifold Z and that it fulfils the equality o-(gg') =c(g) +Ad* g(a(g')).
(2.3.12)
In accordance with the expression (1.5.10), the equality (2.3.12) is a onecocycle of cohomology H*(G;g*) of the group G with coefficients in the Lie coalgebra g*. This cocycle is a coboundary if there exists an element /i £ g* such that a(g) =[x-
Ad*g{ix)
(2.3.13)
(see the expression (1.5.9)). Let J ' be another momentum mapping associated to the same Hamiltonian action of G on Z. Since the difference J' — J' is constant on Z, then the difference of the corresponding cocycles a — a' is
121
Chapter 2 Classical Hamiltonian Systems
the coboundary (2.3.13) where \i = J — J'. Thus, a Hamiltonian action of a Lie group G on a symplectic manifold (Z, Q) defines a cohomology class [a] e H^G-^*) of G. A momentum mapping J is called equivariant if a(g) = 0, g € G. It defines the zero cohomology class of the group G. Example 2.3.1. Let a symplectic form on Z be exact, i.e., Q. = d6, and let 6 be G-invariant, i.e.,
L£.0 = d(&J0) + &jn = O,
eefl.
Then the momentum mapping J (2.3.9) can be given by the relation (J(z),e) = (t;e\e)(z). It is equivariant. In accordance with the relation (10.6.38), it suffices to show that Je(9z) = J A d r ' w W -
(&J0)(ff*) = (£Adff-i(e)J0)(z).
This holds by virtue of the relation (10.6.35). For instance, let T*Q be a symplectic manifold equipped with the canonical symplectic form Cl (2.1.3). Let a left action of a Lie group G on Q have the infinitesimal generators r m = e%m{q)di. The canonical lift of this action onto T*Q has the infinitesimal generators im = rm = ve^di - Pjdie^
(2.3.14)
(10.6.14), and preserves the canonical Liouville form 6 on T*Q. The £ m (2.3.14) are Hamiltonian vector fields of the functions Jm = eim(q)pi, determined by the equivariant momentum mapping J = e\n{q)pi£m. • Now a desired Poisson bracket of functions Je (2.3.10) is established as follows. 2.3.4. A momentum mapping J associated to a symplectic action of a Lie group G on a symplectic manifold Z obeys the relation
THEOREM
{Je,J£,} = J[e<el] - (Tea(e'),e)
(2.3.15)
(see, e.g., [l] where the left Lie algebra is utilized and Hamiltonian vector • fields differ in the minus sign from those here).
122
Geometric and Algebraic Topological Methods in Quantum Mechanics
In the case of an equivariant momentum mapping, the relation (2.3.15) leads to a homomorphism (2.3.16)
{Je,Je'} = J[e,E']
of the Lie algebra g to the Poisson algebra of functions on a symplectic manifold Z (cf. Proposition 2.3.5 below). Now let a Lie group G act on a Poisson manifold (Z, w) on the left by Poisson automorphism. This is a Poisson action. Since G is connected, its action on a manifold Z is a Poisson action if and only if the homomorphism £ M (E, £ e j , (10.6.34) of the Lie algebra g to the Lie algebra TX{Z) of vector fields on Z is carried out by canonical vector fields for the Poisson bivector field w, i.e., the condition (2.1.12) holds. The equivalent conditions are
&({/,}) = {Uf),g} + {/,&()}, k({/,}) = [6, /]()-&, W ) > &.*/] = *«.(/).
f,g € c°°(Z),
where •df is the Hamiltonian vector field (2.1.13) of a function / . A Hamiltonian action of G on a Poisson manifold Z is defined similarly to that on a symplectic manifold. Its infinitesimal generators are tangent to leaves of the symplectic foliation of Z, and there is a Hamiltonian action of G on every symplectic leaf. Proposition 2.3.3 together with the notions of a momentum mapping and an equivariant momentum mapping are also extended to a Poisson action. However, the difference a (2.3.11) is constant only on leaves of the symplectic foliation of Z in general. At the same time, one can say something more on an equivariant momentum mapping (that is also valid for a symplectic action) [426]. PROPOSITION 2.3.5. An equivariant momentum mapping J (2.3.9) is a Poisson morphism to the Lie coalgebra g*, provided with the Lie-Poisson structure (10.6.39). •
A Poisson manifold (Z, w), equipped with a Hamiltonian action of a Lie group G with an equivariant momentum mapping J, possesses the following standard Poisson reduction [426]. THEOREM 2.3.6. Let (Z, w, G, J) be as above. Let q be a point of g* such that
Chapter 2 Classical Hamiltonian Systems
123
• q is a non-critical value for all restrictions of J to symplectic leaves Ft of Z, i.e., level surfaces Zq = J~1{q) are submanifolds of Z and, for each symplectic leaf FL, the overlap Fiq = Zq n FL is a submanifold of F t ; • the overlaps Fiq are clean, i.e., TF^q = TFt n TZg, and so are the overlaps of Zq and the orbits of G in Z. Let Gq c G be the stabilizer of the point q. Then i^g are assembled into a regular foliation of Zq, whose leaves are orbits of the connected component of the unit of Gq, acting on Zq. Furthermore, if this foliation is a fibred manifold Zq —> P, its base P is a reduced Poisson manifold of (Z, w) via (Zq, E), where E is a disjoint union of tangent spaces to orbits of G. D Now one can use Theorem 2.3.1 in order to describe reduction of Poisson Hamiltonian systems with symmetries. Let (Z,w,Zq,E) be a Poisson reduction in Theorem 2.3.6. If "H is a G-invaxiant Hamiltonian on the Poisson manifold (Z,w), the Poisson Hamiltonian system reduces to a Poisson system on the reduced Poisson manifold of (Z, w) via (Zq, E). D. Partially integrable Hamiltonian systems We have seen that, given a Hamiltonian action of a Lie group G on a symplectic manifold Z characterized by the equivariant momentum mapping, there is a homomorphism (2.3.16) of its Lie algebra g to the Poisson algebra of smooth real functions on Z. Moreover, if a Hamiltonian of a symplectic Hamiltonian system on Z is G-invariant, we have a representation of g by first integrals. In reverse, one can start with a Lie algebra A of infinitesimal symmetries of a dynamical (e.g., Hamiltonian) system (see [396] for a survey). We restrict our consideration to a commutative Lie algebra A- This is the case of partially integrable Hamiltonian system (see Definition 2.3.7) below). Since no preferable Poisson structure is associated to a commutative Lie algebra A because its Lie-Poisson structure is zero, the analysis of a commutative integrable system essentially differs from that of the non-commutative ones in [396]. Let we have k mutually commutative vector fields {fix} on a smooth real manifold Z which are independent almost everywhere on Z, i.e., the set of points, where the multivector field A^x vanishes, is nowhere dense. We denote by <S C C°°(Z) the M-subring of smooth real functions / on Z whose derivations ^x\df vanish for all i?^. Let A be the fc-dimensional <S-Lie algebra generated by the vector fields {-&x}- One can think of one
124
Geometric and Algebraic Topological Methods in Quantum Mechanics
of its elements as being a first order dynamic equation on Z and of the other as being the dynamical symmetries. Accordingly, elements of S are regarded as integrals of motion. For the sake of brevity, we agree to call A a dynamical algebra.
Completely and partially integrable systems on symplectic manifolds [12] and broadly integrable dynamical systems of Bogoyavlenkij [42; 148] exemplify finite-dimensional commutative dynamical algebras. Let us recall that, given a symplectic manifold {Z,Q), we have a partially integrable system if there exist 1 < k < dim.Z/2 smooth real functions {H\} in involution which are independent almost everywhere on Z, i.e., the set of k points where thefc-formA dH\ vanishes is nowhere dense. The Hamiltonian vector fields -d\ of functions H\ are mutually commutative and independent almost everywhere. They make up a commutative dynamical algebra over the Poisson subalgebra S of elements of C°°{Z) commuting with all the functions H\. An important peculiarity of a finite-dimensional commutative dynamical algebra A is that its regular invariant manifolds are toroidal cylinders M.k~m x Tm. Since no preferable Poisson structure is associated to a commutative Lie algebra A, one can be free with analyzing different Poisson structures which make A into a Hamiltonian system. Different symplectic structures around invariant tori of commutative integrable systems have been investigated [42; 67; 148; 399]. For instance, the classical Liouville-Arnold theorem [12; 274] and the Nekhoroshev theorem [161; 339] state that, under certain conditions, every symplectic structure making a commutative dynamical algebra into a Hamiltonian system takes a canonical form around a compact invariant manifold. We aim to describe all Poisson structures bringing a commutative dynamical algebra into a partially integrable system near its regular invariant manifold, which need not be compact [179]. DEFINITION 2.3.7. Afc-dimensionalcommutative dynamical algebra on a regular Poisson manifold (Z, w) is said to be a partially integrable system (henceforth a PIS) if: (a) A is generated by Hamiltonian vector fields of k almost everywhere independent integrals of motion H\ G C°°(Z) in involution; (b) all elements of <S C C°°(Z) are mutually in involution. D
It follows at once from this definition that the Poisson structure w is at least of rank 2fc and <S is a commutative Poisson algebra. If 2fc — dim Z,
Chapter & Classical Hamiltonian Systems
125
we have a completely integrable system on a symplectic manifold. Given a fc-dimensional commutative dynamical algebra i o n a smooth manifold Z, let V be the smooth involutive distribution on Z spanned by the vector fields {d\}, and let G be the group of local diffeomorphisms of Z generated by the flows of these vector fields (we follow the terminology of [408]). Maximal integral manifolds of V are the orbits of G, and are invariant manifolds of A [408]. Let z € Z be a regular point of the distribution V, k
k
i.e., f\'d\{z) ^ 0. Since the group G preserves AI?A, the maximal integral manifold M of V through z is also regular (i.e., its points are regular). Furthermore, there exists an open neighborhood U of M such that, restricted to U, the distribution V is regular and yields a foliation # of U. 2.3.8. Let us suppose that: (i) the vector fields $,\ on U are complete, (ii) the foliation 3" of U admits a transversal manifold £ and its holonomy pseudogroup on £ is trivial, (iii) the leaves of this foliation are mutually diffeomorphic. Then the following hold. (I) There exists an open neighborhood of M, say U again, which is the trivial principal bundle
THEOREM
U = N x (Rk-m
x Tm) - ^ N
(2.3.17)
over a domain N C R d i m Z - f c with the structure group Rfc~m x Tm. (II) If 2k < dim Z, there exists a Poisson structure w of rank 2k on U such that (w,A) is a PIS in accordance with Definition 2.3.7. • Let us note the following. Condition (i) states that G is a group of diffeomorphisms of U. Condition (ii) is equivalent to the assumption that U -> U/G is a fibred manifold [316]. Each fibre Mr, r e N, of this fibred manifold admits a free transitive action of the group Gr = G/Kr, where Kr is the isotropy group of an arbitrary point of MT. In accordance with condition (iii), all the groups Gr, r € N, are isomorphic to the toroidal cylinder group Rk~m x Tm for some 0 < m < k. The goal is to define these isomorphisms so that they provide a smooth action of Rfe~m x Tm on U. We follow the proof in [116; 274] generalized to non-compact invariant manifolds. It should be emphasized that a trivial fibration in invariant manifolds is a standard property of integrable systems [12; 42; 66; 148; 155; 161; 176; 339]. However, there exists a well-known obstruction to its global extension in the case of compact invariant manifolds [26; 140], and there is
126
Geometric and Algebraic Topological Methods in Quantum Mechanics
an additional obstruction similar to that in [148] if invariant manifolds are non-compact. We establish a particular trivialization (2.3.17) such that the generators -&\ of the algebra A take the specific form (2.3.21). Part (II) of Theorem 2.3.8 is based on this trivialization Outline of proof. (I). By virtue of the condition (ii), the foliation J of U is a fibred manifold IT : U —> TV, admitting a section a such that and E = cr(N) [316]. Since the vector fields &\ on U are complete and mutually commutative, the group G of their flows is an additive Lie group of diffeomorphism of U. Its group space is a vector space M.k coordinated by parameters (sA) of the flows with respect to the basis {e\ = i!}\}. Since vector fields i?^ are independent everywhere on U, the action of Rk on U is locally free, i.e., isotropy groups of points of U are discrete subgroups of the group R*\ Its orbits are fibres of the fibred manifold U —• TV. Given a point r £ TV, the action of Rfc on the fibre Mr = n~1(r) factorizes as Rfc x Mr -» Gr x Mr -* Mr
(2.3.18)
through the free transitive action on Mr of the factor group GT — Rk/Kr, where Kr is the isotropy group of an arbitrary point of Mr, and the groups Gr are isomorphic to the additive group Rk~m x Tm. Let us bring the fibred manifold U —> N into a principal bundle with the structure group Go, where we denote {0} = ir(M). For this purpose, let us determine isomorphisms pr : GQ —> Gr of the group Go to the groups Gr, r £ TV. Then a desired fibrewise action of Go on U is defined by the law Go x Mr -» p r (G 0 ) x MT -» Mr.
(2.3.19)
Generators of each isotropy subgroup KT of M.k are given by m linearly independent vectors of the group space M.k. One can show that there exist ordered collections of generators (vi(r),... ,vm(r)) of the groups Kr such I > Vi{r) are smooth Revalued fields on TV. Let us consider the that r — decomposition vi(0)=B°(0)ea
+ CJi(0)ej,
a = l,...,k-m,
j =
l,...,m,
where G/(0) is a non-degenerate matrix. Since the fields Vi{r) are smooth, there exists an open neighborhood of {0}, say TV again, where the matrices Cl(r) are non-degenerate. Then
/Id
(B(r) - B(0))C-H0)\
(2320)
Chapter 2 Classical Hamiltonian Systems
127
is a unique linear morphism of the vector space M.k which transforms the frame «A(0) = {eo,«i(0)} into the frame v\(r) = {ea,Vi{r)}. Since it is also an automorphism of the group Rk sending Ko onto Kr, we obtain a desired isomorphism pr of the group Go to the group G r . Let an element g of the group Go be the coset of an element g(sx) of the group Rk. Then it acts on MT by the rule (2.3.19) just as the element g{{A~x)^s^) of the group Rfc does. This action of the group Go on U is smooth. It is free, and U/GQ = N. Then the fibred manifold U —> N is a trivial principal bundle with the structure group Go- Given its section a, the trivialization U = N x Go is defined by assigning the points p~1(gr) of the group space Go to the points gTcr{r), gr € G r , of a fibre Mr. Let us endow Go with the standard coordinate atlas (yx) = (t°,<^) of the group Rk~m x Tm. Then we provide U with the trivialization (2.3.17) with respect to the coordinates (rA,ta,ipl), where (rA), A = 1,...,dimZ — k, are coordinates on the base N. The vector fields i9\ on U relative to these coordinates read *a = da,
$i = -(BC-l)1{r)da
+ (C-^TOflfc.
(2.3.21)
Accordingly, the subring S restricted to U is the pull-back Tr*C°°(N) onto U of the ring of smooth functions on N. (II). Let us split the coordinates (rA) into some k coordinates (I\) and dimZ — 2k coordinates (zA). Then we can provide the toroidal domain U (2.3.17) with the Poisson bivector field w = dx A d\ of rank 2k. The independent complete vector fields da and di are Hamiltonian vector fields of the functions Ha = Ia and Hi = /» on U which are in involution with respect to the Poisson bracket
{/, /'} = dxfdxf - dxfdxf
(2.3.22)
defined by the above mentioned bivector field w. By virtue of the expression (2.3.21), the Hamiltonian vector fields {d\} generate the 5-algebra A. QED
A Poisson structure in Theorem 2.3.8 is by no means unique. Let the toroidal domain (2.3.17) be provided with bundle coordinates (rA,yx),
128
Geometric and Algebraic Topological Methods in Quantum Mechanics
where (rA) are coordinates on N and (yx) = (ta,tpl) are standard coordinates on the toroidal cylinder Rfc~m x T m . It is readily observed that, if a Poisson bivector field on the toroidal domain U satisfies Definition 2.3.7, it takes the form w = w1+w2=
wAX(rB)dA
/\dx+ w^ir8,
T/A)dM A dv.
(2.3.23)
The converse also holds. 2.3.9. For any Poisson bivector field w (2.3.23) of rank 2k on U, there exists a toroidal domain U' C U such that (w,A) is a PIS on U'.
THEOREM
•
It is readily observed that any Poisson bivector field w (2.3.23) fulfills condition (b) in Definition 2.3.7, but condition (a) imposes a restriction on the toroidal domain U. The key point is that the characteristic foliation T of U yielded by the Poisson bivector fields w (2.3.23) is the pull-back of a fc-dimensional foliation Tn of the base N, which is defined by the first summand w\ (2.3.23) of w. With respect to the adapted coordinates (J\,zA), A = 1 , . . . , k, on the foliated manifold (iV, FN), the Poisson bivector field w reads w = < ( J A , zA)du A9 M + vT{Jx,
zA, J/A)9M A dv.
(2.3.24)
Then condition (a) in Definition 2.3.7 is satisfied if N' C N is a domain of a coordinate chart {J\,zA) of the foliation F^. In this case, the dynamical algebra A on the toroidal domain U' = TT~1(N') is generated by the Hamiltonian vector fields dx = -w [dJx = w£0 M
(2.3.25)
of the k independent functions H\ = JxOutline of proof. The characteristic distribution of the Poisson bivector field w (2.3.23) is spanned by the Hamiltonian vector fields vA = -w{drA
= wA>ldil
(2.3.26)
and the vector fields w[dyx = wAXdA + 2w*iXdll. Since w is of rank 2k, the vector fields <9M can be expressed into the vector fields vA (2.3.26). Hence, the characteristic distribution of w is spanned by
129
Chapter 2 Classical Hamiltonian Systems
the Hamiltonian vector fields vA (2.3.26) and the vector fields x
v
= wAXdA.
(2.3.27)
The vector fields (2.3.27) are projected onto N. Moreover, one can derive from the relation [w, w] = 0 that they generate a Lie algebra and, consequently, span an involutive distribution VAT of rank k on N. Let TN denote the corresponding foliation of N. We consider the pull-back T = n*T^ of this foliation onto U by the trivial fibration n [316]. Its leaves are the inverse images TT~X(FN) of leaves F^ of the foliation TN, and so is its characteristic distribution TF=(Tv)-\VN). This distribution is spanned by the vector fields u* (2.3.27) on U and the vertical vector fields on U —> N, namely, the vector fields vA (2.3.26) generating the algebra A. Hence, TT is the characteristic distribution of the Poisson bivector field w. Furthermore, since U —» N is a trivial bundle, each leaf n~1(Fj^) of the pull-back foliation T is the manifold product of a leaf FN of N and the toroidal cylinder Rk~m xTm. It follows that the foliated manifold (U, T) can be provided with an adapted coordinate atlas {{Ut,Jx,zA,yx)},
X = l,...,k,
A=l,...,dimZ-2fe,
are adapted coordinates on the foliated manifold such that (J\,zA) (NJFH). Relative to these coordinates, the Poisson bivector field (2.3.23) takes the form (2.3.24). Let N' be the domain of this coordinate chart. Then the dynamical algebra A on the toroidal domain U' = n~1(N') is generated by the Hamiltonian vector fields d\ (2.3.25) of functions H\ = J\. QED Let us note that the coefficients u>M" in the expressions (2.3.23) and (2.3.24) are affine in coordinates yx because of the relation [u;,u;] = 0 and, consequently, are constant on tori. Furthermore, one can improve the expression (2.3.24) as follows. 2.3.10. Given a PIS (w,A) on a Poisson manifold (w,U), there exists a toroidal domain U' C U equipped with partial action-angle coordinates (Ia,Ii,zA,xa)(pl) such that, restricted to U', a Poisson bivector field takes the canonical form THEOREM
w = da Ada + d'Adu
(2.3.28)
130
Geometric and Algebraic Topological Methods in Quantum Mechanics
while the dynamical algebra A is generated by Hamiltonian vectorfieldsof the action coordinate functions Ha = Ia, Hi = U. • Theorem 2.3.10 extends the Liouville-Arnold theorem to the case of a Poisson structure and a non-compact invariant manifold. In order to prove it, we reformulate the proof of the Liouville-Arnold theorem for noncompact invariant manifolds in [155; 176] in terms of a leafwise symplectic structure. Outline of proof. First, let us employ Theorem 2.3.9 and restrict U to the toroidal domain, say U again, equipped with coordinates (J\,zA,yx) such that the Poisson bivector field w takes the form (2.3.24) and the algebra A is generated by the Hamiltonian vector fields fl\ (2.3.25) of k independent functions H\ = J\ in involution. Let us choose these vector fields as new generators of the group G and return to Theorem 2.3.8. In accordance with this theorem, there exists a toroidal domain U' C U provided with another trivialization U' —> N' C N in toroidal cylinders Rk~m x Tm and endowed with bundle coordinates (J\,zA,y'x) such that the vector fields •&X (2.3.25) take the form (2.3.21). For the sake of simplicity, let U', N' and y' be denoted U, N and y = {ta,Wl) again. Herewith, the Poisson bivector field w is given by the expression (2.3.24) with new coefficients. Let w" : T*U —> TU be the corresponding bundle homomorphism. It factorizes in a unique fashion wi :
T*U ^TF*
^*Tf
^TU
through the bundle isomorphism w^ : TT* -> TT,
w% : a H-> -W(X) [a,
(see the notation in Section 2.2). Then the inverse isomorphisms w^ : TT —> TT* provides the foliated manifold (U, T) with the leafwise symplectic form Q? = W/(Jx,zA,ta)dJliAdJv
fi«ur£ = s%,
+ Q^(Jx,zA)dJv/\dy>1, (2.3.29)
OQ/? = -Qffiwi".
(2.3.30)
Let us show that it is d-exact. Let F be a leaf of the foliation T of U. There is a homomorphism of the de Rham cohomology H*{U) of U to the de Rham cohomology of H*(F) of F, and it factorizes through the leafwise cohomology H^(U) due to (2.2.7). Since iV is a domain of an adapted
Chapter 2 Classical Harniltonian Systems
131
coordinate chart of the foliation .F/v, the foliation FN of VV is a trivial fibre bundle N = V x W -> W. Since T is the pull-back onto U of the foliation FN of N, it is also a trivial fibre bundle U = V x W x (Rk~m x Tm) -» W over a domain W C RdimZ-2k.
(2.3.31)
It follows that
H*(U) = H*{Tm) = H^{U). Then the closed leafwise two-form Q,j? (2.3.29) is exact due to the absence of the term Q^dy'1 A dyv. Moreover, Qj? = d'E where S reads S=
Ea(Jx,zA,yx)dJa+Ei(Jx,zA)diPi
up to a d-exact leafwise form. The Hamiltonian vector fields i?^ = i?^9M (2.3.21) obey the relation
n%4{ = 5J,
0x\nr = -dJx,
(2.3.32)
which falls into the following conditions ft* = dxEt - diZx, Q* = -daZ
x
X
=5 .
(2.3.33) (2.3.34)
The first of the relations (2.3.30) shows that Qjg is a non-degenerate matrix independent of coordinates y*. Then the condition (2.3.33) implies that diEx are independent of
(2.3.35)
di\9,T = -dEi.
(2.3.36)
Let us introduce new coordinates Ia = Ja, It = Ei(J\). By virtue of the equalities (2.3.34) and (2.3.35), the Jacobian of this coordinate transformation is regular. The relation (2.3.36) shows that 9, are Hamiltonian vector fields of the functions Hi = Ii. Consequently, we can choose vector fields d\ as generators of the algebra A. One obtains from the equality (2.3.34) that Ea = -ta +
Ea(Jx,zA)
132
Geometric and Algebraic Topological Methods in Quantum Mechanics
and S* are independent of ta. Then the leafwise Liouville form 2 reads S = (-ta + Ea(Ix, zA))dIa + E\IX, zA)dIi + Iidkp*. The coordinate shifts xa =
_ta + Ea{h> ^
tf
=
^ _
tf^
^
bring the leafwise form Qj? (2.3.29) into the canonical form fijF = dla A dxa + dh A dft which ensures the canonical form (2.3.28) of the Poisson bivector field w. QED
Now, let w and w' be two different Poisson structures (2.3.23) on the toroidal domain (2.3.17) which make a commutative dynamical algebra A into the different PISs (w,A) and (w',A). 2.3.11. We agree to call the triple (w,w',A) a bi-Hamiltonian PIS if any Hamiltonian vector field t? € A with respect to w possesses the same Hamilton representation
DEFINITION
ti = -w [df = -w' [df, relative to w', and vice versa.
f e 5,
(2.3.37) •
Definition 2.3.11 establishes a sui generis equivalence between the PISs (w,A) and (w',A). Theorem 2.3.12 below states that the triple (w,w',A) is a bi-Hamiltonian PIS in accordance with Definition 2.3.11 if and only if the Poisson bivector fields w and w' (2.3.23) differ only in the second terms w2 and w'2. Moreover, these Poisson bivector fields admit a recursion operator as follows. 2.3.12. (I) The triple (w,w',A) is a bi-Hamiltonian PIS in accordance with Definition 2.3.11 if and only if the Poisson bivector fields w and w' (2.3.23) differ in the second terms w-i and w'2. (II) These Poisson bivector fields admit a recursion operator. • THEOREM
Outline of proof. (I). It is easily justified that, if Poisson bivector fields w (2.3.23) fulfil Definition 2.3.11, they are distinguished only by the second summand w2. Conversely, as follows from the proof of Theorem 2.3.9, the characteristic distribution of a Poisson bivector field w (2.3.23) is spanned by the vector fields (2.3.26) and (2.3.27). Hence, all Poisson bivector fields
Chapter 2 Classical Hamiltonian Systems
133
w (2.3.23) distinguished only by the second summand 1112 have the same characteristic distribution, and they bring A into a PIS on the same toroidal domain U'. Then the condition in Definition 2.3.11 is easily justified. (II). The result follows from forthcoming Lemma 2.3.13. QED Given a smooth real manifold X, let w and w' be Poisson bivector fields of rank 2fc on X, and let w^ and -u/" be the corresponding bundle homomorphisms (2.1.10). A tangent-valued one-form i ? o n X yields bundle endomorphisms R:TX
-> TX,
R* : T*X -> T*X.
(2.3.38)
= wioR*.
(2.3.39)
It is called a recursion operator if w'* =RoW*
Given a Poisson bivector field w and a tangent valued one-form R such that Rovfi = w ' o i j * , the well-known sufficient condition for R o ufi to be a Poisson bivector field is that the Nijenhuis torsion of R and the MagriMorosi concomitant of R and w vanish [114; 343]. However, as we will see later, recursion operators between Poisson bivector fields in Theorem 2.3.12 need not satisfy these conditions. LEMMA 2.3.13. A recursion operator between Poisson structures of the same rank exists if and only if their characteristic distributions coincide. • Outline of proof. It follows from the equalities (2.3.39) that a recursion operator R sends the characteristic distribution of w to that of w', and these distributions coincide if w and w' are of the same rank. Conversely, let regular Poisson structures w and w' possess the same characteristic distribution TT —» TX tangent to a foliation T of X. We have the exact sequences (10.6.28) - (10.6.29). The bundle homomorphisms ty" and to'" (2.1.10) factorize in a unique fashion (2.2.11) through the bundle isomorphisms Wj: and w'jr (2.2.11). Let us consider the inverse isomorphisms ntf : TF-> TF*,
w%:TT ^>TT*
(2.3.40)
and the compositions Rr = w% o w^ : TT -> TT,
R*T = w\ o w% : TT* -+ TT*.
There is the obvious relation Wj: = Rjr
O Wp
= W^
O R*r.
(2.3.41)
134
Geometric and Algebraic Topological Methods in Quantum, Mechanics
In order to obtain a recursion operator (2.3.39), it suffices to extend the morphisms R? and R*r (2.3.41) onto TX and T*X, respectively. For this purpose, let us consider a splitting C-.TX -> TT, TX = Tf® (Id - i? o QTX = TF®E, of the exact sequence (10.6.28) and the dual splitting C* : TT* -> T*X, T*X = C(TF) © (Id - C o i*r)T*X = C{TF*) © £", of the exact sequence (10.6.29). Then the desired extensions are R* := (C* o R£) x Id E'. R:=Rry. Id E, This recursion operator is invertible, i.e., the morphisms (2.3.38) are bundle isomorphisms. QED For instance, the Poisson bivector field w (2.3.23) and the Poisson bivector field w0 =
wAX{r)dA/\dx
admit a recursion operator w\ = R o to' whose entries are given by the equalities R%=5%,
RZ = 6Z,
R$=0,
w^ = RxBwB».
(2.3.42)
Its Nijenhuis torsion fails to vanish, unless coefficients w^ are independent of coordinates yx. Let now A be a commutative dynamical algebra associated to a PIS on a symplectic manifold (Z, fi). In this case, condition (b) in Definition 2.3.7 is not necessarily satisfied, unless it is a completely integrable system. Nevertheless, there exists a Poisson structure w of rank 2k on the toroidal domain (2.3.17) such that, with respect to w, all integrals of motion H\ of the original PIS remain to be in involution, and they possess the same Hamiltonian vector fields &\ (see Theorem 2.3.14 below). Therefore, one can think of the triple (fl,io, {H\}) as being a special bi-Hamiltonian system, though it fails to satisfy Definition 2.3.11. Conversely, if Z is evendimensional, any Poisson bivector field w (2.3.23) setting a PIS (w,A) is extended to an appropriate symplectic structure 0 such that (ft, A) is a PIS on the symplectic manifold (Z,ft).
135
Chapter 2 Classical Hamiltonian Systems
2.3.14. Let (fi, {H\},"dx) be a fc-dimensional PIS on a 2ndimensional symplectic manifold (Z, ft). Let the distribution V, its regular integral manifold M, an open neighborhood U of M, and the foliation # of U be as those in Theorem 2.3.8. Under conditions (i) - (iii) of Theorem 2.3.8, the following hold. (I) There exists an open neighborhood of M, say U again, which is the trivial bundle (2.3.17) in toroidal cylinders Rfe~m x Tm over a domain N C R2n-k. (II) It is provided with the partial action-angle coordinates (I\,zA,y*) such that the functions H\ depend only on the action coordinates I\ and the symplectic form ft on U reads THEOREM
ft = dlx A dyx + SlAB(In, zc)dzA
A dzB + nA(I^
zc)dlx
A dzA.
(2.3.43)
(III) There exists a Darboux coordinate chart Q x Rk~m x Tm C U, foliated in toroidal cylinders Rfc~m x Tm and provided with coordinates (I\,Ps,Qs,yX) su ch that the symplectic form fi (2.3.43) on this chart takes the canonical form 0 = dl\ A dyx + dps A dqs.
(2.3.44)
• Outline of proof. This theorem is proved in [179]. Part (I) repeats exactly that of Theorem 2.3.8, while the proof of part (II) follows that of Theorem 2.3.10. The proof of part (III) is a generalization of that of Proposition 1 in [147] to non-compact invariant manifolds. As follows from the expression (2.3.44), the PIS in Theorem 2.3.14 can be extended to a completely integrable system on some open neighborhood of M, but Hamiltonian vector field of its additional local integrals of motion fail to be complete. QED A glance at the symplectic form ft (2.3.43) shows that there exists a Poisson structure w of rank 2k, e.g., w = dx A dx on U such that, with respect to w, the integrals of motion Hx of the original PIS remain to be in involution, and they possess the same Hamiltonian vector fields "&x- Hence, (£l,w, {Hx}) is the above mentioned bi-Hamiltonian system. Conversely, if Z is even-dimensional, any Poisson bivector field w (2.3.24) is extended to an appropriate symplectic structure Q as follows.
136
Geometric and Algebraic Topological Methods in Quantum Mechanics
2.3.15. The Poisson bivector field w (2.3.24) on a toroidal domain U' in Theorem 2.3.9 is extended to a symplectic structure ft on U' such that integrals of motion H\ = J\ remain in involution and their Hamiltonian vector fields with respect to w and ft coincide. • PROPOSITION
Outline of proof. The Poisson bivector field w (2.3.24) on the foliated manifold (U,T) defines a leafwise symplectic form $V (2.3.29). Restricted to the toroidal domain U' in Theorem 2.3.9 where coordinates J\ have trivial transition functions, the exact sequence (10.6.29) admits the splitting C* : TT* - T*U',
C'(dJM) = dJ»,
C W ) = dy\
such that C* o Q,jr is a presymplectic form on U'. Let nz =
QAB(zc)dzAAdzB
be also a presymplectic form on U'. It always exist. Then
is a desired symplectic form on U'.
2.4
QED
Hamiltonian time-dependent mechanics
Subsections: A. Fibre bundles over R, 137; B. Lagrangian time-dependent mechanics, 139; C. Hamiltonian time-dependent mechanics, 140; D. Timedependent completely integrable systems, 146; E. Degenerate systems, 149; F. Systems with time-dependent parameters, 151; G. The vertical extension ofLagrangian and Hamiltonian mechanics, 154. The Hamiltonian formulation of conservative mechanics cannot be extended to time-dependent mechanics in a straightforward manner because the symplectic form (2.1.3) is not invariant under time-dependent transformations. The usual palliative formulation of time-dependent mechanics implies a preliminary splitting of its configuration space Q = Rx M where M is a manifold, while R is a time axis. Prom the physical viewpoint, it means that a certain reference frame holds fixed [294]. Then we have the corresponding splitting of the momentum phase space R x T*M, provided with the presymplectic form pr^ft = dpi A dql
Chapter 2 Classical Hamiltonian Systems
137
which is the pull-back of the canonical symplectic form fl (2.1.3) on the cotangent bundle T*M [88]. The problem is that the above mentioned splittings are also broken by any time-dependent transformation, and so is the presymplectic form pr^fi. At the same time, non-relativistic timedependent mechanics can be adequately formulated as a particular field theory whose configuration space is a fibre bundle over a time axis R [294; 384; 385]. A. Fibre bundles over R Let us point out some peculiarities of fibre bundles over R. Let n : Q —> R be a fibre bundle whose typical fibre M is an m-dimensional manifold. It is endowed with bundle coordinates (t, ql) where t is a Cartesian coordinate on R with the transition functions t' = t+ const, (we are not concerned with time-reparameterized mechanics [212; 379]). The base R is provided with the standard vector field dt and the standard one-form dt which are invariant under the above mentioned coordinate transformations. The same symbol dt also stands for any pull-back of the standard one-form dt onto fibre bundles over R. For the sake of convenience, we sometimes use the compact notation (qx) where q° = t. Since R is contractible, any fibre bundle over R is trivial. Different trivializations
differ from each other in projections Q —> M. The first order jet manifold JXQ of a fibre bundle Q —» R is provided with the adapted coordinates (t, q%, q\). Every trivialization of Q —» R yields the corresponding trivialization of the jet manifold j'QSMxTM.
(2.4.1)
The canonical imbedding (10.6.42) of JXQ takes the form Xi : JlQ -» TQ, i
i
X1:(t,q ,qi)»{t,q ,i
X1=dt=dt
+ q\di,
(2.4.2)
= M * = ?t)-
We will often identify the jet manifold JXQ with its image in TQ. In particular, any connection r = dt ® {dt + r'(t, qj)di)
(2.4.3)
138
Geometric and Algebraic Topological Methods in Quantum Mechanics
on a fibre bundle Q —* E can be associated to a nowhere vanishing horizontal vector field r = dt+Ti(t,qi)di
(2.4.4)
on Q which is the horizontal lift Tdt (10.6.48) of the standard vector field dt on E by means of the connection (2.4.3). Conversely, any vector field T on Q such that dt\T = 1 defines a connection on Q —» E. As a consequence, connections on a fibre bundle Q —> E constitute an affine space modelled over the vector space of vertical vector fields on Q —> E. Accordingly, the covariant differential Dr (10.6.53) associated to a connection T on Q —> E takes its values into the vertical tangent bundle VQ of Q —> E: DT:J1Q-^VQ,
qioDT = q\-Vi.
(2.4.5)
It is a first order differential operator on Q —> E. A connection T (2.4.3) is obviously flat. By virtue of Proposition 10.6.10, it determines an atlas of local constant trivializations of Q —» E such that the associated bundle coordinates (t, q1) on Q possess the transition function independent of t, and T = dt with respect to these coordinates. Conversely, every atlas of local constant trivializations of the fibre bundle Q —> E defines a connection on Q —> E which is equal to dt relative to this atlas. A connection F on a fibre bundle Q —* E is said to be complete if the horizontal vector field (2.4.4) is complete. 2.4.1. [294]. Every trivialization of a fibre bundle Q —> R yields a complete connection on this fibre bundle. Conversely, every complete connection T on Q —> E defines its trivialization such that the vector field (2.4.4) equals dt relative to the bundle coordinates associated D with this trivialization. PROPOSITION
Given a connection T (2.4.3), the kernel of its covariant differential DT (2.4.5) is a first order dynamic equation on a fibre bundle Q —> E. It is given by the equivalent coordinate relations
(2-4-6)
* = 1,
(2.4.7)
in vie of the canonical imbedding (2.4.2). Conversely, any first order dynamic equation on Q —> E is of this type. Thus, one identifies first order dynamic equations and connections on a fibre bundle over E. A dy-
Chapter 2 Classical Hamiltonian Systems
139
namic equation (2.4.6) is autonomous if there exists a trivialization (2.4.1) of Q —> R such that the vector field (2.4.4) is projectable onto M. B. Lagrangian time-dependent mechanics Lagrangian formalism of time-dependent mechanics is particular first order Lagrangian formalism in Section 10.8 on fibre bundles over R. Given a mechanical system on a configuration space Q —> R., its velocity phase space is the jet manifold JXQ of Q —> R. A first order Lagrangian on JlQ is a density L = C{t,qj,q{)dt.
(2.4.8)
In time-dependent mechanics without time reparameterization, one considers projectable vector fields u = uldt + rjdi,
u\dt = u* = const.,
(2.4.9)
on a configuration bundle Q —> R which are generators of local oneparameter groups of automorphism of Q —> M. over translations of the base R. The jet prolongation (10.6.45) of u (2.4.9) is J1u = utdt + uidi + dtuidti.
(2.4.10)
Given a Lagrangian L (2.4.8), its Lie derivative along Jlru (2.4.10) reads LjiuL = {uldt + tfdi + dttfdDCdt.
(2.4.11)
Then the first variational formula (10.8.2) takes the form Jxu\dC = K - u'qDSi + dt(u\HL),
(2.4.12)
where HL = -Kidq* - (mql - C)dt,
n = d\C,
(2.4.13)
is the Poincare-Cartan form (10.8.5) and SL : J2Q -> V*Q,
£L = SiCdq* = {dtC - dtnjdq*,
(2.4.14)
is the Euler-Lagrange operator (10.8.3) of a Lagrangian L. The kernel of the Euler-Lagrange operator (2.4.14) defines the Lagrange equations 6iC = (dt - dtd\)C = 0,
dt = dt + y\di + y\td\.
(2.4.15)
140
Geometric and Algebraic Topological Methods in Quantum
Mechanics
On the shell (2.4.15), the first variational formula (2.4.12) is brought into the weak identity (u*ft + iJdi + dtii'd^C « dtCw-iCu* - u'ql) + ulC),
(2.4.16)
where, by analogy with field theory, the function T = -u\HL
= TT^U* - u*q\) + ulC
(2.4.17)
is said to be the symmetry current along the vector field u. For the sake of convenience, it differs from the expression (10.8.19) in sign minus. If the Lie derivative Lji u L (2.4.11) vanishes, we obtain the weak conservation law 0 ss dt% = dtfciu* - u*q\) + u'C),
(2.4.18)
where the conserved current (2.4.17) is an integral of motion. In particular, if u = uxdi is a vertical vector field, the conservation law (2.4.18) takes the form 0 « dt(ffiu')
where, by analogy with field theory, the integral of motion Xu = i^iU1 is called the Noether current. Besides the Lagrange equations (2.4.15), we also refer to the Cartan equations which are introduced similarly to those in Section 10.8. Being the Lepagean equivalent of the Lagrangian L on J1*?, the Poincare-Cartan form (2.4.13) is also the Lepagean equivalent of the Lagrangian L = (C + (ql- qi)iti)dt,
(2.4.19)
The on the repeated jet manifold J^JlQ, coordinated by (t,ql,ql,qi,qit). Lagrange equations for L are the above mentioned Cartan equations
dl*}(%-qi) = 0, diC -dtn dt = dt+yidi + yitdl
+ (4 - 4)di-nj = 0 , (2.4.20)
They are equivalent to the Lagrange equations (2.4.15) on integrable sections c = J*c of JXQ —> R and in the case of regular Lagrangians. C. Hamiltonian time-dependent mechanics Similarly to Lagrangian time-dependent mechanics, the Hamiltonian time-dependent mechanics is particular Hamiltonian theory in Section 10.9
141
Chapter 2 Classical Hamiltonian Systems
on a fibre bundle Q —> M. The corresponding momentum phase space II (10.8.16) is the vertical cotangent bundle Trn : V*Q - Q,
(2.4.21)
equipped with the holonomic coordinates (t, ql,Pi =
^ =| U
Tt-P+%*-
(2-4.22)
It is the particular homogeneous Legendre bundle Zy (10.8.4) over X = R which admits the canonical Liouville form (2.4.23)
E=pdt + pidq\ the canonical symplectic form dS = dp A dt + dpi A dq\
(2.4.24)
and the corresponding Poisson bracket { / , 9 } T = d?fdtg
- dpgdtf
+ d'fdtg
- d'gdj,
f,ge
C°°{T*Q).
(2.4.25) Provided with the structures (2.4.23) - (2.4.25), the cotangent bundle T*Q of Q plays a role of the homogeneous momentum phase space of timedependent mechanics. There is the canonical one-dimensional fibre bundle C : T*Q -> V*Q
(2.4.26)
(cf. (10.8.15)). A glance at the transformation law (2.4.22) shows that it is a trivial affine bundle. Indeed, given a global section h of £, one can equip T*Q with the fibre coordinate r — p — h possessing the identity transition functions. Let us consider the subring of C°°(T*Q) which comprises the pull-back £*/ onto T*Q of functions / on the vertical cotangent bundle V*Q by the fibration £ (2.4.26). This subring is closed under the Poisson bracket (2.4.25). Then, by virtue of Theorem 2.1.7, there exists the (degenerate) Poisson structure {/, 9}v = d'fdtg
- Pgdif,
f , g e C°°(V*Q),
(2.4.27)
142
Geometric and Algebraic Topological Methods in Quantum Mechanics
on the momentum phase space V*Q such that C{f,9}v = {Cf,C9}T-
(2.4.28)
The holonomic coordinates on V*Q are canonical for the Poisson structure (2.4.27). With respect to the Poisson bracket (2.4.27), the Hamiltonian vector fields of functions on V*Q read df = difdi-difdi,
f£C°°(V*Q).
(2.4.29)
They are vertical vector fields on V*Q —> R. Accordingly, the characteristic distribution of the Poisson structure (2.4.27) is the vertical tangent bundle VV*Q C TV*Q of the fibre bundle V*Q -> R. The corresponding symplectic foliation on the momentum phase space V*Q coincides with the fibration V*Q -> R. Remark 2.4.1. A generic momentum phase space of time-dependent mechanics is a fibre bundle II —> R endowed with a regular Poisson structure whose characteristic distribution belongs to the vertical tangent bundle VII of II —> R [210]. It can be seen locally as the Poisson product over R of the Legendre bundle V*Q -» R and a fibre bundle over R, equipped with the • zero Poisson structure. Remark 2.4.2. The Poisson structure (2.4.27) can be introduced in a different way [294; 384]. Given a section h of the fibre bundle (2.4.26), let us consider the pull-back forms 0 = h*(E A dt) = Vidtf A dt, n = h*(d~.Adt)= dpi A dqi A dt
(2.4.30)
on V*Q. They are independent of a section h, and are the particular tangent-valued Liouville form (10.9.1) and the polysymplectic form (10.9.2), respectively. With ft, the Hamiltonian vector field 1?/ (2.4.29) for a function / on V*Q is given by the relation df\S1 = -dfAdt, while the Poisson bracket (2.4.27) is written as {f,9}vdt = 0g\0f\n.
Chapter 2 Classical Hamiltonian Systems
143
Moreover, one can show that a projectable vector field i9 onV*Q such that •d\dt =const. is a canonical vector field for the Poisson structure (2.4.27) if and only if
• In contrast with autonomous mechanics, the Poisson structure (2.4.27) fails to provide any dynamic equation on the momentum phase space V*Q because Hamiltonian vector fields (2.4.29) of functions on V*Q are vertical vector fields, but not connections. Hamiltonian dynamics of timedependent mechanics is described as a particular Hamiltonian dynamics on fibre bundles in Section 10.9. A Hamiltonian on the momentum phase space V*Q —> R of timedependent mechanics is defined as a global section poh = -H(t,qj,pj),
h:V*Q->T*Q,
(2.4.31)
of the affine bundle £ (2.4.26). It yields the pull-back Hamiltonian form H = h*Z = Pkdqk - Hdt
(2.4.32)
on V*Q, which is the well-known invariant of Poincare-Cartan [12]. Given a Hamiltonian form H (2.4.32), there exists a unique horizontal vector field JH on V*Q (i.e. 7jyjdi = 1) such that 1H\dH = 0.
(2.4.33)
This vector field, called the Hamilton vector field, reads 1H = dt + dk7idk - dkUdk.
(2.4.34)
In a different way (see Remark 2.4.2), the Hamilton vector field JH is defined by the relation 7 /fJfi
= dF
(cf. (10.9.6)). Consequently, it is canonical for the Poisson structure {, }y (2.4.27). This vector field yields the first order Hamilton equations qk = dk7i,
(2.4.35a)
Ptk = ~dkH
(2.4.35b)
144
Geometric and Algebraic Topological Methods in Quantum Mechanics
on V*Q —> R, where (t,qh,pk,qt,Ptk) are the adapted coordinates on the first order jet manifold JXV*Q of V*Q -> R. We agree to call {V*Q, H) a time-dependent Hamiltonian system of m — dim M degrees of freedom. Remark 2.4.3. Any connection F (2.4.4) on the configuration bundle Q -> R defines the global section hr = -piT* (2.4.31) of the affine bundle C (2.4.26) and the corresponding Hamiltonian form HT = Pkdqk -
PiFdt.
(2.4.36)
Furthermore, given a connection T, any Hamiltonian form (2.4.32) admits the splitting H = Hr-
Hrdt,
where HV = H- ftP is a function on V*Q (cf. (10.9.8)).
(2.4.37) D
Remark 2.4.4. If the Hamilton vector field JH (2.4.34) is complete, the Hamilton equations (2.4.35a) - (2.4.35b) admit a unique global solution through each point of the momentum phase space V*Q. By virtue of Proposition 2.4.1, there exist a trivialization of the fibre bundle V*Q —> E (not necessarily compatible with its fibration V*Q —> Q) such that 7# = dt with respect to the associated coordinates (t, Q%, Pi). Moreover, since yn is the canonical vector field for the Poisson structure {, }v, one can show that (t, Ql,Pi) are canonical coordinates for {,}v [294]. The Hamilton equations (2.4.35a) - (2.4.35b) in this coordinates take the form Q^ = 0, Ptfe = 0, i.e., (t,Ql,Pi) are the initial data coordinates. • A first integral of the Hamilton equations (2.4.35a) - (2.4.35b) is defined as a smooth real function FonV*Q whose Lie derivative LlHF = lH\dF = dtF + {H,F}v
(2.4.38)
along the Hamilton vector field JH (2.4.34) vanishes, i.e., the function F is constant on trajectories of JH- One can think of the formula (2.4.38) as being the evolution equation of time-dependent mechanics. In contrast with the evolution equation (2.3.5) of autonomous mechanics, the righthand side of (2.4.38) fails to reduce to the Poisson bracket {, }y which, as it was mentioned above, does not determine evolution of time-dependent mechanics.
145
Chapter 2 Classical Hamiltonian Systems
In order to overcome this difficulty, we use the fact that a timedependent Hamiltonian system of m degrees of freedom can be extended to an autonomous Hamiltonian system of m + 1 degrees of freedom where the time is regarded as a dynamic variable [54; 124; 282; 295]. Let us consider the pull-back £*H of the Hamiltonian form H = h*E onto the cotangent bundle T*Q. It is readily observed that the difference E — £*/i*H is a horizontal one-form on T*Q —• R and that H* =dt\(E-Ch*E))=p
+H
(2.4.39)
is a function on T*Q. Let us regard H* (2.4.39) as a Hamiltonian of an autonomous Hamiltonian system on the symplectic manifold (T*Q,QT)The Hamiltonian vector field of H* on T*Q reads r r
= dt- dtnd° + dkHdk - dkUdk.
(2.4.40)
It is projected onto the Hamilton vector field JH (2.4.34) on V*Q, and we have the relation
C (L 7 */) = {n*, Cfh,
f G c°°(y*Q).
(2.4.41)
As in the case of field theory, a Hamiltonian form H (2.4.32) is the Poincare-Cartan form of the Lagrangian LH = ho(H) = (Piqi - H)dt
(2.4.42)
(cf. (10.9.9)) on the jet manifold JXV*Q such that the Hamilton equations (2.4.35a) - (2.4.35b) are exactly the Lagrange equations for Ljf. Furthermore, let u be a vector field (2.4.9) on the configuration space Q. Due to the canonical lift (10.6.14), it gives rise to the vector field u = utdt + uidi-diujpjdi
(2.4.43)
on the momentum phase space V*Q —* R. Then we have the equality LuH = LjiuLH = (-u'dtH + pidtu* - tfdiH + djuipidiH)dt.
(2.4.44)
The first variational formula (2.4.12) applied to the Lagrangian LH (2.4.42) leads to the weak identity LzH * dt(u\H)dt. If the Lie derivative (2.4.44) vanishes, we come to the conservation law 0 Kdt(u\H)dt
(2.4.45)
146
Geometric and Algebraic Topological Methods in Quantum Mechanics
of the function 5 U = u\dH = piU1 - utrH.
(2.4.46)
By analogy with field theory, we agree to call this function the symmetry current along u. It is easily verified that ~LlH%u = 0 if and only if L^iJ = 0. Consequently, any conserved symmetry current Tu (2.4.46) is a first integral. PROPOSITION 2.4.2. The conserved currents (2.4.46) form a Lie algebra with respect to the Poisson bracket
{£u,£u'}v =£[«,„']•
(2-4.47)
• In particular, if u is a vertical vector field, the conserved current T (2.4.46) is the Noether current %u = u\q = Viu\
q = p^dqi £ V*Q.
(2.4.48)
All (not necessarily conserved) Noether currents (2.4.48) constitute a Lie algebra with respect to the bracket (2.4.47).
D. Time-dependent completely integrable systems A time-dependent Hamiltonian system (V*Q,H) of m degrees of freedom is said to be completely integrable (henceforth a CIS) if the Hamilton equations (2.4.35a) - (2.4.35b) admit m first integrals Fk which are in involution with respect to the Poisson bracket {, }y (2.4.27), and whose differentials dFk are linearly independent almost everywhere (i.e., the set of points where this condition fails is nowhere dense). We aim to show that, similarly to an autonomous CIS, a time-dependent CIS admits the action-angle coordinates around any instantly compact regular invariant manifold [155; 176]. Written relative to these coordinates, its Hamiltonian and first integrals are functions only of action coordinates. We use the fact that a time-dependent CIS (V*Q, H, Fk) of m degrees of freedom on V*Q can be extended to an autonomous CIS (T*Q, H*) of m + 1 degrees of freedom on T*Q where H* is the autonomous Hamiltonian (2.4.39) [176]. Indeed, an immediate consequence of the relation (2.4.41) is the following. THEOREM 2.4.3. (i) Given a time-dependent CIS (H,Fk) on V*Q, the Hamiltonian system (H*, C*Fk) on T*Q is a CIS. (ii) Let M' be a connected
147
Chapter 2 Classical Hamiltonian Systems
regular invariant manifold of (7i, F^). Then h(M') C T*Q is a connected regular invariant manifold of the autonomous CIS (H.*,£*Fk). • Hereafter, we assume that the vector field 7# (2.4.34) is complete. In this case, trajectories of JH define a trivial fibre bundle V*Q —> V*Q over any fibre Vt*Q of V*Q —> K (see Remark 2.4.4)). Without loss of generality, we choose the fibre io : VSQ -> V*Q at t = 0. Since M' is an invariant manifold, the fibration £ : VQ -> V£Q
(2.4.49)
also yields the fibration of M' onto Mo = M'H V0*Q such that M' = R x M 0 is a trivial bundle. Let us introduce the action-angle coordinates around an invariant manifold M' of a time-dependent CIS on V*Q by use of the action-angle coordinates around the invariant manifold h(M') of the autonomous CIS on T*Q in Theorem 2.4.3. Let us note that M' and, consequently, h(M') are never compact because of the time axis. We appeal to Theorem 2.3.14 adapted to a CIS as follows. THEOREM 2.4.4. Let M be a connected invariant manifold of an autonomous CIS {F\}, A = 1 , . . . , n, on a 2n-dimensional symplectic manifold (Z, Q.z)- Let U be an open neighborhood of M such that: (i) the differentials dF\ are independent everywhere on U, (ii) the Hamiltonian vector fields d\ of the first integrals F\ on U is complete, (iii) the submersion xF\ : U —> M.n is a trivial bundle of invariant manifolds over a domain N C W1. Then U is isomorphic to the symplectic annulus
U' = N'x
(Rn-m
x T m ),
(2.4.50)
provided with the action-angle coordinates
( A , . . . , /„; x\ ..., xn~m; (?,...,4>m)
(2-4.51)
such that the symplectic form on U' reads nz = dla A da;a + dJi A dft, and the first integrals F\ depend only on the action coordinates 1^.
D
148
Geometric and Algebraic Topological Methods in Quantum Mechanics
Of course, the action-angle coordinates (2.4.51) by no means are unique. For instance, let Ta, a — 1,..., n — m, be an arbitrary smooth function on Rm. Let us consider the canonical coordinate transformation I'a = Ia- Fa(Ij),
A = h,
x'a = xa,
^ = >*+ xa&Fa{!j).
(2.4.52)
Then (I'a, I'k; x'a, cp'k) are action-angle coordinates on the symplectic annulus which differs from U' (2.4.50) in another trivialization. Now, we apply Theorem 2.4.4 to the CISs in Theorem 2.4.3. THEOREM 2.4.5. Let M' be a connected regular invariant manifold of a time-dependent CIS on V*Q, and let the range Mo of its projection £ (2.4.49) be compact. Then the invariant manifold h(M') of the autonomous CIS on T*Q has an open neighborhood U obeying the condition of Theorem 2.4.4. • Outline of proof. We refer the reader to [176] for the proof.
QED
In accordance with Theorem 2.4.4, the open neighborhood U of the invariant manifold h(M') of the autonomous CIS in Theorem 2.4.5 is isomorphic to the symplectic annulus U' = N' x (R x Tm),
N' = (-e,e) x TV,
(2.4.53)
provided with the action-angle coordinates (Jo,...,/m;^\...,
(2-4.54)
such that the symplectic form on U' reads fi = dlo A dt + dlk A d<j)k,
while To = 7i* and the first integrals C*Fk depend only on the action coordinates 7j. We agree to call (2.4.54) the time-dependent action-angle coordinates. Let us note that the symplectic annulus U' (2.4.53) inherits the fibration U' -^-> U" = TV x (R x Tm).
(2.4.55)
By the relation similar to (2.4.28), the product U" (2.4.55) coordinated by (U; t, (f>1) is provided with the Poisson structure
{/, f'}u = fffdif - djd'f,
f, f e C°°(U").
Chapter 2 Classical Hamiltonian Systems
149
Therefore, one can regard U" as the momentum phase space of the timedependent CIS in question around the invariant manifold M'. It is readily observed that the Hamiltonian vector field 7^ of the autonomous Hamiltonian 7i* = IQ is 7^ = dt, and so is its projection 7# (2.4.34) onto U". Consequently, the Hamilton equations (2.4.35a) (2.4.35b) of a time-dependent CIS with respect to the action-angle coordinates (2.4.54) take the form 7j = 0, 4>% = 0. Hence, (/,;£,>*) are the initial data coordinates. One can introduce such coordinates as follows. Given the fibration £ (2.4.49), let us provide Mo x N C VJQ in Theorem 2.4.5 with action-angle coordinates (Ii\<j>) for the CIS {i^-Ffc} on the symplectic leaf VQQ. Then it is readily observed that (Ii;t,> ) are timedependent action-angle coordinates on U" (2.4.55) such that the Hamiltonian H(Ij) of a time-dependent CIS relative to these coordinates vanishes, i.e., 7i* = IQ. Using the canonical transformations (2.4.52), one can consider time-dependent action-angle coordinates besides the initial date ones. Given a smooth function H on R m , one can provide U" with the actionangle coordinates Jo = 7o - W(Ji),
h = J*,
>* = 7 + td'Hilj)
such that H{Ii) is a Hamiltonian of a time-dependent CIS on U". E. Degenerate systems Referring to general Hamiltonian formalism on fibre bundles in Section 10.9, we consider associated Lagrangian and Hamiltonian systems in timedependent mechanics [294; 295]. Every Lagrangian L (2.4.8) on the velocity phase space JXQ induces the Legendre map L-.J^Q —>V*Q,
Pi°L
= Tri,
(2.4.56)
whose range NL = L{JlQ) c V*Q is called a Lagrangian constraint space. Every Hamiltonian form H (2.4.32) yields the Hamiltonian map H:V*Q
^JlQ,
q\°H = diH.
(2.4.57)
150
Geometric and Algebraic Topological Methods in Quantum Mechanics
Given a Lagrangian L, a Hamiltonian form H is said to be associated to L if H satisfies the relations L o H o L = L,
(2.4.58a)
H*LH = H*L.
(2.4.58b)
A glance at the relation (2.4.58a) shows that L o H is the projector Pi(z)=Tri(t,qi,d
j
H(z)),
zGNL,
from V*Q onto the Lagrangian constraint space JV/,. Accordingly, H o L is the projector from JlY onto H(NL). A Hamiltonian form is called weakly associated to a Lagrangian L if the condition (2.4.58b) holds on the Lagrangian constraint space NLLet us restrict our consideration to almost regular Lagrangians L, i.e., (i) the Lagrangian constraint space NL is a closed imbedded subbundle iN :NL^V*Q of the bundle V*Q -> Q, (ii) the Legendre map L : JlQ —> NL is a fibred manifold, (iii) the inverse image L~1(z) of any point z £ NL is a connected submanifold of JXQ (see Sections 10.8 and 10.9). PROPOSITION 2.4.6. A Hamiltonian form H weakly associated to an almost regular Lagrangian L exists if and only if the fibred manifold JXQ —» NL admits a global section [294]. D
2.4.7. The Poincare—Cartan form HL for an almost regular Lagrangian L is constant on the connected inverse image L~1(z) of any point z€NL [169; 294]. • LEMMA
COROLLARY 2.4.8. All Hamiltonian forms weakly associated to an almost regular Lagrangian L coincide with each other on the Lagrangian constraint space NL, and the Poincare-Cartan form HL (2.4.13) for L is the pull-back HL = L*H, of such a Hamiltonian form H.
Tnql -C = H{t,q\irj),
(2.4.59) •
It follows that, given Hamiltonian forms H and H' weakly associated to an almost regular Lagrangian L, their difference is fdt, / ' e I^L (see (2.1.21)). However, H\NL ^ H'\NL m general. Therefore, the Hamilton equations for H and H' do not necessarily coincide on the Lagrangian
Chapter 2 Classical Hamiltonian Systems
151
constraint space NL- Their solutions can leave the Lagrangian constraint space NL. 2.4.9. Let a section r of V*Q -> R be a solution of the Hamilton equations (2.4.35a) - (2.4.35b) for a Hamiltonian form H weakly associated to an almost regular Lagrangian L. If r lives in the Lagrangian constraint space NL, the section S = TTQOT oiQ —> R satisfies the Lagrange equations (2.4.15), while s = Hor obeys the Cartan equations (2.4.20). • THEOREM
The proof is based on the relation
L = {JlL)*LH, where L is the Lagrangian (2.4.19), while LH is the Lagrangian (2.4.42) (cf. Theorem 10.9.5). This relation is derived from the equality (2.4.59). The converse assertion is more intricate (cf. Theorem 10.9.6). THEOREM 2.4.10. Given an almost regular Lagrangian L, let a section s of the jet bundle JlQ - t i b e a solution of the Cartan equations (2.4.20). Let H be a Hamiltonian form weakly associated to L, and let H satisfy the relation ffoLos = A
(2.4.60)
where s is the projection of s onto Q. Then, the section r = L o s of the Legendre bundle V*Q —> R is a solution of the Hamilton equations (2.4.35a) - (2.4.35b) for H. D F . Systems with time-dependent parameters A configuration space of a mechanical system with time-dependent parameters is a composite fibre bundle Q -> S -> R,
(2.4.61)
coordinated by (t,am,ql) where (t,crm) are coordinates of the fibre bundle E —> R. Sections h of E —> R are time-dependent parameters, and S —> R is called the parameter bundle [175; 294]. Though Q —» R is a trivial bundle, the fibre bundle Q —> E need not be trivial. Let us recall that every section h of the parameter bundle E —> R defines the restriction Qh = h*Q of the fibre bundle Q —• S to h{M) c S, which is a subbundle ih : Qh —> Q of the fibre bundle Q —* R. One can think of the
152
Geometric and Algebraic Topological Methods in Quantum, Mechanics
fibre bundle Qh —• R as being a configuration space of a mechanical system with the fixed parameter function h(t). The velocity phase space of a mechanical system with parameters is the jet manifold JlQ of the composite bundle (2.4.61) which is equipped with the adapted coordinates (t,am,qz,a™,ql) [169]. Let the fibre bundle Q —> S be provided with a connection Av=dt® (dt + A\di) + dam ® (dm + A^fy).
(2.4.62)
Then the corresponding vertical covariant differential 5 : A ? - VxQ,
D = {q\ - A\ - A^a?)^
(2.4.63)
(cf. 10.6.80)) is defined on the configuration space Q (see Section 10.6H). Given a section h of the parameter bundle £ —* R, the restriction of D to Jxih{JlQh) C J 1 Q is precisely the familiar covariant differential on Qh corresponding to the restriction Ah = dt + {(Ain o h)dthm + (A o h)\)di
(2.4.64)
of the connection As to h(M) C E. Therefore, one may use the vertical covariant differential D in order to construct a Lagrangian for a mechanical system with parameters on the configuration space Q (2.4.61). We suppose that such a Lagrangian L depends on derivatives of parameters crj™ only via the vertical covariant differential D (2.4.63), i.e., L = C(t, am, q\ q\ - A{ - A^a^dt.
(2.4.65)
The corresponding Lagrange equations read (di - dtd\)C = 0, l
(dm - dtd m)C = 0.
(2.4.66) (2.4.67)
However, only the Lagrange equations (2.4.66) should be considered since parameter functions hold fixed. One can think of these equations as being the Lagrange equations for the Lagrangian Lh = Jxh*L on the velocity phase space JlQhThe momentum phase space of a mechanical system with parameters is the vertical cotangent bundle V*Q of Q —> R. With a connection (2.4.62), we have the splitting (10.6.79) which reads V*Q = AE{V£Q) ®{Q x V*-E), Pidq' + pmdam
= piidq* - A^da™) + (Pm + A^p^da"1.
153
Chapter 2 Classical Hamiltonian Systems
Then V*Q can be provided with the coordinates Pi=Pi,
Pm=Pm + Aimpi
compatible with this splitting. However, these coordinates fail to be canonical in general. Given a section h of the parameter bundle S —» R, the submanifold {a = h(t), pm = 0} of the momentum phase space V*Q is isomorphic to the Legendre bundle V*Qh of the subbundle h*Q of the fibre bundle Q —• R, which is the configuration space of a mechanical system with the parameter function h(t). Hamiltonian forms on the momentum phase space V*Q associated to the Lagrangian L (2.4.65) read H = (pidq1 + pmdam)
- &(# + A^T™) + PmTm + H]dt,
(2.4.68)
where Y = dt + Ymdm is some connection on the parameter bundle E —> • M and dt + ^dm + iA' + A^T^di is the composite connection (10.6.73) on the fibre bundle Q —> R defined by the connection A% (2.4.62) on Q —» E and the connection F on E —» R. The key point is that the Hamiltonian forms (2.4.68) are afiine in momenta pm. The corresponding Hamilton equations take the form
q\ = A< + ^ r
m
+ a*W,
Pu = -Pi(diAj + diAirm)
(2.4.69a) - dtH,
(2.4.69b)
crtm = T m ,
(2.4.69c)
Ptm = -PiidmA* + r n 9 m O - dmU,
(2.4.69d)
whereas the Lagrangian constraint space is given by the equations Pi = n{t,q\<Jm,&H(t,am,q\pi)),
(2.4.70)
pm + A^pi = 0.
(2.4.71)
Since parameter functions hold fixed, we ignore the equation (2.4.69d) and consider the system of equations (2.4.69a) - (2.4.69c) and (2.4.70) (2.4.71). Let the connection F in the equation (2.4.69c) be complete and admit an integral section h(t). This equation and the equation (2.4.71) define a submanifold V*Qh of the momentum phase space V*Q, which is the momentum phase space of a mechanical system with the parameter function h{t). The remaining equations (2.4.69a) - (2.4.69b) and (2.4.70) are
154
Geometric and Algebraic Topological Methods in Quantum Mechanics
the equations of this system on the momentum phase space V*Q^, which correspond to the Lagrange equations (2.4.66) in the presence of the fixed parameter function h{t). Conversely, whenever h(t) is a parameter function, there exists a connection T on the parameter bundle E —» M. such that h(t) is its integral section. Then the system of equations (2.4.69a) (2.4.69b) and (2.4.70) - (2.4.71) describes a mechanical system with the fixed parameter function h(t). Moreover, we can locally restrict our consideration to the equations (2.4.69a) - (2.4.69b) and (2.4.70). These are the Hamilton equations for the Hamiltonian form Hh = ptdq* - \pi{A' + A^dth™) + H]dt on V*Qh associated to the Lagrangian Lh on J1Qh-
G. The vertical extension of Lagrangian and Hamiltonian mechanics Given a mechanical system on a configuration space Q —> M, its extension onto the vertical tangent bundle VQ —> K of Q —> K describes the Jacobi fields of the Lagrange and Hamilton equations [170; 178; 294]. In this Section, we follow the compact notation in the formula (10.6.14). Remark 2.4.5. Let TZ be the tangent bundle of a manifold Z provided with the induced bundle coordinates (zx,zx). Due to the canonical isomorphism TT*Z = T*TZ [244], any exterior form a = -aXl...xrdzXl
/\---Adzx-
on Z gives rise to the exterior form a = i[i"
Pi^Vi,
Pi^m,
(2.4.73)
Chapter 2 Classical Hamiltonian Systems
155
can be established by inspection of the transformation laws of the holonomic coordinates (xx,yl,Pi) on V*Y and (xx,yi,vt) on VY. Accordingly, any exterior form <j> on Y gives rise to the exterior form 4>v = dv4> =
tfdicf),
(2.4.74)
called the vertical extension of <j>, onto VY so that d(f)y = (d(f))y [294; 296]. Let us also mention the canonical isomorphism J1VY = VJ1Y,
y\ = {y%.
(2.4.75)
As a consequence, given a connection T : Y —> JlY on a fibre bundle Y —> X, the vertical tangent map VT : VY -> V J x y (10.6.10) to F defines the connection VT = cfeA ® (9A + T\di + djTi^di)
(2.4.76)
on the vertical tangent bundle VY —> X. It is called the vertical connection
toT.
•
Let us start with the vertical extension of Lagrangian mechanics. Given a fibre bundle Q —> R, let us consider the extended configuration space VQ —» R, equipped with the holonomic coordinates (t,qi,qi). The corresponding velocity phase space is the jet manifold J 1 VQ, which is provided with the coordinates (t,ql,ql,ql,ql) due to the canonical isomorphism (2.4.75). Because of this isomorphism, Lagrangian formalism on JlQY = VJlQ can be developed as the vertical extension of Lagrangian formalism on JXQ. Namely, given a Lagrangian L on J^Q, its vertical extension onto V^Q is Lv = dyL = dyCdt = (tfdi + q\d\)Cdt = Cydt.
(2.4.77)
The corresponding Lagrange equations (2.4.15) read SiCy = (di - dtd\)C = SiC = 0,
(2.4.78a)
SiCy = OySiC = 0,
(2.4.78b)
8y = q^ di + q\d\ + q\tdf.
156
Geometric and Algebraic Topological Methods in Quantum
Mechanics
The equations (2.4.78a) coincide with the Lagrange equations for the original Lagrangian L, while the equations (2.4.78b) are the well-known variation equation of the equations (2.4.78a) [128; 294]. Substituting a solution s of the Lagrange equations (2.4.78a) into (2.4.78b), one obtains a linear differential equation whose solutions Is are Jacobi fields of the solution s. Indeed, if Q —> R is a vector bundle, there is the canonical splitting VQ = Q © Q over R, and s + s is a solution of the Lagrange equations (2.4.78a) modulo the terms of order > 1 in s. The momentum phase space of a mechanical system on the extended configuration space VQ is the vertical cotangent bundle V*VQ of VQ —> R. Due to the canonical isomorphism (2.4.73), the momentum phase space V*VQ is coordinated by (t,ql,Pi,ql,pi), where (q\pi) and (q\pi) are canonically conjugate pairs. Hamiltonian formalism on the vertical momentum phase space V*VQ is introduced similarly to that on the ordinary momentum phase space V*Q in Section C. In particular, a Hamiltonian form on V*VQ reads Hv = pidq1 + pidcf -
Hvdt.
The associated Hamilton vector field is 1HV =dt + tfUydi - diHvdi + dirHvdi - diHvfr.
(2.4.79)
It is a connection on the fibre bundle VV*Q —> R, and defines the corresponding Hamilton equation on V*VQ. Due to the above mentioned isomorphism (2.4.73), the canonical threeform on V*VQ can be obtained as the vertical extension flv = dvn = [dpi A dqi + dpi A dq*] A dt
(2.4.80)
of the canonical three-form 17 (2.4.30) on V*Q. The corresponding Poisson bracket on V*VQ reads {/, 9}v = d'fdig - dif&g + frfdig - djfrg.
(2.4.81)
It is important that any Hamiltonian system on V*Q gives rise to a Hamiltonian system on V*VQ = VV*Q as follows. 2.4.11. Let 7# be a Hamilton vector field (2.4.34) on the original momentum phase space V*Q —» R for a Hamiltonian form (2.4.32). Then the vertical connection PROPOSITION
VlH = dt + &Hdi - diHdi + dy&Hdi - dydiHtf
(2.4.82)
Chapter 2 Classical Hamiltonian Systems
157
to 7# on the vertical momentum phase space VV*Q —> R is the Hamilton vector field for the Hamiltonian form Hv = dvH = pidj + pidtf - dvndt,
dvH = (4%+&&)?{,
which is the vertical extension of H onto VV*Q.
(2.4.83) •
The corresponding Hamilton equations read 9j
= FHv = d{n,
(2.4.84a)
= -diHv = -5iW,
(2.4.84b)
Pti
$ = 0*« v = Sv^W,
(2.4.84c)
pti = -9iWy = -dydiH.
(2.4.84d)
The equations (2.4.84a) - (2.4.84b) coincide with the Hamilton equations (2.4.35a) - (2.4.35b) for the original Hamiltonian form H, while the equations (2.4.84c) - (2.4.84d) are their variation equation. Substituting a solution r of the Hamilton equations (2.4.84a) - (2.4.84b) into (2.4.84c) (2.4.84d), one obtains a linear dynamic equations whose solutions f are Jacobi fields of the solution r. There is the following relationship between the vertical extensions of Lagrangian and Hamiltonian formalisms [294; 296]. PROPOSITION 2.4.12. Let H be a Hamiltonian form on V*Q associated to a Lagrangian L on JXQ. Then its vertical Hamiltonian extension Hy (2.4.83) is weakly associated to the Lagrangian Ly (2.4.77). •
2.5
Constrained Hamiltonian systems
Subsections: A. Constrained Hamiltonian systems, 158; B. Dirac constrained systems, 160; C. Time-dependent constraints, 164; D. Lagrangian constraints, 167. Let (Z, ft) be a 2m-dimensional symplectic manifold and Ti a Hamiltonian on Z. Let N be a (2m — n)-dimensional closed imbedded submanifold of Z called a primary constraint space or simply a constraint space. We will consider the following two types of conservative constrained systems: • a constrained Hamiltonian system SH\N = \J{vETzN: zGJV
v\n + dH{z) = 0},
(2.5.1)
158
Geometric and Algebraic Topological Methods in Quantum Mechanics
whose solutions are solutions of the Hamiltonian system (Cl,7i) (2.3.3) on a manifold TV which live in TN; • a Dirac constrained system SN-H := U {v G TZN : v\i"N(il + dH{z)) = 0},
(2.5.2)
which is the restriction of the Hamiltonian system (Q, H) to the constraint space N, i.e., it is the presymplectic Hamiltonian system (iNfl, i'^'H) on N. Let us note that, if a non-zero presymplectic form i*N1~l of a Dirac constrained system (2.5.2) is of constant rank, N is necessarily coisotropic. A. Constrained Hamiltonian systems Let us start with the constrained Hamiltonian system (2.5.1). Remark 2.5.1. The following local relations will be useful in the sequel. Let a constraint space N be locally given by the equations a = l,...,n,
fa(z) = 0,
(2.5.3)
where fa(z) are local functions on Z called the primary constraints. Let us consider the ideal IN C C°°(Z) (2.1.21) of functions vanishing on TV. It is locally generated by the constraints fa, and its elements are locally written in the form
/=i>a/a, a=l
(2-5-4)
a
where g are functions on Z. We agree to call {/a} a local basis for the ideal IN. Let dIN be the submodule of the C°°(Z)-module O1(Z) of one-forms on Z which is locally generated by the exterior differentials df of functions f Q IN- Its elements are finite sums i
In view of the formula (2.5.4), they are given by local expressions n
° = J2^adfa + fat"),
(2.5.5)
a=l
where ga are functions and (f>a are one-forms on Z.
•
Chapter 2 Classical Hamiltonian Systems
159
A solution of the constrained Hamiltonian system (2.5.1) obviously exists if a Hamiltonian vector field -d-n, restricted to the constraint space N, is tangent to N. Then integral curves of the Hamiltonian vector field du do not leave N. This condition is fulfilled if and only if (2.5.6)
{H,IN}CIN,
i.e., if and only if the Hamiltonian H belongs to the normalizer I{N) (2.1.24) of the ideal 1^. With respect to the local basis {/a} of the ideal IN, the relation (2.5.6) reads n
tin\dfa = {Kfa} = Y,9Cafc
(2.5.7)
c=l
where <j£ are functions on Z. If the relation (2.5.6) (and, consequently, (2.5.7)) fails to hold, let us introduce secondary constraints
fi2) = {n,fa} = o. If the collection of primary and secondary constraints is not closed (i.e., {W,/o } is not expressed into fa and fa ) let us add the tertiary constraints
fi3) = {H,{H,fa}} = 0, and so on. If a solution of the constrained Hamiltonian system exists anywhere on N, the procedure is stopped after a finite number of steps by constructing the complete system of constraints. This complete system of constraints defines the final constraint space, where the Hamiltonian vector field "d-n is not transversal to the primary constraint space N. From the algebraic viewpoint, we have obtained the minimal extension 7fin of the ideal IN such that {W,/fin} C ifinIn algebraic terms, a solution of a constrained Hamiltonian system can be reformulated as follows. Let N be a closed imbedded submanifold of a symplectic manifold (Z, Q) and IN the ideal of functions vanishing everywhere on N (though any ideal of the ring C°°(Z) can be utilized). All elements of IN are said to be constraints. One aims to find a Hamiltonian, called admissible, on Z such that the symplectic Hamiltonian system (^ij'H) has a solution everywhere on JV, i.e., N is a final constraint space for (fi,W). In accordance with the condition (2.5.6), only an element of the normalizer I(N) (2.1.24) is an admissible Hamiltonian. However, the
160
Geometric and Algebraic Topological Methods in Quantum Mechanics
normalizer I(N) also contains constraints I(N) (1IN- In order to separate Hamiltonians and constraints, let us consider the overlap l'(N) = i(N)niN (2.1.25). Its elements are called the first-class constraints, while the remaining elements of IN are the second-class constraints. As was mentioned above, the set I'(N) of first-class constraints is a Poisson subalgebra of the normalizer I(N) and, consequently, of the Poisson algebra C°°(Z) on the symplectic manifold (Z, Q.). Let us also recall that IN c I'(N), i.e., products of second-class constraints are the first-class ones. Admissible Hamiltonians which do not reduce to first-class constraints are representatives of non-zero elements of the quotient I(N)/I'(N), which is the reduction of the Poisson algebra C°°(Z) via the ideal IN (see Definition 2.1.16). If H is an admissible Hamiltonian, the constrained Hamiltonian systems (Z,Cl,7i,N) is equivalent to the presymplectic Hamiltonian system (N,i*NQ,iNH), i.e., their solutions coincide. Example 2.5.2. If W is a coisotropic submanifold of Z, then IN C I(N) and I'{N) — IN- Therefore, all constraints are of first-class. The presymplectic form i*NQ, on N is of constant rank. Let its characteristic foliation be simple, i.e., define a fibration n : N —> P over a symplectic manifold (P, Clp). In view of the isomorphism (2.1.26), one can think of elements of the quotient I(N)/I'(N) as being the Hamiltonians on the physical phase space P. It follows that the restriction of an admissible Hamiltonian H to the constraint space N coincides with the pull-back onto N of some Hamiltonian H on P, i.e., i*NH = ir*H (cf. Theorem 2.3.1). Thus, (TV, i*NCl, i*NH) is a gauge-invariant Hamiltonian system which is equivalent to the reduced Hamiltonian system (P.flp,Ti,), and the original constrained Hamiltonian system (Z, Cl, H, N) is so if 7i is an admissible Hamiltonian. D B. Dirac constrained systems As was mentioned above, the Dirac constrained system SN'H (2.5.2) is really the pull-back presymplectic Hamiltonian system {UN = ijyfi, Ti-N = ^N^-)
on the primary constraint space N C Z. By virtue of Proposition 2.3.2, it has a solution only at the points of the subset N2 = {z£N
: u\dHN{z)
=0,«£
KerzQjv},
Chapter 2 Classical Hamiltonian Systems
161
which is assumed to be a manifold. Such a solution however need not be tangent to A^. Then the above mentioned constraint algorithm for presymplectic Hamiltonian systems can be called into play. Nevertheless, one can say something more since the presymplectic system SN'H (2.5.2) on N is the pull-back of a symplectic one on Z. Let us assume that a (2m—n)-dimensional closed imbedded submanifold N of Z is already a desired final constraint space of the Dirac constrained system (2.5.2), i.e., the equation v\ nN + <mN (z) = o,
V&TZN,
has a solution at each point z e N. As was mentioned above, this is equivalent to the injection Kerfi,v= TN n Orth^TTV c Ker dUN-
(2.5.8)
Let us reformulate this condition in algebraic terms of the ideal of constraints IN (2.1.21), its normalizer I(N) (2.1.24) and the Poisson algebra of first-class constraints I'{N) (2.1.25). It is readily observed that, restricted to N, Hamiltonian vector fields $/ of elements / of I'(N) with respect to the symplectic form fi on Z take their values into TN n OrthaTN [245]. Then the condition (2.5.8) can be written in the form {H,I'(N)}cIN.
(2.5.9)
At the same time, {H, IN} <£ IN i n general. This relation shows that, though the Dirac constrained system (fijv,Wjv) on N has a solution, the Hamiltonian vector field d-n of the Hamiltonian Ji on Z is not necessarily tangent to N, and its restriction to N need not be such a solution. The goal is to find a constraint f £ IN such that the modified Hamiltonian 7i + f would satisfy the condition {H + f,IN}dN
(2.5.10)
and, consequently, the condition {H + f,I'(N)}cIN.
(2.5.11)
It is called a generalized Hamiltonian system. The condition (2.5.11) is fulfilled for all / e IN, while (2.5.10) is an equation for the second-class constraint / . Therefore, its solution implies separating first- and second-class constraints. The general difficulty lies in the fact that the set of elements generating IN C I'(N) is necessarily
162
Geometric and Algebraic Topological Methods in Quantum Mechanics
infinitely reducible [245]. At the same time, the Hamiltonian vector fields for elements of IN vanish on the constraint space N. Therefore, one can employ the following procedure [328]. Since N is a (2m — n)-dimensional closed imbedded submanifold of Z, the ideal IN is locally generated by a finite basis {/ a }, a = 1 , . . . , n, whose elements determine N by the local equations (2.5.3). Let the presymplectic form Jljv be of constant rank 2m — n — k. It defines a A>dimensional characteristic foliation of JV. Since N c Z is closed, there exist locally k linearly independent vector fields u;, on Z which, restricted to TV, are tangent to the leaves of this foliation. They read n U
6=1,...,*,
b = J2^fa, O= l
where g% are local functions on Z and i?/o are Hamiltonian vector fields of the constraints /„. Then one can choose a new local basis {
6 = 1,...,fc.
Let t?0b be their Hamiltonian vector fields. One can easily justify that tftfjiv = Ub\ff,
b=l,...,k.
It follows that the constraints <j>b, b = 1 , . . . , k, belong to I'(N) \ IN, i.e., they are first-class constraints, while the remaining ones
{
6=1,...,*,
C=l,...,n,
a=l
where C£c are local functions on Z. It should be emphasized that the firstclass constraints <j>i,- • • ,<j>k do not constitute any local basis for I'(N). Now let us consider a local Hamiltonian on Z n
W' = W + ^ A > a ,
(2.5.12)
o=l
where A° are functions on Z. Since H obeys the condition (2.5.11), we find n
{«,&}= X>6&" o=l
6=l...,fc,
Chapter 2 Classical Hamiltonian Systems
163
where B% are functions on Z. Then the equation (2.5.10) takes the form n
{K
Y, a=fe+l
A
n
°fa«.0c}= J 2 D c ^
c = k + l,...n,
(2.5.13)
6=1
where Dbc are functions on Z. It is a system of linear algebraic equations for the coefficients A°, a = k + 1,... n, of second-class constraints. These coefficients are defined uniquely by the equations (2.5.13), while the coefficients Aa, a = l,...,k, of first-class constraints in the Hamiltonian W (2.5.12) remain arbitrary. Then, restricted to the constraint space N, the Hamiltonian vector field of the Hamiltonian 7i' (2.5.12) on Z provides a local solution of the Dirac constrained system on N. We refer the reader to [328] for a global variant of the above procedure. A generalized Hamiltonian system (Z, 0, H + f, N) is a constrained Hamiltonian system with an admissible Hamiltonian H + f. It is equivalent to the original Dirac constrained system. Example 2.5.3. Let the final constraint space N be a coisotropic submanifold of the symplectic manifold (Z, fi). Then IN — I'(N), i.e., there are only first-class constraints. In this case, the Hamiltonian vector fields both of the Hamiltonian H and all the Hamiltonians H + f,f£lN, provide solutions of the Dirac constrained system on N. If the characteristic foliation of the presymplectic form i*NH on N is simple, we have the reduced Hamiltonian system equivalent to the original Dirac constrained one (see Example 2.5.2). • Example 2.5.4. If N is a symplectic submanifold of Z, then I'(N) = 1^. Therefore, all constraints are of second-class, and the Hamiltonian (2.5.12) of a generalized Hamiltonian system is defined uniquely. • Remark 2.5.5. In concrete models, the final constraint space N fails to be given in advance. Therefore, a different procedure of separating first- and second-class constraints is usually applied [407], but its global treatment is under question. D
164
Geometric and Algebraic Topological Methods in Quantum Mechanics
C. Time-dependent constraints Given a time-dependent mechanical system on a configuration bundle Q —» R, time-dependent constraints on the momentum phase space V*Q can be described similarly to those in autonomous Hamiltonian mechanics. Let N be a closed imbedded subbundle iN:N
^V*Q
of the Legendre bundle V*Q —> R, treated as a constraint space. It cannot be neither Lagrangian nor symplectic submanifold with respect to the Poisson structure on V*Q. Let us consider the ideal IN of real functions / on V*Q which vanish on JV, i.e., iNf = 0. Its elements are said to be constraints. There is the isomorphism C°°{V*Q)/IN S* C°°(N)
(2.5.14)
of associative commutative algebras. By the normalize I(N) of the ideal IN is meant the subset of functions of C°°(V*Q) whose Hamiltonian vector fields are restricted to vector fields on N, i.e., I(N) = {fe C°°(V*Q) : {f,g}v
€ IN, g € IN}.
(2.5.15)
It follows from the Jacobi identity that the normalizer (2.5.15) is a Poisson subalgebra of C°°(V*Q). Let us put I'{N) = I(N)nIN.
(2.5.16)
It is naturally a Poisson subalgebra of I(N). Its elements are called the first-class constraints, while the remaining elements of IN are the secondclass constraints. It is readily observed that 7jy C I'(N), i.e., the products of second-class constraints are first-class constraints. Remark 2.5.6. Let N be a coisotropic submanifold of V*Q, i.e., w*(AnnTN) C TN. Then IN = I'(N), i.e., all constraints are of first class. • The relation (2.4.28) enables us to extend the constraint algorithms of conservative mechanics and time-dependent mechanics on a product K x M (see [101; 278]) to mechanical systems subject to time-dependent transformations. Let H be a Hamiltonian form on the momentum phase space V*Q. In accordance with the relation (2.4.41), a constraint / G IN is preserved if the
Chapter & Classical Hamiltonian Systems
165
bracket {Ti*,(*f}T vanishes. It follows that the solutions of the Hamilton equations (2.4.35a) - (2.4.35b) do not leave the constraint space N if {H*,CIN}TCCIN-
(2.5.17)
If the relation (2.5.17) fails to hold, let us introduce secondary constraints {W*,C*/}r, / € IN, which belong to CC°°(V*Q). If the collection of primary and secondary constraints is not closed with respect to the relation (2.5.17), let us add the tertiary constraints {7i*, {W*,C*/a}r}r, and so on. Let us assume that N is a final constraint space for a Hamiltonian form H. If a Hamiltonian form H satisfies the relation (2.5.17), so is a Hamiltonian form Hf=H-
fdt
(2.5.18)
where / e I'{N) is a first-class constraint. Though Hamiltonian forms H and Hf coincide with each other on the constraint space TV, the corresponding Hamilton equations have different solutions on the constraint space N because dH\N^dHf\N. At the same time, we have d(i*NH) = d(i*NHf). Therefore, let us introduce the constrained Hamiltonian form HN = i*NHf
(2.5.19)
which is the same for all / 6 I'(N). Let us note that H^ (2.5.19) is not a true Hamiltonian form on TV —> R in general. On sections r of the fibre bundle N —» E, we can write the equations r*(uN\dHN) = 0,
(2.5.20)
where u^ is an arbitrary vertical vector field on JV —> R. They are called the constrained Hamilton equations. PROPOSITION 2.5.1. For any Hamiltonian form Hf (2.5.18), every solution of the Hamilton equations which lives in the constraint space N is a solution of the constrained Hamilton equations (2.5.20). •
166
Geometric and Algebraic Topological Methods in Quantum Mechanics
Outline of proof. The constrained Hamilton equations can be written as r*{uN\di*NH}) = r*(uN\dHf\N)
= 0.
(2.5.21)
They differ from the Hamilton equations (2.4.35a) - (2.4.35b) for Hf restricted to N which read r*(u\dHf\N)=0,
(2.5.22)
where r is a section of TV —> M. and u is an arbitrary vertical vector field on V*Q —» R. A solution r of the equations (2.5.22) satisfies obviously the weaker condition (2.5.21). QED
Remark 2.5.7. One also can consider the problem of constructing a generalized Hamiltonian system, similar to that for Dirac constrained system in conservative mechanics [294]. Let H satisfy the condition {H*,CI'(N)}T C IN, whereas {H*,CIN}T
{n* + Cf,CiNhcCiN. This is an equation for a second-class constraint / .
•
The construction above, except the isomorphism (2.5.14), can be applied to any ideal J of C°°{V*Q). Then one says that the Poisson algebra J"/J' (see Definition 2.1.16) is the reduction of the Poisson algebra C°°(V*Q) via the ideal J. In particular, an ideal J is said to be coisotropic if it is a Poisson algebra. In this case, J is a Poisson subalgebra of the normalize J " (2.5.15), and it coincides with J' (2.5.16). Let A be a Lie algebra of generators u of gauge symmetries of a Hamiltonian form H. In accordance with the relation (2.4.47), the corresponding symmetry currents i u (2.4.46) on V*Q constitute a Lie algebra with respect to the canonical Poisson bracket on V*Q. Let I_\ denotes the ideal of C°°(V*Q) generated by these symmetry currents. It is readily observed that this ideal is coisotropic. Then one can think of IA as being an ideal of first-class constraints compatible with the Hamiltonian form H, i.e., (2.5.23)
{H*,CIA}TCCIA.
Let us note that any Hamiltonian form HU=H - iudt,
u£A,
Chapter 2 Classical Hamiltonian Systems
167
obeys the same relation (2.5.23), but other currents %u' are not conserved with respect Hu if [u, u'} ^ 0. Let now A be an arbitrary Lie algebra of vertical vector fields u on the configuration bundle Q -* E. The relation (2.4.47) remains true, while the corresponding symmetry currents i u (2.4.48) on V*Q constitute a Lie algebra, and they generate the corresponding coisotropic ideal IA of C°°(V*Q) with respect to the canonical Poisson bracket on V*Q. PROPOSITION 2.5.2. Let A be a finite-dimensional Lie algebra of vertical vector fields on the configuration bundle Q —» R. If there exists a connection F on Q —> E such that [F, A] = 0, then there exists a Hamiltonian form H (2.4.37) on the Legendre bundle V*Q such that Hr J= 0 and A is the algebra of symmetries of H. • Outline of proof. Let A be the universal enveloping algebra of the Lie algebra of the symmetry currents %u, u £ A, (2.4.48). Then each non-zero element C of its center of order > 1 can be written as a polynomial in %u, QED and defines the desired Hamiltonian form H = Hr — Cdt.
D. Lagrangian constraints Let us study the important case of almost regular quadratic Lagrangians. We show that, in this case, there always exist both a complete set of associated Hamiltonian forms and a complete set of non-degenerate weakly associated Hamiltonian forms (see Section 2.4E). Given a configuration bundle Q —> R, let us consider a quadratic Lagrangian L which has the coordinate expression
£ = -TfidA + hqi + c,
(2.5.24)
where a, b and c are local functions on Q. This property is coordinateindependent due to the affine transformation law of the coordinates q\. The associated Legendre map PioL = ai:jqi + bi
(2.5.25)
is an affine morphism over Q. It defines the corresponding linear morphism L:VQ^V*Q,
Pi
oL = cujqi.
(2.5.26)
168
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let the Lagrangian L (2.5.24) be almost regular, i.e., the matrix function aij is of constant rank. Then the Lagrangian constraint space NL (2.5.25) is an affine subbundle of the bundle V*Q —» Q, modelled over the vector subbundle iVL (2.5.26) of V*Q -> Q. Hence, NL^>Q has a global section. For the sake of simplicity, let us assume that it is the canonical zero section 0(Q) of V*Q -> Q. Then ~NL = NL, Accordingly, the kernel of the Legendre map (2.5.25) is an affine subbundle of the affine jet bundle JXQ —> Q, modelled over the kernel of the linear morphism L (2.5.26). Then there exists a connection r : Q -> Ker L C JlQ,
(2.5.27)
aijTi
(2.5.28)
+bi=0,
on Q —> M. Connections (2.5.27) constitute an affine space modelled over the linear space of vertical vector fields v on Q —» K, satisfying the conditions atjvj = 0
(2.5.29)
and, as a consequence, the conditions t/6j — 0. If the Lagrangian (2.5.24) is regular, the connection (2.5.27) is unique. The matrix a in the Lagrangian L (2.5.24) can be seen as a degenerate fibre metric of constant rank in VQ —> Q. Then the following holds. LEMMA 2.5.3. Given a fc-dimensional vector bundle E —> Z, let a be a 2
section of rank r of the tensor bundle V E* —> Z. There is a splitting E = Kera®E',
z
(2.5.30)
where E' = E/Kei a is the quotient bundle, and a is a non-degenerate fibre O metric in E'. Outline of proof. Since a exists, the structure group GL(k, M.) of the vector bundle E —> Z is reducible to the subgroup GL(r, k — r; M) of general linear transformations of Rk which keep its r-dimensional subspace, and to its subgroup GL(r, R) x GL(k - r, R). QED
THEOREM
2.5.4. There exists a linear bundle map a : V*Q -> VQ,
tf
o a = aijPj,
(2.5.31)
169
Chapter 2 Classical Hamiltonian Systems
such that LoaoiN
= iN.
O
Outline of proof. The map (2.5.31) is a solution of the algebraic equations (2.5.32)
CLijC-^akb = aib.
By virtue of Lemma 2.5.3, there exists the bundle slitting (2.5.33)
VQ = Kera®E'
and a (non-holonomic) atlas of this bundle such that transition functions of Kera and E' are independent. Since a is a non-degenerate fibre metric in E', there exists an atlas of E' such that a is brought into a diagonal matrix with non-vanishing components CIAA- Due to the splitting (2.5.33), we have the corresponding bundle splitting F* = (Kera)*eIma.
(2.5.34)
Then the desired map a is represented by a direct sum a\®a0 of an arbitrary section o\ of the bundle VKera* -> Q 2
and the section CFQ of the bundle V £ ' —> Q, which has non-vanishing components aAA = (aAA)~l with respect to the above mentioned atlas of E'. Moreover, a satisfies the particular relations CTQ = O-QO Lo cr0,
a o a\ = 0,
o\ o a = 0.
(2.5.35) QED
COROLLARY
2.5.5. The splitting (2.5.33) leads to the splitting J1Q = S(J1Q)®F{J1Q) Q
=KerL©Im(croZ), Q
qi = Si + Fi= fat - °ok(akj
(2.5.36a)
(2.5.36b) + bk)},
while the splitting (2.5.34) can be written as V*Q = 1l(V*Q) ® V{V*Q) - Ker aQ © NL, Q
Pi=ni+Vi
Q
= \pi-
(2.5.37a) (2.5.37b)
170
Geometric and Algebraic Topological Methods in Quantum Mechanics
• It is readily observed that, with respect to the coordinates <S^ and Tlx (2.5.36b), the Lagrangian (2.5.24) reads £ = -dijPP + c',
(2.5.38)
while the Lagrangian constraint space is given by the reducible constraints Tli=Pi-
ai:jaikpk = 0.
(2.5.39)
Given the linear map a (2.5.31) and the connection T (2.5.27), let us consider the affine Hamiltonian map $ = f + a : V*Q -> JXQ,
$* = P + a^Pj,
(2.5.40)
and the Hamiltonian form H = # $ + $*L =
(2.5.41)
Pidq1 - [p i r i + laoijPiPj + a!ijpiPj (Ki + ViW
- [(Hi + vjr
- c']dt =
+^ v ^
+ a\jpiPj - c']dt.
In particular, if o\ is non-degenerate, so is the Hamiltonian form (2.5.41). THEOREM 2.5.6. The Hamiltonian forms (2.5.41) parameterized by connections F (2.5.27) are weakly associated with the Lagrangian (2.5.24) and constitute a complete set. •
Outline of proof. By the very definitions of Y and a, the Hamiltonian form (2.5.41) satisfies the conditions (2.4.58a) - (2.4.58b). Let us consider the corresponding Hamilton equations (2.4.35a) for a section r of the Legendre bundle V*Q -> R. They are dts = (f + a)or,
(2.5.42)
s = nQor.
Due to the surjections S and T (2.5.36a), the Hamilton equations (2.5.42) break in two parts Sodts
= Yos,
JTo dts = a o r,
dtri-aik(akjdtr> ik
j
a (akjdtr
+bk)=Tios, ik
+ bk) = a rk.
(2.5.43) (2.5.44)
Let s be an arbitrary section of Q -* R, e.g., a solution of the Lagrange equations. There exists a connection T (2.5.27) such that the relation
Chapter 2 Classical Hamiltonian Systems
171
(2.5.43) holds, namely, T = S o T' where V is a connection on Q -» R which has s as an integral section. It is easily seen that, in this case, the Hamiltonian map (2.5.40) satisfies the relation (6.3.1) for s. Hence, the Hamiltonian forms (2.5.41) constitute a complete set. QED It is readily observed that, if a\ = 0, then the Hamiltonian forms (2.5.41) are associated to the Lagrangian (2.5.24). Thus, for different <TI, we have different complete sets of Hamiltonian forms (2.5.41). Hamiltonian forms H (2.5.41) of such a complete set differ from each other in the term vl1Zi, where v are vertical vector fields (2.5.29). If follows from the splitting (2.5.37a) that this term vanishes on the Lagrangian constraint space. The corresponding constrained Hamiltonian form HN = i*NH and the constrained Hamilton equations (2.5.20) can be written. In the case of quadratic Lagrangians, we can improve Proposition 2.5.1 as follows. PROPOSITION 2.5.7. For every Hamiltonian form H (2.5.41), the Hamilton equations (2.4.35b) and (2.5.44) restricted to the Lagrangian constraint • space NL are equivalent to the constrained Hamilton equations.
Outline of proof. Due to the splitting (2.5.37a), we have the corresponding splitting of the vertical tangent bundle VQV*Q of the bundle V*Q —> Q. In particular, any vertical vector field u on y*Q - t R admits the decomposition u = [u — UTN] + UTN,
UTN = u'di + a^a^ Ufcd\
such that UN = UTN\NL is a vertical vector field on the Lagrangian constraint space NL —> IR. Let us consider the equations r*(uTN\dH) = 0,
(2.5.45)
- R and u is an arbitrary vertical vector field where r is a section of V*Q —> on V*Q —> R. They are equivalent to the pair of equations ^(aijO-ifd^dH) = 0,
(2.5.46a)
r*(di\dH)=0.
(2.5.46b)
The equations (2.5.46b) are obviously the Hamilton equations (2.4.35b) for H. Bearing in mind the relations (2.5.28) and (2.5.35), one can easily show that the equations (2.5.46a) coincide with the Hamilton equations (2.5.44). The proof is completed by observing that, restricted to the Lagrangian
172
Geometric and Algebraic Topological Methods in Quantum Mechanics
constraint space NL, the equations (2.5.45) are exactly the constrained QED Hamilton equations (2.5.20). Proposition 2.5.7 shows that, restricted to the Lagrangian constraint space, the Hamilton equations for different Hamiltonian forms (2.5.41) associated to the same quadratic Lagrangian (2.5.24) differ from each other in the equations (2.5.43). These equations are independent of momenta, and they play the role of gauge-type conditions. 2.6
Geometry and cohomology of Kahler manifolds
Subsections: A. Complex structure on a vector space, 172; B. Almost complex manifolds, 174; C. Hermitian manifolds, 179; D. Kahler manifolds, 180; E. Cohomology of Kahler manifolds, 183; F. Hyper-Kahler manifolds, 187. This Section addresses geometry and cohomology of finite-dimensional Kahler manifolds [250; 437; 448]. A. Complex structure on a vector space Let V be a 2m-dimensional real vector space. A complex structure on V is an automorphism J of V such that J2 = —1. With this complex structure, V is naturally brought into an m-dimensional complex vector space by letting (a + i(3)v = av + (3Jv,
v € V,
a,/?e!R.
(2.6.1)
Of course, an m-dimensional complex vector space is a 2m-dimensional real vector space provided with the complex structure J = i l . Equipped with a complex structure J, a vector space V has an adapted basis (vi,...,vm,Jvi,...,Jvm) such that the automorphism J is given by the matrix
j-f J
° M
~{-im o ) '
where lm denotes the unit (m x m)-matrix. An automorphism p of V preserves a complex structure J on V if and only if it commutes with J. These automorphisms form the commutant
173
Chapter 2 Classical Hamiltonian Systems
{J}1 C GL(2m,R) of J which is the image of the group GL(m,C) under the monomorphism GL(m,C) 3 M ~ (^™
" ^ ) £ GL(2m,R).
(2.6.2)
Hence, there is one-to-one correspondence between the complex structures on a 2m-dimensional real vector space V and the elements of the quotient GL(2m,R)/GL(m,C)A complex structure J on V generates a complex structure on the dual V* of V by the law (v, Jw) = (Jv,w),
v G V,
w £ V*.
A scalar product h on a real vector space V with a complex structure J is called Hermitian if it is J-invariant, i.e., h{Jv,Jv') = h(v,v'),
v,v'eV.
It follows immediately that h(Jv, v) = 0 for all v £ V. Moreover, V admits an adapted basis, which is orthonormal with respect to h. We also have the skew-symmetric bilinear form (2.6.3)
tt(v,v') :=h{Jv,v')
on V. It is J-invariant. Let Vc = C
(2.6.4)
of complex holomorphic and antiholomorphic subspaces V1-0 = {v - i Jv : v € V},
V 0 ' 1 = {v + i Jv : v e V}.
These are the eigenspaces of J characterized by the eigenvalues i and —i, respectively. There is the antilinear complex conjugate morphism v = vi + iv2 t-+ v = vi — iv2,
Vr's 3 v i-> u € l^ s ' r ,
r, s = 0,1,
such that Jv = Ju. The complexification V£ of the dual V* of V is the complex dual of Vc • The decomposition (2.6.4) of Vc yields the dual decomposition Vc* = Vii0 © V0,i
(2.6.5)
174
Geometric and Algebraic Topological Methods in Quantum Mechanics
of VQ , where Vifl = {w — iJw : w G V*},
Vb,i = {w + Uw : w € V*}
are the annihilators of V0'1 and V 1 ' 0 , respectively. They are the eigenspaces of the complex structure J on Vc* characterized by the eigenvalues i and —i, respectively. A Hermitian scalar product h on V is uniquely extended to a symmetric complex J-invariant bilinear form on Vc which fulfils the following conditions: • h(v,v') = h(v,v') for all v,v' G Vc; • h(v,v) > 0 for all non-vanishing v G Vc; • h(v, v1) = 0 if v, v' are simultaneously holomorphic or antiholomorphic. This complex bilinear form defines a non-degenerate Hermitian form
(v\v')h := h(v,v') on Vfc such that the subspaces V 0 ' 1 and V 0 ' 1 are mutually orthogonal. Accordingly, the skew-symmetric form (2.6.3) is extended to Vc so that
n(v,v') = n(v,v'), v,v'£Vc, n{v,v') = 0, v,v' GVr's, r,s = 0,1. B. Almost complex manifolds Let Z be a 2m-dimensional smooth real manifold, coordinated by (z*), i = 1 , . . . , 2m. An almost complex structure on Z is defined as a vertical bundle automorphism J of the tangent bundle TZ such that
JoJ = -ldTZ. In accordance with the relation (10.6.22), an almost complex structure J on a manifold Z is represented by a tangent-valued form J = Jlkdzk®di,
(2.6.6)
jy^-8),
on Z. By means of the relation (10.6.23), this tangent-valued form defines an automorphism J of the cotangent bundle T*Z of Z such that {v, Jw) = {Jv,w),
v G TZZ,
w G T*Z,
ze Z.
An almost complex structure provides Z with an orientation associated to the adapted fibre bases for TZ. The pair (Z, J) is called an almost complex
175
Chapter 2 Classical Hamiltonian Systems
manifold. A diffeomorphism / of Z preserves an almost complex structure J on Z if and only if the tangent morphism Tf commutes with J. Let ?r : LZ —» Z be the fibre bundle of linear frames in the tangent bundle over a manifold Z. This is a principal bundle with the structure group GL(2m, E). In view of the monomorphism (2.6.2), there is one-toone correspondence between the almost complex structures J on Z and the global sections hj of the quotient bundle £ = LZ/GL(m, C) —> Z. This means that, in accordance with Theorem 10.10.5, an almost complex structure on a manifold Z exists if and only if the structure group GL(2m, R) of the linear frame bundle LZ (and of the tangent bundle TZ) is reducible to the subgroup GL(m,C) (2.6.2). Given J, the corresponding GL(m,C)principal subbundle LJZ consists of all J-invariant frames in TZ. Example 2.6.1. Any complex manifold Z, coordinated by (zx), A = 1,... ,m, can be seen as a real manifold, provided with coordinates and equipped with the almost complex structure (Rezx,lmzx) \dRezx)
~ dlmzx'
x
J(dRez ) = -dlmz\
\dlmzx) x
dRezx' x
J(dlm z ) = dRez .
K
'
(2.6.8)
Indeed, this almost complex structure commutes with the tangent morphisms to holomorphic transition functions of coordinates (Rez\lm,z A ), i.e., which obey the Cauchy-Riemann conditions dRez'*1 dlmz'v dRez" ~ dlmz1' ~ '
dRez'» dlmzv
dlmz'» _ dRezu ~
• An almost complex structure J on a manifold Z is said to be integrable or a complex structure if the tangent bundle TZ can be endowed with holonomic frames {d/dzl}, i = 1 , . . . , 2m, adapted to J. PROPOSITION 2.6.1. A manifold Z admits a complex structure J if and only if (Z, J) is obtained from a complex manifold just as in Example 2.6.1 [250]. • Therefore, if J is a complex structure on Z, one calls (Z, J) a complex manifold or a holomorphic manifold because of holomorphic transition functions of coordinates (Rez^Imz*).
176
Geometric and Algebraic Topological Methods in Quantum Mechanics
2.6.2. An almost complex structure J on a manifold Z is a complex structure if and only if the Nijenhuis differential THEOREM
N = djj = [J, J ] F N =
(2.6.9) k
{J^dhJi - 4dhJ] - Jfa J% + Jidktydz* A dz ® du of the tangent-valued form J (2.6.6) vanishes [250].
D
The Nijenhuis differential (2.6.9) is called the Nijenhuis torsion of the almost complex structure J. It appears to be related to the torsion of a connection on Z preserving the almost complex structure J. An afnne connection on an almost complex manifold (Z, J) is called almost complex if J is parallel with respect to this connection. 2.6.3. [250]. Any almost complex manifold (Z,J) admits an almost complex affine connection whose torsion is T = —N/8, where N is the Nijenhuis torsion (2.6.9). Furthermore, an almost complex manifold admits a torsionless almost complex connection if and only if it is a complex manifold. • THEOREM
The complexification CTZ —> Z of the tangent bundle of a manifold Z is called the complex tangent bundle, and its elements are said to be complex tangent vectors. An almost complex structure J on a manifold Z is naturally extended to CTZ as the tangent-valued form (2.6.6). Let TZ1'0 (resp. TZ0'1) denote a subbundle of CTZ which consists of eigenvectors of J characterized by the eigenvalue i (resp. — i). Any element of TZ1'0 (resp. TZ0'1) takes the form •& = v - iJv (resp. i9 = v + iJv) where v £ TZ, and vice versa. The subbundle TZ1'0 (resp. TZ0'1) is called a holomorphic tangent bundle (resp. an antiholomorphic tangent bundle), and elements of TZ1'0 (resp. TZ0'1) are said to be holomorphic (resp. antiholomorphic) tangent vectors. We have the decomposition CTZ = TZ1'°@TZ0'1,
(2.6.10)
together with the antilinear bundle morphism TZr's
3 tf ^ d e TZs'r,
r,s = 0 , 1 .
In accordance with this decomposition, each complex tangent vector d G CTZ is represented uniquely by a sum -d — v + u, v,u G TZ1>0, of its holomorphic and antiholomorphic parts. A complex tangent vector is called real if $ = v + v = "#.
177
Chapter 2 Classical Hamiltonian Systems
Let CT*Z —> Z be the complexification of the cotangent bundle of a manifold Z. It is called the complex cotangent bundle. An almost complex structure J on Z is induced on CT*Z by the law (v, Jw) = (Jv,w),
v G CTZZ,
w G CT*Z,
z £ Z.
Then CT*Z is split into the direct sum CT*Z = TZifl © TZ0,i
(2.6.11)
of holomorphic and antiholomorphic subbundles. These are the annihilators of antiholomorphic and holomorphic tangent bundles TZ0'1 and TZ1'0, respectively. Accordingly, the graded commutative algebra C*(Z) of complex exterior forms on Z has the decomposition C*{Z)=
p,q=m
© CP'"{Z)
(2.6.12)
p,q=o
into subspaces CP'9(Z) of exterior (p,q)-forms, defined as sections of the fibre bundle ATZi )0 ATZ 0 ,i -> Z. By (0,0)-forms are meant smooth complex functions on Z. With respect to the decomposition (2.6.12), the conjugation map, the complex structure map and the exterior differential respectively read cp'q{Z)
9^M^gc«'p(Z)1
Jcf> = ( - 1 ) " ^ + ^ ,
J2
dCp>q{Z) c Cp+2'g-1(Z) © Cp+1'q{Z) © Cp'q+1(Z) © C p - l l 9 + 2 (Z).
PROPOSITION 2.6.4. [250]. An almost complex structure J on a manifold Z is a complex (i.e., torsionless) structure if and only if the following equivalent conditions hold: (i) the distribution TZ1'0 is involutive, (ii) the distribution TZ 0 ' 1 is involutive, (iii) dCx-°{Z) C C2-°(Z) ®C1-\Z) and dC^x{Z) c CX<\Z) ©C°-2(Z). D
The condition (iii) in Proposition 2.6.4 implies the relation dCp'q(Z)
C Cp+1'q(Z)
© CP'9+1(Z),
p,q =
0,...,m.
178
Geometric and Algebraic Topological Methods in Quantum Mechanics
As a consequence, the exterior differential d on C*{Z) can be split into a sum d = d + d of holomorphic and antiholomorphic differentials d = i ( d - {-l)p+qUdJ)
: Cp'q(Z) -> C p+1 ' 9 (2'),
3 = l(d+(-l)P+HJdJ)
: Cp'g{Z) -> C P ' 9+1 (Z)
such that
J o 9 = —i9o J,
Jo d = ido J,
d(i> = d~(jj,
d = d~$.
With the differentials 9 and 9, the graded commutative algebra C*(Z) becomes a bicomplex. Let us recall the following notions. • A smooth complex function / on a complex manifold (Z, J) is called holomorphic (resp. antiholomorphic) if df = 0 (resp. df = 0). Accordingly, an exterior (p, 0)-form (resp. a (0, (?)-forin) 0 on a complex manifold (Z, J) is called holomorphic (resp. antiholomorphic) if c?(/> = 0 (resp. dcf> = 0). • A TZ1'°-valued (resp. TZ1'°-valued) vector field i ) o n a complex manifold (Z, J) is said to be holomorphic (resp. antiholomorphic) if $J
*=H^-w)' a
_ 1(
d
, .
dzx = dRe zx + idlm zx,
d
(26 i3)
'-
\ dzx = dRe zx - idlm zx.
(2.6.14)
Chapter 2 Classical Hamiltonian Systems
179
The following relations hold: J(%) = - i % ,
J(dx) = idx, x
x
J(dz ) = idz , d»\dzv = 6£,
J{dzx) = -idzx, dp\dzv = 8%,
d»\dzv = djj\dzv = 0.
The tuples {d\}, {^} play the role of holonomic bases for holomorphic and antiholomorphic tangent bundles TZ1'0 and TZ0'1, while {dzx}, {dzx} are the dual bases for holomorphic and antiholomorphic cotangent bundles TZIQ and TZ$t\, respectively. Relative to these bases, the tangent-valued form J reads J = i(dzx
(2.6.15)
Accordingly, holomorphic and antiholomorphic differentials read d{(j))=dzxAdx
d(
(2.6.16)
In particular, a holomorphic (resp. antiholomorphic) vector field on Z takes the coordinate form $ = tixd\ (resp. i? = i?A%), where dx are holomorphic (resp. antiholomorphic) local functions on Z. Accordingly, a holomorphic (resp. antiholomorphic) exterior form is an element of C*'°(Z) (resp. C°'*(Z)) with holomorphic (resp. antiholomorphic) coefficients.
C. Hermitian manifolds A Riemannian metric g on an almost complex manifold (Z, J) is called Hermitian if it is J-invariant, i.e., g(v,v')=g(Jv,Jv'),
v,v'eTzZ,
z G Z.
(2.6.17)
Clearly, a Hermitian metric yields a Hermitian scalar product on each tangent space TZZ, z € Z. An almost complex (resp. complex) manifold, provided with a Hermitian metric, is called an almost Hermitian (resp. Hermitian) manifold. 2.6.5. Any almost complex manifold (Z, J) admits a Hermitian metric, i.e., it is an almost Hermitian manifold [250]. D THEOREM
Indeed, the existence of an almost complex structure on a manifold Z implies the corresponding contraction of the structure group GL(2m,]R) of the linear frame bundle LZ to the subgroup GL(m, C) (2.6.2) and, consequently, to the unitary group U(m). This means that the structure group
180
Geometric and Algebraic Topological Methods in Quantum Mechanics
of LZ is reduced to the orthogonal group O(2m) D U(m). In accordance with Theorem 10.10.5, this contraction defines a Riemannian metric g on Z. We have monomorphisms LgZ -» LJZ -» LZ of the corresponding reduced subbundles L9Z and L J Z . It follows that L9Z consists of all #-orthonormal and J-invariant frames in TZ. Hence, g is a Hermitian metric with respect to a given almost complex structure J on Z. Moreover, there is one-to-one correspondence between such metrics and the global sections of the quotient bundle LJZ/U{m)
-> Z.
Given an almost Hermitian manifold (Z,J,g), a Hermitian metric g defines the non-degenerate two-form (i.e., an almost symplectic form) Q(0,0')=g(JW),
V,#'eCTzZ,
zeZ,
(2.6.18)
on Z. It is J-invariant, i.e., 9,{v, v') = Sl{Jv, Jv'),
v, v' e TZZ,
ze Z,
(2.6.19)
and is called the fundamental form of the Hermitian metric g. An almost complex structure, a metric and an almost symplectic form, obeying the compatibility conditions (2.6.17) - (2.6.19), are three main "players" blending to form an almost Hermitian structure on a manifold Z. Any one of them is entirely determined by the remaining two. A linear connection on an almost Hermitian manifold (Z,J,g,Q) is called Hermitian if it preserves any two of the forms J, g, Cl and, consequently, the remaining one. It follows that Hermitian connections on an almost Hermitian manifold (Z, J, g, fi) differ from each other in a torsion, and they are classified by the form of their torsion tensor [163; 196]. 2.6.6. A Hermitian manifold admits a unique Hermitian connection whose torsion tensor is totally skew-symmetric [163]. It is called a KT-connection or a Bismut connection. • THEOREM
D. Kahler manifolds Given an almost Hermitian manifold (Z,J,g), a Hermitian metric g is called a Kahler metric if its fundamental form Cl (2.6.18) is closed, i.e., is
Chapter 2 Classical Hamiltonian Systems
181
a symplectic form. A closed fundamental form is called the Kdhler form. Accordingly, an almost Hermitian manifold (resp. a Hermitian manifold) with a closed fundamental form is said to be an almost Kahler manifold (resp. a Kahler manifold). In view of the forthcoming theorem, one can start with a symplectic form on a manifold in order to make it into an almost Kahler manifold [311; 415]. THEOREM 2.6.7. Any symplectic manifold (Z, Q) admits a compatible almost complex structure J and, thus, it is an almost Kahler manifold with • a Kahler metric [199]. At the same time, the existence of a Kahler structure imposes additional topological restrictions on a symplectic manifold (Z, Q) (see Proposition 2.6.10 below). 2.6.8. An almost Hermitian manifold (Z, J,g, Cl) is a Kahler manifold if and only if the Levi-Civita connection of the Hermitian metric g is a Hermitian connection [250]. • THEOREM
Given a Kahler manifold (Z,J,g,£l), the Kahler metric g is uniquely extended to the complex tangent bundle CTZ of Z so that: • g@,ti') = s(i9,tf') for all vectors ti,& G CTZZ, z £ Z (the Hermitian condition); • g(•&,•&) > 0 for all non-vanishing complex tangent vectors i? G CTZ; • g('&,'&') = 0 if #, •&' € CTZZ, z £ Z, are simultaneously holomorphic or antiholomorphic tangent vectors. Accordingly, the Kahler form Q is extended to CTZ so that • Q(d, i?') = fi(j?, i?') for all §, •&' G CTZZ, z G Z; • fi(tf,tf') = 0 if •&,•&' e CTZZ, z e Z, are simultaneously holomorphic or antiholomorphic; • dd = 0. Since the Kahler form Q, is closed, it takes the local form Q = iddF,
(2.6.20)
where F is a real function on Z. It is called the potential function for the Kahler structure on Z. Remark 2.6.2. If Z is a non-compact simply connected manifold, local Kahler potentials can be glued into a global one. If Z is not simply connected, a Kahler potential exists at least on an open subset U obtained
182
Geometric and Algebraic Topological Methods in Quantum Mechanics
from Z by deleting a (real) submanifold of lower (real) dimension [145]. Alternatively, one may pass to the universal covering manifold over Z. • Let the Kahler manifold (Z,J,g,Q,) be equipped with complex coordinates (zx), A = l , . . . , m . We will follow the notation (2.6.13) - (2.6.14). Let the complex tangent and cotangent bundles of Z be provided with the bases {d\,d\} and {dzx,dzx}, respectively. Relative to these bases, the Kahler metric g reads 9=
gapdzaVdz/3-
The Hermitian condition of g implies that (2.6.21)
9a0 =9*0 = 90oiSince the Kahler form fi = iga0-dza A dzp is closed, components ga^ of g also obey the equalities dx9a0 = dagX0,
(2-6.22)
9X9a0 = ^9ax-
In particular, the Kahler metric takes the local form (2.6.23)
9cT0 = da^F,
where F is the potential function in the expression (2.6.20). In view of the relations (2.6.21) - (2.6.22), the Levi-Civita connection of a Kahler metric g has the following non-vanishing components:
Its curvature tensor (10.6.67) reads %//3 = ^ V
= -9aVd^9^
+ Sf^dtffntdxgsj,
Then the Ricci tensor can be brought into the form %/j = % / / 3 = -g^dpd^g^ + g£Vg^d0g^g^ -^^(lndet^p)).
=
(2.6.24)
183
Chapter 2 Classical Hamiltonian Systems
E. Cohomology of Kahler manifolds Let (Z, J, g, Cl) be a Kahler manifold of complex dimension m. The Hodge star operator with respect to the Kahler metric g and the corresponding Hodge inner product are extended to complex exterior forms on Z as follows: * : Cp'q{Z) -> Cm-p'm-g(Z),
** =
(-l)p+g,
(
(2.6.25)
where
are the adjoints of the holomorphic and antiholomorphic differentials d and d with respect to the Hodge inner product (2.6.25). The operators d* and d are nilpotent and satisfy the relations
a* o d * + o* o e* = o,
d o a * + a * o d = o,
d * o a + a o a * = o.
Let us construct the complex Laplacian a = dod* + d* od.
(2.6.26)
One can show that, in the case of a Kahler metric g, the Laplacian (2.6.26) commutes with the conjugation 4> —*
and • = ^A,
(2.6.27)
where A is the standard Laplacian. Thus, for any fixed number p £ N, we obtain the elliptic Dolbeault complex Cp>*(Z) = {Cp'"(Z),d,d\D}, O^UP
i ^ c P l 0 ( Z ) -^CP'\Z)
(2.6.28) - ^ • • • C P ' « ( Z ) - ! + . . . , (2.6.29)
where IAV is the complex space of holomorphic p-forms on Z. The corresponding complex of sheaves n
, 7/p
v/"p'°
^rP'1
v
rP'i
.pP'i+1
d
-
184
Geometric and Algebraic Topological Methods in Quantum Mechanics
is proved to be a fine resolution of the sheaf Uvz of holomorphic p-forms on Z [220]. Hence, by virtue of Theorem 1.7.6, the cohomology group Hq(Z;U^) of Z with coefficients in the sheaf Uvz is isomorphic to the g-th cohomology Hp>q(Z) group of the complex (2.6.29). Namely, HP'°(Z) = Up and Hp>q{Z), q > 0, is the complex space of 9-closed (p, g)-forms module the 9-exact ones. Applied to the Dolbeault complex (2.6.28), the Hodge theorem 10.12.3 implies that, with respect to the inner product (2.6.25), the space CPA{Z) °f (Pi
®~5*Cp'q+1(Z) ®B™(Z),
(2.6.30)
where BPi9(Z) C KerD is the space of harmonic (p,q)-forms. It follows that <j> € Bp>q(Z) if and only if d
(2.6.31)
are said to be the Hodge numbers. In particular, hp'q — 0 if p > m or q > m. With the Hodge numbers (2.6.31), the index (10.12.15) of the Dolbeault complex (2.6.28) reads r(Cp'*) = ^ ( - 1 ) 9 ^ . 9 .
It is important that the Betti numbers, the Euler characteristic and the signature of a compact Kahler manifold are expressed into the Hodge numbers (see the formulae (2.6.34), (2.6.35) and Theorem 2.6.11 below). By means of integration Cm'm(Z)5<j>^
I 4>&
one can define the bilinear form
fizfa, 4>') = j
cf>' 6 C m - " ' m - q ( Z ) ,
Chapter 2 Classical Hamiltonian Systems
185
which depends only on cohomology classes of complex smooth forms and leads to the Poincare duality h? —• <j> yields an anti-isomorphism from Bp'q(Z) to Bq'p(Z), and the Hodge numbers obey the relations (2.6.32)
hP,g = hg,P_
Furthermore, the equality (2.6.27) leads to the cohomology isomorphism F r (Z;C)=
Bp'q(Z).
^
(2.6.33)
p+q=r
Therefore, the r-th Betti number br of Z satisfies
br=
Yl
hP q
'-
(2.6.34)
p+q=r
It should be emphasized that the formulae (2.6.32) - (2.6.34) fail for nonKahler manifolds, in general. For instance, the Kahler form Q, is a harmonic (1, l)-form. Its exterior fc-product £lk, k — 2 , . . . , m, is a non-zero harmonic (k, fc)-form. Hence, the spaces Bk
Betti numbers.
2.6.9. A compact Kahler manifold has non-vanishing even •
As was mentioned above, this property is also satisfied for compact symplectic manifolds. The following assertion provides the so called Thurston's example of symplectic manifolds with no Kahler structure [71; 415] (see, e.g., [112] for other examples). PROPOSITION
fold are even.
2.6.10. The odd Betti numbers of a compact Kahler mani•
As an immediate consequence of the formula (2.6.34), one obtains that
X = Y,(-l)p+qhp'q
(2.6.35)
p. q
is the Euler characteristic of a compact Kahler manifold Z. The following result requires the more sophisticated analysis [220].
186
Geometric and Algebraic Topological Methods in Quantum Mechanics
THEOREM
2.6.11. The signature (the index) of a compact Kahler manifold
is
p,q
• In the proof of this Theorem, the following important construction is utilized. Given the Kahler form Cl, let us introduce the operators (2.6.36)
L+ : CP'9{Z) B
(2.6.37) (2.6.38)
L = [L+,L-}.
If Z is compact, the operators L+ and L~ are mutually adjoint relative to the Hodge inner product (2.6.25). One can also show that L (2.6.38) is the scalar operator L(j> = {V + q-n)(j>,
cPeC^iZ),
[154]. It follows at once that the operators (2.6.36) - (2.6.38) obey the commutation relations [L,L+] = 2L+,
[L,L~] = -2L-
and generate the Lie algebra s/(2,C). The relations [L+,d]=0, [L-,d*} = 0,
[L+,d}=0,
[L+,d*]=id,
[L-,T] = 0,
[L-,d] = id\
[L+,d*] = -id, [L~,d\ = -i&¥
also hold. It follows that the operators (2.6.36) - (2.6.38) commute with the Laplacian A. Therefore, they yield the homomorphisms
L+ L~
:Bp'q(Z)-*Bp+i-q+1(Z), :Bp'q{Z)^Bp-l'q-l{Z)
of spaces of harmonic forms and, consequently, of the Dolbeault cohomology groups Hp'*(Z) = B'«(Z). An important class of harmonic (1, l)-forms on a compact Kahler manifold (Z, J, g, Q) is related to the Ricci tensor of different Kahler metrics on
187
Chapter 2 Classical Hamiltonian Systems
Z. Given the Ricci tensor (2.6.24) of a Kahler metric g, one can introduce the harmonic (1, l)-form R = YR\l3dz!3
Ad X
z
= - j-^(lndet(5 M T7)),
(2.6.39)
called the Ricci form. In accordance with well-known theorem [100], the cohomology class of this form depends only on the complex structure J on Z and equals the first Chern class c\(Z) of Z. Moreover, this result has been improved as follows [80; 449]. 2.6.12. Given a Kahler manifold {Z, J, g, 0), let ^ be a harmonic (1, l)-form which represents the first Chern class c\(Z) of (Z, J). Then one can find a J-compatible Kahler metric g' on (Z, J) such that
For instance, any Kahler manifold (Z, J, g, ft) of the vanishing first Chern class (called a Calabi-Yau manifold) admits a Ricci-flat Kahler metric g' whose Kahler form Q.' belongs to the same de Rham cohomology class as that of O [84; 449]. Usually, most interest centers on Calabi-Yau manifolds of three complex dimension (see [11; 28]).
F. Hyper-Kahler manifolds Three almost complex structures (resp. complex structures) Ji, Ji and J3 on a manifold Z form an almost hypercomplex structure (resp. a hypercomplex structure) if J\ = J2J3 = —J%J2This relation implies the relations Ji = eijkJjJk,
(2.6.40)
i,j,k = l , 2 , 3 ,
where e is the totally skew-symmetric Levi-Civita tensor with the component ei23 = 1. Hence, there is the two-sphere of complex structures
{a\Ji + a-zJi + azJz : aj + 02 + ag = 1}. We will use the notation dkcf> : = (-1)1*1 Jkd(Jk
dk = -(d + idk),
k =
1,2,3.
188
Geometric and Algebraic Topological Methods in Quantum Mechanics
Theorem 2.6.3 on the existence of a torsionless connection on a complex manifold is extended to a hypercomplex manifold as follows [19; 196]. 2.6.13. An almost hypercomplex structure (Jk) on a manifold Z is a hypercomplex structure if and only if there exists a torsionless affine connection which preserves all Jk- It is called the Obata connection. • THEOREM
When g is a Riemannian metric on the hypercomplex manifold (Z,Jk) such that it is a Hermitian with respect to each complex structure Jk, the tuple (Z,Jk,g) is called a hyper-Hermitian manifold. Thus, a hyperHermitian manifold Z is a triple of Hermitian structures (Jk,g) on Z such that the complex structures Jk satisfy the relation (2.6.40). A linear connection on a hyper-Hermitian manifold (Z, Jk,g) is called a hyper-Hermitian connection if it is Hermitian with respect to all the complex structures Jk, i.e., it is a metric connection for g and preserves Jk, k = 1,2,3. A hyper-Hermitian connection is said to be hyper-Kdhler if its torsion tensor is totally skew-symmetric. Clearly, a hyper-Kahler connection is a KT-connection with respect to all the Hermitian structures (Jfc, <;). Therefore, it is also called the HKT-connection. THEOREM 2.6.14. [196]. A hyper-Hermitian manifold (Z,Jk,g) admits a hyper-Kahler connection if an only if the fundamental forms fife, associated to Hermitian structures (Jfc,g), fulfils the equivalent conditions: (i) diOi = d.2^2 = d3rt3, (ii) di(V.2 + in3) = 0 , (iii) 9 i ( f i 2 - t f i 3 ) = 0 , (iv) ditlj = —26ijT — eijkdQk, where T is a three-form denned by the skew-symmetric torsion tensor. If a hyper-Kahler connection exists, it is unique. •
If a hyper-Hermitian manifold admits a hyper-Kahler connection, it is said to be an HKT-manifold. It is readily justified from the equivalent conditions in Theorem 2.6.14 that a hyper-Kahler connection is symmetric (i.e., it coincides with the Levi-Civita connection of the metric g and the Obata connection in Theorem 2.6.13) if and only if fife, k = 1,2,3, are Kahler forms. Then (Z, Jk,g) is a hyper-Kdhler manifold. Let (Z,Jk,g,£lk) be a compact hyper-Kahler manifold. For each k = 1,2,3, one can introduce the operators L^ and Lj~ by the formulae (2.6.36) - (2.6.37), but with Qk playing the role of fi. Let us put
189
Chapter 2 Classical Hamiltonian Systems
Then the following relations hold [154; 432]: [L+ ,LJ}= 2eijkKk + SijL, [Ku Kj] = eijkKk,
[L, L+) = 2L+,
[Kit L+] = eijkL+,
[L, L~] = -2Lr,
[Kit Lj] =
eijkL^,
with all the other brackets equal zero. These are the commutation relations of the Lie algebra so(4,1). Indeed, let Ipq, p, q = 1 , . . . , 5, be the standard basis for the Lie algebra so(5). One can reproduce the above commutation relations by letting
L'l = -(/i5 + Ha),
L~ - - ( i j 5 - ila),
L = 2iJ45,
Kk — tujhj.
We refer the reader to [383] for cohomology of hyper-Kahler manifolds.
2.7
Appendix. Poisson manifolds and groupoids
In this Section, we describe the category Poisson of integrable Poisson manifolds [85; 268; 426; 444], which will be quantized in Section 8.6. We start with the notion of a complete Poisson morphism. Given Poisson manifolds (Pi, W\) and (P21W2), let g : P\ —> Pi be a Poisson morphism. Then, given a function / 6 C°°{P-i) and its Hamiltonian vector field $f (2.1.13), we have the Hamiltonian vector field -de*f on P\ whose trajectories project onto those of flf. A Poisson morphism g is called complete if the Hamiltonian vector field i9 e -/ is complete whenever the Hamiltonian vector field i?/ is complete. The injection of every symplectic leaf in a Poisson Example 2.7.1. manifold is a complete Poisson morphism. More generally, the range of a complete Poisson morphism is a union of symplectic leaves. D Let (P, w) be a Poisson manifold. The symbol P~ further stands for the Poisson manifold (P,—w). Given a Poisson manifold P, a left (resp. right) symplectic P-module is defined as a symplectic manifold S together with a complete Poisson morphism g : S —» P (resp. S —> P~). Just as in algebra (see Section 10.4), one can define the tensor product of right and left symplectic P-modules g : 5 —> P and g' : S' —> P~. Namely, let us consider the fibre product SSx S1 = {(x,y)eSxS'
: g(x) = g'(y)}
(2.7.1)
190
Geometric and Algebraic Topological Methods in Quantum Mechanics
(see the notation (10.4.6)), which is the inverse image of the diagonal of P x P~ under the Poisson morphism {g,g') -.SxS'^PxP-. If either g or g' is a surjective submersion, the fibre product (2.7.1) is a coisotropic smooth submanifold of S x S' and, accordingly, a presymplectic manifold (see Example 2.1.5). Then one can define the tensor product S*S' of right and left symplectic P-modules as the quotient of the presymplectic manifold (2.7.1) by its characteristic foliation. In accordance with Proposition 2.1.5, this tensor product is a symplectic manifold if the characteristic foliation of the presymplectic manifold (2.7.1) is simple. Given Poisson manifolds Pi and P 2 , a symplectic (Pi, P2)-bimodule is denned as a symplectic manifold S and a pair of maps (2.7.2)
Pl^S^P2
which make S both into a left Pi-module and a right P2-module, and satisfy the commutation condition {elh, £2/2} = 0,
/1 e C^iPi),
h e C°°(P2),
where {.,.} is the Poisson bracket on S. Two symplectic (Pi, P2)-bimodules Pl^S^P2,
P1^S'^P2
are said to be isomorphic if there exists a symplectomorphism (j> : S —> 5' such that gi = g\ o (f>. Remark 2.7.2. Let us note that, in [27l], symplectic bimodules, without • assumption of completeness, are called dual pairs. Given symplectic bimodules Pl
^
S
^P
2
,
P2 ^ S ' - ^ P 3 ,
(2.7.3)
one can construct their tensor product S * S' over P 2 . As was mentioned above, it however need not be a symplectic (Pi — P3)-bimodule, unless some extra "regularity" conditions hold. In order to formulate these conditions, let us relate Poisson manifolds with Lie groupoids. A Lie groupoid (25 is called a symplectic groupoid if it is provided with a symplectic formfisuch that the multiplication graph M = {(x,y,xy)
: (x,y) G 0 2 } C <25 x (5 x <2T
(2.7.4)
Chapter 2 Classical Hamiltonian Systems
191
is a Lagrangian submanifold of the product 0 x 0 x <5~ provided with the symplectic form fi © 0, 9 fi. Example 2.7.3. The cotangent bundle T*X of a smooth manifold X endowed with the canonical symplectic form (2.1.3) exemplifies a symplectic groupoid with respect to the fibrewise addition. More generally [426], the cotangent bundle T*(0) of any Lie groupoid © is a symplectic groupoid over the dual V*<8 of the Lie algebroid V<& —> 0° associated to 0 (see Section 10.3). D A symplectic groupoid (0,fi) possesses the following properties. • Its unit space 0° is a Lagrangian submanifold of 0. • The inversion map x —> x~l, x & <&, is the anti-symplectic involution • The fibres 0 U and <5U, u 6 0°, are symplectic orthogonal of one another. • At each point x € 0, we have Ker(Tlx) = {0r.f(x) Ker {Trx) = {^f{x)
: / G C°°(0 0 )} : / e C°°((S0)}.
• The unit space 0° admits a unique Poisson structure such that r : (5 -> (5°,
I : 0 -> <5°~
are Poisson maps. In view of the last property, a Poisson manifold P is called integrable if there exists a symplectic groupoid (0, fi) (called the integration of P) over
0° = P.
THEOREM
2.7.1. If 0 is a symplectic groupoid over a Poisson manifold P,
then r : 0 -> 0° = P is a complete submersion. Conversely, if there exists a symplectic P-module S —> P which is a submersion, then P is an integrable Poisson manifold. Moreover, there exists a symplectic groupoid, integrating P, whose i-fibres are connected and simply-connected, and this groupoid, denoted 0(P), is unique up to isomorphisms. D
192
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let © be a symplectic groupoid acting on a symplectic manifold S on the left with a moment g : S —> ©° (see Section 10.4). Its action © x S-+ S 00
is called symplectic if the graph of this action is a Lagrangian submanifold of <8xSxS~. Let P be an integrable Poisson manifold and © a symplectic groupoid over P. Given a left symplectic action of © on a symplectic manifold S, the moment g : S —» P is a complete Poisson morphism defining a symplectic module over P. Conversely, a symplectic P-module g : S —* P implies a left symplectic action of the ^-connected and Z-simply connected symplectic groupoid ©(P) in Theorem 2.7.1 on S. Consequently, there is one-to-one correspondence between P-modules and left symplectic actions of ©(P). Accordingly, given integrable Poisson manifolds Pi and P2 and a symplectic bimodule (2.7.2), we have a left symplectic action of the symplectic groupoid 6(Pi) and a right symplectic action of the symplectic groupoid ©(P2) on S. Moreover, these actions mutually commute and, consequently, 5 is a symplectic (<5(Pi) — 0(P2))-bibundle (see Section 10.4). A symplectic bimodule (2.7.2) is called regular if the associated symplectic (©(Pi) - 0(P2))-bibundle is left principal. 2.7.2. A symplectic bimodule (2.7.2) is regular if and only if £1 is a submersion, p2 is a. surjective submersion with connected and simplyconnected fibres, and the gi- and £>2-fibres are symplectic orthogonal of one another. •
PROPOSITION
The following theorem states that the tensor product of regular symplectic bimodules is well denned [268]. 2.7.3. The tensor product S*S' of regular symplectic bimodules (2.7.3) is a regular symplectic (Pi — P3)-bimodule which coincides with the tensor product S*S' (10.4.9) of S and S' as principal symplectic (©(Pi) ©(P2))- and (6(P 2 ) - ©(P3))-bibundles. • THEOREM
In particular, it follows that the tensor product of regular symplectic bimodules is associative up to isomorphisms, while the regular symplectic bimodule P^<5(P) -» P
Chapter 2 Classical Hamiltonian Systems
193
associated to the (0(P)-(9(P))-bibundle S = <5(P) plays the role of a twosided unit. As a consequence, the following category of Poisson manifolds is denned. 2.7.4. The category Poisson has integrable Poisson manifolds as objects and classes of isomorphic regular symplectic bimodules as morphisms. •
DEFINITION
Accordingly, the notion of the Morita equivalence of groupoids leads to the following definition of the Morita equivalence of integrable symplectic manifolds in the sense of [444]. Namely, integrable symplectic manifolds Pi and Pi are Morita equivalent if and only if they are isomorphic object in the category Poisson, i.e., their integrations ©(Pi) and <3(P2) in Theorem 2.7.1 are Morita equivalent symplectic groupoids.
Chapter 3
Algebraic quantization
Algebraic quantum theory follows the hypothesis that a quantum system can be characterized by Hermitian elements of a C*-algebra A and positive forms f on A treated as mean values of quantum observables (see Remark 3.1.1 below). In accordance with the Gelfand-Naimark-Segal (henceforth GNS) construction, any positive form on a C*-algebra A determines its cyclic representation by bounded operators in a Hilbert space. This Chapter addresses some modifications of this GNS construction. 3.1
GNS construction I. C*-algebras of quantum systems
Subsections: A. Involutive algebras, 195; B. Hilbert spaces, 198; C. Countably Hilbert spaces and nuclear spaces, 201; D. Operators in Hilbert spaces, 202; E. Representations of involutive algebras, 204; F. The GNS representation, 206. We start with a brief exposition of the conventional GNS representation of C*-algebras [129] A. Involutive algebras A complex algebra A is called involutive, if it is provided with an involution * such that {a*y=a,
{a + Xby =a*+Jb*,
(ab)*=b*a*,
a,b e A,
XeC.
Let us recall the standard terminology. An element a £ A is normal if aa* = a*a, and it is Hermitian or self-adjoint (see Section 3.6) if a* = a. If 195
196
Geometric and Algebraic Topological Methods in Quantum Mechanics
A is a unital algebra, a normal element such that aa* = a*a = 1
is called unitary. Remark 3.1.1. A product of two Hermitian elements need not be Hermitian, unless they mutually commute. As a consequence, Hermitian elements fail to make up an associative algebra. Therefore, one considers Jordan algebras in order to describe quantum observables [3; 144]. Hermitian elements of an involutive algebra constitute a Jordan algebra with respect to the symmetrized product a- a' = -{aa' + a'a).
• An involutive algebra A is called a normed algebra (resp. a Banach algebra) if it is a normed (resp. complete normed) vector space whose norm ||.|| obeys the multiplicative conditions ||ab|| < ||a||||6||,
K | | = ||a||,
a,beA.
An involutive Banach algebra A is called a C* -algebra if \\a\\2 = \\a*a\\
for all a € A. If A is a unital C"*-algebra, then ||1|| = 1. A C*-algebra is provided with the normed topology, i.e., it is a topological involutive algebra. Let X be a closed two-sided ideal of a C*-algebra A. Then X is selfadjoint, i.e., X* = X. Endowed with the quotient norm, the quotient A/X is a C*-algebra. Remark 3.1.2. It should be emphasized that by a morphism of normed algebras is customarily meant a morphism of the underlying involutive algebras, without any condition on the norms and continuity. At the same time, an isomorphism of normed algebras means always an isometric morphism. Any morphism
aeA.
(3.1.1)
•
197
Chapter 3 Algebraic Quantization
Any involutive algebra A can be extended to a unital algebra A = C®A by the adjunction of the identity 1 to A (see Remark 1.1.2). The unital extension of A is also an involutive algebra with respect to the operation (Al + a)* = (Al + a*),
AeC,
a&A.
If A is a normed involutive algebra, a norm on A is extended to A, but not uniquely. If A is a C*-algebra, a norm on A is uniquely prolonged to the norm ||Al + a|| := sup ||Aa'+aa'|| l|a'll
a £ A,
a' G A.
For instance, if A and A' are operator algebras in Hilbert spaces E and E', this norm is exactly the operator norm (3.1.14) of operators in the tensor product E <8> E' of Hilbert spaces E and E'. In general, there are several ways of completing the algebraic tensor product of C*-algebras in order to • obtain a C*-algebra [382].
198
Geometric and Algebraic Topological Methods in Quantum Mechanics
B. Hilbert spaces An important example of C*-algebras is the algebra B(E) of bounded (and, equivalently, continuous) operators in a Hilbert space E. Every closed involutive subalgebra of B(E) is a C*-algebra and, conversely, every C*algebra is isomorphic to a C*-algebra of this type (see Theorem 3.1.2 below). Let us recall the basic facts on pre-Hilbert and Hilbert spaces [56]. A Hermitian form on a complex vector space E is defined as a sesquilinear form (.|.) such that (e\e') =J^\e),
(Ae|e'> = <e|Ae') = X{e\e'),
e,e' G E,
A e C.
Remark 3.1.5. There exists another convention where (e|Ae') = A(e|e') (see Section 8.1). • A Hermitian form (.|.) is said to be positive if (e|e) > 0 for all e € E. Throughout the book, all Hermitian forms are assumed to be positive. A Hermitian form is called non-degenerate if the equality (e|e) = 0 implies e = 0. A complex vector space endowed with a (positive) Hermitian form is called a pre-Hilbert space. Morphisms of pre-Hilbert spaces, by definition, are isometric. A Hermitian form provides E with the topology determined by the seminorm ||e|| = {e\e)ll\
(3.1.2)
Hence, a pre-Hilbert space is Hausdorff if and only if the Hermitian form (.|.) is non-degenerate, i.e., the seminorm (3.1.2) is a norm. In this case, the Hermitian form (.|.) is called a scalar product. A family {e^}/ of elements of a pre-Hilbert space E is called orthonormal if its members are mutually orthogonal and ||ej|| = 1 for all i £ I. Given an element e S E, there exists at most a countable set of elements ej of an orthonormal family such that (e|ej) ^ 0 and
E^}
2
< Ikll2.
A family {e^}/ is called total if it spans a dense subset of E or, equivalently, if the condition (e|ej) = 0 for all i £ I implies e = 0. A total orthonormal family in a Hausdorff pre-Hilbert space is called a basis for E. Given a basis
Chapter 3 Algebraic Quantization
199
{ei}i, any element e £ E admits the decomposition
e=5>| e i > e i , i€l
||e|| 2 =£|<e| ei >| 2 . iel
A basis for a pre-Hilbert space need not exist. 3.1.1. Every Hausdorff pre-Hilbert space, satisfying the first axiom of countability (e.g., if it is second-countable), has a countable orthonormal basis. D PROPOSITION
Remark 3.1.6. The notion of a basis for a pre-Hilbert space differs from that of an algebraic basis for a vector space. • A Hilbert space is denned as a complete Hausdorff pre-Hilbert space. Any Hausdorff pre-Hilbert space can be completed to a Hilbert space so that its basis, if any, is also a basis for its completion. Every Hilbert space has a basis, and any orthonormal family in a Hilbert space can be extended to its basis. All bases for a Hilbert space have the same cardinal number, called the Hilbert dimension. Moreover, given two bases for a Hilbert space, there is its isomorphism sending these bases to each other. A Hilbert space has a countable basis if and only if it is separable. Then it is called a separable Hilbert space. A separable Hilbert space is second-countable. Remark 3.1.7. Unless otherwise stated, by a Hilbert space is meant a complex Hilbert space. A complex Hilbert space (E, {.].)) seen as a real vector space ER is provided with a real scalar product (e,e') := \{(e\e') + (e'|e» = Re <e|e'),
(3.1.3)
which makes EK into a real Hilbert space. It is also a real Banach space. Conversely, the complexification E = C® V of a real Hilbert space (V, (.,.)) is a complex Hilbert space with respect to the Hermitian form (e^ie^+ie'z)
:= (e1;e[) + i{(e2,e[) - {eue'2)) + (e2,e'2).
(3.1.4) D
The following are the standard constructions of new Hilbert spaces from old ones.
200
Geometric and Algebraic Topological Methods in Quantum Mechanics
• Let (EL, (.|.)BO be a set of Hilbert spaces and Yl EL denote the direct sum of vector spaces EL. For any two elements e = (el) and e' = (e") of ^ E ' , the sum
(e|e')e := £ < e V V
C3-1-5)
is finite, and defines a non-degenerate Hermitian form on ^2,El. The completion ®EL of X3 El with respect to this form is a Hilbert space, called the Hilbert sum of {EL}. This is a subspace of the Cartesian product Yl EL which consists of the elements e = (eL) such that
^ H e 1 s '
K M ® = 5>ile2>is Wf>*. i,p
ioi = ^ e i ® f t i ,
u>2 = ^ e f ®/if,
e[,e$ e E,
h\,h% G H.
Let {ej} and {/ij} be bases for E and H, respectively. Then {e^ ® hj} is a basis for E ® H. • Let £" be the topological dual of a Hilbert space E. Then the assignment e w e ( e ' ) := (e'|e),
e,e'e E,
(3.1.6)
defines an antilinear bijection of I? onto JB', i.e., Ae = Ae. The dual E' of a Hilbert space is a Hilbert space provided with the scalar product (e\e')' := (e'\e)
(3.1.7)
such that the morphism (3.1.6) is isometric. The E' is called the dual Hilbert space, and is usually denoted by E. A Hilbert space E and its dual E' seen as real Hilbert and Banach spaces are isomorphic to each other.
201
Chapter 3 Algebraic Quantization
C. Countably Hilbert spaces and nuclear spaces Physical applications of Hilbert spaces are limited by the fact that the dual of a Hilbert space E is anti-isomorphic to E. The construction of a rigged Hilbert space describes the dual pairs (E,E') where E' is larger than E [164]. For instance, let us mention the space of smooth functions of rapid decrease and its dual space of tempered distributions (generalized functions) [4l]. Let a complex vector space E have a countable set of non-degenerate Hermitian forms (.|.)fe, k G N + , such that <e|e)i<-"<(e|e)fc<"for all e £ E. The family of norms ||.|U = <.|.)i/2,
(3.1.8)
k€N+,
yields a Hausdorff topology on E (see Section 10.5A). The space E is called a countably Hilbert space if it is complete with respect to this topology [164]. For instance, every Hilbert space is a countably Hilbert space where all Hermitian forms (.|.)fc coincide. Let Ek denote the completion of £ with respect to the norm ||.||jt (3.1.8). There is the chain of injections (3.1.9)
E1DE2D---EkD--together with the homeomorphism E = nEk-
The dual spaces form the
increasing chain E[ cE'2C---cE'kC---
,
(3.1.10)
and E' = UE'h. k
fe
The dual E' of E can be provided with the weak* and strong topologies. One can show that a countably Hilbert space is reflexive. Given a countably Hilbert space E and m
EnDEBet-^eeEcEm to the continuous map of En onto a dense subset of Em. A countably Hilbert space E is called a nuclear space if, for any m, there exists n such
202
Geometric and Algebraic Topological Methods in Quantum Mechanics
that T™ is a nuclear map, i.e., i l
where: (i) {e n} and {e^} are bases for the Hilbert spaces En and Em, respectively, (ii) A, > 0, (iii) the series Yl^i converges [164]. An important property of nuclear spaces is that they are perfect, i.e., every bounded closed set in a nuclear space is compact. It follows immediately that a Banach (and Hilbert) space is not nuclear, unless it is finite-dimensional. Since a nuclear space is perfect, it is separable, and the weak* and strong topologies (and, consequently, all topologies of uniform convergence) on a nuclear space E and its dual E' coincide. Let E be a nuclear space, provided with still another non-degenerate Hermitian form (.|.) which is separately continuous, i.e., continuous with respect to each argument. It follows that there exist numbers M and m such that (e|e) < M||e|| m ,
e G E.
(3.1.11)
Let E denote the completion of E with respect to this form. There are the injections (3.1.12)
EcEcE',
where E is a dense subset of E and E is a dense subset of E', equipped with the weak* topology. The triple (3.1.12) is called the rigged Hilbert space. Furthermore, bearing in mind the chain of Hilbert spaces (3.1.9) and that of their duals (3.1.10), one can convert the triple (3.1.12) into the chain of spaces (3.1.13)
Ec---cEkC---E1dEcE'1c---(ZE'k(Z---cE'.
Real countably Hilbert spaces, nuclear spaces and rigged Hilbert spaces are similarly described.
D. Operators in Hilbert spaces Unless otherwise stated (see Section 3.6), we deal with bounded operators a £ B(E) in a Hilbert space E. They are provided with the operator norm \\a\\ :=
sup ||ae|| B ,
l|e||B = l
a € B(E).
(3.1.14)
Chapter 3 Algebraic Quantization
203
This norm makes the involutive algebra B(E) of bounded operators in a Hilbert space E into a C*-algebra. The corresponding topology on B{E) is called the normed operator topology. One also provides B(E) with the strong and weak operator topologies, determined by the families of seminorms {pe(a) = \\ae\\, e £ E}, {pe,e.(a) = |(oe|e')|,
e.e'efi),
respectively. The normed operator topology is finer than the strong one which, in turn, is finer than the weak operator topology. The strong and weak operator topologies on the subgroup U(E) C B(E) of unitary operators coincide with each other. Remark 3.1.8. It should be emphasized that B(E) fails to be a topological algebra with respect to strong and weak operator topologies. Nevertheless, the involution in B(E) is also continuous with respect to the weak operator topology, while the operations B{E) 9 a i-> aa' G B(E), B(E) B o H a ' a e B{E), where a' is a fixed element of B(E), are continuous with respect to all the above mentioned operator topologies. D Remark 3.1.9. Let N be a subset of B(E). The commutant N' of N is the set of elements of B(E) which commute with all elements of N. It is a subalgebra of B{E). Let N" = (N1)' denote the bicommutant. Clearly, N C N". An involutive subalgebra B of B(E) is called a von Neumann algebra if B = B". This property holds if and only if B is strongly (or, equivalently, weakly) closed in B(E) [129]. For instance, B(E) is a von Neumann algebra. Since a strongly (weakly) closed subalgebra of B{E) is also closed with respect to the normed operator topology on B(E), any von Neumann algebra is a C*-algebra. D Remark 3.1.10. An operator in a Hilbert space E is called completely continuous if it is compact, i.e., it sends any bounded set into a set whose closure is compact. An operator a € B(E) is completely continuous if and
204
Geometric and Algebraic Topological Methods in Quantum Mechanics
only if it can be represented by the series oo
a{e)=J2Xk(e\ek)ek,
(3.1.15)
k=i
where e^ are elements of a basis for E and Afe are positive numbers which tend to zero as k —> oo. For instance, every degenerate operator (i.e., an operator of finite rank which sends E onto its finite-dimensional subspace) is completely continuous. Moreover, the set T(E) of completely continuous operators in E is the completion of the set of degenerate operators with respect to the operator norm (3.1.14). Every completely continuous operator can be written as a = UT, where U is a unitary operator and T is a positive completely continuous operator, i.e., (Te\e) > 0 for all e £ E. A completely continuous operator a is called the Hilbert-Schmidt operator if the series
IMIHS : =
z2^l' k
called the Hilbert-Schmidt norm, converges. Hilbert-Schmidt operators make up an involutive Banach algebra with respect to this norm, and it is a two-sided ideal of the algebra B(E). A completely continuous operator a in a Hilbert space E is called a nuclear operator if the series
IMITY
:= 5Z A f e ' k
called the trace norm, converges. Nuclear operators make up an involutive Banach algebra with respect to this norm, and it is a two-sided ideal of the algebra B(E). Any nuclear operator is a Hilbert-Schmidt one. Moreover, the product of arbitrary two Hilbert-Schmidt operators is a nuclear operator, and every nuclear operator is of this type. D
E. Representations of involutive algebras In this Section, we restrict our consideration to representations of involutive algebras by bounded operators in Hilbert spaces [129; 351]. It is a morphism IT of an involutive algebra A to the algebra B(E) of bounded operators in a Hilbert space E, called the carrier space of n. Representations throughout are assumed to be non-degenerate, i.e., there is no element e / 0
Chapter 3 Algebraic Quantization
205
of E such that Ae = 0 or, equivalently, AE is dense in E. A representation 7r of an involutive algebra A is uniquely prolonged to a representation n of the unital extension A of A. THEOREM
tation.
3.1.2. If A is a C*-algebra, there exists its isomorphic represenD
Two representations TTI and TT2 of an involutive algebra A in Hilbert spaces Ei and E E% such that ^2(0-)
= 7O7Ti(a) 0 7 " ' ,
a e A
Let {TT1} be a family of representations of an involutive algebra A in Hilbert spaces El. If the set of numbers ||vri'C«.) II is bounded for each a € A, one can construct the continuous linear operator n(a) in the Hilbert sum (BE1 which induces 7r'(a) in each EL. For instance, this is the case of a C*-algebra A due to the property (3.1.1). Then n is a representation of A in ®El, called the Hilbert sum of representations nx. Given a representation •n of an involutive algebra A in a Hilbert space E, an element 0 € E is said to be a cyclic vector for n if the closure of ir(A)9 is equal to E. Accordingly, TT is called a cyclic representation. 3.1.3. Every representation of an involutive algebra A is a Hilbert sum of cyclic representations. •
THEOREM
A representation TV of an involutive algebra A in a Hilbert space E is called topologically irreducible if the following equivalent conditions hold: • the only closed subspaces of E invariant under n(A) are 0 and E; • the commutant of tr(A) in B{E) is the set of scalar operators; • every non-zero element of E is a cyclic vector for TT. Let us recall that irreducibility of n in the algebraic sense means that the only subspaces of E invariant under ir(A) are 0 and E. If A is a C*-algebra, the notions of topologically and algebraically irreducible representations are equivalent. Therefore, we will further speak on irreducible representations of a C*-algebra without the above mentioned qualification. An algebraically irreducible representation n of an involutive algebra A is characterized by its kernel Ker -K C A. This is a two-sided ideal, called primitive. The assignment ASJTM
Ker?r € Prim(A)
(3.1.16)
206
Geometric and Algebraic Topological Methods in Quantum Mechanics
defines the canonical surjection of the set A of the equivalence classes of algebraically irreducible representations of an involutive algebra A onto the set Prim(>l) of primitive ideals of A. It follows that algebraically irreducible representations with different kernels are necessarily inequivalent. The set Prim(^4) is equipped with the so called Jacobson topology (see [129] for detail). This topology is not Hausdorff, but obeys the Frechet axiom, i.e., for any two distinct points of Prim(>l), there is a neighborhood of one of the points which does not contain the other. Then the set A is endowed with the coarsest topology such that the surjection (3.1.16) is continuous. It is called the spectrum of an involutive algebra A. PROPOSITION 3.1.4. If the spectrum A satisfies the Frechet axiom (e.g., A is Hausdorff), the map A —> Pnm(A) is a homeomorphism, i.e., algebraically irreducible representations with the same kernel are equivalent. D
3.1.5. If an involutive algebra A is unital, Prim(A) and A are quasi-compact , i.e., they satisfy the Borel-Lebesgue axiom, but need not be Hausdorff. O PROPOSITION
PROPOSITION 3.1.6. The spectrum A of a C"*-algebra A is a locally quasicompact space. •
If A is a C*-algebra, the topology of its spectrum A can be described as follows. Let {at} be a dense set of elements of A and Zt the set of n € A such that ||7r(at)|| > 1. Then ZL constitute a base for the topology on A. Example 3.1.11. A C*-algebra is said to be elementary if it is isomorphic to the algebra T(E) C B(E) of compact operators in some Hilbert space E. Every non-trivial irreducible representation of an elementary C* algebra A = T(E) is equivalent to its isomorphic representation by compact operators in E [129]. Hence, the spectrum of an elementary algebra is a singleton set. • F. The GNS representation Let / be a complex form on an involutive algebra A. It is called positive if f(a*a) > 0 for all a £ A. Given a positive form / , the Hermitian form (a\b):=f{b*a),
a,b G A,
(3.1.17)
Chapter 3 Algebraic Quantization
207
makes A into a pre-Hilbert space. In particular, the relation \f(b*a)\2
a,b G A,
(3.1.18)
holds. If A is a normed involutive algebra, positive continuous forms on A are provided with the norm
Il/H = sup |/(a)|, ||a|| = l
aeA.
One says that / is a state of A if ||/|| = 1. Positive forms on a C*-algebra are continuous. Conversely, a continuous form / on an unital C*-algebra is positive if and only if / ( I ) = ||/||. In particular, it is a state if and only if /(1) = 1. For instance, let A be an involutive algebra, n its representation in a Hilbert space E, and 0 an element of E. Then the map u8 • a -> (7r(a)0|0>
(3.1.19)
is a positive form on A. It is called the vector form determined by n and 9. This vector form is a state if the vector 0 is normalized. Let w$1 and uj02 be two vector forms on A determined by representations TTI in E\ and TT2 in £'2. If u>0j = ui02, there exists a unique isomorphism of E\ to Ei which sends TK\ to 1x2 and 9\ G E\ to 02 € £2The following theorem states that, conversely, any positive form on a C*-algebra equals a vector form determined by some representation of A called the GNS representation [129]. THEOREM 3.1.7. Let / be a positive form on a C*-algebra A. It is extended to a unique positive form J o n the unital extension A of A such that / ( I ) = ||/||. Let Nf be a left ideal of A consisting of those elements aeA such that f{a*a) = 0. The quotient A/Nf is a Hausdorff pre-Hilbert space with respect to the Hermitian form obtained from f(b*a) (3.1.17) by passage to the quotient. We abbreviate with Ef the completion of A/Nf and with 0/ the canonical image of 1 G A in A/Nf C Ef. For each aeA, let r(a) be the operator in A/Nf obtained from the left multiplication by a in A by passage to the quotient. Then the following holds. (i) Each r(a) has a unique extension to an operator nf(a) in the Hilbert space Ef. (ii) The map a H-+ 7T/(a) is a representation of A in Ef. (iii) The representation iTf admits a cyclic vector Of. (iv) f(a) = (Tr(a)6f\6f) for each aeA. •
208
Geometric and Algebraic Topological Methods in Quantum Mechanics
The representation 7iy and the cyclic vector Of in Theorem 3.1.7 are said to be determined by the form / , and the form / equals the vector form determined by 717 and Of. Conversely, given a representation n of A in a Hilbert space E and a cyclic vector 0 for n, let w be the vector form on A determined by n and 0. Let 7rw and 0w be the representation in Eu and the vector of Eu determined by w in accordance with Theorem 3.1.7. Then there is a unique isomorphism of E to Eu which sends n to n^ and 0 to 0^. Example 3.1.12. In particular, any cyclic representation of a C*-algebra A is a summand of the universal representation ®fftf of A, where / runs through all positive forms on A. • It may happen that different positive forms on a C*-algebra determine the same representation as follows. 3.1.8. (i) Let A be a C*-algebra and / a positive form on A which determines a representation 717 of A and its cyclic vector Of. Then for any b e A, the positive form a —> f(b*ab) on A determines the same representation 717. (ii) Conversely, any vector form / ' on A determined by the representation 717 is the limit a —> F(b*abi), where {6j} is a convergent sequence with respect to the normed topology on A. • PROPOSITION
Now let us specify positive forms on a C*-algebra A which determine its irreducible representations. A positive form / ' on an involutive algebra A is said to be dominated by a positive form / if / — / ' is a positive form. A non-zero positive form / on an involutive algebra A is called pure if every positive form / ' on A which is dominated by / reads A/, 0 < A < 1. THEOREM 3.1.9. The representation of 717 of a C*-algebra A determined by a positive form / on A is irreducible if and only if / is a pure form [129]
•
In particular, any vector form determined by a vector of a carrier space of an irreducible representation is a pure form. Therefore, it may happen that different pure forms determine the same irreducible representation. THEOREM 3.1.10. (i) Pure states f\ and / 2 of a C*-algebra A yield equivalent representations of A if and only if there exists an unitary element U
Chapter 3 Algebraic Quantization
209
of the unital extension A of A such that f2(a) = fi(U*aU),
aeA.
(ii) Conversely, let 7r be an irreducible representation of a C*-algebra A in a Hilbert space E. Given two different elements 6\ and #2 of E (they are cyclic for 7r), the vector forms on A determined by (TT, #i) and (ft, 82) are equal if and only if there exists A G C, |A| = 1, such that #i = A02(iii) There is one-to-one correspondence between the pure states of a C*-algebra A associated to the same irreducible representation n of A in a Hilbert space E and the one-dimensional complex subspaces of E. It follows that these states constitute the projective Hilbert space PE in Example 4.1.5 below. • Let P(A) denote the set of pure states of a C*-algebra A. Theorem 3.1.10 implies a surjection P(A) —> A, where A is the spectrum of A. This surjection is a bijection if and only if any irreducible representation of A is one-dimensional, i.e., A is a commutative C*-algebra. In this case, A is the C*-algebra of continuous complex functions vanishing at infinity on A, while a pure state on A is a Dirac measure ex, x € A, on A, i.e., ex(a) — a(x) for all aEA. Being a subset of the topological dual A' of the Banach space A, the set P(A) is provided with the normed topology. However, one usually refers to P(A) equipped with the weak* topology. In this case, the canonical surjection P(A) —> A is continuous and open [129]. 3.2
GNS construction II. Locally compact groups
Given a locally compact group G provided with a Haar measure, the space LQ(G) of the equivalence classes of complex integrable functions on G is an involutive Banach algebra with an approximate identity. There is oneto-one correspondence between the representations of this algebra and the strongly continuous unitary representations of a group G. Thus, one can employ the GNS construction in order to describe these representations of G [129]. Let a left Haar measure dg on G hold fixed, and by an integrability condition throughout is meant the dg-integrability. A uniformly (resp. strongly) continuous unitary representation of a locally compact group G in a Hilbert space E is a continuous homeomorphism
210
Geometric and Algebraic Topological Methods in Quantum Mechanics
IT of G to the group U(E) C B(E), called the unitary group, of unitary operators in E provided with the normed (resp. strong) operator topology. A uniformly continuous representation is strongly continuous. However, the uniform continuity of a representation is rather rigorous condition. For instance, a uniformly continuous irreducible unitary representation of a connected locally compact real Lie group is necessarily finite-dimensional. Therefore, one usually studies strongly continuous representations of locally compact groups. In this case, any element £ of a carrier Hilbert space E yields the continuous map G B g .-> TT(<7)£ € E.
Since strong and weak operator topologies on the unitary group U(E) coincide, we have a bounded continuous complex function
(3-2.1)
on G for any fixed elements £,77 G E. It is called a coefficient of the representation IT. There is the obvious equality
= VvAd'1)-
The Banach space LQ(G) of the equivalence classes of integrable complex functions on G is provided with the structure of an involutive Banach algebra with respect to the contraction /1 * fy (10.5.20) and the involution
/(ff)-/•(«/) = A( S - 1 )7(F T ), where A is the modular function of G. It is called the group algebra of G. The map / >-* f(9)dg defines an isometric monomorphism of LQ(G) to the Banach algebra M1(G, C) of bounded complex measures on G provided with the involution [j,* = fj,-1. Unless otherwise stated, L^(G) will be identified with its image in M1(G,C). In particular, the group algebra L^(G) admits an approximate identity which converges to the Dirac measure £1 G M1(G, C). The group algebra L^(G) is not a C*-algebra. Its Remark 3.2.1. enveloping C*-algebra C*(G) is called the C* -algebra of a locally compact group G. Let us recall the notion of an enveloping C*-algebra [129]. Let A
Chapter 3 Algebraic Quantization
211
be an involutive Banach algebra A with an approximate identity, and let P(A) be the set of pure states of A. For each a e A, we put \\a\\'= sup f(aa*)1/2, feP(A)
a € A.
(3.2.2)
It is a seminorm on A such that ||a||' < ||a||. If A is a C*-algebra, \\a\\' = \\a\\ due to the relation (3.1.1) and the existence of an isomorphic representation of A. Let I denote the kernel of ||.||'. It consists of o € A such that ||o||' = 0. Then the completion A* of the factor algebra A/1 with respect to the quotient of the seminorm (3.2.2), is a C*-algebra, called the enveloping C*-algebra of A. There is the canonical morphism r : A —> A*. Clearly, A = A* if A is a C*-algebra. The enveloping C*-algebra A* possesses the following important properties. (i) If 7T is a representation of A, there is exactly one representation TT* of A* such that vr = n^ o r. Moreover, the map 7r —> TT* is a bijection of the set of representations of A onto the set of representations of A^. (ii) A representation vr of A is non-degenerate (resp. topologically irreducible) if and only if TT* is non-degenerate (resp. irreducible). (iii) If / is a continuous positive form on A, there exists exactly one positive form p on A* such that / = / * O T . Moreover, ||/t|| = ||/||. The map / —> p is a bijection of the set of continuous positive forms on A onto the set of positive forms on A+. • Unitary representations of a locally compact group G and representations of the group algebra L^(G) are related as follows [129]. Let 7T be a (strongly continuous) unitary representation of G in a Hilbert space E. Given a bounded positive measure /i on G, let us consider the integrals
Vt.vit1) = /
(""(^MM
of the coefficient functions
(3.2.3)
212
Geometric and Algebraic Topological Methods in Quantum Mechanics
The assignment /z —> 7r(/x) provides a representation of the involutive Banach algebra Ml(G, C) in E. Its restriction * L ( / ) = J*(g)f(g)dg
€ B(£?)
(3.2.4)
to Lp(G) is non-degenerate. One says that the representations (3.2.3) of M1(G, C) and (3.2.4) of L^(G) are determined by the unitary representation 7T of G. Conversely, let nL be a representation of the involutive Banach algebra Z>c(G) in a Hilbert space E. There is a monomorphism g i-> eff of the group G onto the subgroup of Dirac measures eg, g € G, of the algebra M1(G, C). Let {ut()}(,e/ be an approximate identity in L^(G). Then {irL(u,,)} converges to an element of B(E) which can be seen as a representation 7rL(ei) of the unit element £i of M 1 (G,C). Accordingly, {irL'{~f{g)uL)} converges to 7rL(es). Thereby, we obtain the (strongly continuous) unitary representation Tc(g) — 7rL(eg) of the group G in the Hilbert space E. Moreover, the representation (3.2.4) of LQ(G) determined by this representation n of G coincides with the original representation TTL of L^(G). Thus, there is one-to-one correspondence between the (strongly continuous) unitary representations n of a locally compact group G and the representations nL (3.2.4) of its group algebra L^G). Furthermore, n and nL have the same cyclic vectors and closed invariant subspaces. In particular, the representation TTL of L^(G) is topologically irreducible if and only if the associated representation TT of G is so. It should be emphasized that, since Lf,(G) is not a G*-algebra, its topologically irreducible representations need not be algebraically irreducible. By irreducible representations of a group G, we will mean only its topologically irreducible representations. The above mentioned correspondence between the unitary representations of a locally compact group G and the representations of the group algebra ^c(^) enables us to apply the GNS construction in order to characterize unitary representations of G by means of positive continuous forms on L^(G). AS was mentioned above, Theorem 3.1.7 is also valid for any involutive Banach algebra A with an approximate identity and positive continuous forms on A. In accordance with Remark 10.5.6, a continuous form on the group algebra LQ(G) is defined as
*(/) = ji>{g)f(g)dg
(3.2.5)
213
Chapter 3 Algebraic Quantization
by an element tp of the space Lg3 (G) of the equivalence classes of integrable complex functions on G of infinite degree. However, the function ip should satisfy the following additional condition in order that the form (3.2.5) to be positive. A continuous complex function ip on G is called positive-definite, if X^teJ19%)CiCj > 0 for any finite set g\,...,gm of elements of G and any complex numbers C\, • • • > Cm- In particular, if m = 2 and g\ = 1, we obtain
i>{g~l) = W),
M(s)
sec,
i.e., ip(l) is bounded. 3.2.1. The continuous form (3.2.5) on LQ(G) is positive if and only if -0 € L^? (G) locally almost everywhere equals a continuous positivedefinite function. • LEMMA
Then cyclic representations of the group algebra L^(G) and the unitary cyclic representations of a locally compact group G are defined by continuous positive-definite functions on G in accordance with the following theorem. 3.2.2. Let 7ty be a representation of LXC{G) in a Hilbert space Eif, and 6^ a cyclic vector for n^, which are determined by the form (3.2.5). Then the associated unitary representation n^ of G in E^ is characterized by the relation THEOREM
1>(9) = «9)h\h)-
(3-2.6)
Conversely, a complex function ip on G is continuous positive-definite if and only if there exists a unitary representation TT,/, of G and a cyclic vector 8$ for 7ty such that the equality (3.2.6) holds. • Example 3.2.2. Let a group G acts on a Hausdorff topological space Z on the left. Let fi be a quasi-invariant measure on Z, i.e., j(g)fi = hg\i where hg is the Radon-Nikodym derivative in Theorem 10.5.1. Then there is the representation G3g:f^
U(g)f,
(U(g)f)(z) = hy\z)f(gz)
(3.2.7)
214
Geometric and Algebraic Topological Methods in Quantum Mechanics
of G in the Hilbert space LQ(Z,/J.) of the equivalence classes of square /xintegrable complex functions on Z [121]. It is a unitary representation due to the equality II/IIM
= / I/(*)IM(*) = / \f{g{z)\2Kg{z)) =
(3.2.8)
Jhg(z)\f(9(z)\2ti(z) = \\n(g)f\\l. A group G can be equipped with the coarsest topology such that the representation (3.2.7) is strongly continuous. For instance, let Z = G be a locally compact group, and let /J, = dg be a left Haar measure. Then the representation (3.2.7) comes to the left-regular representation (Tl(g)f){q) = f{g-1q),
f G L2C(G),
q G G,
(3.2.9)
of G in the Hilbert space LQ(G) of the equivalence classes of square integrable complex functions on G. Let us note that the above mentioned coarsest topology on G is coarser then the original one, i.e., the representation (3.2.9) is strongly continuous. • Let us consider unitary representations of a locally compact group G which are contained in its left-regular representation (3.2.9). In accordance with the expression (3.2.3), the corresponding representation II(/i) of the group algebra L^(G) in L^.(G) reads (H(h)f)(q) = J h(g)(Il(g)f)(q)dg = J h(g)f(g-1q)dg = (h * /)(«/). (3.2.10) Let G be a unimodular group. There is the following criterion that its unitary representation is contained in the left-regular representation. 3.2.3. If a continuous positive-definite function tp on a unimodular locally compact group G is square integrable, then the representation TT,/, of G determined by ip is contained in the left-regular representation II (3.2.9) of G. Conversely, let TT be a cyclic unitary representation of G which is contained in II, and let 6 be a cyclic vector for n. Then continuous positive-definite function (n(g)9\6) on G is square integrable. El PROPOSITION
The representation 7ty in Proposition 3.2.3 is constructed as follows. Given a square integrable continuous positive-definite function ip on G,
Chapter 3 Algebraic Quantization
215
there exists a positive-definite function 6 G L^(G) such that •# = 0*6 = 6*9* = 9* *T. This is a cyclic vector for 7ty. The coefficients (3.2.1) of the representation 7T,/, read
V * $ = 0-
(3.2.11)
An irreducible unitary representation n of a locally compact group G in a Hilbert space £?„. is called square integrable if there exists a vector £ € En such that the coefficient <^£ of n is square integrable over G. This vector is called admissible. If G is a unimodular group, a square integrable representation of G is contained in its left-regular representation in accordance with Proposition 3.2.3. One can show that all coefficients >^i7? of this representation are square integrable over G [129]. Furthermore, there exists a unique constant dv (0 < d^ < oo) such that /
^
*&',„'=<£ W)0£M
(3-2.12)
(3.2.13)
for any £,»7,£V G En. The constant d^ is called the formal dimension of the representation ir. However, its value depends on the choice of the Haar measure on G. If G is compact and dg is the Haar measure of total mass 1, then dn is the (finite) dimension of IT in the usual sense. If -n and n' are inequivalent square integrable representations of G, their coefficients
J
(3.2.14)
(3-2.15)
216
Geometric and Algebraic Topological Methods in Quantum Mechanics
Example 3.2.3. Square integrable representations need not exist. For instance, let G be a commutative group. Its irreducible unitary representation is a character X, i-e., a non-zero homomorphism of G to C such that |x| 2 = 1. If G is compact, each character x of G is square integrable, and so is the corresponding irreducible unitary representation of G. If G is noncompact, it has no square integrable representations. D Now, let G be a connected locally compact (i.e.,finite-dimensional)real Lie group. Any unitary representation of G yields a representation of its right Lie algebra g as follows. A finite-dimensional unitary representation of G in a Hilbert space E is analytic, and the Lie algebra g is represented by bounded operators in E. If IT is an infinite-dimensional (strongly continuous) unitary representation of G in a Hilbert space E, a representation of the Lie algebra g fails to be defined everywhere on E in general. To construct the carrier space of g, let us consider the space /C°°(G,C) C L^(G) of smooth complex functions on G of compact support and the vectors ef=nL(f)e
= J n(g)f(g)edg,
e e E,
f G /C°°(G, C),
(3.2.16)
where nL is the representation (3.2.4) of the group algebra L^(G) [223]. The vectors e/ (3.2.16) exemplify smooth vectors of the representation TT because, for any rj € E, the coefficients (pef,r](g) of vr are smooth functions on G. The vectors ef (3.2.16) for all e £ E and / € /C°°(G,C) constitute a dense vector subspace E^ of E. Let ua be a right-invariant vector field on G corresponding to an element o 6 j . Then the assignment ^ ( a ) : ef -* KL(ua\df)e
(3.2.17)
provides a representation of the Lie algebra g in E^. In particular, let n be a (topologically) irreducible unitary representation of the locally compact Lie group G, and let e = £ be a cyclic vector for 7T. The morphism W^1 (3.3.20) below yields a monomorphism of the space /C°°(G, C) to Eoo C E, and TTOQ (3.2.17) is an irreducible representation of the Lie algebra g in WT1(/COC(G,C)). Given another cyclic vector £'€ J B\^- 1 (/C OO (G,C)) for 7T, we have a different irreducible representation of g. It follows that the representation 7roo (3.2.17) of the Lie algebra g in E ^ is reducible in general.
Chapter 3 Algebraic Quantization
217
Let 7T be a square integrable representation of G. Using the decomposition (3.3.21) below, one can show that EQQ contains all smooth vectors for TT. It is called the Garding space. 3.3
Coherent states
Using unitary square integrable representations of a locally compact group, one conies to the notion of coherent states, which play a prominent role in many quantum models [4; 353; 452]. We start with the more general construction. Let E be a Hilbert space. A family Sz = {£z} of non-zero vectors of E is said to be a reproducing system if it is indexed by points of a locally compact topological space Z, and it determines the decomposition
V = J(v\Z*)tzti*)
(3-3.1)
z of any element 77 of E with respect to some positive measure /x on Z. The integrand in (3.3.1) is brought into the form FZV = (Zzl&PzV, where Pz is the projector onto the vector £z. Hence, the equality (3.3.1) can be rewritten as a resolution of the identity JFZII(Z) = 1.
(3.3.2)
z Remark 3.3.1. In Dirac's bra-ket notation, the integrand in the resolution • of the identity (3.3.2) reads Fz = %){(,z\. A reproducing system need not exist. In a generic case, one deals with a separable Hilbert space E and a positive bounded operator a in E which admits a resolution
JFzii{z) = a z
(3.3.3)
with respect to a positive measure /x on some locally compact topological space Z such that Fz are compact operators in E. The operator a in (3.3.3) need not be invertible, i.e., the domain of a~1 is not dense in E.
218
Geometric and Algebraic Topological Methods in Quantum Mechanics
If the operator a is invertible (but a" 1 need not be bounded), one says that (E,Fz,a) is the reproducing triple. For instance, this is the general case of covariant coherent states below. If the operator a" 1 is bounded and all operators Fz are of the same rank n, one can construct a reproducing system of vectors {/*}, i = 1 , . . . , n, of the Hilbert space E which provides a resolution of the identity. Then, similarly to coherent states below, one also call /* the coherent states [4]. In particular, we throughout are in the case of n = 1. Let us consider a complex function (3.3.4)
K(z,z') = (^zl)
on the Cartesian product Z x Z. It is called a reproducing kernel because of the relation K(z, z') = jK(z, z")K(z", z')n(z"), z
z, z',z" € Z,
(3.3.5)
derived from the decomposition (3.3.1). By the very definition of K(z,z') (3.3.4), we also have K{z,z)>0,
K(z,z') = K{z',z),
z,z'eZ.
(3.3.6)
Let us consider the morphism W:E^Ll(Z,fi),
(Wr,)(z) = (r]\^),
r, G E,
(3.3.7)
of E to the Hilbert space L$.(Z, fi) of the equivalence classes of square integrable complex functions on Z. This is an isometric monomorphism of Hilbert spaces whose range Ew consists of the equivalence classes of square integrable functions f on Z satisfying the equality
f(z)= jK(z',z)f(z')fi(z'). z
(3.3.8)
A glance at this equality shows that the functions K(z,z') = WZz(z'),
*eZ,
constitute a reproducing system in the Hilbert space Ew- Namely, the corresponding decomposition (3.3.1) reads
/(*)= / Jf(z")K(z",z>Mz") K(z',z)Kz'). z Lz
219
Chapter 3 Algebraic Quantization
The inverse W" 1 : Ew -> E to the morphism W (3.3.7) reads
z Then one can assign the operator
J=W~1ofoW=
jf{z)Fzix{z) z
on the Hilbert space E to any element / G Lg^Z,/z) (see Remark 10.5.6), e.g., to any bounded continuous function on Z. Conversely, let a £ B(E) be a bounded operator in the Hilbert space E. One can associate to a a complex function fa(z,z') = (a£z\Sz')
(3-3.9)
/«(*) = {a£x\Zz)
(3-3.10)
on Z x Z. Its restriction
to the diagonal Z in Z x Z is a, mean value function. By virtue of the decomposition (3.3.1), we obtain (fa * fb)(z) := J fa(z', z)fb(z, *>(*') = fab(z), z
a, b € B(E).
(3.3.11)
Let us note that this formula is also true for unbounded operators a and b if Sz c D(a),
Sz C D(b),
bSz C D{a).
A unitary square integrable representation of a locally compact group G exemplifies a reproducing system. Here, G is not necessarily a unimodular group, and not all coefficients of its square integrable representation are square integrable functions on G [4; 20l]. Let 7r be a square integrable representation of G and £ an admissible vector, i.e.,
j M9)m)\2d9 < oo.
(3.3.12)
220
Geometric and Algebraic Topological Methods in Quantum Mechanics
It is readily observed that every vector n(g)£, g e G, is admissible. Hence, the set 5(TT) of admissible vectors is G-invariant and dense in the carrier space E. If G is a unimodular group, then 5(71") = E. 3.3.1. If TT is a square integrable representation of G, there exists a unique closed positive invertible operator C on S(TT) C E such that PROPOSITION
J
V,
i G E.
(3.3.13)
• It follows that coefficients
J\
(3-3.14)
i.e., (C£|C£) = 1. Then one can easily justify that the function K(g, g') = (n(g)^(g')0 = ^Mg^(g)
(3.3.15)
is a reproducing kernel on G, i.e., it obeys the relations K(g,g) >0,
g^G,
K(g,g') = jK(g,g")K(g",g')dg".
(3.3.16)
In particular, if G is a unimodular group, then (£|£) = dn and the relation (3.3.16) for the reproducing kernel (3.3.15) follows from (3.2.13). Thus, we come to the conventional notion of a coherent state. Let 7T be a square integrable representation of a locally compact group G, and let £ G E be an admissible vector obeying the condition (3.3.14). Let us consider the orbit St = {tg=n(g)£:geG}
(3.3.17)
of the group G in E. Its elements £g are called coherent states. They possess the following important properties.
221
Chapter 3 Algebraic Quantization
(i) Since •n is an irreducible representation, coherent states (3.3.17) constitute a total family, but this family is not orthonormal in general. (ii) Coherent states define the reproducing kernel (3.3.15) on the group G. (iii) Let us consider the map Wc-E^LliG),
(W(r,)(g)=J^)
= (r,\7r(g)0,
V e E.
(3.3.18)
By virtue of the relations (3.3.13) and (3.3.14), it is a monomorphism of Hilbert spaces whose range E^ is dense in LQ(G). Moreover, the map W% (3.3.18) is the intertwining operator between •K and the left-regular representation II of G, i.e., Wt«(g) = TI(g)Wz for all g €G. (iv) It follows that a square integrable representation TT of a locally compact group G is contained in its left-regular representation. Using the relations (3.3.13) and (3.3.14), one can show that the space E$ consists of elements / of L^ (G) obeying the condition
f(g) = jK(g',g)f(g')dg',
(3.3.19)
where K{g,g') is the reproducing kernel (3.3.15). (v) The inverse morphism E^ —> E reads
W[1(f)=Jf(g)£gdg,
/ e %
(3.3.20)
In particular, we obtain the decomposition
V = j(W^)(g)^dg
= J(v\Zg)tgdg
(3.3.21)
of any element r) e E with respect to coherent states (cf. the decomposition (3.3.1)). Thus, coherent states constitute a reproducing system. There is the following serious inconsistency of the above construction of coherent states. Let us consider the integral (3.3.12) and suppose that H% is the subgroup of G which keeps £ invariant up to a phase multiplier, i.e., TT(/I)£
= eia(-h\,
h G Hv
(3.3.22)
Clearly, the integrand in (3.3.12) does not depend on g, but only on a coset q € G/H^. Consequently, the finiteness of the integral (3.3.12) forces
222
Geometric and Algebraic Topological Methods in Quantum Mechanics
the subgroup H^ to be compact. In order to avoid this restriction, one introduces the notion of covariant coherent states [4]. Let G be a locally compact group and n its strongly continuous irreducible unitary (not necessarily square integrable) representation in a Hilbert space E. Let H be a closed subgroup of G. We suppose that: • G/H is endowed with a positive G-invariant measure fi (this is the case of a compact group H [23]), • there exists a /i-measurable (not necessarily global) section a of the principal fibre bundle G —> G/H, • there exists a vector £ £ E such that the integral
j IM
V£E,
(3.3.23)
G/H
is finite and aa is a bounded positive invertible operator in E (though a" 1 need not be bounded). Let £ be normalized by (£|aCT£) = 1- Then the family of covariant coherent states is denned as Sa = {&(,) = 7r(orfa))e, q e G/H}.
(3.3.24)
The following summarizes the main properties of covariant coherent states. (i) The family 5CT (3.3.24) is total in E. (ii) Let us introduce the function Ka(q,q') = (t;aiq)\a-1Zaiq,))
(3.3.25)
on the Cartesian product G/H x G/H. There is a monomorphism
Wt:E-^Ll{G/H,n),
(W^)te) = fol&(,)>, V G E.
(3.3.26)
Its range E^ consists of the equivalence classes of square integrable functions / on G/H which satisfy the relation
/()= J W , )/(
similar to the condition (3.3.19) in the case of coherent states.
Chapter 3 Algebraic Quantization
223
(iii) The inverse morphism E% —>• E reads
G/H
(cf. (3.3.20)). In particular, we have
G/H
G/H
or, equivalently,
aaV = I (»?|&(,)>&(,)M9)-
(3-3.27)
G/H
(iv) The integrand in (3.3.27) can be rewritten as
where Pa(q) is the projector onto the vector ^ ^ . Then the equality (3.3.27) takes the form
aa=
J Fa{q)ii{q).
(3.3.28)
G/H
In particular, if G/H is compact, aa is a Hilbert-Schmidt operator. (v) The range E^ of the monomorphism W^ (3.3.26) is complete with respect to the norm
(f\f), = (ftW^a-'W^f)
(3.3.29)
so that the morphism W^ of E onto E^ is isometric. The above mentioned results simplify somewhat if the operator a^1 is bounded. Then Sa is called a frame.. This terminology is borrowed from the theory of non-orthogonal expansions [122J. The further simplification occurs when Sa is a tight frame, i.e., aa — Al, A > 0. This is also the case of generalized coherent states when H = H^ (3.3.22) [353]. If aa = 1, the
224
Geometric and Algebraic Topological Methods in Quantum Mechanics
function Ka(q,q') (3.3.25) fulfils the relation K{q,q)a>0,
qeG/H,
tf(g,<7%=W^)CT, * W ) * = J K(q,q")aK(q",q%Li(q") G/H
and, thus, defines a reproducing kernel on G/H. The norm (3.3.29) coincides with the standard norm on the space L^(G/H, /x) so that Wj (3.3.26) is an isometric monomorphism. The equality (3.3.28) provides a resolution of the identity J Fa{q)n{q) = 1
(3.3.30)
G/H
(cf. (3.3.2)). Thus, generalized coherent states constitute a reproducing system. Of course, the above mentioned construction of covariant coherent states depends on the choice of a particular section a of the principal bundle G —» G/H. However, there is a certain degree of "section-independence" in the method. In particular, if a provides the admissibility condition (3.3.23), the similar condition holds for any translated section <Jg(q) = 9v(9~1q),
g^G.
If two sections a and a' fulfil the admissibility conditions (3.3.23) for a given vector £, there is one-to-one correspondence between the corresponding families of covariant coherent states Sa and Sai carried out by bounded operators £
=Tqt.a>(q).
One is also free to specify a vector £ of the carrier space E in order to obtain a family of covariant coherent states. In many quantum models, there exists a vector £ such that the covariant coherent states produced from £ appear to be minimal uncertainty states. 3.4
GNS construction III. Groupoids
Generalizing the case of a locally compact group, one can associate a C*algebra to a locally compact groupoid equipped with a left Haar system as
Chapter 3 Algebraic Quantization
225
follows [367]. Let a topological groupoid (5 (see Section 10.3) be locally compact. Let /C(<5, C) be the space of continuous complex functions of compact support on (55 endowed with the inductive limit topology. A left Haar system for a groupoid (25 is a family of measures {/xu,u £ <50} on (25 indexed by points of the unit space (25° of (25 such that: • the support of the measure /xu is (25U; • for any function / £ /C((9, C), the function /^e^uHnu(f)
ec
(3.4.1)
is continuous; • for any x € 6 and / G £(<25,C),
/ f(xy)vi(x)(y) = / f(y)vr{x)(y)One can show that, if 0 is a locally compact groupoid with a left Haar system, then the map tC(<5,C)3f^U£tC(®°,C) is continuous and the surjection r : (25 —> <S° is open. Example 3.4.1. Let G be a locally compact group which acts continuously on a locally compact space S. Let us consider the action groupoid 5 x G in Example 10.3.2. It is a locally compact groupoid with the left Haar system {fis, s € S}, where ns = es
(3.4.2)
provided with the relative topology and the following groupoid structure:
.{H,g)-l:={H,g-1); • a pair ((H,g), (H',g')) is composable if and only if H = H';
226
Geometric and Algebraic Topological Methods in Quantum Mechanics
.(H,g)(H,g>) = (H,gg>). The groupoid (3.4.2) is a locally compact group bundle over SQ called the bundle of subgroups of a group G. It admits the left Haar system {/J,JI, H G SG} where /J,H the left Haar measure on H. • Example 3.4.3. A locally compact groupoid 0 is called r-discrete if its unit space is an open subset. This groupoid possesses the following properties. • For any u £ 0°, the subspaces 0 U and 0 U are discrete. • If © admits a left Haar system, one can choose it to be /Uu({u}) := 1 for all u € 0°. Then the invariance condition results in fj,u({x}) = 1 for all • The multiplication 0 2 —> 0 and the map r : 0 —> 0° are local homeomorphisms. • Let 0 be a locally compact groupoid with a left Haar system {fiu}- Let us denote fiu = [i~l the inverse measure. The measures fiu, u € (5°, are assembled into a right Haar system. Given a measure A on the unit space 0°, let us introduce the measure
v=fiiu\{u)
(3.4.3)
on <8 by means of the condition
K / ) = //M(«)M«).
/e/C(0,C),
where /M is the function (3.4.1). A measure A on 0° is said to be quasiinvariant (resp. invariant) if the measure v (3.4.3) is equivalent (resp. equal) to the inverse measure V-i
= j f\{u).
Example 3.4.4. If 0 = G is a locally compact group, then 0° = {1} and the left Haar system for 0 reduces to the left Haar measure on G. A quasi-invariant measure on {1} is exemplified by the Dirac measure, and the measure v (3.4.3) on G coincides with the left Haar measure. • Example 3.4.5. Let us consider the action groupoid 0 = S x G in Example 10.3.2 with the left Haar system in Example 3.4.1. Let A be a
Chapter 3 Algebraic Quantization
111
measure on 5. Then the measure v (3.4.3) on (25 is the product A
{f*9){x) = Jf(xy)g{y-1)a(xy,y-1)fM{x)(y),
(3.4.4)
r(x)=f(x-i)
(3.4.5)
Let us denote this algebra by C(<8,cr). Of course, this algebra depends on the choice of a two-cocycle a. One however can show that, if cocycles a and a' belongs to the same cohomology class, the involutive algebras C(<S,a) and C(<8,a') are isomorphic. The involutive algebra C(<8,a) becomes a normed algebra, called the convolution algebra, if one provides the space /C((5,C) with the norm
||/||w=max(||/|| r , H/IIO,
||/|| P = sup f\f\nu,
/G/C(0,C),
(3.4.6)
Il/U, = sup f \ f \ n \
The corresponding normed topology on /C((25,C) is finer than the inductive limit topology. By a representation n of the convolution algebra C(<25,
lk(/)ll < ll/llw for a l l / e / C ( 0 , C ) . A representation of the convolution algebra C(<S,a) can be constructed as follows. Let the unit space 0° be provided with a quasi-invariant measure
228
Geometric and Algebraic Topological Methods in Quantum Mechanics
A. Let H —> (5° be a group bundle whose fibres are separable Hilbert spaces. Let us consider a morphism
of the groupoid 0 to the frame groupoid Iso H in Example 10.3.4 such that: • ^(x) is represented by some morphism
• the equality
V(x)V(y) =
on<252; • the equality
^(a;)- 1
=a(x,x-1)V(x-1)
holds almost everywhere on <S with respect to the measure v (3.4.3); • the map
is ^-measurable for any pair of A-measurable sections 4> and 4>' of the group bundle H -> <5°. The morphism $ is called the a-representation of the groupoid 0 in H. Two cr-representations ($!,H,\) and ( # ' , # ' , A') of (5 are said to be equivalent if there exists an isomorphism $ of if to H' over (8° such that (*'o$)(a;) = ($o^)(x)
almost everywhere on 0 with respect to the measure v (3.4.3). Let L2(H, A) be the Hilbert space of sections of H —> (5° which are square A-integrable. Then a representation of the convolution algebra C(<8,a) in L2(H,X) is given by the expression 7T*(/)0(«) = j f(x)(V(x)(ci> o l)(x)A-V2v(x),
(3-4.7)
Chapter 3 Algebraic Quantization
229
where A is the modular function of the measure v. Different arepresentations \? and \&' of the groupoid 6 obviously yield different representations 7T* and 7r
Example. Algebras of infinite qubit systems
Let Q be the two-dimensional complex space C2 equipped with the standard positive non-degenerate Hermitian form (.|.)2- Let Mi be the algebra of complex 2 x 2-matrices seen as a C*-algebra. A system of m qubits is usually described by the Hilbert space m
and the C* -algebra m
Am = ®M 2 , called the input algebra, which coincides with the algebra B{Em) of bounded operators in Em [240]. One can straightforwardly generalize this description to an infinite set S of qubits by analogy with a spin lattice [144]. Its algebra As admits non-equivalent irreducible representations. If S = Z + , there is one-to-one correspondence between the representations of As and those of an algebra of canonical commutation relations. Given a system of m qubits, one also considers an algebra of complex m
functions on the set x Z2. It is regarded as the output algebra of a qubit system [240]. There is a natural monomorphism of this algebra to the
230
Geometric and Algebraic Topological Methods in Quantum Mechanics
input one Am. Using the technique of groupoids in previous Section, one can generalize these algebras to infinite qubit systems. We start with the input algebra, and follow the construction of infinite tensor products of Hilbert spaces and C*-algebras in [144]. Let {Qs,s £ S} be a set of two-dimensional Hilbert spaces Qs = C 2 . Let x Qs be s the complex vector space whose elements are finite linear combinations of elements {qs} of the Cartesian product J\QS of the sets Qs- The tensor s product
8 = {e.}e]lQ. s e
such that all 0S ^ 0, let us denote ® Qs the subspace of (g>Qs spanned by s the vectors ®qs where qs ^ 6S only for a finite number of elements s G S. It is called the 9-tensor product of vector spaces Qs, s £ S. Let us choose a family 6 = {0S} of normalized elements 8S £ Qs, i.e., all \63\ = 1. Then ®eQs is a pre-Hilbert space with respect to the positive non-degenerate Hermitian form s€S
Its completion Qes is a Hilbert space whose orthonormal basis consists of the elements
eir = ®qs,
r£S,
i = 1,2,
such that qsitr = 6S and qr = ei, where {e^} is an orthonormal basis for Q. Let now {A3,s € 5} be a set of unital C*-algebras As = Mi. These algebras are provided with the operator norm H I = (AoAo + AiAi + A2A2 + A 3 A3) 1/2 ,
a = i\ol
+
^ A ^ , t=l,2,3
where a1 are the Pauli matrices. Given the family { l s } , let us construct the {l s }-tensor product ®AS of vector spaces As. It is a normed involutive
Chapter 3 Algebraic Quantization
231
algebra with respect to the operations (<8>as)(®a's) = <8>(aaa's),
(<8>as)* =
and t h e norm
I ® a J = JJlKUs
Its completion As is a C*-algebra. Then the following holds [144]. PROPOSITION 3.5.1. Given a family 9 = {6S} of normalized elements 9S E Qs, the natural representation of the involutive algebra ®AS in the pre-Hilbert space ®BQ3 is extended to the representation of the C*-algebra As in the Hilbert space Qes such that As = B(Qes) is the algebra of all bounded operators in Q8S. D PROPOSITION 3.5.2. Given two families 9 = {9S} and 9' = {9's} of normalized elements, the representations of the C*-algebra As in the Hilbert spaces Qs and Qs are equivalent if and only if
£||<6U60|-l|
ses
D
For instance, if S = Zfc, we come to a spin lattice. Let us turn now to the output algebra. One can associate to a system of qubits {Qs,s £ S} the following groupoid. Let Z2 = {1, p : p 2 = 1} be the smallest Coxeter group. Let us consider the set X = Z2S of Zivalued functions on 5. It is a set 2s of all subsets of S, and it can be brought into a Boolean algebra. Let G C X be a subset of functions which equal p G Z2 at most finitely many points of S. Both X and G are commutative Coxeter groups with respect to the pointwise multiplication. Given the action of G on X on the right, the product <8 = X x G is the action groupoid whose unit space 0° is naturally identified with X (see Example 10.3.2). Since G acts freely on X, the action groupoid (25 is principal, i.e., the map (r, I ) : 0 - 4 0 ° x 0 °
232
Geometric and Algebraic Topological Methods in Quantum Mechanics
is an injection. Provided with the discrete topology, X is a locally compact space. Then (5 is a locally compact groupoid. Since (25° C (25 is obviously an open subset, (25 is an r-discrete groupoid. Let us note that the action groupoid (25 is isomorphic to the following one. Let us say that functions x, y € X are equivalent x ~ y if and only if they differ at most finitely many points of S. Let (25' c X x X be the graphic of this equivalence relation, i.e., it consists of pairs (x,y) of equivalent functions. One can provide & with the following groupoid structure: • a pair ((x, y), (y', z)) is composable if and only if y = y', • {x,y)~l -=(y,x), • {z,y)(y,z) = (x,z), • r : ( i , j/) i-> (x,x), I : (x,y) H-> (y,y). The unit space (25/0 of this groupoid is naturally identified with X. The isomorphism of groupoids (25 and (25' is given by the assignment <8B(x,g)^(x,xg)e&.
In particular, 0 * / 0 if and only if x ~ y. Let /C(<25, C) be the space of complex functions on (25 of compact support provided with the inductive limit topology. Since (55 is a discrete space, any function on 0 is continuous. A left Haar system for the groupoid (25 is a family of measures {fj,x, x e X} on 6 indexed by points of the unit space <S° = X such that: • the support of the measure /xx is <25X, • for any (x,g) € (25 and any / € K.(<8,C), we have
J f((x,g)(y,g'))fixg(y,g') = j ' f(v,g')nx(v,g'). Since 0 ° is a discrete space, the function (3.4.1) is always continuous. As was mentioned in Example 3.4.1, a left Haar system for the action groupoid (25 is given by the measures Mx = £x x
MG,
where ex is the Dirac measure on X with support at x G X and HG is the left Haar measure on the locally compact group G. We have
/ f(v,g)i*x{y,g) = / f(x,g)nG(g) =
J
J
^2f(x,g),
g€G
where the sum is finite since / € £(<25,C) are of compact support.
233
Chapter 3 Algebraic Quantization
With this left Haar system, the space /C(<8,C) is brought into the following convolution algebra C(<5). Since the t/(l)-extension of G is trivial up to equivalence [288], the cohomology group H2(<&, 21) is trivial (see Theorem 1.5.1 and Remark 10.3.7), and algebraic operations (3.4.4) - (3.4.5) in /C(<5,C) read
(f*f)(x,9)= ^/(z.^O/W 1 .^" 1 ).
(3.5.1)
g'eG
(3.5.2)
r{x,g) = f((x,g)-i). This algebra is provided with the norm (3.4.6):
Il/H = max(sup Y2\f(x,g)\, sup £ l/O^^fiT1)!)X
9
9
In particular, let us consider a subspace Ac C C(<&) of functions f(x,g) = f(g) independent of x £ X. The algebraic operations (3.5.1) - (3.5.2) on these functions take the form
(/ * /')() = £ fiagV'ia'-1),
rig) = 'W1) = W)-
g'eG
Thus, AQ is exactly the group algebra of the locally compact group G provided with the norm
11/11 = £ 1/(5)1g
There is the monomorphism of this algebra to the algebra ®AS as follows. Let us assign to each element g £ G the element "g = ®as G <8>AS, where aa = 1 if g(s) = 1 and as = cr1 if g(s) = p. Then the above mentioned monomorphism is given by the association
/(s)^ £/()?• g
Thus, one can think of Ac and, consequently, C(&) as being a generalization of the output algebra in [240]. The monomorphism Ac —> ®AS provides the representations of the involutive algebra Ac in Hilbert spaces Qss. One can construct representations of the whole convolution algebra C(<S) as follows. Given the Hilbert space Q6S, let us consider the product € = X x Q%
234
Geometric and Algebraic Topological Methods in Quantum Mechanics
seen as a group bundle in the Abelian groups Q% over X. It is a groupoid such that • a pair
{{x,v),(x',v')),
v,v'eQes,
x,x'£X,
is composable if and only if x' = x, • {x^)-1 :=(x,-v), • (x,v)(x,v') = (x,v + v'), • r((x,v)) = l((x,v)) =x. Its unit space is X, and its fibres €x are isomorphic to Q6S. Let the discrete topological space X be provided with the measure A such that
/e/C(A\C).
ff(x)X(x)=^2f(x), J
x
Let Iso <£ be the set of all isomorphisms 7j:€ w -»Cx,
x,ye<5°.
It is the frame groupoid in Example (10.3.4) whose unit space <£° is naturally i > Id
(7r(/)0)(z) = £ / ( ( * , s))^,?). 9
The algebra C(<3) can be provided with the operator norm ||/|| = ||7r(/)|| with respect to this representation. Its completion relative to this norm is the C*-algebra of the action groupoid <S.
3.6
GNS construction IV. Unbounded operators
There are algebras (e.g., of quantum fields) whose representations in Hilbert spaces need not be normed. Therefore, the generalization of the conven-
Chapter 3 Algebraic Quantization
235
tional GNS representation of C*-algebras to some classes of unnormed topological involutive algebras has been studied. By an operator in a Hilbert (or Banach) space E is meant a linear morphism a of a dense subspace D(a) of E to E. The D(a) is called a domain of an operator a. One says that an operator b on D(b) is an extension of an operator a on D(a) if D(a) C D(b) and &|.D(a) = a. For the sake of brevity, we will write a C b. An operator a is said to be bounded on D(a) if there exists a real number r such that ||oe||
e G £>(a).
If otherwise, it is called unbounded. Any bounded operator on a domain Z)(a) is uniquely extended to a bounded operator everywhere on E. Therefore, by bounded operators in E are usually meant bounded (continuous) operators defined everywhere on E. An operator a on a domain D(a) is called closed if the condition that a sequence {e{\ C D(a) converges to e G E and that the sequence {aej} does to e' G E implies that e £ -D(G) and e' = ae. Of course, any operator defined everywhere E is closed. An operator a on a domain D(a) is called closable if it can be extended to a closed operator. The closure of a closable operator a is defined as the minimal closed extension of a. Operators a and b in E are called adjoint if (ae|e'> = (e\be'), e € £>(a),
e' G D(b).
Any operator a has a maximal adjoint operator a*, which is closed. Of course, a C a** and b* c a* if a C b. An operator a is called symmetric if it is adjoint to itself, i.e., a C a*. Hence, a symmetric operator is closable. One can obtain the following chain of extensions of a symmetric operator: a C a C a** C a* = a* = a***. In particular, if a is a symmetric operator, so are a and a**. At the same time, the maximal adjoint operator a* of a symmetric operator a need not be symmetric. A symmetric operator a is called self-adjoint if a = a*, and it is called essentially self-adjoint if a = a* = a*. It should be emphasized that a symmetric operator a is sometimes called essentially self-adjoint if a** = a*. We here follow the terminology of [360]. If a is a closed operator, the both notions coincide. For bounded operators, the notions of symmetric, self-adjoint and essentially self-adjoint operators coincide.
236
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let E be a Hilbert space. The pair (B, D) of a dense subspace D of E and a unital algebra B of (unbounded) operators in E is called the Op*-algebra (O*-algebra in the terminology of [394]) on the domain D if, whenever b £ B, we have: (i) D(b) = D and bD C D, (ii) D C £>(&*), (iii) 6* |D C B [222; 360]. The algebra B is provided with the involution b\-+ b+ = b*\o, and its elements are closable. A representation TT(A) of an involutive algebra A in a Hilbert space E is an Op*-algebra if there exists a dense subspace D(ir) C E such that D(n) = D(n(a)) for all a £ A and this representation is Hermitian, i.e., 7r(a*) C vr(a)* for all a € A. In this case, one also considers the representations W:a->W(a):=ir(a)\D(^,
D(W) = Q D(7r(a)),
7T* : a -» 7T*(a) := 7r(a*)*|D(7r.);
£ ( 0 = f| £>(ir(a)'), a£A
7T** : a -»7r**(a) := 7r*(aT| D(7r .. );
£ ( O = f ) i?(7r'(a)'),
called the closure of a representation TT, an adjoint representation and a second adjoint representation, respectively. There are the representation extensions 7T C 7f C 7T** C 7T*,
where TTI C TT2 means -D(TTI) C £)(TT2). The representations n and ?r** are Hermitian, while TT* = TT* = TT***. A Hermitian representation n(A) is said to be closed if vr = ?f, and it is self-adjoint if vr = TT*. Herewith, a representation 7T(J4) is closed (resp. self-adjoint) if one of operators ir(A) is closed (resp. self-adjoint). The representation domain D(ir) is endowed with the graph-topology. It is generated by the neighborhoods of the origin U(M,e) = {x€D(n)
: £
||7r(a)z|| < e},
where M is a finite subset of elements of A. All operators of TT(A) are continuous with respect to this topology. Let us note that the graph-topology
Chapter 3 Algebraic Quantization
237
is finer than the relative topology on -D(TT) C E, unless all operators ir(a), a G A, are bounded [394]. Let N denote the closure of a subset N C D(TT) with respect to the graph-topology. An element 6 e D(ir) is called strongly cyclic (cyclic in the terminology of [394]) if
D(TT) C R I ) 0 f . Then the GNS representation Theorem 3.1.7 can be generalized as follows [222; 394]. THEOREM 3.6.1. Let A be a unital topological involutive algebra and / a positive continuous form on A such that / ( I ) = 1 (i.e., / is a state). There exists a strongly cyclic Hermitian representation {^f,0f) of A such that
4>{a) = (7r(a)0*|fy),
a e A.
a We point out the particular class of nuclear barreled "-algebras. Let A be a locally convex topological involutive algebra whose topology is denned by a set of multiplicative seminorms pt which satisfy the condition pL(a*a)=pL(a)2,
a & A.
It is called a b*-algebra. A unital 6*-algebra as like as a C*-algebra is regular and symmetric, i.e., any element (1 + a*a), a € A, is invertible and, moreover, (1 -{-a*^^1 is bounded [5; 227]. The 6*-algebras are related to C*-algebras as follows. PROPOSITION 3.6.2. Any 6*-algebra is the Hausdorff projective limit of a • family of C*-algebras, and vice versa [227].
In particular, every C*-algebra A is a barreled 6*-algebra, i.e., every absorbing balanced closed subset is a neighborhood of the origin of A. Let us additionally assume that A is a nuclear algebra (i.e., a nuclear space). Then we have the following variant of the GNS representation theorem [227] 3.6.3. Let A be a unital nuclear barreled 6*-algebra and / a positive form on A. There exists a unique (up to unitary equivalence) cyclic representation 717 of A in a Hilbert space by operators on a common invariant domain D. This domain can be topologized to conform a rigged THEOREM
238
Geometric and Algebraic Topological Methods in Quantum Mechanics
Hilbert space such that all the operators representing A are continuous on D. D The following is an example of a nuclear barreled b*-algebra which is very familiar from quantum field theory. Let Q be a nuclear space. Let us consider the direct limit §Q = C © Q © Q § Q © - - . Q ® n © . . .
(3.6.1)
of the vector spaces ®~nQ = C © Q © Q®Q • • • © Q®n, where ® is the topological tensor product with respect to the Grothendieck's topology (which coincides with the e-topology on the tensor product of nuclear spaces [356]). The space (3.6.1) is provided with the inductive limit topology, the finest topology such that the morphisms ®~nQ —» ®Q are continuous and, moreover, are imbeddings [417]. A convex subset V of ®Q is a neighborhood of the origin in this topology if and only if V D ®~~nQ is so in ®~ Q. Furthermore, one can show that ®Q is a unital nuclear barreled LF-algebra [31]. The LF-property implies that a linear form / on ®Q is continuous if and only if the restriction of / to each §~"Q is so [417]. If a continuous conjugation * is defined on Q, the algebra ®Q is involutive with respect to the operation *(9i®"-®g n ) = 9n®-"9i
(3-6-2)
on Q®" extended by continuity and linearity to Q®n. One can show that ®Q is a b*-algebra as follows. Since Q is a nuclear space, there is a family ||.||fe, k £ N + , of continuous norms on Q. Let Qk denote the completion of Q with respect to the norm ||.||fe. Then one can show that the tensor algebra ®Qk is a C*-algebra and that ®Q (3.6.1) is the projective limit of these C*-algebras with respect to morphisms ®Qk+i —> ®Qk [227]. In quantum field theory, one usually choose Q the space of functions of rapid decrease or the Schwartz space. 3.7
Example. Infinite canonical commutation relations
The canonical commutation relations (henceforth the CCR) are of central importance in many quantum models as the traditional method of canonical
Chapter 3 Algebraic Quantization
239
quantization. A remarkable result about the CCR for finite degrees of freedom is the Stone-von Neumann uniqueness theorem which states that all irreducible representations of the CCR for n degrees of freedom are unitarily equivalent [354]. On the contrary, the CCR for infinite degrees of freedom admit infinitely many inequivalent irreducible representations (see [157] for a survey). One can find the comprehensive description of representations of the CCR modelled over an infinite-dimensional nuclear space Q in [164]. Let Q be a real nuclear space provided with a non-degenerate separately continuous Hermitian form (.|.). This Hermitian form makes Q into a Hausdorff pre-Hilbert space. A nuclear space Q, the completion Q of the pre-Hilbert space Q, and the (topological) dual Q' of Q make up the rigged Hilbert space Q C Q C Q' (3.1.12). Given a real nuclear space Q together with a non-degenerate separately continuous Hermitian form (.|.), let us consider the group G(Q) of triples 9 = (9i)<72,A) of elements q\, qi of Q and complex numbers A of unit modulus which are subject to multiplications (9i»92, W n g * A') = (qi +
(3-7.1)
It is a Lie group whose group space is a nuclear manifold modelled over W = Q®Q®R.
(3.7.2)
Let us denote T(q) = (q,0,0),
P(q) = (0,q,0).
Then the multiplication law (3.7.1) takes the form T(q)T(q') = T(q + q'), P(q)P(q') = P(q + q'),
(3.7.3)
P(q)T(q')=exp[i(q\q')}T(q')P(q). Written in this form, G(Q) is called the nuclear Weyl CCR group. The complexified Lie algebra of the nuclear Lie group G(Q) is the Heisenberg CCR algebra Q{Q). It is generated by the elements <j>{q), n(q), q G Q, and / which obey the Heisenberg CCR commutation relations [
\cj>(q), <j>(q')} = [*(), 7r( 9 ')] = 0, (3.7.4) (3.7.5)
240
Geometric and Algebraic Topological Methods in Quantum Mechanics
There is the exponential map T{q) = exp[^( 9 )],
P(q) = exp[t7r(g)].
Due to the relation (3.1.11), the normed topology on the pre-Hilbert space Q defined by the Hermitian form (.|.) is coarser than the nuclear space topology. The latter is metric, separable and, consequently, secondcountable. Hence, the pre-Hilbert space Q is also second-countable and, therefore, admits a countable orthonormal basis. Given such a basis {q^} for Q, the Heisenberg CCR (3.7.4) - (3.7.5) take the form [0(9j), 4>{
[n(qj),
-iSjkI.
The CCR group G(Q) contains two nuclear Abelian subgroups T(Q) and P(Q). Following the representation algorithm in [164], we first construct representations of the nuclear Abelian group T(Q). These representations under certain conditions can be extended to representations of the whole CCR group G(Q). One can think of the nuclear Abelian group T(Q) as being the group of translations in the nuclear space Q. Its cyclic strongly continuous unitary representation n in a Hilbert space (E, (.|.)s) with a (normed) cyclic vector 8 £ E defines the complex function Z(q) = (ir(T(q))e\6)E on Q. This function is proved to be continuous and positive-definite, i.e., Z(0) = 1 and
Y^Z{qi-qj)ciCj>Q for any finite set qi,... ,qmoielements of Q and arbitrary complex numbers Ci,.. . , c m .
In accordance with Bochner's theorem for nuclear spaces in Example 10.5.11, any continuous positive-definite function Z(q) on a nuclear space Q is the Fourier transform
Z{q) = J exp{i{q,u)}n(u)
(3.7.6)
of a positive measure \i of total mass 1 on the dual Q' of Q [164]. Then the above mentioned representation n of T(Q) can be given by the operators Tz(q)p(u) = ex.p[i(q, u)]p(u)
(3.7.7)
Chapter 3 Algebraic Quantization
241
in the Hilbert space L%.(Q',[i) of the equivalence classes of square p,integrable complex functions p(u) on Q'. The cyclic vector 9 of this representation is the ^-equivalence class 6 « M 1 of the constant function p{u) = 1. Then we have Z{q) = (Tz(q)e\6)fl = J exp[i(q, u)]fi.
(3.7.8)
Conversely, every positive measure /x of total mass 1 on the dual Q' of Q defines the cyclic strongly continuous unitary representation (3.7.7) of the group T(Q). By virtue of the above mentioned Bochner theorem, it follows that every continuous positive-definite function Z(q) on Q characterizes a cyclic strongly continuous unitary representation (3.7.7) of the nuclear Abelian group T(Q). We agree to call Z(q) a generating function of this representation. It should be emphasized that the representation (3.7.7) need not be (topologically) irreducible. For instance, let p(u) be a function on Q' such that the set where it vanishes is not a jii-null subset of Q'. Then the closure of the set Tz(Q)p is a T(Q)-invariant closed subspace of L^(Q', /i). One can show that distinct generating functions Z(q) and Z'(q) determine equivalent representations Tz and Tz> (3.7.7) of T(Q) in the Hilbert spaces LQ(Q', /x) and L^(Q', pi') if and only if they are the Fourier transform of equivalent measures on Q' [164]. Indeed, let / / = s2fi,
(3.7.9)
where a function s(u) is strictly positive almost everywhere on Q', and fi(s2) = 1. Then the map Ll(Q', fi') 9 p(u) -> s(u)p(u) e L2C(Q', IM)
(3.7.10)
provides an isomorphism between the representations Tz> and TzThe representation Tz (3.7.7) of the nuclear Abelian group T(Q) in the Hilbert space L}.(Q',n) determined by the generating function Z (3.7.6) can be extended to the CCR group G(Q) if the measure y, possesses the following property. Let uq, q £ Q, be an element of Q' given by the condition (ql,ug) = (q'\q),
q' € Q.
(3.7.11)
These elements form the range of the monomorphism Q —> Q' determined by the Hermitian form (.|.) on Q. Let the measure /x in (3.7.6) remains
242
Geometric and Algebraic Topological Methods in Quantum Mechanics
equivalent under translations Q' 3u^u
+ uqeQ',
UqtQcQ',
in Q', i.e., fi[u + ug) = a2(q, u)n{u),
uq £ Q C Q\
(3.7.12)
where a function a(q, u) is square /u-integrable and strictly positive almost everywhere on Q'. This function fulfils the relations a(0,u) = l,
a(q + q',u) = a(q,u)a(q',u + uq).
(3.7.13)
A measure on Q' obeying the condition (3.7.12) is called translationally quasi-invariant, but it does not remains equivalent under an arbitrary translation in Q', unless Q is finite-dimensional. Let a generating function Z of a cyclic strongly continuous unitary representation of the nuclear group T(Q) be the Fourier transform (3.7.6) of a translationally quasi-invariant measure /i on Q'. Then one can extend the representation (3.7.7) of this group to the unitary strongly continuous representation of the CCR group in the Hilbert space L^.(Q',/J,) by operators (3.2.7) in Example 3.2.2. These operators read Pz(q)p(u) = a(q, u)p(u + «,).
(3.7.14)
Remark 3.7.1. Let // (3.7.9) be a /u-equivalent positive measure of total mass 1 on Q'. The equality (i'(u + uq) = s~2(u)a2(q,u)s2(u + uq)n' (u) shows that it is also translationally quasi-invariant. Then the isomorphism (3.7.10) between representations Tz and Tz1 of the nuclear Abelian group T(Q) is extended to the isomorphism Pzr(q) = s~1Pz(q)s : p(u) i-> s~1(u)a(q,u)s(u + ug)p(u + uq) of the corresponding representations of the CCR group G(Q).
O
Similarly to the case of a finite-dimensional Lie group, any strongly continuous unitary representation Tz (3.7.7), Pz (3.7.14) of the nuclear CCR group G{Q) implies a representation of its Lie algebra Q{Q) by (unbounded)
243
Chapter 3 Algebraic Quantization
operators in the same Hilbert space L$.(Q',p,) [389]. This representation reads 1 = 1,
ir{q)p(u) = -i(6q+T)(q,u))p(u),
5gp(u) = lim a~1[p(u + auq) — p(u)],
(3.7.15)
a € M,
r)(q,u) = lim a " 1 [a(a,u) — 1],
(3.7.16)
a—»0
One derives at once from the relations (3.7.13) that SgSq' = Sq-6q,
5g(r){g',u)) = 6g,(r)(q,u)),
5q = -6-g, ij(0,u)=0,
6g({q',u)) = (q'\q), uGQ',
5qe = 0,
qeQ.
With the aid of these relations, it is easily justified that the operators (3.7.15) fulfil the Heisenberg CCR (3.7.4) - (3.7.5). The unitarity condition implies the conjugation rule (q,u)* = (q,u),
6* = -Sq - 2r)(q,u).
Hence, the operators (3.7.15) are Hermitian. Let us further restrict our consideration to representations with generating functions Z(q) such that I S i H Z(tq)
(3.7.17)
is an analytic function on R at t — 0 for all q £ Q. Then one can show that the function (q\u) on Q' is square /z-integrable for all q £ Q and that, consequently, the operators
(3.7.18)
(qi,u)---{qn,u)fi(u). The operators Tr(q) (3.7.15) act in the subspace E^ of all smooth complex functions in L^.(Q',fi) whose derivatives of any order also belongs to LciQ'iP)- However, Eoa need not be dense in the Hilbert space L^(Q',p,), unless Q is finite-dimensional. The space E^ is also the carrier space of a representation of the universal enveloping algebra Q{Q) of the CCR algebra
244
Geometric and Algebraic Topological Methods in Quantum Mechanics
Q{Q). The representations of G{Q) and Q{Q) in E^ need not be irreducible. Therefore, let us consider the subspace Eg = G(Q)0 of E^, where 6 is a cyclic vector for the representation of the CCR group in L^(Q',fi). Obviously, the representation of the CCR algebra Q(Q) in Eg is (algebraically) irreducible. If 6' is another cyclic vector in L^(Q',/x), the representations of G{Q) in Eg and Eg' are equivalent. One also introduces creation and annihilation operators ^(q)
= -T=0(g) T«r(g)] = ~7|[T^ T»?(«,«) + (<*•«)].
(3.7.19)
They obey the conjugation rule (a±(q))* = aT(q) and the commutation relations [o+ (q), a+ (q')} = [a~ (q),a~ (q1)} = 0.
[a-(q),a+(q')} = (q\q')l,
The particle number operator N in the carrier space Eg is defined by the conditions [N,a±(q)}=±a±(q) up to a summand Al. With respect to a countable orthonormal basis {%}, this operator TV is given by the sum N=^a+(qk)a-(qk),
(3.7.20)
k
but need not be denned everywhere in Eg, unless Q is finite-dimensional. Gaussian measures given by the Fourier transform 10.5.24 exemplify the physically relevant class of translationally quasi-invariant measures on the dual Q' of a nuclear space Q. Let fix denote a Gaussian measure on Q' whose Fourier transform is the generating function ZK=exp[-±BK(q)}
(3.7.21)
BK(q) = {K-xq\K-\),
(3.7.22)
with the covariance form
where K is a bounded invertible operator in the Hilbert completion Q of Q with respect to the Hermitian form (.|.). The Gaussian measure fix is translationally quasi-invariant, i.e., HK{U + uq) = a2K(q, u)fj.K(u).
Chapter 3 Algebraic Quantization
245
Using the formula (3.7.18), one can show that aK{q,u) = exp[-±BK(Sq) - ^(Sq,u)],
(3.7.23)
where 5 = KK* is a bounded Hermitian operator in Q. Let us construct the representation of the CCR algebra G(Q) determined by the generating function ZK (3.7.21). Substituting the function (3.7.23) into the formula (3.7.16), we find •n{
--{Sq,u).
Hence, the operators
*(q) = -i{8q-^(Sq,u)).
(3.7.24)
Accordingly, the creation and annihilation operators (3.7.19) read a±(g) = -~=[TSq ± \(Sq,u) + (q,u)}.
(3.7.25)
They act on the subspace Eg, 9 «Mi(. 1, of the Hilbert space L^Q' ,IIK), and they are Hermitian with respect to the Hermitian form (.\.)^K on
Remark 3.7.2. If a representation of the CCR is characterized by the Gaussian generating function (3.7.21), it is convenient for a computation to express all operators into the operators 5q and
K4>tf)] = '!?)• For instance, we have ir(q) = -i6q - -4>{Sq). The mean values {4>{qi) • • •
(3-7.26)
246
Geometric and Algebraic Topological Methods in Quantum Mechanics
where the sum runs through all partitions of the set 1,... ,n in ordered pairs (ii < 12), • • • (in-i < in), and where
imtw)) = (K-WK-W). D In particular, let us put K = \/2 • 1. Then the generating function (3.7.21) takes the form ZF(q)=exp[-^(q\q)},
(3.7.27)
and determines the Fock representation of the CCR algebra G(Q)- It is given by the operators
n(q) =
+
a {q) = -^[-5q+2(q,u)},
-i{6q-(q,u)), a,-{q) =-j=5q.
Its carrier space is the subspace Eg, 9 « w 1 of the Hilbert space L^(Q', /ip), where /XF denotes the Gaussian measure whose Fourier transform is (3.7.27). We agree to call it the Fock measure. The Fock representation up to an equivalence is characterized by the existence of a cyclic vector 9 such that a-(q)9 = 0,
qeQ.
(3.7.28)
For the representation in question, this is 9 « MF 1. An equivalent condition is that the particle number operator TV (3.7.20) exists and its spectrum is lower bounded. The corresponding eigenvector of N in E$ is 9 itself so that N9 = 0. Therefore, one often interprets this eigenvector as a vacuum state. A glance at the expression (3.7-25) shows that the condition (3.7.28) does not hold, unless ZK is Zp (3.7.27). For instance, the particle number operator in the representation (3.7.25) reads
TV" = XV( % >~(<7j) = Y}-6qiSqj + Si(qk,u)dqj + 3
3
(6km - ^S£,)< g f c l u>< 9 m ,u) - (Sjj - ^ ) ] , where {%} is the orthonormal basis for the pre-Hilbert space Q. One can show that this operator is denned everywhere on Ee and is lower bounded
247
Chapter 3 Algebraic Quantization
only if the operator S is a sum of the scalar operator 21 and a nuclear operator in Q, in particular, if
Tr(l - ±S) < oo. This condition is also sufficient for the measures ^K and fip (and, consequently, the corresponding representations) to be equivalent [164]. For instance, the generating function Zc(q)=exp[~(q\q)],
c2 / 1,
determines a non-Fock representation of the nuclear CCR. Since the Fock measure fip on Q' remains equivalent only under translations by vectors uq £ Q c Q', the measure Ha = /iF(u -
a£Q'\Q,
on Q' determines a non-Fock representation of the nuclear CCR. Indeed, this measure is translationally quasi-invariant: ^(u + Ug) =al(q,u)fia(u),
aa(q,u) =
aF(q,u-a),
and its Fourier transform Za(q)=exp[i(q,a)}ZF(q) is a positive-definite continuous function on Q. Then the corresponding representation of the CCR algebra is given by operators a+(q) = -j=(-6q + 2(q,u} - (q,*)),
a~(q) = - L ( J , + (q, a)). (3.7.29)
In comparison with all the above representations, these operators possess non-vanishing vacuum mean values (a±(q)e\e}^=^(q,a). If a G Q C Q', the representation (3.7.29) becomes equivalent to the Fock representation (3.7.25) due to the morphism p{u) H-> exp[-(q', u)]p(u + uq>).
248
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us note that the non-Fock representation (3.7.24) of the CCR algebra (3.7.4) - (3.7.5) in the Hilbert space L%.(Q',fj,K) is the Fock representation
(q,u),
irK(q) = niS-'q) = -i(*f - l(q,u)),
6? = 6s-lq,
of the CCR algebra {4>K(q)^K(q),I}, where [
I<8>I = I
and by the commutation relations (3.7.4) - (3.7.5), written with respect to the tensor product
w G.W,
the counit e(w) = 0, the antipode S(w) = —to, and the universal matrix R=1®1. It is a quasi-triangular cocommutative Hopf algebra. Now let us consider the quotient Qk,c of the tensor algebra ®W by the relations (3.7.4), (3.7.30) and the commutation relations
HqUW)} = -W^k^k-iy
< 3 - 7 - 31 )
where k and c are strictly positive real numbers. Due to the relation (3.7.30), the right-hand side of the relations (3.7.31) is well defined on ®W, and we have
Hq),4>{q')} = - W ^ f c l f c - ! ) 7 '
(3-7'32)
249
Chapter 3 Algebraic Quantization
Hence, Gk,c is the universal enveloping algebra of the Heisenberg CCR algebra Gk,c given by the commutation relations (3.7.4) and (3.7.32). This CCR algebra is modelled over the same nuclear space Q, but provided with the Hermitian form (qW)k,c = CktC(q\Q'),
( 3 - 7 - 33 )
Ck,c = ^ l - i y
The universal enveloping algebra Gk,c admits both the structure of the classical Hopf algebra UGk,c and the Hopf algebra UkGk,c, which differs from the classical one in the comultiplication law A (7) = 7 ® 1 + 1<8) 7, A(#?)) = 4>{q) ® kcI/2 + k~cI/2 ® 4>(q), A(n(q)) = *(q) ® kcI/2 + k~cI'2 ® 7r(g). One can think of UkGk,c as being a Hopf algebra of the /c-deformed CCR. It is readily observe that, if c = 1, the CCR algebras G{Q) and Gk,i coincide for any k, but the Hopf algebra UkGk,i differs from the classical one UGk,i = UG. If k = 1, then Gi,c = G{Q) and C/fc^i,c = W for any c. Since the Hopf algebra UkGk,c is the universal enveloping algebra of the CCR algebra Gk,c, its representations are determined in full by representations of Gk,c Comparing the commutation relations (3.7.5) and (3.7.32), one can show that, given a representation p of the CCR algebra G(Q), the CCR algebra Gk,c admits a representation pk,c by the operators pkA4>(q)) = p(0(«)).
pk,c(*(q)) = p(*{Ck,cg)),
PkAi) = PW =!.
where Ck,c is given by the expression (3.7.33). For instance, if p is the Fock representation of the CCR algebra G(Q), the representation pkiC is not equivalent to the Fock representation of the CCR algebra Gk,c, unless Q is finite-dimensional. 3.8
Automorphisms of quantum systems
Let us consider uniformly and strongly continuous one-parameter groups of automorphisms of C*-algebras. In particular, they characterize evolution of quantum systems. Forthcoming Remarks 3.8.1 and 3.8.2 explain why we restrict our consideration mainly to these automorphism groups.
250
Geometric and Algebraic Topological Methods in Quantum Mechanics
Remark 3.8.1. Let V be a Banach space and B(V) the set of bounded endomorphisms of V. The normed, strong and weak operator topologies on B(V) are defined in the same manner as in Section 3.ID. Automorphisms of a C*-algebra are obviously its isometries as a Banach space. Any weakly continuous one-parameter group of endomorphism of a Banach space is also strongly continuous and their weak and strong generators coincide with each other [63]. • Remark 3.8.2. There is the following relation between morphisms of a C*-algebra A and the set E{A) of its states which is a convex subset of A'. A linear morphism 7 of a C*-algebra A as a vector space is called the Jordan morphism if the relations -y(ab + ba) =7(0)7(6) +7(6)7(0),
a,b e A.
hold. One can show the following [144]. Let 7 be a Jordan automorphism of a unital C*-algebra A. It yields the dual weakly* continuous affine bijection 7' of E(A) onto itself, i.e., 7 '(A/
+ (1 - A)/') = A 7 '(/) + (1 - A)7'( A f,f',£E(A), Ae[0,l]. Conversely, any such a map of E(A) is the dual to some Jordan automorphism of A. However, we are not concerned with groups of Jordan automorphisms because of the following fact. If G is a connected group of weakly continuous Jordan automorphisms of a unital C*-algebra A which is provided with a weak operator topology, then it is a weakly continuous group of automorphisms of A. D One says that a one-parameter group G(K) is a uniformly (resp. strongly) continuous group of automorphisms of a C*-algebra A if it is a range of a continuous map of M. to the group Aut (A) of automorphisms of A which is provided with the normed (resp. strong) operator topology and whose action on A is separately continuous. The problem is that, if a curve G(M.) in Aut (A) is continuous with respect to the normed operator topology, then the curve G(M)(a) for any a £ A is continuous in the C*algebra A, but the converse is not true. At the same time, a curve G(R) is continuous in Aut (A) with respect to the strong operator topology if and only if the curve G(M)(a) for any a £ A is continuous in A. By this reason, strongly continuous one-parameter groups of automorphisms of C*-algebras
Chapter 3 Algebraic Quantization
251
are most interesting. However, the infinitesimal generator of such a group fails to be bounded, unless this group is uniformly continuous. Remark 3.8.3. If G(R) is a strongly continuous one-parameter group of automorphisms of a C*-algebra A, there are the following continuous maps [63]: • R 3 t n (Gt(a), / ) G C is continuous for all a G A and / G A'; • A 3 a —» Gt(a) G A is continuous for all t G M; • R 3 t i - » Gt(a) G A is continuous for all o G A. • Let A be a C*-algebra. Without loss of generality, we will assume that A is a unital algebra. The space of derivations of A is provided with the involution u >—> u* defined by the equality 5*{a) := -6(a*)*,
a £ A.
(3.8.1)
Throughout this Section, by a derivation S of A is meant an (unbounded) symmetric derivation of A (i.e., S(a*) = S(a)*, a £ A) which is defined on a dense involutive subalgebra D(S) of A. If a derivation 6 on D(S) is bounded, it is extended to a bounded derivation everywhere on A. Conversely, every derivation defined everywhere on a C*-algebra is bounded [129]. For instance, any inner derivation S(a) = i[b,a], where b is a Hermitian element of A, is bounded. There is the following relation between bounded derivations of a C*-algebra A and one-parameter groups of automorphisms of A [63]. THEOREM 3.8.1. Let 5 be a derivation of a C*-algebra A. The following assertions are equivalent: • S is defined everywhere and, consequently, is bounded; • S is the infinitesimal generator of a uniformly continuous oneparameter group [Gt] of automorphisms of the C*-algebra A. Furthermore, for any representation n of A in a Hilbert space E, there exists a bounded self-adjoint operator H G it {A)" in E and the uniformly continuous representation n(Gt) = exp(-itH),
t £ R,
(3.8.2)
252
Geometric and Algebraic Topological Methods in Quantum Mechanics
of the group [Gt] in E such that 7r(<5(a)) = -i[H,7r(o)], itH
a £ A, itn
7r(Gt(a)) = e- n{a)e ,
t e R.
(3.8.3) (3.8.4)
a A C*-algebra need not admit non-zero bounded derivations. For instance, no commutative C*-algebra possesses bounded derivations. The following is the relation between (unbounded) derivations of a C*-algebra A and strongly continuous one-parameter groups of automorphisms of A [62; 361]. THEOREM 3.8.2. Let 5 be a closable derivation of a C*-algebra A. Its closure 5 is an infinitesimal generator of a strongly continuous one-parameter group of automorphisms of A if and only if (i) the set (1 + XS)(D(5) for any A G R \ {0} is dense in A, • (ii) ||(1 + A<5)(a)|| > ||a|| for any A 6 M and any a € A.
It should be noted that, if A is a unital algebra and 5 is its closable derivation, then 1 G D(5). Let us mention a more convenient sufficient condition of a derivation of a C*-algebra to be an infinitesimal generator of a strongly continuous one-parameter group of its automorphisms. A derivation S of a C*-algebra A is called well-behaved if, for each element a G D(5), there exists a state / of A such that /(o) = ||a||,
/(*(*)) = 0.
If 6 is a well-behaved derivation, it is closable [247], and obeys the condition (ii) in Theorem 3.8.2 [62; 361]. Then we come to the following. 3.8.3. If 5 is a well-behaved derivation of a C*-algebra A and obeys the condition (i) in Theorem 3.8.2, its closure 5 is an infinitesimal generator of a strongly continuous one-parameter group of automorphisms of A. D PROPOSITION
For instance, a derivation 5 is well-behaved if it is approximately inner, i.e., there exists a sequence of self-adjoint elements {bn} in A such that 6(a) = limi[6 n ,o], n
a & A.
Chapter 3 Algebraic Quantization
253
In contrast with the case of a uniformly continuous one-parameter group of automorphisms of a C*-algebra A, a representation of A does not imply necessarily a unitary representation (3.8.2) of a strongly continuous oneparameter group of automorphisms of A, unless the following. PROPOSITION 3.8.4. Let Gt be a strongly continuous one-parameter group of automorphisms of a C*-algebra A and S its infinitesimal generator. Let A admit a state / such that
|/(<5(a))| < A[/(a*a) + f{aa*)\1'2
(3.8.5)
for all a £ A and a positive number A, and let (it/,Of) be a cyclic representation of A in a Hilbert space Ef determined by / . Then there exist a self-adjoint operator H o n a domain D{H) C AOj in Ej and a strongly continuous unitary representation (3.8.2) of Gt in Ef which fulfils the relations (3.8.3) - (3.8.4) for 7r = vr/. • Let us note that the condition (3.8.5) in Theorem 3.8.4 is sufficient in order that the derivation 5 to be closable [247]. There is a general problem of a unitary representation of an automorphism group of a C*-algebra. For instance, let B{E) be the C*-algebra of bounded operators in a Hilbert space E. All its automorphisms are inner. Any (unitary) automorphism U of a Hilbert space E yields the inner automorphism a-+UaU-\
a£B(E),
(3.8.6)
of B{E). Herewith, the automorphism (3.8.6) is the identity if and only if U = Al, |A| = 1, is a scalar operator in E. It follows that the group of automorphisms of B(E) is the quotient PU(E) = U(E)/U(1),
(3.8.7)
called the projective unitary group of the unitary group U(E) with respect to the circle subgroup U(l). Therefore, given a group G of automorphisms of the C*-algebra B(E), the representatives Ug in U(E) of elements g £ G constitute a group up to phase multipliers, i.e., UgUg. = exp[ia(g,g')}Ugg/,
a(g,g') £ K.
Nevertheless, if G is a one-parameter weakly* continuous group of automorphisms of B(E) whose infinitesimal generator is a bounded derivation
254
Geometric and Algebraic Topological Methods in Quantum Mechanics
of B(E), one can choose the multipliers exp[ia(g,g')} = 1. Representations of groups by unitary operators up to phase multipliers are called projective representations [91; 429]. In a general setting, let G be a group and A a commutative algebra. An A-multiplier of G is a map £ : G x G —> A such that £(1G,3) = £(S,1G) = :U,
geG,
^(91,9293)^(92,93) = ^(91,92)^(9192,93),
9i G G.
For instance, £:GxG->lA£A is a multiplier. Two A-multipliers £ and £' are said to be equivalent if there exists a map f : G —> A such that
An ,4-multiplier is called exact if it is equivalent to the multiplier £ = 1A. The set of ^.-multipliers is an Abelian group with respect to the pointwise multiplication, and the set of exact multipliers is its subgroup. Let HM(G, A) be the corresponding factor group. If G is a locally compact topological group and A a Hausdorff topological algebra, one additionally requires that multipliers £ and equivalence maps / are measurable maps. In this case, there is a natural topology on HM(G, A) which is locally quasi-compact, but need not be Hausdorff [321]. PROPOSITION 3.8.5. [91]. Let G be a simply connected locally compact Lie group. Each f/(l)-multiplier £ of G is brought into the form £ = exp ia, where a is an R-multiplier. Moreover, £ is exact if and only if a is well. Any R-multiplier of G is equivalent to a smooth one. •
Let G be a locally compact group of strongly continuous automorphisms of a C*-algebra A. Let M(A) denote the multiplier algebra of A, i.e., the largest C*-algebra containing A as an essential ideal, i.e., if a € M(A) and ab = 0 for all b G A, then a — 0 [441]). For instance, M(A) = A if A is a unital algebra. Let £ be a multiplier of G with values in the center of M(A). A G-covariant representation n of A [132; 336] is a representation 7T of A (and, consequently, M(A)) in a Hilbert space E together with a
Chapter 3 Algebraic Quantization
255
projective representation of G by unitary operators U(g), g G G, in E such that <9{a)) - U(g)n(a)U*(g),
U(g)U(g') = n(t{g, g'))U(gg').
Chapter 4
Geometry of algebraic quantization
Algebraic quantum theory usually deals with Hilbert spaces. This Chapter addresses quantum models involving Hilbert manifolds and Hilbert bundles, but we restrict our consideration to particular Hilbert bundles over smooth finite-dimensional manifolds. For instance, this is the case of timedependent quantum systems and quantum models depending on classical parameters. Since a Hilbert space E is a real Banach space and the unitary group U(E) is a Banach-Lie group, we in fact deal with Banach manifolds and bundles. Their differential geometry is similar to differential geometry of finite-dimensional manifolds and bundles in main. In particular, the inverse mapping theorem and the Frobenius theorem hold. For instance, this is not the case of Frechet manifolds. We refer the reader to [262] for the wider class of infinite-dimensional smooth manifold and principal bundles modelled on so called convenient locally convex topological vector spaces. 4.1
Banach and Hilbert manifolds
We start with the notion of a real Banach manifold [273; 422]. Banach manifolds are defined similarly to finite-dimensional smooth manifolds, but they are modelled on Banach spaces, not necessarily finite-dimensional. Passing later to complex Hilbert manifolds, we also refer the reader to Section 2.6 for analogy tofinite-dimensionalHermitian and Kahler manifolds. Let us recall some particular properties of (infinite-dimensional) real Banach spaces (see Section 10.5A for a general case of topological vector spaces). Let us note that afinite-dimensionalBanach space is always provided with an Euclidean norm. • Given Banach spaces E and H, every continuous bijective linear map 257
258
Geometric and Algebraic Topological Methods in Quantum Mechanics
of E to H is an isomorphism of topological vector spaces. • Given a Banach space E, let F be its closed subspace. One says that F splits in E if there exists a closed complement F' of F such that E = F®F'. In particular, finite-dimensional andfinite-codimensionalsubspaces split in E. As a consequence, any subspace of a finite-dimensional space splits. • Let E and H be Banach spaces and / : E —> H a continuous injection. One says that / splits if there exists an isomorphism g:H-*H1xH2 such that g o f yields an isomorphism of E onto Hi x {0}. • Given Banach spaces (E, \\.\\E) and (H, ||.||ff), one can provide the set Hom°(£, H) of continuous linear morphisms of E to H with the norm ||/||:=
sup \\f(z)\\H,
f€Rom°(E,H).
(4.1.1)
||*|| E =i
In particular, the norm (10.5.1) on the topological dual E' of E is of this type. If E, H and F are Banach spaces, the bilinear map Hom°(£,F) x Horn °(F, H) -» Horn °(E,H), obtained by the composition / o g of morphisms 7 € Horn (E, F) and / £ Hom°(jF, H), is continuous. Let us note that this assertion is false for more general spaces, e.g., the Frechet ones. • Let (E, \\.\\E) and (H, \\.\\H) be real Banach spaces. One says that a continuous map / : E —> H (not necessarily linear and isometric) is a differentiable function between E and H if, given a point z G E, there exists an R-linear continuous map df(z)
:E^H
(not necessarily isometric) such that
IV) = /M + <M(')V -») + o(z' - z),
lim
y-«)a. = O i
for any 2' in some open neighborhood U of z. For instance, any continuous linear morphism / of E to H is differentiable and df(z)z = f{z). The linear map d/(z) is called a differential of / at a point z £ U. Given an element v £ E, we obtain the map E 3 z >-+dvf(z) = df(z)v € H,
(4.1.2)
Chapter 4 Geometry of Algebraic Quantization
259
called the derivative of a function / along a vector v G E. One says that / is two-times differentiable if the map (4.1.2) is differentiable for any v G E. Similarly, r-times differentiable and infinitely differentiable (smooth) functions on a Banach space are denned. The composition of smooth maps is a smooth map. The following inverse mapping theorem enables one to consider smooth Banach manifolds and bundles similarly to the finite-dimensional ones. 4.1.1. Let / : E —> H be a smooth map such that, given a point z € E, the differential df (z) : E —> H is an isomorphism of topological vector spaces. Then / is a local isomorphism at z. • THEOREM
Let us turn to the notion of a Banach manifold, without repeating the statements true both for finite-dimensional and Banach manifolds. 4.1.2. A Banach manifold B modelled on a Banach space B is defined as a topological space which admits an atlas of charts V&e = {(Ui,4>i)}, where the maps 4>L are homeomorphisms of UL onto open subsets of the Banach space B, while the transition functions (j>^(j)^1 from >t(C/t n U{) C B to 4>^{Utr\U^) C B are smooth. Two atlases of a Banach manifold are said to be equivalent if their union is also an atlas. • DEFINITION
Unless otherwise stated, Banach manifolds are assumed to be connected paracompact Hausdorff topological spaces. A locally compact Banach manifold is necessarily finite-dimensional. Remark 4.1.1. Let us note that a paracompact Banach manifold admits a smooth partition of unity if and only if its model Banach space does. For instance, this is the case of (real) separable Hilbert spaces. Therefore, we restrict our consideration to Hilbert manifolds modelled on separable Hilbert spaces. • Any open subset U of a Banach manifold B is a Banach manifold whose atlas is the restriction of an atlas of B to U. Morphisms of Banach manifolds are defined similarly to those of smooth finite-dimensional manifolds. However, the notion of the immersion and submersion need a certain modification (see Definition 4.1.3 below). Tangent vectors to a smooth Banach manifold B are introduced by analogy with tangent vectors to a finite-dimensional one. Given a point Z G B , let us consider the pair (v; (Ut, >,,)) of a vector v £ B and a chart (£/t 9 z, 4>t)
260
Geometric and Algebraic Topological Methods in Quantum Mechanics
on a Banach manifold B. Two pairs (v; {Ut,4>L)) and {v'\ (f/ c ,0 c )) are said to be equivalent if
v'= dfc^XM*))*-
(4-1-3)
The equivalence classes of such pairs make up the tangent space TZB to a Banach manifold B at a point z e B. This tangent space is isomorphic to the topological vector space B. Tangent spaces to a Banach manifold B are assembled into the tangent bundle TB of B. It is a Banach manifold modelled over the Banach space B® B which possesses the transition functions
(^rSdo^:1))Any morphism / : B —> B' of Banach manifolds yields the corresponding tangent morphism of the tangent bundles Tf : TB —> TB'. 4.1.3. Let / : B —> B' be a morphism of Banach manifolds. (i) It is called an immersion at a point z £ B if the tangent morphism Tf at z is injective and splits. (ii) A morphism / is called a submersion at a point z £ B if Tf at z is surjective and its kernel splits. • DEFINITION
In the case of finite-dimensional smooth manifolds, the split conditions are superfluous, and Definition 4.1.3 recovers the notion of the immersion and submersion in Section 10.6A. The range of a surjective submersion / of a Banach manifold is a submanifold, though / need not be an isomorphism onto a submanifold, unless / is an imbedding. One can think of a surjective submersion n : B —» B' of Banach manifolds as a fibred Banach manifold. For instance, the product B x B' of Banach manifolds is a fibred Banach manifold with respect to prx and pr2. Let B be a Banach manifold and E a Banach space. The definition of a (locally trivial) vector bundle with the typical fibre E and the base B is a repetition of that of finite-dimensional smooth vector bundles. Such a vector bundle Y is a Banach manifold and Y —> B is a surjective submersion. The above mentioned tangent bundle TB of a Banach manifold exemplifies a vector bundle over B. Let Bnh be the category of Banach spaces and Vect(S) denotes the category of vector bundles over a Banach manifold B with respect to their
Chapter 4 Geometry of Algebraic Quantization
261
morphisms over idB. Let F : Bnh x Bnh -> Bnh
(4.1.4)
be a functor of two variables which is covariant in the first and contravariant in the second. Then there exists a functor VF : Vect(tf) x Vect(B) -> Vect(£) such that, if YE —> B and YR —» B are vector bundles with the typical fibres E and H, then VF(YE, YH) is a vector bundle with the typical fibre F(E,H). For instance, the Whitney sum, the tensor product, and the exterior product of vector bundles over a Banach manifold are defined in this way. In particular, since the topological dual E' of a Banach space E is a Banach space with respect to the norm (10.5.1), one can associate to each vector bundle YE —* B the dual Yg = YE> with the typical fibre E'. For instance, the dual of the tangent bundle TB of a Banach manifold B is the cotangent bundle T*B. Sections of the tangent bundle TB —> B of a Banach manifold are called vectorfieldson a Banach manifold B. They form a locally free module T\{B) over the ring C°° (B) of smooth real functions on B. Every vector field •# on a Banach manifold B determines a derivation of the R-ring C°°(B) by the formula f{z) -> dtf(z) = df(z)ti(z),
z G B.
Different vector fields yield different derivations. It follows that T\ (B) possesses a structure of a real Lie algebra, and there is its monomorphism T1(B)^-0Coo(B)
(4.1.5)
to the derivation module of the R-ring C°°(B). Let us consider the Chevalley-Eilenberg complex of the real Lie algebra 71 (S) with coefficients in C°°(B) and its subcomplex O*[TX{B)\ of C°°(S)multilinear skew-symmetric maps by analogy with the complex C*[0^4] in Section 1.6. This subcomplex is a differential calculus over a R-ring C°°(B) where the Chevalley-Eilenberg coboundary operator d (1.6.6) and the prod-
262
Geometric and Algebraic Topological Methods in Quantum Mechanics
uct (1.6.9) read r
# ( 0 O , . . . , 0 r ) = ^ ( - 1 ) ^ Wtfo,.-., £ , . . . ,
tfr))+
(4.1.6)
i=0
X J ( - l ) i + J > ( [ t f i , ^ l . ^ o , . . . , * , . . . , 0 , , . . . , fc ), 0A^'(i?i,...,i? P+a )=
(4.1.7)
*l<-"
0 G Or[7i(B)],
0'GO'[TI(B)],
^GTi(B).
There are the familiar relations *\
= d*f,
feC°°(B),
0G7I(B),
d(4> A 4>') = d4>A<j>' + (-1)1*10 Ad<j>\
The differential calculus O*[Ti(B)] contains the following subcomplex. Let O1(B) be the C°°(B)-module of global sections of the cotangent bundle T*B of B. Obviously, there is its monomorphism O1{B)-^-QC°°{B)*
(4.1.8)
to the dual of the derivation module T)C°°(B). Furthermore, let f\T*B be the r-degree exterior product of the cotangent bundle T*B and Or{B) the C°°(S)-module of its sections. Let O*(B) be the direct sum of C°°(B)modules Or{B), r £ N, where we put O°(B) := C°°(B). Elements of O*(B) are obviously C°°(S)-multilinear skew-symmetric maps of T\(B) to C°°(B). Therefore, the Chevalley-Eilenberg differential d (4.1.6) and the exterior product (4.1.7) of elements of O*(B) are well denned. Moreover, one can show that dcp and <j> h<j>',
(4.1.9)
It follows that the differential calculi 0*[7i(B)], O*{B) and Ol[DC°°{B)\ over the M-ring C°°(B) are not mutually isomorphic in general. However, it is readily observed that the minimal differential calculi in O* [7i (B)] and
263
Chapter 4 Geometry of Algebraic Quantization
O*(B) coincide with the minimal Chevalley-Eilenberg differential calculus O*CC°(B) over the R-ring C°°(B) because they are generated by the elements df, f G C°°{B), where d is the restriction (4.1.6) to 71 (B) of the Chevalley-Eilenberg coboundary operator (1.6.6). A connection on a Banach manifold B is defined as a connection on the C°°(B)-module 71 (B) [422]. In accordance with Definition 1.3.2, it is an R-module morphism V : 7i(B) -• O1COO(B) ®Ti(B), which obeys the Leibniz rule V ( / 0 ) = # ® 0 + /V(tf),
f£C°°(B),
0€7i(B).
(4.1.10)
In view of the inclusions, OlC°°{B) c O\B) c TX{B)*,
Ti(B) C 71(B)** c
O\B)*,
it is however convenient to define a connection on a Banach manifold as an R-module morphism V:r 1 (B)-»O 1 (B)®Ti(B),
(4.1.11)
which obeys the Leibniz rule (4.1.10). Let us turn now to Hilbert manifolds. These are particular Banach manifolds modelled on complex Hilbert spaces, which are assumed to be separable (see Remark 4.1.1). An important case of Hilbert manifolds are infinite-dimensional Kahler manifolds [105; 327] (see Section 2.6 for geometry of finite-dimensional Kahler manifolds). Remark 4.1.2. We refer the reader to [273] for the theory of real Hilbert and (infinite-dimensional) Riemannian manifolds. A real Hilbert manifold is a Banach manifold B modelled on a real Hilbert space V. It is assumed to be connected Hausdorff and paracompact space admitting the partition of unity by smooth functions (this is the case of a separable V). A Riemannian metric on B is defined as a smooth section g of the tensor bundle V2T*B such that g(z) is a positive non-degenerate continuous bilinear form on the tangent space TZB. This form yields two maps TZB —» T*B. It is said to be non-degenerate if these maps are continuous isomorphisms. In infinite-dimensional geometry, the most of local results follow from general arguments analogous to those in the finite-dimensional case. In particular, a Riemannian metric makes B into a metric space. As
264
Geometric and Algebraic Topological Methods in Quantum Mechanics
in the finite-dimensional case, B admits a unique Levi-Civita connection. The global theory of real Hilbert manifolds is more intricate. For instance, an infinite-dimensional real (and, consequently, complex) Hilbert space V is diffeomorphic to V \ {0}, and the unit sphere in V is a deformation retract of V [36]. • A complex Hilbert space (E, {•].)) can be seen as a real Hilbert space E 3 v i-> vR e
CUR,«R)
ER,
=Re(v\v'),
in Remark 3.1.7 equipped with the complex structure JUR = (W)R. We have {Jvu, Jv^) = (vu,vR),
(JVR,I4)
= Im(vR,vR).
Let E
(4.1.12)
10
E ' = {vu- Uvm : vu G Eu}, which are mutually orthogonal with respect to the Hermitian form (.|.)cSince {vR - iJvK)\v'R - Uv^) = 2(v\v'),
{vR + iJMvii + iJvn) = 2 ( w » ,
there are the following linear and antilinear isometric bijections EBv^vu-*
-/=(VR
-
IJVR)
e E1'0,
They make ^ l l 0 and E0'1 isomorphic to the Hilbert space E and the dual Hilbert space E", respectively. Hence, the decomposition (4.1.12) takes the form (4.1.13)
EC = E®~E. The complex structure J on the direct sum (4.1.13) reads J : E © E 3 i ) + u i - > iv-iu€
E®E~,
(4.1.14)
Chapter 4 Geometry of Algebraic Quantization
265
where E and E are the (holomorphic and antiholomorphic) eigenspaces of J characterized by the eigenvalues i and —i, respectively. Let / be a function (not necessarily linear) from a Hilbert space E to a Hilbert space H. It is said to be differentiable if the corresponding function /K between the real Banach spaces J5R and H^ is differentiable. Let dfR(z), z G £7R, be the differential (4.1.2) of /R on £ R which is a continuous linear morphism EM 3 vu H-> dfR(z)vu
E Hu
between real topological vector spaces _ER and H$L- This morphism is naturally extended to the C-linear morphism £ C 3 B C H dfa(z)vc G Hc
(4.1.15)
between the complexifications of ER and HR- In view of the decomposition (4.1.13), one can introduce the C-linear maps
dfo(z)(v + u) := dfa{z)v,
df(z){v + u) := dfR{z)u
from E © E to .He such that
<#kO)«c = dfu(z)(v + u) = dfR(z)v + dfR(z)u. Let us split
fu(z) = f(z)+J(Z) in accordance with the decomposition He = H ® H. Then the morphism (4.1.15) takes the form dfa{z)(v + u) = df(z)v + df{z)u + d]{z)v + dj{z)u,
(4.1.16)
where df = df, d f = df. A function / : E —> H is said to be holomorphic (resp. antiholomorphic) if it is differentiable and df(z) = 0 (resp. df(z) = 0) for all z & E. A holomorphic function is smooth, and is given by the Taylor series. If / is a holomorphic function, then the morphism (4.1.16) is split into the sum
dfa(z)(v + u)= df(z)v + dj(z)u of morphisms E —> H and E~ —>H~. Example 4.1.3. Let / be a complex function on a Hilbert space E. Then /R = (Re/,Im/)
266
Geometric and Algebraic Topological Methods in Quantum Mechanics
is a map of E to M2. The differential dfa{z), z € E, of / R yields the complex linear morphism E®T5 3vc^
(dRef(z)vc,dLmf(z)vc)
>-> d(Re/ + ilm/)(z)v c e C ,
which is regarded as a differential df(z) of a complex function / on a Hilbert space E. O A Hilbert manifold V modelled on a Hilbert space E is denned as a real Banach manifold modelled on the Banach space Em. which admits an atlas {(Ui, (pi)} with holomorphic transition functions (j>$<j>~1. Let CTV denote the complexified tangent bundle of a Hilbert manifold V. In view of the decomposition (4.1.13), each fibre CTZV, z e V, of CTV is split into the direct sum CTZV = TZV®TZV of subspaces TZV and TZV, which are topological complex vector spaces isomorphic to the Hilbert space E and the dual Hilbert space E, respectively. The spaces CTZV, TZV and TZV are respectively called the complex, holomorphic and antiholomorphic tangent spaces to a Hilbert manifold V at a point z £ V. Since transition functions of a Hilbert manifold are holomorphic, the complex tangent bundle CTV is split into a sum CTV = TV®TV of holomorphic and antiholomorphic subbundles, together with the antilinear bundle automorphism TV®TV3v + u^>v + ueTV®TV and the complex structure J : TV ®TV 3 v + u H^ iv - iu € TV © TV.
(4.1.17)
Sections of the complex tangent bundle CTV —» V are called complex vector fields on a Hilbert manifold V. They constitute the locally free module CT\{V) over the ring C°°(V) of smooth complex functions on V. Every complex vector field d + v on V yields a derivation
f(z)^df(z)(0 + v) = df(z)4(z) + df(z)v(z),
feC°°(V), z&V,
oftheC-ring
267
Chapter 4 Geometry of Algebraic Quantization
complex vector spaces isomorphic to E © E. Since Hilbert spaces are reflexive, the complex tangent bundle CTV is the dual of CT*V. The complex cotangent bundle CT*V is split into the sum CT*V = T*V © T*V
(4.1.18)
of holomorphic and antiholomorphic subbundles, which are the annihilators of antiholomorphic and holomorphic tangent bundles TV and TV, respectively. Accordingly, CT*V is provided with the complex structure J via the relation (v, Jw) = (Jv,w),
v G CTZV,
w G CT^V,
zeV.
Sections of the complex cotangent bundle CT*V —> V constitute a locally free C°°(-p)-module Ol{V). It is the C°°(P)-dual O1(V) = CTi(Vy
(4.1.19)
of the module CTi(V) of complex vector fields on V, and vice versa. Similarly to the case of a Banach manifold, let us consider the differential calculi O*[TiCP)], 0*{V) (further denoted by C*{V)) and O1[5COC(P)] over the C-ring C 0 0 ^ ) . Due to the isomorphism (4.1.19), 0*{Ti(V)] is isomorphic to C*(V), whose elements are called complex exterior forms on a Hilbert manifold V. The exterior differential d on these forms is the Chevalley-Eilenberg coboundary operator k
d
(4.1.20)
t=0
x ; ( - i ) i + W i . ^ - ] . 0 o , . . . , & , . . . , £ , • , . . . ,,>fc),
^ G cTi(p).
In view of the splitting (4.1.18), the differential graded algebra C*(V) admits the decomposition C*{V) = © Cp'"(V) p,q=o
pq
into subspaces C ' {V) of p-holomorphic and q-antiholomorphic forms. Accordingly, the exterior differential d on C* (T7) is split into a sum d = d + 9 of holomorphic and antiholomorphic differentials d : Cp'q(V) -> Cp+ll9(73),
9 : Cp'q{V) -> C P ' 9+1 (P),
268
Geometric and Algebraic Topological Methods in Quantum Mechanics
A Hermitian metric on a Hilbert manifold V is denned as a complex bilinear form g on fibres of the complex tangent bundle CTV which obeys the following conditions: • g is a smooth section of the tensor bundle CT*V
WM
= g@zX),
g(J#z, J#'z) = 9(W*)-
A Hermitian metric exists, e.g., on paracompact Hilbert manifolds modelled on separable Hilbert spaces. The above mentioned properties of a Hermitian metric on a Hilbert manifold are similar to properties of a Hermitian metric on a finite-dimensional complex manifold in Section 2.6. Therefore, one can think of the pair (V, g) as being an infinite-dimensional Hermitian manifold. A Hermitian manifold (V, g) is endowed with a non-degenerate exterior two-form tyW,)
= 9(J#z,K),
tiz,ti'z£CTzV,
z€V,
(4.1.21)
called the fundamental form of the Hermitian metric g. This form satisfies the relations
WM
= to@*X), £l(Jz#z,Jz#'z)=n(#z^'z)-
If ft (4.1.21) is a closed (i.e., symplectic) form, the Hermitian metric g is called a Kdhler metric and 0. a Kdhler form. Accordingly, (V, g, Q) is said to be an infinite-dimensional Kdhler manifold. A Kahler metric g and its Kahler form ( l o a a Hilbert manifold V yield the bundle isomorphisms £ : CTV 9 ^ 1 / j f l e CT*V, nb : CTV 3v»-v\Q,£ CT*V. Let g" and fi" denote the inverse bundle isomorphisms CT*V —> CTV.
269
Chapter 4 Geometry of Algebraic Quantization
They possess the properties fi" = Jg*,
fi«(«;,).K = -n J K)jw z , wz,w'z G cr*p. In particular, every smooth complex function / € C 0 0 ^ ) on the Kahler manifold (V,g) determines: • the complex vector field 9Hdf) = g\df) + g*(df),
(4.1.22)
which is split into holomorphic and antiholomorphic parts g^(df) and
gHdf);
• the complex Hamiltonian vector field n«(#) = J(gHdf)) = -ig*(df) + igHdf);
(4.1.23)
• the Poisson bracket {/,/'} =
fi«(d/)J4f',
/ , / ' e C°°(P).
(4.1.24)
By analogy with the case of a Banach manifold, we modify Definition 1.3.2 and define a connection V on a Hilbert manifold V as a C-module morphism V : C7I0P) -» C\V) ® CTi(P), which obeys the Leibniz rule V(/0) = # ® 0 + /V(i9),
fsC°°(V),
tiGCT^P).
Similarly, a connection is introduced on any C°°(P)-module, e.g., on sections of tensor bundles over a Hilbert manifold V. Let D and D denote the holomorphic and antiholomorphic parts of V, and let V^ = #J V, D$ and D-g be the corresponding covariant derivatives along a complex vector field 1? on V. For any complex vector field i? = v + v on 7>, we have the relations A? = V,,,
A , = Vu,
£>j,9 = i£>0,
Dj^ = -iDtf.
There is the following generalization of Theorem 2.6.8 to infinitedimensional Kahler manifolds.
270
Geometric and Algebraic Topological Methods in Quantum Mechanics
PROPOSITION 4.1.4. Given a Kahler manifold (P,g), there always exists a metric connection on V such that Vg = 0,
Vfi = 0,
V J = 0,
where J is regarded as a section of the tensor bundle CT*V ® CTV-
•
Example 4.1.4. If V = E is a Hilbert space, then CTV = Ex ( £ © £ ) . A Hermitian form (.|.) on E defines the constant Hermitian metric g : (E © E) x (E © E) -> C, 5(0,0') = H u ' } +
0 = v + u, tf' = T / + <
(4.1.25)
onV = E. The associated fundamental form (4.1.21) reads fi(#,tf')= j( v | u ') _ i( w '|u).
(4.1.26)
It is constant on E. Therefore, dO = 0 and g (4.1.25) is a Kahler metric. • The metric connection on E is trivial, i.e., V = d, D — d, D = d. Example 4.1.5. Given a Hilbert space E, a projective Hilbert space PE is made up by the complex one-dimensional subspaces (i.e., complex rays) of E. This is a Hilbert manifold with the following standard atlas. For any non-zero element x £ E, let us denote by x a point of PE such that x G x. Then each normalized element h £ E, \\h\\ = 1, defines a chart (Uh,4>h) of the projective Hilbert space PE such that 0h(x) = T ^ T - h. (4.1.27) (x\h) The image of Uh in the Hilbert space E is the one-codimensional closed (Hilbert) subspace Uh = {x£PE
: (x\h)^0},
Eh = {z£E
: {z\h)=0},
(4.1.28)
where z(x) + /i G x. In particular, given a point x £ PE, one can choose the centered chart Eh, h £ x, such that 4>h(x) = 0. Hilbert spaces Eh and Eh> associated to different charts Uh and Uh' are isomorphic. The transition function between them is a holomorphic function (4.1.29)
271
Chapter 4 Geometry of Algebraic Quantization
from (f>h(Uh nUh>) C Eh to <j>h>{Uh r\Uh>) C Eh>. The set of the charts
{(Uh,
v' = -^[(x\h)v-x{v\h)].
The projective Hilbert space PE is homeomorphic to the quotient of the unitary group U(E) equipped with the normed operator topology by the stabilizer of a ray of E. It is connected and simply connected [104]. The projective Hilbert space PE admits a unique Hermitian metric g such that the corresponding distance function on PE is p(x,x') = v/2arccos(|(a;|3;/)|),
(4.1.31)
where x,x' are normalized elements of E. It is a Kahler metric, called the Fubini-Studi metric. Given a coordinate chart (Uh,4>h), this metric reads _ (A A \ fol|"2) + fal"l) 9FS{VUM2) = 1 + ||Z||2
(Z\u2)(vi\z} + (z\Ul){V2\z) (1 + INI2)2 '
(4.1.32) for any complex tangent vectors $i = V\ + u\ and ^2 = V2+U2 in CTZPE. The corresponding Kahler form is given by the expression
nFS(#M
=r
fTpp
i
(1+ N | 2 ) 2
• (4-1-33)
It is readily justified that the expressions (4.1.32) - (4.1.33) are preserved under the transition functions (4.1.29) - (4.1.30). Written in the coordinate chart centered at a point z(x) = 0, these expressions come to the expressions (4.1.25) and (4.1.26), respectively. •
4.2
Dequantization
The standard approach to canonical quantization of classical Hamiltonian systems implies the assignment of Hermitian operators / to smooth functions / on a Poisson manifold (Z,{,}) such that the Poisson bracket of functions {/,/'} is replaced with the commutator of operators [/, /'] and Dirac's condition llf'} = -i(fJr},
fi=l,
(4.2.1)
272
Geometric and Algebraic Topological Methods in Quantum Mechanics
holds. Here, we show that, conversely, any C*-algebra can be reproduced as a Poisson algebra of smooth functions on a projective Hilbert space [104; 105]. One can think of this procedure as being dequantization [238]. As is well-known (see Theorem 4.2.1 below), any commutative C*-algebra is represented by continuous functions on its spectrum. The above mentioned dequantization in fact generalizes this theorem to non-commutative C*algebras. Let X be a locally compact space. Continuous complex functions vanishing at infinity on X are assembled into a C*-algebra A with respect to the norm | | / | | = sup 1/(^1, feA. (4.2.2) xex If X is compact, A = C°(X) is the unital algebra of continuous complex functions on X. Remark 4.2.1. If X is a compact smooth manifold, the algebra C°°(X) of smooth complex functions on X is a dense subalgebra of the C*-algebra C°(X) of continuous functions on X, but it is not a C*-algebra. D Let A be an arbitrary commutative C*-algebra. Any its irreducible representation is one-dimensional and, consequently, is a character of A, i.e., a homomorphism \ : A —> C of a C-algebra A to the field C. Therefore, the spectrum A of a commutative C*-algebra reduces to the set of characters of A provided with the weak topology as a subset of the dual A' of A. The spectrum A is Hausdorff and, consequently, locally compact. Then A — Prim(j4) in accordance with Proposition 3.1.4. Given an element a G A, the complex function Ta : X - X(a)
(4.2.3)
on A is the well-known Gelfand map. THEOREM 4.2.1. The Gelfand map (4.2.3) is an isomorphism of A to the C*-algebra of continuous complex functions vanishing at infinity on the • spectrum A [129].
In particular, the spectrum of the C*-algebra of continuous complex functions vanishing at infinity on a locally compact space X is homeomorphic to X. Let us turn now to the above-mentioned dequantization procedure. We start with the algebra B(E) of bounded operators in a Hilbert space
273
Chapter 4 Geometry of Algebraic Quantization
(E, (.|.)). Let PE be the projective Hilbert space in Example 4.1.5. Let us assign to each element a G B(E) the complex function / o (x) = (ax\x)
(4.2.4)
on PE, where x £ E is a normalized representative of x G PE. This assignment is a monomorphism of the complex vector space B(E) to the complex vector space C°°(PE) of smooth complex functions on PE such i > fa. In order to make the monomorphism (4.2.4) into a morphism that a* — of algebras, let us provide C°°(PE) with the multiplication
/ * / ' = / / ' + i4s(9/W.
(4-2-5)
where fips is the Kahler form (4.1.33). Then one can show that (4.2.6)
fab = fa*fb for all a, b G B(E) [104; 105]. Moreover, we have \\a\\2= s u p ( 7 a * / o ) ( x ) . x€P£
It follows from the expressions (4.1.24) and (4.2.5) that {/,/'}FS
= -t(/*/'-/'*/),
(4.2.7)
where {,}FS is the Poisson bracket with respect to the Kahler form f2pg (4.1.33) on the projective Hilbert space PE. Then combining the formulae (4.2.6) and (4.2.7) gives the relation {/a,/fc}FS = -»/[o,6],
a,b G B{E).
(4.2.8)
This relation is viewed as the converse Dirac 's condition, when one assigns to operators in a Hilbert space E the smooth functions on the projective Hilbert space PE such that a^
fa,
b>-+ fb,
[a, b] i-> i{fa,
fb}Fs-
Remark 4.2.2. The relation (4.2.8) is converse to Dirac's condition (4.2.1) up to complex conjugation, that can be corrected by passing to • complex conjugate functions a —> (x\ax). Let now A be an arbitrary unital C*-algebra, and let us consider the open surjection P{A) -> A, where P{A) is the space of pure states of A and A is the spectrum of A. In accordance with item (iii) of Theorem 3.1.10,
274
Geometric and Algebraic Topological Methods in Quantum Mechanics
the inverse image PV{A) of P(A) —> A over a point TT £ A, (i.e., the set of pure states of A associated to the same irreducible representation n) is isomorphic to the projective Hilbert space PEn, where En the carrier space of IT. This isomorphism is a homeomorphism if P{A) is provided with the weak* topology and P^(A) C p(A) is endowed with the relative one [105; 398]. Let us assign to each element a £ A the complex function 4> - » fa(d>) =
(4.2.9)
on P(A) —» A. These functions can be subject to the fibrewise multiplication * (4.2.5) on each fibre P^(A) = PEn, where Qps is the Kahler form of the Fubini-Studi metric on PEn [105]. Then a H /„ is a monomorphism of A to the space of continuous complex functions C°(P(A)) on P(A) such that fab = fa*fb, 2
a,b€A,
sup (fa*fa){4>).
\\a\\ =
4>€P(A)
In particular, if A is a commutative C*-algebra, then P(A) = A and the function (4.2.9) restarts the Gelfand map (4.2.3). If a Hilbert space E admits a reproducing system, we have observed a different variant of dequantization of the C*-algebra B{E) by the product (3.3.11) of the mean values (3.3.9). Another dequantization is exemplified by Berezin's quantization in forthcoming Section.
4.3
Berezin's quantization
Berezin's quantization [4; 32; 145] in fact is a particular variant of dequantization. Given a complex Poisson algebra C°°(Z) on a symplectic manifold (Z,£l), one aims to construct an involutive algebra A whose dequantization by functions on Z belongs to C°°(Z). Herewith, A must satisfy the following conditions. (1) There is a family of involutive algebras A^ such that: (la) the index h runs through a subset R C M+ of strictly positive numbers which has 0 in its closure; (lb) A is a subalgebra of © AhheR
Chapter 4 Geometry of Algebraic Quantization
275
(2) There is a linear morphism $ : A —> C°°(Z) with the following properties: (2a) for any two distinct points z, z' € Z, there exists a function / e $(Jl) such that f{z) ^ f{z'); (2b) any two elements a, a1 £ A fulfil the condition ih $([a,a']) = - - { $ ( a ) , $ ( a ' ) } ; (2c) $(a*) = $(a) for all a € A. Prom the physical viewpoint, h plays the role of Plank's constant. The properties (2a) - (2b) are called the correspondence principle. Berezin's quantization is said to obey the weak correspondence principle if the morphism $ is defined on a linear subspace A' of A. Berezin's quantization is called special if it satisfies the following additional properties. (3) Each Ah is a subalgebra of the algebra of complex (not necessarily continuous) functions on Z. (4) The algebra A consists of complex functions a(h, z) on R x Z such that their restrictions to {h} x Z belong to A^ for any h € R. (5) The morphism *:J4-»C°°(Z)
is given by the formula $(a) = )im a(h,z),
a € A.
(4.3.1)
h—>0
It is readily observed that Berezin's quantization axioms do not direct to finding a map / —» / from a classical observable to its quantum counterpart, but to obtaining a quantum algebra A with the appropriate dequantization properties. These axioms are rather restrictive. Therefore, one mainly provides Berezin's quantization only of a particular class of Kahler manifolds (see Remark 4.3.2 below). Let Z be a non-compact Kahler manifold of complex dimension m. Let Z be coordinated by (za,~za) and equipped with a Riemannian metric 9 = 9apdza V dzp
and the Kahler form 9. = iga-pdza
hdz13.
276
Geometric and Algebraic Topological Methods in Quantum Mechanics
The corresponding Poisson bracket of smooth complex functions on Z reads {/, /'} = -ig^idafdpf - daf'd0f).
(4.3.2)
Let U C Z be an open dense subset, where the Kahler metric g is expressed
into a potential function F (see Remark 2.6.2). Let us consider the Hilbert space L^(U, fj,h) of the equivalence classes of complex functions on U which are square integrable with respect to the measure (4.3.3)
Let us provide L^U,^)
with the scalar product
(f\f')h = ch Jfj'vh, u which differs from the standard one (10.5.13) in a normalization constant Ch to be specified later. Let Eh c L^iUi^h) be the subspace of holomorphic functions (called the Bergman space). However, it may happen that Eh consists of only the zero constant. One can show that Eh is a closed Hilbert subspace of L^(U,/J,h), and that it admits the reproducing kernel Kh{z,z')=YJfk{z')h&),
(4.3.4)
where {fk}^L\ is a countable orthonormal basis for Eh such that
]T|/ fc (z)| 2 < oo,
zGZ,
k
[4; 145; 216]. It follows that the set Sh = {L=Kh(z,z'), ||k||£= £|/ fc (z)| 2 , zeZ}
(4.3.5)
k
is a reproducing system in the Hilbert space Eh with respect to the measure iih (4.3.3). Let us note that the reproducing kernel (4.3.4), called the Bergman kernel function, is independent of the choice of an orthonormal basis, and it is antiholomorphic in the first variable and holomorphic in the second
277
Chapter 4 Geometry of Algebraic Quantization
one. The relations (3.3.5) - (3.3.6) hold. As a consequence, any function f E Eh obeys the equality (3.3.8) which reads
/(*) = /, = / f(z')K(z', z)nh(z').
(4.3.6)
Given the reproducing system (4.3.5), one can associate to each bounded operator a in the Hilbert space Eh the mean value function fa{z) (3.3.10) on Z. However, the product (3.3.11) of these functions need a minor modification in order to be utilized in Berezin's quantization procedure. Let us put
alz) = A£f) = (<**&)* _ A a{z) HU 2 «.IOh " El/*(*)l a
'
(4 3 7 )
(
}
fc
It is an element of Eh called the covariant symbol of an operator a. The product of covariant symbols (4.3.7) reads (a*b)(z) = ||6|| 2 (/a * fb)(z) = ab(z).
(4.3.8)
Example 4.3.1. Let Ph denote a projector onto the Hilbert subspace Eh- Then any bounded measurable function g on Z yields the bounded operator
Tg : f -» Ph(gf) = J(gmz)kZ*Hk{z)
(4.3.9)
in Eh- It is called the Toeplitz operator. The covariant symbol (4.3.7) of the Toeplitz operator Tg (4.3.9) coincides with the Berezin transform fg = Bhg(z) = fg(z')lK^Z'^\h(z') J Kh(z,z)
(4.3.10) •
of the function g.
Here, we restrict ourselves to the most interesting case when a potential function F is the logarithm of the reproducing kernel (4.3.4) and when, in accordance with the hypothesis (la), there exists a set R C K + such that Kh(z,z)=Xhex.p\—1F(Z)\,
L "
J
Afc = const.,
(4.3.11)
278
Geometric and Algebraic Topological Methods in Quantum. Mechanics
for any h e R. With some additional technical assumptions, it can be shown that, choosing a normalization constant Ch in an appropriate way, one obtains A^ = 1. Let Ah denote the algebra (under the product (4.3.8)) of covariant symbols of all bounded operators in the Hilbert space Eh- One can think of elements of © Ah as being complex functions on R x Z. Let A be the linear h
subspace of all functions a in © A^ which can be represented in the form h
a(h,z) = a{0,z) + ^^(z)
+ —a2(z),
where a, ai, a-i are functions admitting analytic continuations to Z x Z. One can justify that (a*b)(h,z) =a{0,z)b(0,z) + o(h), ~ ~ ih ~ (a*b-b* a)(h, z) = - — {a, 6}(0, z) + o(h2), where {, } is the Poisson bracket (4.3.2). Consequently, A is desired special Berezin's quantization obeying the weak correspondence principle. Remark 4.3.2. It should be emphasized that the condition (4.3.11) is very restrictive. There are only a few instances where it is satisfied, e.g., Z = C m and Z is a bounded symmetric domain (i.e., it is a bounded open connected subset of Cn whose each point is the isolated fixed point of an involutive holomorphic automorphism of the domain; it is homogeneous, i.e., the group of its holomorphic automorphisms acts transitively) [51; 79; 248]. In order to make Berezin's quantization procedure applicable to a wide class of Kahler manifolds, one studies the case when the condition (4.3.11) is fulfilled only asymptotically as h -> 0 [145]. •
4.4
Hilbert and C*-algebra bundles
Due to the inverse mapping Theorem 4.1.1 for Banach manifolds, fibred Banach manifolds similarly to the finite-dimensional smooth ones can be defined as surjective submersions of Banach manifold onto Banach manifold. Locally trivial fibred Banach manifolds are called (smooth) Banach vector bundles. They are exemplified by vector (e.g., tangent and cotangent) bundles over Banach manifolds in Section 4.1. Here, we restrict our
Chapter 4 Geometry of Algebraic Quantization
279
consideration to particular Banach vector bundles whose fibres are C*algebras (seen as Banach spaces) and Hilbert spaces, but the base is a finite-dimensional smooth manifold. Note that sections of a Banach vector bundle B —> Q over a smooth finite-dimensional manifold Q constitute a locally free C°° (Q)-module B(Q). Following the proof of Proposition 1.8.9, one can show that it is a projective C°° (<2)-module. In a general setting, we therefore can consider projective locally free C°°(Q)-modules, locally generated by a Banach space. In contrast with the case of projective C°°(X) modules of finite rank, such a module need not be a module of sections of some Banach vector bundle. Let C —+ Q be a locally trivial topological fibre bundle over a finitedimensional smooth real manifold Q whose typical fibre is a C*-algebra A, regarded as a real Banach space, and whose transition functions are smooth. Namely, given two trivializations charts (Ui,ipi) and (U2,ip2) of C, we have the smooth morphism of Banach manifolds ipi o ip'1 : Ui n U2 x A -> Ui n U2 x A, where "01 °1p21\q€U1nU2
is an automorphism of A. We agree to call C ^ Q a bundle of C* -algebras. It is a Banach vector bundle. The C°°(Cj)-module C(Q) of smooth sections of this fibre bundle is a unital involutive algebra with respect to fibrewise operations. Let us consider a subalgebra A(Q) C C(Q) which consists of sections of C —> Q vanishing at infinity onQ. It is provided with the norm ||a|| = sup||a( 9 )||
a€A(Q),
(4.4.1)
but fails to be complete. Nevertheless, one extends A(Q) to a C*-algebra of continuous sections of C —» Q vanishing at infinity on a locally compact space Q. Remark 4.4.1. Let C —* Q be a topological bundle of C*-algebras over a locally compact topological space Q, and let C°(Q) denote the involutive algebra of its continuous sections. This algebra exemplifies a locally trivial continuous field of C* -algebras in [129]. Its subalgebra A°(Q) of sections vanishing at infinity on Q is a C*-algebra with respect to the norm (4.4.1). It is called a C* -algebra defined by a continuous field of C*-algebras. There
280
Geometric and Algebraic Topological Methods in Quantum Mechanics
are several important examples of C*-algebras of this type. For instance, every liminal C*-algebra A with a Hausdorff spectrum A is isomorphic to a C*-algebra defined by a continuous field (not necessarily locally trivial) of C*-algebras on A [129]. By virtue of Theorem 4.2.1, any commutative C*-algebra is isomorphic to the algebra of continuous complex functions • vanishing at infinity on its spectrum. Remark 4.4.2. One can consider a locally trivial bundle of C*-algebras C —> Q as a fibre bundle with the structure topological group Aut(A) of automorphisms of A. If a fibre bundle C is smooth, this group is necessarily • provided with a normed operator topology. Hilbert bundles over a smooth manifold are similarly defined. Let £ —> Q be a locally trivial topological fibre bundle over a finite-dimensional smooth real manifold Q whose typical fibre is a Hilbert space E, regarded as a real Banach space, and whose transition functions are smooth functions taking their values in the unitary group U(E) equipped with the normed operator topology. We agree to call £ —> Q a Hilbert bundle. It is a Banach vector bundle. Smooth sections of £ —> Q constitute a C°°(Q)-module £(Q), called a Hilbert module. Continuous sections of £ —> Q constitute a locally trivial continuous field of Hilbert spaces [129]. There are the following relations between bundles of <7*-algebras and Hilbert bundles. Let T(E) C B(E) be the C*-algebra of compact (completely continuous) operators in a Hilbert space E (see Remark 3.1.10). It is called an elementary C* -algebra. Every automorphism <j> of E yields the corresponding automorphism
T{E) ->
q£U,D
Uo
over a cover {Ut} of Q, we have the associated locally trivial bundle of elementary C*-algebras T(E) with transition functions T(E) - pa0(q)T(E)(paP(q))-\
qeUan
Up,
(4.4.2)
which are proved to be continuous with respect to the normed operator topology on T(E) [129]. The proof is based on the following facts.
Chapter 4 Geometry of Algebraic Quantization
281
• The set of degenerate operators (i.e., operators of finite rank) is dense in T(E). • Any operator of finite rank is a linear combination of operators P(,*ta • C - (C\v)S,
t,vX£E,
and even the projectors Pj onto £ 6 E. • Let £i,...,£2™ be variable vectors of E. If &, i — 1,..., 2n, converges to 77$ (or, more generally, (&|£j) converges to (r]i\r]j} for any i and j), then
uniformly converges to Pm,V2 + " ' " + -Pf)2n-l,f?2n-
Note that, given a Hilbert bundle E —> Q, the associated bundle of C*algebras B(E) of bounded operators in E fails to be constructed in general because the transition functions (4.4.2) need not be continuous. The opposite construction however meets a topological obstruction as follows [73; 86]. Let C —> Q be a bundle of C*-algebras whose typical fibre is an elementary C*-algebra T(E) of compact operators in a Hilbert space E. One can think of C —> Q as being a topological fibre bundle with the structure group of automorphisms of T(E). This is the projective unitary group PU(E) (3.8.7. With respect to the normed operator topology, the groups U(E) and PU(E) are the Banach Lie groups [213]. Moreover, U(E) is contractible if a Hilbert space E is infinite-dimensional [263]. Let {Ua,pap) be an atlas of the fibre bundle C —> Q with PU(E)-va\ued transition functions pa0- These transition functions give rise to the maps pap :UanUp^
U(E),
which however fail to be transition functions of a fibre bundle with the structure group U(E) because they need not satisfy the cocycle condition. Their failure to be so is measured by the [/(l)-valued cocycle ealpt =90-y9ay9a0-
This cocycle defines a class [e] in the cohomology group H2(Q;U(1)Q) of the manifold Q with coefficients in the sheaf U(1)Q of continuous maps of Q to U(l). This cohomology class vanishes if and only if there exists a
282
Geometric and Algebraic Topological Methods in Quantum Mechanics
Hilbert bundle associated to C. Let us consider the short exact sequence of sheaves 0 - + Z —>$ JL, y ( i ) Q _> o, where CQ is the sheaf of continuous real functions on Q and the morphism 7 reads 7 : C°Q 3 f .-» exp(27ri/) e
U{\)Q.
This exact sequence yields the long exact sequence (1.7.16) of the sheaf cohomology groups O ^ Z —+ C Q —»tf(l)g —>ff 1 (Q;Z) —»••• tfp(Q;Z)
_ > f r P ( Q ; C § ) — > f P ( Q ; t f ( l ) g ) —>if p + 1 (Q;Z) — > . . . .
Since the sheaf C Q is fine and acyclic, we obtain at once from this exact sequence the isomorphism of cohomology groups
H2(Q,U(1)Q) = H3(Q,Z). The image of [e] in H3(Q,Z) is called the Dixmier-Douady class [129]. One can show that the negative — [e] of the Dixmier-Douady class is the obstruction class of the lift of Pf/(£;)-principal bundles to the C/(£')-principal ones [86]. Thus, studying Hilbert and C*-algebra bundles, we come to fibre bundles with unitary and projective unitary structure groups. 4.5
Connections on Hilbert and C*-algebra bundles
There are different notions of a connection on Hilbert and C* -algebra bundles which need not be obviously equivalent, unless bundles are finitedimensional. These are connections on structure modules of sections, connections as a horizontal splitting and principal connections. Given a bundle of C* -algebras C —> Q with a typical fibre A over a smooth real manifold Q, the involutive algebra C(Q) of its smooth sections is a C°° (Q)-algebra. Therefore, one can introduce a connection on the fibre bundle C —> Q as a connection on the C°°(Q)-algebra C(Q). In accordance with Definition 1.3.4, such a connection assigns to each vector field r on Q a symmetric derivation VT of the involutive algebra C(Q) which obeys the
283
Chapter 4 Geometry of Algebraic Quantization
Leibniz rule VT(fa) = (T\df)a + fVTa,
/ € C°°(Q),
aeC(Q),
and the condition VTa* - (VTa)*. Let us recall that two such connections V r and V'r differ from each other in a derivation of the C°°(Q)-algebra C(Q). Then, given a trivialization chart C\u^UxA of C —> Q, a connection on C(Q) can be written in the form VT=Tm(q)(dm-6m(q)),
(4.5.1)
q£U,
where (qm) are local coordinates on Q and Sm(q) for all q € U are symmetric bounded derivations of the C*-algebra A. Remark 4.5.1. Bearing in mind the discussion in Section 4.2, one should assume that, in physical models, the derivations 8m(q) in the expression (4.5.1) are unbounded in general. This leads us to the notion of a generalized connection on bundles of C*-algebras [15]. • Let £ —> Q be a Hilbert bundle with a typical fibre E and £{Q) the C°°((3)-module of its smooth sections. Then a connection on a Hilbert bundle 8 —> Q is defined as a connection V on the module £{Q). In accordance with Definition 1.8.7, such a connection assigns to each vector field r on Q a first order differential operator V r on £{Q) which obeys both the Leibniz rule V r W ) = (T\df)1> + /V T V,
/ G C°°(Q),
V G S(Q),
and the additional condition ((VrVOfa) W«)> + W?)l(VTV)(g)> = T(q)}dU>(q)\1>(q)). Given a trivialization chart £\u =UxEoi£—>Q,a, reads Vr = rm(q)(dm + iHm(q)),
(4.5.2)
connection on £{Q) q € U,
(4.5.3)
where TCm(q) for all q £ J7 are bounded self-adjoint operators in the Hilbert space E.
284
Geometric and Algebraic Topological Methods in Quantum Mechanics
In a general setting, let B —> Q be a Banach vector bundle over a finitedimensional smooth manifold Q and B(X) the locally free C°°(Q)-module of its smooth sections s(q). By virtue of Definition 1.8.7, a connection on B(Q) assigns to each vector field r on Q a first order differential operator VT on B(Q) which obeys the Leibniz rule VT(fs) = (T\df)s + fVTs,
/£CM(0),
seB(Q).
(4.5.4)
In accordance with Proposition 1.8.11, such a connection exists. Connections (4.5.1) and (4.5.3) exemplify connections on Banach vector bundles C —> Q and S —> Q, but they obey additional conditions because these bundles possess additional structures of a C* -algebra bundle and a Hilbert bundle, respectively. In particular, the connection (4.5.3) is a principal connection whose second term is an element of the Lie algebra of the unitary group U(E). In a different way, connection on a Banach vector bundle B —> Q can be defined as a splitting of the exact sequence 0 -> VB -^ TB -> TQ®B -> 0, Q
where VB denotes the vertical tangent bundle of B —> Q. In the case of finite-dimensional vector bundles, both definitions are equivalent (see Section 1.8). This equivalence is extended to the case of Banach vector bundles over a finite-dimensional base. We leave the proof of this fact outside the scope of our exposition because it involves the notion of jets of Banach fibre bundles. Turn now to principal connections. Given a Banach-Lie group G, a principal bundle over a finite-dimensional smooth manifold Q, a principal connection, its curvature form and a holonomy group are defined similarly to those in the case offinite-dimensionalLie groups. The main difference lies in the facts that there are Banach-Lie algebras without Lie groups and the holonomy group of a principal connection need not be a Lie group. Referring the reader to [262] for theory of Lie groups and principal bundles modelled over so called convenient locally convex vector spaces (including Frechet spaces), we here formulate some statements adapted to the case of Banach-Lie groups and Banach principal bundles over a finite-dimensional manifold. • Any Banach-Lie group G admits an exponential mapping which is a diffeomorphism of a neighborhood of 0 in the Lie algebra g of G onto a neighborhood of the unit in G. In a general setting, one can always
Chapter 4 Geometry of Algebraic Quantization
285
associate to a Banach-Lie algebra a local Banach-Lie group which however fails to be extended to the global one in general [213]. • Let G be a Banach-Lie group and fl its Lie algebra. If f) is a closed Lie subalgebra offl,there exists a unique connected closed Banach-Lie subgroup H of G with the Lie algebra I) [373]. • Given a Banach-Lie group, the definition of a G-principal bundle P —> Q over a finite-dimensional smooth manifold Q, a principal connection and its curvature form in [262] follows those in the case of a locally compact Lie group [250]. A principal connection T on P defines the global parallel transport and a holonomy group. In particular, the following generalizations of the reduction theorem ([250], Theorem 7.1) and the AmbroseSinger theorem ([250], Theorem 8.1) to Banach principal bundles hold [291; 431]. THEOREM 4.5.1. Let P -* Q be a principal bundle with a Banach-Lie structure group G over a simply connected finite-dimensional manifold Q. Let H be & Banach-Lie subgroup of G. Let us assume that there exists a principal connection on P whose curvature form u possesses the following property. For any smooth one-parameter family of horizontal paths Hcs starting at a point p e P and arbitrary smooth vector fields u, u' on Q,
[0,l}23s,t^ujC3it)(u,u')
(4.5.5)
is a smooth (j-valued map. Then the structure group G of P is reduced to
•
H.
4.5.2. Let us consider closed Lie subalgebras of the Lie algebra Q which contain the range of the map (4.5.5). Their overlap is the minimal closed Lie subalgebra gred of Q possessing this property. The corresponding Banach-Lie group Gred is the minimal Banach-Lie group which contains the holonomy group of a connection T. By virtue of Theorem 4.5.1, the structure group F of P is reduced to GredD THEOREM
• Theorem 10.10.5 holds when G and H are Banach-Lie groups [86]. • Given a trivialization chart of a Banach principal bundle P —> Q with a structure Banach-Lie group G, a principal connection on P is represented by a g-valued local connection one-form Tmdqm with the corresponding transition functions. Let B = {Px V)/G
286
Geometric and Algebraic Topological Methods in Quantum Mechanics
be a Banach vector bundle associated to P whose typical fibre V is a Banach space provided with a continuous effective left action of the structure group G. Then a principal connection T on P yields a connection on B given by the first order differential operators Vr = r m ( 9 m - r m )
(4.5.6)
on the C°°(Q) module B(Q) of sections of B —> Q which obey the Leibniz rule (4.5.4). For instance, let G = U(E) be the unitary group of a Hilbert space E. Its Lie algebra consists of the operators i7i, where Ji are bounded selfadjoint operators in the Hilbert space E. It follows that a t/(£')-principal connection takes the form (4.5.3). In conclusion, let us mention the straightforward definition of a connection on a Hilbert bundle as a parallel displacement along paths lifted from a base [228]. Roughly speaking, such a connection corresponds to parallel displacement operators whose infinitesimal generators are (4.5.3). Due to the condition (4.5.2), these operators are unitary. If a path is closed, we come to the notion of a holonomy group of a connection on a Hilbert bundle. 4.6
Example. Instantwise quantization
As will be shown in Section 5.3, geometric quantization of time-dependent Hamiltonian mechanics takes the form of the instantwise quantization, and results in a quantum system described by a Hilbert bundle over the time axis R. This Section addresses the evolution of such quantum systems which can be viewed as a parallel displacement along time. It should be emphasized that, in quantum mechanics based on the Schrodinger and Heisenberg equations, the physical time plays the role of a classical parameter. Indeed, all relations between operators in quantum mechanics are simultaneous, while computation of mean values of operators in a quantum state does not imply integration over time. It follows that, at each instant ( € l , there is an instantaneous quantum system characterized by some C*-algebra At. Thus, we come to instantwise quantization. Let us suppose that all instantaneous C*-algebras At are isomorphic to some unital C*-algebra A. Furthermore, let they constitute a locally trivial smooth bundle C of C*-algebras over the time axis M. Its typical fibre is A. This bundle of C*-algebras is trivial, but need not admit a canonical trivial-
287
Chapter 4 Geometry of Algebraic Quantization
ization in general. One can think of its different trivializations as being associated to different reference frames (see Section 2.4). Let us describe evolution of quantum systems in the framework of instantwise quantization. Given a bundle of C*-algebras C —» R, this evolution can be regarded as a parallel displacement with respect to some connection on C —> R [15; 228; 385]. Following previous Section, we define V as a connection on the involutive C°° (R)-algebra C(R) of smooth sections of C —> M. It assigns to the standard vector field dt on R a symmetric derivation V t of C(R) which obeys the Leibniz rule V t (/a) = dtfa + / V t a ,
a e C(R),
/ € C°°(R),
and the condition V t a* = (V t a)'. Given a trivialization C — R x A, a connection V t reads Vt = $ - * ( « ) ,
(4.6.1)
where 6(t), t G R, are symmetric derivations of the C*-algebra A, i.e., <5t(ab) = 6t(a)b + aSt(b),
5t{a*) = 6t(a)*,
a,b £ A.
We say that a section a of the bundle of C*-algebras C —> R is an integral section of the connection Vt if V t o(t) = [dt - 6(t)]a(t) = 0.
(4.6.2)
One can think of the equation (4.6.2) as being the Heisenberg equation describing quantum evolution. In particular, let the derivations S(t) = 5 in the Heisenberg equation (4.6.2) be the same for all £ £ M, and let 5 be an infinitesimal generator of a strongly continuous one-parameter group [Gt] of automorphisms of the C*algebra A (see Theorem 3.8.2). The pair (A, [Gt]) is called the C*-dynamic system. It describes evolution of a conservative quantum system. Namely, for any a £ A, the curve a(t) — Gt(a), t £ R, in A is a unique solution with the initial value a(0) — a of the Heisenberg equation (4.6.2). It should be emphasized that, if the derivation 5 is unbounded, the connection Vt (4.6.1) is not defined everywhere on the algebra C(R). In this case, we deal with a generalized connection. It is given by operators of a parallel displacement, whose generators however are ill defined [15]. Moreover, it may happen that a representation w of the C*-algebra A does not
288
Geometric and Algebraic Topological Methods in Quantum Mechanics
carry out a representation of the automorphism group [Gt] (see Proposition 3.8.4). Therefore, quantum evolution described by the conservative Heisenberg equation, whose solution is a strongly (but not uniformly) continuous dynamic system (A, Gt), need not be described by the Schrodinger equation (see Remark 4.6.1 below). If 6 is a bounded derivation of a C*-algebra A, the Heisenberg and Schrodinger pictures of evolution of a conservative quantum system are equivalent. Namely, by virtue of Theorem 3.8.1, S is an infinitesimal generator of a uniformly continuous one-parameter group [Gt] of automorphisms of A, and vice versa. For any representation TT of A in a Hilbert space E, there exists a bounded self-adjoint operator Tim E such that n(5{a)) = -i[H,ir(a)],
n(Gt) = exp(-iiW),
a e A,
The corresponding conservative Schrodinger equation reads (dt + iH)ip = 0,
tgt. (4.6.3) (4.6.4)
where ip is a section of the trivial Hilbert bundle R x E —» R. Its solution with an initial value tp(O) S E is ip(t) = exp[-itH]tp(0).
(4.6.5)
Remark 4.6.1. If the derivation S is unbounded, but obeys the assumptions of Proposition 3.8.4, we also obtain the unitary representation (4.6.3) of the group [Gt], but the curve ip(t) (4.6.5) need not be differentiate, and the Schrodinger equation (4.6.4) is ill denned. • Let us return to the general case of a non-conservative quantum system characterized by a bundle of C*-algebras C —> R with the typical fibre A. Let us suppose that a phase Hilbert space of a quantum system is preserved under evolution, i.e., instantaneous C*-algebras At are endowed with representations equivalent to some representation of the C*-algebra A in a Hilbert space E. Then quantum evolution can be described by means of the Schrodinger equation as follows. Let us consider a smooth Hilbert bundle £ —> R with the typical fibre E and a connection V on the C°°(R)-module £ (R) of smooth sections of £ —> R (see previous Section). This connection assigns to the standard vector field dt on R an M-module endomorphism Vt of £ (R) which obeys
Chapter 4 Geometry of Algebraic Quantization
289
the Leibniz rule
V t (/V) = dtfip + /V t V,
i> e £{R),
/ e C°°(R),
and the condition
((vtv)(i)|VW) +
(4.6.6)
where W(t) are bounded self-adjoint operators in E for all t £ l . It is a C/(£)-principal connection. We say that a section ip of the Hilbert bundle II —» M is an integral section of the connection V t (4.6.6) if it fulfils the equation VtiKO = (dt + iH{t))ijt(t) = 0.
(4.6.7)
One can think of this equation as being the Schrodinger equation for the Hamiltonian H. Its solution with an initial value ^(0) G E exists and reads (4.6.8)
1>{t) = U(t)1>{0),
where U(t) is an operator of a parallel displacement with respect to the connection (4.6.6). This operator is a differentiable section of the trivial bundle
R which obeys the equation
v
TJ(
T7l\
TIT)
X U \-tli) —> K
Uo = 1.
dtU{t) = -iH(t)U(t),
(4.6.9)
The operator U(t) plays the role of an evolution operator. It is given by the time-ordered exponential t
U{t)=Texp -ifn(t')dt' o
,
(4.6.10)
which uniformly converges in the operator norm [118]. Under certain conditions, U(t) can be written as a true exponential U(t) = exp S(t)
290
Geometric and Algebraic Topological Methods in Quantum Mechanics
of an anti-Hermitian operator S(t) which is expressed as the Magnus series oo
s(*)=5>fc(t) fc=l
of multiple integrals of nested commutators [266; 348]. It should be emphasized that the evolution operator [/(£) has been defined with respect to a given trivialization of a Hilbert bundle £ —> R. 4.7
Example. Berry connection
We refer the reader to [7; 43; 246; 320; 443] and references therein for the geometric and topological analysis of the Berry's phase phenomenon in quantum systems depending on classical time-dependent parameters. In Section 2.4F, classical mechanical systems with time-dependent parameters have been described in terms of composite fibre bundles and composite connections. Here, this description is extended to quantum systems. Let us consider a quantum system depending on a finite number of real classical parameters given by sections of a smooth parameter bundle £ —> R. For the sake of simplicity, we fix a trivialization S = R x Z, coordinated by (t, am). Although it may happen that the parameter bundle E —> R has no preferable trivialization. In previous Section, we have characterized the time as a classical parameter in quantum mechanics. This characteristic is extended to other classical parameters. In a general setting, one assigns a C*-algebra Aa to each point a £ £ of the parameter bundle E, and treat Aa as a quantum system under fixed values (£, am) of the parameters. However, we will simplify repeatedly our consideration in order to single out a desired Berry's phase phenomenon. Let us assume that all algebras Aa are isomorphic to the algebra B{E) of bounded operators in some Hilbert space E, and consider a smooth Hilbert bundle £ —> S with the typical fibre E. Smooth sections of £ —> S constitute a module £(E) over the ring C°°(E) of real functions on S. A connection V on £(E) assigns to each vector field r on E a first order differential operator VTGDiff !(£(£),£(£))
(4.7.1)
291
Chapter 4 Geometry of Algebraic Quantization
which obeys the Leibniz rule V T (/s) = (rJd/) S + / V r s ,
SGf(E),
/€C°°(£).
Let r be a vector field on £ such that dt\ r = 1. Given a trivialization chart of the Hilbert bundle £ —> E, the operator V T (4.7.1) reads V r (s) = (ft + iW(t, ff'))s + r m ( 9 m - ^ ( t ,
(4.7.2)
where H(t,al), Am(t,al) for each cr G S are bounded self-adjoint operators in the Hilbert space E. Let us consider the composite Hilbert bundle £ —+ £ —> R. Similarly to the case of smooth composite fibre bundles (see Section 10.6A), every section h(t) of the fibre bundle £ —> R defines the subbundle £h — h*£ —> R of the Hilbert bundle £ —> R whose typical fibre is E. Accordingly, the connection V (4.7.2) on the C°°(S)-module £(£) yields the pull-back connection Vh(V) = [ft - <(An(t, V(t))ftftro - W(t, ^(i))]V
(4.7.3)
on the C°°(R)-module £h(M) of sections V of the fibre bundle £h -» R (cf. (2.4.64)). As in previous Section, we say that a section ip of the fibre bundle £h —> R is an integral section of the connection (4.7.3) if
Vh(V) = [ft - i(Am(t, h'itydth™ - H{t, h'it))}^ = 0.
(4.7.4)
One can think of the equation (4.7.4) as being the Shrodinger equation for a quantum system depending on the parameter function h(t). Its solutions take the form (4.6.8) where U(t) is the time-ordered exponent "
t
U{t)=Texp i f(Amdt'hm-H)dt' . o
.
(4.7.5)
The term iAm(t, hl(t))dthm in the Shrodinger equation (4.7.4) is responsible for the Berry's phase phenomenon, while "H. is treated as an ordinary Hamiltonian of a quantum system. To show the Berry's phase phenomenon clearly, we will continue to simplify the system under consideration. Given a trivialization of the fibre bundle £ —» R and the above mentioned trivialization £ = R x Z of the parameter bundle E, let us suppose that the components Am of the connection V (4.7.2) are independent of t and that
292
Geometric and Algebraic Topological Methods in Quantum Mechanics
the operators 7i(a) commute with the operators Am{a') at all points of the curve h(t) C S. Then the operator U(t) (4.7.5) takes the form
r * £/(*) = Texp i
.Texp -i jH{t')dt'
f Am{ais)dam
. h([o,t})
\
.
(4.7.6)
o
I
One can think of the first factor in the right-hand side of the expression (4.7.6) as being the operator of a parallel transport along the curve h([0, i\) C Z with respect to the pull-back connection V = i*V = dm-iAm(t,ai)
(4.7.7)
on the fibre bundle £ —» Z, defined by the imbedding i : Z <-> {0} x Z c E. Note that, since Am are independent of time, one can utilize any imbedding of Z to {*} x Z. Moreover, the connection V (4.7.7), called the Berry connection, can be seen as a connection on some principal fibre bundle P —» Z for the unitary group U(E). Let the curve /i([0,£]) be closed, while the holonomy group of the connection V at the point h(t) = h(0) is not trivial. Then the unitary operator Texp i
I Amio-^da™
(4.7.8)
. M[o,t]) is not the identity. For example, if (4.7.9)
iAmio-') = iA^a^lds
is a [/(l)-principal connection on Z, then the operator (4.7.8) is the wellknown Berry phase factor exp i
f Am{ai)dam
.
. h([o,t])
If (4.7.9) is a curvature-free connection, Berry's phase is exactly the Aharonov-Bohm effect on the parameter space Z. The following variant of the Berry's phase phenomenon leads us to a principal bundle for familiar finite-dimensional Lie groups. Let a Hilbert
Chapter 4 Geometry of Algebraic Quantization
293
space E be the Hilbert sum of n-dimensional eigenspaces of the Hamiltonian H(a), i.e., oo
£=0£fc,
Ek = Pk(E),
where Pk are the projection operators, i.e., H(a)oPk=\k(a)Pk
(in the spirit of the adiabatic hypothesis). Let the operators Am(z) be time-independent and preserve the eigenspaces Ek of the Hamiltonian H, i.e., Am(z)='£Akm(z)Pk,
(4.7.10)
k
where A!^(z), Z € Z, are self-adjoint operators in Ek- It follows that Am(a) commute with Ti.(<j) at all points of the parameter bundle S —> R. Then, restricted to each subspace .£&, the parallel transport operator (4.7.8) is a unitary operator in Ek- In this case, the Berry connection (4.7.7) on the [/(-E)-principal bundle P —> Z can be seen as a composite connection on the composite bundle P -* P/U(n) - Z, which is defined by some principal connection on the [/(n)-principal bundle P —» P/U(n) and the trivial connection on the fibre bundle P/U(n) —> Z. Note that, since U(E) is contractible, the [/(n)-principal bundle U(E) —» U(E)/U(n) is universal (see Section 10.10) and, consequently, the typical fibre U(E)/U(n) of P/U(n) —> Z is exactly the classifying space B(U(n) (10.10.32) of [/(n)-principal bundles. Moreover, one can consider the parallel transport along a curve in the bundle P/U(n). In this case, a state vector ip(t) acquires a geometric phase factor in addition to the dynamical phase factor. In particular, if S = R (i.e., classical parameters are absent and Berry's phase has only the geometric origin) we come to the case of a Berry connection on the [/(n)-principal bundle over the classifying space B(U(n)) (see [43]). If n = 1, this is the variant of Berry's geometric phase of [7].
Chapter 5
Geometric quantization
The geometric quantization procedure follows the principle of canonical quantization by replacing a Poisson bracket of smooth functions on a Poisson manifold with a commutator product of operators in a Hilbert space such that Dirac's condition (4.2.1) holds. In the framework of geometric quantization, these operators are constructed by means of a suitable U(l)connection. 5.1
Leafwize geometric quantization
Subsections: A. Prequantization, 296; B. Polarization, 302; C. Quantization, 303. We refer the reader to [141; 257; 401; 438] for the basics on geometric quantization of symplectic manifolds. This quantization technique has been generalized to Poisson manifolds in terms of contravariant connections [425; 426; 427] (see Section 5.3). Though there is one-to-one correspondence between the (regular) Poisson structures on a smooth manifold and its symplectic foliations, geometric quantization of a Poisson manifold need not imply quantization of its symplectic leaves [427]. • Firstly, contravariant connections fail to admit the pull-back operation. Therefore, prequantization of a Poisson manifold does not determine straightforwardly prequantization of its symplectic leaves. • Secondly, polarization of a Poisson manifold is defined in terms of sheaves of functions, and it need not be associated to any distribution. As a consequence, its pull-back onto a leaf is not polarization of a symplectic manifold in general. • Thirdly, a quantum algebra of a Poisson manifold contains the center 295
296
Geometric and Algebraic Topological Methods in Quantum Mechanics
of a Poisson algebra. However, there are models where quantization of this center has no physical meaning. For instance, the center of the Poisson algebra of a mechanical system with classical parameters consists of functions of these parameters [175]. Geometric quantization of symplectic foliations disposes of these problems. The quantum algebra of a symplectic foliation is also the quantum algebra of the associated Poisson manifold such that its restriction to each symplectic leaf is denned and quantized. Thus, geometric quantization of a symplectic foliation provides the leafwise quantization of a Poisson manifold. This is the case of systems whose symplectic leaves are indexed by non-quantizable variables. In particular, if a Poisson manifold is symplectic, we come to the familiar geometric quantization of symplectic manifolds (see its examples in Sections 5.2, 5.3 and 5.6). Geometric quantization of a symplectic foliation is phrased in terms of leafwise connections on a foliated manifold [175]. Firstly, we have seen that homomorphisms of the de Rham cohomology of a Poisson manifold both to the de Rham cohomology of its symplectic leaf and the LP cohomology factorize through the leafwise de Rham cohomology (Propositions 2.2.1 and 2.2.2). Secondly, any leafwise connection on a complex line bundle over a Poisson manifold is proved to come from a connection (see Theorem 5.1.3 below). Using these facts, we state the equivalence of prequantization of a Poisson manifold to prequantization of its symplectic foliation, which also yields prequantization of each symplectic leaf. On the contrary, polarization of a symplectic foliation is associated to a particular polarization of a Poisson manifold, and its restriction to any symplectic leaf is polarization of this leaf. Therefore, we define metaplectic correction of a symplectic foliation so that its quantum algebra restricted to each leaf is quantized. It is represented by Hermitian operators in the pre-Hilbert space of leafwise half-forms, integrable over the leaves of this foliation. A. Prequantization Prequantization of a symplectic foliation {T, Cl?) of a manifold Z provides a representation
/ -»if,
[/, /'] = -iUJ^r,
(5-1.1)
of the Poisson algebra (C°°(Z), {/, f'}?) by first order differential operators on sections of a complex line bundle TT : C —> Z, called the prequantization bundle. These operators are given by the Kostant-Souriau prequantization
297
Chapter 5 Geometric Quantization
formula / = -tV£+e/,
0/=n$r(d/),
£>O,
(5.1-2)
where V^ is a leafwise connection on C —» Z (see Definition 5.1.1 below) such that its curvature form obeys the prequantization condition (5.1.3)
R = ieQjr.
Using the above mentioned fact that any leafwise connection comes from a connection, we provide the cohomology analysis of this condition, and show that prequantization of a symplectic foliation yields prequantization of its symplectic leaves. DEFINITION 5.1.1. In the framework of the leafwise differential calculus, a (linear) leafwise connection on the complex line bundle C —> Z is defined as a connection V^ on the C°°(Z)-module C(Z) of global sections of this bundle, where C°°(Z) is regarded as a Sf(Z)-ring. It associates to each element r G T\{F) an 5^(Z)-linear endomorphism Vf oiC(Z) which obeys the Leibniz rule
Vf(/«) = (Tjd7)s + /Vf00,
f€C°°{Z),
seC(Z).
(5.1.4)
• A linear connection on C —> Z can be equivalently defined as a connection on the module C{Z) which assigns to each vector field r & T\(Z) on Z an R-linear endomorphism of C(Z) obeying the Leibniz rule (5.1.4). Restricted to Tx(J-), it obviously yields a leafwise connection. In order to show that any leafwise connection is of this form, we will appeal to an alternative definition of a leafwise connection in terms of leafwise forms. The inverse images 7r -1 (F) of leaves F of the foliation T of Z provide a (regular) foliation C? of the line bundle C. Given the (holomorphic) tangent bundle TCjr of this foliation, we have the exact sequence of vector bundles 0->VC —>TCr - ^ C x T f ^ O , c c z where VC is the (holomorphic) vertical tangent bundle of C —> Z.
(5.1.5)
DEFINITION 5.1.2. A (linear) leafwise connection on the complex line bundle C —> Z is a splitting of the exact sequence (5.1.5) which is linear over C. O
298
Geometric and Algebraic Topological Methods in Quantum Mechanics
One can choose an adapted coordinate atlas {(U^;zx;zi)} (10.6.27) of a foliated manifold (Z, !F) such that U^ are trivialization domains of the complex line bundle C —> Z. Let (zx\zt;c), c € C, be the corresponding bundle coordinates on C —» Z. They are also adapted coordinates on the foliated manifold (C, C?). With respect to these coordinates, a (linear) leafwise connection is represented by a TCV-valued leafwise one-form Ayr = dz*
(5.1.6)
where Ai are local complex functions on C. The exact sequence (5.1.5) is obviously a subsequence of the exact sequence 0 -» VC —^ TC —»C x TZ -» 0, C
C
Z
where TC is the holomorphic tangent bundle of C. Consequently, any connection r = dzx ® {dx + Txcdc) + dzi ® {di + Ticdc)
(5.1.7)
on the complex line bundle C —> Z yields a leafwise connection IV = dz* ® (di + Yicdc).
(5.1.8)
5.1.3. Any leafwise connection on the complex line bundle C —> Z comes from a connection on it. •
THEOREM
Outline of proof. Let A? (5.1.6) be a leafwise connection on C —> Z and IV (5.1.8) a leafwise connection which comes from some connection V (5.1.7) on C —> Z. Their affine difference over C is a section Q = Ayr - IV = dzi ® {Ai - Ti)cdc of the vector bundle TT* ®VC^C. c Given some splitting B:dzi<-> dzi - B{dzx
(5.1.9)
of the exact sequence (10.6.29), the composition (B ®Idvc)oQ
= (dzi - B\dzx) ® (Ai - Ti)cdc :C-^T*Z®VC
299
Chapter 5 Geometric Quantization
is a soldering form on the complex line bundle C —> Z. Then T + (B®Idvc)oQ
=
x
dz ® {dx + [rA - B\{Ai - Ti)]cdc) + dz* ® (ft + AiC9c) is a desired connection on C —> Z which yields the leafwise connection Ajr (5.1.6). QED In particular, it follows that Definition 5.1.1 and Definition 5.1.2 of a leafwise connection are equivalent, namely, V:Fs = ds-Aisdzi,
seC(Z).
The curvature of a leafwise connection V^ is defined as a C°°(Z)-linear endomorphism R(T, T') = V£ T ,, - [Vf, V?] = T ' 7 % - ,
Rij = ftyl,- - a ^ i , (5.1.10)
of C(Z) for any vector fields r, r ' € 7i (^"). It is represented by the complex leafwise two-form R=-Rijdzi
NdzK
(5.1.11)
If a leafwise connection V ^ comes from a connection V, its curvature leafwise form R (5.1.11) is the image R = i*-pR
of the curvature form R of the connection V with respect to the morphism %*T (2.2.2). Now let us turn to the prequantization condition (5.1.3). LEMMA 5.1.4. Let us assume that there exists a leafwise connection F> on
the complex line bundle C —> Z which fulfils the prequantization condition (5.1.3). Then, for any Hermitian form g on C —> Z, there exists a leafwise connection A9-p on C —> Z which: (i) satisfies the condition (5.1.3), (ii) preserves g, (iii) comes from a L/(l)-principal connection on C —> Z. • Outline of proof. For any Hermitian form g on C —> Z, there exists an associated bundle atlas * 9 = {(zx; z%, c)} of C with [/(l)-valued transition
300
Geometric and Algebraic Topological Methods in Quantum Mechanics
functions such that (5.1.12)
g(c,c')=cc'.
Let the above mentioned leafwise connection Tj? come from a linear connection r (5.1.7) on C —> Z written with respect to the atlas $9. The connection T is split into the sum A9 + 7 where A9 = dzx
(5.1.13)
is a [/(l)-principal connection, preserving the Hermitian form g. The curvature forms R of T and R9 of A9 obey the relation R9 = lm(R). The connection A9 (5.1.13) defines the leafwise connection A% = i*TA = dzi ® (di + iA?cdc),
iA? = ha(Ti),
(5.1.14)
preserving the Hermitian form g. Its curvature fulfils a desired relation R? = ifR9
= Im(i^iZ) = iefijr.
(5.1.15) QED
Since A s (5.1.13) is a [/(l)-principal connection, its curvature form R9 is related to the first Chern form of integer de Rham cohomology class by the formula ci = i{2-K)-lR9 (10.10.34). If the prequantization condition (5.1.3) holds, the relation (5.1.15) shows that the leafwise cohomology class of the leafwise form (27r)^1e0jr is the image of an integer de Rham cohomology class with respect to the cohomology morphism [ij-] (2.2.3). Conversely, if a leafwise symplectic form 0.? on a foliated manifold (Z,!F) is of this type, there exist a complex line bundle C —> Z and a (/(l)-principal connection A on C —* Z such that the leafwise connection i*TA fulfils the relation (5.1.3). Thus, we have stated the following. 5.1.5. A symplectic foliation {T,Q.jr) of a manifold Z admits prequantization (5.1.2) if and only if the leafwise cohomology class of (2IT)~1EQ.F is the image of an integer de Rham cohomology class of Z. D PROPOSITION
In particular, let (Z,w) be a Poisson manifold and (!F, f2jr) its characteristic symplectic foliation. As is well-known, a Poisson manifold admits prequantization if and only if the LP cohomology class of the bivec-
Chapter 5 Geometric Quantization
301
tor field (2ir)~lew, e > 0, is the image of an integer de Rham cohomology class with respect to the cohomology morphism [iu"] (2.1.34) [425; 426]. By virtue of Proposition 2.2.2, this morphism factorizes through the cohomology morphism [ip] (2.2.3). Therefore, in accordance with Proposition 5.1.5, prequantization of a Poisson manifold takes place if and only if prequantization of its symplectic foliation does well, and both these prequantizations utilize the same prequantization bundle C —> Z. Herewith, each leafwise connection V^ obeying the prequantization condition (5.1.3) yields the admissible contravariant connection V£ := V £ , w >
<j> G O\Z),
on C —> Z whose curvature bivector equals iew. Clearly, V^ and V™ lead to the same prequantization formula (5.1.2). Let F be a leaf of a symplectic foliation {T, Cl?) provided with the symplectic form
In accordance with Proposition 2.2.1 and the commutative diagram of cohomology groups H*{Z;Z) —>tf-(Z)
I
H*(F;Z)
1 •
-^H*{F)
the symplectic form (2-K)~ SQ.F belongs to an integer de Rham cohomology class if a leafwise symplectic form Q.j: fulfils the condition of Proposition 5.1.5. This states the following. 1
PROPOSITION 5.1.6. If a symplectic foliation admits prequantization, each its symplectic leaf does well. •
The corresponding prequantization bundle for F is the pull-back complex line bundle i*FC, coordinated by (zi,c). Furthermore, let A9^ (5.1.14) be a leafwise connection on the prequantization bundle C —> Z which obeys Lemma 5.1.4, i.e., comes from a {/(l)-principal connection A9 on C —> Z. Then the pull-back AF = i*FA9 = dz* ® (dt + iiF{A9)cdc)
(5.1.16)
302
Geometric and Algebraic Topological Methods in Quantum Mechanics
of the connection A9 onto i*FC —> F satisfies the prequantization condition RF = i*FR = ieflF, and preserves the pull-back Hermitian form i*Fg on i*rC —» F.
B. Polarization Let us define polarization of a symplectic foliation (F, £V) of a manifold Z as a maximal (regular) involutive distribution T c TF on Z such that n?(u,v) = 0,
U,VGTZ,
zeZ.
(5.1.17)
Given the Lie algebra T(Z) of T-subordinate vector fields on Z, let Ar C C°°(Z) be the complexified subalgebra of functions / whose leafwise Hamiltonian vector fields •#/ (2.2.10) fulfil the condition
l#f,T(Z)]cT(Z). It is called the quantum algebra of a symplectic foliation (F, Qj?) with respect to the polarization T. This algebra obviously contains the center Sj?(Z) of the Poisson algebra (C°°(Z), {,}?), and is a Lie 5^(Z)-algebra. PROPOSITION 5.1.7. Every polarization T of a symplectic foliation (F, Q?) • yields polarization of the associated Poisson manifold (Z, w^i).
Outline of proof. Let us consider the presheaf of local smooth functions / on Z whose leafwise Hamiltonian vector fields $f (2.2.10) are subordinate to T. The sheaf $ of germs of these functions is polarization of the Poisson manifold (Z, wn). Equivalently, $ is the sheaf of germs of functions on Z whose leafwise differentials are subordinate to the codistribution fi^T. QED
Let us note that the polarization $ need not be maximal, unless T is of maximal dimension dim^-"/2. It belongs to the following particular type of polarizations of a Poisson manifold. Since the cochain morphism i*^ (2.2.2) is an epimorphism, the leafwise differential calculus 5* is universal, i.e., the leafwise differentials df of functions / € C°°(Z) on Z make up a basis for the C°°(Z)-module &{Z). Let $(Z) denote the structure M-module of global sections of the sheaf $. Then the leafwise differentials of elements of $(Z) make up a basis for the C°°(Z)-module of global sections of the codistribution fi^-T. Equivalently, the leafwise Hamiltonian vector fields of elements of $(Z) constitute a basis for the C°°(Z)-module T(Z). Then
303
Chapter 5 Geometric Quantization
one can easily show that polarization T of a symplectic foliation (!F, Qyr) and the corresponding polarization $ of the Poisson manifold (Z, WQ) in Proposition 5.1.7 define the same quantum algebra ATLet (F, SIF) be a symplectic leaf of a symplectic foliation (T, CIT)- Given a polarization T —> Z of (F, CIT), its restriction T F = i*FT c i*FTT = TF to F is an involutive distribution on F. It obeys the condition i*FSlr(u,v)=0,
u,v€
TFz,
zeF,
i.e., it is polarization of the symplectic manifold (F, £lF). Thus, we have stated the following. PROPOSITION 5.1.8. Polarization of a symplectic foliation defines polar• ization of each symplectic leaf.
Clearly, the quantum algebra AF of a symplectic leaf F with respect to the polarization TF contains all elements iFf of the quantum algebra A? restricted to F.
C. Quantization Since AT is the quantum algebra both of a symplectic foliation {T, QT) and the associated Poisson manifold (Z, WQ), let us start with the standard metaplectic correction technique [141; 438]. Assuming that Z is oriented and that H2(Z;Z2) = 0, let us consider the metalinear complex line bundle T> —» Z characterized by an atlas #z = { ( % z V ; c ) } with the transition functions d = 5c such that 5 5 is the inverse Jacobian of coordinate transition functions on Z. Global sections of this bundle are half-forms on Z. The metalinear bundle V belongs to the category of natural bundles, and the Lie derivative L r = rxdx + Tldi + i(9 A r A + diT*)
(5.1.18)
of its sections along any vector field r on Z is defined. The quantization bundle is the tensor product Y = C <8> V. The space YK(Z) of its sections
304
Geometric and Algebraic Topological Methods in Quantum Mechanics
of compact support is provided with the non-degenerate Hermitian form / . v dimZ/2
^
J PP1,
(5.1.19)
p,p'eYK(Z),
z written with respect to the atlases * s of C and * z of T>. Given the leafwise connection AgT (5.1.14) and the Lie derivative L (5.1.18), one can assign the first order differential operator / = -i[(V#f + ief) ® Id + Id
(5.1.20)
on YK(Z) to each element of the quantum algebra Ap. These operators obey the Dirac condition (5.1.1), and provide a representation of the quantum algebra A? by (unbounded) Hermitian operators in the pre-Hilbert space YK{Z). Finally, this representation is restricted to the subspace E of sections p e YK{Z) which obey the condition (Vj ® Id + Id ® L#)p = ( V j + \di^)p
=0
for all T-subordinate leafwise Hamiltonian vector fields $. However, it may happen that the above quantization has no physical sense because the Hermitian form (5.1.19) on the carrier space E and, consequently, the mean values of operators (5.1.20) are defined by integration over the whole manifold Z. For instance, it implies integration over classical parameters. Therefore, we suggest a different scheme of quantization of symplectic foliations. Let us consider the exterior bundle A TT*, m = dim T. Its structure group GL(m,R) is reducible to the group GL+(m,R) since a symplectic foliation is oriented. One can regard this fibre bundle as being associated to a GL(m, C)-principal bundle P —> Z. As earlier, let us assume that H2(Z;Z2) = 0. Then the principal bundle P admits a two-fold covering principal bundle with the structure metalinear group ML(m,C) [141]. As a consequence, there exists a complex line bundle V? —> Z characterized by an atlas yr = {{Ui;zx;zi;c)}
305
Chapter 5 Geometric Quantization
with the transition functions c' = Sjrc such that
One can think of its sections as being leafwise half-forms on Z. The metalinear bundle Vj: —> Z admits the canonical lift of any T-subordinate vector field T on Z. The corresponding Lie derivative of its sections reads Lf = T% + \diT\
(5.1.21)
We define the quantization bundle as the tensor product Yjr = C® Vjr.
Given a leafwise connection A9T (5.1.14) and the Lie derivative ~LT (5.1.21), let us associate the first order differential operator / = - i [ ( v £ , + t e / ) ® I d + I d ®L£ / ] = -ilV^+ief+^drf)],
(5.1.22)
feAr,
on sections p? of Y? to each element of the quantum algebra A?. A direct computation with respect to the local Darboux coordinates on Z proves the following. LEMMA
5.1.9. The operators (5.1.22) obey the Dirac condition (5.1.1). •
LEMMA 5.1.10. If a section p? fulfils the condition ( V j ® H + Id ® l4)pF
= (Vj + i W A F = 0
(5.1.23)
for all T-subordinate leafwise Hamiltonian vector field •&, then fp? for any / G A?- possesses the same property. • Let us restrict the representation of the quantum algebra A? by the operators (5.1.22) to the subspace E? e YF(Z) of sections p? which obey the condition (5.1.23) and whose restriction to any leaf of T is of compact support. The last condition is motivated by the following. Since i*FTT* = T*F,
306
Geometric and Algebraic Topological Methods in Quantum Mechanics
the pull-back iFV? of V? onto a leaf F is a metalinear bundle of halfforms on F. By virtue of Propositions 5.1.6 and 5.1.8, the pull-back ipY? of the quantization bundle Y? —> Z onto F is a quantization bundle for the symplectic manifold (F, ipd^). Given the pull-back connection AF (5.1.16) and the polarization T> = i*FT, this symplectic manifold is subject to the standard geometric quantization by the first order differential operators / = -i(iF^f
+ ief + i040>),
/ G A F,
(5.1.24)
on sections pF of iFYy? —» F of compact support which obey the condition (»fVj + ldi^)pF
=0
(5.1.25)
for all Tir-subordinate Hamiltonian vector fields $ on F. These sections constitute a pre-Hilbert space EF with respect to the Hermitian form
/ 1\m/2 f
(PF\P'F) = ( ^
)
J PFP'FF
The key point is the following. PROPOSITION
5.1.11. We have i*FEj: C EF, and the relation
iF(fpT) = WTWFP?) holds for all elements / 6 Ap and py? G Ejr.
(5-1-26) D
Outline of proof. One can use the fact that the expressions (5.1.24) and (5.1.25) have the same coordinate form as the expressions (5.1.22) and (5.1.23) where zx =const. QED The relation (5.1.26) enables one to think of the operators / (5.1.22) in Ef as being the leafwise quantization of the 5^(Z)-algebra Ayr in the pre-Hilbert Sy(Z)-module.
5.2
Example. Quantum completely integrable systems
We study geometric quantization of a CIS on a 2m-dimensional symplectic manifold (Z, fi) around a regular compact invariant manifold M. By virtue of the classical Arnold-Liouville theorem [12] (see Theorem 2.3.14 for the
Chapter 5 Geometric Quantization
307
general case of a PIS), a small neighborhood U of M in the ambient symplectic manifold Z is isomorphic to the symplectic annulus U = N x T m , where N c M™ is a non-empty domain and Tm is an m-dimensional torus. The product Nx Tm is equipped with the action-angle coordinates (Ik, (f>k). With respect to these coordinates, the symplectic form on TV x Tm reads Sl = dIkAd(j)k,
(5.2.1)
while a Hamiltonian of a CIS is a function H{Ik) of action variables only. Let us note that there are different approaches to quantization of autonomous CISs [185; 208]. The peculiarity of the geometric quantization procedure is that it remains equivalent under symplectic isomorphisms, but essentially depends on the choice of a polarization [38; 364]. Geometric quantization of CISs around an invariant tori has been studied with respect to polarization spanned by Hamiltonian vector fields of first integrals [329]. In fact, the well-known Simms quantization of the harmonic oscillator is also of this type [l4l]. The problem is that the associated quantum algebra includes functions which are not globally defined and that elements of its carrier space are not smooth sections of a quantization bundle. Indeed, written with respect to the action-angle variables, this quantum algebra consists of functions which are afnne in angle coordinates. Here, we use a different polarization spanned by locally Hamiltonian vector fields of angle variables [177]. It provides the Schrodinger representation of action variables by first order differential operators on functions of angle coordinates. Remark 5.2.1. Since the action-angle coordinates, by definition, are canonical for the symplectic form (5.2.1), geometric quantization of the symplectic annulus (N x T m ,fi) in fact is equivalent to geometric quantization of the cotangent bundle T*Tm = R m x Tm of the torus T m provided with the canonical symplectic form (2.1.3). In particular, the above mentioned angle polarization of N x Tm corresponds to the familiar vertical polarization VT*Tm of T*Tm -> Tm which leads to Schrodinger quantization of the cotangent bundle T*Tm. The associated quantum algebra A of N x Tm consists of functions which are affine in action variables /&. Therefore, applications of Schrodinger geometric quantization are limited by the fact that a Hamiltonian fails to belong to the quantum algebra A, unless it is afnne in momenta. In the case of a CIS, a
308
Geometric and Algebraic Topological Methods in Quantum Mechanics
Hamiltonian H depends only on action variables which mutually commute. Therefore, if Ti(h) is a polynomial function, it is uniquely represented by an element of the enveloping algebra A of the Lie algebra A, and is quantized as the operator H(Ik)- Moreover, this quantization is also extended to Hamiltonians which are analytic functions on R m because the action operators Ik are diagonal. • Remark 5.2.2. One usually mentions a harmonic oscillator as the simplest CIS whose quantization in the action-angle variables looks notoriously difficult because the eigenvalues of its action operator is expected to be lower bounded [209]. However, a harmonic oscillator quantized with respect to action-angle coordinates is defined on the momentum phase space R2 \ {0}, but it is not the standard oscillator on R2. There is a monomorphism, but not an isomorphism of the Poisson algebra of smooth complex functions on R2 to that on R2 \ {0}. Furthermore, the angle polarization on R2 \ {0} is not extended to R2. As a consequence, the quantum algebra associated to this polarization contains functions on R2 \ {0} which are not extended to R2, and the carrier space of this quantization consists of tempered distributions. • We follow the standard geometric quantization procedure in order to quantize the symplectic annulus (N x Tm, O) [401; 438]. Since the symplectic form fi (5.2.1) is exact, the prequantization bundle is a trivial complex line bundle C —> N x Tm. Let its trivialization C^(N
xTm)xC
(5.2.2)
hold fixed. Any other trivialization leads to equivalent quantization of N x Tm. Given the associated bundle coordinates (7fc,^fe,c), c G C, on C (5.2.2), one can treat its sections as smooth complex functions on N x Tm. The Kostant-Souriau prequantization formula associates to each smooth real function / on N x Tm the first order differential operator
f = -iV*,+f
(5.2.3)
on sections of C, where "df (2.1.4) is the Hamiltonian vector field of / and V is the covariant differential with respect to a [/(l)-principal connection A on C whose curvature form obeys the prequantization condition R = id. This connection reads A = A o + iclkd
(5.2.4)
309
Chapter 5 Geometric Quantization
where Ao is a flat [/(l)-principal connection on C —> V x Tm. The classes of gauge conjugate (i.e., linked by vertical automorphisms) flat principal connections on C are indexed by the set M m /Z m of homomorphisms of the de Rham cohomology group
F^JV xTm)
=Rm
of the annulus N x Tm (which equals its first homotopy group) to U(l) (see Example 10.6.4). We choose their representatives of the form A0{(Xk)\
= dlk ®dk + d<j>k g> (dfc + i\kcdc),
Afc e [0,1).
Then the connection A (5.2.4) on C up to gauge conjugation reads A[{Xk)} = dlk ®dk + d<j>k ® (dk + i(h + Xk)cdc).
(5.2.5)
For the sake of simplicity, let Afc in the expression(5.2.5) be arbitrary real numbers, but we will bear in mind that connections >l[(Afc)] and ^4[(A'fc)] with Afc — Afe £ Z are gauge conjugate. Given a connection (5.2.5), the prequantization operators (5.2.3) read f = -i#f
+
(f-(Ik
+
Xk)dkf).
(5.2.6)
It is readily observed that the prequantization operators / also keep their form (5.2.6) on sections of the quantum bundle C
V1/2
-+NxTm
is a metalinear bundle whose sections are half-forms on N x Tm. Let us choose the above mentioned angle polarization. It is the vertical tangent bundle Vix of the fibration •K :
N x Tm -> Tm
spanned by the vectors dk. It is readily observed that the corresponding quantum algebra
A C C°°(N x Tm) consists of afnne functions
f = ak(4>T)h + b{4>T)
(5.2.7)
310
Geometric and Algebraic Topological Methods in Quantum Mechanics
of action coordinates Ik. The carrier space £ of its representation (5.2.6) consists of sections p of the quantization bundle C ® D1/2
-^NxTm
of compact support which obey the condition V^/9 = 0 for any Hamiltonian vector field -d subordinate to the distribution Vn. This condition reads dkfdkp
= 0,
/ e C°°(r m ).
It follows that elements of £ are independent of action variables and, consequently, fail to be of compact support, unless p = 0. This is the well-known problem of Schrodinger geometric quantization. It is solved as follows [39; 174]. Given an imbedding
iT:Tm -> Nx Tm, let CT = ij-C be the pull-back of the prequantization bundle C (5.2.2) onto the torus T m . It is a trivial complex line bundle CT = Tm x C whose sections are smooth complex functions on Tm. Let A[Xk] = iTA[Xk} = d<j>k ® {dk + i(h + \k)cdc) be the pull-back of the connection A[Xk] (5.2.5) onto CT, and let V denote the corresponding covariant differential. Let T>T be a metalinear bundle of complex half-forms on the torus Tm. It admits the canonical lift of any vector field r on T m , and the corresponding Lie derivative of its sections reads
L T = r kdk +
\dkrk.
Let us consider the tensor product Y = CT®/DT->Tm.
(5.2.8)
Since the Hamiltonian vector fields df = akdk - (Irdkar + dkb)dk of functions / (5.2.7) are projectable onto Tm, one can associate to each
Chapter 5 Geometric Quantization
311
element / of the quantum algebra A the first order differential operator / = (-iV w t f / + / ) ® Id + Id ® L Ttf/ =
(5.2.9)
-iafe9fc - -dkak - ak\k + b on sections of Y. A direct computation shows that the operators (5.2.9) obey Dirac's condition (4.2.1). Sections s of the quantization bundle Y _> I'm (5.2.8) constitute a pre-Hilbert space ET with respect to the non-degenerate Hermitian form r
( 1\
(s\s ) = ( — J
m
f
-
I ss,
S,SG£T.
Then / (5.2.9) are Hermitian operators in ET- They provide the desired geometric quantization of a CIS on the annulus N x Tm. Of course, this quantization depends on the choice of a connection ^[(^fc)] (5.2.5) and a metalinear bundle VT [312]. The latter need not be trivial. If T>T is trivial, sections of the quantization bundle Y —> Tm (5.2.8) obey the transformation rule s(0fc + 2TT) = s(4>k) for all indices k. They are naturally complex smooth functions on Tm. By virtue of the multidimensional Fourier theorem, the functions Wr)=exp[2(nr0r)],
(n r ) = ( n i , . . . , n m ) € Z " \
(5.2.10)
constitute an orthonormal basis for the pre-Hilbert space ET = C°°(T m ). The action operators h = -idk - Afc
(5.2.11)
4VVr) = K - Afc)VVr)
(5.2.12)
(5.2.9) are diagonal
with respect to this basis. Other elements of the algebra A are decomposed into the pull-back functions 7r*V'(rir) on N x Tm which act on C°°(T m ) by multiplications T ^ V W ^ K . ) = V>(nr)V>K.) = Anr+K)-
(5.2.13)
312
Geometric and Algebraic Topological Methods in Quantum Mechanics
If T>T is a non-trivial metalinear bundle, sections of the quantization bundle Y -> Tm (5.2.8) obey the transformation rule p r ( ^ + 27r) = -pr(0»)
(5.2.14)
for some indices j . In this case, the orthonormal basis of the pre-Hilbert space £T can be represented by double-valued complex functions iP(nunj)
= e x p [ i ( n ^ + (rij + ^)<jP)]
(5.2.15)
on Tm. They are the eigenvectors Iii>(ni,nj) = (ni ~ Xi)ip(nunj),
^{nunj)
= («j ~ Aj +
^(ni,nj)
of the operators Ik (5.2.11), while the pull-back functions Tr*^^,.) act on the basis (5.2.15) by the above mentioned law (5.2.13). It follows that the representation of the quantum algebra A determined by the connection ^4[(Afc)] (5.2.5) in the space of sections (5.2.14) of a nontrivial quantization bundle Y (5.2.8) is equivalent to its representation determined by the connection A[(Xi,Xj - |)] in the space C°°(T m ) of smooth complex functions on Tm. Therefore, one can restrict the study of representations of the quantum algebra A to its representations in C°°(T m ) associated to different connections (5.2.5). These representations are nonequivalent, unless Xk ~ X'k eZ for all indices k. Given the representation (5.2.9) of the quantum algebra A in C°°(r m ), a polynomial Hamiltonian Ti(Ik) and polynomial first integrals fi(h) of a CIS are uniquely quantized as Hermitian elements •H(h) = W(Jfc),
fi(h) = fi{h)
(5.2.16)
of the enveloping algebra A of A. They have the countable spectra ii{h)il)(nT) = ft(nfc ~ Afc)VVr)>
fi(h)^(nr)
= fi(nk - Afc)VW)-
Let us note that, since the action operators / are diagonal, the quantization formula (5.2.16) can be extended to analytic functions of Ik on R m . 5.3
Quantization of time-dependent mechanics
We study the covariant (frame-independent) geometric quantization of nonrelativistic mechanics, subject to time-dependent transformations (see Sec-
Chapter 5 Geometric Quantization
313
tion 2.4). In contrast with the existent geometric quantizations of nonrelativistic mechanics [401; 438], we do not fix a trivialization Q = RxM,
V*Q = R x T*M.
(5.3.1)
The key point is that, in this case, the evolution equation is not reduced to the Poisson bracket on V*Q, but can be expressed into the Poisson bracket on the cotangent bundle T*Q of Q. Therefore, covariant geometric quantization of time-dependent mechanics on a configuration bundle Q —> R requires compatible geometric quantization both of the cotangent bundle T*Q and the vertical cotangent bundle V*Q of Q. The relation (2.4.28) defines the monomorphism of Poisson algebras C : (C°°(V*Q), {, }v) - (C°°(T*Q), {, } T ).
(5.3.2)
Therefore, a compatibility of geometric quantizations of T*Q and V*Q implies that this monomorphism is prolonged to a monomorphism of quantum algebras of V*Q and T*Q. Of course, it seems natural to quantize C°°(V*Q) as a subalgebra (5.3.2) of the Poisson algebra C°°(T*Q). However, geometric quantization of the Poisson algebra (C°°(T*Q), {, }T) need not imply that of its Poisson sub-
algebra CC°°(V*Q). We show that the standard prequantization of the cotangent bundle T*Q yields the compatible prequantization of the Poisson manifold V*Q such that the monomorphism £* (5.3.2) is prolonged to a monomorphism of prequantum algebras. However, polarization of T*Q need not induce any polarization of V*Q, unless it contains the vertical cotangent bundle V^T*Q of the fibre bundle C (2.4.26) spanned by vectors d°. A unique canonical real polarization of T*Q, satisfying the above condition V(T*Q c T,
(5.3.3)
is the vertical tangent bundle VT*Q of T*Q —> Q. The associated quantum algebra AT consists of functions on T*Q which are affine in momenta p,\We show that this vertical polarization of T*Q yields the polarization of the Poisson manifold V*Q such that the corresponding quantum algebra Ay consists of functions on V*Q which are affine in momenta pk- It follows that Av is a subalgebra of AT under the monomorphism (5.3.2). After metaplectic correction, the compatible Schrodinger representations of AT and Ay by operators on complex half-forms on Q is obtained.
314
Geometric and Algebraic Topological Methods in Quantum Mechanics
The physical relevance of the Schrodinger quantization of T*Q however is open to question. The scalar product of half-forms on Q implies integration over time, though the time plays the role of a classical evolution parameter in quantum mechanics, based on Schrodinger and Heisenberg equations. At the same time, the Schrodinger quantization of V*Q provides instantwise quantization of non-relativistic mechanics. Indeed, a glance at the Poisson bracket (2.4.27) shows that the Poisson algebra C°°(V*Q) is a Lie algebra over the ring C°° (R) of functions of time alone, where algebraic operations in fact are instantwise operations depending on time as a parameter. We show that the Schrodinger quantization of the Poisson manifold V*Q induces geometric quantization of its symplectic fibres Vt*Q, ( £ K , such that the quantum algebra At of V*Q consists of elements / £ Ay restricted to Vt*Q. This agrees with the instantwise quantization of symplectic fibres {t} x T*M of the direct product (5.3.1) in [401]. Moreover, the induced geometric quantization of fibres V*Q, by construction, is determined by their injection to V*Q, but not projection of V*Q. Therefore, it is independent of a trivialization (5.3.1). Let us turn now to quantization of the evolution equation in timedependent mechanics. The problem is that the Poisson structure (2.4.27) fails to provide any dynamic equation on the momentum phase space V*Q because all Hamiltonian vector fields $/ (2.4.29) for functions / on V*Q are vertical. Hamiltonian dynamics of time-dependent mechanics is described as a particular Hamiltonian dynamics on fibre bundles where the evolution equation is brought into the form (2.4.41). The problem is that the covariant Hamiltonian H* (2.4.39) does not belong to the algebra AT, unless it is affine in momenta. Let us assume that H* is a polynomial of momenta. This is the case of all physical models. Then we show below that H* can be represented by a finite sum of products of elements of AT, though this representation by no means is unique. Thereby, it can be quantized as an element of the enveloping algebra AT of the Lie algebra ATRemark 5.3.1. An ambiguity of the operator representation of a classical Hamiltonian is a well-known technical problem of Schrodinger quantization as like as any geometric quantization scheme, where a Hamiltonian does not preserve a polarization (see [40l] for a general, but rather sophisticated analysis of such Hamiltonians). One can include the covariant Hamiltonian Ji* (2.4.39) in a quantum algebra by choosing polarization of T*Q which contains the Hamiltonian vector field of 7i*. This polarization always exists,
315
Chapter 5 Geometric Quantization
but does not satisfy the condition (5.3.3) and, therefore, does not define any polarization of the Poisson manifold V*Q. Let us note that, given a trivialization (5.3.1), symplectic fibres Vt*Q, t e R, of the Poisson bundle V*Q —> R can be provided with the instantwise polarization spanned by vectors {diHd1
- d ^ d i , ••• ,dmndm
-
dmHdm).
However, this polarization need not be regular and, by construction, is frame-dependent. It is a standard polarization in conservative Hamiltonian mechanics of one-dimensional systems, but it requires an exclusive analysis of each physical model. • Given a covariant Hamiltonian H* (2.4.39) and its representative H in AT, the map V:/^{W*,/}T is a derivation of the enveloping algebra Av C AT of the Lie algebra Av • Moreover, this derivation obeys the Leibniz rule V(r/) = dtrf + rV/,
r e C°°(R),
and, consequently, is a connection on the instantwise algebra Ay- Since this property is preserved under quantization, geometric quantization of nonrelativistic time-dependent mechanics leads to its instantwise quantization (see Section 4.6). We start with the standard prequantization of the cotangent bundle T*Q [141; 401; 438]. Since the symplectic form fi on T*Q is exact and, consequently, belongs to the zero de Rham cohomology class, a prequantization bundle is the trivial complex line bundle C = T*Q x C -> T*Q
(5.3.4)
of zero Chern class. Coordinated by (qx,p\, c), this bundle is provided with the admissible linear connection A = dpx ® dx + dqx ® (d\ + ipxcdc)
(5.3.5)
whose strength form equals — ifl. The ^-invariant Hermitian fibre metric on C —> Q is given by the expression (5.1.12). The covariant derivative of sections s of the prequantization bundle C (5.3.4) relative to the connection
316
Geometric and Algebraic Topological Methods in Quantum Mechanics
A (5.3.5) along the vector field u on T*Q reads V u s = ux{dx - ip\)s + uxdxs.
(5.3.6)
Given a function / € C°°(T*Q) and its Hamiltonian vector field
uf = dxfdx - dxf3x,
uf\n = -df,
the covariant derivative (5.3.6) along Uf is VUfs =
dxf(dx-ipx)s-dxfdxs.
Then, in order to satisfy the Dirac condition (4.2.1), one assigns to each function / e C°°(T*Q) the first order differential operator f(s) = - i ( v u / + if)s = [-i{dxfdx - dxfdx) + (/ - pxdxf)]s
(5.3.7)
on sections s of the prequantization bundle C (5.3.4). The prequantum operators (5.3.7) for elements / of the Poisson subalgebra CC°°(V*Q) C C°°(T*Q) read f(s) = [-i(dkfdk - dxfdx) + (/ -
Pkd
k
f)]s.
(5.3.8)
Let us turn now to prequantization of the Poisson manifold (V*Q, {, }v)The Poisson bivector w of the Poisson structure (2.4.27) on V*Q is w = dk Adk = -[w,-d]SN,
(5.3.9)
where [, ]SN is the Schouten-Nijenhuis bracket and "9 = pkdk is the Liouville vector field on the vertical cotangent bundle V*Q —> Q. The relation (5.3.9) shows that the Poisson bivector w is exact and, consequently, has the zero Lichnerowicz-Poisson cohomology class. Therefore, let us consider the trivial complex line bundle Cv = V*Q x C -> V*Q
(5.3.10)
of zero Chern class. Since the line bundles C (5.3.4) and CV (5.3.10) are trivial, C can be seen as the pull-back C*CV of CV, while Cy is isomorphic to the pull-back h*C of C with respect to a section h (2.4.31) of the affine bundle (2.4.26). Since Cy = h*C and since the covariant derivative of the
Chapter 5 Geometric Quantization
317
connection A (5.3.5) along the fibres of £ (2.4.26) is trivial, let us consider the pull-back h*A = dpk
(5.3.11)
of the connection A (5.3.5) onto Cy —> V*Q. This connection defines the contravariant derivative V
(5.3.12)
of sections sy of Cy —> V*Q along one-forms <j> on V*Q. This contravariant derivative corresponds to a contravariant connection Ay on the line bundle Cy —> y*Q [426]. Since the vector fields w^cfi =
= [-i(dkfdk-dkfdk)
+ (f-pkdkf)}sy, (5.3.13)
defines prequantization of the Poisson manifold V*Q. In particular, the prequantum operators of functions r e C°°(R) of time alone are reduced to the multiplication fsy = rsy. Consequently, the prequantum algebra of V*Q inherits the structure of a C°°(R)-algebra. It is immediately observed that the prequantum operator / (5.3.13) coincides with the prequantum operator £*/ (5.3.8) restricted to the pullback sections s = C,*sv- Thus, the above mentioned prequantization of the Poisson algebra C°°(V*Q) is equivalent to its prequantization as a subalgebra of the Poisson algebra C°°(T*Q). Let us note that, since the complex line bundles C (5.3.4) and Cy (5.3.10) are trivial, their sections are simply smooth complex functions on T*Q and V*Q, respectively. Then the prequantum operators (5.3.7) and (5.3.13) can be written in the form / = -*L u / + ( / - L t f / ) ,
(5.3.14)
where i? denotes the Liouville vector field •& = p\dx on T*Q —+ Q or •d=pkdk on V*Q^Q. Given compatible prequantizations of the cotangent bundle T*Q and the vertical cotangent bundle V*Q, let us now construct their compatible polarizations and quantizations. We assume that Q is an oriented manifold and that the cohomology group H2(Q,Z2) is trivial.
318
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us recall that polarization of a Poisson manifold (Z, {, }) is defined as a sheaf T* of germs of complex functions on Z whose stalks T*, z £ 2 , are Abelian algebras with respect to the Poisson bracket {,} [427]. Let T*(Z) be the structure algebra of global sections of the sheaf T*; it is also called a Poisson polarization [425; 426]. A quantum algebra A associated to the Poisson polarization T* is defined as a subalgebra of the Poisson algebra C°°{Z) which consists of functions / such that {f,T*(Z)}cT*(Z). Polarization of a symplectic manifold yields its Poisson one. Let T* be a polarization of the Poisson manifold (T*Q, {, }T)- Its direct image in V*Q with respect to the fibration £ (2.4.26) is polarization of the Poisson manifold (V*Q,{,}v) if the germs of T* are constant along the fibres of C [427], i.e., are germs of functions independent of the momentum coordinate p. It follows that the corresponding symplectic polarization T of T*Q is vertical with respect to the fibration T*Q -> V*Q. The vertical polarization T = VT*Q of T*Q obeys this condition. The associated quantum algebra AT C C°°(T*Q) consists of functions which are affine in momenta p\. The algebra AT acts by operators (5.3.14) on the space of smooth complex functions s on T*Q which fulfill the relation V u s = 0 for any T-valued (i.e, vertical) vector field u = u\dx on the cotangent bundle T*Q —•» Q. Clearly, these functions are the pull-back of complex functions on Q with respect to the fibration T*Q —» Q. Following the general metaplectic technique, we come to complex half-forms on Q which are sections of the complex line bundle P1/2 —> Q with the transition functions c' = Sc such that SS is the Jacobian of coordinate transition functions on Q. Then the formula (5.3.14), where L^ / is the Lie derivative of half-forms, defines the Schrodinger representation of the Lie algebra AT by operators fp = {-iLa,dx f = a\q»)px +
+ b)p = (-iaxdx - l-dxax + b)p,
(5.3.15)
b{q»)£AT,
in the space £>i/2(Q) of complex half-forms p on Q. From now on, we assume that a coordinate atlas of Q and a bundle atlas of 2?i/2 —> Q are defined on the same covering of Q, e.g., by contractible open sets. Let E C T>i/2(Q) consists of half-forms of compact support, and let E be its
Chapter 5 Geometric Quantization
319
completion with respect to the non-degenerate Hermitian form (P|p')=(^)m/w'-
(5-3.16)
The (unbounded) operators (5.3.15) on the domain E in the Hilbert space E are Hermitian. The vertical polarization of T*Q defines the polarization Tv of the Poisson manifold V*Q which contains the germs of functions, constant on the fibres of V*Q —> Q. The associated quantum algebra Av consists of functions on V*Q which are affine in momenta. It is a C°°(R)-algebra. This algebra acts by operators (5.3.14) on the space of smooth complex functions Sy on V*Q which fulfill the relation V u sy = 0 for any vertical vector field u = ukdk on V*Q —» Q. These functions are also the pull-back of complex functions on Q with respect to the fibration V*Q —> Q. Similarly to the case of AT, we obtain the Schrodinger representation of the Lie algebra Av by the operators fp = {-iLakdk
+ b)p = {-iakdk - l-dkak + b)p,
(5.3.17)
on half-forms on Q and in the above Hilbert space E. Moreover, a glance at the expressions (5.3.15) and (5.3.17) shows that (5.3.17) is the representation of Av as a subalgebra of the Lie algebra ATAs was mentioned above, the physical relevance of the space of halfforms on Q with the scalar product (5.3.16) is open to question. At the same time, the representation (5.3.17) preserves the structure of Av as a C°°(K)-algebra. Therefore, let us show that this representation defines the instantwise quantization of AvFirstly, the prequantization (5.3.13) of the Poisson manifold V*Q yields prequantization of its symplectic leaves Vt*Q, t G R, as follows. The symplectic structure on V*Q is nt = (hoityn = dPkAdqk,
(5.3.18)
where h is an arbitrary section of the fibre bundle f (2.4.26) and it : Vt*Q —> V*Q is the natural imbedding. Since w^
320
Geometric and Algebraic Topological Methods in Quantum Mechanics
It is the pull-back At = i*th*A = dpk ®dk + dqk ® (dk + ipkcdc) of the connection h* A (5.3.11) onto the trivial pull-back line bundle i*Cv
= Vt*Q X C H
Vt*Q.
It is readily observed that this connection is admissible for the symplectic structure (5.3.18) on Vt*Q, and provides prequantization of the symplectic manifold (V*Q,Q.t) by the formula j t = -»L* /t + (ft - L*t) = -i(dkft3k
- dkftdk) + (ft~Pkdkft),
(5.3.19)
where 0 / t = dkftdk - dkftdk is the Hamiltonian vector field of a function ft on Vt*Q with respect to the symplectic form Clt (5.3.18). The operators (5.3.19) act on smooth complex functions st on V*Q. In particular, let ft, st and (fs)t be the restriction to V*Q of a real function / and complex functions s and f(s) on V*Q, respectively. We obtain from the formulae (5.3.13) and (5.3.19) that (fs)t = ftstThis equality shows that the prequantization (5.3.13) of the Poisson manifold V*Q is a leafwise prequantization. Let Ty be the above mentioned polarization of the Poisson manifold V*Q. It yields the pull-back polarization TJ1 = i*tT*v of a fibre Vt*Q with respect to the Poisson morphism it : Vt*Q -> VQ. The corresponding distribution T ( coincides with the vertical tangent bundle of the fibre bundle Vt*Q —> Qt. The associated quantum algebra At consists of functions on V*Qt which are affine in momenta. In particular, the restriction to V*Q of any element of the quantum algebra Ay of V*Q obeys this condition and, consequently, belongs At- Conversely, any element of At is of this type. For instance, using a trivialization (5.3.1) and the corresponding surjection nt : V*Q —> V*Q, one can define the pull-back 7Tt*/t of a function ft G At which belongs to the quantum algebra Ay and ft = i*(7Tj/t). Thus, At = i*Av and, therefore, the polarization Tv of the Poisson bundle V*Q —> M. is a fibrewise polarization.
321
Chapter 5 Geometric Quantization
In order to provide metaplectic correction and to complete geometric quantization of symplectic fibres of the Poisson bundle V*Q —> R, one can use the following fact. Any atlas {(£/; t, qk)} of bundle coordinates on the fibre bundle Q —> R induces a coordinate atlas {(Qt fl U; qk)} of its fibre Qt, t G R. Since
the Jacobian J of the transition function between coordinate charts (U;t,qk) and (U';t,q'k) on Q coincides with the Jacobian Jt of the transition function between coordinate charts (Qt fl U;qk) and (Qt H U';q'k) on Qt at points oi QtC\U n U'. It follows that, for any fibre Qt of Q, the pull-back ijP —> Qt of the complex line bundle V —> Q of complex densities on Q with transition functions c' = Jc is the complex line bundle of complex densities on Qt with transition functions Jt = J\Qt- Accordingly, any density L on Q yields the pull-back section Lt = L o it of the line bundle i*X> —> Q t , i.e., Lt is a density on Q t . The pull-back L —> Lt takes the coordinate form L = C(t, qk)dmq A dt -> Lt = £(*, 9fe)dm9|t=const = £(t, ^Jd""?, where {d9fc} are holonomic fibre bases for V*Q. It is maintained under transformations of bundle coordinates on Q. Let now T>\ji —> Q be a complex line bundle of complex half-forms on Q with transition functions S such that 55 = J on J7 D [/'. Its pull-back i*X>i/2 is a complex line bundle over a fibre Qt, t G R, with transition functions St = 5|Q 4 . These transition functions obey the relation StSt = J|Qt = Jt, i.e., i*T>\/2 —* Qt is the fibre bundle of half-forms on Qt. Then the formula (5.3.19) defines the Schrodinger representation of the quantum algebra At of the symplectic fibre Qt by (unbounded) Hermitian operators ft/H = (-iLafd,. + b)Pt = {-iakdk - %^dkak + b)pu
(5.3.20)
ft = ak(j)pk + btf) G Au
(5.3.21)
in the Hilbert space Et which is the completion of the pre-Hilbert space Et
322
Geometric and Algebraic Topological Methods in Quantum Mechanics
of half-forms on Qt of compact support with respect to the scalar product
MM=(-^y j ptptQt
If Qt is compact, the operators (5.3.20) in Et are self-adjoint. Pre-Hilbert spaces Et constitute a trivial bundle ER over R. As in the above case of densities, any half-form p on Q yields the section p o it of the pull-back bundle i*Pi/ 2 —» Qt, i-e, a half-form on Qt. Given an element / £ Av and its pull-back ft = i*f £ At, we obtain from the formulae (5.3.17) and (5.3.20) that fp°H = ft(p°it)This equality shows that the Schrodinger quantization of the Poisson manifold V*Q can be seen as the instantwise quantization. Following this interpretation, let us choose the representation space ER(B.) for Av which consists of complex half-forms p on Q such that, for any ( e l , the halfform poit on Qt is of compact support. It is a C°°(R)-module. The ER(R) is also the carrier space for the Lie algebra AT, but its action on ER(R) is not instantwise. Let us turn now to quantization of the evolution equation (2.4.41). As was mentioned above, the problem is that, in the framework of the Schrodinger quantization, the covariant Hamiltonian 7i* (2.4.39) does not belong to the quantum algebra AT, unless it is affine in momenta. Let us restrict our consideration to the physically relevant case of 7i*, polynomial in momenta. We aim to show that such TC* is decomposed in a finite sum of products of elements of the algebra AT • Let / be a smooth function on T*Q which is a polynomial of momenta p\. A glance at the transformation laws (2.4.22) shows that it is a sum of homogeneous polynomials of fixed degree in momenta. Therefore, it suffices to justify a desired decomposition of an arbitrary homogeneous polynomial F of degree k > 1 on T*Q. We use the fact that the cotangent bundle T*Q admits a finite bundle atlas (see Theorem 10.6.2). Let {C/j}, £ = 1,... ,r, be the corresponding open cover of Q and {/^} a smooth partition of unity subordinate to this cover. Let us put
It is readily observed that {l^} is also a partition of unity subordinate to
Chapter 5 Geometric Quantization
323
{{/,}. Let us consider the local polynomials
{ai-ctfc)
Then we obtain a desired decomposition F
= E ^ = E E ^ ^ r ^ ^ ^ H ^ J • • • ikPak}, (5-3.22)
where all terms ^a? 1 '" afc p ai and /^pa are smooth functions on T*Q. Clearly, the decomposition (5.3.22) by no means is unique. The decomposition (5.3.22) shows that one can associate to a polynomial covariant Hamiltonian 7i* an element H of the enveloping algebra AT of the Lie algebra AT- Let us recall that A consists of finite sums of tensor products of elements of AT modulo the relations /®/'-/'®/-{/,/'}r = 0.
To be more precise, a representative TC belongs to AT + Av, where Av is the enveloping algebra of the Lie algebra Av C AT (see the decomposition (10.8.16) below). The enveloping algebra Av is provided with the antiautomorphism * : /i ® • • • ® fk -> (-l)fc/fc ® • • • ® / l . and one can always make a representative H Hermitian. Since the Dirac condition (4.2.1) holds, the Schrodinger representation of the Lie algebras AT and Av is naturally extended to their enveloping algebras AT and Av, and provides the quantization H* of a covariant Hamiltonian H* • Given an operator Ji*, the bracket V / = t[W*,/]
(5.3.23)
defines a derivation of the quantum algebra Av- Moreover, since p = —idt, the derivation (5.3.23) obeys the Leibniz rule V(r/j = dtrf + r V / ,
r 6 C°°(R).
Therefore, it is a connection on the instantwise algebra Av- In particular, / is parallel with respect to the connection (5.3.23) if [H*,/3 = 0.
(5.3.24)
324
Geometric and Algebraic Topological Methods in Quantum Mechanics
One can think of this equality as being the Heisenberg equation in timedependent mechanics, while the quantum constraint H*p = 0,
p e ER(R),
(5.3.25)
plays a role of the Schrodinger equation. It is readily observed that an operator / is a solution of the Heisenberg equation (5.3.24) if and only if it preserves the subspaces of solutions of the Schrodinger equation (5.3.25).
5.4
Example. Non-adiabatic holonomy operators
We address the Berry phase phenomena (Section 4.7) in a completely integrable system (CIS) of m degrees of freedom around its invariant tori Tm. The reason is that, being constant under an internal evolution, its action variables are driven only by a perturbation holonomy operator A. We construct such an operator for an arbitrary connection on a fibre bundle ExT^E,
(5.4.1)
without any adiabatic approximation [l8l]. In order that a holonomy and a dynamic Hamiltonian mutually commute, we first define a holonomy operator with respect to the initial data action-angle coordinates and, afterwards, return to the original ones. The key point is that both classical evolution of action variables and mean values of quantum action operators relative to original action-angle coordinates are determined in full by the dynamics of initial data action and angle variables. A generic phase space of a Hamiltonian system with time-dependent parameters is a composite fibre bundle
n
v"1 \^ TCP
i TCP
where FI -» S x R is a symplectic bundle and £ x R —> R. is a parameter bundle whose sections are parameter functions [175; 294; 385; 442]. Here, we assume that all the bundles are trivial and, moreover, their trivializations hold fixed. In the case of a CIS with time-dependent parameters, we have the product n = E x £ / = £ x l / x ( R x Tm)
-^ExR^l,
equipped with the coordinates (aa, Ik,t, ipk). Let us suppose for a time that parameters are also dynamic variables. The phase space of such a system
Chapter 5 Geometric Quantization
325
is the product
n' = T*S x u coordinated by (aa, aa, Ik,t, yfc), and the dynamics is characterized by the Hamiltonian form Hx = aadaa + Ikd
-Hs(al3,a0,Ij,t,
Hx=vaZ? + h(A$ + AkaZ?) + H,
(5.4.2)
where H is a function, dt + £f da is a connection on the parameter bundle S x R -> R, and A = dt ® {dt + Afdfc) + d
(5.4.3)
is a connection on S xRxr
-> S x I
[175; 181]. Bearing in mind that aa are parameters, one should choose the Hamiltonian 7is (5.4.2) to be affine in their momenta aa- Then, a Hamiltonian system with a fixed parameter function aa = £ a (t) is characterized by the connection E° = dt£a, and is described by the pull-back Hamiltonian form ff£ = ?HJ: = Ikdvk - (Jfc[A*(t,ipi) +
hka{e,t^)dtia] +
(5.4.4)
n{ejht^))dt
on the Poisson manifold U (2.4.55). Let H = 7i(Ii) be a Hamiltonian of an original autonomous CIS on the toroidal domain U (2.4.55) equipped with the action-angle coordinates (Ik,t,(pk). We introduce a desired holonomy operator by the appropriate choice of the connection A (5.4.3). For this purpose, let us choose the initial data action-angle coordinates (Ik,t,(f>k) by the converse to the canonical transformation (2.4.52). With respect to these coordinates, the Hamiltonian of an original CIS vanishes and the Hamiltonian form (5.4.4) reads Hz = Ikd4>k - Ik[Ak(t,
(5.4.5)
Let us put A* = 0 by the choice of a reference frame associated to the initial data coordinates (j>k, and let us assume that coefficients A£ are independent
326
Geometric and Algebraic Topological Methods in Quantum Mechanics
of time, i.e., the part A E = daa ® (da + Akadk)
(5.4.6)
of the connection A (5.4.3) is a connection on the fibre bundle (5.4.1). Then, the Hamiltonian form (5.4.5) reads Hz = Ikd
4P)dt(,adt.
(5.4.7)
Its Hamilton vector field (2.4.34) is lH
= dt + Aidt^di - Ikdikkadted\
(5.4.8)
and it leads to the Hamilton equations
dt<^ = A ^ ( 0 , < m r , k
l
(5.4.9)
dth = -IkdiA a^(t),
(5.4.10)
V*AE = daa ® (da + A^di - hdiKa&)
(5.4.11)
Let us note that
is the lift of the connection A s (5.4.6) onto the fibre bundle S x (V x Tm) -> S, seen as a subbundle of the vertical cotangent bundle V*{E x Tm) = E x T*Tm of the fibre bundle (5.4.1). It follows that any solution /»(£), 4>l(t) of the Hamilton equations (5.4.9) - (5.4.10) (i.e., an integral curve of the Hamilton vector field (5.4.8)) is a horizontal lift of the curve £(i) C £ with respect to the connection V*A-£ (5.4.11), i.e.,
Thus, the right-hand side of the Hamilton equations (5.4.9) - (5.4.10) is the holonomy operator A = (AadtSa, -IkdiAadtta)-
(5.4.12)
It is not a linear operator, but the substitution of a solution >(£(£)) of the equation (5.4.9) into the Hamilton equation (5.4.10) results in a linear holonomy operator on the action variables Ii. Let us show that the holonomy operator (5.4.12) is well defined. Since any vector field $ on R x Tm such that $J dt = 1 is complete, the Hamilton
Chapter 5 Geometric Quantization
327
equation (5.4.9) has solutions for any parameter function £(£). It follows that any connection As (5.4.6) on the fibre bundle (5.4.1) is an Ehresmann connection, and so is its lift (5.4.11). Let us recall that a connection on a fibre bundle Y —> X is said to be an Ehresmann connection if, given an arbitrary smooth curve £([0,1]) C X, there exists its horizontal lift through any point of Y over £(0). Because V*A^ (5.4.11) is an Ehresmann connection, any curve £([0, l ] ) c S can play a role of the parameter function in the holonomy operator A (5.4.12). Now, let us return to the original action-angle coordinates (Ik,t,ipk) by means of the canonical transformation (2.4.52). The perturbed Hamiltonian reads
H' = /*A* (£(*), ¥>* -
t&nitfdteit)+n{ii),
while the Hamilton equations (5.4.9) - (5.4.10) take the form dtp* = d'Wj) + A\,m,
him),
¥>*(*) =
l
where /*(£(*)), (p (^(t)) is a solution of the Hamilton equations (5.4.9) (5.4.10). We observe that the action variables Ik are driven only by the holonomy operator, while the angle variables ipl have a non-geometric summand. Let us emphasize that, in the construction of the holonomy operator (5.4.12), we did not impose any restriction on the connection AE (5.4.6). Therefore, any connection on the fibre bundle (5.4.1) generates a holonomy operator in a CIS. However, a glance at the expression (5.4.12) shows that this operator becomes zero on action variables if all coefficients A^ of the connection A s (5.4.6) are constant, i.e., A s is a principal connection on the fibre bundle (5.4.1) seen as a principal bundle with the structure group i-pm
In order to quantize a time-dependent CIS on the Poisson toroidal domain (U,{,}v) (2.4.55) equipped with action-angle coordinates (/*,*,/), one may follow the instantwise geometric quantization of time-dependent mechanics (see Section 5.3). As a result, we can simply replace functions
328
Geometric and Algebraic Topological Methods in Quantum Mechanics
on Tm with those on R x Tm [155]. Namely, the corresponding quantum algebra A C C°°(U) consists of affine functions f = ak{t,
(5.4.13)
of action coordinates Ik represented by the operators (5.2.9) in the space E = C°°(R x Tm)
(5.4.14)
of smooth complex functions ip(t,ip) on R x Tm. This space is provided with the structure of the C°°(R)-module endowed with the non-degenerate C°°(R)-bilinear form
W > = (JA I W'dm
rp,i>' G C°°(R x Tm).
We call it a pre-Hilbertian module. Its basis consists of the pull-back onto R x Tm of the functions V(«r) (5.2.10). Furthermore, this quantization of a time-dependent CIS on the Poisson manifold (U, {, }y) is extended to the associated homogeneous CIS on the symplectic annulus (U',0.) (2.4.53) by means of the operator Io — —idt in the pre-Hilbertian module E (5.4.14). Accordingly, the homogeneous Hamiltonian H is quantized as H = -idt + U. It is a Hamiltonian of a quantum time-dependent CIS. The corresponding Schrodinger equation is H-0 = -idtip + Hip = 0,
(5.4.15)
ip£E.
For instance, the quantum Hamiltonian of the original autonomous CIS seen as the time-dependent one is H = -idt + H(Tj). Its spectrum Hip{nr)
=
E(nr)il>lnr)
on the basis {^(nr)} for E (5.4.14) coincides with that of the autonomous Hamiltonian (5.2.16). The Schrodinger equation (5.4.15) reads H^ = -idtip + H(-idk - Xk)tp = 0,
ipGE.
329
Chapter 5 Geometric Quantization
Its solutions are the Fourier series V" = J2BM (nr)
s
exp[-t*S(nr)]V(nr),
(rv) € C.
Now, let us quantize this CIS with respect to the initial data actionangle coordinates (7j, <j>%). As was mentioned above, it is given on a toroidal domain U (2.4.55) provided with another fibration over R. Its quantum algebra A$ C C°°(U) consists of afnne functions f = ak(t,<j>j)Ik + b(t,
(5.4.16)
The canonical transformation (2.4.52) ensures an isomorphism of Poisson algebras A and AQ. Functions / (5.4.16) are represented by the operators / (5.2.9) in the pre-Hilbertian module EQ of smooth complex functions #(£, ^ ) o n R x r . Given its basis
the operators h and tp(nr) take the form (5.2.12) and (5.2.13), respectively. The Hamiltonian of a quantum CIS with respect to the initial data variables is H o = —idt. Then one easily obtains the isometric isomorphism R(1>M) = exp[tt£(nr)]tf ( n r ) ,
(fl(V) W ) > = W > ,
(5-4.17)
of the pre-Hilbertian modules E and -Bo which provides the equivalence
Ti = R'1TiR,
£ (nr) = j r
1
*^,
U = R-1HOR
(5.4.18)
of the quantizations of a CIS with respect to the original and initial data action-angle variables. In view of the isomorphism (5.4.18), let us first construct a holonomy operator in a quantum CIS (Ao, H o ) with respect to the initial data actionangle coordinates. Let us consider the perturbed homogeneous Hamiltonian H S = H 0 + H! = /„ + dtC(t)Aka (£(*), 4?)Ik of the classical perturbed CIS (5.4.7). Its perturbation term Hi is of the form (5.4.13) and, therefore, is quantized by the operator Hi = -idtZa&a = -idtCl^adk
+ \dk(Aka) - iAfeA*].
330
Geometric and Algebraic Topological Methods in Quantum Mechanics
The quantum Hamiltonian Hf = H 0 + Hi defines the Schrodinger equation 3ttf + dt£a[Akadk + \dk{Kka) - iAfeA*]tf = 0.
(5.4.19)
If its solution exists, it can be written by means of the evolution operator U{t) which reduces to the geometric factor •
t
Ui(t) = Texp i Idt'Za{t')Ka{t')dt' . o
.
(5.4.20)
The latter can be viewed as a displacement operator along the curve £[0, l ] c S with respect to the connection Ax=daa(da + Aa)
(5.4.21)
m
on the C°°(S)-module C°°(ExT ) of smooth complex functions o n S x T " 1 (see Section 4.6). Let us study weather this displacement operator exists. Given a connection As (5.4.6), let $%(t,
dtp-yit,*) = - a t f AiCU*-1)*^)) = -dtFAfaftdkp-1)^,*). Let ^o be an arbitrary complex half-form \Po on Tm possessing identical transition functions, and let the same symbol stand for its pull-back onto R x Tm. Given its pull-back ($-!)•*o = det (^Qjr-)
MQ'Ht,4)),
(5-4-22)
it is readily observed that $ = (S-Y^oexptiAfc^]
(5.4.23)
obeys the Schrodinger equation (5.4.19) with the initial data ^o- Because of the multiplier exp[iAfc<^fc], the function ^ (5.4.23) however is ill defined, unless all numbers Afe equal 0 or ±1/2. Let us note that, if some numbers Afc are equal to ±1/2, then ^o exp[*Ajt>fc] is a half-density on T m whose
331
Chapter 5 Geometric Quantization
transition functions equal ±1, i.e., it is a section of a non-trivial metalinear bundle over Tm. Thus, we observe that, if A^ equal 0 or ±1/2, then the displacement operator always exists and A = iHi is a holonomy operator. Because of the action law (5.2.13), it is essentially infinite-dimensional. For instance, let As (5.4.6) be the above mentioned principal connection, i.e., A^ =const. Then the Schrodinger equation (5.4.19) where A^ = 0 takes the form dt*(t,
*(t,^) = w - (rw-no))Ai). The corresponding evolution operator U(t) reduces to Berry's phase multiplier Ui*{nr) = e x p [ - m , ( r (t) - C(0))A£)]tf(np)>
nd G (nr).
It keeps the eigenvectors of the action operators Jj. In order to return to the original action-angle variables, one can employ the morphism R (5.4.17). The corresponding Hamiltonian reads H=
R-^R.
The key point is that, due to the relation (5.4.18), the action operators Ii have the same mean values
(7fcV#> = (4*1*),
* =%),
with respect both to the original and the initial data action-angle variables. Therefore, these mean values are denned only by the holonomy operator. In conclusion, let us note that, since action variables are driven only by a holonomy operator, one can use this operator in order to perform a dynamic transition between classical solutions or quantum states of an unperturbed CIS by an appropriate choice of a parameter function £. The key point is that this transition can take an arbitrary short time because we are entirely free with time parametrization of £ and can choose it quickly changing, in contrast with slowly varying parameter functions in adiabatic models. This fact makes non-adiabatic holonomy operators in CISs promising for several applications, e.g., quantum control and quantum computation. It also looks
332
Geometric and Algebraic Topological Methods in Quantum Mechanics
attractive that quantum holonomy operators in CISs are essentially infinitedimensional, whereas both the existent quantum control theory [119; 393] and the theory of quantum information and computation [240] involve only finite-dimensional operators. 5.5
Geometric quantization of constrained systems
We start with autonomous constrained systems. Let (Z, Q) be a symplectic manifold and ijv : N —» Z its closed imbedded submanifold such that the presymplectic form i^fi on TV is non-zero. We assume that N is a final constraint space and H. an admissible Hamiltonian on Z. In this case, the constrained Hamiltonian system (Z, Cl, H, N) is equivalent to the Dirac constrained system (N,i*NQ,i*NH). Therefore, it seems natural to quantize a symplectic manifold (Z, Cl) and, afterwards, replace classical constraints with the quantum ones. In algebraic quantum theory, quantum constraints are described as follows [203; 204]. Let £ be a Hilbert space and H € B(E) a Hermitian operator in E. By a quantum constraint is meant the condition He = 0,
eeE.
(5.5.1)
A Hermitian operator H defines the unitary operator exp(iW). Then the quantum constraint (5.5.1) is equivalent to the condition exp(iH)e = e. In a general setting, let A be a unital C*-algebra and 1 some subset of its unitary elements called state conditions. Let Sj denote a set of states / of A such that f(a) = 1 for all a eX. They are called Dirac states. One has proved that / G Sx if and only if f(ba) = f(ab) = f(b) for any a £ 1 and b £ A [203]. In particular, if / £ Sj, it follows at once from the relation (3.1.18) that |/(6(a - 1))|2 < f(bb*)f((a - l)(a* - 1)) = 0. One can similarly show that, if a, a' £ 2 and / £ Sj, then / ( ( O - l ) ( a ' - l ) ) = 0.
Chapter 5 Geometric Quantization
333
Thereby, elements a—I, a Gl, generate an algebra which belongs to Ker / for any / G Sj. The completion of this algebra in A is a C*-algebra A% such that f(a) = 0 for all a e Ax and / g l The following theorem provides the important criterion of the existence of Dirac states [203]. THEOREM 5.5.1. The set of Dirac states S% is not empty if and only if D 1 g AT. Let us return to quantization of constrained systems. In a general setting, one studies geometric quantization of a presymplectic manifold via its symplectic realization. There are the following two variants of this quantization [14; 40; 187]. (i) Let (N, u>) be a presymplectic manifold. There exists its imbedding iN:N^Z
(5.5.2)
into a symplectic manifold (Z, Q) such that u> = iNQ. This imbedding is not unique and different symplectic realizations (Z,Q.) of (N,w) fail to be isomorphic. They lead to non-equivalent quantizations of a presymplectic manifold (N,u>). For instance, if a presymplectic form ui is of constant rank, one can quantize a presymplectic manifold (N,u) via its canonical coisotropic imbedding in Proposition 2.1.4 [187]. Geometric quantization of (N,u>) via its imbedding into T*N has been studied in [40]. Given an imbedding i^ (5.5.2), we have a constrained system where classical constraints are smooth functions on Z vanishing on TV. They constitute an ideal IN of the associative ring C°°(Z). Then one usually attempts to provide geometric quantization of a symplectic manifold (Z, Q) in the presence of quantum constraints, but meets the problem how to associate quantum constraints to the classical ones. • Firstly, prequantization procedure / H-> / does not preserve the associative multiplication of functions. Consequently, prequantization IN of the ideal IN of classical constraints fails to be an ideal in a prequantum algebra, i.e., if / e IN then / ' / e IN for any / ' e C°°(Z), but / ' / £ ?N in general. Therefore, one has to choose some set of constraints <j>i,...
334
Geometric and Algebraic Topological Methods in Quantum Mechanics
• Secondly, given a set {fa} of classical constraints, one should choose a compatible polarization of the symplectic manifold (Z, Q) such that prequantum operators fa belong to the quantum algebra. Different sets of constraints imply different compatible polarizations in general. Moreover, a compatible polarization need not exist. (ii) If a presymplectic form u is of constant rank and its characteristic foliation is simple, there is a different symplectic realization (P, Q) of (N, u>) via a fibration N —> P (see Proposition 2.1.5 and Example 2.5.2). Then the reduced symplectic manifold (P, fi) is quantized [14]. Let us apply the above mentioned quantization procedures (i) - (ii) to the Poisson manifold (V*Q, {, }y) in Section 5.3. A glance at the equation (2.4.33) shows that one can think of the vector field 7# as being the Hamiltonian vector field of a zero Hamiltonian with respect to the presymplectic form dH on V*Q. Therefore, one can examine quantization of the presymplectic manifold (V*Q,dH). Given a trivialization (5.3.1), this quantization has been studied in [438]. (i) We use the fact that the range Nh = h(V*Q) of any section h (2.4.31) is a one-codimensional imbedded submanifold and, consequently, is coisotropic. It is given by the constraint H*=p + H(t,qk,Pk)=0. Then the geometric quantization of the presymplectic manifold (V*Q,dH) consists in geometric quantization of the cotangent bundle T*Q and setting the quantum constraint condition n*ip = 0
(5.5.3)
on physically admissible quantum states. It serves as the Shrodinger equation. The condition (5.5.3) implies that, in contrast with geometric quantization in Section 5.3, the Hamiltonian H* always belongs to the quantum algebra of T*Q. This takes place if one use polarization of T*Q which contains the Hamiltonian vector field uH. =dt + dkHdk - dk7idk.
(5.5.4)
Such a polarization of T*Q always exists. Indeed, any section h (2.4.31) of the afHne bundle £ (2.4.26) defines the splitting axdx = ak{dk - dknd°) + (o0 + akdkH)d°
335
Chapter 5 Geometric Quantization
of the vertical tangent bundle VT*Q of T*Q -> Q. Then elements (dk — dkHd°) together with the Hamiltonian vector field (5.5.4) span a polarization of T*Q. Clearly, this polarization does not satisfy the condition (5.3.3), and does not define any polarization of the Poisson manifold V*Q. (ii) In application to (V*Q, dH), the reduction procedure leads to quantization along classical solutions as follows. The kernel of dH is spanned by the vector field 7# and, consequently, the presymplectic form dH is of constant rank. Its characteristic foliation is made up by integral curves of this vector field, i.e., solutions of Hamilton equations. If the vector field 7/f is complete, this foliation is simple, i.e., is a fibration of V*Q over a symplectic manifold N of initial values. In this case, we come to the instantwise quantization when functions on V*Q at a given instant t G R are quantized as functions on N.
5.6
Example. Quantum relativistic mechanics
Quantum relativistic Hamiltonian mechanics exemplifies geometric quantization of a constrained system [391]. A configuration space of relativistic mechanics is an oriented pseudoRiemannian manifold (Q,g), coordinated by (qx). The space of relativistic velocities of relativistic mechanics on Q is the tangent bundle TQ of Q equipped with the induced coordinates (qx,qx) with respect to the holonomic frames {d\}. Relativistic motion is located in the subbundle Wg of hyperboloids 9»u{qWqv
- 1 = 0
(5.6.1)
of TQ. It is described by a second order dynamic equation gA = 3 V , 9 " )
(5-6.2)
on Q which preserves the subbundle (5.6.1), i.e.,
(qxdx + E ^ X s ^ T - 1) = 0,
dx=
d/dq\
This condition holds if the right-hand side of the equation (5.6.2) takes the form
336
Geometric and Algebraic Topological Methods in Quantum Mechanics
where {,/„} axe Christoffel symbols of a metric g, while Fx obey the relation g^vF^q" — 0. In particular, if the dynamic equation (5.6.2) is a geodesic equation qx = Kxq» with respect to a (non-linear) connection K = dqx® (dx +
K^)
on the tangent bundle TQ —> Q, this connections splits into the sum ^
= {A}«"+fl%
(5-6-3)
of the Levi-Civita connection of g and a soldering form F = ^"F^dq"
® dx,
F^ = -FVil.
The momentum phase space of relativistic mechanics on Q is the cotangent bundle T*Q provided with the canonical symplectic form n = dpxAdqx.
(5.6.4)
Let us note that one also considers another symplectic form Q + F where F is the strength of an electromagnetic field [401]. A relativistic Hamiltonian is defined as follows [294; 380; 384]. Let H be a smooth real function on T*Q such that the morphism H : T*Q -> TQ,
(f o H = d»H,
(5.6.5)
is a bundle isomorphism. Then the inverse image
of the subbundle of hyperboloids Wg (5.6.1) is a one-codimensional (consequently, coisotropic) closed imbedded subbundle N oiT*Q given by the constraint HT = g^d^Hff'H - 1 = 0.
(5.6.6)
We say that H is a relativistic Hamiltonian if the Poisson bracket {H, HT} vanishes on N. This means that the Hamiltonian vector field 7 = dxHdx - dxHdx
(5.6.7)
337
Chapter 5 Geometric Quantization
of H preserves the constraint N and, restricted to N, it obeys the Hamilton equation (5.6.8)
•y\QN+i*NdH = 0
of a Dirac constrained system on N with a Hamiltonian H. The morphism (5.6.5) sends the vector field 7 (5.6.7) onto the vector field 7T
= qxdx + {dfiH8xdliH - d^Hdxd^H)dx
on TQ. This vector field defines the second order dynamic equation qx = dfiHdxdtlH - d^Hdxd"H
(5.6.9)
on Q which preserves the subbundle of hyperboloids (5.6.1). Example 5.6.1. The following is a basic example of relativistic Hamiltonian systems. Let us put
# = ^
Tit
-9au{p»-K),
Pa = -^-dag^(Pll
- bjfrv - K) + - < T ( P M -
bJdcK.
The corresponding second order dynamic equation (5.6.9) on Q is 9A = { / A M V - ^9XuF^, { M M = —j9Xfi{dn9pv + 9^90^ - dpg^), Ffj,u = dpbi, — dvb^.
It is a geodesic equation with respect to the affine connection K
of type (5.6.3).
n - U u)q - —9 tp.v flC
(5.6.10)
338
Geometric and Algebraic Topological Methods in Quantum Mechanics
For instance, let g be a metric gravitational field and let b^ = eA^, where A^ is an electromagnetic potential whose gauge holds fixed. Then the equation (5.6.10) is the well-known equation of motion of a relativistic massive charge in the presence of these fields. D Let us turn now to quantization of relativistic mechanics [391]. We follow the standard geometric quantization of a cotangent bundle [39; 401; 438]. Because the canonical symplectic form fi (5.6.4) on T*Q is exact, the prequantum bundle is denned as a trivial complex line bundle C over T*Q. Let us note that this bundle need no metaplectic correction since T*X is endowed with canonical coordinates for the symplectic form Q. Thus, C is a quantum bundle. Let its trivialization C^T*QxC
(5.6.11)
x
hold fixed, and let (q ,p\,c), c £ C, be the associated bundle coordinates. Then one can treat sections of C (5.6.11) as smooth complex functions on T*Q. Let us note that another trivialization of C leads to an equivalent quantization of T*Q. The Kostant-Souriau prequantization formula associates to each smooth real function / G C°°(T*Q) on T*Q the first order differential operator f = -iVtf+f
(5.6.12)
on sections of C, where
0, = dxfdx - dxfdx is the Hamiltonian vector field of / and V is the covariant differential with respect to a suitable [/(l)-principal connection A on C. This connection preserves the Hermitian metric g(c, c') (5.1.12) on C, and its curvature form obeys the prequantization condition R = i£l. For the sake of simplicity, let us assume that Q and, consequently, T*Q is simply connected. Then the connection A up to gauge transformations is A = dpx ® dx + dqx ® (dx + icpxdc),
(5.6.13)
and the prequantization operators (5.6.12) read f = -Wf + {f-pxdxf).
(5.6.14)
Let us choose the vertical polarization on T*Q. It is the vertical tangent bundle VT*Q of the fibration n : T*Q —> Q. The corresponding quantum
Chapter 5 Geometric Quantization
339
algebra .4 C C°°(T*Q) consists of affine functions of momenta (5-6.15)
f = a\q*)px + b(q*) on T*Q. They are represented by the Schrodinger operators / = -iax6x - %-dxax - -^axdxH-g)
+ b,
g = det(ga0),
(5.6.16)
in the space E of sections p of the quantum bundle C of compact support which obey the condition V^p = 0 for any vertical Hamiltonian vector field •d on T*Q. This condition takes the form dxfdxp = Q,
f£C°°{Q).
It follows that elements of E are independent of momenta and, consequently, fail to be compactly supported, unless p = 0. This is the above mentioned problem of Schrodinger quantization which is solved as follows. Let IQ : Q —> T*Q be the canonical zero section of the cotangent bundle T*Q. Let CQ = i*QC be the pull-back of the bundle C (5.6.11) over Q. It is a trivial complex line bundle CQ = Q x C provided with the pull-back Hermitian metric g(c, c') = cc' and the pull-back AQ = i*QA = dqx ® (dx + icpxdc) of the connection A (5.6.13) on C. Sections of CQ are smooth complex functions on Q, but this bundle need metaplectic correction. Let the cohomology group H2(Q; Z2) of Q be trivial. Then a metalinear bundle V of complex half-forms on Q is defined. It admits the canonical lift of any vector field r on Q such that the corresponding Lie derivative of its sections reads LT = TX8x + \dXTX. Let us consider the tensor product Y — CQ
340
Geometric and Algebraic Topological Methods in Quantum Mechanics
of functions / (5.6.15) project onto Q, one can assign to each element / of the quantum algebra A the first order differential operator / = (-iV**, + /) ® Id + Id
PQ = (-s) v V, where ip are smooth complex functions on Q. Then the quantum algebra A can be represented by the operators / (5.6.16) in the space C°°(Q) of these functions. It is easily justified that these operators obey the Dirac condition
lf,f'} = -i{fJ7}. Remark 5.6.2. One usually considers the subspace EQ C C°°(Q) of functions of compact support. It is a pre-Hilbert space with respect to the non-degenerate Hermitian form
(V|V') = /#'(-) 1/2 ^ m+ V Q
It is readily observed that / (5.6.16) are symmetric operators / = /* in EQ, i.e.,
(/V#'> = WV/>. In relativistic mechanics, the space EQ however gets no physical meaning.
•
Let us note that the function HT (5.6.6) need not belong to the quantum algebra A. Nevertheless, one can show that, if HT is a polynomial of momenta of degree k, it can be represented as a finite composition HT=Y^fii---fki i
(5-6.17)
341
Chapter 5 Geometric Quantization
of products of affine functions (5.6.15), i.e., as an element of the enveloping algebra A of the Lie algebra A [174]. Then it is quantized 5Z/ii---/fci
HT^HT=
(5.6.18)
i
as an element of A. However, the representation (5.6.17) and, consequently, the quantization (5.6.18) fail to be unique. The quantum constraint HTip = 0.
(5.6.19)
serves as a relativistic quantum equation. Example 5.6.3. Let us consider a massive relativistic charge in Example 5.6.1 whose relativistic Hamiltonian is H=
±g»'(jpli-eAll)(pv-eAv).
It defines the constraint HT = - ^ " " ( P M - eAM)(jpv
- eAv) - 1 = 0 .
(5.6.20)
Let us represent the function H? (5.6.20) as the symmetric product HT
=^
^
• (pM - eA.) • H O V4 • « T . {-g) V4 .
{pv - eAv) • ( ~ g ) ~ 1 / 4 - 1 m of affine functions of momenta. It is quantized by the rule (5.6.18), where {-9)1/4odao(-g)-V* = -ida. Then the well-known relativistic quantum equation (-
- ieAu) + m2)^ = 0. •
342
5.7
Geometric and Algebraic Topological Methods in Quantum Mechanics
Geometric quantization of holomorphic manifolds
This Section addresses geometric quantization of holomorphic manifolds. The key point is that holomorphic manifolds admit the canonical complex polarization. Let Z be an m-dimensional holomorphic manifold and C - > Z a complex line bundle equipped with complex coordinates (c, z*). It means that the coordinate transition functions on Z obey the condition dzH _ dzH
_
Let us assume that the line bundle C admits a linear connection r = dzi ® (di + Ticdc) + dzi ® (di - Ticdc) with the curvature form R = iQ, where f2 is a non-degenerate real 2-form on Z. It implies that the manifold Z is provided with a symplectic structure, given by the symplectic form Cl, and with a F-invariant Hermitian fibre metric g. With respect to the local complex canonical coordinates, this symplectic form reads n = 2i ^dzj i
Adzj,
while the Hermitian metric g is given by the expression (5.1.12). Therefore, geometric quantization of a holomorphic manifold is similar to that of a real symplectic manifold if complex and real canonical coordinates are connected with each other by the relations ^ = l^Vi+W)The holomorphic manifold Z has the canonical polarization T = {v e TZ : Jv = -iv}, whose sections i? e TQ are complex vector fields d = i? di on Z (see Section 2.6 for the notation). As in the case of a real symplectic manifold, let us consider the subalgebra AT of the Poisson algebra A{Z) which consists of smooth complex functions / on Z such that [i?/,Tf,]cr fl ,
Chapter 5 Geometric Quantization
343
where i
is the Hamiltonian vector field for a function / . It is readily observed that holomorphic functions on Z always belong to Aria order to construct the carrier space of the algebra AT, we take the sections s of the complex line bundle C —» Z such that V,?s = 0,
i? G Tn,
i.e., {di+Ti)3 = 0, and the holomorphic sections P = Pi.^mdz1 A • • • A dzm
of the exterior product m
A TZi )0 -> z.
Let g
-
s
(g, p — spx...mdzx A • • • A dzm
(5.7.1)
be sections of the tensor product C®ATZioz Then the operator f(s ®p) = -i[(V tf/ + z/)(s) ® /o + s ® Ltf//o]
(5.7.2)
can be assigned to any function / G AT- With respect to local complex canonical coordinates, this operator reads
?{s®p) = \( Y^dkfidk -Tk + if)s) ® p + \s ® ( J2 dk(pdkf)\ . The operators (5.7.2) obey Dirac's condition (4.2.1).
344
Geometric and Algebraic Topological Methods in Quantum Mechanics
Sections (5.7.1) of compact support form a pre-Hilbert space Sz with respect to the Hermitian form i™2 f g(si,s2)pip2 {si ® pi\s2 ® p2) = — z
(5.7.3)
The operators / (5.7.2) fulfill the relation fSz C Sz, and they are Hermitian in SzThere is the standard procedure of geometric quantization of a holomorphic manifold Z which admits an imbedding to a projective Hilbert space. Let Z be an m-dimensional holomorphic manifold provided with a symplectic form fi and C —> Z a, complex line bundle over Z with a Hilbert space E realized as a vector subspace of the space C of smooth sections of the dual C —> Z to the complex conjugate C of C —> Z. Given a trivialization chart Ua of C —> Z, let us assume that, for each section ip £ E, the evaluation functionals ea,z{ija) = ipa{z) : E ^ C are non-zero and continuous, i.e.,
hM*)l < Ak.zllV'll,
Ma,zeR.
In accordance with the Riesz theorem, there exists non-zero Ka{z) e E such that
[345]. Then we have the smooth map K : Z 9 Z H
[Ka(z)\ € PE,
where [Ka(;z)] denotes the one-dimensional subspace spanned by Ka(z) ^ 0 and PE is the projective Hilbert space in Example 4.1.5. It is readily observed that the definition of K. is independent of a trivialization chart. Since a projective Hilbert space admits the canonical complex line bundle (see (5.7.4) below), we have the following standard procedure of its geometric quantization [4]. Let be E a Hilbert space and PE the corresponding projective Hilbert space. For the sake of simplicity, E is assumed to be separable. Let us
345
Chapter 5 Geometric Quantization
consider the fibre bundle C = {(z, z) € E x PE : z G z}
(5.7.4)
together with the corresponding projections C : C -» PE,
K:C-+E.
It is called the universal line bundle. There is its subbundle C o = E \ { 0 } - > PE,
n : z ^ z ,
with the typical fibre C \ {0}. Let {Uh,iph) be a chart (4.1.27) of the projective Hilbert space PE. Recall that its image in E is the subspace Eh (4.1.28). We provide this chart with the following coordinate system. Let us take the orthonormal basis {h, efc}, k = 1,..., dimE — 1, for the Hilbert space E. Then {e^} is an orthonormal basis for its subspace Eh, while the coordinates zfc with respect to this basis are the coordinates on the above mentioned chart of the projective Hilbert space PE. The projective Hilbert space PE is provided with the metric Fubini-Studi (4.1.32) and with the corresponding Kahler form u (4.1.33). The universal line bundle (5.7.4) has the connection T = dz>® (dj + j ^ j p ^ c ) <& ® (dj - YVWCd)
'
which preserves the Hermitian fibre metric (5.1.12) in C and satisfies the condition R = iui. Now let Z be a holomorphic manifold and K. its imbedding into the projective Hilbert space PE. The above manifested standard procedure of geometric quantization of Z consists in the following. Let us propose that the pull-back form K.*u> on Z is non-degenerate, i.e., this is a symplectic form. Then one can consider the pull-back line bundle fC*C over Z. This line bundle admits the pull-back connection K.*T with the curvature form iK,*ui, i.e., K,*C is a prequantization bundle.
Chapter 6
Supergeometry
Supergeometry is an ingredient in many quantum models, e.g., SUSY mechanics and BRST formalism. Supergeometry is phrased in terms of Z2graded modules and sheaves over Z2-graded commutative algebras. Their algebraic properties naturally generalize those of modules and sheaves over commutative algebras, but this is not a particular case of non-commutative geometry because of the peculiar definition of graded derivations. Here, we restrict our consideration to geometry of graded manifolds. They are not supermanifolds, though every graded manifold determines a DeWitt JY^-supermanifold, and vice versa [21] (see Theorem 6.9.5 below). 6.1
Graded tensor calculus
Unless otherwise stated, by a graded structure throughout this Chapter is meant a Z2-graded structure, and the symbol [.] stands for the Z2-graded parity. Let us recall some basic notions of the graded tensor calculus [21; 102]. An algebra A is called graded if it is endowed with a grading automorphism 7 such that 7 2 = Id. A graded algebra seen as a Z-module falls into the direct sum A = Ao © Ai of two Z-modules AQ and A\ of even and odd elements such that 7 (a)
= (-l) i a,
aeAu
t = 0,l.
One calls Ao and A\ the even and odd parts of A, respectively. In particular, if 7 = Id, then A = Ao- Since -y(aa') = 7(0)7(0'), 347
348
Geometric and Algebraic Topological Methods in Quantum Mechanics
we have [aa'j = ([o] + [o'])mod2 where a £ A[a], a' € A[aiy It follows that A§ is a subalgebra of A and A\ is an .Ao-module. If A is a graded ring, then [1] = 0. A graded algebra A is said to be graded commutative if aa' = (-i)MM o 'a, where a and a' are arbitrary homogeneous elements of A, i.e., they are either even or odd. Given a graded algebra A, a left graded A-module Q is a left .4-module provided with the grading automorphism 7 such that -y(aq) = 7(0)7(9).
a€A,
q&Q,
i.e., H = ([a] + [<7])mod2. A graded module Q is split into the direct sum Q = Qo ® Qi of two AQmodules Qo and Qi of even and odd elements. Similarly, right graded modules are denned. If /C is a graded commutative ring, a graded /C-module can be provided with a graded K-bimodule structure by letting qa := {-l)^^aq,
a G AC,
q£Q.
A graded /C-module is called free if it has a basis generated by homogeneous elements. This basis is said to be of type (n, m) if it contains n even and m odd elements. In particular, by a (real) graded vector space B — BQ ® B\ is meant a graded K-module. A graded vector space is said to be (n, m)-dimensional if Bo = R" and Bx = Rm. The following are standard constructions of new graded modules from old ones. • The direct sum of graded modules over the same graded commutative ring and a graded factor module are defined just as those of modules over a commutative ring.
349
Chapter 6 Supergeometry
• The tensor product P®Q of graded /C-modules P and Q is an additive group generated by elements p®q, p G P, q G Q, obeying the relations {p + p') ® q = p ® q + p' ® q, p®{q + q') = p®q+p®q', ap®q = (-l)[p][a]pa ®q = (-l) [pI[al p ® aq = aeK.. (-l)(W+[9])Wp®9o, In particular, the tensor algebra
+ {-l)[q][q']q'®q,
9,«'eQ,
is the bigraded exterior algebra of a graded module Q with respect to the graded exterior product qAq'
= -(-1)WI»VA?.
• A morphism $ : P —> Q of graded /C-modules is said to be even (resp. odd) if $ preserves (resp. change) the graded parity of all elements P. It obeys the relations $(ap) = (-l)l*" a l$(p),
p£P,
OG/C.
The set Hom^(P, Q) of graded morphisms of a graded /C-module P to a graded /C-module Q is naturally a graded /C-module. The graded /C-module P* = Horn £(P, /C) is called the duaJ of a graded /C-module P. A graded commutative JC-ring A is a graded commutative ring which is also a graded /C-module. A graded commutative R-ring is said to be of rank N if it is a free algebra generated by the unit 1 and iV odd elements. A graded commutative Banach ring A is a graded commutative R-ring which is a real Banach algebra whose norm obeys the additional condition ||ao + ai|| = ||ao|| + ||ai||,
a0 G Ao,
ai G Ai-
Let V be a real vector space. Let A = AV be its (N-graded) exterior algebra provided with the Z2-graded structure
A = A o eAi,
Ao = R 0 A V , fc=i fc=i
Ai=®2fc/\V.
(6.1.1)
350
Geometric and Algebraic Topological Methods in Quantum Mechanics
It is a graded commutative K-ring, called the Grassmann algebra. A Grassmann algebra, seen as an additive group, admits the decomposition A = R®R = R®R0®R1=R®
(A^2 © Ai,
(6.1.2)
where R is the ideal of nilpotents of A. The corresponding projections a : A —» R and s : A —> R are called the body and soul maps, respectively. Remark 6.1.1. Let us note that there is a different definition of a Grassmann algebra [231] which is equivalent to the above one only in the case of an infinite-dimensional vector space V [102]. Let us mention the Arens-Michael algebras of Grassmann origin [69] which are most general graded commutative algebras, suitable for superanalysis (see Remark 6.9.3 below). • Hereafter, we restrict our consideration to Grassmann algebras of finite rank. Given a basis {c1} for the vector space V, the elements of the Grassmann algebra A (6.1.1) take the form
« = E E
a il ... it c il ---cS
(6.1.3)
fc=o (n-ifc)
where the second sum runs through all the tuples (ii • • • ik) such that no two of them are permutations of each other. The Grassmann algebra A becomes a graded commutative Banach ring if its elements (6.1.3) are endowed with the norm fc=0 (u-i f c )
Let B b e a graded vector space. Given a Grassmann algebra A of rank N, it can be brought into a graded A-module AB = (AB)0 © (AS)! = (Ao ® Bo 8 Ax ® B{) ® {Al ®BQ®AQ®
B{),
called a superspace. The superspace Bn\m =
[ (( | ^
e (™ A i ) ] e
[(£ Ai) © (© A0)]
is said to be (n, m)-dimensional. The graded A0-module 5"-" l = (ffiAo)©(©A1) is called an (n, m)-dimensional supervector space.
(6.1.4)
Chapter 6 Supergeometry
351
Whenever referring to a topology on a supervector space Bn'm, we will mean the Euclidean topology on a 2N~1[n + m]-dimensional real vector space. Given a superspace _B"lm over a Grassmann algebra A, a A-module endomorphism of £?™lm j s represented by an (n + m) x (n + m) matrix (6.1.5) with entries in A. It is called a supermatrix. One says that a supermatrix L is • even if L\ and L4 have even entries, while L 2 and L3 have the odd ones; • odd if L\ and L4 have odd entries, while L2 and L3 have the even ones. Endowed with this gradation, the set of supermatrices (6.1.5) is a graded A-ring. Unless otherwise stated, by supermatrices are meant homogeneous ones. The familiar notion of a trace is extended to supermatrices (6.1.5) as the supertrace StrL = T r L i - ( - l ) [ L l T r £ , 4 . For instance, Str( 1) — n — m. A supertransposition Lst of a supermatrix L is the supermatrix t
_(
L\
(-l)MLl\
where L* denotes the ordinary matrix transposition. There are the relations Str(Ls*) = StrL, (6.1.6)
(LL')st = (-i)MlL']L'stLs\ Str(LL') = (-l) lL1[L/1 Str(L'L)
or Str([L,L']) = 0.
(6.1.7)
In order to extend the notion of a determinant to supermatrices, let us consider invertible supermatrices L (6.1.5). They are never odd. One can show that an even supermatrix L is invertible if and only if either the matrices Li and L4 are invertible or the real matrix a(L) is invertible, where a is the body morphism. Invertible supermatrices constitute a group GL(n\m; A), called the general linear graded group. Then a superdeterminant of L € GL(n\m; A) is
352
Geometric and Algebraic Topological Methods in Quantum Mechanics
defined as SdetL = det(L! -L 2 ^4 1 i3)(detL4 1 ). It satisfies the relations Sdet(LL') = (SdetL)(SdetZ/)> Sdet(L st )=SdetZ,, Sdet(exp(L)) = exp(Sdet (L)). Let K. be a graded commutative ring. A graded commutative (nonassociative) /C-algebra g is called a Lie IC-superalgebra if its product, called the superbracket and denoted by [.,.], obeys the relations ( _l)W[e"] [e7 [ £ V / ] ] + ( _l)[ £ 'l[e] [£ ' ) [e»>e]]
+ ( - 1 ) 1 ^ ' ] [e ", [£)£']] = 0 .
Obviously, the even partfloof a Lie /C-superalgebra g is a Lie /Co-algebra. A graded /C-module P is called a g-module if it is provided with a ^-bilinear map g x P B (e, p) t-> ep e P, [ep] = ([e] + [p])mod2) [e,e']p = ( e o e / - ( - l ) I e " e V o e ) p . 6.2
Graded diflferential calculus and connections
Linear differential operators and connections on graded modules over graded commutative rings are denned similarly to those in commutative geometry. Let K, be a graded commutative ring and A a graded commutative /C-ring. Let P and Q be graded ,4-modules. The graded /C-module Horn A; (P, Q) of graded /C-module homomorphisms $ : P —•» Q can be endowed with the two graded ,4-module structures (a$)(p) := a$(p),
($ • a)(p) := $(ap),
a € A,
p € P,
(6.2.1)
called A- and .4#-module structures, respectively. Let us put 5a$:=a$-(-l)[al[*I$«a,
a & A.
(6.2.2)
353
Chapter 6 Supergeometry
An element A e Hom;c(P, Q) is said to be a Q-valued graded differential operator of order s on P if <5a o °---°^ s A = 0 for any tuple of s +1 elements ao, • • •, as of A- The set Diff S(P, Q) of these operators inherits the graded module structures (6.2.1). In particular, zero order graded differential operators obey the condition SaA(p) = aA(p) - (-l)W( A ]A(ap) = 0,
a £ A,
p € P,
i.e., they coincide with graded ,4-module morphisms P —» Q. A first order graded differential operator A satisfies the condition 5a o Sb A(p) = abA(p) - (-l)(l6WA»Ia)&A(ap) - {-lf^aA(bp) peP. (_l)[b][A]+([A]+[6])H = 0> a,bzA,
+
For instance, let P = A. Any zero order Q-valued graded differential operator A on A is defined by its value A(l). Then there is a graded .4-module isomorphism DiSQ(A,Q) = Q via the association QBq^Aq
eDiff o (.4,Q),
where A, is given by the equality A g (l) = q. A first order Q-valued graded differential operator A on ^4 fulfils the condition A{ab) = A{a)b + (-l)t°][A]aA(6) - (-l)([ b ] + [ Q ]" A la6A(l),
a.fisA
It is called a Q-valued graded derivation of A if A(l) = 0, i.e., the graded Leibniz rule A{ab) = A(a)b + {-l)WWaA{b),
a,b e A,
(6.2.3)
holds (cf. (1.2.10)). One obtains at once that any first order graded differential operator on A falls into the sum A(o) = A(l)a + [A(a)-A(l)o] of a zero order graded differential operator A(l)a and a graded derivation A(a) — A(l)a. If d is a graded derivation of A, then ad is so for any a G A. Hence, graded derivations of A constitute a graded .4-module D(A,Q), called the graded derivation module.
354
Geometric and Algebraic Topological Methods in Quantum Mechanics
If Q = A, the graded derivation module T)A is also a Lie superalgebra over the graded commutative ring /C with respect to the superbracket [uItt/]=uou'-(-i)M[«Vou,
u,u'eA.
(6.2.4)
We have the graded .4-module decomposition (6.2.5)
DiS1(A)=A®DA.
Let us turn now to jets of graded modules. Given a graded ,4-module P, let us consider the tensor product A ®tc P of graded ^-modules A and P. We put 8b(a®p) :=(ba)®p-(-l)[a]lb]a®(bp),
p e P,
a,b e A.
(6.2.6)
The k-order graded jet module Jk{P) of the module P is defined as the quotient of the graded /C-module A ®K P by its submodule generated by elements of type 5bo
o---o5bk{a®p).
In particular, the first order graded jet module Jl{P) consists of elements a®ip modulo the relations ab®lP-
(—I)ta][6l6 ®t (ap) -a®x (bp) + 1
(6.2.7)
For any h G Horn A(A ® P,Q), the equality Sb(h(a®p)) = (-l)WWh(6b(a®p)) holds. By analogous with Theorem 1.2.3, one then can show that any Qvalued graded differential operator A of order A; on a graded ^-module P factorizes uniquely A : P ^Jk(P)
—*Q
through the morphism Jk
:p3p^l®kpeJk(P)
and some homomorphism fA : Jk(P) —> Q. Accordingly, the assignment A w f A defines an isomorphism Diff s (P,Q)=Hom^(J s (P),Q).
(6.2.8)
355
Chapter 6 Supergeometry
Let us focus on the first order graded jet module Jx of A consisting of the elements a ®\ b, a, b G A, subject to the relations ab
(6.2.9)
It is endowed with the A- and .4*-module structures c(a ®i b) := (ca)
c • (a (g>i b) :— a ®i (c6).
There are canonical .4- and ^'-module monomorphisms
ii : A3 a>-* a®il € Jl, :A3a^>l®laeJ1, J1 such that Jl, seen as a graded «4-module, is generated by the elements Jla, a e A. With these monomorphisms, we have the canonical ^.-module splitting J1 = ii(A)®O\ 1
aJ ^)
(6.2.10)
= a 0i b — ab
where the graded ,4-module Ol is generated by the elements I®i6-6®il,
JeA
Let us consider the corresponding ,4-module epimorphism h1 : J1 3 1 <s>i b H^ 1 0 ! 6 - b ®j 1 e O1
(6.2.11)
and the composition d = h1oJ1:A3bt-^l®1b-b®1leO1.
(6.2.12)
The equality d(a6) = a ®i 6 + 6 <&! a - a6
A)
(6.2.13)
356
Geometric and Algebraic Topological Methods in Quantum Mechanics
ofVA to the dual 01* of the graded .4-module O1. It is given by the duality relations X)A3u^(j)u£
Ou,
a € A.
(6.2.14)
Using this fact, let us construct a differential calculus over a graded commutative /C-ring A. Let us consider the bigraded exterior algebra O* of a graded module Ol. It consists of finite linear combinations of monomials of the form cj) = aodai A • • • A dak,
Oj G A,
(6.2.15)
whose product obeys the juxtaposition rule (agdai) A (podbi) = aodiaibo) A db\ — aoa\dbo A db\ and the bigraded commutative relations 4> A <£' = (_l)l*ll*'l+[*H*y A >.
(6.2.16)
In order to make O* to a differential algebra, let us define the coboundary operator d : Ol -> O2 by the rule # ( « , « ' ) = -u'(u(4>)) + (-1)H[«'] U ( U '(0)) + [u',u](
(6.2.17)
It makes O* into a differential bigraded algebra, called a graded differential calculus over a graded commutative /C-ring A. Furthermore, one can extend the duality relation (6.2.14) to the graded interior product of u S VA with any monomial <j> (6.2.15) by the rules u\(bda) =
(-l)M[b]u(a),
u}(') = (u\cj>) Acj>' + (_1)I*I+MM0 A {u\cj>').
(6.2.18)
357
Chapter 6 Supergeometry
As a consequence, any graded derivation u £ DA of A yields a derivation ~Lu4> = u\d<j) + d{u\4)),
(j>£O*,
u€DA,
(6.2.19)
Lu(<j> A 4/) = Lu(
^C^D^A]
^•••Ck[VA;A]
-^-> • • •
(6.2.20)
where C* [DA; A] = Horn
K(Al>A,A)
are J).4.-modules of /C-linear graded morphisms of the graded exterior prodk
ucts A ?)A of the /C-module T)A to A. Let us bring homogeneous elements k DA into the form of A
ei A •••£,. A e r + i A--- Aefc,
e4 e D^o,
£j G ?>A\.
Then the coboundary operators of the complex (6.2.20) are given by the expression 5r+a~1c(ex
A • • • A er A ei A • • • A es) =
(6.2.21)
r
5^(—l) i ~ 1 £»c(ei A • • • £; • • • A e r A ei A • • • e s ) + i=l
s ^ ( - l ) r £ i c ( e i A • • • A er A ei A • • • e,- • • • A es) +
J=I
^2
(-l) i+j c([£i> £j] A £i A • • • £i • • • £, • • • A £ r A ei A • • • A e3) +
l
X]
c([ej, Cj] A £i A • • • A £ r A ei A • • • ei • • • e"j • • • A e s ) +
l
J2
( - l ) i + r + 1 c ( [ e i , €j) A ex A • • • £i • • • A er A ea A • • • e) • • • A es).
l
The subcomplex O*[5^4] of the complex (6.2.20) of ,4-linear morphisms is the graded Chevalley-Eilenberg differential calculus over a graded commutative /C-ring A. Then one can show that the above mentioned graded
358
Geometric and Algebraic Topological Methods in Quantum Mechanics
differential calculus O* is a subcomplex of the Chevalley-Eilenberg one •
O*[X>A}.
Following t h e construction of a connection in commutative geometry in Section 1.3, one comes t o t h e notion of a connection on modules over a graded commutative IR-ring A. T h e following is a straightforward counterp a r t of Definitions 1.3.3 a n d 1.3.4. D E F I N I T I O N 6 . 2 . 1 . A connection morphism
on a graded ,4-module P is an ,4-module
oA B u ^ V u e Diff! (P, P)
(6.2.22)
such that the first order differential operators V u obey the Leibniz rule Vu(ap)=u(a)p+{-l)MMaVu{p),
a e A,
p € P.
(6.2.23)
• 6.2.2. Let P in Definition 6.2.1 be a graded commutative Aring and X)P the derivation module of P as a graded commutative /C-ring. A connection on a graded commutative ,4-ring P is a «4-module morphism DEFINITION
*A3ui->Vue
1>P,
(6.2.24)
which is a connection on P as an .4-module, i.e., obeys the Leibniz rule • (6.2.23).
6.3
Geometry of graded manifolds
By a graded manifold of dimension (n,m) is meant a local-ringed space (Z, 21) where Z is an n-dimensional smooth manifold Z and 21 = 2l0 © 2ti is a sheaf of graded commutative algebras of rank m such that [2l]: (i) there is the exact sequence of sheaves 0-4ft->2l^C|°->0,
U = 2ii + (2ii)2,
(6.3.1)
where CJ? is the sheaf of smooth real functions on Z; (ii) TZ/1Z2 is a locally free sheaf of C|?-modules of finite rank (with respect to pointwise operations), and the sheaf 21 is locally isomorphic to the the exterior product Acg=(7£/7?.2).
Chapter 6 Supergeometry
359
The sheaf 21 is called a structure sheaf of the graded manifold (Z, 21), while the manifold Z is said to be a body of (Z, 21). Sections of the sheaf 21 are called graded functions. They make up a graded commutative C°°(Z)ring 2l(Z). A graded manifold (Z, 21) has the following local structure. Given a point z G Z, there exists its open neighborhood U, called a splitting domain, such that 2l([/) S C°°(£/)
(6.3.2)
It means that the restriction 2l|y of the structure sheaf 21 to [/ is isomorphic to the sheaf Cg?
TT
v
AID 7 7 1
Ail/jy — C X A K
i
TV
—> U.
The well-known Batchelor's theorem [21; 24] states that such a structure of graded manifolds is global. 6.3.1. Let (Z,2l) be a graded manifold. There exists a vector bundle E —> Z with an m-dimensional typical fibre V such that the structure sheaf 21 of (Z, 21) is isomorphic to the structure sheaf 21E of sections of the exterior bundle AE*, whose typical fibre is the Grassmann algebra THEOREM
D
AV*.
Outline of proof. The local sheaves Cg? <8> ARm are glued into the global structure sheaf 21 of the graded manifold (Z, 21) by means of transition functions in Proposition 1.7.1, which are assembled into a cocycle of the sheaf Aut (ARm)°° of smooth mappings from Z to Aut (ARm). The proof is based on the bijection between the cohomology sets if 1 (Z; Aut (ARm)°°) QED &ndH1(Z;GL(m,Rm)o°). It should be emphasized that Batchelor's isomorphism in Theorem 6.3.1 fails to be canonical. At the same time, there are many physical models where a vector bundle E is introduced from the beginning. In this case, it suffices to consider the structure sheaf 2l# of the exterior bundle AE*. We agree to call the pair (Z, 2tg) a simple graded manifold. Its automorphisms are restricted to those induced by automorphisms of the vector bundle E —• Z, called the characteristic vector bundle of the simple graded manifold (Z, 21^). Accordingly, the structure module 21B (Z) = AE*(Z)
360
Geometric and Algebraic Topological Methods in Quantum Mechanics
of the sheaf QiE (and of the exterior bundle AE*) is said to be the structure module of the simple graded manifold (Z,$IE)Given a simple graded manifold ( Z , 2 1 E ) , every trivialization chart (U;zA,ya) of the vector bundle E —> Z is a splitting domain of (Z,21E). Graded functions on such a chart are A-valued functions m
j
fc=0
'
(6.3.3)
where fai--ak(z) are smooth functions on U and {ca} is the fibre basis for E*. In particular, the sheaf epimorphism a in (6.3.1) is induced by the body map of A. We agree to call {zA,ca} the local basis for the graded manifold (Z,$IE). Transition functions y'a =
pab{zA)yb
of bundle coordinates on E —> Z induce the corresponding transformation c'a = pab(zA)cb
(6.3.4)
of the associated local basis for the graded manifold (Z, 21E) and the according coordinate transformation law of graded functions (6.3.3). Remark 6.3.1. Although graded functions are locally represented by A-valued functions (6.3.3), they are not A-valued functions on a manifold Z because of the transformation law (6.3.4). • Remark 6.3.2. Let us note that general automorphisms of a graded manifold take the form (6.3.5)
c>a = pa(zA,c»),
where pa(zA,cb) are local graded functions. Considering simple graded manifolds, we actually restrict the class of graded manifold transformations (6.3.5) to the linear ones (6.3.4), compatible with given Batchelor's isomorphism. • Let E -» Z and E' -> Z be vector bundles and $ : £ - > £ " their bundle morphism over a morphism C, : Z —> Z'. Then every section s* of the dual bundle E'* —> Z' defines the pull-back section $*s* of the dual bundle E* -> Z by the law
^ J $ V ( z ) = S(r>,)K(C(*)),
vzGEz.
Chapter 6 Supergeometry
361
It follows that a linear bundle morphism $ yields a morphism S$:(Z,*E)^(Z',*E.)
(6.3.6)
of simple graded manifolds seen as local-ringed spaces (see Section 1.8). This is the pair ((, £* o $*) of the morphism £ of the body manifolds and the composition of the pull-back %E> 3 / ^ $ 7 G 2tB of graded functions and the direct image £* of the sheaf 2l# onto Z'. Relative to local bases {zA,c°) and (z'A,da) for (Z,%LE) and (Z',VLE>) respectively, the morphism (6.3.6) reads S*(z) = C(z),
S$(c'a) = $ab(z)cb.
Accordingly, the pull-back onto Z of graded exterior forms on Z' is denned. Given a graded manifold (Z,2l), by the sheaf 021 of graded derivations of 21 is meant a subsheaf of endomorphisms of the structure sheaf 21 such that any section u of 1)21 over an open subset U C Z is a graded derivation of the graded algebra 2l(C/). Conversely, one can show that, given open sets U' CU, there is a surjection of the derivation modules 0(2l(£/)) -> 5(2l([/')) [21]. It follows that any graded derivation of the local graded algebra 2l(f7) is also a local section over U of the sheaf 021. Sections of 521 are called graded vector fields on the graded manifold (Z, 2t). They make up the graded derivation module 02l(Z) of the graded commutative R-ring 2l(Z). This module is a Lie superalgebra with respect to the superbracket (6.2.4). In comparison with general theory of graded manifolds, an essential simplification is that graded vector fields on a simple graded manifold (Z, 2lg) can be seen as sections of a vector bundle as follows. Due to the vertical splitting VE^ExE, the vertical tangent bundle VE of E —> Z can be provided with the fibre bases {d/dca}, which are the duals of the bases {ca}. These are the fibre bases for pr2VE £* E.
362
Geometric and Algebraic Topological Methods in Quantum Mechanics
Then graded vector fields on a trivialization chart (U; zA,ya)
of E read
u = uAdA + ua—,
(6.3.7)
where ux, ua are local graded functions on U. In particular,
d
d
d
d
d
d
The derivations (6.3.7) act on graded functions /
•••cb)=
uAdA(fa...b)ca
• • • cb + ukfa...b-^\(c°
• • • cb).
(6.3.8)
This rule implies the corresponding coordinate transformation law
u'A = uA,
u'a =Py + uAdA(p°)c>
of graded vector fields. It follows that graded vector fields (6.3.7) can be represented by sections of the vector bundle VE —> Z which is locally isomorphic to the vector bundle VE\u~AE*®(E®TZ)\u, z z and is characterized by an atlas of bundle coordinates
(6.3.9)
( A *£...«,*. <...** )> fc = 0 , . . . , m , possessing the transition functions 7lA
^ii-..ik
— n~lai
—H
»!
• • • n~lak
V
ik
rA
6ai...a.ki
<~i> = P~Xh • ••P~Xl [p>L.bk + J^y<..M-^AP\k\
,
which fulfil the cocycle condition (10.6.2). Thus, the derivation module 02tjs(Z) is isomorphic to the structure module VE(Z) of global sections of the vector bundle VE —* Z. There is the exact sequence 0 ^ A£*
(6.3.10)
(6-3.11)
363
Chapter 6 Supergeometry
transforms every vector field r on Z into the graded vector field T = TAdA^VT = rA(dA + TA-^),
(6.3.12)
which is a graded derivation of the graded commutative R-ring 2l# (Z) satisfying the Leibniz rule V T (s/) = (7-Jd5)/ + aV T (/),
feZE(Z),
seC°°(Z).
It follows that the splitting (6.3.11) of the exact sequence (6.3.10) yields a connection on the graded commutative C°°(Z)-ring 21B(^) in accordance with Definition 6.2.2. It is called a graded connection on the simple graded manifold (Z,21B). In particular, this connection provides the corresponding horizontal splitting U
= UAdA+U«^=uHdA+lA-^)
+ {ua-UATA)~
of graded vector fields. In accordance with Theorem 10.6.5, a graded connection (6.3.11) always exists. Remark 6.3.3. By virtue of the isomorphism (6.3.2), any connection 7 on a graded manifold (Z, 21), restricted to a splitting domain U, takes the form (6.3.11). Given two splitting domains U and U' of (Z,2l) with the transition functions (6.3.5), the connection components ~jA obey the transformation law l'l = lA^Pa
+ dAPa.
(6.3.13)
If U and U' are the trivialization charts of the same vector bundle E in Theorem 6.3.1 together with the transition functions (6.3.4), the transformation law (6.3.13) takes the form 1 A = Pb{z)lA + dAPab(z)cb.
(6.3.14)
• Remark 6.3.4. It should be emphasized that the above notion of a graded connection is a connection on the graded commutative ring 2ts(Z) seen as a C°°(Z)-module. It differs from that of a connection on a graded fibre bundle {Z, 21) -> (X, B) in [6]. The latter is a connection on a graded B(X)-module represented by a section of the jet graded bundle J1(Z/X) —> (Z,2l) of
364
Geometric and Algebraic Topological Methods in Quantum Mechanics
sections of the graded fibre bundle (Z, 21) —> (X, B) (see [381] for formalism of jets of graded manifolds). • Example 6.3.5. Every linear connection 7 =
dzA®(dA+7Aabybda)
on the vector bundle E —> Z yields the graded connection 75
= dzA ® (dA + 7Aabcb^)
(6.3.15)
on the simple graded manifold (Z, 21B). In view of Remark 6.3.3, 7s is also a graded connection on the graded manifold (Z,2t)S(Z,2l £ ), but its linear form (6.3.15) is not maintained under the transformation law (6.3.13). • The curvature of the graded connection VT (6.3.12) is denned by the expression (1.8.22): #(T,T') = [VT,VT,]-V[T,T,,, R(T,T') RAB
= TAT'BRAB-^
: *E(Z) - 2l £ (Z),
= 9A% - OBTA + 1A-^%
- 1% ^TA-
(6-3.16)
It can also be written in the form (10.6.58): iJ:5a B (Z)->O a (Z)®a B (Z), R=\RaABdzA
AdzB ®-^.
(6.3.17)
Let now Vjg —> Z be a vector bundle which is the pointwise A.E*-dual of the vector bundle VE —> Z. It is locally isomorphic to the vector bundle V%\v ?z AE* ®{E* ®T*Z)\u. z z With respect to the dual bases {dzA} for T*Z and {dcb} for pi2V*E^E*, sections of the vector bundle V% take the coordinate form (j> = 4>AdzA
+ (j>adca,
(6.3.18)
365
Chapter 6 Supergeometry
together with transition functions K = P~lbab,
4>'A = A +
p-lbadA(j>^)
They are regarded as graded exterior one-forms on the graded manifold (Z,2lE), and make up the QlE(Z)-dual
cE = wE(zy of the derivation module
m.B(z) = vE{z). Conversely,
mE{z) = {cEy. The duality morphism is given by the graded interior product u\<j> = uA(j>A + {-l^ua
(6.3.19)
In particular, the dual of the exact sequence (6.3.10) is the exact sequence 0-> A£*®T*Z-> VE-* AE*®E* -> 0. (6.3.20) z z Any graded connection 7 (6.3.11) yields the splitting of the exact sequence (6.3.20), and determines the corresponding decomposition of graded oneforms
lAdzA).
Higher degree graded exterior forms are defined as sections of the extek
rior bundle A V%. They make up a bigraded algebra C*E which is isomorphic z to the bigraded exterior algebra of the graded module CE over CE = 2l(Z). This algebra is locally generated by graded forms dzA, dc* such that dzA Adci = -dci AdzA,
dc{ A dc? = dc? A dc*.
(6.3.21)
The graded exterior differential d of graded functions is introduced by the condition u\df = u(f) for an arbitrary graded vector field u, and is extended uniquely to graded exterior forms by the rule (6.2.17). It is given by the coordinate expression d(f> = dzA A dA4> + dca A T — <j>,
366
Geometric and Algebraic Topological Methods in Quantum Mechanics
where the derivatives d\, d/dca act on coefficients of graded exterior forms by the formula (6.3.8), and they are graded commutative with the graded forms dzA and dca. The formulae (6.2.16) - (6.2.19) hold. The graded exterior differential d makes CE into a bigraded differential algebra whose de Rham complex reads 0 - » R - > a B ( Z ) -^->4 - ^ • • • C | - ^ • • • .
(6.3.22)
Its cohomology HQR{Z) is called the graded de Rham cohomology of the graded manifold (Z, 2l#). One can compute this cohomology with the aid of the abstract de Rham theorem. Let Ofc2l£ denote the sheaf of germs of graded k-forms on (Z, 21E). Its structure module is CE. These sheaves make up the complex O ^ R —-.a B -^D1^
-^•••D f e 2l B - ^ • • - •
(6.3.23)
k
Its members D 2lE are sheaves of C^-modules on Z and, consequently, are fine and acyclic. Furthermore, the Poincare lemma for graded exterior forms holds [2l]. It follows that the complex (6.3.23) is a fine resolution of the constant sheaf R on the manifold Z. Then, by virtue of Theorem 1.7.6, there is an isomorphism H*GR{Z) = H*{Z;R) = H*(Z)
(6.3.24)
of the graded de Rham cohomology HQR(Z) to the de Rham cohomology H*(Z) of the smooth manifold Z. Moreover, the cohomology isomorphism (6.3.24) accompanies the cochain monomorphism O*(Z) —> CE of the de Rham complex 0*{Z) (1.8.9) of smooth exterior forms on Z to the graded de Rham complex (6.3.22). Hence, any closed graded exterior form is split into a sum
Lagrangian formalism on graded manifolds
In order to describe Lagrangian systems of odd variables, one usually calls into play fibre bundles over graded manifolds and supermanifolds [90; 103; 319]. In Section 6.2, the differential calculus, jet modules and connections for graded modules over graded commutative rings have been introduced. At the same time, the Lagrangian BRST theory [20; 61; 60] involves jets of odd variables only with respect to even space-time coordinates. This is also the case of supermechanics where the time coordinate remains even [ill;
367
Chapter 6 Supergeometry
265]. Therefore, in order to develop Lagrangian formalism on a graded manifold (Z, 21B), we follow the differential calculus and geometry introduced on the structure ring 2lg(Z) seen as a C°°(Z)-module in previous Section. As a consequence, we can derive Lagrangian formalism of odd variables from the variational bicomplex over a graded manifold by analogy with Lagrangian formalism on smooth fibre bundles in Section 10.8. Let Q —> X be a vector bundle over an n-dimensional smooth manifold Q equipped with bundle coordinates (xx,qa) together with transition functions q'a = p^q3. Let (X, 21Q) be a simple graded manifold constructed from Q —» X. Its local basis is {xx,ca}. Let 0*21Q be the differential bigraded algebra of graded exterior forms on the graded manifold (X, 21Q). It is locally a free C°°(X)-algebra finitely generated by the elements (l,ca,dxx,dca). Since the jet bundle JrQ —> X of a vector bundle Q —> X is also a vector bundle, let us consider the simple graded manifold (X, 2ljrg) constructed from JrQ —> X. Its local basis is {xx,cA}, 0 < |A| < r, together with the transition functions C'A+A
- dx(pK).
dx
= dx + E
C
|A|
A+A8a ,
(6-4.1)
where d£ are the duals of c\. Let C*JrQ be the differential bigraded algebra of graded exterior forms on the graded manifold (X, 21 jr^). It is locally a free C°°(X)-algebra finitely generated by the elements (l,caA,dx\0aA
= dcak ~ cax+kdxx),
0 < |A| < r.
Since < - i : JrQ -> Jr~lQ is a linear bundle morphism over X, it yields the morphism of graded manifolds
(6.3.6) and the monomorphism of the differential bigraded algebras Cjr-iQ —» C*JTQ [21; 296]. Hence, there is the direct system of differential bigraded algebras CQ ^ C } 1 Q - ^ . . . C } ^ ' ^ . . . ,
(6.4.2)
whose direct limit C^ consists of graded exterior forms on graded manifolds (X, Sljr-g), 0 < r, modulo the pull-back identification. The direct limit C^
368
Geometric and Algebraic Topological Methods in Quantum Mechanics
inherits the operations of a differential bigraded algebra which commute with the monomorphisms 7r£+1*. It is locally a free C°°(X)-algebra countably generated by the elements (l,d}i,dxx,6fi), 0 < |A|. This definition of odd jets differs from that of jets of a graded fibre bundle in [218], but reproduces the heuristic notion of jets of ghosts in the above mentioned BRST theory. Moreover, it enables one to describe odd and even variables on the same footing. Namely, let a smooth fibre bundle Y —» X be affine. Then X is a strong deformation retract of Y and the de Rham cohomology of Y equals that of X. Let V^ be the C°°(X)subalgebra of the differential graded algebra 0*^ which consists of exterior forms on the infinite order jet space J°°Y whose coefficients are polynomial in the fibre coordinates on J°°Y —> X (see Section 10.7). Let us consider the graded product c* _ f>* A -p* °oo ~ '--oo A ' o o
of graded algebras C^ and V^ over their common graded subalgebra O*(X). It consists of the elements i
i
of the tensor products ££,
(6.4.3)
and the multiplication (ip ® 4>) A (ip' ® >') := (-l)l*'H*l(^ A V') ® (^ A 0')-
(6-4.4)
Elements V"8;l+M[v>'y
A
^
^ y,' G S*x,
and makes C^, AP^, into a bigraded C°°(X)-algebra 5^,, where the asterisk means the total form degree. Due to the algebra monomorphisms c;3^^i = i®^s;, ^ 9 0 - » 0 ® l = l ® 0 6 5i ) ,
369
Chapter 6 Supergeometry
one can think of S^ as being an algebra generated by elements of C^ and V^,. For instance, elements of the ring S^ are polynomials of c\ and y\ with coefficients in C°°(X). Let us provide S^ with the exterior differential d(1,<8>),
* e C
0 € 7 £ , , (6.4.5)
where dc and d-p are exterior differentials on the differential algebras C^ and V^, respectively. We obtain at once from the relation (6.4.3) that d{(j>®i)) = {dv
V - e C
4>^V^.
The exterior differential d (6.4.5) is nilpotent. It obeys the equalities d{
if, iff G S£>,
and makes S^ into a differential bigraded algebra, which is locally generated by the elements (1,c%,y\,dx\6 a h = del ~ cl+Adx\6\
= dy\ - y\+Adxx),
|A| > 0.
Hereafter, let the collective symbols s^ and 9^ stand both for even and odd generating elements c^, y\, Q\, 6\ of the C°°(X)-algebra S^ which, thus, is locally generated by (l,s^,dxx,6£), |A| > 0. We agree to call elements of S^ the graded exterior forms on X. Similarly to O^, the differential bigraded algebra S^ is decomposed into S^ -modules S^r of fc-contact and r-horizontal graded forms together with the corresponding projections hk and hr. Accordingly, the exterior differential d (6.4.5) on 5 ^ is split into the sum d = dn + dy of the total and vertical differentials dH{(j>)=dxx
hdx{<j)),
dv(cj)) = e£Ad%ct>,
The projection endomorphism Q of S^ is given by the expression ?o/lfeo/l Q= X^T "' K
fc>o
m
= Y, ( - i ) | A | ^ A [dA(^j,)],
|A|>0
similar to (10.7.18). The graded variational operator 6 = god is introduced. Then the differential bigraded algebra ££, is split into the Z2-graded vari-
370
Geometric and Algebraic Topological Methods in Quantum Mechanics
ational bicomplex (O*(X),S^,Ek
= g(S^);d,dH,dv,Q,6),
(6.4.6)
analogous to the variational bicomplex (10.7.20). It should be emphasized that, in contrast with the Remark 6.4.1. differential graded algebra O*^ of exterior forms on J°°Y, the differential bigraded algebra C^ consists of sections of sheaves on X. In order to regard these algebras on the same footing, let us consider the open surjection 7r°° : j°°y -» X and the direct image T T ^ Q ^ on X of the sheaf Q ^ °f exterior forms on J°°Y (see Theorem 10.7.3). Its stalk at a point x G X consists of the equivalence classes of sections of the sheaf £2^ which coincide on the inverse images {K°°)~1{UX) of open neighborhoods Ux of X. Since (n°a)~1(Ux) is the infinite order jet space of sections of the fibre bundle ir~1(Ux) —> X, every point x G X has a base of open neighborhoods {Ux} such that the sheaves £2^,* of Q^-modules are acyclic on the inverse images (Tr°°)~1(Ux) of these neighborhoods. Then, in accordance with the Leray theorem [184], cohomology of J°°Y with coefficients in the sheaves £2^* is isomorphic to that of X with coefficients in their direct images TT^Q^*, i.e., the sheaves •^oo-doo* on X are acyclic. If Y —> X is an affine bundle, then X is a strong deformation retract of J°°Y and the inverse images (7roo)~1(t/x) of contractible neighborhoods Ux are contractible and 7r£°R. = R. Then, by virtue of the algebraic Poincare lemma, the variational bicomplex H ^ of sheaves on (7roo)~1(C/x), except the terms R, is exact, and the variational bicomplex TT^O^, of sheaves on X is so. There is the R-algebra isomorphism of the differential graded algebra of sections of the sheaf n^Q^ on X to the differential graded algebra Q^ in the proof of Theorem 10.7.3. Thus, Q£o and its subalgebra O^ can be regarded as algebras of sections of sheaves on X, and they keep their d-, dn- and <5-cohomology expressed • into the de Rham cohomology H*{X) = H*(Y) of X [172]. Remark 6.4.2. If Y —> X is a vector bundle, one can obtain the differential bigraded algebra S^ = C^ A P^ in a different way. Let us consider the Whitney sum S = Q © Y of vector bundles Q —> X and Y —> X regarded as a bundle of graded vector spaces Qx © Yx, x € X. Let us define the
371
Chapter 6 Supergeometry k
quotient S of the tensor product sk =
R®S*®®S*®---®®S*
xx by the elements q*®q'*+q'*®q*,
xx
y* ®y'* - y" ® y*',
q* ®y* -y* ®q*
for all q*,q'* G Q*, jy*,y'* £ y^, and x E X. The C°° (X)-modules ^ | of sections of the vector bundles S —> X make up a direct system with respect to the natural monomorphisms As —> Ag1. Its direct limit A'g' is endowed with a structure of a graded commutative C°° (X)-ring generated by odd and even elements. Generalizing the above technique for a graded manifold (X, AQ) to (X, Ag3), one obtains the differential bigraded algebra • isomorphic to S ^ [296; 387]. We aim to study the cohomology of the short variational complex 0—•R_+S£ ^ S ^ 1 - - - ^ S £ , n -*-£!
(6.4.7)
and the complex of one-contact graded forms U —> 6 ^ — • " o o " "
— >< - ) oo
—
>
£'i^U
(b.4.8)
of the differential bigraded algebra S^ • For this purpose, one however must: (i) enlarge the differential bigraded algebra 5^, to the differential bigraded algebra r(6J o ) of graded exterior forms of locally finite jet order, (ii) compute the cohomology of the corresponding complexes of r(SJ 0 ), (iii) prove that this cohomology of r(6J o ) coincides with that of S^. Following this procedure, we show that cohomology of the complex (6.4.7) equals the de Rham cohomology of X, while the complex (6.4.8) is globally exact. Note that the exactness of the short variational complex (6.4.7) on X = Rn has been repeatedly proved [20; 60; 134]. One has also considered its subcomplex of graded exterior forms whose coefficients are constant on R n . Its dn-cohomology is not trivial [20]. One can think of the elements L = Cu e S^n,
5(L)=
^2(-l)^eAAdA(d^L)eE1
|A|>0
372
Geometric and Algebraic Topological Methods in Quantum Mechanics
of the complexes (6.4.7) - (6.4.8) as being a graded Lagrangian and its Euler-Lagrange operator, respectively. Therefore, the exactness of the complex (6.4.8) enables us to generalize the first variational formula and Lagrangian conservation laws in the calculus of variations on fibre bundles to graded Lagrangians and contact supersymmetries. THEOREM 6.4.1. The cohomology of the complex (6.4.7) equals the de Rham cohomology H*(X) of X. The complex (6.4.8) is exact. • The proof of Theorem 6.4.1 falls into the three steps, (i) We start by showing that the complexes (6.4.7) - (6.4.8) are locally exact. LEMMA
6.4.2. The complex (6.4.7) on X = Rn is exact.
•
Referring to [20], Theorems 4.1 - 4.2, for the proof, we summarize a few / S S^* admits formulae quoted in the sequel. Any horizontal graded form <> the decomposition l
4> = d>0 + l
^ = / y E s A a ^> 0
(6.4.9)
|A|>0
where cf>o is an exterior form on R™. Let 4> £ <S^ m < n be dn-closed. Then its component
Jo
fc>°
(6.4.10)
The relation
holds, and leads to the desired expression (6.4.11)
P 0 = l,
Pk=dVl.--dVkD+^---D+v*.
Now let <j> 6 S^m
Chapter 6 Supergeometry
373
by the expression
£= E
E
|A|>O£+E=A
(-l) |E| *2d E 0£ +A &V
(6-4-12)
Remark 6.4.3. Since elements of S^ are polynomials in s^, the sum in the expression (6.4.11) is finite. However, the expression (6.4.11) contains a d/f-exact summand which prevents its extension to O^. In this respect, we also quote the homotopy operator (5.107) in [346] which leads to the expression l
£= Jl(
(6.4.13)
o
i(4,) = y y — ^ t ± l —
(6.4.14)
d
At E (- 1 ) 5 ( r + A A vS ) ! s A ^^ + A + E ^^^i' |H|>0
VM
;
'~"
where A! = AMl! • • • A/in! and AM denotes the number of occurrences of the index /i in A. The graded forms (6.4.12) and (6.4.13) differ in a d^-exact graded form. D LEMMA
6.4.3. The complex (6.4.8) on X = W1 is exact.
D
Outline of proof. The fact that a dn-closed graded (1, m)-form (f> £ S^7l
(6.4.15)
where <$>\ 6 S^m are horizontal graded m-forms. Let us introduce additional variables s^ of the same Grassmann parity as s^. Then one can associate to each graded (l,m)-form <j> (6.4.15) a unique horizontal graded m-form
^=E^5A>
(6-4-16)
whose coefficients are linear in the variables s^, and vice versa. Let us consider the modified total differential
dH = dH+dxXA
E^A+A^l |A|>0
374
Geometric and Algebraic Topological Methods in Quantum Mechanics
acting on graded forms (6.4.16), where dA is the dual of ds^. Comparing the equality dHs£ = dxxs*+A and the last equality (10.8.30), one can easily justify that dH
^H4>-
Let a graded (l,m)-form
It follows that <j> = dif£ where
It remains to prove the exactness of the complex (6.4.8) at the last term Ex. If
Q(a)= ^2(-l)^9AA[dA(di\a)}
=
|A|>0
53 (-l)lAl0A A [d^]u = 0,
ae S^n,
|A|>0
a direct computation gives
c =- E
E (-i) |E| ^Ad E ^ + V-
|A|>0S+E=A
(6-4-17) QED
(ii) Let us associate to each open subset U C X the differential bigraded algebra Sjj of elements of the C°° (X)-algebra S^ whose coefficients are restricted to U. These algebras make up a presheaf over X. Let & ^ be the sheaf of germs of this presheaf and F(©^o) its structure module of sections. One can show that 6 ^ inherits the variational bicomplex operations, and r(S^ o ) does so. For short, we say that F(SJO) consists of polynomials
375
Chapter 6 Supergeometry
in s\, ds\ of locally bounded jet order |A|. There is the monomorphism S^ —» r(6^ o ). Let us consider the complexes of sheaves
o — R _ ^ 6 ^ ^ e ^ 1 - - - *£+&£ - ^ d ,
(6.4.18)
0 -> ©M> ^
(6 4 1 9 )
g u . .. ^
gi,n ^
£i
_
0;
over X and the complexes of their structure modules
O ^ R —>r(62j ^r(e°J)--- ^ r ( s ^ " ) -^r(€!),(6.4.20) o-+r(e^°) ^ r ( s ^ ) . . . ^ r ( s ^ " ) - ^ ( d j - o . (6.4.21) By virtue of Lemmas 6.4.2 - 6.4.3 and Theorem 1.4.2, the complexes of sheaves (6.4.18) - (6.4.19) are exact. The terms 6^,* of the complexes (6.4.18) - (6.4.19) are sheaves of C°°(X)-modules. Therefore, they are fine and, consequently, acyclic. By virtue of Theorem 1.7.7, the cohomology of the complex (6.4.20) equals the cohomology of X with coefficients in the constant sheaf M, i.e., the de Rham cohomology H*(X) of X, whereas the complex (6.4.21) is globally exact. (iii) It remains to prove the following. 6.4.4. Cohomology of the complexes (6.4.7) - (6.4.8) equals that of the complexes (6.4.20) - (6.4.21). • PROPOSITION
Let the common symbol D stand for the operators dn, 5 and g in the complexes (6.4.20) - (6.4.21), and let F ^ denote the terms of these complexes. Since cohomology groups of these complexes are either trivial or equal to the de Rham cohomology of X, one can say that any D-closed element cf> G F ^ takes the form 4> = i, + D^
CeC
(6.4.22)
where ip is a closed exterior form on X which is not necessarily exact. Since all D-closed elements of F ^ of finite jet order are also of form (6.4.22), it suffices to show that, if an element (p £ «S£> ^s -f-exact in the module F ^ (i.e.,
376
Geometric and Algebraic Topological Methods in Quantum Mechanics
It follows that, if
property. • Outline of proof. Let <j> e S^ be a D-exact graded form on X. The finite exactness on UUa holds since > = D<pa on every Ua and all [ipa] < JV( [<£]). QED
6.4.6. Suppose that the finite exactness of an operator D takes place on open subsets U, V of X and their non-empty overlap U(~\V. Then it is also true on U U V. • LEMMA
Outline of proof. Let
where a is also of bounded jet order. Lemma 6.4.7 below shows that a = °~u + °~v where o\j and av are graded forms of bounded jet order on U and V, respectively. Then, putting y'u =
Dav,
we have the graded form <j>, equal to Dip'y on U and D
I
377
Chapter 6 Supergeometry
is of bounded jet order.
QED
6.4.7. Let U and V be open subsets of X and a a graded form of bounded jet order on U D V. Then a splits into the sum o\j + oy of graded exterior forms ou on U and ay on V of bounded jet order. • LEMMA
Outline of proof. By taking a smooth partition of unity on U U V subordinate to its cover {U, V} and passing to the function with support in V, we get a smooth real function / on U U V which is 0 on a neighborhood Uu-v of U — V and 1 on a neighborhood Uv-u of V — U in U U V. The graded form fa vanishes on Uu-v n{UDV) and, therefore, can be extended by 0 to U. Let us denote it a\j. Accordingly, the graded form (1 — f)a has an extension ay by 0 to V. Then a = au +ay is a desired decomposition because au and ay are of finite jet order which does not exceed that of a. QED Lemma 9.5 in [65], Chapter V, states that, if some property holds on a domain and obeys the conditions of Lemmas 6.4.5 and 6.4.6, it holds on any open subset of R n . Hence, the operator D has the jet exactness property on any open subset of R™ and, consequently, on any chart of the fibre bundle QxY-^X. x Since the latter admits a finite bundle atlas with the transition functions (10.7.11) and (6.4.1) preserving the jet order, the finite exactness of D takes place on the whole manifold X in accordance with Lemma 6.4.6. This proves Proposition 6.4.4 and, consequently, Theorem 6.4.1. Remark 6.4.4. Let us consider the complex O - R ^ S ^ J L ^ - . - J U ^
—>•••,
(6.4.23)
which we agree to call the de Rham complex because («!>£,, d) is the differential calculus over the R-ring ££,. If X = W1, it is exact [134]. Similarly to the proof of Theorem 6.4.1, one can show that the cohomology of the de Rham complex (6.4.23) equals the de Rham cohomology of X. O
378
Geometric and Algebraic Topological Methods in Quantum Mechanics
COROLLARY 6.4.8. Every ^ - c l o s e d graded form <$> G <S^m<™ falls into the sum 4> = il> + dHK,
(6.4.24)
^€S^m-\
where tp is a closed m-form on X. Every 5-closed graded Lagrangian L £ S^n is the sum £e<S£"-\
cf> = iP + dHt,,
(6.4.25)
where ip is a non-exact n-form on X.
D
The global exactness of the complex (6.4.8) at the term S^n results in the following. PROPOSITION
6.4.9. Given a graded Lagrangian L =
CUJ,
there is the
decomposition dL = 5L-dHZ,
(6.4.26)
EGS^1'1,
3= EC.^^ 8 "" 1 ^,
(6A27)
s=0
where local graded functions h obey the relations h"a = 0,
/ i l " ' " ' - 1 ' - " 1 = 0.
• Outline of proof. The decomposition (6.4.26) is a straightforward consequence of the exactness of the complex (6.4.8) at the term 5^," and the fact that Q is a projector. The coordinate expression (6.4.27) results from a direct computation
- d H E = - d H \ 6 A F \ + OfF?
+ ••• + tfM...UlF%'--v* +
[9AdxF>l + Q${F\ + dxF]?) + ••• + 0 A t + l V . . . . V l ( K ' + 1 ' " " " 1 + d x F ? - + 1 ' " - V l ) + ---]Aw = [U d\tA +tiv{oAL) -\
h oVa+il/a
Ul
[OA
L)-\ ] A w -
6A(dxF% - dAC) Au> + dL = -6L + dL. QED
379
Chapter 6 Supergeometry
Proposition 6.4.9 states the existence of a global finite order Lepagean equivalent Hi, = H + L of any graded Lagrangian L. Locally, one can always choose H (6.4.27) where all functions h vanish. By analogy with a contact symmetry (see Proposition 10.8.2), an infinitesimal contact supersymmetry or, simply, a contact supersymmetry is defines as a graded derivation v € 05^, of the K-ring 5^, such that the Lie derivative L v preserves the ideal of contact graded forms of the differential bigraded algebra S^ (i.e., the Lie derivative Lv of a graded contact form is a graded contact form). 6.4.10. With respect to the local basis (xx,sA,dxx,6A) for the differential bigraded algebra S^, any contact supersymmetry takes the form PROPOSITION
v = vH + vv = vxdx + (vAdA + J2 d^Ad^),
(6.4.28)
|A|>0
where vx, vA are local graded functions.
•
Outline of proof. The key point is that, since elements of C^, can be identified as sections of a finite-dimensional vector bundle over X, so can elements of the C°°(X)-algebra 5£,. Moreover, any graded form is a finite composition of df, f e S^. Therefore, the proof follows that of Proposition 10.8.2. QED The interior product v\
teSl,, +
v\(
Lvcf> = v\d(f) + d(v}(/>), Lv{
A
Lv{a),
as those on a graded manifold. Following the proof of Lemma 10.8.3, one can justify that any vertical contact supersymmetry v (6.4.28) satisfies the relations (6.4.29)
v\dH(f> = -dH(v\(f>), Lv(dH
4>eSZ,.
(6.4.30)
380
Geometric and Algebraic Topological Methods in Quantum Mechanics
6.4.11. Given a graded Lagrangian L 6 S^1, its Lie derivative LVL along a contact supersymmetry v (6.4.28) fulfills the first variational formula PROPOSITION
LVL = w\SL + dH{h0{v]EL)) + dv{vH\u)C,
(6.4.31)
where S L = 2 + L is a Lepagean equivalent of L given by the coordinate expression (6.4.27). D Outline of proof. The proof follows that of Proposition 10.8.4 and results QED from the decomposition (6.4.26) and the relation (6.4.29). In particular, let v be a variational symmetry of a graded Lagrangian L, i.e., LvL = dHa,
aeS0^-1.
Then the first variational formula (6.4.31) restricted to Ker<5L leads to the weak conservation law 0*dH(h0(v\ZL)-a).
(6.4.32)
Remark 6.4.5. Let us consider the gauge theory of principal connections on a principal bundle P —> X with a structure Lie group G. These connections are represented by sections of the quotient C = JlP/G —> X (see Section 10.10). This is an afflne bundle coordinated by (xx,arx) such that, given a section A of C —» X, its components A\ — arx o A are coefficients of the familiar local connection form (i.e., gauge potentials). Let J°°C be the infinite order jet manifold of C —> X coordinated by {xx,ar A ) , 0 < |A|, and let P^,(C) be the polynomial subalgebra of the differential graded algebra 0^(0). Infinitesimal generators of one-parameter groups of vertical automorphisms (gauge transformations) of a principal bundle P are G-invariant vertical vector fields on P —» X. They are associated to sections of the vector bundle VGP = VP/G -* X of right Lie algebras of the group G. Let us consider the simple graded manifold (X, 2lyGy) constructed from this vector bundle. Its local basis is (xx, Cr). Let CJTVGY be the differential bigraded algebra of graded exterior
Chapter 6 Supergeometry
381
forms on the graded manifold (X,2ljryoP), and C^VQP) the direct limit of the direct system (6.4.2) of these algebras. Then the graded product S^VcC)
= C(VCP)AP;(C)
(6.4.33)
describes gauge potentials, odd ghosts and their jets in the BRST theory. With respect to a local basis (ar\a^,C r ) for the differential bigraded algebra S^VcC) (6.4.33), the BRST symmetry is given by the contact supersymmetry
v = v\$ + vrdr + £ (dAVrxOy + dAvrd£),
(6.4.34)
|A|>0
where c£g are structure constants of the Lie algebra of G and d£, dr, <9^'A d£ are the duals of darx, dCr, darA x and dCrK, respectively. A remarkable peculiarity of this contact supersymmetry is that the Lie derivative Lv along v (6.4.34) is nilpotent on the module S^* of horizontal graded forms.
•
In a general setting, a vertical contact supersymmetry v (6.4.28) is said to be nilpotent if
ML^) =
£
(vid%(v£)d% + (-l)[sB][vA]viv£d%db
|E|>0,|A|>0
(6.4.35) for any horizontal graded form (f> £ S^*. 6.4.12. A contact supersymmetry v is nilpotent if and only if it is odd and the equality LEMMA
U{vA) - Y, vld%(vA) - 0 |S|>0
holds for all vA. Outline of proof. There is the relation dxoV\d$=v\d$odx,
D (6.4.36)
similar to (10.8.31). Then the lemma follows from the equality (6.4.35) where one puts <j> — sA and 0 = sAs^. QED
382
6.5
Geometric and Algebraic Topological Methods in Quantum Mechanics
Lagrangian supermechanics
Lagrangian supermechanics exemplifies Lagrangian systems on graded manifolds in previous Section. Moreover, following the BRS mechanics of [191; 192], one can construct the canonical SUSY extension of Lagrangian and Hamiltonian time-dependent mechanics [294; 386]. The first step toward this extension is the vertical extension of Lagrangian and Hamiltonian mechanics on a configuration space Q —> R onto the vertical tangent bundle VQ - • R in Section 2.4G. The SUSY-extended mechanics is formulated in terms of simple graded manifolds whose characteristic vector bundles are
VJXVQ -* JlVQ in Lagrangian formalism (see the isomorphism (2.4.75)) and
VV*VQ -> VVQ in Hamiltonian formalism (see the isomorphism (2.4.73)). The SUSY extension adds to (t,gl,ql) the odd variables (c\cl). The corresponding SUSY extensions of a Lagrangian and a Hamiltonian, by construction, are invariant under the one-parameter supergroup with the infinitesimal generator UQ=c%+iqi-^.
(6.5.1)
Let us consider the vertical tangent bundle VVQ —> VQ of VQ —> R and the simple graded manifold (VQ,QIVVQ) whose body manifold is VQ and the characteristic vector bundle is VVQ —> VQ. Its local basis is (t, g 4 ,^ 1 ,^,?) where {cl,c1} is the fibre basis for V*VQ, dual of the holonomic fibre basis {di,di} for VVQ —> VQ. Graded vector fields and graded exterior one-forms are introduced on (VQ,$IVVQ) as sections of the vector bundles VVVQ and VyVQ, respectively. We complexify these bundles as C
VVJXQ -> VJlQ of VJlQ -> R. Its local basis is
383
Chapter 6 Supergeometry
where {c1 ,c*, c^cl} is the fibre basis for V*VJ1Q dual of the holonomic fibre basis {di,di,dj,dl} for
VVJlQ -> VJlQ. The affine fibration TT* : VJXQ - • 1/g and the corresponding vertical tangent morphism V-KI : V V ^ Q -» VVQ
yields the associated morphism (6.3.6) of graded manifolds
(VJlQ,KVVJiQ)
-»
(VQ,*VVQ).
In a similar way, the simple graded manifold with the characteristic vector bundle
VVJkQ -» VJkQ is defined. Its local basis is the collection {t,q\qiq\qlc\T?,^,c\),
A = (t...t),
0 < |A| < fc.
Let us introduce the operator of the total derivative dt = dt + qidi+ci—.+7?t-^r
+...,
With this operator, the coordinate transformation laws of c\ and cj read d? = dtc'\
c'j = dtcH.
(6.5.2)
Then one can treat c\ and c[ as the jets of the odd variables c% and c\ The transformation law (6.5.2) shows that the graded vector field UQ (6.5.1) on (VQ,QIVVQ) gives rise to the complex graded vector field
JluQ = uQ + cidt + iyi-jjg
(6.5.3)
on (VJ^Q, ^•VVJ1Q)- The Lie derivative along this graded vector field plays a role of the BRS operator in SUSY mechanics. In accordance with Lemma 6.4.12, it is nilpotent.
384
Geometric and Algebraic Topological Methods in Quantum Mechanics
As in the BRS mechanics [191; 192], the main criterion of the SUSY extension of Lagrangian mechanics is its invariance under the BRS transformation (6.5.1). Let us introduce the operators
dc = cidi + etd\ + ettdf + ---, dc = cidi +
(6.5.4)
dtdc — dcdt.
Then the desired BRS-invariant extension of the vertical Lagrangian Ly (2.4.77) is the graded horizontal density (6.5.5)
Ls = Csdt = LV + idcdcCdt such that IJJUQLS
=
0.
The corresponding Euler-Lagrange equations are defined as the kernel of the Euler-Lagrange operator
They read
SLs = (dq% + dq% + dc'^7 + dc'^Cs ocl 6c
A dt.
5iCs = 5iC = 0, 5iCs = SiCv + idcdAC = 0,
(6.5.6a) (6.5.6b)
-^-Cs = -idcSiC = 0, oc1 ^-Cs = idcSiC = 0,
(6.5.6c) (6.5.6d)
where the relations (6.5.4) are used. The equations (6.5.6a) are the EulerLagrange equations for the original Lagrangian L, while (6.5.6b) - (6.5.6d) can be seen as the equations for a Jacobi field 5yl = ecl + c*£ + ieey% modulo terms of order > 2 in the odd parameters e and e.
385
Chapter 6 Supergeometry
6.6
Graded Poisson manifolds
Let us start with the notion of a graded symplectic manifold. Let (Z,%IE) be a simple graded manifold, E its characteristic vector bundle and C*E the differential bigraded algebra of graded exterior forms on (Z,21JS). A graded two-form
(6.6.1)
has the zero kernel Ker<^>. In this case, the map (6.6.1) is an isomorphism of graded modules. Let <j>0 and 4>i be even and odd parts of (p. A graded two form cfi is said to be homogeneously non-degenerate if Kerfo n K e r ^ i = 0. Of course, any non-degenerate graded two-form is also homogeneously nondegenerate, while a homogeneous (even or odd) graded two-form is nondegenerate if and only if it is homogeneously non-degenerate. In a general setting, a closed graded two-form is called the graded symplectic form if it is homogeneously non-degenerate. However, by graded symplectic form is often meant a non-degenerate closed graded two-form. For the sake of simplicity, we here restrict our consideration to even symplectic forms (see, e.g., [420; 243] for the non-homogeneous and odd ones). Let Q, be an even symplectic form on a simple graded manifold (Z, 21E). Similarly to the case of symplectic manifolds in Section 2.1 A, one can associate to any graded function / GCE the graded vector field tf/jn = -df,
-Of = « » ( # ) ,
(6.6.2)
where fi" is the inverse map to the even map fib (6.6.1). It is called the Hamiltonian graded vector field of a graded function / . Given another graded function g and its Hamiltonian graded vector field dg, one can define the (even) graded Poisson bracket {f,g} = $f\dg =
tig\$f\n.
(6.6.3)
386
Geometric and Algebraic Topological Methods in Quantum Mechanics
It obeys the relations {/,} = -(-l) [/1[9l {,/}, (-1)[/H*J{/, {g, h}} + (-l)W[/]{fl, {h, /}} + (-l)[hM{h,{f,9}}
(6.6.4a) (6.6.4b)
= 0,
{/, 9h) = {/, g}h + (-1)1/1 W p {/, h}, [^/A] =#{f,g}-
(6.6.4c) (6-6.4d)
One can prove that a graded manifold (Z, 21E) can be endowed with a graded symplectic structure if its body manifold Z admits an exact symplectic form [182; 318]. The graded version of the Darboux theorem assures that one can find local canonical basis for a graded symplectic structure. One says that a graded symplectic form on a simple graded manifold (Z, 21JE) is projectable onto Z if the restriction of the graded Poisson bracket (6.6.3) to the R-subring C°°(Z) C C% of functions on Z is a Poisson bracket associated to some symplectic structure on Z. Theorem 6.6.1 below shows that a graded symplectic structure cannot be projectable, unless the characteristic bundle E —» Z admits a flat connection. Turn now to the notion of a graded Poisson manifold. As before, we restrict our consideration to the case of even Poisson brackets (see, e.g., [241; 242] for the odd ones). A glance at the relation (6.6.4c) shows that a graded Poisson bracket on a simple graded manifold (Z, 2lg) can be defined by associating to each graded function / £ CE a graded derivation
#f.CEBf~{f,f}eC°E
(6.6.5)
of the M-ring C\ which satisfies the graded skew-symmetry condition (6.6.4a) and the graded Jacobi identity (6.6.4b). Herewith, "df (6.6.5) is the Hamiltonian graded vector field of / . A graded Poisson bracket is said to be projectable if its restriction to the M-subring C°°(Z) C C°E of functions on Z is a Poisson bracket on Z. We quote the following theorem [225]. 6.6.1. Let (Z,QlE) be a simple graded manifold and {-,-}z a (regular) Poisson bracket on Z. Then a graded Poisson bracket on (Z, 21^) projected to {.,.}% on Z exists if and only if the characteristic vector bundle E —> Z of (Z, QIE) admits a connection whose curvature vanishes along the symplectic leaves of {., .}zd THEOREM
Chapter 6 Supergeometry
387
The proof of this theorem is based on the following facts. Bearing in mind that the ring C% is the ring /\E*(Z) of sections of the exterior bundle AE*, let us consider the following two types of derivations of the ring AE*{Z). (i) Let 7 (6.3.5) be a linear connection on E and 75 (6.3.15) the corresponding graded connection on (Z, Slg). Given a section 5 = g
€ AE*(Z),
(6.6.6)
where V r is given by the expression (6.3.12). Clearly, V s (6.6.6) is a graded derivation of the E-ring AE*(Z). (ii) Let us consider the tensor product AE*®E. Given a section r = g®e of AE* ®E -> Z, one can associate to r the endomorphism Tr : AE*(Z) 9 / ' -» ge\
(6.6.7)
This is a graded derivation of the C°°(Z)-ring AE*(Z). 6.6.2. Given a connection 7 on E —> Z, each graded derivation A of the K-ring AE*(Z) is uniquely split into the sum A = V s + Tr of the • graded derivations (6.6.6) and (6.6.7) [317].
LEMMA
In particular, for each graded function / G C%, there exist the above mentioned sections Sf and rf such that the graded Hamiltonian vector field •df (6.6.5) of / is brought into the form tif = VSf+Trj. Then the graded skew-symmetry condition (6.6.4a) and the graded Jacobi identity (6.6.4b) lead to the statement of Theorem 6.6.1.
388
6.7
Geometric and Algebraic Topological Methods in Quantum Mechanics
Hamiltonian supermechanics
This Section addresses the SUSY extension of Hamiltonian time-dependent mechanics on a configuration space Q —> R. The momentum phase space of the SUSY-extended mechanics is the complexified simple graded manifold (V*VQ,QIVVVQ) whose characteristic vector bundle is VV*VQ^ V*VQ. Its local basis is (t,ql,Pi,ql,Pi,cl,^,Ci,Ci),
where Cj and Cj have the same transformation laws as Pi and pi, respectively. The corresponding graded vector fields and graded one-forms are introduced as sections of the vector bundles on the graded manifold {V*VQ,%IVVVQ) C®VVVVQ and C<8)VVV.VQ, respectively. X
X
In accordance with the above mentioned transformation laws of c^ and c*, the graded vector field UQ (6.5.1) on (VQ,%LVVQ) can give rise to the complex graded vector field (6.7.1)
uQ=dc + iq'-^r + ipi—
on {V*VQ,%VV*VQ)The BRS-invariant extension of the canonical threereads form fly (2.4.80) onto the graded manifold (V*VQ,QIVVVQ) Qs = [dpi A dqi + dpi A dql + i(ddi A dc* - d ? A dci)} A dt,
(6.7.2)
where (cl, —ici) and (c*,icj) are the conjugate pairs. Let 7 H be the Hamilton vector field (2.4.34) for a Hamiltonian form H (2.4.32) on the original momentum phase space V*Q. Its double vertical prolongation VV-y on VV*VQ -> R (see the formula (2.4.76)) defines the graded connection 7S
= v,H + 8&H^ - a ^ A + de#H£ _
dAn^
on (V*V<2,2lyv*VQ)) where V^H is given by the expression (2.4.82) and dc = 0% + Cid\
dc = tfdi + Cid\
This graded connection satisfies the relation 75 Jfi s
= -dHs,
389
Chapter 6 Supergeometry
and can be regarded as the Hamilton graded vector field for the Hamiltonian graded form Hs = \pidqi + pidc? + iicidd + d?a)] - Hsdt, Hs = {dv + idcdc)H,
(6.7.3)
on (V*VQ,%LVV*VQ)- It is readily observed that this graded form is BRSinvariant, i.e., LizQHs = 0. Thus, it is the desired SUSY extension of the original Hamiltonian form H (2.4.32). The Hamiltonian graded form Hs (6.7.3) defines the corresponding SUSY extension of the Lagrangian L# (2.4.42) as follows. The fibration JlV*VQ->V*VQ yields the pull-back of the Hamiltonian graded form Hs (6.7.3) onto J1V*VQ. Let us consider the graded generalization of the operator h0 (10.7.15) such that ho '• dcl H+ c\dt,
dci >—> cudt.
Then the graded horizontal density LHS = ho(Hs) = LHV + i[{ciCi + cla) - dedjQdt = LHV +
(6.7.4)
i[ci(4 ~ aca'w) + (cj - dc&tyci + 2ic^aiajw - cicjdidjn}dt on JXV*VQ -» X is the SUSY extension (6.5.5) of the Lagrangian LH (2.4.42). The Euler-Lagrange equations for Lus coincide with the Hamilton equations for Hs, and read q\ = &HS = dirH, Pti = -diHs = -diH, Pu = -diHs = q\ = &Hs = (dv + idcdJ&H, ~{dv + idcdc)diH,
(6.7.5a) (6.7.5b)
4 = i ^ - = -dc&H,
(6.7.5c)
3 = - t ^ =-^W, dci
cu = i ^
= -dcdiH,
cti = -i^f = -ckdiH. (6.7.5d) d?
The equations (6.7.5a) are the Hamilton equations for the original Hamiltonian form H, while (6.7.5b) - (6.7.5d) describe the Jacobi fields 5ql = ecl + cle + ieeq1,
dpi = ecj + Cj£ + ieepi.
390
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let now Q —> M. be an affine bundle. In this case, transition functions of the holonomic coordinates q1 on VQ are independent of q%, the transformation laws of the frames {di} and {di} coincide, and so do the transformations laws of the coframes {c1} and {}. Then the graded vector fields (6.7.6) UK==CiW
U«=~Cid*
are globally denned. The graded vector fields (6.5.1) and (6.7.6) constitute the Lie superalgebra of the supergroup ISp(2): [UQ,UQ] = [UQ,UQ] = [UQ,UQ] = {UK,UQ} = [UW,UQ] [UK,UQ]
= uQ,
[uc, uK) = 2uK,
[u-j?, uQ] = U-Q,
[UK, ujr] = uc,
= 0, (6.7.7)
[uc, U-K\ = -2uj(.
Similarly to (6.5.3), let us consider the jet prolongation of the graded vector fields (6.7.6) onto VJlY. Using the compact notation u = uada, we have the formula
Jlu = u + dtuadl and, as a consequence, obtain
(6.7.8) Jiuc=uc
= cl—-c\
—.
It is readily observed that the SUSY-extended Lagrangian Ls (6.5.5) is invariant under the transformations (6.7.8). The graded vector fields (6.5.3) and (6.7.8) make up the Lie superalgebra (6.7.7). The graded vector fields (6.7.6) can give rise onto ( V * V Q , 2 1 V V V Q ) by the formula 5 = u _(_i)[«-](w+[« 6 ])a o U «'p 6 ^_.
OPa
391
Chapter 6 Supergeometry
We have
(6.7.9)
id ac*
3 9Q
,
8 . 3 9c* oci
A direct computation shows that the BRS-extended Hamiltonian form Hs (6.7.3) is invariant under the transformations (6.7.9). Accordingly, the Lagrangian LHS (6.7.4) is invariant under the jet prolongation Jlu of the graded vector fields (6.7.9). The graded vector fields (6.7.1) and (6.7.9) make up the Lie superalgebra (6.7.7). With the graded vector fields (6.7.1) and (6.7.9), one can construct the corresponding graded currents I{u) = u\Hs = u\Hs, which read I{uQ) = dpi - fa, I(UK)
= -icjC*,
I(UQ) = tin /(JJJ)
- q%,
= iCiZ\
I{uc) = ifac1 - clCi).
These currents form the Lie superalgebra (6.7.7) with respect to the product [/(«),/(«')]=/([«,«'])•
The following construction is similar to that in the SUSY and BRS mechanics. Given a function F on the Legendre bundle V*Q, let us consider the operators Fp = e0F OUQO e~0F = uQ- (3dcF, Fp = e-0FouQoe0F =v,Q+pdcF, called the SUSY (V*VQ,$IVV*VQ)-
/3 > 0,
charges, which act on graded These operators are nilpotent, i.e., Ff}oFi3 = 0,
functions on
TpoFp = 0.
(6.7.10)
By the BRS-invariant extension of a function F is meant the graded function fs = ~(f/joF/J+F/,oF/1). We have the relations F0oFs-FsoF(s
= 0,
T(3 o Fs - Fs o ~F0 = 0.
392
Geometric and Algebraic Topological Methods in Quantum Mechanics
These relations together with the relations (6.7.10) provide the operators Fp, Ffj, and Fs with the structure of the Lie superalgebra sl(l/l) [94]. In particular, let F be a local function % in the expression (10.9.4). Then W5 = -i(WioW 1 +W 1 oWi) is exactly the local function Hs in the expression (6.7.3). The similar splitting of a super-Hamiltonian is the corner stone of the SUSY mechanics [ill; 265]. 6.8
BRST complex of constrained systems
We start with conservative constrained systems on a Poisson manifold and, afterwards, consider Lagrangian constraints (2.5.39) in time-dependent mechanics in Section 2.5D. Given a constrained system on a Poisson manifold (Z, {,}), one can associate to this system a BRST charge 0 by introducing a set of ghosts and antighosts and extending the Poisson bracket {,} to a superbracket [, ] [156; 245; 404]. The BRST charge obeys the condition [0,0] = 0 and defines the nilpotent classical BRST operator D = [6,.], which can be quantized [153]. In order to obtain the BRST charge, one starts with constructing the Koszul-Tate complex of constraints. Let Z be a (finite-dimensional) smooth manifold and J an ideal of the associative ring CO°(Z) whose elements are regarded as constraints. Let J be finitely generated by elements {/i}, i = 1,... ,n. One associates to J the Grassmann C°°(Z)-algebra B° freely generated by odd elements {&?}. These elements are treated as antighosts of level 0, and are endowed with the antighost number 1. The algebra B° is provided with the bigraded structure B° = C°°(Z)®B%®---B°
(6.8.1)
with respect to the antighost number agh= l , . . . , n and the Grassmann parity [&] = (agh(6) + l)mod2,
b G B°.
The algebra B° is brought into the chain complex
0 ^C°°(Z) J-B° J-
B° *— 0
(6.8.2)
393
Chapter 6 Supergeometry
with respect to the C°° (Z)-linear boundary operator 6(b°) = fh
(6.8.3)
S(b A b') = 5{b) A b' + (-lfh
A 6{b'),
b, b' e B°.
This is exactly the Koszul boundary operator (1.5.19) - (1.5.20), and the chain complex (6.8.1) exemplifies the Koszul complex (1.5.21) in Section 1.5B. One obtains at once its homology group (6.8.4)
H0(Bi) = C°°(Z)/J.
For instance, if J = /JV is the ideal of functions vanishing on a closed imbedded submanifold N of Z, then (6.8.5)
H0(B°)=C°°(N).
One says that the constraints {/j} are irreducible if the equality g'fi = 0,
g{£ C°°(Z),
(6.8.6)
implies the equalities 7 ij = - 7 j i ,
9* =-?*!,,
lijeC°°(Z).
In this case, the tuple {ft} is regular in the ring C°°(Z) and, in accordance with Theorem 1.5.6, the homology groups Hp>o(B®) of the Koszul complex (6.8.2) are trivial. It follows that the Koszul complex (6.8.2) is acyclic and, thus, it provides a finite resolution of the ring CCO(Z)/J. In the case of reducible constraints {/j}, one can also obtain the resolution of the ring C°°(Z)/J by extending the chain complex (6.8.2) to the Koszul-Tate one as follows [156; 245]. If constraints {/j} are reducible, there exists a collection of functions x \ G C°°(Z), i\ = 1,... , m , which do not belong J \ 0, but XUiJi=0,
ii = l , . . . , n i ,
(6.8.7)
and the equality (6.8.6) implies the equalities Si = X 1 ! 1 P i l +7 i j 7j,
7
ij
= -7ji,
ph,Jij
eC°°(Z).
(6.8.8)
In this case, the term B® of the Koszul complex (6.8.2) contains the homologically non-trivial cycles generated by the elements X1^,
i1 = l,...,n1.
(6.8.9)
394
Geometric and Algebraic Topological Methods in Quantum Mechanics
In order to remove these non-trivial cycles, let us extend the algebra B° to the C°°(Z)-algebra B1 of polynomials in the odd variables &9, i = 1,... ,n, of antighost number 1 and the even ones b}^, i\ = 1,... , m, of antighost number 2. Then B1 is a bigraded algebra B1 = C°°(Z) © Bj © B* e • • •
(6.8.10)
with respect to the antighost number and the Grassmann parity [b] = (agh(b) + l)mod 2,
beB1.
For instance, B\ = Bj. The Koszul boundary operator 5 (6.8.3) is extended to B1 as a C°°(Z)-linear Z2-graded derivation S1 mapping b\t into the nontrivial cycles (6.8.9), i.e., ii = 1 ni, S\bil)=xUilbl S\b A b') = 8(b) A 6' + (-l)M& A 5(b'),
(6.8.11) b, b' e B 1 .
This boundary operator makes the bigraded algebra B1 (6.8.10) into the chain complex 0 ^— C°°(Z) £- B\ £- B\ A
.
(6.8.12)
The homology group HQ{B\) of this chain complex equals the homology group H0(B°) (6.8.4) of the Koszul complex, while the first homology group Hi(Bl) is trivial. However, there may be nontrivial cycles in B\. This procedure can be iterated to higher antighost numbers. Let us assume that we have constructed the collection of chain complexes Bm, m = 0,..., L, graded in the antighost number k = 0 , 1 , . . . such that: • the complex Bm consists of polynomials with coefficients in C°°(Z) over the antighosts bik,
H = l,...,nfc,
k =
0,...,m,
of antighost number k + 1 and Grassmann parity (fc + I)mod2; • B^ = B*-1 for all k = 1,... m; • Hk{Bm) = 0 for 0 < k < m, while B%+1, m = 1,... ,L, contains non-trivial cycles generated by the elements of the form X
m+lim tm _i_ im + l »m ~l~
pm+1 tm+l'
where functions X m+1 i" +1 € C°°(Z) do not belong to J \ 0, but
xm+i£+1xrir1=0modJ
Chapter 6 Supergeometry
395
and the equality
implies the equalities
• the coboundary operator 6m of the complex Bm is the extension of the coboundary operator 5m~1 of the complex B"1'1 by the rule
Given the chain complex (BL,8L), one can prove the existence of the complex (BL+1, SL+1) which obeys the above mentioned conditions for m = L + 1 [156]. Let us emphasize the key points of the proof. (i) One assumes that the constraints {fi} can be locally split into the independent and dependent ones. For instance, this is the case of the ideal J — IN of functions vanishing on a closed imbedded submanifold N of Z. (ii) One shows that any <5fe-closed element of Bk is a (5fc+1-exact element of£ f c + 1 . (iii) If all functions XL+17i+1
= 0,
(6.8.13)
then the complex BL is acyclic. If the equalities (6.8.13) hold, the constraints are called finitely reducible. In this case, the iteration procedure is finite, and the complex (BL,SL) forms an acyclic resolution of C°°(Z)/J. It is called the Koszul-Tate complex with respect to the Koszul-Tate differential SL. If the constrained system in question if infinitely reducible, the above mentioned procedure is iterated an infinite number of times by introducing an infinite number of antighosts. It also gives the Koszul-Tate resolution B°° of C°°(Z)/J. For instance, this is the case of second-class constraints [245]- We here restrict our consideration to finitely reducible constraints. Given the Koszul-Tate complex (BL,SL), one then aims to extend the algebra BL in order to replace the graded derivation SL with the inner one D = [9,.]. For this purpose, given the C°°(Z)-module Vfc of antighosts of level k, let us consider its dual Vfe* whose elements are called the ghosts of level k. It is a free module generated by the set of elements {c^}. The ghosts c^° possess the same Grassmann parity as the antighosts b\k and the antighost number 0, but they are characterized by the ghost number
396
Geometric and Algebraic Topological Methods in Quantum Mechanics
gh(c^) = k + 1. Accordingly, the antighosts bkk are endowed with the negative ghost number — (k + 1). Let us consider the C°°(Z)-aigebra CL of polynomials in ghosts Ck and antighosts bk, k = 0,... ,L. It is called the BRST algebra. This algebra is provided with the C°°(Z)-bilinear superbracket [, ]gh denned by the relations
M V = [c,c']gh = 0,
[ # , & £ > = *£•<&£,#) = SI-.
(6.8.14)
Given a Poisson bracket {,} on a manifold Z, the superbracket (6.8.14) is extended to the M-bilinear superbracket [»] = {.} +Ugh
(6-8.15)
The Koszul-Tate differential 8L is prolonged onto CL by letting 5Lc = 0 for all ghosts c. Let C denote the C°°(Z)-linear restriction morphism CL —> BL by One can show the following [156; putting ck = 0, k = Q,...,L. 245]. 6.8.1. Let J be a coisotropic ideal of the Poisson algebra (C°°(Z),{,}). There exists an element 6 e CL of ghost number 1 called the BRST charge such that: • [0,0] = 0 where [,] is the superbracket (6.8.15); THEOREM
• the operator, called the BRST differential, aeCL, £>(a) = [0,a],
(6.8.16)
is a nilpotent graded derivation of the R-algebra CL; • the operator BL9&^C([e,&]gh)e£L coincides the Koszul-Tate differential 8L on BL; • the BRST charge takes the coordinate form
e = 4/i+Ectx fc S- 16 t- 1l + efcc fc=i
where etc. consists of terms with at least two ghosts and one antighost or terms with at least two antighosts and one ghost; • the BRST charge is defined up to the Koszul-Tate boundary 5La, D where a E CL is of ghost number 2.
397
Chapter 6 Supergeometry
The BRST differential makes the BRST algebra CL into a cochain complex graded by the ghost number. It is called the BRST complex. It is important that, at the zero ghost number, the cohomology group H°(CL) of the BRST complex equals J"/J, where J" is the normalize (2.1.24) of the ideal J, i.e., H°(CL) coincides with the reduction of the Poisson algebra C°°(Z) via the ideal J (see Definition 2.1.16). Example 6.8.1. Let {/,} be irreducible constraints on a manifold Z provided with the zero Poisson structure, i.e., {/,g} = 0 for all f,g £ C°°{Z). In this case, the BRST algebra C° is generated by odd ghosts CQ of ghost number 1 and odd antighosts 6° of antighost number 1 and ghost number -1. The Koszul-Tate differential S° is the Koszul boundary operator 5°(b°i) = fi,
S°(4) = 0-
(6-8.17)
In this case, the BRST charge and the BRST differential (6.8.16) read
0 =/4,
D = 6°. a
Example 6.8.2. Let Z be a symplectic manifold which admits a Hamiltonian action ££ of a finite-dimensional real Lie algebra $j. Let {e,} be a basis for Q and fi the functions on Z possessing £Ei as their Hamiltonian vector fields (see Proposition 2.3.3). The functions fi constitute the Lie algebra g with respect to the Poisson bracket {fiJj} = <$jfk, where c^ are constant on Z. Let the representation e i-> ££ be a Lie algebra isomorphism. Then {/»} is an irreducible system. As in Example 6.8.1, the BRST algebra C° is generated by ghosts cj, and antighosts b°, and the Koszul-Tate differential 5° on CL takes the form (6.8.17). The BRST charge is
© = fA - \ckiA4A-
(6-8.18)
The corresponding BRST differential reads
D(f) = {/„ f}cl
D($) =U- 4^6°,
D(4) =
-\AA^
398
Geometric and Algebraic Topological Methods in Quantum Mechanics
It can be written in the form D = 5 + d + etc., where 5 is the Koszul boundary operator (6.8.17) and d is the Chevalley-Eilenberg differential • (1.5.55). Turn now to Lagrangian constraints Ki=Pi-Oij<4kPk=0
(6.8.19)
(2.5.39) in Hamiltonian time-dependent mechanics on a configuration bundle Q —> R (see Section 2.5D). We aim to construct their Koszul-Tate and BRST complexes. In accordance with the splitting Pi = Hi + Vi = fa - Oij4kPk] + [oijOo'Pk]-
(6-8.20)
(2.5.37b) and the relation aij
(6.8.21)
(2.5.32), let us introduce the projection operators
R = Si- aijalk,
P* =
k aij4
such that PfUk = 0,
RkUk = TZi.
(6.8.22)
A glance at the relations (6.8.22) shows that the constraints TZi (6.8.20) are reducible. Moreover, since P*R) = 0,
(6.8.23)
these constraints are infinitely reducible in general. The corresponding algebra B of antighosts is generated by the elements &£, r = 0 , 1 , . . . of antighost number agh(&£) = r + 1 and Grassmann parity (r + I)mod2. They are introduced as follows [295]. Given the vertical tangent bundle VQV*Q of V*Q —> Q, let us consider the infinite Whitney sum E = E0®E1
= VQV*Q © VQV*Q®--V*Q
(6.8.24)
over V*Q of the copies of VQV*Q. This vector bundle is provided with the holonomic coordinates (t,qi,pi,p'i), r = 0 , 1 , . . . , where (t,q\pi,pfr+1), are
399
Chapter 6 Supergeometry
coordinates on Eo, while {t,ql,Pi,pfr) are those on E\. Following Remark 6.4.2, let us define the quotient E of the tensor product Er
= R © E* V*Q
© ®E*
© ••• © <£>E*
V'Q
V'Q
V'Q
by the elements ei
e0 ® e'o - e'o ® e0
e0 ® ei - ei ® e0
for all eo,e(, G ^op, ei,ei G Ei p , and p G V*Q. The Croo(V*Q)-modules Br of sections of the vector bundles E —> V*Q make up a direct system with respect to the natural monomorphisms Br —> Br+1. Its direct limit B is endowed with a structure of a graded commutative C°°(V*Q)-ring generated by odd and even elements b\, r = 0,1,..., which have the same linear coordinate transformation law as the coordinates pi. Furthermore, one defines the graded differential calculus over the graded commutative C°°(y*Q)-ring B. The C°°(^*(5)-module B is graded by the antighost number as B = C°°{V*Q) ®B° ®Bl © • • • , and is brought into a chain complex ^Af1
0 ^C°°(V*Q)
<
(6.8.25)
with respect to the Koszul-Tate differential 5 : C°°(V*Q) 6(b2ir+1)=Plkb2kr,
-
0,
S(b°)=R^pk, r+2
5{bl
)
= RkiblT+\
r>0.
(6.8.26)
The nilpotency property 8 o S = 0 of this differential is the corollary of the relations (6.8.23). 6.8.2. The chain complex (6.8.25) with respect to the differential (6.8.26) is the Koszul-Tate resolution, i.e., its homology groups are
PROPOSITION
#fc>o = O,
HQ = C°°(V*Q)/IN
= COO(NL).
a Let us note that, in particular cases of the degenerate quadratic Lagrangian (2.5.24), the complex (6.8.25) may have a subcomplex, which is also the Koszul-Tate resolution. For instance, if the fibre metric a in
400
Geometric and Algebraic Topological Methods in Quantum Mechanics
VQ —> Q is diagonal with respect to a holonomic atlas of VQ, the constraints (6.8.19) are irreducible and the complex (6.8.25) contains a subcomplex which consists only of the antighosts 6°. Now, we aim to construct the BRST charge G such that {e,9}
s
= 0,
8(b) = { Q , b } s ,
b € B
with respect to some Poisson superbracket {, }s- The problem is to find the Poisson superbracket such that {f,g}s = 0 for all f,g e C°°(V*Q). In order to solve this problem, we follow the SUSY extension of Hamiltonian formalism in Section 6.7. We assume that Q —> R is a vector bundle, and further denote II = V*Q. Let us consider the vertical tangent bundles VII and W I I . The latter admits the canonical decomposition
vvn = vn © vn ^ vn.
(6.8.27)
Let choose the bundle E = Eo © E\ over Vn which is the infinite Whitney sum over VII of the copies of VVII. In view of the decomposition (6.8.27), we have
E = vn©vn©---p-^vn. R
This vector bundle is provided with the holonomic coordinates {t,q\Pi,qr,Vri), r = 0' 1 )---! where {qlr+nPT^) a r e bundle coordinates on EQ —> VII and {q\r,p\T) are those on Ex -* VII. As a repetition of the construction of B, we define the graded commutative C°°(VII)-ring C generated by odd and even elements (6£, bt, clr, clr), r = 0,1,..., where: • c*, clr are ghosts of ghost number r + 1 and Grassmann parity (r + I)mod2, • b\, VI are antighosts of antighost number r +1, ghost number — (r+1) and Grassmann parity (r + I)mod2. One can think of C as being the BRST algebra. Furthermore, we define the graded differential calculus over the graded commutative C°°(V*Q)ring C. It contains the graded three-form tts = [dpi A dqi + dpi A dql + oo
i Yffi r=0
A ~H
- dbl A d°r)\ A dt>
401
Chapter 6 Supergeometry
generalizing the graded three-form (6.7.2). The corresponding Poisson superbracket on C reads
{c,c'}s = {c,c'}v + i J2(-l)[f]{r+1)[^~ r=0 (_1)r+1dc_d£
_ dc_ dd_ _
(6-8.28)
+
Ot)i
r
+1
dc_dd_
where {, }y is the Poisson bracket (2.4.81). Then the desired BRST charge takes the form
e = i[cjii*pfc + £ ( 4 r + i W + 4- + 2*W +1 )]r=0
The corresponding BRST differential D = {6, .}s reads D ( ^ ) = 0,
D(bl) = 6(bl), i?(4-) = ^ r + l .
6.9
r = 0,l,...,
^>(4r+l) = - 4 4 + 2 ,
£>(<£) = 0.
Appendix. Supermanifolds
Subsections: A. Superfunctions, 401; B. Supermanifolds, 405; C. DeWitt supermanifolds, 409; D. Supervector bundies, 411; E. Superconnections, 415; F. Principal superconnections, 417. There are different types of supermanifolds. These are H°°-, G°°-, GH°°-, G-, and DeWitt supermanifolds [21]. By analogy with smooth manifolds, supermanifolds are constructed by gluing together of open subsets of supervector spaces Bn'm with the aid of transition superfunctions. Therefore, let us start with the notion of a superfunction. A. Superfunctions Though there are different classes of superfunctions, they can be introduced in the same manner as follows. Let Bn<m = A£©A?\
n,m>0,
402
Geometric and Algebraic Topological Methods in Quantum Mechanics
be a supervector space, where A is a Grassmann algebra of rank 0 < N > m. Let an,m
_>
. Bn,m
Rn^
g
. Q^m
_>
Rn,m
=
Rn
Q
Rrn
be the corresponding body and soul maps (see the decomposition (6.1.2)). Then any element q £ Bn>m is uniquely split as q = (x,y) = (a(i 4 ) + s(xi))e? +yje),
(6.9.1)
where {e?, e]} is a basis for Bn'm and a(xi) € R, six*) £ Ro, yj £ Ri. Let A' be another Grassmann algebra of rank 0 < N' < N which is treated as a subalgebra of A, i.e., the basis {ca}, a — 1,..., N', for A' is a subset of the basis {c*}, i = 1,..., N, for A. Given an open subset U C M", let us consider a A'-valued function N' fc=0
(6.9.2) '
on U with smooth coefficients fai-.akiz), z £U. It is a graded function on U. Its prolongation to (cr"' 0 )-" 1 ^) C Bn'° is defined as the formal Taylor series i r
N>
fc=o
N
' LP=O
i
F>vf
y'
(6.9.3) Then a superfunction F(q) on (cr"' 7 ")-^^) C Bn'm is given by a sum m
F{q) = F(a;,i/) = ]T - / ^ . . j , ^ • • V ' , r=0
(6.9.4)
'
where fj1...jrix) are functions (6.9.3). However, the representation of a superfunction Fix, y) by the sum (6.9.4) need not be unique. The germs of superfunctions (6.9.4) constitute the sheaf SJV of graded commutative A'-algebras on Bn<m, but it is not a sheaf of Cg^.m-modules since superfunctions are expressed in Taylor series. Using the representation (6.9.4), one can define derivatives of superfunctions as follows. Let fix) be a superfunction on Bn<°. Since / , by
403
Chapter 6 Supergeometry
definition, is the Taylor series (6.9.3), its partial derivative along an even coordinate xl is defined in a natural way as dif(x) = (dif)(a(x), s(x)) = N
'
i r
N
(6.9.5)
1
i
8P+ 1
L, k\ [L p\ dzidzii...dziP
MX)MX )
*& Y
c .
This even derivative is extended to superfunctions F on Bn'm in spite of the fact that the representation (6.9.4) is not necessarily unique. However, the definition of odd derivatives of superfunctions is more intricate. Let &°N, C &N1 be the subsheaf of superfunctions F(x, y) = f(x) (6.9.3) independent of the odd arguments yi. Let AMm be a Grassmann algebra generated by ( a 1 , . . . , am). The expression (6.9.4) implies that, for any open subset U C Bn'm, there exists the sheaf morphism A : 6 ^ ,
X X
( ^)
:
E 3 / * - * • ( * ) ® K • --ajr) r=0
m
(6.9.6)
-
(6-9-7)
1
r=0
'
n m
over 5 - . Clearly, the morphism A (6.9.6) is an epimorphism. One can show that this epimorphism is injective and, consequently, is an isomorphism if and only if iV - TV' > m
(6.9.8)
[21]. Roughly speaking, in this case, there exists a tuple of elements y*1,..., yir € A for each superfunction / such that \(f®(ajl at t h e point (x, yjl,...,
•••«>)) ^ 0
yjm) of Bn<m.
If the condition (6.9.8) holds, the representation of each superfunction F(x, y) by the sum (6.9.4) is unique, and it is an image of some section / ® a of the sheaf 6 ^ , ® ARm with respect to the morphism A (6.9.7). Then an odd derivative of F is defined as
^y(A(/®y)) = A(/®A(a)).
404
Geometric and Algebraic Topological Methods in Quantum Mechanics
This definition is consistent only if A is an isomorphism, i.e., the relation (6.9.8) holds. If otherwise, there exists a non-vanishing element / ® a such that A(/®a) = 0 , whereas A(/®a,(a))#0. For instance, if N-N'
=
m-\,
such an element is /<8>a = c 1 ---c w '®(a 1 ---a i n ). In order to classify superfunctions, we follow the terminology of [21; 377]. • If N' = N, one deals with G°° -superfunctions, introduced in [376], In this case, the inequality (6.9.8) is not satisfied, unless m = 0. • If the condition (6.9.8) holds, SJV = QHN' is the sheaf of GH°°superfunctions. • In particular, if N' = 0, the condition (6.9.8) is satisfied, and &N> — H°° is the sheaf of H°° -superfunctions
r=0
Lp=O
(6.9.9) where fj-i...jr are real functions [25; 126]. Superfunctions of the above three types are called smooth superfunctions. The fourth type of superfunctions is the following. Given the sheaf GTi-N' of Gfi"°°-superfunctions on a supervector space Bn'm, let us define the sheaf of graded commutative A-algebras QN, =gnN> ®A, A'
(6.9.10)
where A is regarded as a graded commutative A'-algebra. The sheaf Q^1 (6.9.10) possesses the following important properties [2l]. • There is the evaluation morphism 5:GN' 3 F®a>-> Fa£C£n,m,
(6.9.11)
Chapter 6 Supergeometry
405
where
is the sheaf of continuous A-valued functions on Bn>m. Its image is isoraorphic to the sheaf G°° of G°°-sup erfunctions on Bn'm. • For any two integers N' and N" satisfying the condition (6.9.8), there exists the canonical isomorphism between the sheaves GN' and GN" • Therefore, one can define the canonical sheaf Gn,m of graded commutative A-algebras on the supervector space Bn'm whose sections can be seen as tensor products F($>a of iJ°°-superfunctions F (6.9.9) and elements a £ A. They are called G-superfunctions. • The sheaf 5^ n , m of graded derivations of the sheaf Gn,m is a locally free sheaf of 5 nm -modules of rank (n, m). On any open set U C Bn'm, the Gn,m{Uymod\i\e dGn,m{U) is generated by the derivations d/dxl, d/dyj which act on Gn,m{U) by t h e rule 7— ( F <8> a) = 77--
^:Uc^Bn'm,
such that the transition functions (p^ o (p'T1 are supersmooth. Obviously, a smooth supermanifold of dimension (n, m) is also a real smooth manifold of dimension 2N~1(n + m). If transition superfunctions are H00-, G°°- or G#°°-superfunctions, one deals with H°°-, G°°- or Gff^-supermanifolds, respectively. By virtue of Theorem 1.8.2 extended to graded local-ringed spaces, this definition is equivalent to the following one. 6.9.1. A smooth supermanifold is a graded local-ringed space (M,&) which is locally isomorphic to (Bn'm,S), where S is one of the sheaves of smooth superfunctions on Bn'm. The sheaf S is called the structure sheaf of a smooth supermanifold. • DEFINITION
406
Geometric and Algebraic Topological Methods in Quantum Mechanics
In accordance with Definition 6.9.1, by a morphism of smooth supermanifolds is meant their morphism (
M — ° M <& A
is sheaf of continuous A-valued functions on M, and S is locally isomorphic to the evaluation morphism (6.9.11). • Example 6.9.1. The triple (Bn'm,Gn,m,5), where S is the evaluation morphism (6.9.11), is called the standard G-supermanifold. For any open subset U c Bn
C°°(a ' {U))
(6.9.13)
N+m
®AR
.
a Remark 6.9.2. Any Gif^-supermanifold (M, GH^) with the structure sheaf GH'M is naturally extended to the G-supermanifold (M, GH'jfi
Chapter 6 Supergeometry
407
(M,5(GM)), where S(GM) — G^ is the sheaf of G°°-superfunctions on M. 0 As in the case of smooth supermanifolds, the underlying space M of a G-supermanifold (M,GM) is provided with the structure of a real smooth manifold of dimension 2N~1(n + m), and morphisms of G-supermanifolds are smooth morphisms of the underlying smooth manifolds. However, it may happen that non-isomorphic G-supermanifold have isomorphic underlying smooth manifolds. Remark 6.9.3. Let us present briefly the axiomatic approach to supermanifolds which enables one to obtain all the previously known types of supermanifolds in terms of i?°°-supermanifolds [21; 22; 69]. This approach to supermanifolds refines that in [378]. The i?°°-supermanifolds are introduced over the above mentioned Arens-Michael algebras of Grassmann origin [69], but we here omit the topological side of their definition, though just the topological properties differ i?°°-supermanifolds from Rsupermanifolds in [378]. Let A be a graded commutative algebra of the above mentioned type (for the sake of simplicity, the reader can think of A as being a Grassmann algebra). A superspace over A is a triple (M,?R°°,6), where M is a paracompact topological space, 91°° is a sheaf of graded commutative Aalgebras, and S : $H°° —> C^ is an evaluation morphism to the sheaf C^ of continuous A-valued functions on M. Sections of EH°° are called R°°superfunctions. Let Mq denote the ideal of the stalk fH£°, q e M, formed by the germs of i?°°-superfunctions / vanishing at a point q, i.e., such that <5(/)(?) = 0. An i?°°-super-manifold of dimension (n, m) is a superspace (M, 91°°, S) satisfying the following four axioms [69]. Axiom 1. The graded O^-dual (59^°°)* of the sheaf of derivations is a locally free sheaf of graded ^""-modules of rank (n,m). Every point q 6 M has an open neighborhood U with sections x1, • • • ,xn G 9\°°(U)o, y1, • • • , ym € m°°(U)i such that {dx\ dyj} is a graded basis for (W\°°)*(U) over m°°(U). Axiom 2. Given the above mentioned coordinate chart, the assignment determines a homeomorphism of U onto an open subset of Bn'm. Axiom 3. For every q G M, the ideal Mq is finitely generated. Axiom 4. For every open subset U
408
Geometric and Algebraic Topological Methods in Quantum Mechanics
An .R-supermanifold over a graded commutative Banach algebra satisfying Axiom 4 is an -R^-supermanifold. The standard G-supermanifold in Example 6.9.1 is an i?°°-supermanifold. Moreover, in the case of a finite Grassmann algebra A, the category of i?°°-supermanifolds and the category • of G-supermanifolds are equivalent. Let (M, GM) be a G-supermanifold. As was mentioned above, it satisfies the Axioms 1-4. Sections u of the sheaf C)GM of graded derivations are called super-vectorfieldson the G-supermanifold (M,GM), while sections <j> of the dual sheaf VG*M are one-superforms on (M, GM). Given a coordinate chart (
<j> =
where coefficients ul and fa are G-superfunctions on U. The graded differential calculus in supervector fields and superforms obeys the standard formulae (6.2.4), (6.2.16), (6.2.17) and (6.2.18). Given a GLet us consider cohomology of G-supermanifolds. supermanifold (M,GM),
let S A M = OJW® A
be the sheaves of smooth A-valued exterior forms on M. These sheaves are fine, and they constitute the fine resolution 0^A-*Cfi®A->D1M®A-*--of the constant sheaf A on M. We have the corresponding de Rham complex
0-A->CnM)-»OyUM)-"of A-valued exterior forms on M. By virtue of Theorem 1.7.6, the cohomology H^ (M) of this complex is isomorphic to the sheaf cohomology H* (M; A) of M with coefficients in the constant sheaf A and, consequently, is related to the de Rham cohomology as follows: HX(M) = H*{M;A) = H*(M)®A.
(6.9.14)
Thus, the cohomology groups of A-valued exterior forms do not provide us with information on the G-supermanifold structure of M. Let us turn to cohomology of superforms on a G-supermanifold k
(M,GM)-
The sheaves ADG^f of superforms constitute the complex 0-+ A ^ G M - ^ * G M - * ••• •
(6.9.15)
Chapter 6 Supergeometry
409
The Poincare lemma for superforms is proved to hold [21; 68], and this complex is exact. However, the structure sheaf GM need not be acyclic, and the exact sequence (6.9.15) fails to be a resolution of the constant sheaf A on M in general. Therefore, the cohomology H*S{M) of the de Rham complex of superforms are not equal to cohomology H*(M; A) of M with coefficients in the constant sheaf A, and need not be related to the de Rham cohomology H*{M) of the smooth manifold M. In particular, cohomology Hg(M) is not a topological invariant, but it is invariant under G-isomorphisms of G-supermanifolds. 6.9.3. The structure sheaf Qn,m of the standard Gsupermanifold (Bn>m, Gn,m) is acyclic, i.e.,
PROPOSITION
n
' , yn,m)
\B
— u.
D The proof is based on the isomorphism (6.9.13) and some cohomological constructions [21; 69]. C. DeWitt super manifolds There exists a particular class of supermanifolds, called DeWitt supermanifolds. Their notion implies that a supervector space Bn'm is provided with the De Witt topology, which is coarser than the Euclidean one. This is the coarsest topology such that the body map an,m
. Bn,m _^ R n
(6.9.16)
is continuous. The open sets in the DeWitt topology are of the form V x Tln'm, where V are open sets in Rn. Clearly, this topology is not Hausdorff. DEFINITION 6.9.4. A smooth supermanifold (resp. a G-supermanifold) is said to be a De Witt supermanifold if it admits an atlas such that the local morphisms <j>^ : U^ —> Bn'm in Definition 6.9.1 (resp. Definition 6.9.2) are continuous with respect to the DeWitt topology, i.e., ^(U^) C Bn'm are open in this topology. •
Given an atlas (U^,<j>^) of a DeWitt supermanifold in accordance with Definition 6.9.4, it is readily observed that its transition functions 4>^o(p~1 must preserve the fibration an'm (6.9.16) whose fibre (an'm)~1 (z) over z €
410
Geometric and Algebraic Topological Methods in Quantum Mechanics
Rn is equipped with the coarsest topology, where only 0 and (
Then by virtue of Batchelor's Theorem 6.3.1 and Theorem 6.9.5, there is one-to-one correspondence between the isomorphism classes of DeWitt H°°supermanifolds of odd rank m with a body manifold Z and the equivalence classes of m-dimensional vector bundles over Z. This result is extended to DeWitt GH°°-, G°°- and G-supermanifolds because their isomorphism classes are in one-to-one correspondence with isomorphism classes of DeWitt i?°°-supermanifolds [2l]. Let us say something more on DeWitt G-supermanifolds. PROPOSITION 6.9.6. The structure sheaf GM of a DeWitt G-supermanifold is acyclic, and so is any locally free sheaf of graded GM-modules [21; 69].
•
PROPOSITION 6.9.7. There is an isomorphism of the cohomology Hg(M) of superforms on a DeWitt G-supermanifold to the cohomology (6.9.14) of A-
411
Chapter 6 Supergeometry
valued exterior forms on its body manifold [363].
ZM,
i-e., Hg(M) =
H*{ZM)®^-
•
These results are based on the fact that the structure sheaf GM on M, provided with the DeWitt topology, is fine. However, this does not imply automatically that GM is acyclic since the DeWitt topology is not paracompact. Nevertheless, it follows that the image
(6.9.17)
gn,m(U)®Gr,s(V),
where
M',GM§>GM'),
where
412 GM®GM'
Geometric and Algebraic Topological Methods in Quantum Mechanics
is the sheaf determined by the presheaf
UxU'
^GM(U)®GM>(U'),
5 : GM{U)®GMI{U') - CSfogC^,) =
C^myiauVJI),
for any open subsets U C M and U' C M ' . This product is a Gsupermanifold of dimension (n + r,m + s) [21]. Furthermore, there is the epimorphism pri
: {M,GM) x (M',GM>) -* (M,GM).
One may define its section over an open subset U C M as the Gsupermanifold morphism su • {U,GM\u)
-> ( M , G M ) x
(M'.GM')
such that pr1 o sy is the identity morphism of (C/, G M It/)- Sections su over all open subsets U C M determine a sheaf on M . This sheaf should be provided with a suitable graded commutative GM-structure. For this purpose, let us consider the product (M,GM)x(Br^,gris),
(6.9.18)
where Br^s is the superspace (6.1.4). It is called a product G-supermanifold. are isomorphic, Br^s has a natural Since the Ao-modules Br\" and Br+3'r+s structure of an (r + s,r + s)-dimensional G-supermanifold. Because Br\8 is a free graded A-module of the type (r, s), the sheaf S^ of sections of the fibration (M, GM) x (Br^, gr]s) -> (M, G M )
(6.9.19)
has the structure of the sheaf of free graded GM-modules of rank (r,s). Conversely, given a G-supermanifold (M, GM) a n d a sheaf S of free graded Gjw-modules of rank (r, s) on M, there exists a product G-supermanifold (6.9.18) such that S is isomorphic to the sheaf of sections of the fibration (6.9.19). Let us turn now to the notion of a supervector bundle over Gsupermanifolds. Similarly to smooth vector bundles (see Theorem 1.8.5), one can require of the category of supervector bundles over Gsupermanifolds to be equivalent to the category of locally free sheaves of graded modules on G-supermanifolds. Therefore, we can restrict ourselves to locally trivial supervector bundles with the standard fibre Br^s.
413
Chapter 6 Supergeometry
DEFINITION 6.9.8. A supervector bundle over a G-supermanifold (M,GM) with the standard fibre (Br^,Gr\s) 1S denned as a pair ((Y,Gy),7r) of a G-supermanifold (Y,Gy) and a G-epimorphism
n : (Y,GY) -+ (M,GM)
(6.9.20)
such that M admits an atlas {(U^, V'c) °f local G-isomorphisms A • (^" 1 (^c)^y| 7 r -i(c/ c )) -> {U(,GM\U()
x
(Br^,gr]s).
a It is clear that sections of the supervector bundle (6.9.20) constitute a locally free sheaf of graded Gjvf-modules. The converse of this fact is the following [21]. 6.9.9. For any locally free sheaf S of graded GM-modules of rank (r, s) on a G-supermanifold (M, GM), there exists a supervector bundle over (M, GM) such that 5 is isomorphic to the structure sheaf of its sections. •
THEOREM
The fibre Yq, q G M, of the supervector bundle in Theorem 6.9.9 is the quotient SJMq
<* S%/{Mq • SMsq) * Br^
of the stalk Sg by the submodule Aiq of the germs s € Sq whose evaluation "K/X?) vanishes. This fibre is a graded A-module isomorphic to Br^s, and is provided with the structure of the standard G-supermanifold. Remark 6.9.4. The proof of Theorem 6.9.9 is based on the fact that, given the transition functions p^£ of the sheaf S, their evaluations 9a = HPtt)
(6-9.21)
define the morphisms U(nUt-+GL{r\s;A), and they are assembled into a cocycle of G°°-morphisms from M to the general linear graded group GL(r\s;A). Thus, we come to the notion of a G°°-vector bundle. Its definition is a repetition of Definition 6.9.8 if one replaces G-supermanifolds and G-morphisms with the G°°- ones. Moreover, the G°°-supermanifold underlying a supervector bundle (see Remark 6.9.2) is a G°°-supervector bundle, whose transition functions g^ are related to
414
Geometric and Algebraic Topological Methods in Quantum Mechanics
those of the supervector bundle by the evaluation morphisms (6.9.21), and are GL(r\s; A)-valued transition functions. • Since the category of supervector bundles over a G-supermanifold (M, GM) is equivalent to the category of locally free sheaves of graded C?M-modules, one can define the usual operations of direct sum, tensor product, etc. of supervector bundles. Let us note that any supervector bundle admits the canonical global zero section. Any section of the supervector bundle TT (6.9.20), restricted to its trivialization chart (U,GM\u)x(Br^,grls),
(6.9.22)
is represented by a sum s = sa(q)ea, where {ea} is the basis for the graded A-module Br\s, while sa(q) are G-superfunctions on U. Given another trivialization chart U' of n, a transition function s'b(q)e'b = sa(q)hba(q)eb,
(6.9.23)
q€UnU', b
is given by the (r + s) x (r + s) matrix h whose entries h a(q) are Gsuperfunctions on U D U'. One can think of this matrix as being a section of the supervector bundle over U D U with the above mentioned group GL(r\s; A) as a typical fibre. Example 6.9.5. Given a G-supermanifold (M,GM), let us consider the locally free sheaf QGM of graded derivations of GM- In accordance with called supertanTheorem 6.9.9, there is a supervector bundle T(M,GM), gent bundle, whose structure sheaf is isomorphic to QGM- If (q1, • • •, qm+n) and (q11,..., q'm+n) are two coordinate charts on M, the Jacobian matrix , dqH /i*. = — ,
i,j = l , . . . , n + m,
(see the prescription (6.9.12)) provides the transition morphisms for T(M,GM). It should be emphasized that the underlying G°°-vector bundle of the called G°°-supertangent bundle, has the supertangent bundle T(M,GM), transition functions <$(/iJ) which cannot be written as the Jacobian matrices since the derivatives of G°°-superfunctions with respect to odd arguments • are ill denned and the sheaf OG^ is not locally free.
415
Chapter 6 Supergeometry
E. Superconnections Given a supervector bundle n (6.9.20) with the structure sheaf S, one can follow suit of Definition 1.8.10 and introduce a connection on this supervector bundle as a splitting of the the exact sequence of sheaves 0 -> X>G*M ® 5 -^{GM®X>G*M) ®S-^S^0.
(6.9.24)
Its splitting is an even sheaf morphism V :S->-0*GM®S
(6.9.25)
satisfying the Leibniz rule V ( / s ) = d / ® a + /V(s),
feGM(U),
s€S(U),
(6.9.26)
for any open subset U € M. The sheaf morphism (6.9.25) is called a superconnection on the supervector bundle n (6.9.20). Its curvature is given by the expression R = V 2 : S -> A dG*M ® S,
(6.9.27)
similar to the expression (1.8.26). The exact sequence (6.9.24) need not be split. One can apply the criterion in Section 1.8 in order to study the existence of a superconnection on supervector bundles. Namely, the exact sequence (6.9.24) leads to the exact sequence of sheaves 0 -» Horn (5, T)G*M ® S) -> Horn (5, (G M © QG*M) ® S) -» Horn (5,5) -> 0 and to the corresponding exact sequence of the cohomology groups 0 -> # ° ( M ; Horn (5, clC^ ® S)) -> ff°(M; Horn (5, (G M © 5G^) <8> S)) -> if o (M; Horn (5, 5)) -> F : ( M ; Horn (51, QG*M ® S1)) -» • • • . The exact sequence (6.9.24) defines the Atiyah class At(7r) Gi? 1 (M;Hom(S,DGM®S)) of the supervector bundle TT (6.9.20). If the Atiyah class vanishes, a superconnection on this supervector bundle exists. In particular, a superconnection exists if the cohomology set H1(M;Hom(S,QG*M ® S)) is trivial. In contrast with the sheaf of smooth functions, the structure sheaf GM of a G-supermanifold is not acyclic in general, cohomology
416
Geometric and Algebraic Topological Methods in Quantum Mechanics
H*(M; Horn (S, VG*M ® S)) is not trivial, and a supervector bundle need not admit a superconnection. Example 6.9.6. In accordance with Proposition 6.9.3, the structure sheaf of the standard G-supermanifold (Bn'm, Gn,m) is acyclic, and the trivial supervector bundle {Bn'm, Qn,m) X (B r|S , Qr\s) - (Bn'm, Gn,m)
(6.9.28)
has obviously a superconnection, e.g., the trivial superconnection.
•
Example 6.9.7. By virtue of Proposition 6.9.6, the structure sheaf of a DeWitt G-supermanifold (M,GM) is acyclic, and so is the sheaf Horn (S, f)G*M (g> S). It follows that any supervector bundle over a DeWitt G-supermanifold admits a superconnection. • Example 6.9.6 enables one to obtain a local coordinate expression for a superconnection on a supervector bundle n (6.9.20), whose typical fibre is Br\s and whose base is a G-supermanifold locally isomorphic to the standard G-supermanifold (B n ' m , Gn,m)- Let 17 C M (6.9.22) be a trivialization chart of this supervector bundle such that every section s of n\u is represented by a sum sa(q)ea, while the sheaf of one-superforms X)*GM\U has a local basis {dq1}. Then a superconnection V (6.9.25) restricted to this trivialization chart can be given by a collection of coefficients Via&: V(ea) = dq1 ® (Vi6oe6),
(6.9.29)
which are G-superfunctions on U. Bearing in mind the Leibniz rule (6.9.26), one can compute the coefficients of the curvature form (6.9.27) of the superconnection (6.9.29). We have R(ta) Rij\
= ^dq{ A dqj
Rijbaeb,
= (-l)Mb-]0. V / 6 - djViab + (-l)Wb]+M+W) V / f c VA (_1)W(H+W)Via)cVjfeb.
In a similar way, one can obtain the transformation law of the superconnection coefficients (6.9.29) under the transition morphisms (6.9.23). In particular, any trivial supervector bundle admits the trivial superconnection Viba = 0.
Chapter 6 Supergeometry
417
F. Principal superconnections In contrast with a supervector bundle, the structure sheaf Gp of a principal superbundle (P, Gp) —» (M, GM) is not a sheaf of locally free G^-modules in general. Therefore, the above technique of connections on modules and sheaves is not applied to principal superconnections in a straightforward way. Principal superconnections are introduced on principal superbundles by analogy with principal connections on smooth principal bundles [2l]. For the sake of simplicity, let us denote G-supermanifolds (M, GM) and their morphisms (ip:M->N,
$ : GN-+
by M and
DEFINITION
together with the natural identifications e x H = H x e = H, which satisfy the associativity m o (Id x in) = m o (m x Id) : H x H x H —> H x H —> H, the unit property (mo (ex Id))(e x H) = (mo (Id x e))(H x e) = Id if, and the inverse property (m o (S,ld))(H) = (fho (Id, S))(H) = e(e).
• Given a point g e H, let us denote by g : e —> H the G-supermanifold morphism whose range in H is g. Then one can introduce the notions of
418
Geometric and Algebraic Topological Methods in Quantum Mechanics
the left translation Lg and the right translation Rg as the G-supermanifold morphisms Lg:H = exHd^HxH I
Rg:H = Hxe ^HxH
^H, ^H.
Remark 6.9.8. Given a G-Lie supergroup H, the underlying smooth manifold H is provided with the structure of a real Lie group of dimension 2 (n + m), called the underlying Lit group. In particular, the actions on the underlying Lie group H, corresponding to the left and right translations • by g, are ordinary left and right translations by g. Let us reformulate the group axioms in Definition 6.9.10 in terms of the structure sheaf Ti of the G-Lie supergroup (H, "H). We observe that H has properties of a sheaf of graded Hopf algebras as follows. If {H,H) is a G-Lie supergroup, the structure sheaf % is provided with the sheaf morphisms: • a comultiplication in* :H —> m*(W
Chapter 6 Supergeometry
419
such that F(g,g') >-> F(gg'). It follows that GL(n\m;A) is an H°°-Lie supergroup. It is trivially extended to the G-Lie supergroup GL(n | m; A), O called the general linear supergroup. A Lie superalgebra h of a G-Lie supergroup H is defined as an algebra of left-invariant supervector fields on H. Let us recall that a supervector field u on a G-supermanifold H is a derivation of its structure sheaf H. It is called left-invariant if (Id
(6.9.30)
i.e., the sheaf of supervector fields on a G-Lie supergroup H is the globally free sheaf of graded W-modules of rank (n,m), generated by left-invariant supervector fields. The Lie superalgebra of right-invariant supervector fields on H is introduced in a similar way. Let us consider the right action of a G-Lie supergroup H on a Gsupermanifold P. This is a G-morphism p: P x H -> P such that po (p x Id) = p o (Id x m) : P x H x H -> P, po(ldxe)(Pxe)=ldP. The left action of H on P is defined similarly. Example 6.9.10. Obviously, a G-Lie supergroup acts on itself both on the left and on the right by the multiplication morphism m. The general linear supergroup GL(n\m; A) acts linearly on the standard supermanifold Bn\n on the left by the matrix multiplication which is a G• morphism. Let P and P' be G-supermanifolds that are acted on by the same G-Lie supergroup H. A G-supermanifold morphism (p : P —> P' is said to be
420
Geometric and Algebraic Topological Methods in Quantum Mechanics
if-invariant if
7rop = 7ropr1 :PxH^>M,
(6.9.31)
(ii) for every morphism (p : P —> • M' such that (p o p = <J? o pi-j, there is a unique morphism 7j : M —> M' with (p = 7j OK. D The quotient (M, 7?) does not necessarily exists. If it exists, there is a monomorphism of the structure sheaf GM of M into the direct image 7r*Gp. Since the G-Lie group H acts trivially on M, the range of this monomorphism is a subsheaf of ir^Gp, invariant under the action of H. Moreover, there is an isomorphism GM = {^GP)H
(6.9.32)
between GM and the subsheaf of Gp of if-invariant sections. The latter is generated by sections of Gp on •K~1{U), U C M, which are if-invariant as G-morphisms U —> A, where one takes the trivial action of H on A. Let us denote the morphism in the equality (6.9.31) by -d. It is readily observed that the invariant sections of Gp{-K~l(U)) are exactly the elements which have the same image under the morphisms p* : GP{K-X{U))
-+
pr? : GP(n-\U)) -
{H%GP)W-\U)), (WSGPXIT1^)).
Then the isomorphism (6.9.32) leads to the exact sequence of sheaves of A-modules on M 0 -^GM ^n*GPP*-^l-d*(GM®H).
(6.9.33)
DEFINITION 6.9.12. A principal superbundle of a G-Lie supergroup H is defined as a locally trivial quotient n : P —> M, i.e., there exists an open covering {U^} of M together with if-invariant isomorphisms
$(:P\0(-+U(xH,
Chapter 6 Supergeometry
421
where H acts on C/c x # -> [/c by the right multiplication.
(6.9.34) •
Remark 6.9.11. In fact, we need only the condition (i) in Definition 6.9.11 of the action of H on P and the condition of local triviality of P. • In an equivalent way, one can think of a principal superbundle as being glued together of trivial principal superbundles (6.9.34) by ^-invariant transition functions
which fulfill the cocycle condition. As in the case of smooth principal bundles, the following two types of supervector fields on a principal superbundle are introduced. DEFINITION 6.9.13. A supervector field u on a principal superbundle P is said to be invariant if
p* o u = (u
• One can associate to every open subset V C M the GM (V)-module of all iJ-invariant supervector fields on n~1(V), thus defining the sheaf DH(7r*Gp) of Gjw-niodules. DEFINITION 6.9.14. A fundamental supervector field v associated to an element v S f) is defined by the condition
v = (Id ® v) o p* : GP -• GP®e*(A) = GP.
n Fundamental supervector fields generate the sheaf VGp of Gp-modules of vertical supervector field on the principal superbundle P, i.e., UOTT* = 0 . Moreover, there is an isomorphism of sheaves of Gp-modules G p ® h 9 F ® w i - > F u G VGP, which is similar to the isomorphism (6.9.30).
422
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us consider the sheaf (n*VGP)H
= 7T*(VGP) n VH(n*GP)
on M whose sections are vertical if-invariant supervector fields. PROPOSITION 6.9.15. [21]. There is the exact sequence of sheaves of GMmodules 0 -> (n*VGp)H
-* UH{n*GP)
-» DGM -» 0.
(6.9.35)
D The exact sequence (6.9.35) is similar to the exact sequence (10.10.23) and the corresponding exact sequence of sheaves of C°°-modules 0 -> (VGP)x
- (TGP)x
- VC% - 0
in the case of smooth principal bundles. Accordingly, we come to the following definition of a superconnection on a principal superbundle. DEFINITION 6.9.16. A superconnection on a principal superbundle IT : P —> M (or simply a principal superconnection) is defined as a splitting V : DGM - * QH{K*GP)
(6.9.36)
of the exact sequence (6.9.35).
D
In contrast with principal connections on smooth principal bundles, principal superconnections on a //-principal superbundle need not exist. A principal superconnection can be described in terms of a h-valued one-superform w : DGp ^GP®\]
on P called a superconnection form. defines the morphism of Gp-modules
=
VGP,
Indeed, every splitting V (6.9.36) H
TT*{DGM) -> 7f*(5 (7r*GF)) ^ X)GP
which splits the exact sequence 0 -> VGP -> OGp -> Tf*(T)GM) -> 0. Therefore, there exists the exact sequence 0 -> ? * ( D G M ) V G P -» *GP
^ VGP
-» 0.
423
Chapter 6 Supergeometry
Let us note that, by analogy with associated smooth bundles, one can introduce associated superbundles and superconnections on these superbundles. In particular, every supervector bundle of fibre dimension (r, s) is a superbundle associated with GL(r\s; A)-principal superbundle [2l].
6.10
Appendix. Graded principal bundles
Graded principal bundles and connections on these bundles can be studied similarly to principal superbundles and principal superconnections, though the theory of graded principal bundles preceded that of principal superbundles [6; 258]. Therefore, we will touch on only a few elements of the graded bundle technique (see, e.g. [405] for a detailed exposition). Let (Z, 21) be a graded manifold of dimension (n, m). A useful object in the graded manifold theory, not mentioned above, is the finite dual 2l(Z)° of the algebra %l(Z) which consists of elements a of the dual 2l(Z)* vanishing on an ideal of 2l(Z) of finite codimension. This is a graded commutative coalgebra with the comultiplication (A°(o))(/®/'):=a(//'),
/,/'€»(£),
and the counit c°(a):=a(l a ). In particular, 2l(Z)° includes the evaluation elements 5Z such that
Given an evaluation element Sz, elements u € 2l(Z)° are called primitive elements with respect to 5Z if they obey the relation (6.10.1)
A°(v)=u®8z+Sz®u. These elements are derivations of 2l(Z) at z, i.e., « ( / / ' ) = ( « / ) & / ' ) + (-i)M[/]
(Szf)(uf).
DEFINITION 6.10.1. A graded Lie group (G,Q) is defined as a graded manifold such that G is an ordinary Lie group, the algebra Q(G) is a graded Hopf algebra (A, e, S), and the algebra epimorphism a : Q(G) -> C°°(G) is a morphism of graded Hopf algebras. •
424
Geometric and Algebraic Topological Methods in Quantum Mechanics
One can show that Q{G)° is also equipped with the structure of a Hopf algebra with the multiplication law o*d:=(o«i)oA,
a,beQ(G)°.
(6.10.2)
With respect to this multiplication, the evaluation elements Sg, Q € G, constitute a group 6g*5g> = 5ggi isomorphic to G. Therefore, they are also called group-like elements. It is readily observed that the set of primitive elements of Q(G)° with respect to 5e, i.e., the tangent space Te(G,Q) is a Lie superalgebra with respect to the multiplication (6.10.2). It is called the Lie superalgebra g of the graded Lie group (G, Q).
One says that a graded Lie group (G, Q) acts on a graded manifold (Z, 21) on the right if there exists a morphism (¥>,$): (Z, 21) x ( G , S ) ^ ( Z , 21) such that the corresponding algebra morphism $ : 2l(Z) -+ 2l(Z) ® Q(G) defines a structure of a right C/(G)-comodule on 2l(Z), i.e., (Id
(Id ® e) o $ = Id.
For a right action (tp, <E>) and for each element a G G(G)°, one can introduce the linear map $ a = (Id®a)o$:2l(Z)-^2l(Z).
(6.10.3)
In particular, if a is a primitive element with respect to 5e, then $ a G t)2l(Z). Let us consider a right action of (G, Q) on itself. If $ = A and a = 5g is a group-like element, then $ o (6.10.3) is a homogeneous graded algebra isomorphism of degree zero which corresponds to the right translation G —> Gg. If a Gfl,then $ a is a derivation of Q{G). Given a basis {UJ} for JJ, the derivations $ Ui constitute the global basis for dG(G), i.e., T)G{G) is a free left £/(G)-module. In particular, there is the decomposition g{G) = Q'{G) ®R G"(G), G'(G) - {/ G G(G) : $„(/) = 0, u G fl0}, G"(G) = {feG(G) : $ « ( / ) = 0, uGfli}. Since G'{G) = C°°{G), one finds that every graded Lie group (G,G) is the sheaf of sections of some trivial exterior bundle G x Q\ -* G [6; 59; 258].
Chapter 6 Supergeometry
425
Let us turn now to the notion of a graded principal bundle. A right action ((p, $) of (G, G) on (Z, 21) is called free if, for each z £ Z, the morphism $ z : 2l(Z) -> G(G) is such that the dual morphism $** : G(G)° -> 2l(Z)° is injective. A right action (?, $) of (G, G) on (Z, 21) is called regular if the morphism (if x prx) o A : (Z,21) x (G, 5) - (Z, 21) x (Z, 21) defines a closed graded submanifold of (Z,2l) x (Z,2l). Remark 6.10.1. Let us note that (Z',2t') is said to be a graded submanifold of (Z,2l) if there exists a morphism (Z',21') —> (Z, 21) such that the corresponding morphism 2l'(Z')° —> 2l(Z)° is an inclusion. A graded • submanifold is called closed if dim (Z1,21') < dim (Z, 21). Then we come to the following variant of the well-known theorem on the quotient of a graded manifold [6; 405]. THEOREM 6.10.2. A right action (?,$) of (G,G) on (Z,2l) is regular if and only if the quotient (Z/G, QL/G) is a graded manifold, i.e., there exists an epimorphism of graded manifolds (Z, 21) —> (Z/G, Ql/G) compatible with • the surjection Z —» Z/G.
In view of this Theorem, a graded principal bundle (P, 21) can be defined as a locally trivial submersion (P,B)->(P/G,a/0) with respect to the right regular free action of (G,Q) on (P, 21). In an equivalent way, one can say that a graded principal bundle is a graded manifold (P, 21) together with a free right action of a graded Lie group (G, Q) on (P, 21) such that the quotient (P/G, 2l/£) is a graded manifold and the natural surjection
(P,a)->(P/G,a/e) is a submersion. Obviously, P —> P/G is an ordinary G-principal bundle. A graded principal connection on a graded (G, S)-principal bundle (P, 21) —> (X, *8) can be introduced similarly to a superconnection on a
426
Geometric and Algebraic Topological Methods in Quantum Mechanics
principal superbundle. This is defined as a (G, £)-invariant splitting of the sheaf 921, and is represented by a g-valued graded connection form on (P, 21) [405]. Remark 6.10.2. In an alternative way, one can define graded connections on a graded bundle (Z, 21) —> (X, 03) as sections T of the jet graded bundle J\Z/X)^{Z,%) of sections of (Z, 21) -> {X, 03) [6], which is also a graded manifold [381]. In the case of a (G, £)-principal graded bundle, these sections T are required D to be (G, C/)-equivariant.
6.11
Appendix. The Ne'eman-Quillen superconnection
In this Section, we consider the class of superconnections introduced by Y.Ne'eman in the physical literature [337; 338] and by D.Quillen in the mathematical literature [306; 362]. The fibre bundles that they consider belong neither to above studied graded manifolds and bundles nor superbundles. Ne'eman-Quillen superconnections have been applied to computing the Chern character in K-theory (see below), to non-commutative geometry [342], BRST formalism [275] and some particle unification models [337], Let X be an iV-dimensional smooth manifold and /\T*X the exterior bundle. Let EQ, and E\ be two vector bundles over X of dimensions n and m, respectively, One constructs the vector bundle Q = /\T*X®E = AT*X®(E0®E1), (6.11.1) xx x called hereafter the body of an NQ-superbundle (see Definition 6.11.1 below) or simply an NQ-superbundle. The typical fibre of the NQ-superbundle Q is V = ARN®(B0®Bi),
(6.11.2)
where Bo and Bx are the typical fibres of the vector bundles Eo and E± respectively. This typical fibre can be provided with the structure of the superspace B"l m over the Grassmann algebra A — /\RN. This is the graded envelope of the graded vector space B = Bo © Bi, where Bo and Bi are regarded as its even and odd subspaces, respectively. The NQ-superbundle
427
Chapter 6 Supergeometry
Q inherits this gradation since transition functions of EQ and E\ are mutually independent, while the transition functions of T*X preserve the Zgradation of the Grassmann algebra AR^. Nevertheless, the NQ-superbundle (6.11.1) is not a supervector bundle over X since its transition functions are not A-module morphisms. Obviously, one can think of the NQ-superbundle Q as the tensor product of the vector bundle E and the characteristic vector bundle AT*X of the simple graded manifold (X,OX), where Ox is the sheaf of exterior forms on X. The vector bundle Vg of graded vector fields and the vector bundle V% of graded one-forms (see Section 6.3) have a local structure of an NQ-superbundle (see local isomorphisms (6.3.9) and (6.3.18), respectively). Let us denote by Q(X) and Qx the space of global sections of the NQ-superbundle Q and the sheaf of its sections, respectively. Of course, Q(X) = Qx(X). The space Q(X) has the natural structure of a locally free C°°(X)-module, while Qx is the locally free sheaf of C^-modules of rank 2N(n + m). At the same time, bearing in mind that (X, Ox) is a graded local-ringed space, one can provide Q(X) = O*{X)®E(X) with the structure of the graded locally free 0*(X)-module, while Qx = CX°<E) ARN ®B = O*X®B can be seen as the graded locally free sheaf of 0^-modules of rank (n + m). We will denote Q(X) and Qx endowed with the above mentioned structures by Q{X) and Qx, respectively. DEFINITION 6.11.1. The pair Q = (Q,Q(X)) (or the pair (Q,QX)) is called an NQ-superbundle. •
Given a trivialization domain U C X of the vector bundle E —» X, let {CA} and {CJ} be fibre bases for the vector bundles Eo and E\ over U, respectively. Then every element q of Q{X) reads q = qAcA + qlCi,
where qA, ql are local exterior forms on U. An element q is homogeneous, if its Grassmann degree is [?] = [iA\ = fe'] + 1.
[A = |
[
428
Geometric and Algebraic Topological Methods in Quantum Mechanics
Given another trivialization domain U' C X of E, the corresponding transition functions read
Q'A = PUB,
qH = pW>
(6-H-3)
where p^, p*- are local smooth functions on U D U'. We call the triple (U; qA, q1) together with the transition functions (6.11.3) a splitting domain of the NQ-superbundle Q. A connection on the the NQ-superbundle (Q,Q(X)) can be defined in accordance with Definition (1.3.2). It is easily seen that, in the case of the ring O*(X), the derivation d1 (1.3.4), is exactly the exterior differential d. DEFINITION 6.11.2. A connection on the NQ-superbundle {Q,Q(X)), called an NQ-superconnection, is denned as a morphism V : Q —> Q which obeys the Leibniz rule
V(to) = (d0)g + (-l)M/V(g),
qeQ(X),
^O'(I).
(6.11.4)
• It should be emphasized that an NQ-superconnection is defined as a connection on C*(X)-module Q(X), but not on the C°°(X)-module Q(X). Therefore, it is not an ordinary connection on the smooth vector bundle Q->X. The Leibniz rule (6.11.4) implies that an NQ-superconnection is an odd morphism. For instance, if E —> X is a trivial bundle, we have the trivial NQ-superconnection V = d. Let To and Fi be linear connections on the vector bundles Eo and Ei, respectively, and F o © Fi a linear connection on E —> X. Then the covariant differential V r (10.6.54) relative to the connection F is an NQ-superconnection Vr(q) = dxxA(dx-T)(q).
(6.11.5)
Given a splitting domain U, this superconnection reads
Vr() = (dqA - dxx A ToxABqB)cA + W - dxx A T^rf)*,
(6.11.6)
where FOA^B, ^ixzj a r e local functions on U with the familiar transformation laws under the transition morphisms p^ and /?* of the vector bundles Eo and Ei, respectively. As it follows directly from the Leibniz rule (6.11.4), the NQsuperconnections on an NQ-superbundle constitute an affme space modelled over the C*(X)0-module ~End(Q(X))i of odd degree endomorphisms
429
Chapter 6 Supergeometry
ofQ(X),i.e., V' = V + L,
(6.11.7)
where L is an odd element of End(Q(X)). It is easy to see that the O* (X)-module End(Q(X)) is a C°°(X)-module of sections of the vector bundle AT*X®E®E* ^ X. X
X
Given a splitting domain U of the NQ-superbundle Q, every element of End(Q(X)) is represented by a supermatrix function (or simply a supermatrix) (6.11.8)
whose entries are local exterior forms on U. The transformation law of this supermatrix under the transition morphisms (6.11.3) is
L' = pLr\ where p is the (n + m) x (n + m) matrix
?-(f°)
(6.U,,
whose entries are the transition functions p^ and pj. Due to the relation (6.11.7), any NQ-superconnection on a splitting domain of the NQ-superbundle Q can be written in the form V = d+
tf,
(6.11.10)
where r? is a local odd supermatrix (6.11.9), i.e., entries of $i, $4 are exterior forms of odd degree, while those of $2, i?3 are exterior forms of even degree. Obviously, the splitting (6.11.10) is not maintained under the transition morphisms (6.11.3), and we have the transformation law ti^p&p-i-dpp-1, where dp is the supermatrix whose entries are one-forms dp^ and dp*. Similarly to the expression (1.8.26), the curvature of the NQsuperconnection V is the morphism fl=V2 :Q-+Q.
(6.11.11)
430
Geometric and Algebraic Topological Methods in Quantum Mechanics
For instance, the curvature of the trivial NQ-superconnection V = d is equal to zero. The curvature of the NQ-superconnection V r (6.11.6) is f \R\»ABdxx V
A dx» A qBcA 0
0 \ ±RXlSjdxxAdxi1Aq1ciJ' (6.11.12) where R\^,AB, R\^lj are the curvatures (10.6.63) of the linear connections To and Fi, respectively. Given the local splitting (6.11.10) of an NQ-superconnection V, its curvature (6.11.11) takes the local form B(a) W
'
=
R = d(-d) +
ti2,
(6.11.13)
and has the transformation law R' =~pRp~lunder the transition morphisms (6.11.3). It follows that the curvature of an NQ-superconnection is an even endomorphism of the O*(X)-module Q(X). Remark 6.11.1. The notions of an NQ-superbundle and an NQsuperconnection are extended in a straightforward manner to the case of a complex vector bundle E —> X and the exterior algebra C ® O*(X) of complex exterior forms on X. D Let us discuss briefly the application of NQ-superconnections to the computation of the Chern character [285; 306; 362]. Given an NQsuperbundle Q (6.11.1), let us consider an NQ-superconnection V = Vr + iI,
(6.11.14)
where V r is the superconnection (6.11.6), L is an odd element of End(Q(X)), and t is a real parameter. The curvature (6.11.11) of the connection (6.11.14) reads R = t2L2 + t[Vr,L] + (Vr)2, where ( V r ) 2 is the curvature (6.11.12) of the NQ-superconnection V r (6.11.6). Then we have cht = Str(exp(i?)) = J^ Str((V r + L)2k).
(6.11.15)
fc=0
This series converges since R is a section of a bundle of finite-dimensional algebras over X. It is readily observed that ch t=0 = c h ( £ 0 ) - c h ( £ i ) ,
(6.11.16)
431
Chapter 6 Supergeometry
where ch(I?o), ch(Ei) are the Chern characters (10.10.39) of the complex vector bundles Eo and E\ (with accuracy to the customary factor i/2n). This is the Chern character of the difference element EQ Q E\ in the Ktheory of complex vector bundles on a manifold X [16; 237] (see the relation (10.11.2)). The key point is that the De Rham cohomology classes of the exterior forms cht (6.11.15) and chj=o (6.11.16) are the same. This fact issues from the following two assertions. 6.11.3. Given an NQ-superconnection V and an odd endomorphism T G End(Q(X)), one has
PROPOSITION
d(Str(T)) = Str([V,T]).
(6.11.17)
• Outline of proof. Since the relation (6.11.17) is local, one can assume that V is split as in (6.11.10) and T is a sum of supermatrices of the form 4>kTk where 4>k is are exterior forms, while Tk are constant supermatrices. Bearing in mind the relation (6.1.7), we have Str([d + 0,
d(Sti(
Let T = Rk, where R is the curvature of the NQ-superconnection V. Since Str([V,-R/c]) = 0, we obtain from the relation (6.11.17) that the exterior form Str(Rk) is closed. [362]. Given the curvature R of an NQPROPOSITION 6.11.4. superconnection V, the De Rham cohomology class of the exterior form Str(Rk) is independent of the choice of the NQ-superconnection V. • The coincidence of the De Rham cohomology classes of the forms cht (6.11.15) and cht=o (6.11.16) enables one to analyze the Chern character under the different choice of the supermatrix L and the parameter t in the expression (6.11.14).
Chapter 7
Deformation quantization
In a general setting, a deformation of an algebra A over a commutative ring K. is its Gerstenhaber extension to an algebra Ah over the ring K.[[h\] of formal power series in a real variable h such that Ah/hAh — A [167]. By deformation quantization is meant a deformation of a Poisson algebra of functions on a Poisson manifold where h is treated as a Plank constant [30]. One also considers a generalized deformation where a deformation parameter no longer commutes with elements of the original algebra [330; 332; 358]. 7.1
Gerstenhaber's deformation of algebras
Subsections: A. Formal deformation, 433; B. Deformation of associative algebras, 436; C. Relative deformation, 439; D. Commutative deformation; 441; E. Deformation of Lie algebras. 442. This Section summarizes the relevant material on deformations of algebraic structures, especially, associative algebras and Lie algebras [152; 167; 168]. A deformation parameter h throughout is real. A. Formal deformation Let A be a (not necessarily associative) algebra. Let us consider the set of formal power series ah = a + hai + h2a2 H
,
ai & A,
in a real variable h whose coefficients are elements of A. This set is customarily denoted by A[[/i]]. It is naturally an algebra, called the power series 433
434
Geometric and Algebraic Topological Methods in Quantum Mechanics
algebra, with respect to the formal sum and product of power series. For instance, let K, be a commutative ring. Then /C[[/i]] is also a commutative ring. Without a loss of generality, we further assume that any algebra A under consideration is an algebra over some commutative ring K.. Then the set A[[h]] is a /C[[/i]]-module and, in particular, the power series algebra A[[h\] is /C[[/i]]-algebra. It is also a /C-algebra, and there is a /C-algebra monomorphism A —> ^4[[/i]]. The power series algebra A[[/i]] may be provided with a different algebraic structure. The key point is that a multiplication ah{a,b) : A x A —> A in any /C[[/i]]-algebra A is automatically represented by a power series oo
ah = a+ J~^hrar, r=l
where a* are /C[[/i]]-bilinear maps Ay- A—* A. Given a /C-algebra A over a commutative ring /C, let A[[h]] be the power series module over the power series ring /C[[/i]]. Let (7.1.1)
B:AxA^A
be a /C-bilinear map. It is naturally extended to a /C[[/i]]-bilinear map B : A[[h}} x A[[h]] -> A[[h}\.
(7.1.2)
By a formal deformation of B (7.1.1) is meant a K.[[h]]-bilinear map Bh : A[[h]] x A[[h]] -> A[[h}}, oo
hrBr(a, b),
Bh(a, b) = B(a, b) + £
(7.1.3)
r=l
where Bi are /C-bilinear maps A x A —• A. In particular, if B(a,b) = ab (7.1.1) is a multiplication in A, then its formal deformation oo
a * b = ah(a, b) = ab + ^
hrar(a, b)
(7-1-4)
r=l
(7.1.3) is a multiplication in the /C[[/i]]-module A[[h}] which makes it into a /C[[/i]]-algebra. This algebra is called a formal deformation of A. This term is also abbreviated with the power series (7.1.4).
Chapter 7 Deformation Quantization
435
For instance, the power series algebra J4[[/I]] with the multiplication law a * b = ab is a formal deformation of A called the null deformation. Let us note that a /C[[/i]]-module P is isomorphic to a power series module Q[[h]} for some /C-module Q if an only if Q = P/hP. Therefore, given a /C-algebra A, a /C[[/i]]-algebra A is a formal deformation of A if and
only if A/hA = A. Two formal deformations Ah and A'h of a /C-algebra A are called equivalent if they are isomorphic £[[/i]]-algebras, i.e., there exists an isomorphism of /C[[/i]]-modules fa : A[[h}\ -> A[[h]\
(7.1.5)
such that the relation
fa(a'h(a, b)) = ah(fa(a),
(7.1.6)
holds. For the sake of brevity, let us write a'h — ah o fa. Any /C[[/i]]-linear morphism (7.1.5) is necessarily a formal power series
fa =
(7.1.7)
whose coefficients are /C-linear maps >j : A —> A. Substituting this power series into the relation (7.1.6), one easily obtains that (7.1.8)
4> = IdA, a b
0i(a6) + ai(a,&) = ai(a,b) +
(7.1.9)
A formal deformation Ah of A is said to be trivial if it is equivalent to the null deformation A[[h]] of A. Remark 7.1.1. A /C[[/i]]-algebra automorphism fa of the power series algebra A[[h}} obeys the relation
fa(ab) =
fa(a)fa(b),
a,beA.
(7.1.10)
In particular, the equality (7.1.9) takes the form 0 i {ab) = cj>i (a)b + a<j>i (6),
i.e., 0i is a derivation of the algebra A into itself. Furthermore, if K. is a Q-ring, any formal automorphism fa of A[[h}] is brought into the form
fa =
e^.e^;
where 5it i — 1 , . . . , are derivations of A.
•
436
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let S —> A be an algebra morphism. A deformation Ah of A is said to be the S-relative deformation if a*s = as,
s*a = sa
(7.1.11)
for all o G A and s G S (see Subsection C below). For instance, let A be a unital algebra and S = {1}. One can show that any formal deformation of A has a unit (see Theorem 7.1.6 below). Given an algebra A, one usually consider its formal deformations Ah possessing some properties of A. Namely, let 8. be a full subcategory of the category of (arbitrary) algebras, and let A £ Ob£. One considers the formal deformations of A in the category .8, but such a deformation need not exists. We further restrict our consideration to deformations of associative algebras and Lie algebra. B. Deformation of associative algebras Let A be a unital /C-algebra over a commutative ring fC. We aim to study its formal deformations which are associative algebras. They are called associative deformations or, simply, deformations. It is readily observed that the null deformation of an associative algebra is associative. An algebra which admits only trivial deformations is called rigid. One can show the following. • Any deformation of A is equivalent to that where the unit element coincides with the unit element of A. • Invertible elements of A remain invertible in any deformation. In order to be associative, a formal deformation Ah of A must satisfy the associativity condition oo
(ai * a2) * a3 - ai * (a2 * a3) = ^ hkDk(ai,a2, a3) = 0, (7.1.12) fc=i
(7.1.13)
Dr{ai,a2,a3)= ^2
[ar(aa(ai,a2),a3) - a r (ai,a 5 (a2,a 3 ))],
s+r=k,s,r>0
i.e., A ( a i I O 2 , a 3 ) = 0)
k = l,...,
a* £ A.
(7.1.14)
This condition is phrased in terms of the Hochschild cohomology as follows. Let Ek{ai, a 2 , a3) denote the sum of the terms with indices 1 < s,r < k in the right-hand side of the expression (7.1.13). It is easily observed that
437
Chapter 1 Deformation Quantization
Ei = 0, while each Ek, k = 2 , . . . , depends on the terms a,, i = 1 , . . . ,fc— 1. Then the condition (7.1.14) takes the form Dk(ai,a2,a3) = Ek(ai,a2,a3) - (dak)(ai,a2,a3)
= 0,
(7.1.15)
where (dak)(ai,a2,a3)
= aiak(a2,a3) - ak(aia2,a3) +
(7.1.16)
afc(oi,O2a3) - afc(ai, 02)03 is the Hochschild coboundary operator (1.5.40) of the Hochschild complex B*(A, A) (see Section 1.5C). Thus, one can think of the terms ak of the power series (7.1.4) as being two-cochains of the above mentioned Since E\ = 0, a glance at the condition Hochschild complex B*(A,A). (7.1.15) shows that a\ is a two-cocycle of the Hochschild complex. Then we can obtain an associative deformation in the framework of the following recurrence procedure. Let us assume that a,, i < k, are two-cochains such that Di = 0 for all i < k. One can show that dEk+i = 0, i.e., Ek+\ is a three-cocycle whose cohomology class depends only on that of Q!j, i = 1, ...,fc. If this cocycle is a coboundary, i.e., it belongs to the zero element of the Hochschild cohomology group H3(A, A), then a desired cochain ak+x can be found. For instance, let Ah be a deformation of A such that oti = 0,
i = 1 , . . . ,k - 1,
k > 1.
Then Ek = 0, and the condition (7.1.15) shows that ak is a two-cocycle. As was mentioned above, this is also the case of fc = 1. The Hochschild cohomology class [ak] of ak (or, if there is no danger of confusion, ak itself) is called the infinitesimal of a deformation. Let us assume that ak = d<j> is a coboundary. It is easily verified that a'h = aho (U A-hk4>)
(7.1.17)
is an equivalent deformation such that a< = 0, i = 1 , . . . , fc. If H2(A, A) = 0 and ah is a deformation, one can use the equivalences (7.1.17) in order to remove all the terms cti follows. THEOREM 7.1.1. Equivalent deformations of A possess the same infinitesimal [Qi] e H2{A, A). •
438
Geometric and Algebraic Topological Methods in Quantum Mechanics
7.1.2. If A is a unital /C-algebra with H2{A, A) = 0, then any deformation of A is trivial, i.e., A is rigid. • THEOREM
An algebra A is called absolutely rigid if H2(A, A) = 0. Let H3(A, A) = 0. For any two-cocycle a, there exists a deformation of A such that ot\ = a. Indeed, ai = a defines E2 by the formulae (7.1.13) and (7.1.15). This three-cocycle is a coboundary. Therefore, there exists a twocochain a2 such that the term D2 (7.1.15) vanishes. The two-cochains «i and a2 define the three-cocycle E3 by the formulae (7.1.13) and (7.1.15). It is also a coboundary. Consequently, there exists a two-cochain a$ such that the term £>3 (7.1.15) vanishes, and so on. Thus, elements of the Hochschild cohomology group H3 (A, A) provide the obstruction to a Hochschild twococycle be the infinitesimal of a deformation. Example 7.1.2. Let K, be a Q-ring. Let u and v be derivations of a /C-ring A. They are Hochschild one-cocycles. Their cup-product u -^ v (1.5.41) is a two-cocycle. This two-cocycle need not be the infinitesimal of a deformation of A, unless u and v mutually commute. If the derivations u and v commute, they define a deformation of A given by the formula a*b = exp{hu^-v)(a,b)
°° hr
:=ab+ ] T — ur{a)vr{b),
a,b e A. (7.1.18)
r=l
It should be emphasized that the notation exp(/m •—' v) is somewhat misleading. It may happen that u —- v = v! •—' v' for another pair u' and v' of derivations, while exp(hu ^ v) ^ exp(hu' --- v'). D Since deformations of a /C-ring A are characterized by its Hochschild cohomology H*(A, A), there is the following relation between deformations of Morita equivalent rings [167]. PROPOSITION 7.1.3. If/C-rings A and A' are Morita equivalent, then there is an isomorphism of their Hochschild cohomology groups
H*(A, A) -> H*(A', A'), which preserves the cup-product. The isomorphism (7.1.19) leads to the following.
(7.1.19) •
439
Chapter 7 Deformation Quantization
PROPOSITION 7.1.4. If/C-rings A and A' are Morita equivalent, then there is a bijection between their sets of equivalence classes of deformations. •
C. Relative deformation In order to describe the above mentioned relative deformations, we start with relative Hochschild cohomology, which is a straightforward generalization of Hochschild cohomology in Section 1.5C. Let /C be a commutative ring, A a AC-ring and Q an .4-bimodule, which a commutative /C-bimodule. Let B*(A, Q) be the Hochschild complex (1.5.36). Given a JC-subring S of A, the S-relative Hochschild complex is the subcomplex
(7.1.20)
B*{A,S,Q)->B*(A,Q)
of the Hochschild complex B*(A, Q) where: • B°(A, S, Q) consists of elements q € Q = B°(A, Q) such that sq = qs for all s G <S, • Bk>0(A, 5 , Q) consists of cochains fk € Bk(A, Q) obeying the conditions fk{sau..., k
f (ai,...
ak) = sfk(au ,aiS,ai+i,...
k
.. .,ak), k
,ak) = f (a\,..
s€S, .,a,i,sai+i,...
,ak),
k
f (ai,...,aks)
=f (ai,...,ak)s.
Cohomology H*(A,S,Q) Hochschild cohomology.
of B*(A,S,Q)
Example 7.1.3. If S = /C, clearly Bk>0(A,
are called the
JC, Q) = Bk(A, Q).
S-relative
•
The inclusion (7.1.20) is a cochain morphism which yields a homomorphism of cohomology groups
H*(A,S,Q)-^H*(A,Q).
(7.1.21)
It however need not be a monomorphism, but we always have
H°(A,S,Q) = H°(A,Q). THEOREM 7.1.5. If the algebra <S is separable (see its definition below), the morphism (7.1.21) is an isomorphism [167]. •
440
Geometric and Algebraic Topological Methods in Quantum Mechanics
Generalizing the notion of an allowable morphism in Section 1.5C, we say that, given /C-ring A and its subring B, a morphism 3> : Q —> P between A-modules is B-allowable if there is a B-module morphism A : P —> Q such that $ o A o $ = $.
(7.1.22)
For an epimorphism $ , the condition (7.1.22) reduces to $ o A = $. Likewise, for a monomorphism $ , it becomes A o $ = $ . In general, $ is Ballowable if and only if, in some (and, hence, every) factorization $ through an epimorphism and a monomorphism, both these maps are B-allowable. It should be emphasized, that a composition of allowable maps need not be allowable. A /C-ring A is called separable (in the terminology of [167]) if any /C-allowable map of modules over A is A-allowable. A finite product of separable algebras is separable. Moreover, if /C is a field, a separable /C-ring is necessarily finite-dimensional. The 5-relative Hochschild cohomology coincides with cohomology of the normalized subcomplex C*(A, S, Q) of the relative Hochschild complex , a*;) vanish if any a* € <S. In particB*(A,S, Q) whose elements fk(a,i,... ular, any normalized 5-relative one-cocycle f1 is a Q-valued derivation of A regarded as a <S-ring, i.e., fl(a) = 0 for all a E S. The relative Hochschild complex B*(A,S,Q) is functorial in Q, i.e., every ,4-bimodule morphism Q —> P yields a cochain /C-module morphism B*(A,S,Q)-*B'(A,S,P) and, consequently, a homomorphism of 5-relative Hochschild cohomology groups H*(A,S,Q)^H*(A,S,P). Each short exact sequence 0 -> Q' —> Q —-> Q" -> 0 of <S-allowable morphisms yields the long exact sequence •••Hk(A,S,Q') k+1
H
(A,S,Q')
—+Hk(A,S,Q)
-^Hk{A,S,Q")
—>
—....
Turn now to the <S-relative associative deformations of a /C-ring A [167]. THEOREM
7.1.6. Let S be a /C-subring of A.
441
Chapter 7 Deformation Quantization
(i) If the map (7.1.21) is an isomorphism for k = 2 and k = 3, then any deformation of A is equivalent to an S-relative deformation. (ii) Let Ah and A'h be equivalent 5-relative deformations of A. If the map (7.1.21) is an isomorphism for k = 1 and k = 2, then there is an S-relative equivalence
D. Commutative deformation Let A be a commutative /C-ring. Its deformations need not be commutative, but its null deformation is certainly commutative. Commutative deformations of a commutative ring are characterized by its Harrison cohomology. Let Q be an j4-bimodule. The Harrison cohomology Har*(yl, Q) is defined as a cohomology of a certain subcomplex S* (A, Q) of the Hochschild complex B*(A, Q). Unless K, is a Q-algebra, there are problems with the Harrison cohomology theory in degrees greater than 3. Therefore, we here restrict our consideration to the terms S-3(A,Q) of the Harrison complex. They are defined as follows: . S i ( A Q ) = B ' ( i , Q ) , « = O,l; • S2(A, Q) consists of the symmetric Hochschild two-cochains f2 € 2 B (A,Q), i.e., / 2 (ai,a 2 ) - / 2 ( o 2 , a i ) = 0,
a\,a2 G A;
• S3(A, Q) consists of the Hochschild three-cochains / 3 G B2(A, Q) such that / 3 (a x , o 2) o3) + / 3 (a 3 , a l5 o2) - / 3 (a 2 , oi, o3) = 0, / 3 (ai,a 2 ,a 3 ) + / 3 (a2,a3,ai) - / 3 ( a i , a 3 , a 2 ) = 0
a i , a 2 , a 3 € A;
As a consequence, we have Har°(A Q) = H°(A, Q) 3* Q,
Bar1 (A, Q) = Hl{A, Q).
A glance at the formula (1.5.39) shows that, if A is commutative and Q is an A-bimodule, any Hochschild two-coboundary is symmetric. It follows
442
Geometric and Algebraic Topological Methods in Quantum Mechanics
that the Harrison cohomology group Ha,i2(A,Q) is a submodule of the Hochschildonetf2(A,Q). With the Harrison cohomology groups Uai2(A, A) and Har3(A, A), the commutative deformations of a commutative /C-ring A are characterized as follows. • The infinitesimal of a commutative deformation of A is a Harrison two-co cycle. • Elements of the Harrison cohomology group Hai3(A, A) provide the obstruction to a Harrison two-cocycle be the infinitesimal of a deformation. • If Hai3(A,A) = 0, every Harrison two-cocycle is the infinitesimal of a deformation of A and, moreover, this deformation is equivalent to a commutative one. E. Deformation of Lie algebras Let A be a Lie algebra over a commutative ring /C. Let us study its formal deformations which are Lie algebras over the power series ring JC[[/i]]. They are called the Lie deformations. It is readily observed that the null deformation of a Lie algebra is a Lie deformation. In order to be a Lie algebra, a formal deformation Ah (7.1.4) of A must be skew-symmetric a*b=—b*a,
ar(a,b) = —ar(b,a),
a,b € A,
and must satisfy the Jacobi identity (ai * o2) * a3 + (a2 * a3) * a\ + (a3 * ai) * a2 = oo
J2hkTk(ai,a2,a3)
(7.1.23)
= 0,
fc=i
T fc (ai,a 2 ,a 3 ) =
^
[ar(ai,as(a2,a3)) +
(7.1.24)
r+s=fc, s,r>0
ar(a2, as(a3, ax)) + ar(a3, aa{a\, a2))], i.e., Tfc(a1)o2>a3) = 0> fc = l , . . . ,
a* e A.
(7.1.25)
This condition is phrased in terms of the Chevalley-Eilenberg cohomology just as the associativity condition (7.1.14) has been done in terms of the Hochschild cohomology.
443
Chapter 7 Deformation Quantization
Let J/c(ai, 0,2,0,3) denote the sum of the terms with indices 1 < s,r < k in the right-hand side of the expression (7.1.24). It is easily observed that Jj = 0, while each Jk, k = 2,..., depend on the terms on, % = 1,..., k — 1. Then the condition (7.1.25) takes the form Tfc(ai,a2,a3) = Jk(ai,a2,a3)
- (5ak)(ai,a2,a3)
= 0,
(7.1.26)
where (5ak)(ai, 0,2,0,3) = aia fc (a2,a 3 ) + a2ak(a3,ai) + a3ak(ai,a2) ak{a\, a2a3) + ak{a2, a 3 ai) + ak[a3, aia2)
+
is the Chevalley-Eilenberg coboundary operator (1.5.47) of the ChevalleyEilenberg complex C* [A] of A-valued cochains on A in item (i) of Section 1.5D. Thus, one can think of the terms ak of the power series (7.1.4) as being two-cochains of the Chevalley-Eilenberg complex C*[A]. Since Tx = 0, the condition (7.1.26) shows that a\ is a two-cocycle of the above mentioned Chevalley-Eilenberg complex. Then one can obtain a Lie deformation of A in the framework of the following recurrence procedure. Let us assume that oci, i < k, are two-cochains such that % = 0 for all i < k. One can show that 5Jk+i = 0, i.e., Jk+i is a three-cocycle whose cohomology class depends only on that of ai} i = 1,..., k. If this cocycle is a coboundary, i.e., it belongs to the zero element of the Chevalley-Eilenberg cohomology group H3(A), then a desired cochain afc+i can be found. THEOREM 7.1.8. If H2(A) = 0, then any Lie deformation of A is trivial.
•
Example 7.1 A. Let K be a Q-ring and A a semisimple finite-dimensional Lie algebra. Then Hl{A) = H2{A) = 0. • Let H3(A) = 0. For any two-cocycle a, there exists a deformation of A such that ai = a. Indeed, ai = a defines J2 by the formulae (7.1.24) and (7.1.26). This three-cocycle is a coboundary. Therefore, there exists a twocochain a2 such that the term T2 (7.1.26) vanishes. The two-cochains a^ and 0.2 define the three-cocycle JS3 by the formulae (7.1.24) and (7.1.26). It is also a coboundary. Consequently, there exists a two-cochain a3 such that the term T3 (7.1.26) vanishes, and so on. Thus, elements of the ChevalleyEilenberg cohomology group H3(A) provide the obstruction to a ChevalleyEilenberg two-cocycle be the infinitesimal of a Lie deformation.
444
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us note that, if A is an associative algebra provided with the Lie algebra structure with respect to the bracket [a,b] = ab — ba,
a,b € A,
any associative deformation a * b of A yields the Lie deformation [a, 6]* =a*b-b*a
(7.1.27)
of the Lie algebra A. 7.2
Star-product
Let (Z, {,} be a Poisson manifold. By a star-product on Z is meant an associative deformation oo
/ * / ' = / / ' + $>rar(/,/')
(7.2.1)
r=l
of the R-ring C°°(Z) of smooth real functions on Z such that <*i(/,/') -<*(/',/) = 2{/,/'},
/,/'eC°°(Z).
(7.2.2)
One usually requires that a star-product (7.2.1) - (7.2.2) is a differential are bidifferential operators of finite deformation of C°°{Z), i.e., ar(f,f) order. Moreover, up to the equivalence, one can always choose * such that / * l = l * / = /,
feC°°(Z).
(7.2.3)
Given a star-product (7.2.1) - (7.2.3), the commutator
[/, f% = {2hT\f * / ' - / ' * / ) = {/, /'} + 1
(7.2.4)
oo
- J> r (a r+1 (/,/')-a r+ i(/',/)) r=l
provides a Lie deformation of the Poisson bracket {,} on Z. This deformation is treated as deformation quantization. Remark 7.2.1. Due to formulae like (7.2.1), it is sometimes convenient to take the Laurent series £[/i~\/i]] in h, i.e., a polynomial in h~l and a formal series in h. O In order to study star-products, we start with associative deformations of the R-ring C°°(Z) of smooth real functions on a smooth manifold Z.
Chapter 7 Deformation Quantization
445
Since one usually ignores the algebraic case which is too general to be handled, there are two cohomology and deformation theories of interest, namely, the continuous and differential ones. Because C°°(Z) is a Prechet ring (see Remark 1.8.4), by its continuous deformation is meant the deformation (7.2.1) where the bilinear maps ar are continuous in each argument. Continuous cochains constitute a subcomplex B*(C°°{Z),COO{Z)) of the Hochschild complex B*(C°°{Z), C°°(Z)) whose cohomology H*{C°°{Z), C°°{Z)) is called the continuous Hochschild cohomology. Another subcomplex B*A{CCO{Z),CCO{Z)) of the Hochschild complex B*(C°°(Z), C°°(Z)) consists of cochains which are multidifferential operators (i.e., differential operators in each argument) of finite order. Since every differential operator is continuous with respect to the Prechet topology of C°°(Z), the complex B^C00 (Z), C°° (Z)) is also a subcomplex of the complex B*(C°°(Z),CO°(Z)) of continuous Hochschild cochains. Cohomology HZ(C°°{Z),C°°{Z)) of the complex B*d(Cco{Z),C°°(Z)) is called the differential Hochschild cohomology. A deformation (7.2.1) of C°°(Z) is said to be differential if all cochains ar are bidifferential operators of finite order. LEMMA 7.2.1. Any multivector field i? e %{Z) (10.6.16) of degree r defines a differential Hochschild r-cocycle 0(/i, • • • Jr) := * W i A • • • Ad/ r € Brd(C°°(Z),C°°(Z)).
(7.2.5)
• Thus, there is the inclusion %{Z) c B*d(C°°(Z), C°°(Z)).
(7.2.6)
Moreover, any differential Hochschild cocycle is cohomologous to some multivector field (7.2.5) as follows [221; 433]. THEOREM
7.2.2. The inclusion (7.2.6) induces C°°(Z)-module isomor-
phisms Hrd(C°°(Z), C°°(Z)) = %(Z).
(7.2.7)
• This result is also extended to continuous Hochschild cohomology [331; 352; 357].
446
Geometric and Algebraic Topological Methods in Quantum Mechanics
THEOREM 7.2.3. Continuous and differential Hochschild cohomology groups are isomorphic
Hrc(C°°(Z),C°°{Z)) = H^(COO(Z),COO{Z)) =%{Z).
(7.2.8)
• As a consequence, one gets the following theorem [331; 357]. THEOREM 7.2.4. Every continuous deformation of C°°(Z) is equivalent to a differential one. A continuous equivalence between two differential deformations of C°°(Z) is differential (i.e., the maps fa (7.1.7) are linear differential operators). • In view of Theorem 7.2.4, by deformations of the ring C°°(Z) of smooth functions are further meant its differential deformations. Let us turn to a star-product * (7.2.1) - (7.2.3) on a Poisson manifold (Z, {, } defined by a differential associative deformation of the M-ring C°°(Z). It is a Lie deformation of the Poisson algebra C°°(Z). Of course, an arbitrary deformation (7.2.1) need not be a star-product. The key point is that, in contradistinction with the Hochschild cohomology, the needed Chevalley-Eilenberg cohomology is small [433]. Remark 7.2.2. By virtue of Theorem 7.2.2, any Hochschild two-cocycle is cohomologous to its skew-symmetric part. Therefore, we can restrict our consideration to deformations of C°°(Z) whose infinitesimal is the Poisson bracket «i(/, /') = {/, / ' } .
/, / ' e C°°(Z),
(7.2.9)
in order to obtain a star-product. The fact that a (regular) Poisson bracket is a Hochschild two-cocycle is easily justified when it is written with respect to local Darboux coordinates (pi,ql) and, thus, is a sum of cup-products dl •-' di of mutually commutative vector fields dl and di (see Example 7.1.2). • Let us consider star-products on a symplectic manifold (Z,Q). The following assertions provide their comprehensive description. THEOREM 7.2.5. Any symplectic manifold admits a star-product [125], e.g., Fedosov's one [149]. •
Chapter 7 Deformation Quantization
447
THEOREM 7.2.6. Any star-product on a symplectic manifold is equivalent to some Fedosov star-product [34; 446], described in next Section. • 7.2.7. The equivalence classes of star-products on a symplectic manifold constitute an affine space modelled on the linear space //2(Z)[[/i]] of power series in h whose coefficients are elements of the de Rham coho• mology group H2(Z) of Z [123; 206; 340]. THEOREM
Namely, given two different star-products * and *', one associate to them a unique Cech cohomology class *(*',*) Gff2(Z;R)[[/i]],
(7.2.10)
called Deligne's relative class. It is defined as follows [123; 206]. 7.2.8. Star-products * and *' are equivalent if they are equivalent as associative deformations, i.e., LEMMA
f,g&C°°(Z),
oo
(j) = Id + ] T hr
(7.2.11)
(7.2.12)
r=l
where <j>r are differential operators.
•
One calls 4> (7.2.12) the formal differential operator. PROPOSITION 7.2.9. Let (Z,Cl) be a symplectic manifold, and let us suppose that the Cech cohomology group H2(Z;1BL) is trivial. Then any two star-products on Z are equivalent. D PROPOSITION 7.2.10. Let * be a star-product on (Z, Q), and let us suppose that HX(Z;R) = 0. Then any self-equivalence
4> = exp{ad,7>, for some 7 G C°°{Z)[[h}}, where ad, 7 (/) •=\9,f\.=9*f-f*9, is the star adjoint representation.
/ G C°°(Z)[[h]], •
Let {Ui} be a locally finite open cover of Z by Darboux coordinate charts such that Ui and all their non-empty intersections are contractible.
448
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us denote d = C°°(Ui). A star-product * in C°°(Z)[[h]] certainly yields a star-product * in Ci on a symplectic manifold {Ui, £1). Let * and *' be two star-products on Z. By virtue of Proposition 7.2.9, their restrictions to Ui are equivalent star products, i.e., there exists the formal differential operator 4>i : Ci[[h]] -» d[[h]]
(7.2.12) such that & ( / * s ) = &/*&,
f,geCi{[h}}.
On Ui fl Uj, we accordingly have a self-equivalence (pj1 o >» of * in Cjj[[/i]]. By virtue of Proposition 7.2.10, this self-equivalence is inner, i.e., there exists an element fji = — 7y € ^[[/i]] such that fij1 ° 4>i — exp{ad»7jj}. On Ui fl Uj n C/fc, the composition ad*7fcji = ad*7jfc o a d , 7 ^ o ad*7ji is the identity morphism of Cijjt[[/i]] and, consequently, is represented by an element jkji in the center R[[h}] of Ctjfc[[/i]]. The standard arguments show that the set of the elements jkji define a Cech two-cocycle with values in K[[/i]]. Its cohomology class [jkji] G H2(Z;R) is desired Deligne's relative class (7.2.10). PROPOSITION 7.2.11. If *, *' and *" are three star-products on (Z,Q,), then
D The moduli space of equivalent star-products on a symplectic manifold (Z, fi) is usually identified with
±{a} + H\z)[[h}}. Let us point out to the isomorphism H2(C°°(Z)[[h]],C°°(Z)[[h\]) = Z\Z)
+ H\Z)[[h\],
where Z2(Z) is the space of closed two-forms on Z.
(7.2.13)
Chapter 7 Deformation Quantization
449
Remark 7.2.3. One also introduces the notion of Morita equivalent star-products [77]. Star-products * and *' are said to be Morita equivalent if the associative algebras (C°°{Z)[[h]],*) and (C°°{Z)[[h]},*') are Morita equivalent. Two star products * and *' on a symplectic manifold (Z,Q) are proved to be Morita equivalent if and only if there exists a symplectomorphism ip such that Deligne's relative class t(*,ip*(*')) is integral. It provides an isomorphism [*]~t(*,r
(*')) + [*]
of the Hochschild cohomology space (7.2.13).
D
Remark 7.2.4. Given a symplectic manifold (Z, fi), one also considers a star-product * of smooth complex functions on Z which obeys the additional condition 7^9 = 9*7,
/,S€C°°(Z),
where / —> / denotes the pointwise complex conjugate. Such a star-product exists. It is utilized in order to define the GNS representation of deformation quantization [48; 434]. • Remark 7.2.5. Let (Z,Q) be a 2m-dimensional symplectic manifold. A star-product * on Z is said to be closed if / ( / * < 7 - S * / ) A O = 0. z Closed star-products exist on any symplectic manifold [347], and any starproduct is equivalent to the closed one. For instance, the Moyal product on R 2m is closed. Cyclic cohomology replaces the Hochschild one for closed star products. Obstructions to the existence and equivalence of closed star• products live in the third and second cyclic cohomology spaces [106].
Remark 7.2.6. Let * be a star-product on a 2m-dimensional symplectic manifold (Z,fi). Given the ideal Cf{Z) c C°°(Z) of compactly supported smooth real functions on Z, by a trace of * is meant an R[[/i]]-linear map
450
Geometric and Algebraic Topological Methods in Quantum Mechanics
satisfying the relation
T(f*g) = T(g*f),
f,geC™(Z)[(h}}.
Any star-product on Z has a trace which is unique up to multiplication by an element of E[/i - 1 , h]} [150; 207; 340]. It takes the form f
AO
r(/)= J fp-f^, z
PeC°°{Z)[h-\h)\.
a Let us turn to star-products on Poisson manifolds. A star-product on an arbitrary (smooth, not necessarily regular) Poisson manifold exists. Its general construction was first suggested in [252; 256]. If (Z, {,} is a regular Poisson manifold, there exists a tangential starproduct on Z [150; 302]. It is denned as follows. Let S?{Z) be the center of the Poisson algebra C°°(Z). Let us consider an <S^(Z)-relative associative deformation of the R-ring C°°{Z). A star-product on Z and a Lie deformation of C°°(Z)) which come from an <Sj-(2")-relative deformation are called tangential because Sp{Z) consists of functions constant on leaves of the characteristic foliation of (Z, {,})• A tangential star-product can be introduced either as the tangential version of Vey's work [281; 302] or the straightforward generalization of Fedosov's deformation quantization of symplectic manifolds to symplectic foliations.
7.3
Fedosov's deformation quantization
The Moyal product on Z = R 2 m was the first example of a star-product [211; 236]. Let R 2 m be provided with the coordinates (g\Pi) and the canonical symplectic form fi = dpi Adq%.
Let us consider a differential associative deformation of C°°(R 2 m ) whose infintesimal is the Poisson bracket (7.2.2). This infinitesimal is a sum of cup-products of mutually commutative vector fields d% and <9*. Then, generalizing Example 7.1.2, one can show that such a deformation exists and
451
Chapter 7 Deformation Quantization
it is given by the expression
/*/'=exp[^{/,/'}]=/exp[^(?a i -a i ?)]/'= i r D - 1 ) 7 " ^ • • • d^dir+1
£
fc=0 ' r=0
• • • dik^dh
(7.3.1) • • • d i r d i ^ • • • diJ')-
This is a star-product, called the Moyal product. Since the de Rham cohomology group H2{R2m) of R 2m is trivial, all star-products on (K 2m ,fi) are equivalent to the Moyal one (7.3.1). This star-product defines the corresponding Lie deformation (7.2.4). Fedosov's deformation quantization [149] generalizes the construction of the Moyal product (7.3.1) to an arbitrary symplectic manifold as follows. Let Z be a 2m-dimensional symplectic manifold coordinated by (zA) and provided with the symplectic form n = -£lapdza
A dz13.
A formal Weyl algebra Az over the tangent space TZZ is the unital associative algebra whose elements are the formal sums a(y,h)=
J2 2fc+r>0
hkak,a1...*lyai---yar,
(7-3.2)
where yM = iM are holonomic coordinates on the tangent space TZZ. The algebra Az is provided with the Weyl product aoa' = exp \^na" (z)-^-~^
/c=0
v
a(y, h)a'(y', h)\v=y. =
(7.3.3)
'
a
where U P is the inverse to the matrix Q,a0- Of course, this definition is independent of a coordinate system. The disjoint union of the Weyl algebras Az, z £ Z, is the Weyl algebra bundle A —> Z whose sections read a{z,y,h)=
Yl 2fc+r>0
hka
k,ai...ar(z)yai---yar,
(7.3.4)
where afclQl...Qr(^) are sections of the tensor bundles VTZ. The set A of sections (7.3.4) is also an associative algebra with respect to the fibrewise
452
Geometric and Algebraic Topological Methods in Quantum Mechanics
multiplication (7.3.3). Its unit element is a(z,y,h) = 1. The center Z of the algebra A consists of the elements oo
a=^2hkak(z),
(7.3.5)
fc=o
independent of coordinates yx. There is a filtration A D Ai D • • • in the algebra A with respect to the total degree 2k + r of the terms of the series (7.3.4). Let us consider the tensor product A® O*(Z) whose elements are Avalued exterior forms on the manifold Z: a(z,y,dz,h) =Y^hkak,al...ar,fi1...i3s{z)yai
• • -y^dz01 A- • • Adz0-, (7.3.6)
called simply .4-forms. Their multiplication is defined as the exterior product A of exterior forms and the Weyl product o (7.3.3) of polynomials in ya. Let the symbol o also stand for this multiplication. With this multiplication law, the algebra A®O*(Z) has the structure of a graded algebra over the graded commutative ring O*(Z). The corresponding bracket of two «4-forms a, a' (7.3.6) is defined as [a,a'}=aoa'-
( - l ) l a " ° V oa,
(7.3.7)
where \a\ e A® OM(Z), \a'\ € A® O^a'\Z). One says that an element a belongs to the center of the algebra A ® O*(Z) if its bracket (7.3.7) with any element of this algebra vanishes. This center is Z ® O*(Z). There are two projections of an ,4-form a(z,y,dz, h) (7.3.6) to this center. These are a0 =o(0,O, dz,h),
aOo = a(z,0,0,h).
We also have the following two operators acting on A-forms: 6 : A ® OS(Z) -> Ar-i ® OS+1(Z), a, 5a — dzx A - — dya 5*:AT® 03{Z) -» A-+1 ®
O'~\Z),
These operators possess the following properties: • the operator 5 is an antiderivation, i.e., 5(aoa') = (5a)oa'+ (-l)Mao6a',
(7.3.8)
Chapter 7 Deformation Quantization
453
and is represented by the bracket 6a =
-[^na0yadzl3,a};
• 62 = (5*)2 = 0; • for any monomial a = y a i • • • ya-dz01
A • • • A dz0",
(7.3.9)
we have (65* +6*6)a = (r + s)a. One also introduces the operator (T 1 : Ar ® OS{Z) -» Ar+l ® Os~l{Z) which acts on monomials (7.3.9) by the law x-i
_ ({r + s)~15*a ~ \8-1a = 0
r + s>0 r+s= 0'
Then any ^-form (7.3.6) has the decomposition a = (66-1 + d-15)a + a0o.
(7.3.10)
Now let us consider a symplectic connection (7.3.11) on a symplectic manifold (Z,Q). It defines a connection on the graded O*(Z)-algebra A <8> O*(Z) of .4-forms (7.3.6) by the rule
Vo = dzx A V\a. In local canonical coordinates (za), we have
Va = da+ \TG,a\ , [h J where
d = dzxA^,
G=-kXlxvy»y»dz\
454
Geometric and Algebraic Topological Methods in Quantum Mechanics
and kx^L, are the covariant coefficients (2.1.7). In accordance with Definition 1.3.4 extended to an algebra over a graded ring, this connection is a graded derivation V(a o a') = V a o a ' + ( - l ) | a | a o Va' which obeys the Leibniz rule V(0 Ao) = # A a + ( - l ) V A Va for any exterior form <j> G Ok(Z). Its curvature reads
R = \R\a^yvdzx
Adza,
RXa^ =
O^-RAA.
We have the relation (V<5 + <5V)a = 0. Given a symplectic connection K on the symplectic manifold (Z, Q), let us consider more general connections on the algebra A <8> O*(Z), namely, connections of the form Va = Va + [JT, a}=da+ [i(G + r), a],
(7.3.12)
where T is an ^.-valued one-form on Z. The .A-form T in the expression (7.3.12) is not denned uniquely. For the uniqueness of r, we require that its projection To (7.3.8) to the center Z®O*(Z) vanishes. Then the curvature
of the connection V (7.3.12) takes the form ^R=^(R
+ VT + T2).
(7.3.13)
The connection (7.3.12) is called an Abelian connection if V2a = 0 for all elements a of the algebra A ® O*(Z), i.e., the curvature form R (7.3.13) belongs to the center of this algebra.
455
Chapter 7 Deformation Quantization
PROPOSITION 7.3.1. For any symplectic connection K o n a symplectic manifold (Z,Q), there exists an Abelian connection on the algebra A <S> O*(Z) which takes the form
V = V - 5 + [ir,.] = V + [±(na/3yadzP + r),.],
(7.3.14)
where r is an .4-valued one-form such that r 0 = 0.
•
The curvature form of the connection V (7.3.14) reads
R = -lna0yadz13 2i
+ R-8r + Vr + \r2. ft
The connection V is Abelian if
Sr = R + Vr + ]-r2.
(7.3.15)
h Then R = —fl is a central form. LEMMA
7.3.2. The equation (7.3.15) has a unique solution r such that 5 - ^ = 0.
(7.3.16)
• Let V (7.3.14) be a desired Abelian connection where r obeys the relations (7.3.15) and (7.3.16). Let us consider the subalgebra Ay of the algebra A which consists of the elements a, called flat, such that Va = 0. 7.3.3. For any element b G Z (7.3.5), there exists a unique flat element a(z, y, 7) G A\r such that THEOREM
a (a) :=a o (z, 0,7) = b.
a Then the associative deformation for elements a, a' £ Z (7.3.5) on the symplectic manifold Z is defined as a*a'=
(7.3.17)
In particular, let Z = K2m and K be the zero symplectic connection. The corresponding Abelian connection (7.3.14) takes the form V = -S + d.
456
Geometric and Algebraic Topological Methods in Quantum Mechanics
It is readily observed that, in this case, the associative deformation (7.3.17) restarts the Moyal product (7.3.1). As was mentioned above, Fedosov's deformation construction has been generalized to regular Poisson manifolds. In a general setting, one extends it to symplectic ringed spaces [428]. Let K. be a commutative ring and A a commutative /C-ring. Let g be a Lie /C-algebra which is also an ,4-module, and is endowed with a Lie algebra and ^.-module homomorphism (7.3.18)
Q:Q^-0A
to the Lie algebra of derivations of the /C-ring A such that [u,fv} = f[u,v] + (g(u)f)v,
u,v£$,
feA.
(7.3.19)
One says that g is a Herz-Reinhart Lie algebra (henceforth a HRL-algebra) over the pair (/C, A). Of course, g = DA is a HRL-algebra. Example 7.3.1. Let L -* Z be a Lie algebroid with an anchor g on a smooth manifold Z (see Section 10.3). Then the Lie algebra g = L(Z) of sections of L is a HRL-algebra over the pair (/C = R, C°°(Z)). • Example 7.3.2. Let L = T be a regular involutive distribution on Z in Example 10.3.11. Section of T -> Z make up a HRL-algebra T(Z) over the pair (/C = R,A = C°°(Z)). In particular, the Lie algebra T\{Z) of vector fields on a manifold Z is such a HRL-algebra. D Example 7.3.3. Let Z be a smooth manifold endowed with a regular foliation F. One can associate to F the following two HRL-algebras. (i) Let L = TJ- be the tangent bundle to IF. It exemplifies a Lie algebroid in Example 7.3.2. The corresponding HRL-algebra TIF{Z) consists of sections of TIF —> Z, i.e., vector fields tangent to leaves of IF. (ii) Let VF —• Z be the normal bundle to F and VFN(Z) the space of its sections which are constant on leaves of F. It is both a real Lie algebra and a module over the R-ring Syr(Z) of smooth real functions which are constant on leaves of F. Moreover, there is the homomorphism g of VFN(Z) to the Lie algebra of derivations of S?(Z) given by the relation g(u)f := u\df,
u G VFN{Z),
(7.3.20)
457
Chapter 7 Deformation Quantization
where u is a vector field on Z which projects onto u in accordance with the exact sequence (10.6.28). Then we have a HRL-algebra g = VT^{Z) over the pair (K. = 9., A = Sp{Z)). In contrast with the previous examples of HRL-algebras, it does not come from a Lie algebroid. • Given a HRL-algebra g over (/C, .4), one can associate to g the following differential calculus over A which coincides with the Chevalley-Eilenberg one if g = liA. Due to the homomorphism g (7.3.18), the ring A is a gmodule. Let us consider a subcomplex O*[g] of the Chevalley-Eilenberg complex C*[JJ;.4.] (1.5.46) which consists of ^-multilinear cochains. Its coboundary operator reads k
dcj>(u0,...,
ujb) = ^ ( - l ^ u O M u o , • • • > « * > • • • . « * ) ) +
(7-3.21)
t=0
^(-l)t+^(/)([
U i
,U
: 7
- ] , u 0 , ...,Ui,...,Uj,..
.,Uk),
i<j
(cf. the coboundary operator (1.6.6)). One can think of elements of O*[g] as being differential forms over a HRL-algebra g whose exterior product (1.6.9) obeys the relations (1.6.10) - (1.6.11). Let w € H2[fl] be a two-form. It yields an A-module morphism tob :g-^n1[g],
J(v)(u) =u(u,v),
u,v e g.
(7.3.22)
A two-form u> is called non-degenerate if J" is an isomorphism. Let w" denote the inverse morphism to LJ° . If a non-degenerate two form u> is closed (i.e., dui = 0), it is called a symplectic form. Clearly, a symplectic form need not exists. Example 7.3.4. Let L —» Z be a Lie algebroid and g = L(z) the HRLalgebra in Example 7.3.1. If the differential calculus Q*[L(Z)] over C°°(Z) contains a symplectic form, one says that L is a symplectic Lie algebroid.
•
Example 7.3.5. Let (Z,w) be a symplectic manifold and g = T\{Z) the HRL-algebra in Example 7.3.2. Then the differential calculus ft*[7I(Z)] over C°°(Z) is exactly the algebra O*(Z) of exterior forms on Z, and w is a symplectic form over the HRL-algebra g = T\(Z). • Example 7.3.6. Let (Z, w) be a regular Poisson manifold, J- its characteristic foliation and wy? the corresponding symplectic leafwise form (2.2.13).
458
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let TT(Z) be the HRL-algebra in Example 7.3.3, item (i). Then the differential calculus Q*[TF(Z)] over C°°(Z) is the leafwise differential calculus y(Z) (2.2.1), and wjr is a symplectic form over the HRL-algebra TT(Z). This is also the case of Example 7.3.5. • Example 7.3.7. Let w be a presymplectic form of constant rank on a manifold Z and T its characteristic foliation. Let V!FN{Z) be the HRLalgebra in Example 7.3.3, item (ii). The differential calculus £l*[VFN(Z)} over Sf(Z) is the subcomplex of the de Rham complex O*(Z) which consists of exterior forms <j> such that: (i) u\4> = 0 whenever u is a vector field tangent to the foliation T (see the exact sequence (10.6.29)), and (ii)
df=J{df)£2 (cf. (2.1.2)), and can define the Poisson bracket {/,/'} : = ^ / ) ( / ' ) >
(7.3.24)
fJ'eA,
on the ring A (cf. (2.1.14)). This bracket obeys the Jacobi identity, but the relation (2.1.15) need not hold. We have
P/Al-'WeKere. For instance, if (Z, {, }) is a Poisson manifold and Q is the HRL-algebra TJ-(Z) provided with the symplectic tojr in Example 7.3.6, then the Poisson bracket (7.3.24) coincides with the original one. Given the differential calculus n* [9] over A, one can introduce a connection on an ,4-module g as an ^-module morphism 0 B u h-» Vu € Diff ! (fl) g)
(7.3.25)
such that the first order differential operators V u obey the Leibniz rule V u (/i;) = u(/)v + /V u «,
f£A,
veg,
(cf. Definition 1.3.3). In this case, one can also define the torsion T(u, v) — Vuv - Vvu - [u, v],
u, v G 0,
(7.3.26)
459
Chapter 7 Deformation Quantization
of a connection on g and the torsionless connection V£v = -(V u i> + V,,u+ [«,«]). If a HRL-algebra g = L(Z) comes from a Lie algebroid, a connection on g is called a connection on a Lie algebroid. Any connection V on g yields a connection on the ,4-module £1*[Q\ of differential forms over the HRL-algebra g by the law (V u >)(ui,...,u r ) :=u(4>(vi,...,vr))
-
r
^^(ui,...,VuVi,...,ur),
u,ViCg,
(jxEQ.r[g}.
i=l
In particular, given a symplectic form u> G fi2[fl], a torsionless connection V on g is called a symplectic connection if Vuw = 0 for all u £ g. Let g be a HRL-algebra over the pair (fC,A) such that the morphism Q (7.3.18) is a monomorphism. Let w be a symplectic form over g and {,} (7.3.24) the corresponding Poisson bracket on A- For instance, this is the case of regular Poisson manifolds and presymplectic manifolds. With a symplectic connection V, one can straightforwardly generalize Fedosov's deformation procedure on a symplectic manifold in order to obtain a starproduct on the ring A. We refer the reader to [428] for the detailed exposition in terms of sheaves of iJi?L-algebras called HRL-ringed spaces (or sheaves of twisted Lie algebras in the terminology of [235]). 7.4
Kontsevich's deformation quantization
Subsections: A. Differential graded Lie algebras, 460; B. L^-algebras, 461; C. Formality theorem, 464; D. Kontsevich's formula, 468; E, Globalization of Kontsevich's deformation, 470; F. Deformation of algebraic varieties, 471. Kontsevich's deformation [252; 256] generalizes the Moyal star-product on R2m to a generic Poisson structure on M2m which fails to be regular and does not admit the Darboux coordinates. Recently, Kontsevich's deformation quantization has been extended to an arbitrary Poisson manifold [92; 130; 255]. The key point of Kontsevich's deformation is the formality theorem (see Theorem 7.4.24 below). This theorem and Theorem 7.4.1 establish
460
Geometric and Algebraic Topological Methods in Quantum Mechanics
the relations between algebras of multivector fields and multidifferential operators on a smooth manifold. A. Differential graded Lie algebras We start with the relevant algebraic constructions (see, e.g., [93; 127]). Unless otherwise stated, all algebras are over a field K of characteristic zero. By a graded structure is meant a Z-graded structure, and the symbol |. stands for the Z-graded parity. A differential graded Lie algebra (henceforth a DGLA) is a differential graded algebra 3 = © fl\ iez whose multiplication operation is a graded Lie bracket [,]:fl®fl->fl,
[a,&]€fll°l+'6l,
satisfying the relations [o,6] = -(-l)l-ll 6 l[6,o], a,b,ceg, (-1)I°HCI [a, [b, c\] + (-l) |6| l a l [b, [c, a}} + (-l)lcHfcl [c, [a, b}} = 0, (7.4.1) and the differential d of degree one obeys the graded Leibniz rule d[a,b] = [da,&] + (-l) |a| [a,d6].
(7.4.2)
The degree zero part g° and the even part of a DGLA g are Lie algebras. Any Lie algebra is a DGLA concentrated in degree zero with d = 0. A morphism of DGLAs is a graded linear map of degree zero which commutes with differentials and brackets. In particular, it is a cochain morphism. The cohomology H*(g) of a DGLA g is a DGLA with respect to the zero differential d = 0 and the bracket [ a , b ] H ••= [a,b],
a,b£g,
where a denotes the cohomology class of a € g. It is evident that the cohomology of the DGLA H*(g) coincides with H*(g), i.e., H*(H*(2)) = H*(2).
(7.4.3)
461
Chapter 7 Deformation Quantization
Every morphism $ : g —> g' of DGLAs yields a morphism of the cohomology DGLAs $:fr( f l )->JT(fl').
(7.4.4)
A DGLA morphism $ is called the quasi-isomorphism if the induced morphism $ (7.4.4) is an isomorphism. Let us emphasize that the existence of a quasi-isomorphism fl —>fl'does not imply the existence of the 'quasi-inverse' Q1 —> 0 which induces the inverse morphism $ . If the quasi-inverse exists, the DGLAs g and Q' are called quasi-isomorphic. A DGLA Q is called formal if it is quasi-isomorphic to the DGLA H*(g). A graded coalgebra is a graded vector space f provided with the following graded co-operations: • a graded comultiplication A:f-»f®f such that
A(f)c
© f®f*;
• a counit e : f -> K such that e(fi>0) = 0. These operations obey the relations (10.2.1). A graded coalgebra is called graded cocommutative if P o A = A, where P{a®b)
= (-l)^%®a,
a,bef,
is the graded transposition operator. A coderivation of degree r on a graded coalgebra f is a graded linear map d : f —> f+r which satisfies the co-Leibniz rule: (A o d){v) = {{d ® Id + (-l) r Hld ® d) o A)(v),
v G f.
By a codifferential on a graded coalgebra is meant a nilpotent coderivation of degree one. B. Loo-algebras The fact that the quasi-inverse of DGLAs need not exists has motivated one to call into play a wider class of algebras, which are Loo-algebras. Let V be a graded vector space and ®V its tensor algebra (1.1.7). It is a graded vector space such that v®v'\ = \v\ + \v'\,
v,v'GV,
462
Geometric and Algebraic Topological Methods in Quantum Mechanics
and |A| = 0, A £ K. We call ®+V = ®V \ K the reduced tensor algebra. The tensor algebra ®V is brought into a graded coalgebra with respect to the counit e : ®V —> K and the comultiplication A(t>i ® • • •
The graded symmetric algebra \/V and the graded exterior algebra f\V are defined as the the quotients of &V by the two-sided ideals generated by homogeneous elements of the form v <S>v' — P(v
v®v' + P(v®v'), respectively. Accordingly, we have the reduced algebras \J+V and A+V^. The graded symmetric algebra VV is brought into a graded coalgebra provided with the comultiplication given on elements of V by A(u) = l ® u + t ; ® l , and extended as an algebra homomorphism with respect to the tensor product. Given a graded vector space V, one can obtain a new graded vector space V[k] by shifting the degree byfe,i.e.,
V[kY = Vi+k,
|i7|[fc] = M-fc,
v&V.
Then we have the decalage isomorphism between the graded symmetric and exterior algebras. It is given on the fc-symmetric power of V shifted by one by the expression (7.4.5)
Ck:\fV[l]^AV[k], Cfc : vi V • • • V vk ^ ( - l ) £ t i ( k - * ) ( K I - i V l
A...AVk_
An Loo-algebra is a graded vector space g endowed with a codifferential D on the reduced symmetric space V+fl[l]. Let us abbreviate the symbol Dj with the projection of D to the com%
3
ponent V_|_g[l] of the target vector space and V + g[l] of the domain space.
463
Chapter 7 Deformation Quantization
A coderivation D is proved to be uniquely determined by its components D]., and it is a codifferential if and only if n
5 3 ^ 0 ^ = 0,
n>l.
(7.4.6)
i=i
In particular, D\ o D\ = 0, i.e., a codifferential D defines a complex on Q. One can show that any DGLA (g, d) can be brought into an Loo-algebra where D\a := (-l)[alda, Dl{a V b) := {-l)^^'^[a,b], Dln = 0, n > 3. Remark 7.4.1. If one had chosen £>! non-vanishing, the bracket [, ] would fulfill the graded Jacobi identity up to a term of the form dh(a, b, c) ± h(da, b, c) ± h(a, db, c) ± h(a, b, dc),
a,b,c G g,
where h: Afl->fl[-l]. In this case, g is said to become a homotopy Lie algebra.
•
An Loo-algebra morphism (a pre-Loo-algebra morphism in another terminology)
$:( fl) D)->( fl M>') is a morphism of graded coalgebras $ : V+fl[l] - V +0 '[l],
(7.4.7)
which intertwines the codifferentials, i.e., $ o £) = L>' o $. An L^-algebra morphism is uniquely determined by its components $] : V+fl[l] ->
fl'[l]
(7.4.8)
which obey the relations (7.4.9) i=l
t=l
464
Geometric and Algebraic Topological Methods in Quantum Mechanics
In particular, a morphism $ of DGLAs induces their morphism $ as Loo-algebras such that $J = $. Let us note that Loo-algebras are customarily introduced as strongly homotopy algebras [297]. Due to the decalage isomorphism (7.4.5), the codifferential D defines a sequence {Ik} of morphisms ** : Afl->fl[2-fc],
k>l.
The relation (7.4.6) puts an infinite family of conditions on the Ik's. These conditions imply that h :0 -
fl[l]
(7.4.10)
is a differential of degree one obeying the graded Leibniz rule. It follows that (g,h) is a complex. The morphism l^ defines a graded bracket [•,.]:Afl->fl
(7.4.11)
on g compatible with l\ (i.e., the relation (7.4.2) holds). This bracket obeys the graded Jacobi identity up to homotopy given by I3. Hence, any Looalgebra (g, D) such that Ik = 0 for k > 3 is a DGLA. Every Loo-algebra morphism $ provides a cochain morphism $1 between the complexes (g,h) and (g',l[). One says that an Loo-morphism $ (7.4.7) of Loo-algebras is a quasi-isomorphism if $1 is a quasi-isomorphism of complexes (g,h) and (g'Ji), i.e., it yields an isomorphism of their cohomology. In contrast with morphisms of DGLAs, any quasi-isomorphism of Loo-algebras possesses the quasi-inverse. Thus, quasi-isomorphisms of Loo-algebras define an equivalence relation, i.e., two Loo-algebras are quasiisomorphic if and only if there is an Loo-quasi-isomorphism between them. The notion of formality of Loo-algebras is formulated similarly to that of DGLAs. C. Formality theorem First, let us show that multivector fields on a smooth manifold Z make up a DGLA. The graded commutative algebra T*(Z) of multivector fields is customarily provided with the Schouten-Nijenhuis bracket [., .]SN (10.6.17) which obey the relations (10.6.18) - (10.6.20). However, there is another sign convention used in the definition of the Schouten-Nijenhuis bracket [298]. This bracket, denoted by [., .]SN', is [i?,v]sN'=-(-l) |tf| [i?,u]sN-
(7-4.12)
465
Chapter 7 Deformation Quantization
The relation (10.6.18) for this bracket reads [MSN' = - ( - ^ " " " ^ ' - " M l s N ' -
(7-4.13)
The relation (10.6.19) keeps its form, while the relation (10.6.20) is replaced with the one (-l)(M-i)(M-i) [ l / ) [ t f ) t ; ] S N / ] S N , + ( _l)(M-i)(l*l-D [ i ; i M ] S N
+
,
(_ 1 )(W-i)(M-i) [ t f ) [ t ; j I / ] S N / ] s N ,
] S N
, =
o.
(7.4.14)
The equalities (7.4.13) and (7.4.14) show that, with the modified SchoutenNijenhuis bracket (7.4.12), the graded vector space V=T,(Z)[1]
(7.4.15)
of multivector fields on a manifold Z is precisely a graded Lie algebra. This graded Lie algebra is brought into a DGLA by setting the differential d to be identically zero. Clearly, this DGLA is formal. In particular, let (Z, w) be a Poisson manifold. Then the Poisson bivector w obeys the Maurer-Cartan equation dw + ^[w,w]sw = 0
(7.4.16)
on the DGLA V* (7.4.15). Let us describe the DGLA of multidifferential operators on the ring C°°(Z) of smooth real functions on a manifold Z. In a general setting, let A be a K-ring and B*(A, A) its Hochschild complex. Let us consider the complex (7.4.17)
C* = B*{A,A)[l}. It inherits the Hochschild coboundary operator (1.5.37): (6
J +
V
f e
( a o , . . . , CLjdj+u
•••, ak+i)
(7.4.18) +
i
(-l) f e + V(a 0 ,...,afc)afc+i, the composition product (1.5.43): <^ m o (? !,"(ao,...,a m+n )=
(7.4.19)
ro
^(-l)
m
0
m
( a o , . . . , Oi-i,it>n(ai,...,
an+i),
an+i+i,...,
am+n),
466
Geometric and Algebraic Topological Methods in Quantum Mechanics
and the Gerstenhaber bracket (1.5.44):
[0,0'] G = W - ( - l ) l * N * V o 0 ,
(t>,
(7.4.20)
One can show that this bracket obeys the graded Jacobi identity (7.4.1). Thus, (C*, [., .]G) is a graded Lie algebra. Furthermore, let a one-cocycle / ^ a o , ai) = m(a0, a{) = a o ai be the multiplication in A. We consider the operator dm
cj>eC*,
(7.4.21)
of degree one on C*. A direct computation shows that
dm
>eC*,
and the relation (7.4.2) holds. Then the complex C* (7.4.17) is a DGLA with respect to the bracket (7.4.20) and the differential (7.4.21). Let A = C°°(Z), and let us consider a subcomplex T>* of the complex B*(C°°(Z), C°°(Z))[1] whose cochains are multidifferential operators on C°°(Z). This subcomplex is closed with respect both to the Gerstenhaber bracket (7.4.20) and the action of dm. Thus, it is a desired DGLA of multidifferential operators. Given a bidifferential operator a £ P 1 , one can think of
/ * / ' = (m + a)(/,/') = / / ' + «(/./'),
f,f'eC°°(Z),
(7.4.22)
as being a deformation of the original product m in C°°(Z). One can show that the associativity constraint on this deformation is given by the equality [m + a,m + a\G = 0 , which takes the form of the Maurer-Cartan equation dma + ^{a,a}G=0.
(7.4.23)
Now, the goal is to construct a morphism of the DGLA V* (7.4.15) of multivector fields to the DGLA V* of multidifferential operators which intertwines their differential graded Lie algebra structures and solutions of the Maurer-Cartan equations (7.4.16) and (7.4.23). One has proved the following [221].
Chapter 7 Deformation Quantization
467
7.4.1. For any smooth manifold Z, there is an isomorphism between the cohomology H*(D*) of the algebra P and the algebra V*. Since V* coincides with its cohomology H*(V*), we have an isomorphism THEOREM
H*(V*) -» V* S #*(V*).
(7.4.24)
• The next step is the above mentioned formality theorem. 7.4.2. The DGLA P* of multidifferential operators on a smooth manifold Z is formal. •
THEOREM
It follows that there exists a quasi-isomorphism of the DGLA P* to H*{V*) = V* and, consequently, there is the inverse Loo-quasi-isomorphism Wof V* toP*. Remark 7.4.2. There is a natural quasi-isomorphism U{ of complex V* to the complex P*. This morphism associates to a multivector field i?o A • • -Ai?n the multidifferential operator whose action on functions fo, • • • ,fn & C°°(Z) is given by the expression
where s runs through all permutations of the numbers (0,..., n) and sgn(s) is the sign of a permutation s. For instance, li[ assigns to a Poisson bivector w the Poisson bracket | { , } . However, the morphism li[0' fails to preserve the Lie structure. • An Loo-quasi-isomorphism U of V* to P* intertwines their differential graded Lie algebra structures and solutions of the Maurer-Cartan equations (7.4.16) and (7.4.23). Let us note that this Loo-quasi-isomorphism fails to be canonical. It is represented by a power series whose first term of differential operators of minimal order coincides with the morphism U^ in Remark 7.4.2. This morphism associates to each Poisson bivector field ID 6 V1 on Z a certain bidifferential operator aw G P 1 which obeys the Maurer-Cartan equation (7.4.23) and, thus, defines a differential associative deformation m + aw of the ring C°°(Z).
468
Geometric and Algebraic Topological Methods in Quantum Mechanics
D. Kontsevich's formula An Loo-quasi-isomorphism U\ of V1 to V1 on Z = Rr in the explicit form has been obtained in [252; 256]. Given a Poisson bivector field w eV1, the corresponding deformation of C°°(Z) reads oo
« » = E E
«=i reG n , 2
^Sr,
(7.4.25)
where Bp, T £ Gni2, are bidifferential operators of order n in each argument and Wr are the weight coefficients. These operators and weight coefficients are indexed by the elements F of a suitable subset C?2,n of the set of graphs on n + 2 vertices. They are called admissible graphs. An oriented graph F belongs to the set G n , m , n, m > 0, In — m + 2 > 0, if the following holds. • The set of its vertices has n + m elements indexed, by {1,... ,n; 1,... , m}, where the vertices {1,... ,n} are said to be of first type, while {1,... ,m} are said to be of the second type. • The £Y is the set of 2n — m + 2 oriented edges of F. There is no edge starting at a vertex of the second type. The cardinality of the set of edges starting at a vertex of the first type k is denoted by (j(fc); hence
J2 lt(fc) = 2n + m - 2. l
The edges of F starting at vertex k are denoted by {e\,..., ejj. }. We also abbreviate (s(v),e(v)) with an edge possessing the starting vertex s(u) G {1,... ,n} and the ending one e(v) £ {1,... ,n : 1,... ,m}. • The F has no loop and no parallel multiple edges. In the case of a star-product, only the sets Gn,2, n>0, are of interest. Moreover, there are exactly 2 edges starting at each vertex of the first type, i.e., tt(fc) = 2, k = 1,... ,n. Hence, £p admits 2n edges for V € Gn,2- For the sake of convenience, we write the pair of edges {e£, e^} as {ik,jk}- The set Gn<2 has (n(n + 1))" elements, while Go,2 consists of only one element, which is the graph having as set of vertices {1,2} and no edges. Given a Poisson manifold (W, w) coordinated by (z 1 ,..., zr), one associates to each graph F € Gn,2, n > 1, the following bidifferential operator ^r(/iflO) f,9 £ C°°(Rr). The components w%k'ik of a Poisson bivector w are assigned to each vertex k of first type where the edges {ik,jk} start from. Functions / and g are associated to the vertices 1 and 2, respec-
469
Chapter 7 Deformation Quantization
tively. Every edge ik acts as dik on its ending vertex. Then Br is given by the following sum over all maps I : ET —> {1, • • •, r}:
\fc=ifc'=i
i
fft^vf) (ftdn^) •
/
The weight W(T) of a graph T in the formula (7.4.25) is determined as follows. Let H = {x G C : Im (a;) > 0} be the upper half-plane with its Lobachevsky hyperbolic metric. The configuration space of n distinct points in 7i is denoted by Hn. It is an open submanifold of C n with its natural orientation. Let
W9T) = ,/n .„
/ A nn
feAfo),
where {e\, e\}, 1 < k < n, is the set of edges of T. It should be emphasized that the weights depend neither on a Poisson structure nor the dimension r. With the deformation aw (7.4.25), we obtain the star-product f*g
= fg + aw(f,9)
(7.4.26)
470
Geometric and Algebraic Topological Methods in Quantum Mechanics
on the Poisson manifold (W,w). It should be emphasized that, if a is a solution of the Maurer-Cartan equation (7.4.23), ha is not so, unless h = 0,1. Therefore, athw / haw, but takes the form oo
ahw= $ > n n=i
£
W B
^ ^
(7.4.27)
J2 WTBr{f,g).
(7.4.28)
reG n , 2
The corresponding star-product reads oo
f*9 = f9+Y,hn n=i
reG n , 2
E. Globalization of Kontsevich's deformation As was mentioned above, the globalization of Kontsevich's deformation quantization to an arbitrary Poisson manifold has been performed both as a corollary of the formality theorem [255] and in an explicit form [92] similar Fedosov's one [150] (see also [130]). In the last case, local data given by the Poisson structure are described in terms of infinite order jets of functions, multivector fields and multidifferential operators. Let EQ be the fibre bundle of infinite jets of smooth functions on Z. It is endowed with the canonical flat connection Vo (see Remark 10.7.3) whose integral sections are the jets of globally defined functions. Moreover, the jets of a Poisson bracket w yield a Poisson bracket {, }oo on each fibre of Eo such that, whenever / , g are sections of Eo —> Z, the relations Vo(/s) = Vo(/)<7 + /Vo(<7), V 0 ({/, 9}oo) = {V 0 (/), g}oo + {/, V0()}oo hold. Thus, EQ is a fibre bundle of Poisson algebras with a flat connection. Moreover, there is an isomorphism of the Poisson algebra C°°{Z) onto the algebra H°(EQ,VO) of integral sections of Vo- Then the goal is to deform (Eo, Vo). For this purpose, one consider the fibre bundle E —> Z of associative R[[/i]]-algebras using Kontsevich's local star-product * (7.4.28) in each fibre. In order to avoid the dependence of coordinates, the fibre bundle E is defined by means of the quotient of jets of coordinate systems on Z by the action of the group GL(dim Z, R) of linear diffeomorphisms. Afterwards, the connection VQ is deformed to a flat connection Vo = VQ +
471
Chapter 7 Deformation Quantization
o(h) on E obeying the Leibniz rule V o (/ * 9) = V o (/) * 9 + f * Vo(s),
/, 9 G E{Z).
The star-product * induces a product on the space H°(E,V) of integral sections of the connection V. Finally, one obtains the R[[/i]]-module isomorphism Q:H°(Eo,Vo)[[h}}^H°(E,V){[h}},
which induces the star-product
in H°(Eo,Vo) and, consequently, C°°(Z). F. Deformation of algebraic varieties Kontsevich's deformation of a smooth Poisson manifold is generalized to deformation of algebraic varieties [255]. First, let us recall that, given a commutative ring .4, an algebroid over K is a small category £ such that: • £ is non-empty and all objects of C are isomorphic; • the homsets of C are endowed with the structure of ^4-modules: • composition of morphisms are ^.-bilinear. For instance, an algebroid over /C with one object is an associative /C-ring. Let A be a field of characteristic zero and V a smooth affine algebraic variety. Given its function field R\>, let us consider the module AdRy of multiderivations of Ry provided with the Schouten-Nijenhuis bracket [, ]SN2
Then a Poisson biderivation w £ A0i?v such that [W,U;]SN = 0 defines an algebraic Poisson structure on V. Let us consider a formal deformation of the Poisson algebra Ry, i.e., a non-commutative associative /C[[/i]]-linear product on i?y[[h]]. For this purpose, one is to choose a connection on f?y. Such a connection exists. Then one can associate to (V,w) an algebroid £-v,w o v e r ^[[^l] whose objects are connections V on i?y and the homsets Hom(V, V ) are canonically identified as ^t[[/i]]-modules with flv[[/i]]. Composition of morphisms is given by power series in h whose terms are bidifferential operators.' Furthermore, one considers the canonical Abelian category of -4[[/i]]linear functors from £yitu to the category of .4[[/i]]-modules, and treats this Abelian category as a deformation of the the category of quasi-coherent sheaves on V.
472
Geometric and Algebraic Topological Methods in Quantum Mechanics
7.5
Deformation quantization and operads
A different proof of the formality theorem on R r has been suggested in [41l]. One of its main idea is to consider the Lie algebras of multivector fields and multidifferential operators not as just DGLAs, but as homotopy Gerstenhaber algebras [183; 219; 411]. Another important perculiarity of Tamarkin's proof of the formality theorem is that it is phrased in terms of operads [253; 411]. Roughly speaking, one states that, for any associative algebra over a field of characteristic zero, its Hochschild complex and its Hochschild cohomology are algebras over the same operad. As was mentioned above, this observation has been the starting point of 'operad renaissance'. In a general setting, operads and algebras over operads are defined in terms of monoidal categories (see next Section), but we start with the following particular variant of (symmetric) operads. 7.5.1. An operadP of vector spaces consists of the following: • a collection of vector spaces P(n) indexed by integers n > 0, • an action of the symmetry group Sn (of permutations of n elements) on P(n) for every n, • an identity element 1 £ P(l), • compositions DEFINITION
m{ni...nk) • P(k) ® (P(ni) ® • • • ® P{nk)) -> P(m + • • • nk)
(7.5.1)
for every k > 0 and n 1 ; . . . ,nk > 0 obeying certain conditions [307].
D
There are the following basic examples of operads. Example 7.5.1. Given a vector space V, let us put P(n) = Horn (®nV,
V)
and m{ni...nk)(e
(7.5.2) (8) • • • (8)
vni+...nk)),
where e € P(k) and e, G P(rii), i — 1,..., k. The symmetry group Sn acts on ei ® • • • ® en € P(n) by permutations. D
Chapter 7 Deformation Quantization
473
Example 7.5.2. The nth component of the operad Associ is denned as the collection of functorial n-linear operations A®" —> A defined on all associative rings A, and it is spanned by the operations a i
t-> aa(i)
• • • aff(n),
where a £ Sn is a permutation.
•
Example 7.5.3. Let C be the symmetric monoidal category of Z-graded cochain complexes (cf. Example 7.6.5 below). Operads in this category are called differential graded operads or, simply, dg-operads. Passing from complexes to their cohomology, one obtains the cohomology operad H* (P) of complexes with zero differential. • An algebra over an operad P (or, simply a P-algebra) consists of a vector space A and a collection of polylinear maps /„ : P{n) ® A®n -» A,
n G N,
obeying the following axioms: • the map /„ is Sn-equi variant; • fi (Id
Y^riifi,
N>0,
of continuous maps fi-.
modulo the following relations
[0,l}k^X
riiSZ,
474
Geometric and Algebraic Topological Methods in Quantum Mechanics
• / o a = sgn)o-)/ for any a G S^ acting on the standard cube [0, l]k by permutations of coordinates; • / ' o pr^fe,!) = 0, where pr^fc-DrlO.l]*-^!]*"1 is the projection onto first (k — 1) coordinates and f':[0,l]k-1-+X
is a continuous map. The boundary operator on cubical chains is introduced in the usual way. There is the map ® (Chains^)) -> Chains( TT X4)
i€l
tki
which is a natural homomorphisms of complexes for any finite collection (Xj)j e / of topological spaces. Let now P be a topological operad. Then the collection of complexes Chains(P(n)) possesses a natural operad structure in the category of complexes of Abelian groups. It is the operad of chains Chains(P). The compositions in this operad are defined using the external tensor product of cubical chains. Turn now to the definition of the little discs operad. Let d > 1 and Gd the (d + l)-dimensional Lie group acting on Rd by affine transformations u h-» \u + v, where A > 0 is a real number and v £ Rd is a vector. This group acts simply transitively on the space of closed discs with a center v and a radius A in M.d. The little discs operad C4 is a topological operad defined as follows: • Cd(0) = 0; • Cd(l) = {IdCd} is a singleton set; • Cd(n > 2) is the space of configurations of n disjoint discs inside the standard unit disc centered at the zero point of Rd; • the composition Cd(fc) x (Cd(rai) x • • • x Cd(rifc)) -> C d (ni + • • • + nk) is obtained applying elements of Gd associated with discs in the configuration in Cd(k) to configurations in all Cd(rij), i = 1,..., k, and putting the resulting configurations together;
475
Chapter 7 Deformation Quantization
• the symmetry group Sn acts on Cd(n) by renumerations of indices of discs in Cd(n). With these preliminaries, we can formulate the main steps of Tamarkin's proof of the formality theorem [253; 411]. (i) There exists a natural action of the operad of chains Chains(C2) of the little discs operad C2 on the Hochschild complex B*(A,A) for an arbitrary associative algebra (Deligne's conjecture). (ii) The operad of chains of the little discs operad is formal, i.e., it is isomorphic to its cohomology. This is true only for rings over a field of characteristic zero. (iii) It follows from items (i) and (ii) that, for any ring A over a field of characteristic zero, its Hochschild complex B'*(A,A) and its Hochschild cohomology H*(A, A) are algebras over the same operad up to a homotopy. Herewith, if one chooses an explicit homotopy between these complexes, one obtains two different structure of a homotopy Gerstenhaber algebra on H*(A, A) which are always equivalent for the case of A = M.[[h]]. These two structures give equivalent structures of homotopy Lie algebras. It follows that B*(A,A) is equivalent to H*(A,A) as a homotopy Lie algebra which is the statement of the formality Theorem 7.4.2. Let us note that Deligne's conjecture has several proofs [254; 309]. Tamarkin's proof is based on the Etingof-Kazhdan theorem of quantization of Lie bialgebras [146],
7.6
Appendix. Monoidal categories and operads
In a general setting, operads and algebras over operads are defined in terms of monoidal categories [276; 297; 307; 308].
DEFINITION 7.6.1. A class of objects of a monoidal category 27T is provided with a product K
(K®K')®K" ->K®{K'®K"),
lm®K-*K,
K®\m-^
K (7.6.1)
476
Geometric and Algebraic Topological Methods in Quantum Mechanics
for all K, K', K" e ObSD? which obey the natural coherence relations (K®lm)®K' \
—* K®(lm®K') / K®K' (lm®K)®K' —* lm ® (K ® K') \
JRT (8) i=f'
/
(if ® K') ® 1OT —> K ® (K1 ® lm) \ / K®K' (lm ® K) ® 1OT —> lm ® (K ® lm) K®lm
I
—>
I
-FST <—lart®-^ D
If the isomorphisms (7.6.1) are actually the identities, 9Jt is said to be a strict monoidal category. A monoidal category or a strict monoidal category 971 is said to be symmetric if there are isomorphisms K ® K' —» K' ® K for all K, K' € SDT. Let us note that the word "strict" qualifies "monoidal", but not "symmetric" in the term " symmetric monoidal category". The coherence theorem [234] states that any monoidal category (resp. symmetric monoidal category) is equivalent in a suitable sense to a strict monoidal category (resp. strict symmetric monoidal category). An object M in a monoidal category 9Jt is said to be a monoid if there are morphisms M®M -> M,
lgn -+ Af,
which obey the commutative diagrams (M®M)®M
I
—>M®(M®M)
I
M®M —> M <—M®M lm®M —>M®M <—M®lm
\
j /
M
Chapter 7 Deformation Quantization
477
Example 7.6.1. The category of sets Ens in Example 10.1.3 is a symmetric monoidal category where ® is the Cartesian product x of sets and the unit object is some singleton set. • Example 7.6.2. Similarly to Ens, the symmetric monoidal category Top of topological spaces is denned. Top, is the category of topological based spaces. Its object are topological spaces with a base point whose maps are continuous functions preserving base points. This category can be provided with the following two symmetric monoidal structures. Firstly, there is that given by the Cartesian product x and a singleton set. Secondly, there is the join V of two spaces by their base points whose unit is also a singleton set. • Example 7.6.3. The objects of the sceletal category N of finite sets are finite sets (n) = (0,..., n — 1) for each integer n € N and the empty set 0. Their morphisms are functions. The disjoint union (the addition) provides N with a monoidal product, whose unit object is 0. Then N becomes a symmetric monoidal category. Let N* be the category whose objects are also finite sets (n), but their morphisms are order-preserving functions. Then the disjoint union makes it into a monoidal category, which however is not symmetric. • Example 7.6.4. The category Mod of modules over a commutative ring K is a symmetric monoidal category where
(7.6.2)
We will use the latter because this is the case of the category of chain complexes in Example 7.6.5 below. D Example 7.6.5. Let us denote ChC the category of Z-graded chain complexes of JC-modules. The tensor product and the unit object are the usual ones. The symmetry is given by the formula (7.6.2) because the other one is not a chain map. •
478
Geometric and Algebraic Topological Methods in Quantum Mechanics
DEFINITION 7.6.2. Let Wl and £ be monoidal categories. A monoidal functor consists of a functor F : 9Jt —» £ together with the isomorphisms Co : F(lm)
CKK>-F{K®K')->F{K®K'),
-> 1 £
(7.6.3)
such that the following diagrams commute for any objects K, K'', K" of 971: F(K ®K'® K")
I
—> F(K ® K')
I
F(K) ® F(if' O A"") — • F{K) ® f (JFT') <8> /^(/if")
F(/r) ® F(1 OT ) <—-F(tf ® IOR)
I ^
If 371 and £ are symmetric monoidal categories, the additional commutative diagram
a
F{K®K') —>F(K)®F{K') a monoidal transformation F —»orFnot) ' is amonoidal natural transformation such—» that must hold. Given two (symmetric functors F, F' : 271 £, a monoidal transformation F —>F' is a natural transformation such that F{K'®K) the following coherence diagrams —*F{K')®F(K) commute:
I
I
F{K
—> F'(K ® K')
I
F{K)®F{K')
I
-^F'{K)®F'{K') -^F'(lm) F(lm) \ / Thus, if 97T and £ are monoidal categories l(resp. symmetric monoidal cat£ egories), there is a category Mon(27T, £) (resp. SMon(27l, £)) of monoidal functors from 971 —> £.
479
Chapter 7 Deformation Quantization
Replacing isomorphisms C,KK' (7.6.3) in Definition 7.6.2 with maps, we come to the notion of a colax monoidal functor. DEFINITION 7.6.3. A non-symmetric operad P consists of a sequence {P(n)}, n G N, of sets (objects in Ens) together with an element 1 G P(l) and a function P{n)
x • • • x P(kn)
x Pih)
- > P{h
+
---kn)
satisfying the natural unit and associativity axioms. A symmetric operad consists of a non-symmetric operad P together with the right action of the symmetry group Sn on P(n), n G N, satisfying sone compatibility laws [307]. • Let us note that, in contrast with Definition 7.5.1, the operad P in Definition 7.6.3 is an operad of sets. Though a symmetric operad is usually just called an operad, we further abbreviate this term with non-symmetric operads. There is no requirement that P(0) is a singleton set. Example 7.6.6. • There is a unique operad Obj with
• There is a unique operad Sem with fln>l a Sem = < rt n . \0 n = 0 • There is a unique operad Mon with Mon(n) = 1 for all n G N. • There is a unique operad P t with f l n = 0,l n
"|«n>2
'
• Let G be a monoid in Ens. There is an operad Algo with .,
, .
(G n = 1
•
480
Geometric and Algebraic Topological Methods in Quantum Mechanics
7.6.4. Let P be an operad and DJt a monoidal category. An algebra over P (or, simply, a P-algebra) in 971 consists of an object A in 9Xt with a function DEFINITION
P{n) -> Horn (A®n, A),
neN,
satisfying some axioms [307]. If P is a symmetric operad, a P-algebra in a symmetric monoidal category is considered. • There is an obvious notion of maps of P-algebras, and we come to the category of algebras over operads Alg(P,9Jl). Example 7.6.7. The following are some examples of algebras over the operads in Example 7.6.6. • Alg(Obj,9Jt) is isomorphic to 9Jt. • Alg(Mon, Ens) is the category of monoids and Alg(Mon, Top) is the category of topological monoids. • Alg(Mon, Cat) is the category of strict monoidal categories. • Alg(Sem, Ens) is the category of semigroups and Alg(Sem, Top) is the category of topological semigroups. • Alg(Pt,Ens) is the category of pointed sets (i.e., sets with a distinguished element). • Alg(Pt,Top) is isomorphic to Top,,. • Alg(Pt, ChC) is a chain complex together with a chosen zero-cycle. • Alg(Actc, Mod) is the category of /C-linear representations of G. D
DEFINITION 7.6.5. A monoidal category with equivalence is a monoidal category 9Tt equipped with a subclass £ of morphisms in 371, called equivalences or homotopy equivalences, such that the following properties hold: (El) any isomorphism is an equivalence; (E2) if 7 = a o (3 is a composition of morphisms in 9Jt and if any two of 7, a, (3 are equivalences, then so is the third; (E3) if a : Ki —> K[ and (3 : K K'2 are equivalences then so is
a ® (3 : Ki ® K2 -» K[ ® K'2.
D Clearly, the class of isomorphisms (resp. all morphisms) in 371 is the smallest (resp. largest) possible class of equivalences in 9Jt.
Chapter 7 Deformation Quantization
481
Example 7.6.8. • The 9JI is (Top, x, 1) and equivalences are homotopy equivalences. • The 9Jt is (ChC,®,/C) and equivalences are chain homotopy equivalences. • The SDT is (ChC,®,/C) and equivalences are quasi-isomorphisms of chain complexes. Axioms El and E2 are easily justified, but axiom E3 • holds only if K. is a field. One can associate to an operad P the following monoidal category P. Its objects are the natural numbers 0,1,... and the monoidal structure on these objects is addition, with the unit 0. The homsets are given by TT
Horn (m, n) = P(m, n) = w»H
P{mi) x • • • x P(mn).
\-mn=m
Thus, if 6i G P(i) and mi H h mn = m, there is an element (#i,..., 6n) of P(m, n). The identity element of P(m, m) is (1,... ,1), consisting of m copies of the unit 1 of P. The tensor product of morphisms is defined by (e1,...,en)®(6'1,...,e'nl)
=
(61,...,dn,6'1,...,eln,).
In order to define the composition rule, let us take fa G P{ki) with k\ + • • • + km — k so t h a t (
9n) G
P(fc, m) as above. One can rewrite the collections (fci,..., km) as I / £ } , . . . , K^
, . . . , Kn, . . . , Kn
)
and, similarly, do (<j>i,..., <j>m) as
where >J G P(kjf).
T h e n we have
6io{4>l...,(j)^)eP{kl
+ --- + kT%
and the composition rule reads
(«i,...,y»fc..,w = («iow!r.,r)r..,»i°(^..,ff")). 7.6.6. Let P be an operad and Tt a monoidal category with equivalences. A homotopy P -algebra in 9JI is a colax monoidal functor DEFINITION
where £o and (,m,n
are
equivalences.
•
482
Geometric and Algebraic Topological Methods in Quantum Mechanics
It is evident that, if the only equivalences in 371 are isomorphisms, the homotopy P-algebra in 971 is just a P-algebra in OJl. Conversely, any P-algebra is a homotopy P-algebra, and the categories Alg(P,9)I) and Mon(P,9Jt) are equivalent.
Chapter 8
Non-commutative geometry
Non-commutative geometry is developed in main as a generalization of the calculus in commutative rings of smooth functions [107; 194; 267; 290]. Accordingly, a non-commutative generalization of differential geometry is phrased in terms of the differential calculus over a noncommutative ring which replaces the exterior algebra of differential forms. The Chevalley-Eilenberg differential calculus over a commutative ring in Section 1.6 is straightforwardly generalized to a non-commutative /C-ring A- However, the extension of the notion of a differential operator in Section 1.2 to modules over a non-commutative ring meets difficulties (see Section 8.3). In a general setting, any non-commutative ring can be called into play, but one often follows the more deep analogy to the case of commutative smooth function rings. In Connes' commutative geometry, A is the algebra CCO(X) of smooth complex functions on a compact manifold X. It is a dense subalgebra of the C*-algebra of continuous complex functions on X. Generalizing this case, Connes' non-commutative geometry [107; 109] addresses the differential calculus over an involutive algebra A of bounded operators in a Hilbert space E and, furthermore, studies a representation of this differential calculus by operators in E (see Section 8.5). Section 8.6 is devoted to another variant of non-commutative geometry, where one assigns to a Poisson manifold P = ^4*(<$) coming from a Lie groupoid <5 the C*-algebra of <J5 (see Section 3.4). The key point is that these assignments are functorial if one considers certain categories of Poisson manifolds, groupoids and C*-algebras.
483
484
8.1
Geometric and Algebraic Topological Methods in Quantum Mechanics
Modules over C*-algebras
Let us point out some features of modules over non-commutative algebras and, in particular, C*-algebras. Let K. throughout be a commutative ring and A a /C-ring which need not be commutative. Let Z^ denote its center. An ,4-bimodule throughout is assumed to be a commutative ,2,4-bimodule. Sometimes, it is convenient to use the following compact abbreviation [137]. We say that right and left .A-modules, .4-bimodules and Z^-bimodules are A-modules of type (1,0), (0,1), (1,1) and (0,0), respectively (or (Ai — A,)-modules where AQ = Z^ and A\ = A). Of course, A-modules of type (1,1) are also of type (1,0) and (0,1), while ^-modules of type (1,0), (0,1) (1,1) are also of type (0,0). With this abbreviation, the basic constructions of new modules from old ones are phrased as follows. • If P and P' are ^-modules of the same type (i,j), so is its direct sum P®P'. • Let P and P' be ^-modules of type (i,k) and (k, j), respectively. Their tensor product P <8> P' is an *4j-module of type (i,j) (see Remark 10.4.1). • Given an «4j-module P of type (i,j), its .4-dual P* = Horn Ai-Aj (P, -A) is a module of type (i + 1, j + l)mod 2. Let A be a complex involutive algebra. Any module over A is also a complex vector space. An ylj-module of type (1,1) is called an involutive module if it is equipped with an antilinear involution p t-> p* such that (apb)* = b*p*a*,
a,b £ A,
p £ P.
Due to this relation, an involutive module is reconstructed by its right or left module structure. In particular, an involutive module is said to be a projective module of finite rank if, seen as a right (or left) module, it is a finite projective module. Given a right module P over an involutive algebra A, a Hermitian form on P is defined as a sesquilinear A-valued form (8.1.1)
(.\.):PxP^A, (pa|pV> = a*(p|pV.
{p\p') = (p'\p)*,
P,p'eP,
a,a'£A,
(see Remark 3.1.5). A Hermitian form (8.1.1) on P yields an antilinear morphism h of P to its A-dual P* given by the formula (hp)(p') := (p\p'),
p,p' e P
(8.1.2)
Chapter 8 Non-Commutative Geometry
485
A Hermitian form (8.1.1) is called invertible if the morphism h is invertible. Let A be a C*-algebra. A Hermitian form (8.1.1) on a right A-module P is called positive if {p\p} for all p G P is a positive element of a C*-algebra A, i.e., (p\p) = aa*,
a £ A.
Let A be a unital C*-algebra. Any projective A-module P of finite rank admits an invertible positive Hermitian form. Moreover, all these forms on P are isomorphic [313]. A positive Hermitian form on a right ^4-module P endows P with the semi-norm
l|plh=ll
peP,
(8.1.3)
where ||(p|p)|| is the C*-algebra norm of (p\p) G A. Equipped with this seminorm and the corresponding topology, P is called the (right) pre-Hilbert module. It is a Hilbert module (a C*-module in the terminology of [107]) if the seminorm (8.1.3) is a complete norm. Example 8.1.1. A C*-algebra A is provided with the structure of a Hilbert ^-module with respect to the action of A on itself by right multiplications and the positive Hermitian form (a\a') := a*a',
a,a' G A.
(8.1.4)
• Example 8.1.2. Let A = C°(X) be the C*-algebra of continuous complex functions on a compact space X, and let E —> X be a (topological) complex vector bundle endowed with a Hermitian fibre metric {.\.)x- Then the space E{X) of continuous sections of E —> X is a Hilbert C°(X)-module with respect to the C°(X)-valued Hermitian form (s\s')(x) := (s(x)\s'(x))x,
s,s'€E(X).
a Given a Hilbert A-module P, by its endomorphism T is meant a continuous ^-linear endomorphism of a right module P which admits the adjoint endomorphism T*, which is uniquely given by the relation
p,p'&P
486
Geometric and Algebraic Topological Methods in Quantum Mechanics
The set BA(P) of yl-linear endomorphisms of a Hilbert A-module P is a C*-algebra with respect to the operator norm (3.1.14). Compact endomorphisms of P are defined as the closer of its endomorphisms of finite rank (see Remark 3.1.10). Let us consider endomorphisms TP:Q € BA(P) of P of the form TP,gp'=p(p'\q),
(8.1.5)
p,p',q£P.
They obey the relations S
p,q
=
J
9,P>
J-p,q-l-p',q' — 1p(q\p'),q'
=
J
p,T,,p/
The linear span of endomorphisms (8.1.5) is a two-sided ideal of BA(P) [370]. Its closure is the set TA(P) of compact A-linear endomorphisms of P. In conclusion, let us turn to projective Hilbert modules of finite rank over a unital C*-algebra A. One can show the following [313; 370]. • Let P be a right Hilbert A-module such that I d P G TA{P)- Then P is a projective module of finite rank. • Conversely, let P be a projective right ^4-module of finite rank. Then P admits a positive Hermitian form which makes it into a Hilbert module such that IdP &TA(P). • Given two positive Hermitian forms (.|.) and (.|.)' on a projective right yl-module P, there exists an invertible A-linear endomorphism of P such that (p\p'Y = (Tp\Tp'),
8.2
p,p' G P.
Non-commutative differential calculus
The notion of a differential calculus in Section 1.6 has been formulated for any /C-ring A. Let us generalize the Chevalley-Eilenberg differential calculus over a commutative ring in Section 1.6 to a non-commutative Airing .4. For this purpose, let us consider derivations u £ Z>A of A. They obey the Leibniz rule u(ab) =u(a)b + au(b),
a,b £ A,
(8.2.1)
(see Remark 1.2.1). By virtue of the relation (8.2.1), the set of derivations QA is both a Z^-bimodule and a Lie /C-algebra with respect to the Lie
487
Chapter 8 Non-Commutative Geometry
bracket [u, u'} = uu' - u'u.
(8.2.2)
It is readily observed that derivations preserve the center Z^ of A. Remark 8.2.1. If A is an involutive ring, the differential calculus over A fulfills the additional relations (a-P)*=(3*-a*,
a,/3eft*.
(6a)* = -5a*,
(8.2.3)
In particular, the second relation (8.2.3) shows that, in contrast with derivations in Section 3.8, 5 is an antisymmetric derivation of an involutive ring
•
A.
Let us consider the extended Chevalley-Eilenberg complex (1.6.4) of the Lie algebra DA with coefficients in the ring .4, regarded as a T)Amodule. This complex contains a subcomplex O*[5<4] of -Z^-multilinear skew-symmetric maps (1.6.5) with respect to the Chevalley-Eilenberg coboundary operator d (1.6.6). Its terms (^[O-A] are .4-bimodules. The graded module 0*[i).4] is provided with the product (1.6.9) which obeys the relation (1.6.10) and makes 0*[D.A] into a differential graded algebra. Let us note that, if A is not commutative, there is nothing like the graded commutativity of forms (1.6.11) in general. Since O^HA] = Horn ZA (It A, A),
(8.2.4)
we have the following non-commutative generalizations of the interior product
(u\
u£T)A,
and the Lie derivative
Lu(>)=d(WJ>)+uJ/(<£). Then one can think of elements of O 1 ^^] as being the non-commutative generalization of exterior one-forms. The minimal Chevalley-Eilenberg differential calculus O*A over A consists of the monomials aodai A • • • A dak,
&% € A,
whose product A (1.6.9) obeys the juxtaposition rule (aodai) A (bodbi) = a0d(a,ib0) A db\ — aoaxdbo A db\,
aiy 6j e A.
488
Geometric and Algebraic Topological Methods in Quantum Mechanics
For instance, it follows from the product (1.6.9) that, if a, a' £ ZA, then da A da' = -da' A da, PROPOSITION
ado! = (da')a.
(8.2.5)
8.2.1. There is the duality relation J>A = 1OomA-.A(O1A,A),
(8.2.6)
generalizing the relation (1.3.5) to non-commutative rings.
•
Outline of proof. It follows from the definition (1.6.6) of the ChevalleyEilenberg coboundary operator that (da)(u) = u(a),
a & A,
u£X)A.
(8.2.7)
This equality yields the morphism UA 3 u H-*
A),
<j>u(da) := u(a),
a £ A.
This morphism is a monomorphism because the module O U is generated by elements da, a £ A. At the same time, any element > £ HomA_A(O1A,A) induces the derivation Uj,(a) := <j>(da) of A. Thus, there is a morphism B.omA_A{OlA,A)^*A, which is a monomorphism since OlA is generated by elements da, a £ A. QED
Example 8.2.2. Matrix geometry over the algebra A = Mn of complex nxn matrices provides an important example of a non-commutative system of finite degrees of freedom [135; 290]. Let {er}, 1 < r < n 2 — 1, be an anti-Hermitian basis for the (right) Lie algebra su(n). All derivations of the algebra Mn are inner, and ur = ade r constitute a basis for the complex Lie algebra X)Mn of derivations of Mn, together with the commutation relations [ur,uq]
=csrqus,
where csrq are structure constants of the Lie algebra su{n). Since the center ZMU of Mn consists of matrices cl, c £ C, the derivation module X)Mn is an (n2 — l)-dimensional complex vector space. Let us consider the minimal Chevalley-Eilenberg differential calculus (O*Mn,d) over the algebra Mn with respect to the Chevalley-Eilenberg coboundary operator d (1.6.6). In
489
Chapter 8 Non-Commutative Geometry
particular, 0°Mn = Mn, while OlMn is a free left Mn-module of rank n 2 — 1 whose basis {9r} is the dual of the basis {ur} for the complex Lie algebra DMn, i.e., er{uq) = srqi. It is readily observed that elements 8r of this basis belong to the center of the Mn-bimodule O1Mn, i.e., a6r = 6ra,
a e Mn.
(8.2.8)
It also follows that er A 9" = -6q A 9r.
(8.2.9)
The morphism
d:Mn^
OlMn,
given by the formula (1.6.7), reads der(uq) = &deq{er) = csqres,
that is, (8.2.10)
der = cqres6«. The formula (1.6.8) leads to the Maurer-Cartan equations d6r = -\crq3eq
K6S.
(8.2.11)
If we put 6 = er9r, the equality (8.2.10) can be brought into the form da = aB — 9a,
a £ Mn.
It follows that the Mn-bimodule O1Mn is generated by only one element
•
0.
Let us turn now to a different differential calculus over a noncommutative ring which is often used in non-commutative geometry [107; 267]. Let A be a. (non-commutative) /C-ring over a commutative ring /C. Let us consider the tensor product A® A of /C-modules. It is brought into K,
an ,4-bimodule with respect to the multiplication b(a
a, a', b,cG A .
490
Geometric and Algebraic Topological Methods in Quantum Mechanics
Let us consider its submodule ^(A) generated by the elements l®a-a
a&A.
It is readily observed that d:A3a^l®a-a®l€fl1(A)
(8.2.12)
is a f21(^4)-valued derivation of A. Thus, Q1(A) is an .4-bimodule generated by the elements da, a £ A, such that the relation (da)b = d(ab) - adb,
a,b e A ,
(8.2.13)
holds. Let us consider the tensor algebra SI* (A) of the ,4-bimodule ft1 (A). It consists of the monomials aodai • • • dak,
o-i £ A ,
(8.2.14)
whose product obeys the juxtaposition rule (aodai)(bodbi) = aad{a-\bo)db\ — aaa\dbob\,
a,,6j £ A,
because of the relation (8.2.13). The operator d (8.2.12) is extended to SI* {A) by the law d(aoda\
• • • dak) '•— da^dai
• • • dak,
(8.2.15)
that makes Q* (A) into a differential graded algebra. Its de Rham cohomology groups are H°(n*(A)) = /C,
Hr>0(Sl*(A)) = 0.
If A is not a unital algebra, one can consider its unital extension A in Remark 1.1.2, and then construct the differential graded algebra Q*(A). This algebra contains the differential graded subalgebra Cl*(A) of monomials (8.2.14). The de Rham cohomology groups of fl*(A) are trivial. Of course, Q.*(A) is a minimal differential calculus. One calls it the universal differential calculus over A because of the following property [267]. Let P be an ,4-bimodule. Any P-valued derivation A of A factorizes as A = fA o d through some (A — .4)-module homomorphism fA : Q^A) -> P.
(8.2.16)
Moreover, let A' be another /C-algebra and (SI1*, 6') its differential calculus over a /C-ring A'. Any homomorphism A —> A' is uniquely extended to a
Chapter 8 Non-Commutative Geometry
491
morphism of differential graded algebras
P*
: n*(A) -> n"
such that /9fe+1 od = 5' o pk. Indeed, this morphism factorizes through the morphism of $1* (A) to the minimal differential calculus in Q,'* which sends da —> 5'p(a).
Elements of the universal differential calculus Q* (A) are called universal forms. However, they can not be regarded as the non-commutative generalization of exterior forms because, in contrast with the Chevalley-Eilenberg differential calculus, the monomials da, a e Z^, of the universal differential calculus do not satisfy the relations (8.2.5). In particular, if A is a commutative ring, the module O1 (1.3.2) of exterior one-forms over A is the quotient of the module ^(A) (8.2.12) of universal forms by the relations (1.3.1). At the same time, if P — A, the morphism (8.2.16) takes the form fA{da) = A(a). This relation defines the monomorphism of fix(«4) to C 1 ^^.] (8.2.4) by the formula (8.2.7). Therefore, its range coincides with the term O1A of the minimal Chevalley-Eilenberg differential calculus, i.e., there is an isomorphism Q1(A)=O1A.
(8.2.17)
R e m a r k 8.2.3. There is the following relation between the universal differential calculus £l*(A) over a /C-ring A and the Hochschild chain complex B*{A,A) in Remark 1.5.3 [283; 368]. Let us assign to each Hochschild chain a 0 <8> • • • ® afc €
Bk(A,A)
the universal cochain UJ - a 0 ® • • • ® afc e £lh(A). This assignment provides an isomorphism of graded /C-modules £?* (A, A) and Q*(A). Let us write u = u'dak. Then the Hochschild boundary operator (1.5.35) on SI*(A) reads 3fc(w) = (-l) fc (o fc w' - w'ofc).
(8.2.18)
•
492
8.3
Geometric and Algebraic Topological Methods in Quantum Mechanics
Differential operators in non-commutative geometry
It seems natural to regard derivations of a non-commutative /C-ring A and the Chevalley-Eilenberg coboundary operator d (1.6.6) as particular differential operators in non-commutative geometry. Definition 1.2.1 provides a standard notion of differential operators on modules over a commutative ring. However, there exist its different generalizations to modules over a non-commutative ring [50; 136; 286]. Let P and Q be ,4-bimodules over a non-commutative /C-ring A. The K,-module Horn K. (P, Q) of /C-linear homomorphisms $ : P —> Q can be provided with the left A- and ^'-module structures (1.2.1) and the similar right module structures ($o)(p) := $(p)a,
(a«$)(p) :=$(pa),
a £ A,
p G P. (8.3.1)
For the sake of convenience, we will refer to the module structures (1.2.1) and (8.3.1) as the left and right .4 —.4* structures, respectively. Let us put 5o$:=$a-a»$,
a G A,
$eHomic(P,Q).
(8.3.2)
It is readily observed that 5aoSb
= Sb° Sa,
a,b e A.
The left ^.-module homomorphisms A : P —» Q obey the conditions 5aA = 0, for all a G A and, consequently, they can be regarded as left zero order Q-valued differential operators on P. Similarly, right zero order differential operators are defined. Utilizing the condition (1.2.3) as a definition of a first order differential operator in non-commutative geometry, one however meets difficulties. If P = A and A(l) = 0, the condition (1.2.3) does not lead to the Leibniz rule (1.2.8), i.e., derivations of the /C-ring A are not first order differential operators. In order to overcome these difficulties, one can replace the condition (1.2.3) with the following one [136]. 8.3.1. An element A e Homx;(-P, Q) is called a first order differential operator on a bimodule P over a non-commutative ring A if it obeys the condition DEFINITION
a,beA, <5ao5hA = <56o<5aA = 0, aA{p)b - aA(pb) - A{ap)b + A{apb) = 0,
p G P.
(8.3.3)
493
Chapter 8 Non-Commutative Geometry
• First order Q-valued differential operators on P make up a Z^-module Diffi(P,Q). If P is a commutative bimodule over a commutative ring .4, then Sa = 5a and Definition 8.3.1 comes to Definition 1.2.1 for first order differential operators. In particular, let P = A. Any left or right zero order Q-valued differential operator A is uniquely defined by its value A(l). As a consequence, there are left and right .A-module isomorphisms Q 9 ^ A « S Diff ${A, Q), Q3q^A^£Di$^(A,Q),
Af(a) = qa,
a G A,
A^(a)=aq.
A first order Q-valued differential operator A o n ^ l fulfils the condition A(o6) = A{a)b + aA(b) - aA(l)6.
(8.3.4)
It is a derivation of A if A(l) = 0. One obtains at once that any first order differential operator on A is split into the sums A(a) = oA(l) + [A(a) - oA(l)], A(o) = A(l)a + [A(a) - A(l)a] of the derivations A(a) - oA(l) or A(a) — A(l)a and the left or right zero order differential operators aA(l) and A(l)a, respectively. If u is a Qvalued derivation of A, then au (1.2.1) and ua (8.3.1) are so for any a G Z^. Hence, Q-valued derivations of A constitute a Z^-module T)(A,Q). There are two Z^-module decompositions JM1(A,Q) = BiS^{A,Q)®J)(A,Q), Di$1(A,Q) = DiS*(A,Q)®*(A,Q). They differ from each other in the inner derivations a — i > aq — qa. Let Hom5(P, Q) and Hom^(P, Q) be the modules of right and left Amodule homomorphisms of P to Q, respectively. They are provided with the left and right ,4-.4#-module structures (1.2.1) and (8.3.1), respectively. 8.3.2. An element A G Hom c (F, Q) is a first order Q-valued differential operator on P in accordance with Definition 8.3.1 if and only if it obeys the condition
PROPOSITION
A(apb)
= (da)(p)b
+ aA(p)b + a(db)(p),
p G P,
a,b G A,
(8.3.5)
494
Geometric and Algebraic Topological Methods in Quantum Mechanics
where d and d are Kom^(P, Q)- and Hom^(P, Q)-valued derivations of A, respectively. Namely, (da){pb) = (da)(p)b,
(db)(ap)=a(db)(p). D
Outline of proof. It is easily verified that, if A obeys the equalities (8.3.5), it also satisfies the equalities (8.3.3). Conversely, let A be a first order Qvalued differential operator on P in accordance with Definition 8.3.1. One can bring the condition (8.3.3) into the form A(apb) = [A(ap) - aA(p)]b + aA(p)b + a[A(pb) - A(p)b], and introduce the derivations (9a)(p) := A(op) - oA(p),
(db)(p) := A(pb) - A(p)b. QED
For instance, let P be a differential calculus over a /C-ring A provided with an associative multiplication o and a coboundary operator d. Then d exemplifies a P-valued first order differential operator on P by Definition 8.3.1. It obeys the condition (8.3.5) which reads d(apb) = (da op)b + a{dp)b + a({-l)Mp
o db).
For instance, let P = O*A be the Chevalley-Eilenberg differential calculus over A. In view of the relations (8.2.4) and (8.2.6), one can think of derivations u £ T)A as being vector fields in non-commutative geometry. A problem is that QA is not an «4-module. One can overcome this difficulty as follows [50]. Given a non-commutative /C-ring A and an .A-bimodule Q, let d be a Q-valued derivation of A. One can think of Q as being a first degree term of a differential calculus over A. Let Q^ be the right ^4-dual of Q. It is an .A-bimodule: (bu){q) :=bu(q),
(ub){q) := u{bq),
b £ A,
q£Q.
One can associate to each element u G Q^ the /C-module morphism u : i 6 a H u(da) € A.
(8.3.6)
Chapter 8 Non-Commutative Geometry
495
This morphism obeys the relations (bu)(a) = bu(da),
u(ba) = u{b)a + (ub)(a).
(8.3.7)
One calls (QR,U H-> U) the Aright Cartan pair, and regards u (8.3.6) as an A-valued first order differential operator on A [50]. Let us note that u (8.3.6) need not be a derivation of A and fails to satisfy Definition 8.3.1, unless u belongs to the two-sided .A-dual Q* C QR of Q. Morphisms u (8.3.6) are called into play in order to describe (left) vector fields in noncommutative geometry [50; 232]. In particular, if Q = OlA, then au for any u € DA and a G A is a left non-commutative vector field in accordance with the relation (1.6.7). Similarly, the ,4-left Cartan pair is denned. For instance, ua for any u G T)A and a G A is a right non-commutative vector field. If .A-valued derivations u\,.. .ur of a non-commutative /C-ring A or the above mentioned non-commutative vector fields u\,... ur on A are regarded as first order differential operators on A, it seems natural to think of their compositions ui o • • • ur or u\ o • • • uT as being particular higher order differential operators on A. Let us turn to the general notion of a differential operator on ,4-bimodules. By analogy with Definition 1.2.1, one may try to generalize Definition 8.3.1 by means of the maps 5a (1.2.2) and Sa (8.3.2). A problem lies in the fact that, if P = Q = A, the compositions Sa o <5b and 5a o Sb do not imply the Leibniz rule and, as a consequence, compositions of derivations of A fail to be differential operators. This problem can be solved if P and Q are regarded as left ^-modules [286]. Let us consider the /C-module Horntc{P,Q) provided with the left A — A* module structure (1.2.1). We denote by ZQ its center, i.e., 5Q$ = 0 for all $ g Zo and a & A. Let J o = ^o be the A - A* submodule of Horn A; (P, Q) generated by Zo. Let us consider: (i) the quotient Hom^P, Q)/Io, (ii) its center Z\, (iii) the A-A* submodule ~Z\ of Hom K (P, Q)/Xo generated by Zx, (iv) the A — A' submodule 1\ of Hom^(P, Q) given by the relation I\/1Q = Z\. Then we define the A — A* submodules 2r, r — 2,..., of Horn K(P, Q) by induction as J r / I r _ i = ZT, where ZT is the A- A' module generated by the center Zr of the quotient HornK.{P,Q)/Z r -iDEFINITION
8.3.3. Elements of the submodule 1T of Rom ,c(P, Q) are said
496
Geometric and Algebraic Topological Methods in Quantum Mechanics
to be left r-order Q-valued differential operators on an *4-bimodule P [286].
•
PROPOSITION 8.3.4. An element A G Horn K(P, Q) is a differential operator of order r in accordance with Definition 8.3.3 if and only if it is a finite sum A(p) = bi&{p) + A P _i(p),
bi G A,
(8.3.8)
where A r _i and 8a$l for all a G A are (r - l)-order differential operators if r > 0, and they vanish if r = 0. • Outline of proof. If r = 0, the statement is a straightforward corollary of Definition 8.3.3. Let r > 0. The representatives $ r of elements of Zr obey the relation 5c$r = A'r_v
(8.3.9)
ceA,
where A'r_1 is an (r — l)-order differential operator. Then representatives $ r of elements of ZT take the form
$ r (p) = Y, ci&(ciP) + A;'_I(P),
i
where $* satisfy the relation (8.3.9) and A"_! is an (r —l)-order differential operator. Due to the relation (8.3.9), we obtain
^r(p) = ^ i ( p ) + A;/L1(p), ^ = c , 4
A - ^ - ^ C ^ ^ + A;'^.
(8.3.10) Hence, elements of Ir modulo elements of Ir~i take the form (8.3.10), i.e., they are given by the expression (8.3.8). The converse is obvious. QED If A is a commutative ring, Definition 8.3.3 comes to Definition 1.2.1. Indeed, the expression (8.3.8) shows that A G iiomic(P,Q) is an r-order differential operator if and only if 6a A for all a G A is a differential operator of order r — 1. PROPOSITION 8.3.5. If P and Q are ,4-bimodules, the set Ir of r-order Qvalued differential operators on P is provided with the left and right A —A' module structures. • Outline of proof. This statement is obviously true for zero order differential operators. Using the expression (8.3.8), one can prove it for higher order differential operators by induction. QED
497
Chapter 8 Non-Commutative Geometry
Let P = Q = A. Any zero order differential operator on A in accordance with Definition 8.3.3 takes the form a H-» cac' for some c, c' £ A. PROPOSITION 8.3.6. Let Ai and A2 be n- and m-order .4-valued differential operators on A, respectively. Then their composition Ai o A2 is an (n + m)-order differential operator. •
Outline of proof. The statement is proved by induction as follows. If n = 0 or m = 0, the statement issues from the fact that the set of differential operators possesses both left and right A — A' structures. Let us assume that A o A' is a differential operator for any fc-order differential operators A and s-order differential operators A' such that k + s < n + m. Let us show that Ai o A2 is a differential operator of order n + m. Due to the expression (8.3.8), it suffices to prove this fact when 6aAi and <5aA2 for any a £ A are differential operators of order n — 1 and m — 1, respectively. We have the equality 6a(A1 o A2)(6) = o(A! o A2)(6) - (A, o A2)(o6) = Ai(oA2(6)) + {8aA1 o A2)(6) - (Ax o A2)(o6) = (A1o«5aA2)(6) + (5 a A 1 oA 2 )(6), whose right-hand side, by assumption, is a differential operator of order QED n + m — 1. Any derivation u G QA of a /C-ring A is a first order differential operator in accordance with Definition 8.3.3. Indeed, it is readily observed that (Sau)(b) = au{b) — u(ab) = —u(a)b,
b € A,
is a zero order differential operator for all a G A. The compositions au, u» a (1.2.1), ua, a»u (8.3.1) for any u £ DA, a € A and the compositions of derivations uio- • -our are also differential operators on A in accordance with Definition 8.3.3. At the same time, non-commutative vector fields do not satisfy Definition 8.3.3 in general. First order differential operators by Definition 8.3.1 also need not obey Definition 8.3.3, unless P = Q = A. By analogy with Definition 8.3.3 and Proposition 8.3.4, one can define differential operators on right ,4-modules as follows. 8.3.7. Let P and Q be seen as right ^-modules over a noncommutative /C-ring A. An element A e Hom^(P, Q) is said to be a right
DEFINITION
498
Geometric and Algebraic Topological Methods in Quantum Mechanics
zero order Q-valued differential operator on P if it is a finite sum A = $'&», bi G A, where <Sa$* = 0 for all a £ A. An element A G Hom;c(P, Q) is called a right differential operator of order r > 0 on P if it is a finite sum &(p) = &(p)bi
+ Ar-1(p),
bte A ,
(8.3.11)
where A r _i and (5a$l for all a G A are right (r — l)-order differential operators. • Definition 8.3.3 and Definition 8.3.7 of left and right differential operators on .4-bimodules are not equivalent, but one can combine them as follows. 8.3.8. Let P and Q be bimodules over a non-commutative /C-ring A. An element A G Horn K.(P, Q) is a two-sided zero order Q-valued differential operator on P if it is either a left or right zero order differential operator. An element A G Horn jt(P, Q) is said to be a two-sided differential operator of order r > 0 on P if it is brought both into the form DEFINITION
A ^ b i ^ + Ar-i,
he
A,
A = $ % + A r _i,
he
A,
and
where A r _i, A r _j and 5a$l, <5a$ for all a G A are two-sided (r — l)-order differential operators. • One can think of this definition as a generalization of Definition 8.3.1 to higher order differential operators. It is readily observed that two-sided differential operators described by Definition 8.3.8 need not be left or right differential operators, and vice versa. At the same time, »4-valued derivations of a /C-ring A and their compositions obey Definition 8.3.8.
8.4
Connections in non-commutative geometry
This Section is devoted to the definitions of a connection in noncommutative geometry. We follow the notion of an algebraic connection in Section 1.3 generalized to modules over non-commutative rings [107; 137; 138].
499
Chapter 8 Non-Commutative Geometry
Let (fi*, 5) be a differential calculus over a /C-ring A, and let P be a left .A-module. Following Definition 1.3.2, one can construct the tensor product of modules Q,1
VL : P-> Q,1 <8> P, A
which obeys the Leibniz rule V L (ap) = 5a®p + dVL(p),
p £ P,
a £ A,
[267; 430]. If fi* = Q* (.4) is the universal differential calculus over A, the connection (8.4.1) is called a universal connection. Similarly, a right non-commutative connection on a right .4-module P is defined as a /C-module morphism VR:P^P®n\ A
which obeys the Leibniz rule VR(pa) =p®5a + VR(p)a,
p £ P,
a £ A.
The forthcoming theorem shows that a connection on a left (or right) module over a non-commutative ring need not exist [115; 267]. 8.4.1. A left (resp. right) universal connection on a left (resp. right) module P of finite rank exists if and only if P is projective. • THEOREM
In contrast with connections on left and right modules, there is a problem how to define a connection on an .4-bimodule over a non-commutative ring. If A is a commutative ring, then the tensor products Q1 (g) P and P ® fi1 are naturally identified by the permutation morphism Q : a®p — i > p<S> a,
asfi1,
p £ P,
so that any left connection V^ is also the right one goVL, and vice versa. If A is a non-commutative ring, neither left nor right connections are connections on an .4-biinodule P since V L (P) £ fi^P, whereas V f i (P) £ P&H1. As a palliative, one may assume that there exists some ,4-bimodule isomorphism £>:ft 1
(8.4.2)
500
Geometric and Algebraic Topological Methods in Quantum Mechanics
Then a couple (V L , V fl ) of right and left non-commutative connections on P is said to be a ^-compatible if go\7L =VR [138; 267; 324] (see also [117] for a weaker condition). However, this couple is not a true connection on an ,4-bimodule. The problem is not simplified even if P —ft1,together with the natural permutation <j> ®