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 + V(x), v(E)) is any observable associated with j£? such that the range of v is contained in j£?0, there exists a real valued Borel function 9 on Rn such that v(E) — u(y ~ 1(E)) for all E. Proof. If u is an observable with range ££u, S£u is obviously the smallest sub -a- algebra of ££ containing all the u(E), where E is any open interval of R1 with rational end points. This shows that ££u is separable. Suppose conversely that ££x<=,££ is a separable sub-a-algebra of ££. By theorem 1.3 there exists a set X, a a-algebra £f of subsets of X, and a (T-homomorphism h of Sf onto J^. Let {^4n:% = l , 2 , - - - } b e a countable family of sets of Sf such that ££^ *s the smallest sub a-algebra of ££ ~1(E)) is an observable associated with j£? whose range is obviously contained in j£?0. Suppose now that v is an observable associated with 3? whose range «5?vcjg?0. If we use the notations of the previous paragraph, and define £f~ by &>- = { / - i ( J p ) :Fe@{Rn)}, then h maps £f~ onto J5?0. Applying theorem 1.4 to £f~ and v9 we infer the existence of a real valued y ~ -measurable Borel function c o n l such that h(c~1(E)) = v(E) for all Borel sets E of R1. By lemma 1.5, since c is S?~ -measurable, there exists a real valued Borel function (?{x)) for all x e X. If now 2£ is any Borel set on the line, we have: i be a sequence of mutually orthogonal elements of j£? and let a = \/nan. Let a; be a discrete observable such that ^({0})=^ and x({n}) = an, n = 1, 2, • • •. Let fn be the function on E1 whose value is 1 for the point n and 0 otherwise. Then/ n o x is the question qUn and we have, by (21) and (13), p(an) = Px({n}). Since Px is a probability measure, it follows that 0) given by 9(x) = ( ^ ) % x p { - ( y / 2 S ) (*-*)* + (p/h)ix}; then oM = («/2y)*, S-(p is called the wave function, and since the spectral measure of q is E-*qs where qs is multiplication by the characteristic function of E, the probability distribution of q in the state g> is the measure fi :E-^ Q(c; E) be the spectral measure of A(c). If cp G Jtf is a unit vector, and if we define ^ ( c ; E) by (63) q^^M) is bounded and *-bilinear on J f x«?f. To verify this, let 9^, 902, is a D-valued measure on <^(R2). However, if M = E(c), where EeSKJk1), then g(?)(if) = . Consequently, v(M) = 0 whenever M = E(c). Hence v = 0. Similar arguments complete the proof that 99, q^^M) is *-bilinear. Since Q(c; E) = q0(c; E) is a nonnegative measure on ^ ( R 1 ) . We define (70) 0(ci, |2 evidently depends only on r and r' and not on cp and cp'. We denote it by [r,r']. It is called the transition probability between the states p and p' which correspond (cf. theorem 4.23) to the rays r and r'. If a physical system © is in the state p, [r,rr] is the probability that it will be actually found in the state p' when an experiment is performed to ••• • If the initial state is not pure but mixed and U is the von Neumann operator representing it, the operator of the changed state is (30) x) 0 be a Borel function such t h a t a — dd\da. Let (IP\P') be the system of imprimitivity associated with a and 95. Then W:f->a~ll2f z) = v) x) n(g,g-1 •x)w{g~1 -x)-1 ', it is clear that (U,P) is equivalent to a system induced by m. The only thing that remains to be proved is the statement about the commuting rings. We assume (as we may) that X = X'. Let 9 be a strict (6r,X,ikfj-cocycle defining the homomorphism m at x0 of G0. We choose a quasi-invariant cr-finite measure a on X and write 3^ = 3%^^. In our usual notation, for each g e G and Borel E^X, p 9''Xo))' Suppose that T is an operator in J f which commutes with all PE and all Ug. By lemma 6.4 we can choose a Borel map t, (64) g-x, is an a-measurable subset of X. We define (71) gt be a one-parameter group of diffeomorphisms of M such that t, x->t -x — gt(x) is O00 from R1 x M into M. Let Z be the vector field induced by t^gt. Suppose that v is invariant under all the gt. Then, the operator Bz is symmetric on the domain CC*>(M), and its closure is self-adjoint in 3tf. / / we denote this closure by Bz also, then for any t, exj)( — itBz) is the operator f-^fW in Jtf*, i.e., (e*V(-itBz)f)(x)=f(gt-1x). (M)i let u 0 q> be the function s, x -> u(s)cp(x). Then, ff(Z'(tt®9)))(*,a?)e^/o(«)*(a;)*? = f°° | ( s ) e ^ [ —iZcp is selfadjoint. Now the operators Vt, where g-x of GxM into M is 0 00 . Let v be a Borel measure invariant under G. We then take J f = J2?2(JI4» and define (26) tk(p(tl9 • • •, y , and the linear momentum in the direction of the jth coordinate axis, say pj9 is represented by the operator —i(dldtj): (34) ; (ii) [AM,A(ffl = 0; (iii) [A(9),A(+y] = (>l>> ),A{i)} = 0; (iii) {A( Y} = (Ml- yA(ip) + A( — d(q>)u ' = ( ) (um A un) = (d((p) um) A un + ( - 1 )m um A d(q>) un (umeAmiuneAn). ,.) U ?/_ * transfers the measure a on 0 to a quasi-invariant measure j8 on O'1. Choose any anti-unitary automorphism S of Jf* and define J by q>(p)) is simply the integral Po~\ d<Xm + (P)JX + ), 2dfo'2(v). JR3 dA+'1(p), ndfa-n(j>), Jp8 a acts as follows: w2->R,e(z2)w2-Im(z2)w3, w3 - > Im(z 2 )w 2 + Re(z 2 )w 3 . Equations (174) show t h a t the representation of E* in question is t h e direct sum of the two characters f )|. peP4 P)2po' 1 is Lebesgue integrable in the domain p0>0 of P 3 . But as P -+f(Po,V)Pi(s 0) of the group Aut(«j2?) as t h e automorphism induced by the symmetry Sx, where (%>)(^o>p) = <M:Po,-p), (200) ) 1/2 ,p) and reduce this to the form (213). Corollary 9.14 is then applicable and leads to the following necessary and sufficient condition for localizability: © is localizable if and only if there exists a representation TT' of K* such that, for almost all r > 0 , the representation of the stability subgroup at ((m 2 + ^ 2 ) 1/2 ,0,0,r), = r*} = r2}; in fact, (236) shows that in the Hilbert space CtiC being the space of 7r,. VU acts as multiplication by the function p —> exp *'P> and hence QE must be the operator of multiplication by XE> proving that Q vanishes for the complement of the set = r2. Therefore, the projection valued measure WQW'1 also vanishes for the complement of the set {p : = ^ 2 }- On the other hand, we use the Fourier-Plancherel isomorphism (cf. (209)) of J^2(R3,Jf,dx) with ^ 2 (P 3 ,JT,^p) to find (3TWVuW-^-y)(V) P) = r2 } °f z e r o Lebesgue measure. Therefore, the systems corresponding to the representations Ln are not localizable. Our analysis therefore shows that we must consider the projective representations of the inhomogeneous Galilean group in order to be able to build a theory of Galilean invariant particles. We shall now indicate the details of this theory. We shall use the results and the theory of multipliers as developed in Chapter VII. According to the results we have obtained there, any nonexact multiplier of the covering group of the Galilean group is similar to the multiplier mz which is given by (239) + 2 ^ r p 0 = + 2 ^ + 2wTp0 is invariant under Hx for each fixed n. Let us now define the sets ZUtP by (246) + 2 ^ ),p; n) which is, of course, the general point of ZUtP. The stability subgroup of Hx a t an p is the group of all elements of the form ( M ; 1) which is isomorphic t o K* (the unitary group). The m a p ^:p-^((2^r)-1(p- ; n) to the element (p0 — J ; ].
where V is a C°° function on if. The function x,p-> (p*,p*}x then represents the kinetic energy of the system in question. It may be pointed out that one can introduce the concept of the momenta of the system in this setup. Let (7)
V-t-+yt
4
GEOMETRY
OF QUANTUM
THEORY
be a one-parameter group of symmetries of the configuration space M, i.e., y(t - > yt) is a one-parameter group of C00 diffeomorphisms of i f onto itself such t h a t the m a p t, x - > y t (#) oi RxxM into i f is O00. The infinitesi00 mal generator of y is a (7 vector field, say X y , on M; for a n y # e i f and a n y real valued C00 function / defined around x,
WKX) = {a/(^*))}t.oX y defines, in a natural fashion, a 0 0 0 function ytxy on 8. I n fact, if a; e M and # e Mx*, fiy(xtp) = 2>(Xy(z)) (here Xy(x) denotes the tangent vector to M a t x which is the value of XY a t x). The observable corresponding to the function /zy is called .the momentum of the system corresponding to the one-parameter group of symmetries y. If M = Rn, if xlf • • •, xn are the global affine coordinates on if, and if yte(xl9 • • •, xn) = &!-&>!, • • •, xn-tcn, then t h e observable corresponding to /xyc is called the component of the linear momentum along (clt • • •, c n ). I n the same case, if y!''(si»-"»*n) = (2/i>"->2/n)> where i/r = z r ,
r ^ i, j ,
^ = xt cos £ + a;y sin t,
y5 = —x{ sin t + # y cos £,
then t h e observable corresponding to juy.* is called the angular momentum with respect to a rotation in the i-j plane. A straightforward calculation shows t h a t in the case when M = Rn, S = Bnx Rn, and x1,---,xn>p1,'--ipn are global coordinates on 8 ((xl9 • • • , # „ , pl9 • • •, pn) depicts 2?= iPi(dXi)x)9 fiYc(x,p) = 0^
+- -
-+cnpn,
and Suppose now t h a t M is a general C00 manifold and 8 its cotangent bundle. I f / and g are two 0°° functions on $, then we can form J~ 1(df)9 which is a O00 vector field on 8, and apply it to g to get another (700 function on #, denoted by [f9g]: (8)
[/,] =
(J-Hdf))g.
[f9g] is called the Poisson Bracket o f / with g. If we use local coordinates xl9 • • •, xn on i f and the induced coordinates #!, • • •, xn, pl9 • • •, pn on 8 (so t h a t (xl9 - • •, xn9pl9 • - ',pn) represents ^iP^dXi)), then J goes over
BOOLEAN
ALGEBRAS
ON A CLASSICAL
PHASE
SPACE
into the map which (locally) sends ^iA^djdx^+^B^djdp^ — 2t B% dxt + 2 i A{ dpi, and [f,g] becomes
5
into
The m a p / , gr —> [f,g] is bilinear, skew symmetric, and satisfies the identity
as is easily verified from (9). If X is any O00 vector field on M and /xx is the <7°° function on 8 denned by
PxfaP) = P(x(x))> then one can verify, using (9), that Pax + bY =
fl/zx+6/xy
(a, b constants),
where [-X,F] is the Lie bracket of the vector fields X and Y. If / is any C00 function on M a n d / 0 is the lifted function on 8, i.e.,
f°(x,p) =/(*), then we may use (9) once again to check that
for any C°° vector field X on M. In many problems, there is a Lie group 6r which acts on M and provides the natural symmetries of the problem. For g e G we write a; -» g-x for the symmetry associated with g and assume that g,x->g-x is C00 from GxM into .AT. In such problems, one restricts oneself to the momenta specified by the one-parameter groups of M. If g is the Lie algebra of G (cf. Chevalley [1]) and if we associate for l e g , the vector field on M denoted by X also and defined by (X/)(x) = ( | / ( e x p t X . * ) ) t _ o , then we obtain the relations (10)
MEX.Y] — [MXJ/^YL
[/xX)/°] = (X/)° between the configuration observables / ° and the momentum observables fjux. These relations are usually referred to as commutation rules.
6
GEOMETRY
OF QUANTUM
THEORY
2. T H E LOGIC OF A CLASSICAL SYSTEM We shall now examine the algebraic aspects of a general classical system. I n view of the discussion carried out just now, it is clear t h a t for a n y classical system (B there is associated a space 8 called t h e phase space of ©. The states of the system are in one-one correspondence with the points of 8. The notion of a state is so formulated t h a t if one knows the state of the system a t an instant of time t0, and also the dynamical law of evolution of the system, then one can determine t h e state of t h e system a t time t + t0. The observables or physical quantities which are of interest to t h e observer are then represented by real valued functions on S. If / is t h e function corresponding to a particular observable, its value f(s) a t the point s of S is interpreted as t h e value of the physical quantity when the system is in the state s. If s is the state of the system a t time t0, we can write D(t)s for the state of the system a t time t + t0. We thus have a transformation D(t) of 8 into itself. For each t, D(t) is invertible and maps 8 onto itself. The correspondence t -> D(t) satisfies t h e equations (3). t - > D(t) is then a one-parameter group of transformations of 8. I t is called the dynamical group of t h e system <3. These concepts make sense in every classical system. I n t h e case of a n y such system t h e most general statement which can be made about it is one which asserts t h a t the value of a certain observable lies in a real number set E. If the observable is represented by the function / on 8, t h e n such a statement is equivalent to the statement t h a t the state of t h e system lies in the setf~1(E) of the space 8. I n other words, the physically meaningful statements t h a t can be made about the system are in correspondence with certain subsets of 8. The inclusion relations for subsets naturally correspond t o implications of statements. I n mathematical terms, this means t h a t at t h e background of t h e classical system there is a Boolean algebra of subsets of the space S, t h e elements of which represent t h e statements about the physical system. I t is natural to call this Boolean algebra the logic of the system. Suppose now t h a t @ is a system which does not follow the laws of classical mechanics. Then one cannot associate with it a phase space in general. I t is nevertheless meaningful to consider t h e totality of experimentally verifiable statements which m a y be made about the system. This collection, which m a y be called t h e logic of @, comes equipped with t h e relations of implication and negation which convert it into a complemented partially ordered set. For a classical system this partially ordered set is a Boolean algebra. Clearly, it is possible to conceive of mechanical systems whose logics are not Boolean algebras. We take the point of view that quantum mechanical systems are those whose logics form some sort of projective geometries and which are consequently nondistributive
BOOLEAN ALGEBRAS
ON A CLASSICAL
PHASE SPACE
7
lattices. With such a point of view it is possible to understand the role played by simultaneously observable quantities, the uncertainty relations, and the complementarity principles. These phenomena, which are so peculiar to quantum systems, will then be seen to be consequences of the nondistributive nature of the logic in the background of the system (&. I t might seem a bit surprising t h a t the basic assumption on a quantum system is t h a t its logic is not a distributive lattice. I t would be natural to argue t h a t statements about a physical system should obey the same rules as the rules of ordinary set theory. The well known critiques of von N e u m a n n and Heisenberg address this question (von Neumann [1], Birkhoff-von Neumann [1], Heisenberg [1]). The point is t h a t only experimentally verifiable statements are to be regarded as members of the logic of the system. Consequently, as it happens in m a n y questions in atomic physics, it m a y be impossible to verify experimentally statements which involve the values of two physical quantities of the system—for example, measurements of the position and momentum of an electron. One can verify statements about one of them but not, in general, those which involve both of them. W h a t the basic assumptions imply is t h a t the statements regarding position or momentum form two Boolean subalgebras of the logic b u t t h a t there is in general no Boolean algebra which contains both of these Boolean subalgebras. Before beginning an analysis of the logic of general quantum mechanical systems it would be helpful to recast at least some of the features of the formulation given in section 1 in terms of the logic of the classical system. I n the first place it is natural to strengthen the hypothesis and assume t h a t the logic of a given classical system © is a Boolean cr-algebra, say j£?, of subsets of 8, the phase space of <S. Suppose now, t h a t an observable associated with the system is represented by the real valued function / on 8. The statements concerning the observable are then those which assert t h a t its value lies in an arbitrary Borel set E of the real line and these are represented by the subsets f~x(E) of 8. The observable can thus be represented, without any loss of physical content, equally by the m a p E -^f~1(E) of the class of Borel subsets of the real line into ££. The range of this mapping is a sub-a-algebra, say 3Pf. Suppose g is a real valued Borel function on the real line. Then, the observable represented by the function g of (s ->g(f(s))) can also be represented by the m a p E ->f~1(g"1(E)) from which we conclude t h a t J?gof is contained in Jjff. In order to formulate the general features of a classical mechanical system in terms of its logic J?, it is therefore necessary to determine to what extent an abstract a-algebra ££ can be regarded as a a-algebra of subsets of some space 8; further to determine the class of mappings from t h e a-algebra of Borel sets of the real line into S£ which correspond to real valued functions on 8; and to clarify the concept of functional
8
GEOMETRY
OF QUANTUM
THEORY
dependence in this general context. We shall now proceed to a discussion of these questions.
3. BOOLEAN ALGEBRAS Let j£? be a nonempty set. ££ is said to be 'partially ordered if there is a relation < between some pairs of elements of j£? such t h a t (i) a < a for all a in S£\ (ii) a < b and b < a imply a = b; (hi) a < b and b < c imply a
0 and 1 exist in j£? and 0 ^ 1 , \] a and f\ a exist for all finite subsets F of «£?. aeF
aeF
Suppose t h a t j£? is a lattice. Given any element a of j£?, an element a' of j£? is said to be a complement of a if a A a' = 0 and a v a' = 1. a is t h e n a complement of a'. j£? is said to be complemented if, given any element, there exists at least one complement of it. I t is obvious t h a t 0 and 1 have the unique complements 1 and 0, respectively. A lattice ££ is said to be distributive if for any three elements a, 6, c of j£?, the identities a A (b v c) = (a A b) v (a A c), a V (b A c) = (a V b) A (a V c) are satisfied. A complemented distributive lattice is called a Boolean algebra. A Boolean a-algebra J? is & Boolean algebra in which /\aeF a and N/aeF a exist for every countable subset F of j£\ Every element in a Boolean algebra has a unique complement. Suppose in fact t h a t j£? is a Boolean algebra and t h a t a is an element with two complements a± and a 2 . Then, one has a1 = ax A (a v a 2 ) = ( a i A a) v (#i A a 2 ) = «i A a 2 < a 2 ;
BOOLEAN ALGEBRAS
ON A CLASSICAL
PHASE SPACE
9
similarly, a2
VaeFa'
and (\ZaeF^Y = A
aep o/ for a n y finite subset F of J*f. If j£? is a Boolean cx-algebra, then t h e same identities remain valid even when F is countably infinite. If ££ is a n y Boolean algebra and a, b are elements in it with a, x e A},
is easily seen to be a maximal dual ideal; it is also easy t o check t h a t t h e correspondence x - > Jt{x) is one-one if we notice t h a t X is Hausdorff. The concept of maximal dual ideals is central in t h e proof of Stone's theorem.
10
GEOMETRY
OF QUANTUM
THEORY
Suppose that j£? is an arbitrary Boolean algebra. Using Zorn's lemma one can show easily that maximal dual ideals of -£? exist. Let X — X(S£) be the set of all maximal dual ideals of ££\ For any a G ^ w e define Xa by Xa = {Jt : Jt e X, a e Jt), X0= 0, the null set, and X± = X. We shall say that a subset A^X is open if A is the union of sets of the form Xa. This definition defines the structure of a topology on X called the Stone topology. We now have: Theorem 1.1 (Stone [1]). Let ££ be a Boolean algebra and let X = X(JP) be the space of all maximal dual ideals of ££\ Then, equipped with the Stone topology, X becomes a Stone space. The map a -> Xa is then an isomorphism of ££ with the Boolean algebra of all open-closed subsets of X. X is determined by ££, among the class of Stone spaces, up to a homeomorphism. More generally, let X and Y be Stone spaces and let J?(X) and <¥{Y) be their respective Boolean algebras of open-closed subsets. If u is any isomorphism of &{Y) onto S£(X), there exists a homeomorphism h of X onto Y such that (12)
u(A) = h-\A)
{A e &(Y))\
moreover, h is uniquely determined by (12). This theorem is very well known and we do not give its proof. The reader may consult the books of Birkhoff [1], Sikorski [1], and the paper of Stone [1] for the proof. Corollary 1.2. Let X be a Stone space and let <£ = J£(X) be the Boolean algebra of open-closed subsets of X. If t -> Dt(— oo
BOOLEAN ALGEBRAS
ON A CLASSICAL
PHASE SPACE
11
and j£?2 are two Boolean cr-algebras and h a cr-homomorphism of ££x onto ££2. The set tAr = {a : a e ££x, h(a) = 0} is a subset of Sff1 with t h e properties: (i) 0 e ./T, 1 ^ JV\ (ii) if a e JV and b < a, then b e Jf\ (iii) if F is a countable subset of JV, \/a€F a e JV*. ^ is called t h e kernel of &. Suppose conversely ££ is a Boolean cr-algebra a n d i/T a subset of ££ with properties (i) t o (iii) listed above. W e shall say t h a t elements a a n d b of 3? are equivalent, a~b, if a A b' a n d 6 A a' are in c/T. I t is easily verified t h a t ~ is an equivalence relation. Let J ^ be the set of all equivalence classes, and for a n y a in «£?, let a denote t h e unique equivalence class containing a. W e define aa is a ahomomorphism of 3? onto ££ with kernel JV*. We write JS? = ££\Jf. Theorem 1.3 (Loomis [1]). Let 3? be a Boolean a-algebra. Then there exists a set X, a a-algebra Sf of subsets of X, and a a-homomorphism h of Sf onto &. Proof. Let X be a Stone space such t h a t t h e lattice <JS?' = J?(X) of openclosed subsets of X is isomorphic t o t h e Boolean algebra j£?. Let Sf denote t h e smallest cr-algebra of subsets of X containing ££'. We denote by U and n t h e operations of set union and set intersection for subsets of X, and by V and A t h e lattice-theoretic operations in J£ a n d S£'. If Al9 A2, - • - is a n y sequence of sets in j£?', then \/n An = A exists in ££' since ££' is isomorphic to ££ and 3? is a cr-algebra. Since A is t h e smallest element of ££' containing all t h e An, it follows t h a t t h e set A — {Jn An cannot contain a n y element of ££' as a subset. The sets in 3" form a base for t h e topology of X and hence we conclude t h a t A — [Jn An cannot contain a n y nonnull open set. Since ( J n An is open, this shows t h a t A — U n ^ 4 n i s a closed nondense set. Consider now t h e class y x of all sets A e Sf with t h e property t h a t for some B in J*?', (A — B) U (B — A) is of t h e first category. If Bx and B2 are elements of j£?' such t h a t (A — Bt) U (Bi — A) is of t h e first category (i = l , 2), then it will follow t h a t (Bx-B2) U (B2-B±) is of the first category, which is not possible (by the category theorem of Baire) unless BX = B2. Thus, for a n y A in S^1 there exists a unique B = hx(A) in <£' such t h a t (A-B)U(B-A) is of the first category. Clearly <£''c^, and for 4 e-Sf", ^ ( ^ ) = ^ 1 . We claim t h a t ^ (A-B)
is a cr-algebra. Since u (J3-4) = ( ^ ' - 5 ' ) u (£'-.4')
(primes denoting complementation in X), we see t h a t for a n y A in , 9 ^ ,
GEOMETRY
12
OF QUANTUM
THEORY
A' is in Sf-L and h^A^ — h^A)'. Suppose Al9 A2i • • • is a n y sequence in yx. Write J8n = * 1 (4 n ) > A = \Jn An, B=\/n Bni B0 = {Jn Bn. B y what we said above, B — B0 is a closed nondense set. Moreover, as B0^B, we have (A-B)
U (B-A)
£ {(il-Bo) U ( £ 0 - 4 ) } U ( B - B 0 ) £ U Un-Bn)
U (B»-iin)} U (B-B0)-
n
As all members of t h e right side are of the first category, this proves t h a t A e £fx and h1(A) = \/nh1(An). I n a similar fashion we can show t h a t p|n An lies in Sfx and ^ ( f l n An) = /\n h^An)The conclusions of t h e preceding paragraph show t h a t £Pl9 is a Boolean cr-algebra <^<Sf. Since ^ contains JSP', S?1 = Sf. Moreover, we see at t h e same time t h a t hx is a cr-homomorphism of Sf onto JS?'. If we write h = koh± where k is an isomorphism of JSf" onto <£?, then & is a or-homomorphism of y onto if. Remark. Let £f be t h e cr-algebra of Borel sets on the unit interval [0,1], JV the class of Borel sets of Lebesgue measure 0, and S£ — SPjJf. Then 3? is a Boolean cr-algebra. We can obviously define Lebesgue measure A as a countably additive function A on jSf; A is strictly positive in the sense t h a t for any a^O of j£f, A(a) is positive. From this it follows t h a t any family of mutually disjoint elements of ££ is countable. On t h e other hand, since £f is countably generated, so is ££. However, any cr-algebra of subsets of some space X which is countably generated can be proved to have atoms, t h a t is, minimal elements. Since j£? does not have atoms, j£? cannot be isomorphic to a n y cr-algebra of sets.
4. F U N C T I O N S We now take u p the second question raised in section 2, namely, the problem of describing the calculus of functions on a set X entirely in terms of the Boolean a-algebra of subsets of X with respect to which all these functions are measurable. The results are summarized in theorems 1.4 and 1.6 of this section. Let X be any set of points x and Sf a Boolean cr-algebra of subsets of X. A f u n c t i o n / f r o m X into a complete separable metric space Y is said to be ^ - m e a s u r a b l e iff~1{E) e ¥ for all Borel sets E^ Y. I f / i s ^ - m e a s u r a b l e , t h e mapping E ->f~x(E) is a cr-homomorphism of the cr-algebra 3S(Y) of Borel subsets of Y into £P. Suppose now £f is an abstract Boolean cralgebra. We shall then define a Y-valued observable associated with S£ to be a n y cr-homomorphism of 38 (Y) into ££. If Y = B1, the real line, we call these observables real valued and refer to them simply as observables. F r o m our definition of cr-homomorphisms we see t h a t a m a p u(E -> u(E))
BOOLEAN ALGEBRAS
ON A CLASSICAL
PHASE SPACE
13
of «^( Y) into ££ is a Y-valued observable associated with 3? if and only if (i) u(0) = O, u(Y) = l; (ii) u(Y-E) = u(E)f for all E in ^ ( 7 ) ; (iii) if Elt E2, • • • is any sequence of Borel sets in Y, u([Jn En) = \/n u(En) and u(C)nEn) = /\nu(En). Theorem 1.4 Let X be a set, £f a Boolean a-algebra of subsets of X and h a a-homomorphism of 6? onto a Boolean a-algebra ££\ Suppose further that u(E -> u(E)) is any (real valued) observable associated with <£?. Then there exists an 6f-measurable real valued function f defined on X such that (13)
u(E) =
h{f-\E))
for all Borel sets E^R1. f is essentially unique in the sense that if g is any ^-measurable real valued function defined on X such that u(E) = h(g~1(E)) for all Borel sets E^R1, the set {x : x e X, f{x)^g(x)} belongs to the kernel
ofh. Proof. We begin with a simple observation. Suppose A and B are two subsets of X in £f such that A^B, and c any element of j£f such that h(A)
A(fV,) = A «W) = «(A *>/) = 0, we may, by replacing Ak by Ak — f\jAi
if necessary, assume that
14
GEOMETRY
OF QUANTUM
THEORY
p | ; Aj = 0. Further h{[Jj A5) = N/y u(D,) = u(V, Dt) = 1 so that h(N) = 0, where N = X — (Jj Aj. We now define a function/ as follows: TO if /(a?) = | i n f ^
(14)
zeiV, :xeAj}
if ^ e g
^
Clearly, / is finite everywhere. Moreover, for any k, f-i(Dk)n(X-N)=
(J
Ap
so that / is ^-measurable. Further,
Hf-W*)) = M U A,\ \y.rj
)
= V «W i:ry
so that h(f~1(E)) = u(E) whenever E = Dk for some &. Since the class of all E for which this equation is valid is a Boolean a-algebra, we conclude that h(f~1(E)) = u(E) for all Borel sets E. It remains to examine the uniqueness. Let g be any real valued immeasurable function on X such that h(g~1(E)) = u(E) for all Borel sets E. Then, if we write Dk for R1 - Dkt M = {x:xeX,f(x)
* g{x)}
= U {(/"MAc) n
so that A(J0 = V {«(£*)
Au(Dk')}
k
= 0.
This shows that M belongs to the kernel of h. This completes the proof of the theorem. Lemma 1.5. Let X be a set, Sf a o-algebra of subsets of X and f an Immeasurable making of X into Rn. Suppose £f~ ={f~1(F) : F e &(Rn)}. Then to any £f ~-measurable real function c on X there corresponds a real valued Borel function c~ on Rn such that c(x) = c~(f(x)) for all x e X. Proof. Since c is £f~ -measurable, there exists a sequence cn (?i = 1, 2, • • •) of ^"-measurable functions such that (i) each cn takes only finitely many values; (ii) cn(x) -> c(x) for all x e X. For any n, let Anl, An2, • • •, Ank be disjoint subsets of X whose union is X such that cn is a constant, say ani, on Ani, the ani being distinct for i = 1, 2, • • •, kn. Since cn is ^ ~ -measurable, Anie Sf~\ so there exists a Borel set Bni of Rn such that ^4 ni =/ _1 (jB ni )
BOOLEAN ALGEBRAS
ON A CLASSICAL
PHASE SPACE
15
(* = 1, 2, • • •, kn). Replacing Bni by Bni — \Jj
»~W
=
J0
if
teBni,
t$[JBni.
Clearly cn~ is Borel and cn(x) = cn~(f(x)) for all x in X. Let us define c~ on Rn as follows: lim cn~(t) if the limit exists,
{
n->oo
0 otherwise.
Clearly c~ is Borel. Since cn(x) = cn~(f(x)) and lim cn(x) exists and is equal to c(x) for all a: in X, c~(f(x)) = c(x) for all x in JT. Let = ^ b e a Boolean a-algebra. ££x<^££ is said to be a sub-a-algebra if (i) 0, 1 e jSfij (ii) if a e S?^ then a' e ££^\ (iii) if a1? a2, • • • are in ££x, then \/n a n a n d An an a r e m «^i- A sub-d-algebra J^x is said to be separable if there exists a countable subset D of j£? such that j£?x is the smallest sub-aalgebra of <£ containing D. Theorem 1.6. (i) Let ££ be a Boolean a-algebra and u(E -> u(E)) an observable associated with ££\ Then the range J?u = {u(E) : E e ^(R1)} of u is a separable Boolean sub-a-algebra of ££. Conversely, if ££x is a separable Boolean sub-a-algebra of ££, there exists an observable u associated with ££ such that S£x is the range of u. (ii) Let ut (i = l, 2, • • •, n) be observables associated with <Sf, and S^{ (i = ly 2, • • •, n) their respective ranges. Suppose j£?0 is the smallest sub-aalgebra of ££ containing all the J^. Then there exists a unique o-homomorphism u of &(Rn) (the a-algebra of Borel subsets of the n-dimensional space Rn) onto J£Q such that for any Borel set E of R1, ui(E) = u(pi~1(E)), where p{ is the projection (t1,t2,- • •, tn) -» tt of Rn onto R1. If y is any real valued Borel function on Rn, the map E->u(
16
GEOMETRY
OF QUANTUM
THEORY
containing all the h(An). We denote by £P0 the smallest a-algebra of subsets of X containing all the An. The function c : a->
(XA^*),
XA2{*)> •" •, XAS*)>
"')
(where XA denotes the function which is 1 on A, and 0 on X — A) is immeasurable from X into the compact metric space Y which is the product of countably many copies of the 2-point space consisting of 0 and 1. Moreover, it is obvious that each An is of the form c~1(F) for some Borel set J ^ c 7 , and hence £f0 = {c-1(F) : F Borel in 7}. Now, by a classical theorem (Kuratowski [1]), there exists a Borel isomorphism d of Y onto R1, so that the function cx : x -> d(c(x)) is an ^-measurable real valued function and 6^0 = {c1~1(E) : E Borel in R1}. If we now define, for any Borel set E of R1, u(E) by the equation
u(E) = McrHE)), then u is an observable associated with ££ whose range is ££x. This proves (i). We now come to the proof of (ii). Suppose ult u2, • • •, un are observables associated with j£?, having ranges Sfl9 • • •, j£?n, respectively. Each j£?t is separable and hence j£?0, the smallest sub-a-algebra of ££ containing all the ££{, is also separable. Let X, £f\ and h have the same significance as in the proof of (i). By theorem 1.4, there exists a real valued immeasurable function ft on X such that ui(E) = h(fi~1(E)) for all Borel subsets E of R1. Let / be the map x-> (fx(x), • • -,/ n (#)) of X into Rn. Then / is ^-measurable. The map u : F ->h(J~1(F)) is then a ahomomorphism of &(Rn) into <£ such that ui(E) = u{pi-1(E)) for all E etffliR1). Since &(Rn) is the smallest cr-algebra of subsets of Rn containing all the sets pi~1(E)) it is clear that the range of u is j£?0. The uniqueness of u is obvious. For any real Borel function
u(
BOOLEAN ALGEBRAS
ON A CLASSICAL
PHASE SPACE
17
Remark. The uniqueness of u, guaranteed by (ii) of theorem 1.6 shows that u is independent of the constructs X, <9*, and h, used in its construction. Consequently, for any real Borel function cp on Rn, the map E -> u((p~1(E)) is uniquely determined by ul9 u2, • • •, un and cp. It is natural to denote this observable associated with ££ by
CHAPTER II PROJECTIVE GEOMETRIES 1. COMPLEMENTED MODULAR LATTICES The point of departure of our discussion of quantum phenomenology is the observation that the partially ordered set of experimentally verifiable statements associated with an atomic system cannot be expected to possess the distributivity properties characteristic of the Boolean algebras associated with classical systems. The simplest and most interesting of the mathematical structures that model such systems are the projective geometries, namely, the lattices of subspaces of vector spaces. More precisely, let D be a division ring, and let V be a vector space of finite dimension n ^ 2 over D; we shall always suppose our vector spaces to be left vector spaces (cf. Jacobson [1]). Then j£?=j£?(F,D), the lattice of linear subspaces of V partially ordered by inclusion, is called the projective geometry of V. This chapter will begin with a brief review of the basic properties of projective geometries. Although these are too simple to serve as models for realistic quantum systems, they already possess many of the fundamental features of the more complex systems. The first example of an abstract projective geometry that anyone comes across is the projective plane, described axiomatically in terms of its points, lines, and their incidence properties. To define higher dimensional projective spaces one replaces incidence by partial order and uses a dimension function to characterize the hierarchies of points, lines, planes, and so on; the properties of incidence that are characteristic of a projective geometry are then summarized by certain natural properties of the dimension function. We recall that a lattice 3? is complemented if for any aeJ? there is a b e S£ such that a Kb = 0 and a yb = 1. A complemented lattice JS? is modular if (1)
a,2>,CGJ^,
c < a => a A (bye) = (a A 6) Vc
Boolean algebras are modular; for any division ring D and any finite dimensional vector space V over D, ^(V,D) is a modular lattice which is not a Boolean algebra if dim(F) > 2. We next introduce the notion of a lattice of finite rank. A chain in a lattice is a sequence (a^i^i^n of elements at of the lattice such that ax < a2 < ... < an, with ax ^ 0 and a{ / a , for i <j; n is called the length of the chain, and the rank of the lattice is the supremum of 18
PROJECTIVE
GEOMETRIES
19
the lengths of all possible chains. An element of a lattice is called a point if it is nonzero and minimal. If i' 2' * * *' n are points, they are called independent if x{<£ V j V t ^ for * = 1,2,..., n. The lattice ^(VJ)) has rank dim(V), its points are the one-dimensional subspaces, and independence is just t h e usual linear independence. For any lattice 3? and a n y a / 0 in iff, let us write J?[0,a] for the set of all elements b of ££ with 0dim(a) has the following properties: (i) dim(0) = 0, dim(a) is an integer ^ 0 for all a; (ii) if a < 6, dim (a) ^ dim(6), with equality only for a = b; (2) (hi) dim(ayb) + dim(aAb) — dim(a) -f dim(&) for all a, b; (iv) if s = dim(a) > 0, then s is the maximum number of independent points contained in a; and a family of independent points contained in a has sum equal to a if and only if it has s elements. I n particular dim(l) = n and dim(#) = 1 for any point x. The converse is also true; if ££ is a complemented lattice of finite rank admitting a function d:J?->Jl satisfying (i)-(iii) of (2) and such t h a t d(x) = l for all points x, then ££ is modular and d(a) =dim(a) for all a. The lattice J£?(F,D) is modular, dim being the usual linear space dimension. At the other extreme is the Boolean algebra of all subsets of a finite set, dim being the cardinality. To single out the geometries we need the notion of irreducibility. An element of a complemented modular lattice j£P is called central if it has a unique complement in ££\ which is written as a'; the reason for this terminology is the easily proved fact t h a t a e j£? is central if and only if for all x e J*? (3)
x =
{x/\a)\f{x/\a').
The set of all central elements is called the center oi<£; it is a Boolean algebra, which coincides with 3? only when J$? itself is a Boolean algebra. & is called irreducible if 0 and 1 are its only central elements. A geometry is an irreducible complemented modular lattice of finite rank. The lattices JS?(F,D) (dim(F)^2) are geometries. Every complemented modular lattice of finite rank m a y be viewed in a canonical manner as a direct sum of geometries over its centre. More precisely, let ££ be a complemented modular lattice of finite rank and let cv...,ck be the minimal (nonzero) elements of the center of 3?. Let J^,- = JSf[0,cJ
and let
& = ££X x ... x Seh
20
GEOMETRY
OF QUANTUM
THEORY
where the partial order in ££ is denned by saying that (xl9...,xk)
<
(yl9...,yk)
if and only if xt < y{ for 1 ^ i^ k. Then each J^- is a geometry, and the map that takes x e££ to (xAcv xAc2, ..., xAck)eJ? is an isomorphism of ££ with J*?, the inverse map taking (x1,...,xk) to xx\/...\fxk. For arbitrary geometries S^i9 ££ denned as above is a complemented modular lattice whose center has (0, ...0,1,0, ...,0) (1 in the j t h place) as its minimal elements (l^j^k). The usual geometrical terminology with which we are familiar may be introduced as follows. Let JS? be a complemented modular lattice of finite rank. If aeS£ and dim(a) = l, a is a 'point', if dim(a) = 2, a is a line; if dim(a) = 3, a is a plane. If a = b vc where b and c are distinct points, a is called the line joining these points, and 6, c are said to be on a; if d is a point < a, 6, c, d are said to be collinear. It can be proved that ££ is a geometry if and only if every line has at least three points on it. The fundamental theorem of classical geometry is that every geometry of rank ^ 4 is isomorphic to some j£?( F,D) wher the division ring D is determined uniquely up to isomorphism; and that if the rank is 3, this is true if (and only if) the plane in question is Desarguesian. We shall discuss this question a little later. At this time we shall take a closer look at the geometries S£ (F,D). 2. ISOMORPHISMS OF PROJECTIVE GEOMETRIES. SEMILINEAR TRANSFORMATIONS. The first problem that arises naturally is to obtain a description of all possible isomorphisms between j£?(F,D) and J?(V',D'). We suppose the vector spaces to be always finite dimensional and that the scalars act from the left. To formulate the answer we need the concept of semilinear transformations. Let D, D' be two division rings which are isomorphic and let (4)
G:C-+C<>-
(ceD)
be an isomorphism of D onto D'. Let V (resp. V) be a vector space over D (resp. D'). A map L (v->Lv) of V into V is said to be semilinear relative to a or a-linear if L(vx + v2) = Lvx + Lv2 L(cv) = c^Lv
(vv
v2eV),
(c eD, v e F).
If L is in addition a bijection of V onto Y\ we say that L is a a-linear isomorphism,
PROJECTIVE
GEOMETRIES
21
If L is a <7-linear isomorphism of V with V, then for any linear subspace M <= 7, its image X[lf] under L is a linear subspace of V, and Jf ^ ^ [ i l f ] is an isomorphism of ^ ( F , D ) with j£?(F',D'), denoted by £z,. We can now formulate the main result. Theorem 2.1. There exists an isomorphism between if?(F,D) and ^(F'jD') if and only if D and D' are isomorphic and dim( V) = dim( 7'). / / £ ^s aw i'somorphism of J*?(F,D) wi£A J5?(F',D'), there is an isomorphism a(D ^ D ' ) awd a a-linear isomorphism L(V^> V) such that | = £L. If ^-(D^D') is another isomorphism and L'( V ^ V) a r-linear isomorphism, then £L = £L' if and only if for some d e D, d ^ 0, we have cr = dc*d-1 (ceD),
(6)
L'v = dLv
(veV).
For the proof we refer to Baer [1], If D = D' it makes sense to call an isomorphism £(j£?(F,D) ^ J^(F,D')) linear if there is a linear isomorphism L(V^> V) such that £ = £L. Theorem 2.1 leads to Theorem 2.2. In order that every isomorphism ofJ?(V,D) with j£?(F',D) be linear it is necessary and sufficient that every automorphism of D be inner. The division rings of greatest importance in physics are R, the field of real numbers, C, the field of complex numbers, and H, the division ring of "Hamiltonian" quaternions. (a) R. The identity map of R is its only automorphism, i.e., R is rigid. So every isomorphism between real projective geometries is linear and the linear map is determined up to a multiplicative constant. (b) C. One knows from the theory of algebraic fields that there are infinitely many automorphisms (cf. Bourbaki [1], pp. 114-115). However, the identity and complex conjugation (c^c*) are the only analytically wellbehaved ones (e.g., measurable, bounded, etc.). Since C is commutative, any isomorphism between two complex geometries determines uniquely the automorphism of C associated with it. (c) H, the quaternions. Let 1=
(o i ) '
jl=
(o
-if'
J2=
( - i o)'
h==
[i
o)'
where i is the usual square root of — 1 in the complex number field. We define H to be the R-linear span of 1 and j a (1 ^ a < 3 ) ; since j a 2 = —1 (1 ^ a ^ 3 ) and j«j&= — j&ja=jc whenever (abc) is an even permutation of (123), H is an (associative) algebra. It is easy to verify that it is the algebra generated by j a with the above relations. Let |q| = +det(q)*= +(?o2 +
22
GEOMETRY OF QUANTUM
THEORY
q ^ 0. Thus H is a division ring. For q e H , say q = ^01 + 2i i t s conjugate q* is the quaternion g^ —2i
so that the unit quaternions constitute the group SU(2,C) of 2 x 2 unitary matrices of determinant 1. Two quaternions are said to be in the same class if they can be transformed into each other by an inner automorphism; for q, q' to be in the same class it is necessary and sufficient that tr(q) = tr(q') (tr = trace) and |q| = |q'|. All the automorphisms of H are inner so that isomorphisms of quaternionic projective geometries are linear. The map x-+x\ (#eR) imbeds R as the center of H. H may be viewed as a vector space over C, although in a noncanonical manner. We identify C as a subfield of H by the correspondence
H is then a vector space over C, scalar multiplication being from the left. Then dim c (H) = 2, and {l,j2} may be taken as a basis. For any q e H , R q :q'->q'q* is then an endomorphism of H over C, q->R q is a faithful representation, and the matrix of j 0 (1 ^ a < 3 ) is c a j a , where ea= — 1 for a = 1,3 and + 1 for a = 2. 3. DUALITIES AND POLARITIES We shall now examine the anti-automorphisms and orthocomplementations of the projective geometries j£?(F,D). It will turn out that orthocomplementations essentially arise only from Hilbert space structures, at least when D is one of R, C, and H. We begin with the concept of the dual of a division ring. Let D be a division ring. We define D° to be the ring with the same elements as D, with the same rule for addition, but reversing the order of multiplication. D° is a division ring, said to be dual to D. Isomorphisms of J)1 with D2° may be viewed as anti-isomorphisms of T>1 with D2. If V is a vector space over D, its dual, defined as the space of D-linear maps of V into D, becomes a vector space over D° if we define scalar multiplication by writing
(«'/)W=/Wa
(aeD°,veV9feV*);
addition in V* is defined in the usual way. F* is then a (left) vector space over D° and has the same dimension as V. Clearly (D°)° = D and so V can
PROJECTIVE
GEOMETRIES
23
be canonically identified with F**. If (vv ..., vn) is a basis of V, the corresponding dual basis (v±*,...,«„*) of F* is defined by v-*(vk) = Sjk. If i f <= F is a linear subspace, its annihilator is the linear subspace of F* defined by (7)
Jf° = {/:/eF*,/(v) = 0 for all veM}.
It is clear that (8)
dim( Jf) + dim( Jf °) = dim( V) 00
and so M = M . The map M->M° is obviously an inclusion-reversing bisection of J£?( F,D) with «J^( F*,D°). We call it a duality, and use the same term to refer to any inclusion-reversing bijection of one geometry onto another. To determine the most general duality we need the concept of a nonsingular semibilinear form. Let 6 be an anti-automorphism of D, F a vector space over D. A 0-bilinear form is a map <. , . ) : x, y ~> (x, y) of F x F into D with the following properties: (9)
(i) {x1 + x
,
m
(ii)
(c,detof a;,yGF).
We call <. , . ) nonsingular if (lOr) and
VzeF=>*/ = 0
(101)
<s,y> = 0
VyeF=>a? = 0.
For any anti-automorphism 0 of D, the map (»!,...,rr n ), (yls ...,yn)->Xiyie + — +Xnyn0 is a nonsingular 0-bilinear form on D n x D n . Theorem 2.3. Let V be a vector space of dimension n over D; 6, an antiautomorphism of D; and (.,.), a 6-bilinear form on Vx F. Then (. , . ) is nonsingular if and only if it satisfies either of (101) or (lOr). For any linear subspace M of V let (11)
M' = {u:ue F, (x,u) = 0 for all zeilf}.
Then, if(.,.)is nonsingular, M->M' is a duality of\££(V ,D). If n^3, every duality of j£?(F,D) arises in this manner, for suitable 6 and (.,.). The pair 6', (.,.)' induces the same duality as 6,(.,.) if and only if there is a nonzero deD such that for all x, y e V, c e D, (12)
{x, yy = (x, y)d,
Co'
= d-Wd.
In particular, if D is commutative, 6' = 6. Proof. For y eV let tyeV* be defined by ty(u) = (u,y) (ueV). Then t (y->ty) is a 0-linear map of F into F*. Clearly (lOr) is equivalent to the
24
GEOMETRY
OF QUANTUM
THEORY
statement t h a t t is injective while (101) is equivalent to the statement t h a t (0) is the annihilator in V of the range of t; i.e., t h a t t is surjective. The first statement is now clear. Suppose now t h a t ( . , . ) is nonsingular. Then t is a 0-linear isomorphism and so, as M' = t~1(M°) we see t h a t M->M' is a duality of &(F,D). Conversely, if f is a duality of J^(F,D), v:M-> (^(M))0 is an isomorphism of J*?(F,D) onto ^ ( F * , D ° ) . This shows already t h a t (13)
dim(M) + dim(i(M))
= n.
Moreover, theorem 2.1 gives the existence of an anti-automorphism 6 of D and a 0-linear isomorphism t of V with V* such t h a t M° = t[£(M)] for all M in«S?(F,D). If we set {x,y) = (ty)(x) (x,ye V), it is immediate t h a t ( . , . ) is a nonsingular 0-bilinear form and (j(M) = Mf for all M. If 6' and ( . , . ) ' is another pair and ty'{x) = (x,y)' (x, y e V), 0', (.,.}' also give rise to £ if and only if t and t' induce the same isomorphism of j£?(F,D) with «^f(F*,D°). Theorem 2.1 now leads to the required result. This completes the proof of theorem 2.3. A duality £ of j£?(F,D) which is involutive is known as a polarity, this means i f = |(£(Jf)) for all M. A polarity is called isotropic if ilf <= | ( J f ) for all one-dimensional M. Theorem 2.4. Let V be a vector space of dimension n^3 over D. Then J?(V,D) admits an isotropic polarity if and only if D is commutative and dim(F) is even. In this case, if£is an isotropic polarity and 2N = dim(V), we canfind a nonsingular skew-symmetric bilinear form (.,.) such that i-(M) = M' for all M. Moreover, we can find a basis {x1,y1,...,xN,yN} for V such that (xi,xj) = (yi,yj) = 0, (xi,yj)=-(yj,xi) = Sij (l^i,j^N). In particular, if £ and £' are two isotropic polarities of3?{V',D), there is a linear automorphism a ofSf( F,D) such that £' = af cr 1 . Proof. Let f be an isotropic polarity of j£?(F,D). By theorem 2.3 there exists an anti-automorphism 6 of D and a nonsingular 0-bilinear form < . , . ) giving rise to f. Since f is isotropic (x,x) = 0 for all xe V. Replacing x b y x+y we see t h a t < . , . ) is skew symmetric. If now c e D,
=
c(x,y);
choosing x, y such t h a t (x,y) = 1 we see t h a t 0 is the identity. This means t h a t D is commutative and < . , . > is bilinear. The remaining statements are standard. Theorem 2.4 is proved. We now consider nonisotropic polarities. A 0-bilinear form ( . , . } is called symmetric if (x,y)0 = (y,x) for all x,yeV. Notice t h a t this property is relative to 6. We observe t h a t if ( . , . > is not identically zero, 6 is necessarily involutive; then the set of values of (.,.) is all of D, and
(x,y)62 = (y,xy = (x,y).
PROJECTIVE
GEOMETRIES
25
Lemma 2.5. Let dim( V) ^ 2, ( . , . ) a nonsingular d-bilinear form such that (x,x) = (x,x)d^O for some xeV. Suppose (.,.) induces a polarity. Then this polarity is nonisotropic, 9 is involutive, and (.,.) is symmetric. VrooLIfu,veV9(u9v) = OoB'V^ (D-u)'o(D-u) g (D-v)'o(v,u) Let c = (x,x). Then x $(D -x)' and dim((D •#)') = % — 1, so that
= 0.
F = (D-z)0(D-a;)'. If u,ve(D'x)', ((u,v)x — u,x + cv) = 0, so that {x-\-cv,(u,v)x — u) = 0, proving that {u,v)e — (v,u). Now (.,.) is not identically zero on (D.z)'x(D.z)', and so this already proves that 9 is involutive. The symmetry of {.,.) V xV now follows from direct computation. Lemma 2.5 is proved.
on
Theorem 2.6. Let 9 be an involutive anti-automorphism of D, V a vector space of dimension n over D, and (.,.) a nonsingular symmetric 6-bilinear form on V xV. Then the duality corresponding to ( . , . ) is a nonisotropic polarity unless D is commutative and of characteristic 2. Conversely, let n^3 and let £bea nonisotropic polarity o/J§?(F,D). Then there exists an involutive anti-automorphism 9 of D and a nonsingular symmetric 9-bilinear form ( . , . ) on V x V such that £ is induced by {.,.). If 9' and {.,.)' is another pair, £ is also induced by them if and only if there is a d^O in D such that for all x,yeV,ceD (x,y)f = (x,y)d, c6' = d~xced. In this case de — d. Proof. Given 9 and ( . , . ) , the corresponding duality is a polarity. If this were isotropic we can argue as in theorem 2.4 to conclude that D is commutative, 9 = identity, and <. , . ) is skew symmetric. As (. , . ) is symmetric it follows that D has characteristic 2. Conversely, let n^3 and let | be a nonisotropic polarity of J>?(F,D). Then £ corresponds to a pair 90, ( . , . ) . Choose x e V such that (x,x)0 = c0^0. Writing (u,v)
= (U^QCQ-1,
ce =
C^CQ-1,
we see that £ is also induced by 9, ( . , . ) , and that (x,x) = l. Lemma 2.5 shows that 6 is involutive, and {.,.) is symmetric; clearly, 0, {.,.) induce £. Suppose 9' is an anti-automorphism of D and ( . , . ) ' a symmetric 0'-bilinear form. Then, by theorem 2.3, 9', (.,.)' induce £ if and only if for some d^O in D, (x,yY = (x,y)d, (x,yeV) and cd' = d-1ced for all ceD. If (x,y) = 1, then (y,x) — 1 and so {{x,yy)e' — {y,x)' = d while {{x,yyf' = dd' = d-1ded, also. So we have dd = d.
26
GEOMETRY
OF QUANTUM
Remark. It is useful to note that the form (.,.) erty that (w,w) = 1 for some weV.
THEORY
inducing £ has the prop-
4. ORTHOCOMPLEMENTATIONS AND HILBERT SPACE STRUCTURES An orthocomplementation ofJ?( F,D) is a polarity £ such that M n £(M) = 0 for all M, i.e., an inclusion reversing bijection £ of J^(F,D) with itself such that g(i;(M)) = i f and M n f (ilf) = 0 for all M. It is easy to see that a polarity is an orthocomplementation if and only if (D-x) n £(D-a:) = 0 for all xe F. The preceding discussion leads to Theorem 2.7 (Birkhoff-von Neumann [1]). Le£ w = dim(F)^3 and £ an orthocomplementation of JS?(F,D). TAew £Aere is aw- involutive antiautomorphism Oof J) and a nonsingular symmetric O-bilinear form (.,.) on V x V such that 6 and {.,.) induce £. We have (x,x) = 0ox = 0, and {.,.) can be chosen so that (w,w) = 1 for some weV. Let a symmetric 0-bilinear form < . , . ) be called definite if (14)
(i) (x,x) = 0 o x = 0, v \ >/ (ii) (w,w) = 1 for some
weV.
A definite form is necessarily nonsingular. Assume now that D is one of R, C, or H. (a) D = R. Then the definite forms are precisely the positive definite quadratic forms. So the definite form (.,.) of theorem 2.7 defines a scalar product and f is the orthocomplementation in the corresponding Hilbert space structure. (b) D = C. Here we must assume that 6 is continuous or measurable (analytically well behaved). Since complex bilinear forms are not definite, 6 must be the complex conjugation.The definite forms are thus the Hermitian positive definite forms, and in theorem 2.7, £ is the orthocomplementation defined bv the Hilbert space structure corresponding to the scalar product (c) D = H, the quaternions. A scalar product for a vector space V over H is a *-bilinear form on F x F such that: (i) for any x-+V, (x,x} is in the center of H and coincides with a nonnegative real number; (ii) (x)x) = 0 if and only if x = 0; if \\x\\ = -f- (x,x)%, || • || is a norm for V. If V is complete under ||-1|, it is called a Hilbert space; this is the case when dim(F)
PROJECTIVE
GEOMETRIES
27
ueVbe such that (u,u) is a unit quaternion, say a, and let ueg (H• u) be such that (u,u) is also a unit quaternion, say b; of course V = H • u © £(H *u) • From (u,u)6 = <w,^) we get, after a small calculation, <->«< = (_%
r
c)
where c e R , |c2| + |r| 2 = l. Hence tr(aq0) = 0. Similarly, tr(bq0) = 0. So aq0 and — bq0 are in the same class, proving that qiaq0q1*q0~1 + b = 0 for some unit quaternion qv But then (x,x) = 0 for x = qxu + v. So, q 0 2 = +1 is the only possibility. Since det(q0) = 1, we must have q0 = + 1. So 6 must be conjugation and (.,.) is *-bilinear. In particular, (u,u) is real for all ueV, and >0 for ue V\(0). V thus becomes a Hilbert space under < . , . ) , and £ is the associated orthocomplementation. We thus have Theorem 2.8. Let D = R or H and let V be a vector space of dimension n^3 over D. / / £ is any orthocomplementation in J?(V,D), there is a scalar product for (.,.) converting V into a Hilbert space such that £ is the corresponding orthocomplementation. If D = C and f is regular in the sense that the anti-automorphism of C defined by £ is continuous, then it is the complex conjugation, and £ is induced by a Hermitian scalar product {.,.) that converts V into a Hilbert space over C. Finally {.,.) is determined by £ up to multiplication by a positive real number. We shall now assume that D is one of R, C, or H, and that V is a Hilbert space of dimension n ^ 3 over D, with scalar product ( . , . ) . Theorem 2.9. Suppose £ is an automorphism of J?(V,D) which preserves the orthocomplementation of V. Then there exists a semilinear automorphism Lof V inducing £ such that: (i)
if D = R or H, L is linear, and for D = C, L is either linear or conjugate linear;
(ii)
for all
x,yeV,
(15)
(Lx,Ly) = (x,y) if L is linear, and
(16)
(Lx,Ly)=(y,x) ifJ) = C and L is conjugate linear. In particular, L is an isometry of V.
Proof. We begin with the following observation.Let 9? be an automorphism of D and Lx a ^-linear isomorphism of V onto itself such that the induced automorphism of ^(V,D) preserves the orthocomplementation of V; then, for some d^0 in D, we have, for all x,yeV, (17)
(L1x>L$) = {x,y)<ed
28
GEOMETRY OF QUANTUM
THEORY
Indeed, (x,y) = 0 if and only if {Lxx,Lxy) = 0; so, if we define
(x,y)~ = (LiXtLtfyv1) (x,yeV)
and 0 = y i ^
where 6 is the anti-automorphism corresponding to ( . , . ) , then the pair ifj, (.,.)~ also induces the orthocomplementation in J£( V,D). Theorem 2.3 now leads to (17). This said, let us first suppose that f is induced by a linear automorphism Lx of V. Then (17) becomes (L1xiL1y) = (x,y)d. This shows that d is a real positive number, and we get the required result on taking L — d~^Lv If £ is not linear, then D = C and there is an automorphism
5. COORDINATES IN PROJECTIVE AND GENERALIZED GEOMETRIES The study of the projective geometries =£?(F,D) gains significance in view of the fundamental theorem of classical geometry which asserts that these are the only geometries, at least when the rank is ^ 4. When the rank is 3, i.e., when we are dealing with a projective plane, the geometries j£?(F,D) are characterized by the property of being Desarguesian, namely, that the classical theorem of Desargues is true for them. Of course a projective plane imbedded in a geometry of higher rank is always Desarguesian; this can be proved directly and in a simple manner. However, as was mentioned at the beginning of the chapter, the geometries ££{ F,D) and, more generally, lattices of finite rank, are not adequate to serve as models for the proposition calculi of complex quantum systems. We shall therefore enlarge the scope of our discussion by including a class of partially ordered sets of infinite rank, and proving for them a coordinization theorem that will include the classical result. The objects that we shall study will be called generalized geometries; and the main result concerning them is that they are isomorphic to the partially ordered sets of finite dimensional subspaces of possibly infinite dimensional vector spaces over division rings.
PROJECTIVE
GEOMETRIES
29
Let ££ be a set with partial ordering <; we do not assume that ££ has a unit element. 3? is called a generalized geometry if the following conditions are satisfied: (i) for any finite subset F of J^, \/a&Fa, /\a£Fa, exist in <£?, (18) (ii) if a e 3? and a ^ 0, j£?[0,a] is a geometry. Thus J^7 is a geometry if and only if it has a unit element. If D is a division ring and V a not necessarily finite dimensional vector space over D, the partially ordered set J^(F,D) of finite dimensional subspaces of V is a generalized geometry, which is a geometry if and only if dim( V) < oo. If & is a generalized geometry, ££ has points and every element of ££ is a sum of points. The dimension functions of the J?[0,a] are mutually compatible and define a dimension function on all of J§P. We write d i m ^ ) for sup aGif (dim(a)). We can obviously speak of lines, planes, etc., in j£f. In the rest of this section we shall sketch a proof that any generalized geometry is isomorphic to some J2?( F,D), provided dim(oSf) ^ 4. Our proof is a straightforward variant of the classical one; but we have decided to include a sketch since it is not easy to locate it in the literature in the precise form we need. The reader who wants more details on the part dealing with Desarguesian planes should consult Seidenberg [1] and Heyting [1] (see the Notes at the end of this chapter). Let S£ be a generalized geometry with d i m ^ ) ^ 4. Then all the planes in ££ are Desarguesian. The classical method of constructing the division rings associated to 3? is as follows. We take a line t in ££ and a plane a in 5f containing t, and we choose three distinct points 0, E, W on t. To this data we shall associate a division ring D whose elements will be the points of the line t that are different from W. The operations of addition and multiplication in D will be defined with the help of geometrical constructions (in the plane a) that imitate the familiar constructions in the Euclidean plane. Addition. Let AGD; then we define the map aA'B-+B + A ofD with itself in the following manner. In the plane a we take distinct lines q, r through W9 different from t, and distinct points X, Y on q, different from W9 such that if n(X; r,t) (resp. n(Y; t,r) is the perspectivity from r to t with center X (resp. from t to r with center Y), then n(X; r,t) n(Y; t,r) takes 0 to A; then we define (Fig. 1) aA =
7r(X;r,t)7T(Y;t,r).
If we view q as the "line at infinity" then this construction resembles the usual Euclidean one (Fig. 1(a)).
30
GEOMETRY
OF QUANTUM
FIG.
1(a)
THEORY
PROJECTIVE
GEOMETRIES
31
Multiplication. Given A e D we define right multiplication TYIA by A; if A=0, MAB = 0 by definition for all B and so we take A^O. We take in the plane a a line g through W different from t, and a line p through 0, different from t. We choose distinct points X, Y on q different from W such that n(X; p,t)n{Y; t,p) takes E to A (see Fig. 2); then m^ = n(X;pft)7r(Y;
t,p).
If we view g as the "line at infinity", this resembles the usual construction in the Euclidean plane (Fig. 2(a)). Of course we could have equally well taken this as the definition of A -B; our choice is dictated by the requirement of getting isomorphisms with the geometry of subspaces of a left vector space. Although these definitions involve the choices of auxiliary lines and points, the Desarguesian nature of the plane can be used to prove that addition and multiplication are well defined and convert D into a division ring with 0 and E as its null and unit elements. If t, 0, E, W are fixed, this division ring is independent of the choice of the plane a that contains t; for, if 6 is a different plane containing t, and X is a point in the three-dimensional space avb which does not lie on a or 6, the perspectivity from X gives an
FIG.
2
32
GEOMETRY OF QUANTUM
FIG.
THEORY
2(a)
isomorphism of ^f[0,a] with J£[0,b] which is the identity on t — a/\b, and it is clear that addition and multiplication on D are the same whether we use a or b. The division ring is also unchanged up to isomorphism if we change t 0y E, W. More precisely, let t' be another line in the plane a, and O', E', W, three distinct points on t'. Then any projectivity in the plane that takes t to t' and 0, E, W, respectively to 0', E', W', is an isomorphism of the associated division rings; this follows from the well-known fact that any projectivity of t with t' extends to an automorphism of the plane j£?[0,a]. We remark finally that D is commutative if and only if the geometry of J&?[0,a] is Pappian for any plane a in jSf, i.e., the theorem of Pappus is valid in o£?[0,a] for any plane a. The first step in the coordinatization of ££ is to define" afnne " coordinates for all points of JSf that do not "lie at infinity". If S£ is a geometry we can choose an arbitrary hyperplane to be at infinity. For a generalized geometry we use a simple modification of this procedure. To obtain an intuitive picture of the definition of coordinates the reader is asked to picture everything in a Euclidean context, interpreting incidences at infinity in terms of parallellism. Let us call a set of points of j£? independent if every finite subset of the set is independent. If {P3} (j e J) is an independent set of points, then we have a map u (K->u(K)) from finite subsets of J into j£? defined by (19)
u(K)=
V Pi9
u(4>)=0.
PROJECTIVE
GEOMETRIES
33
I t is easy to verify t h a t (20)
u(Kt
u K2) = u{Kx) v u(K2),
u(K± n K2) = u{Kx) A tt(X2).
An independent set C of points of J§? is called a 6cms if it is maximal. Every independent set is contained in a basis in view of Zorn's lemma. If C is a basis of J£? and P is a point of if, it follows from the maximality of C t h a t there is a finite set F <= (7 such t h a t P < \/QeFQ; hence for any ae^, there is a finite set (r c 0 with a <\/Qe0Q. Let 0 be a point of if. By a frame a t 0 we mean a pair (#,{PJ ; . G t / ) such t h a t {0,{P,} yet/ } is a basis for if. There are frames at any point of i^7. From now on we fix a frame (0,{Pj}jeJ) a t 0. An element a e i f is said to lie at infinity if there is a finite set K^J such t h a t adim(u(K)) so t h a t avu(K)
=
Ovu(K);
if a^ = aAu(K), then dim(a 00 ) = dim(a) — 1, from which the characterizations of a^ foliow immediately. The uniqueness of a^ is obvious and a^ is called the hyperplane at inifinity in a. I n particular, if P is a point not in ifoo a n d K ^ J a finite set, then for a = P\/u(K), u(K) is the hyperplane at infinity. Every line not in ££^ has a unique point at infinity on it, every plane not in ^^ has a unique line at infinity on it, and so on. Let rrij be the line 0\fP3 (jsJ). For each j e J we choose a point Ej in m3distinct from O and P3. As explained earlier we convert the set of points of nij different from Fj into a division ring Dy with O and E3 as its null and unit elements, respectively. The division rings D^ are of course mutually isomorphic b u t we can do better and introduce a " c o m p a t i b l e " system of isomorphisms. Let j , keJ, j ^ h. The lines E3vEk and P3\/Pk, being in t h e plane m3\imk, meet at some point, say P3k; this is the point at infinity of EjVE^ and the perspectivity from P3k, of the line mk onto the line m3, fixes 0 and takes Ek (resp. Pk) to Ej (resp. P3). I t thus induces an isomorphism of Dfc onto D ; , which we denote by (21)
d3k:Dk^D3,
6jk = n(P3k;
mk,m3).
We define d33 to be the identity. Then the "compatibility" of the 6jk may be formulated as the following lemma. Lemma 2.10. Ifj, (22)
k,reJ,
then &rj ° Ojk = 6rk-
34
GEOMETRY
OF QUANTUM
THEORY
Proof. From (21) it is immediate t h a t 0kj = 6jk~ . Suppose now t h a t j , Jc, r are three distinct indices from J, XeBk, Y = Ojk(X), Z = 6rj(Y). W e must prove t h a t Z = 6rk(X). W e m a y assume X^O, X^Ek. Desargue's theorem applied t o the triangles X YZ and EkEjEr which are in perspective from 0, gives a line t on which t h e points (X V Y) A (Ek V Et) = Pik, and
(YvZ)A
(XvZ)A(EkvEr)
(E, V Er) = Pjr,
=Q
lie. Obviously t is a t infinity a n d so Q = P rfc , proving t h a t Z = 6rk(X). From (22) we see t h a t there is a division ring D a n d isomorphisms (23)
>,-: D ^ > D ,
such t h a t t h e diagrams
&. D
(24)
^
D
'
are commutative. W e write 0 a n d E for t h e null a n d unit elements of D. Let ty be the set of points of «£? which do not lie a t infinity. Let us denote b y W the vector space of all functions with finite supports on J with values in D. W is a left vector space over D in t h e usual fashion. Suppose now P ety. We can choose a finite setK^J such t h a t P < Ovu{K). For any j e J, u(K-{j}) is t h e hyperplane a t infinity on Pyu(K-{j}) a n d so t h e latter does n o t contain P y ; thus MjKP =
mjA(Pvu(K-{j}))
is a n element of D,. If Kx <^K is such t h a t P
MjP = mj A ( P V !*(#-{j}))
( P < 0 V w(JC)).
p
Transferring i f , t o D now gives t h e coordinates of P , i.e., let fp be t h e function from J t o D defined b y (26)
tpU)=vr1(M,p)
(jeJ).
Lemma 2.11. For any P e ^ , fp is in W; fo = 0, the origin ofW; and P-> fp is a bijection ofty with W . Proof. If P
U(j'-{jr})
=0
PROJECTIVE
GEOMETRIES
35
by (20). So, P e ty and it is easy to verify that fP = g. If P' e $ is also such that fP, = g, it is clear that P' < Ovu(J'), and hence, for 1 < r < 8,
Qr=MJrr'
P
M^^MfJeK.
Fixj,keK,j^k.
The triangles PMjpMkp and QM^Mjfi
are in perspective from O and so have an axis of perspectivity. This is a line at infinity (cf. the remark above) and so (M,p V Mkp) A (Mfl vMkQ) = T is a point at infinity, which lies on the line Pj\/Pk because all four points are in the plane a = 0vP,vP fc - If n denotes the perspectivity, in the plane a = OvPj\/Pkiofmjto mk with center T, we can use the definition of multiplication in-D, to conclude that if Nj = (6jkon) (Ej), then for any UeDjt (6jko7T)(U)=U'Nj. Thus djk(MkQ) = MjQ-Nj and djk{Mkp) = Mjp>Nii proving that 0Jk(Mk) = Mj. For the converse, suppose Q ^ 0 is in ^ with fQ = M • ip. Assume as before, to avoid trivialities, that K has at least two elements. Choose a>j0eK and let Q' be the point on the line 0\jP such that the lines Q'vMJoQ and PyMJop meet at infinity. Obviously, Q'ety, Q'j-O, and Jf, o «' = Jfy G. Since
GEOMETRY
36
fQ/ = j|f'-f P for some M'eD, Le.,Q = Q'.
OF QUANTUM
THEORY
we have
Lemma 2.13 ("parallelogram law of addition"). Let 0' e ^S, 0 V 0, and let t' be a line through 0' not containing 0. Let U, R be the points at infinity on the lines OyO' and t', respectively, and let t be the line OyR. If PeSfi lies on tand (27)
P' =
then
(UyP)At',
fp/ = fp-ffo-.
Conversely, suppose P' e ^} lies ont'\ then fp' — fcr =fp for some Pont, P and P' are related by (27).
and
Proof. For the first part we must show that Mf^Mf + Mfi* for all jeK where X g J i s a finite set such that O', P, P' are all <0\lu(R). If U
FIG.
3
As 0" < 0\ju{K - {j}) we have U' < u(K - {j}) and so Mf
= Mjp",
Mf
=
M?K
It is clear from Fig. 3 that with respect to addition on the line t (minus R), Pm = P + 01\ so, if t = mh Mf'^Mf + MpK Ift^mj, let T be the common
PROJECTIVE
GEOMETRIES 0
37 P
point at infinity on the lines O^siM, *, PvMf, P"yMj ". Consider the plane tymj9 and in it, the perspectivity from t to m5 with center T. As this takes Ov P, P" to if,0*, M$p, and Mf", respectively, we get Mf^Mf
+ M,0*.
For the converse let g = fp'-fo-. Define P = tA(UvP'). Then P and P ' are related by (27) and so g = fp. Lemma 2.14. Let P x ', P 2 ', P 3 'feeJAree distinct points inSfi. Then they are collinear if and only if for some A^O inD we have (28)
tps^A-tps + iE-Aytp,.
Proof. Assume first that the P{ are on a line t'. Then (28) follows from lemma 2.12 if t' contains 0. If 0 <£ t\ we can apply lemma 2.13 with 0' = P 2 'If P x and P 3 are the points on the line 0\lE corresponding to PX' and P 3 ', we have fp3, = fp2, -|- fp3> fPi, = fp2, -f fPi) while lemma 2.12 gives a nonzero ^. eD with fp3=A -fp,. This implies (28). For the converse, assume (28) and let t' be the line P^yP^. We may suppose that Oj;tf, as the case 0
gp(i) = g ( «
(29) and (30)
y(P) = D-gp.
Suppose P is a point at infinity on ££. We choose a point Q on the line 0\lP such that # ^ P , Q ^ 0 and define g P , Q e \ , y(P) eS£' by (31)
g
3 eJ, 3 = 00,
and (32)
y(P) = D- &P.Q-
Lemma 2.12 shows that y(P) in (32) is independent of the choice of Q and so is well defined.
38
GEOMETRY
OF QUANTUM
THEORY
Lemma 2.15. The map P -> y(P) is a bijection, of the set of all points of J£? with the set of all points of j£?', which preserves collinearity. Proof. The bijective nature of the map P->y(P) is a straightforward consequence of the definitions and we sketch the argument for proving that y preserves collinearity. Fix three points of j£?, say Pf (i— 1,2,3); we must show that they are collinear in j£? if and only if y(Pi) are collinear in -Sf". Case 1. P{• G^} for all i. In this case lemma 2.14 gives what we want. Case 2. Exactly one of the Piy say P 3 , lies at infinity. Let Q be a point on OvP 3 , Q¥^0,Q^Pz. It is trivial that y{Px), y(P2), a n d 7(^3) a r e collinear if and only if for some M eD, £Q = M -(fp2 — fpj. Lemmas 2.12 and 2.13 show that this is precisely the condition for Q to lie on the line joining 0 to the point at infinity on Px\/P2, i.e., for P 3 to lie on P1yP2. Case 3. P 2 and P 3 are both at infinity. Suppose P l 5 P 2 , P 3 are collinear, so that P1 is also at infinity. Chooce Q*G^P> Qi^O (i = 1,2) such that: (i) Qi
M,N^O
mD}
By lemma 2.12 we can find points Q/ ^ 0, such that Then *b3' = ^ ' —f^x'. If X is the point at infinity on Qi\zQ2', lemma 2.13 and this relation tell us that Q3 is on OyX. So X = P 3 , i.e., the P z (^ = 1,2,3) are collinear. This proves the lemma. These lemmas imply the following result which is the fundamental theorem on coordinatization of geometries. Theorem 2.16. Let & be a generalized geometry with dim(J^) ^ 4. Then there exists a division ring D and a left vector space V over D such that 3? is isomorphic to the generalized geometry of all finite dimensional subspaces of V. V is finite dimensional if and only if 3? is a geometry.
NOTES ON CHAPTER II 1. For more detail on quaternionic geometry see Chapter I of Che valley [1]. The identification of H with C2 in Section 2 gives rise to a bijection
PROJECTIVE n
GEOMETRIES
39
2n
H ^>C which is C-linear and associates q = (qv ...,#«) to x=^{x1,...ix2n) where qk = x]^-\-xn+k}i. We may regard H as a Hilbert space with scalar product (q,r)=21
S
(Zkyn+k-yk%n+k),
where x = (xv ...,#2«) and y = (yv ...,y2n). 2. The definitions in Section 5 are classical. For any line t in a Desarguesian plane n let G be the set of all projectivities of t with itself which are compositions of two perspectivities: from t to a second line s followed by a perspectivity from s to t. The line s, as well as the centers of perspectivity, are arbitrary. Assume that s^t and that the two centers, X, Y are distinct. Then the projectivity has one or two fixed points according as the line Xv Y passes through sAt or not. In the former case the projectivity is called special and the point W, which is on s, t and XV Y, is the unique fixed point and is called its canonical fixed point. If 0 = sAt and W = (XyY)At are distinct, they are the two fixed points and are respectively called the first and second canonical fixed points of the projectivity. For any W on t (resp. distinct 0, W, on £), GStw (resp. Gg>o, w) is the set consisting of the identity and the special (resp. general) projectivities of t with itself, with W as canonical fixed point (resp. 0 and W as the first and second canonical fixed points). The key fact, which follows from the Desarguesian nature of 77, is that if aeGStw (resp. aeGgio, w), a is determined by its action on one point oit other than W (resp. its action on one point of t other than 0 and W), independently of the choices of s, X, Y. It is immediate from this that the operators a A and rriA are well defined. The possibility of using different s, X, Y to define the same projectivity then allows one to prove that GSt w and Gg>0>w are groups, the former being abelian. This will prove that + converts D into an abelian group with O as its null element, and • converts D \ 0 into a group with E as unit element. Right multiplication is of course in Gg, o. w\ left multiplication is also a projectivity of t with itself fixing 0 and W, which however is a composition of three perspectivities. Distributivity laws are special cases of the more general result that if a is any projectivity of t fixing 0 and W, a is an automorphism of the additive group of D. This is an immediate consequence of the extension principle stating that any projectivity from one line of n to another can be extended to an automorphism of 7T. This principle, in the case of a perspectivity, is an easy application of the following lemma: Lemma. Let C be a point of n; t, t' distinct lines through C; and A, A' distinct points ont', different from C. Then there is a unique automorphism a
40
GEOMETRY
OF QUANTUM
THEORY
of the plane n such that: (i) a fixes each point of t\ (ii) a(A) = A'; (iii) for any point X not on t} a(X) is on the line joining C to X. Proof. Consider first a line u through C distinct from t and two points X, X' on u, distinct and different from C. Let P(u) be the set of points of the plane not on u. We define dx, x' to be the m a p P(u) -> P(u) such t h a t for a n y B in P(u), B' = dx,x'(B) is the point where the line KvX' meets the line CvB, K being the point where the line ByX meets t (draw a figure). I t is easy to see t h a t dx, x' is bijective and its fixed points are precisely those on t (except for C). If v is the line ByB' for some B in P(u), 6B, B' is well defined on P(v), and a suitable application of Desargues' theorem shows t h a t QB,B' = QX,X' on P(u) nP(v). We now define a as follows. For a n y point P<^,a(P) = P;ifPH:^andP^r,a(P) = ^,^(P);ifP
PROJECTIVE
GEOMETRIES
41
of principal right ideals to form a complemented partially ordered set; and finally, that in this case the set of principal right ideals is a complemented modular lattice (so that principal = finitely generated). Of course in all of this "right" can be replaced by "left". The lattice will be irreducible if and only if R is indecomposable in the sense that R does not admit any decomposition of the form R = 7^©R 2 where R^O are two-sided ideals; and this is equivalent to requiring that the center of R is a field. Finally, if a ring R has the property that its principal right (or left) ideals satisfy the ascending and descending chain conditions, then R is regular if and only if it is semisimple, thus tying in the classical coordinatization theory with the Wedderburn theory. It was also established by von Neumann that dualities of 3? correspond naturally to anti-automorphisms of R, and that orthocomplementations correspond to involutive anti-automorphisms x->x* which are definite in the sense that x*x=0oa; = 0. Rings of bounded operators in a complex Hilbert space are regular only when they are finite dimensional; however, the ring of not necessarily bounded operators affiliated (in the sense of Murray and von Neumann) to a type IIX factor is a regular ring, and is in fact the regular ring whose principal right ideals form the lattice isomorphic to the lattice of projections of the factor. Occupying a special place among the complemented modular lattices are the continuous geometries also discovered by von Neumann. He characterized completely the associated regular rings. The lattices of projections of type IIX factors are orthocomplemented continuous geometries. Roughly speaking the axioms for an irreducible complemented modular lattice to be a continuous geometry require the completeness of the lattice and the continuity of the lattice operations, and lead to the existence of a dimension function with values in the unit interval [0,1]. Let D be a division ring, Vk and Vmk (m an integer ^ 2) be left vector spaces of dimensions k and mk, respectively, over D. We then have a map M-+M<$>...®M (ra summands) of ^(VkiD) into J^(Fmfc,D). Using a succession of such mappings (with changing "ratios" m), von Neumann constructed continuous geometries associated with D which, when D = C, were not isomorphic to the projection lattices of type IIX factors. These scattered remarks should convey a glimpse of the beautiful generalization of classical projective geometry that von Neumann erected. The reader who wants to get a sharper perspective on these things should begin with a study of von Neumann's book (Continuous Geometry, Princeton University Press, Princeton, New Jersey, 1960, which has annotations by I. Halperin that illuminate many aspects of the theory and contain references to further literature), and his papers on the subject that appear in volume IV of his Collected Works (6 volumes, Pergamon, 1961).
CHAPTER III THE LOGIC OF A QUANTUM MECHANICAL SYSTEM 1. LOGICS Let © be any quantum mechanical system. We have seen that one can associate with <& the partially ordered set ££(!&) of all experimentally verifiable propositions concerning <&. The partial ordering is that induced by the implication relation. Moreover, the map, which associates with any element of ^ ( G ) its negation, behaves very much like an orthocomplementation. This leads one to introduce axiomatically a class of partially ordered sets and to study their properties. We call these systems logics. The basic assumption of modern quantum theory can then be described (if we anticipate some terminology) by saying that ££(<&) is a standard logic. Let j£? be a lattice under a partial ordering < . By an orthocomplementation of j£? we mean a mapping a-^a1
(_L) :
of & into itself such that (i) J_ is one-one and maps «£? onto itself, (ii) a < b implies b1 < a1, (1) (hi) a 1 1 = a for all a, (iv) a A a 1 = 0 for all a, (v) a V a1 = 1 for all a. We note that the relations 0 1 = 1 and 1 1 = 0 are easy consequences of (ii). Moreover it follows easily from (iii) that J_ is one-one and onto, so that (iii) => (i). Finally, we observe that (ii), (iii), and (iv) imply (v). In fact, let a e 3? and b = a\jaL. Since both a and a1 are
(i) for any countably Vn «n and / \ n an (ii) if al9 a2e J? and that b
infinite sequence al9 a2, - - - of elements of JSf, exist in &, a1
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
43
Before proceeding further we make the observation t h a t if a1
SP[a,b] = {c : c G J^,
a < c < b}.
Then under the partial ordering inherited from <j£f, j£?[0,6] becomes a lattice, in which countable unions and intersections exist and whose zero element is 0 and unit element is b. If we define, for any x in ^ [ 0 , 6 ] , its orthogonal complement x' by x/ = x1Ab, then it can be shown t h a t j£?[0,6] equipped with the orthocomplementation x —> x' is a logic (cf. corollary 3.3). The central assumption t h a t one makes in any quantum mechanical application is t h a t the set of experimentally verifiable propositions is a logic. The only thing t h a t m a y be open to serious question in this is assumption (i) of (2) which forces any two elements of S£ to have a lattice sum; the others m a y be regarded as technical necessities. We can offer no really convincing phenomenological argument to support this (cf. BirkhofT-von Neumann [1]). If we omit the assumption t h a t 3? is a lattice the axioms become so weak t h a t it is very difficult to avoid pathology in any mathematical discussion. We point out also t h a t every calculation in quantum mechanics is based on assumptions which not only imply t h a t the set of experimentally verifiable propositions is a logic but in fact a very special one. We shall now make a few general remarks. If 3?^ and ££2 are two logics, an injection of 3?x into j£?2 is a m a p / ( a - > / ( « ) ) of 3?x into 3?2 such t h a t ( i ) / i s one-one,/(0) = 0 , / ( l ) = l ; (ii) if al9 a2, • • • is a n y at most countable sequence of elements of &l9 f(\/nan) = V n / K ) and f(/\an) = A / K ) ; (iii) f(a1)=f(a)1 for all a e S£x. An isomorphism of S£x on S£2 is an injection which maps <£x onto J^ 2 - If <^i = o^ 2 , isomorphisms are called automorphisms. The set of all automorphisms of a logic J£ is a group under composition; we denote it by Aut(J^). If ££x and ££ are two logics, we say t h a t ££x is a sublogic of S£ if (i) ^ c ^ and (ii) the identity m a p of =£?! into ££ is an injection.
GEOMETRY
u
OF QUANTUM
THEORY
Any Boolean cr-algebra is a logic provided we define, for any element a, a1 to be the complement of a. These logics are of course not very interesting from the point of view of quantum mechanics. To obtain a more typical example, let us consider a Hilbert space £F over the real, complex or quaternionic division rings. We denote by J^(Jf) the collection of all closed linear manifolds of 3tf. If we now define < to mean set inclusion, and J_ to mean the usual operation of orthogonal complementation in ^ ( ^ f 7 ) , then it can easily be verified t h a t ^f (Jf) is a logic. The isomorphism class of ^(Jti?) depends only on the field D of definition of 34? and the dimension of ^f \ A logic ££ is said to be standard if it is isomorphic to the logic j£?(c^) of a separable infinite dimensional Hilbert space over one of t h e three division rings R, C, or H. Modern quantum theory works with the assumption t h a t the logic of any atomic system is, if not standard, at least a sublogic of a standard logic. We shall make a deep study of standard logics in Chapter IV. We shall also describe there a number of logics t h a t are sublogics of standard logics, b u t are not standard themselves. We shall now proceed to derive a few simple consequences of the axioms denning a logic. Lemma 3.1. Let ££ be a logic and ava2, ...a sequence of elements of ££. Ifb e S£ and b\_anfor all n, then &JL.Vn an- Moreover we have the identities:
(V OnY = A «n\
Finally,
if ax
c
Proof. Since b
b
(a V 6) A c = (a A c) V b.
Proof. Write d = (aAc)vb
and e = (avb) Ac. I t is obvious t h a t
d<e.
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
45
Hence, by (ii) of (2), there exists an element g of J? such t h a t d_\_g, and dvg = e. Since a±b and the element g satisfies the relations gx' defined above is a logic. Proof. We must verify (ii), (hi), (iv) of (1) and (ii) of (2). (ii), (iv) of (1) are obvious. We shall now prove (hi) of (1). Let x
2. OBSERVABLES We shall introduce in this section the concept of an observable associated with a logic. The formal definition is given in the next paragraph, but we shall first t r y to motivate the definition. Suppose t h a t (5 is a physical system, S£(&) its logic, and £ a physical quantity or observable. Then in any experiment which an observer performs, the statements t h a t can be made concerning £ are of the type which asserts t h a t the value of £ lies in some set E of real numbers. I t is natural and harmless to require
46
GEOMETRY
OF QUANTUM
THEORY
that the sets E be Borel. If we denote by x(E) the statement that the value of £ lies in the Borel set E^R1, then one has a mapping x : E -> x(E) of ^(R1) into J^(©). We shall regard two observables as "identical" if and only if the corresponding mappings are the same. If / is a real valued Borel function of a real variable, then we mean by / o £ the observable whose value is f(r) whenever £ takes the value r; to this observable clearly corresponds the mapping E -> x(f~1(E)) (cf. Chapter I, sections 2 and 4). Motivated by these remarks we introduce the following definitions. Let j£? be a logic. An observable associated with ££ is a mapping x : E -> x(E) of the o--algebra ^(R1) of Borel sets of the real line into ££ such that (6)
(i) a.( 0 ) ==o,a?(J81) = 1, (ii) if E, F e ^(R1) and E n ^ = 0 , x(E) _[_ x(F), (hi) if 2£1? 2£ 2 >' ' ' i s a sequence of mutually disjoint Borel sets in R1,
x(y j?n) = y ^ j . We write 0 or 0(j£?) for the set of all observables associated with J£f. We remark that the properties (i) through (iii) are natural if we want to interpret x(E) as the statement that the value of the observable £ lies in E. If X is any set and j£? is an arbitrary cr-algebra of subsets of X, then j£? is obviously a logic provided we define < to be set inclusion and _[_ to be set complementation. Theorem 1.4 asserts that the set of observables is in canonical one-one correspondence with the algebra of all real valued J^-measurable functions on X; if x(E —> x(E)) is any observable, then the corresponding function / on X is the unique ^-measurable one such that f~1(E) = x(E) for all E e ^(R1). Suppose that x is an observable and / a real valued Borel function of a real variable. Then (7)
fox:E^x{f-\E))
is also an observable. Lemma 3.4. Let S£ be a logic and x an observable associated with £\ Then for any sequence El9 E2, • • • of Borel sets (8)
U
/
»
*(n En) = A *{En).
THE LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
47
Suppose further that fx and f2 are two real valued Borel functions of a real variable t. Let f± of2 denote the function t ->/i(/ 2 W)- Then (9)
(/i°/2)°z=/i°(/2°*).
Proof. Since x(En)±x(R1-En) and x(En) vx(Rl-En) = l, it follows 1 1 that x(R — En)=x(En) . Consequently, in view of the equation (4), the second equation in (8) will follow if we prove the first. To prove the first, we begin by observing that if A c= B, then x(A) <x(B); in fact, x(B) = x(A) v
x(B-A).
This said, let E = \Jn En. Then x(En) < x(E) for all n so that \Jn x(En) < x(E). On the other hand, let F1 = E1 and for n>l, let Fn = En-\Jj
<
n
\/z(En). n
This completes the proof of (8). Equation (9) is trivial. Let x be an observable associated with ££. A real number A is said to be a strict value of x if x({\}) ^ 0 . x is said to be discrete if there exists a countable set C = {cu c2, • • •} of real numbers such that x(C) = 1. a; is said to be constant if there exists a real number a such that x({a}) = l; we shall then write x — a. x is said to be bounded if there exists a compact set K such that x(K) = 1. If we define a(x) by (10)
a{x)=
p|
0,
C closed, x(C)= 1
then G(X) is a closed set called the spectrum of x. Since the topology of R1 satisfies the second countability axiom, there exists a sequence of closed sets C1,C2i--' such that x(Cn) = l for all n and a(x) = P| n C n . Since x(Rx-Cn) = 0 for nil n, x(Ri-*(x))
=
x(\J(Ri-Cn)}
=
\/x(Ri-Cn) n
= o, which proves that x(o(x)) = 1. or(#) is thus the smallest closed set C such that x(C) = l. The numbers A GCT(#)are called the spectral values of #. A strict value is a spectral value but the converse is not true in general, x is bounded if and only if G(X) is a compact set. A e a(x) if and only if for any open set U containing A, x(U)^0. Since the defining axioms of a logic are somewhat weak, it is not reasonable to hope for any incisive description of the set of all observables associated with ££\ In the next chapter we shall examine this question
48
GEOMETRY
OF QUANTUM
THEORY
for logics which are associated with Hilbert spaces. For the moment we confine ourselves to a remark on discrete observables. Suppose ££ is an arbitrary logic. Let # be a discrete observable and {cl5 c2, • • •} the set of its strict values. Then x{{c$) and #({c;}) are orthogonal whenever i^j and
V*({c,}) = l. i
Conversely, let {Ci}ieD be an at most countably infinite set of distinct real numbers and {at}ieD a family of mutually orthogonal elements of j£? (with the same indexing set D) such that
v «f = i. ieD
Then there exists a unique discrete observable x associated with £? such that x({ci})=ai for all i e D. In fact, we have only to define, for E e^(R1), x(E) =
V «iiiCieE
It is easy to verify that x is an observable. It is clear that x({ci}) = ai for all ieD and that x is discrete. The uniqueness of x is obvious. Even though many interesting and important observables are discrete, there are quantum observables which are not discrete. The construction of observables which are not discrete is, however, quite complicated. We shall indicate how it can be done in the next chapter for logics which are associated with Hilbert spaces. If j£? is an orthocomplemented geometry of finite dimension N, then it is easy to see that every observable associated with j£? is discrete and has at most N values. Clearly, such logics are not suitable models to represent the experimentally verifiable propositions of complicated atomic systems.
3. STATES If © is a classical mechanical system, the concept of a state of © is so defined (cf. Chapter I) that if © is in a state p, and £ is any observable, then £ has a value in p. This leads one to postulate that the states are points of the phase space and observables are real valued functions on it. In quantum theory this has been rejected as irreconcilable with the available experimental facts regarding the behavior of atomic systems when subjected to refined microscopic observations (cf. von Neumann [1], Heisenberg [1]). The modern approach to atomic physics is based on the principle that in a given state of the system, an observable has only a probability distribution of values and in general no sharply defined value; and no matter how carefully the state is prepared, there will be some observables whose values are distributed according to probability distributions
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
49
having arbitrarily high variance. Accordingly, if (B is in a state or, one can associate, with each observable x, a probability distribution, say Px , on the line, which will be interpreted as the probability distribution of the values of x when the system is in the state a. If a; is a discrete observable, and cl9 c 2 , • • • are its possible values, then Px would have mass concentrated on the set {c±, c2, • • •}; the probability mass which is then assigned by Px to the point cn will be the probability t h a t an experiment on @, designed to yield an exact value of x, gives the value cn, when @ is in t h e state a. I f / is any real valued function of a real variable t, then, for t h e o b s e r v a b l e / o x, we would have the equation: P r o b { / o x = a, when (5 is in the state o-}
(ii)
P
= 2
**(K}).
n:f(cn) = a
With these remarks serving as our motivation we proceed to the formal definitions. Let 3? be any logic and 0 the set of all observables of ££\ A state function of ££ is a m a p (12)
P :x->Px
{xeO)
which assigns to each observable x e 0 a> probability distribution Px on ^(R1) such t h a t for any real valued Borel function / on R1 and a n y observable x, (13)
Pf0X(E)
=
Px(f-HE)).
Equation (13) has an interesting consequence. We claim first t h a t if 0 is the zero observable, then P0 is the probability measure whose entire mass is concentrated at the origin. I n fact, if g is the function identically zero, then g o 0 = 0; since g~1(E) = R1 or 0 according as 0 e E or 0 <£ E, we have, by (13), P0(E) = P0(g-1(E) = l or 0 according as 0 e E or 0 £ E. We next observe t h a t if x is a n y observable and E is a Borel set such t h a t x(E) = 0, then Px(E) = 0. I n fact, ifx{E) = 0, then, for the f u n c t i o n / whose value is 1 on E and 0 on R1 — E, it is obvious t h a t y —f o x is the zero observable and hence Py is the measure concentrated at the origin. As E = {t:f(t) = l}, we have Px(E) = Py({l}) = 0. I n other words, (14)
PX(E) = 0
(Ee ^(R1),
x(E) = 0).
(14) implies t h a t if x is discrete, the mass of Px is concentrated on t h e set of strict values of x. I n turns out to be possible to construct all the state functions of ££ in a particularly simple manner. In order to describe this construction we introduce, with Mackey [1], the notion of a question. An observable x
50
GEOMETRY
OF QUANTUM
THEORY
is said to be a question if #({0,1}) = 1; x is then discrete. If a = x({\}), it is clear that x is the only question such that x({l}) = a. We shall call it the question associated with a and denote it by qa. Suppose now that x is an observable associated with the logic 3? and E any Borel set ^R1. If XE *S the function which takes the value 1 on E and 0 on R1 — E, then XE ° x *s the question associated with x(E), i.e., (15)
XE°x
= qx(Ey
In fact, since XE takes only the values 0 and 1, it is clear that XE ° x i s discrete and (xE ° #)({0,1}) = 1. XE ° # is therefore a question, and, as E is the set where XE takes the value 1, (XEOX)({1})=X(E).
The description of all state functions can now be given in terms of the concept of a measure on 3. A measure on S£ is a function (16)
j9:a-H^(a)
(a
eg)
such that (i) # is real valued and 0 < p(a) < 1 for all a e ^ , (H) 3>(0) = 0, i»(l) = 1, (17) (hi) if ax, a2, • • • is a sequence of mutually orthogonal elements of Se and a = Vn «n> then p(a) = 2n ^ K ) Notice that if al9 a2 e j£? and ax
p(a±) < p(a2)
(ax < a2).
The concept of a measure on ££ is a generalization of the well known concept when j£? is a Boolean cr-algebra. However ££ need not be distributive and hence the structure of a measure on & is much more complex than that of a measure on a Boolean cr-algebra. Theorem 3.5. Let 3 be a logic and 0 the set of all observables associated with ££. Let p be a measure on J&f. / / we define, for any observable xe(9 and any Borel set Eg^R1, (19)
P/(E)
= p(x(E)),
then Pxp is a probability measure on ^(R1) and (20)
Pp : x -> P /
is a state function of ££. Conversely, if P(x -> Px) is an arbitrary state
THE LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
51
function of ££\ there exists one and only one measure p on 3? such that Px(E)=p(x(E))forallx. Proof. Suppose that p is a measure on ££\ If El9 E2, • • • are mutually disjoint Borel sets of E1 with union E, then x(Ex), x(E2), • • • are mutually orthogonal elements of ££ with x(E) = \Jn x(En); moreover,
PX»(E) = 2
p p
x {KY
n
Since Pxp(E1)=p(l) = l) we see that Pxp is a probability measure on ^(E1). If / i s a real valued Borel function of a real variable, then P?ox(E)=p((foX)(E)) =
p{x{f-\E)))
=
Px*(f-i(E)).
This proves that x ~> Pxp is a state function of ££\ Conversely, let P(x —> Px) be a state function of ££. If a e <£?, P a is a probability distribution on the real line and as qa is a question, (14) ensures that PQa({0,l}) = l. Define (21)
p(a) = P, o ({l}).
^)(a->^(a)) is clearly a well-defined, real valued function on ££ and 0
J p(an) = PX(Z), n=l
where Z is the set {1, 2, • • •}. If/ denotes the function whose value is 1 on Z and 0 on E1 — Z, f o x is the question ga and hence, exactly as before, p(a) = Px(Z), so that
(22)
*>(a) = 2 ^ ^ ) . n= l
52
GEOMETRY
OF QUANTUM
THEORY
p is thus a measure on 3?. If x is any observable and E e ^(R1), we have, on writing g for the function whose value is 1 on E and 0 on E1 — E, PX(E) = PgoX({l}) = p{x{E)). Finally, equation (21) implies easily that p is unique. This completes the proof of the theorem. In view of the theorem we proved just now, it is natural to define a state of a logic to be a measure on 3?. The set of all states of j£? will be denoted by £f. If p{a-^p{a)) is a state of S£ and x(E -> x(E)) any observable, E -> p(x(E)) is a probability measure on ^(R1). It is called the probability distribution of x in the state p and is denoted by Pxp. Using the probability distribution Pxp, we may define the concepts of moments of an observable. The (mean) expected value S(x\p) of an observable x in a state p is defined by
(23)
*(x\p) = r
tdP/(t).
J - 00
The variance of x in the state p is defined when >(x2\p) < oo; in this case, (24)
var^)
= £{x2\p)-[£(x\ p)f.
Clearly va,T(x\p)>0 and is = 0 if and only if PXP is concentrated at the single point t0 = (o(x\p), and then it is natural to say that, in the statep, x has a sharply defined value, namely t0.
4. PURE STATES. SUPERPOSITION PRINCIPLE Let ££ be a logic and Sf the set of all states of ££. If pl9 p2, • • • is a sequence of elements of S? and cu c2, • • • is a sequence of constants such that c n >0 for n = l, 2, • • • and c1-j-c2+ • • • = ] . , then the function p(a->p(a)) denned by (25)
p(a) = ^ cnVn{a) n>l
is easily verified to be a state of ££\ We shall write (26)
p = ^ n>l
cnPn.
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
53
Sf is therefore a convex set. If x is any observable, it follows at once from (26) t h a t (27)
c P Pn
PJ> = ^
«* -
n>l
The equation (27) may be given the following interpretation. Suppose t h a t one knows t h a t the state of a system is one of px, p2, • • • with probabilities cl9 c2, • • •, respectively. Then, for any bounded observable x, the expectation value of x is 2n;>i cn^(x\Pn) which is equal to ${x\p). I n other words, assuming t h a t the system is in the state p is, statistically speaking, equivalent to assuming t h a t its state is one of px, p2, • • • with respective probabilities c1, c 2 , • • •. I t is customary to say t h a t the state p is a classical superposition of the states pn(n> 1). I n view of the probability interpretation of (26) and (27), interest centers around states for which no representation of the form (26) is possible. A state p is said to be pure if the equation (28)
P = cp1 + {l-c)p2
(0
pl9p2eST)
implies t h a t p=p1=p2' We write SP for the set of all pure states of j£?. 3P is the set of extreme points of the convex set SP. We may now introduce the notion of superposition. Let 2 be a set of states and p0 an arbitrary state. We say t h a t p0 is a superposition of states in 2 if the following property is satisfied: (29)
a eg,
p(a) = 0
for all
p e 2 => p0(a) = 0.
If p = clp1 + c2p2+ - • • where cl9 c 2 - • • > 0 , and c1 + c2+ • • • = 1 , t h e n p is a superposition of the states in the set {px, p2, • • •}. As we shall presently see, if 3? is a Boolean a-algebra, this is, roughly speaking, the only kind of superposition possible, in particular, no pure state can be a superposition of other pure states distinct from it. This is in contradistinction to quantum mechanics, where the structure of J^, namely the fact t h a t it is standard, forces the concept of superposition to be nontrivial even among pure states. Theorem 3.6. Let ££ be a Boolean o-algebra of subsets of a space X such that (i) ££ is separable (cf. Chapter I) and (ii) {a} e ££ for allae X. For any aeX let Sa be the state defined by (\
if
[fi
if
as
A a$A
Then {Ba : a e X} is precisely the set of all pure states of ££. If 2 is any set of pure states, and p0 an arbitrary pure state, p0 is a superposition of states of 2) if and only if p0 e 2.
54
GEOMETRY
OF QUANTUM
THEORY
Proof. Let {Al9 A2, • • •} generate J*?. T h a t 8a is pure is trivially verified. Suppose t h a t p is a pure state. If for some A0 E J?, 0
= (l-p(A0))-ip(A
n
(X-A0))9
we would obtain the decomposition p =
p(A0)'p1+(l-p(A0))'p2;
p1y£p2 as p1(A0) = l and p2(A0)=0. This is not possible as p is pure b y assumption. Therefore we m a y conclude for any A e J?, t h a t ^(^4) = 0 or 1. B y replacing An by X — An if necessary, we may assume t h a t p(An) = 1 for all n, where {Al9 A2, • • •} generates ££\ Let
B = C]An. n
Then p(B) — \. I n particular, B is nonempty. We assert t h a t B cannot contain more t h a n a single point. I n fact, the collection of all sets G e ££ with the property t h a t either B^C or B n C = 0 , is a a-algebra which contains all the An. Hence it coincides with j£f. Therefore, as {x} e J?, this is possible only if B = {a} for some a e X. But then p = Sa. Finally, let p0 be a superposition of the states of 2) where each element of 2 is pure. If p0 = 8aQ, but pQ<£2, then p({aQ}) = 0 for all p e 3i but Po({a0}) = l; a contradiction. The theorem is completely proved. Let ££ be an arbitrary logic and 0 the set of all pure states of 3?. For a n y subset Ss^&, we write S for the set of all p e 0* such t h a t p is a superposition of the states of S. Let us now consider Jt, the collection of all 8^0 with the property t h a t S=S. Under set inclusion, Jt becomes a partially ordered set. If J£? satisfies the assumptions of theorem 3.6 for example, then jft is the class of all subsets oi0 and is therefore a Boolean algebra. However, in the case of more complex j£f, Jt is not a Boolean algebra. I n general, the geometric structure of Jt seems somewhat hard to determine. For the so-called standard logics to be introduced in the next chapter, this can be done and it becomes possible to determine the structure of Jt completely. For standard ££\ Jt then becomes isomorphic to j£f. I t is this fact which confers on the set of pure states of a standard logic a geometric structure characteristic of q u a n t u m theory. I n the physical literature (cf. Dirac [1]) this is elevated to a physical principle, the so-called superposition principle. From our point of view, the geometry of pure states will appear as a consequence of the structure of the logic 3?. 5. SIMULTANEOUS O B S E R V A B I L I T Y One of the features which sets apart quantum mechanics from classical mechanics is the existence of observables whose values cannot always be
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
55
simultaneously measured. We shall now introduce this idea formally. We begin with the elements of ££ themselves. Let a9b e ££. We shall say t h a t a and b are simultaneously verifiable, a <—> b in symbols, if there are elements al9 bl9 and c e i f such t h a t *
(i) al9 bl9 and c are mutually orthogonal, (ii) a = a1 V c, b = b± V c.
'
If a: and y are two observables, they are said to be simultaneously observable if for any two Borel subsets E and F of R1, x(E)
Lemma 3.7. Let a,b e ££, 3? being any logic. Then the following
statements
are equivalent: (a) a <-» b9 (b) a-(a A b) J_ b, ^C) 6 ~ ^ a A 6 ) - L a ' (d) ^ e r e erciste an observable x and two Borel sets A and B of the real line such that x(A) = a and x(B) = b, (e) there exists a Boolean subalgebra of ££ containing a and b. Proof. Suppose a <-> b. We m a y then write a = a1y c, b = b±vc, where al9 bl9 and c are mutually orthogonal. Clearly c
ax = a— (a A b) ±_ b± v c = b. This proves t h a t (a) => (b). By symmetry, (a) => (c). If a — (a A b)_\_b, then, on writing a x = a — (aA6), b1 = b — (aAb) and c = aAb9 we find t h a t a = a1v c and b = b1v c. Since %_[_&> w e have «i_L^i a n d a i_L c > while, by
56
GEOMETRY
OF QUANTUM
THEORY
definition, b1_[_c. Therefore a<-> 6, proving that (b) => (a), (a), (b) and (c) have thus been proved equivalent. We shall complete the proof by showing (a) => (d) => (e) => (a). If a = a1vc, b = b1vc, where al9 bl9 and c are mutually orthogonal, we write d = a1vb1\/ c and define x to be the discrete observable such that x({0}) = a1, #({!}) = &!, x({2}) = c and x({3}) = d1. Then x({0,2}) = a, x({l,2}) = b. This proves that (a) => (d). Suppose that (d) is satisfied. Then x(A-Ar\B) = a~-aAb, x(B — AnB) = b — aAb. Writing a1 = a — aAb,a2 = aAb,a3 = b — aAb,a± = (avb)1iwe see that the a{ are mutually orthogonal and a± V a2 V a3 V a4 = 1. If sf is the collection of elements which are the lattice sums a
h V ai2 V • • • V aik
(k < 4, 1 < t\ < i2 < • • • < ik < 4),
then it is easy to verify that s/ is a Boolean subalgebra of ££. Since a,b e jtf, we see that (d) => (e). Finally, let (e) be satisfied, and let $0 be a Boolean subalgebra of ££ containing a and 6. Clearly [a — (a A b)] A 6 = 0. But, as a— (a A 6), 6 and 6 1 are in the Boolean algebra s>/, a-(a
A b) = {(a-(a
A 6)) A b} V {(a-(a A 6)) A &1}
= {a-(aA6)} A b1 (b) and hence that (e) * (a). Corollary 3.8. If a<-*b, the elements ax, bl9 and c of (31) are uniquely determined. In fact, c = a Ab, a±=a — aAb, and b1 — b — aAb. Lemma 3.7 shows that in the presence of simultaneous verifiability of two elements of ££\ one can operate with the statements corresponding to these elements as if they were classical. We shall now take up the question of generalizing this fact to observables. Our aim is to prove that simultaneous observability is essentially a characteristic of classical systems. We shall deduce, as a corollary, that, if x, y, z, • • • are observables, associated with a logic, they are mutually simultaneously observable if and only if they are all functions of a single observable. This was first proved by von Neumann (cf. [1]) when the logic in question was a standard one. For a general logic, it was proved by Varadarajan [1]. Theorem 3.9. | Let J? be any logic and {xA}AeD a family of observables. Suppose that A
—^ *^A'
for all A, X' e D. Then, there exists a space X, a f The definition of a logic given in Varadarajan [1] does not require it to be a lattice. I t was pointed out to us by Dr. S. P . Gudder [1] that theorem 3.9 does not remain valid for these more general classes of partially ordered sets considered in our paper [1].
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
57
o-algebra & of subsets of X, real valued ^-measurable functions gK on X(X e D), and a a-homomorphism T of & into ££ such that (33)
r{gA-\E))
= xK(E)
1
for all A e D and E e^{R ). Suppose further that either ££ is separable in the sense that every Boolean sub-a-algebra of ££ is separable, or that D is countable. Then there exists an observable x and real valued Borel functions f\ of a real variable such that for all A e D, (34)
xK = fA o x.
The proof of this theorem (cf. Varadarajan [1]) depends on a series of lemmas. J? is a fixed logic in these lemmas. Lemma 3.10. Let 6, al9 a2, • • • be elements of <S£.Ifb
(35)
&<->V«n, n
b+-> /\an. n
Moreover, we then have the distributivity laws (36)
o A A / < | = V (h A an), U n ' o V (A «») = A (6 V an).
Proof. We use the criteria of (32). It follows from the equivalence of (a) and (d) of (32) that if c <-> d, then any two of the four elements c, c1, d, d1 are simultaneously verifiable. This proves the first of the three relations of (35). Now, we come to the proof of the rest of (35). Write a = N/m am. Since b A am a. Changing all the an into anL we deduce from the relation b <-> \/n an that b <-> An an- Next we come to (36). Once again it is enough to prove only the first of the two relations. Clearly \/n (b A an) < b A a, where a = \Jn an. Let d < b A a be such that d_|_Vn (b^an) a n d (Vn (bAan))yd = bAa. Since d_j_&Aan and 6? an, we conclude that dj_a n , from (c) of (32). Since this is true for all n, d\_a. As d
58
GEOMETRY
OF QUANTUM
THEORY
P | a ££\ is also a sublogic of <£\ From this it follows that if {^a}aeA i s a n y family of sublogics of ££, there exists a smallest sublogic containing all the j£?a. We denote it by \Ja j£?a. 9ft is thus a complete lattice. If each j£?a is a Boolean subalgebra of ££\ it is however not true in general that \/a J?'a is a Boolean subalgebra. We shall now prove that this is so if and only if the relations ££\ <—> ££\> are satisfied for all a, a e A. We begin by considering the case when there are only two Boolean subalgebras involved. Lemma 3.11. Let Mx and 0t2 be two Boolean subalgebras of ££\ Then 8%x <—> ^ 2 if and onty if &i V ^ 2 is a Boolean subalgebra of ££. Proof. Suppose that ^t1 M 0t2 is a Boolean subalgebra of ££. Then Si1 <—» ^ 2 m view of (e) of (32). We shall now prove the converse. We shall first consider the case when the Boolean subalgebras are finite. Let F% (* = 1, 2) be finite Boolean subalgebras of 3? such that Fx <->#*2. Then, their Stone spaces (cf. Chapter I) are finite and hence there are elements al9 • • •, ar e Fx and bl9 • • •, bs e F2 s u c n that W ail.
cw = af A 6,
(* = 1, 2, • • •, r, j = 1, 2, • • •, 5).
Obviously c{j±_cvr unless i = i' and j = j ' . Since ai^-^bj we infer from (36) that
for all i and j ,
s
ij = at>
c
V
(38)
V % = &/• i=l
Consequently,
(39)
V V % = Ii=l
;= 1
If ^" denotes the collection of all lattice sums of the ci;, then F is a Boolean subalgebra of ££ and (38) implies that J j C j fori = l, 2. The formulas (37) show that if IF' is any sublogic of ££ containing both Fx and F2, then J ^ c J ' . Clearly, therefore, F^Fxy^2, Note that •^"1 V ^"2 *s a ^ s o finite. Let ^ (i = l, 2) be Boolean subalgebras of ££ such that ^ <->^ 2 - Let ^ be defined by (40)
» = U ( * ' i V ^2),
where the union is over all pairs [FX,F2) of finite Boolean subalgebras of S£ such that F^M^ i — 1, 2. We assert that ^ is a Boolean subalgebra
THE LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
59 1
of «£? containing 0kx and 0k2. If a e 0k, then it is obvious that a e 0k. Clearly, 0, 1 e &. Suppose now that a,b e0k. Then there are finite Boolean subalgebras J^ l5 J ^ ' , ^ 2 , J*Y of & such that J ^ , . F / are c # x , J*"2, J ^ ' are ^ ^ 2 , a e ^ V ^ a n d fte^/v^'Since ^ and «^Y are contained in the Boolean algebra 0kx, tFx <—> <^Y and hence ^ \ v ^ Y = ^ Y ' is finite and contained in ^ x . Similarly, ^ 2y ^ 2 =: ^ 2" is finite and contained in 0 2 . As mi
60
GEOMETRY
OF QUANTUM
THEORY
valid for all s<s0 and let s = s 0 + l. Since Sti
»{F)
a
= V #AAeF
Obviously, St{F)<^St{F')
(42)
if F^F'.
Define St by
St =
U
St(F).
F finite
Obviously, StK<^St for all A e D. We shall now complete the proof by showing t h a t SI is a Boolean subalgebra of j£?. I t is enough to prove t h a t any finite subset of St is contained in a Boolean subalgebra of ££ contained in St. If a1,a2,'-'iar are elements of St there are finite subsets Flt F2, • • •, Fr of D such t h a t a{ e St{F{). If F = ( J t .F t , then a t e ^ ( ^ ) for all *. This proves t h a t ^ is a Boolean subalgebra of J^. Corollary 3.15. Let {StK}KeD &e a family of Boolean sub-a-algebras of 3? such that StK <—> StK' for all A, A ' e D . Then, there exists a Boolean sub-aalgebra St of <£ such that StK<^St for all XED. Proof. By lemma 3.14 there are Boolean subalgebras St' such t h a t St^St' for all A. By Zorn's lemma we can choose a maximal such, say St. We claim t h a t St is a Boolean a-algebra. I t is enough to prove t h a t if al9 a2> * * * is a sequence of elements of St, then a=\/m am is also in S$. Since am eSt, b <-» a m for all 6 G ^ . Hence a < - > ^ by (35). If j / = {0,a,a 1 ,l} then j / is a Boolean subalgebra of «£? and stf ±-*St. Hence si M St is a Boolean subalgebra of j£? containing S%. Since ^ is maximal, stf'v St=St. Therefore a e i For any observable # associated with 3? we write
(43)
St(x) = {z(#) : # e ^(i? 1 )}.
We shall call S%(x) the ran^e of a:. Lemma 3.16. Let St be a Boolean sub-o-algebra of j£f'. Then, in order that St=St(x) for some observable x, it is necessary and sufficient that St be separable.
THE LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
61
Proof. If 0k=0k(x) and @ = {x(E) : E an open interval with rational end points}, then ^ is a countable subset of 0k and 0k is the smallest sub-cralgebra of itself containing Q). 0k is thus separable. Conversely, let 0k be a separable Boolean sub -a- algebra of j£f\ By theorem 1.3, there exists a space X, a a-algebra 01 of subsets of X, and a a-homomorphism h of 01 onto ^ . Let {^4^ ^42> • • •} be a sequence of subsets of X such that An e 08 for all n and {^(^4n) : n = l, 2, • • •} generates ^ . Let «^0 be the sub-o-algebra of 9S generated by {An : w = l, 2, • • •}. By theorem 1.6(i), there exists a cr-homomorphism w of ^(R1) onto «^0. Define a; by (44)
x(E) = A(M(JS))
(# G ^(i? 1 )).
It is obvious that x is an observable and 0k=0k(x). Now to the proof of theorem 3.9. Suppose that x is an observable and xA=fA ox (A G D), where each /A is a real Borel function on R1. Then 0k(x^)<^0k(x) and hence &(xK) <->0k{xK>) for all A, A' e D. This proves that xx <-> # v , for all A, A' e D. Conversely, suppose that {xA} is a family of observables and that #A<->#A, for all A, X'eD. By corollary 3.15, there exists a Boolean sub -a- algebra 01' of ££ such that 0k(xK)^0k' for all A. By theorem 1.3, there exists a space X, a cr-algebra 01 of subsets of X, and a cr-homomorphism r of 0? onto ^ ' . It is now clear that for each A G D, there exists, by theorem 1.4, a real valued ^-measurable function gA on X such that, for all E e ^{R1), xx(E) = r[gK-\E)}
(XeD).
This proves the equation (33). For proving (34) let 0t be the smallest sub-cr-algebra of 0k' containing all the 0k(xK). If 3? is separable, so is 0k. On the other hand, if D is countable, and i^A is a countable set generating &(x\)> UA ®\ i s a countable set generating £%. Hence 0k is separable in either case. By lemma 3.16 there exists an observable x such that 0t(x) =0k. As 0k(x-^ <^0k(x)i we conclude from theorem I.6(ii) that there exists a real Borel function / A of a real variable such that xx =/ A o x. This proves (34) and completes the proof of the entire theorem. Remark. It is obvious that if the xK satisfy either (33) or (34), A ^ ^
A'
for all A, A' ED. Theorem 3.9 therefore tells us that the classical logics are precisely those for which any two observables are simultaneously observable. As soon as ££ is not a Boolean algebra, it follows that the properties of a physical system which are embodied in the logic ££ have too complex a structure to be exhausted by a single set of simultaneously observable quantities. In general, when certain propositions are experimentally verified, certain others cannot be so at the same time.
62
GEOMETRY
OF QUANTUM
THEORY
6. FUNCTIONS O F S E V E R A L OBSERVABLES The characterization given in theorem 3.9 of simultaneous observability enables us t o construct a calculus of functions of several observables which are mutually simultaneously observable. Theorem 3.17, Let ££ be any logic and observables such that x{ <-> Xj for all i, j . Then there exists one and only one a-homomorphism r of @(Rn) into & such that for all E e ^{R1) and all i = 1, 2, • • •, n, (45)
xx(E) =
rfrc^E)),
where iri is the projection (tl9 t29 • • •, tn) - > tt of Rn on R1. If g is any real valued Borel function on Rn, (46)
g o (Xl,x29.
- .9zn) : E ->
rlg-HE))
is an observable. If gl9 g2, • • •, gk are real valued Borel functions on Rn and yi — gi o (xl9 - - -, xn), then yl9 • • •, yk are simultaneously observable, and for any real valued Borel function h on Rk, (47)
ho (yl9--,yk)
= (h(gl9 • • •, gk)) o (xl9 • • ., xn),
where h(gl9 • • •, gk) is the function t =
(t1,--,tn)->h(g1(t),...,gk(t)).
Proof. B y theorem 3.9, there exists a n observable x and real valued Borel functions fl9 • • - , / n of a real variable such t h a t xi=fi o x for all i. Let t h e Borel m a p f of R1 into Rn be defined by f : * - M / i ( 0 , •••,/»(*)). If we define, for M e &(Rn), (48)
r(M) = xtf-^M)),
t h e n r is a a-homomorphism of &(Rn) into j£? which satisfies (45). If r and T' are two a-homomorphisms of &(Rn) into ££ satisfying (45), then T(M)
=
T\M)
whenever M is of the form Ex x E2 x • • • x En9 the Ej being Borel subsets of R1. Since the class of such sets generates &(Rn), it follows t h a t T = T'. The fact t h a t g o (xl9 • • •, xn) is a n observable is trivial. Note t h a t its range is contained in t h e range 0t of r which is a Boolean sub-a-algebra oiX.
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
63
Suppose now that yi — qi o (xx, • • •, xn). Then the range of y{ is contained in 0t and hence y1, • • •, yk are simultaneously observable. Let g be the map t -» (^(t), • • •, gk(t)) of Rn into Rk. Then
y. M-> r(g-\M)) is a <7-homomorphism of &(Rk) into J£ such that
y\^c\m
= AgrHE)) = yt(E),
TTX being the map (sl9 s2, • • •, sk) -> s{ of Rk on R1 (i = 1, 2, • • •, &). Therefore we conclude from the uniqueness of the map r in (45) that, for w=
h°(yi>y2>--,yk)>
u{E) = y{h-\E))
(E e ^{R1)).
But y(A~1 (E)) = T(g~1(A~1 (2£))), so that we may conclude that A o (2/i, • • •, yk) = ( % i , • • •, &)) ° (»i, • • •, s n ) .
This proves the theorem. Theorem 3.17 enables us to define the joint distributions of xlt ...,#„. Let ^(a-^^)(a)) be a state of J*? and xl9 • • •, xn observables such that X^ ^—^ »£y for all i, j . We may then clearly define the probability measure P* 1> ..., Xn on^( J R»)by (49)
P»l.....Xn(M) = p(r(M)). s
Pxlf-,xn * called the joint probability distribution of (xl9 • • •, xn) in the state p. It follows easily from (47) that for Borel Mg:Rk, (50)
P5lt...iiyfc(Jlf) =
P*lt...tXn{g-*{M)),
where g is the map t =
{h,--,tn)-+(g1(t),...,gk(t)).
We see from (50) that the rules for the calculation of the joint probability distributions are the standard ones of probability theory.
7. THE CENTER OF A LOGIC We begin by recalling the discussion of Chapter II in which we singled out the geometries among the modular lattices of finite rank by demanding that their centers be trivial. We shall now do the same for logics. The irreducible logics would then have properties very far removed from those of Boolean algebras and would consequently serve as plausible models for the logics of atomic systems.
GEOMETRY
u
OF QUANTUM
THEORY
Let ££ be any logic. We define the center of 3? to be the set %>, where (51)
V = {a:ae&,
a^b,
for all b e &}.
*& is nonempty since 0 and 1 are in *€. The elements of ^ are thus simultaneously verifiable with every element of ££. If x is any observable with &(x) $=:<£, then for any observable y, x+-+y. Such observables x are called central. ££ is said to be irreducible if <€={0,1}- Clearly, ££ is irreducible if and only if the only central observables are the constants. Theorem 3.18. Let ££ be any logic and <€ its center. Then <$ is a Boolean sub-o-algebra of ££\ Proof. The equations (35) and (36) reveal immediately that ^ is a Boolean subalgebra of «£?. If ax, • • •, ar, • • • is a sequence of elements in < €, then b
p~(a) =p(a A an)
(ae&),
then p~ is a pure state of ££. If (53)
^ n = {p~ : p a pure state of J?n},
then the {0*n : ne D) are disjoint subsets of & and (54)
0> = U &„
If *& is continuous and separable, & is empty, i.e., J? has no pure states.
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
65
Proof. We have already noted that for any n, j£?[0,an] = «£?n is a logic (corollary 3.3). We prove that ££n is irreducible. In fact, let c
qM = (*>(««)) " ^ ( a A an), ?2(a) = ( l - ^ ( a m ) ) - ^ ( a A amL).
Since am e # , it can be easily verified using (35) and (36) that qx and q2 are states of ££\ Since qi(am) = l, and g ^ m ) ^ * <7i#(?2- As 2> = ^ K ) • ?i + (1 -!>(««)) • q2> p is not pure, a contradiction. Therefore p(an) = 1 for some ne D. The argument given in the previous paragraph can now be repeated to enable us to conclude that the restriction of p to «j£?n, say q, is a pure state of 5£n and that p = q~. This proves that pe&n and completes the proof of (54). We now come to the proof of the last assertion. Assume that ^ is continuous and separable. Suppose that p is a state of ££'. We assert first that for some c e # , 0
GEOMETRY
66
OF QUANTUM
THEORY
L
cn if necessary, assume that p(cn) = l for all n. Let c0 = /\ncn. Clearly, p(c0) = l. In particular, c 0 ^ 0 . We assert that c0 is an atom of <€. In fact, the subset of all elements c in ^ such that either c_[_c0 or c> c0, is a sub-o-algebra of ^ containing all the cn and so coincides with # ; this shows that if a
g2(a) = (l-p{d)y1p(a
A d 1 ).
Then p=p(d)q1 + (1 — p(d))g2 and p is not pure. The proof of the theorem is complete. We shall end this section with a few remarks of a general nature regarding the Heisenberg uncertainty principle. We must point out that a considerable part of its meaning is physical and we do not go into this here (cf. the discussions of von Neumann [1] and Heisenberg [1]). We shall be concerned only with mathematical formulations. Suppose that (x2\p)
op*{x) =
£{x*\p)-{£{x\p)f.
If crp(x) = 0, then the probability distribution of x in the state p is concentrated at the point &(x\p). x has then a sharply defined value in the state p. One might ask the question whether there are states in which every observable has, if not a sharply defined value, at least a small variance. The Heisenberg principle of uncertainty asserts that even this is not possible. Roughly speaking, it asserts that in any state, no matter how prepared, there are observables whose distributions have large variances. Notice that this is not possible if the logic is a Boolean a-algebra of subsets of a space. In fact, let «j£? be the cr-algebra of subsets of a set X. Then, for any observable with corresponding function/, the observable has a sharply defined value f{x0) whenever the state p is a measure concentrated in a point x0 of X. On the other hand, let us assume that © is the physical system of some atomic particle and that x and y are the position and momentum observables along some direction in space. If we assume that the logic .if of © is the standard logic associated with the complex number field, then (57)
°p(x)*p(y) > P ,
where h is Planck's constant divided by 2TT (cf. H. Weyl [1] pp. 77 and 393-394 for a rigorous derivation). From (57) it follows quite simply that
THE LOGIC OF A QUANTUM MECHANICAL
SYSTEM
67
in any state in which the distribution of a: has small variance (i.e., x is "localized"), the distribution of y has very large variance. I t is obvious that this is a property of the structure of the logic JS? of <3. 8. AUTOMORPHISMS Let ££ be a logic, 0 the set of observables of ££, and Sf the set of all states of ££\ The set of all automorphisms of ££ forms a group in a natural fashion, denoted by Aut(j£f). We shall now introduce the notion of an automorphism of Sf. By a convex automorphism of S? we mean a map (58)
(:p-+p*
(peS?)
such that (i) $ is one-one and maps Sf onto itself, (ii) if cl5 c2, • • • are numbers > 0 with 2 n cn —1> and jpl5 p 2 , • • • e £?, fecnPnY V n
= 2 /
C
»^'
n
From (ii) of (59) it follows at once that convex automorphisms of £P send pure states of J£? into pure states of ££? and that the correspondences induced in the set of pure states are one-one and onto. If a : a -> aa
(60)
(ae^f)
is an automorphism of j£? and if, for p e Sf we define pa by pa(a)
(61)
=p(a^~%
a
then p -> p is a convex automorphism of t9^. We shall call it the automorphism induced by a. The set of all convex automorphisms of Sf also forms a group denoted by A u t ( ^ ) . Let ££ be a a-algebra of subsets of a space X, and let * : x -> <(a)
(a? e Z)
be any one-one map of X onto itself such that for any set A<=iX, A e J? if and only if t" X{A) e J^7. When X and «^f satisfy some technical regularity conditions, it can be shown that every sufficiently well-behaved automorphism of Sf is induced by a mapping t such as the one described above. Suppose X is the phase space of a classical mechanical system with a C00 manifold M as its configuration space. X thus appears (cf. Chapter I) as the cotangent bundle of M. In this case, the subgroup of those automorphisms of ££ induced by (both ways) differentiate homeomorphisms of X onto itself and which preserve the canonical 2-form on X, plays undoubtedly a very crucial role (contact transformations).
68
GEOMETRY
OF QUANTUM
THEORY
The axioms satisfied by an arbitrary logic are too weak to allow any reasonable determination of its automorphisms. In the next chapter we shall determine the automorphisms associated with standard logics. In the meantime, we shall briefly illustrate with two (among many) examples the important role played by the group Aut(^) in physical problems. As a first example we consider the problem of describing the dynamics of a physical system <&. Let j£? be the logic of ©, £P the convex set of all its states. Then, for each real number t, there exists a unique one-one map of SP onto itself, say D(t), with the following physical interpretation; if p e Sf is the state of <5 at time t0, then D(t)p is the state of <S at time t+t0. The dynamical group Qj where (62)
9 : t -> D(t)
satisfies the conditions (3) of Chapter I. We now postulate that each D(t) is in fact an automorphism of SP, i.e., if plt p2i • • • is a sequence of elements of SP and clt c2, • • • a sequence of nonnegative numbers such that c i + c 2 + • • • = 1, then, for each t, (63)
D(t)(2ciPi)
=2^(*)3>i.
This postulate may be justified by the following argument. Suppose one does not know the state at time 0 but only that it is one ofp1,p2, • • • with probabilities c1? c2, • • •, respectively. Then from the fact that the state at time t is D(t)p{ if the state time 0 is pt, it follows that at time t the state of the system (3 is one of D(t)pu D{t)p2i • • • with the same probabilities c l5 c2, • • •. But the statement that the state of S£ (or (&) is one of qx, q2i • • • with probabilities cl5 c2, • • • is physically equivalent to the statement that the state of ££ is c1q1+c2q2+ • • •. This proves the validity of (63). It follows now that each D(t) leaves invariant the subset of pure states of j£? and induces a one-one mapping on it. Furthermore, if a e ££, then, for each p e £f, (64)
t^(D(t)p)(a)
is a real valued function of t; it is physically reasonable to assume that for any 3 , this function is Borel for each p e SP and a E &. If this condition is satisfied, we call 2 a Borel one-parameter group. Thus the dynamical group of (3 may be assumed to be a Borel one-parameter group of convex automorphisms of 6f. This leads us to introduce the concept of representations of groups. Let ( j b e a locally compact topological group satisfying the second axiom of countability. The sets of the smallest a algebra of subsets of G
THE
LOGIC OF A QUANTUM
MECHANICAL
SYSTEM
69
containing the open sets are called the Borel sets in G. By a representation of G in Aut(y) we mean a map (65)
T:g^Tg
(geG)
of G such that
(66)
(i) each Tg e Aut(^), (ii) Tgig2 = TgiTg2 (gltg2e€f), (iii) for each a e ££ and p e Sf, the function g -> (Tgp)(a) is Borel on£.
With these definitions we may say that the dynamical group of a system © is determined by a representation of the additive group of real numbers into Aut(^), where £f is the set of states of the logic ££ of ©. As our second example, we consider the problem of describing the quantum mechanical theory of a free particle. The configuration space of the particle in classical mechanics is the three-dimensional space, denoted by M. Let ££ be the logic of the physical system under consideration and Sf the convex set of its states. Suppose now an observer 0 wants to describe the system. He would then select a point of M, an orthogonal frame at 0, and map M onto R3 by a map (67)
y0'-M->
R3.
Let 0' be another observer and let yQ> be the corresponding map. Then there exists a unique Euclidean motion D0t0, of R3 such that for all meM (68)
yo,{m)
=
D0t0,(y0(m)).
Suppose now 0 describes the state of the system © by an element p0 e £?. Then, 0', using his orientation, would describe the same physical state by an element p0, of SP. It is clear that the map pQ —> pQ, is an automorphism, say a0t0> of Sf. In other words, a00, has the following physical interpretation: if the physical state of © is described by 0 by the element ge£f, then a0t0\g) is the element of £f describing the same physical state of © from the point of view of 0'. It is obvious that a0tQ, is the identity for 0 — 0\ and (69)
a0i0» = a0,f0»a0t0>.
We now remark that the basic physical laws of © are independent of the choice of observers and their frames. Therefore, we may conclude that a o,o' depends only on the transformation D0,o'- Let G be the group of all
70
GEOMETRY
OF QUANTUM
THEORY
3
Euclidean motions of R . Then, from (68), (69), and the above remarks on the a00,, we infer the existence of a homomorphism (70)
T:g->Ta
of G into A u t ( ^ ) with the property t h a t whenever 0 and 0' are two observers of
«o.o> =
TD0,0,-
We m a y reasonably make the assumption t h a t for each a e & and p e £f, g - > (Tgp)(a) is a Borel function on G. T is thus a representation of G in Aut(^). We see thus t h a t the description of <& gives rise to a representation of G in A u t ( ^ ) . The determination of such representations is therefore the basic problem to be solved in the mathematical description of the physical system of a free particle. The reader might notice t h a t the arguments given above are very general. I n fact, the same type of arguments lead to the remarkable fact t h a t with every relativistic free particle there is associated a representation of t h e inhomogeneous Lorentz group. These representations were first completely determined by Wigner under t h e assumption t h a t t h e logic of <3 was standard. This led in t u r n to his celebrated classification of the relativistic wave equations [1].
N O T E S ON C H A P T E R I I I 1. As a general reference on lattice theory we refer to the book of F . Maeda and S.Maeda, Theory of Symmetric Lattices, Springer-Verlag, Berlin, 1970. 2. There has been a good amount of work by m a n y people in recent years on the structure of logics and probability measures on them, motivated mostly b y the relevance of such problems to the foundations of Quantum Mechanics. The encyclopedia volume of E . G. Beltrametti and G. Cassinelli (The Logic of Quantum Mechanics, Volume 15, Encyclopedia of Mathematics and Its Applications, Addison-Wesley, Reading, Mass., 1981) gives an excellent overview of this work (Chs. 10-20). I n addition, this book presents a well-organized, comprehensive discussion of the foundations of Quantum Mechanics from different points of view with m a n y references to current literature, examples, and exercises. Especially illuminating are the critiques of controversial topics such as hidden variables (Chs. 15, 25) consistency of physical interpretation (Chs. 7, 8), the process of measurement (Ch. 26), and so on. As additional references we mention the following: J . M. J a u c h , Foundations of Quantum Mechanics, Addison-Wesley, Reading, Mass., 1968.
THE LOGIC OF A QUANTUM MECHANICAL
SYSTEM
71
C. Piron, Foundations of Quantum Physics, Benjamin, Reading, Mass., 1976. M. Jammer, The Philosophy of Quantum Mechanics, Wiley, New York, 1974. C.A.Hooker (Ed.), The Logico-algebraic Approach to Quantum Mechanics, Vol. 1, Reidel, Dordrecht, 1975. E. G. Beltrametti and B. C. van Frassen (Eds.), Current Issues in Quantum Logic, Plenum, New York, 1981. 3. The structure of the states of orthocomplemented continuous geometries is still mysterious. One of the most striking results is due to von Neumann who showed that in the presence of a transition probability function with suitable invariance properties, such a lattice is nothing but the lattice of projections of a type 1^ factor (Continuous geometries with a transition probability, Mem. Amer. Math. Soc. No. 252, Providence, R.I., 1981). 4. It is possible to determine, in the standard model, the states for which the lower bound of the Heisenberg uncertainty relations (57) is attained, for the quantum mechanical system of a one-dimensional particle. Let je = L2 (-00,00),
q = operator of multiplication by xt p = - j - ; then ov(g)-M#)=2 for the family of states
a9(p) = («y/2)*
(cf. von Neumann [1], pp. 233-237). As we shall see in greater detail in Chapter IV, any vector
\cp\ JE
dx.
Thus the wave function allows one to calculate, by this formula, for any region E^(~ao,co), the probability/*(!£) of finding the particle inside E. This is the simplest instance of the statistical interpretation of Quantum Mechanics which is due to Max Born (Z. Physik, 37 (1926), pp. 863-867).
CHAPTER IV LOGICS ASSOCIATED WITH HILBERT SPACES 1. T H E LATTICE O F SUBSPACES O F A BANACH SPACE The general theory t h a t was developed in Chapter I I I is concerned mainly with the broad principles of Quantum Theory. Further developm e n t of the theory, such as a discussion of simple quantum mechanical systems, leads to problems of a more technical nature. These problems are, however, difficult to answer in the context of abstract logics, and therefore, in dealing with them, it becomes necessary to restrict the class of logics under consideration. I n this chapter we shall examine some of these logics and describe, in a somewhat more concrete fashion t h a n in Chapter I I I , the observables, states, and symmetries associated with t h e m . We have seen in Chapter I I t h a t a large class of orthocomplemented geometries are obtained by considering Hilbert space structures on finite dimensional vector spaces. These logics are, however, not of much use in discussing even very simple atomic systems, since every observable associated with t h e m is finite. Consequently, if one wants to obtain realistic examples, one is forced to deal with infinite dimensional vector spaces. I t is known (cf. Baer [1]) t h a t the lattice of all linear manifolds of an infinite dimensional vector space does not admit any polarity. Therefore it becomes necessary to consider lattices j£? such t h a t the elements of <£? are certain linear manifolds of a given vector space V and, wherein, the partial ordering is the inclusion relation. I t must be noted t h a t the lattice sum of two elements of such a lattice contains, but in general does not coincide with, the algebraic sum in V. A simple and interesting class of such lattices may be obtained in the following way. Let D be a topological division ring, say, for example, one of R, C, or H. Let V be a topological vector space over D (cf. Bourbaki [2] for the definitions and elementary properties). Then, the set of all closed linear manifolds of V is a lattice—in fact, a complete lattice. One m a y then construct examples of logics whose underlying lattices are these lattices of closed manifolds. The description of such logics would provide us with a large class of examples of logics which are complex enough to serve as models for the logics of atomic systems. 72
LOGICS ASSOCIATED
WITH HILBERT
SPACES
73
The main purpose of this section is to prove a theorem of Kakutani and Mackey ([1]), which asserts t h a t if V is a Banach space, the above procedure leads to no new examples of logics other t h a n the standard ones. I n particular, if the underlying lattice of a logic is isomorphic to the lattice of all closed linear manifolds of a separable infinite dimensional Banach space over one of R, C, or H, then the logic is standard. Let D be one of R, C, or H and V any vector space over D. We shall say t h a t a topology 3T on V is compatible with the vector space structure of F i f it is Hausdorff, and, if the map, a, b,x, y^ax+byof&x'Dx VxV into V, is continuous. A topological vector space over D is a pair (F,^"), where V is a vector space over D, and «^", a topology over V compatible with its vector space structure. By the usual abuse of language we shall say t h a t V is a topological vector space. We shall denote by «j£?(F,D) the lattice of all closed linear manifolds of V. This is a complete lattice. If we denote by V and A the lattice operations in j£?(F,D), then, for any family {Ma} of closed linear manifolds of V, we have:
(i)
A Ma = n Mtt> a
(2)
a
\JMa=^May,
where 2 denotes the algebraic sum, and the bar denotes closure. If N is a n y finite dimensional linear manifold of V, then N is necessarily closed and for any closed linear manifold M, N v M = N -\-M. Moreover, if M is a closed linear manifold whose codimension ( = dim VjM) is finite and N any algebraic complement of M, i.e., such t h a t M n N = 0, M +N= V, then N is necessarily closed and V = M y N (for these facts cf. Bourbaki [2], Ch. 1). We shall recall the notion of Banach spaces. Let D be one of R, C, or H and V a vector space over D. A norm over V is a function x - > ||a?|| of V into the real numbers such t h a t
(3)
(i)
||a || > 0, = 0
(h)
\\x +y\\ < \\x\\ + \\ y\\
(in)
||ca;|| = |c| • ||#||
if and only if for
for
x = 0,
x, y e V,
c e D, x e V.
A Banach space over D is a pair (V, \\ • ||), where V is a vector space over D and || • || a norm on V such t h a t V is complete under the metric d defined by d(x,y) =\\x — y\\. A Banach space over D is obviously a topological vector space over D. We are now in a position to formulate the theorem of Kakutani and Mackey. We recall the notion of Hilbert spaces and inner products associated with D (cf. Chapter I I , section 4).
74
GEOMETRY
OF QUANTUM
THEORY
Theorem 4.1 (Kakutani-Mackey [1]). Let D be one of the three division rings R, C, or H, V an infinite dimensional Banach space over D, and jSf = o^(F,D) the lattice of all closed linear manifolds of V. Suppose that ±_{M-> M1) is an orthocomplementation of j£f. Then there exists a Dvalued inner product <. , . > on V xV such that (i) V becomes, under <. , . >, a Hilbert space over D; (ii) the topology of V, induced by the norm associated with <. , .>, coincides with its original topology; and (iii) the map M -> ML coincides with the orthocomplementation induced by <. , .>. In particular, if the underlying lattice of a logic is isomorphic to the lattice of closed linear manifolds of a separable Banach space of infinite dimension over D, then the logic is standard. The proof is rather long and depends on a series of lemmas. With later applications in mind we formulate our first lemma somewhat more generally than is necessary for our immediate purposes. Lemma 4.2. Let K be a division ring and W a vector space of infinite dimension over K. Let Jtbe a lattice of linear manifolds of W such that (i) Jt contains all finite dimensional linear manifolds of W, and (ii) if M,N e Jt and at least one of them is finite dimensional, then My N = M+N. Suppose Jt{M -> ML) is an orthocomplementation in Jt. Let w0e W be any nonzero vector. Then there exists an involutive anti-automorphism 6 of K and a symmetric 6-bilinear form <. , .) on W x W such that (i) (w0,woy = 1 and (ii) (x,y} = 0 if and only if x e (K • y)L • 6 and (. , . > are uniquely determined. The form <. , .) is definite, i.e., (x,x) = 0 if and only if x = 0. Proof. Let F^ W be a finite dimensional linear manifold. Then F e Jt'. For any linear manifold L^F we have L e J also. We now define (4)
(F(L) =
L^nF.
Note that if M is finite dimensional and iV" e Jt, then M AN = MnN. In fact, M AN^MOLN always while the finite dimensionality of Mc\N implies, by (i), that MnN e Jt, so that MnN^M A N. Let JtF be the projective geometry of all linear manifolds L^F. £F is a map of JtF into itself. We shall show first that £F is an orthocomplementation of J/F. Clearly ^(0) = ^ , £F(F) = F1nF = F1 A F = 0. Since M —> ML is order inverting, it is obvious that fF is order inverting. We observe next that for LeJtF, Ln£F(L)^LnL1 = L A L1 = 0. On the L L other hand, W = L v L = L + L so that F = L+£F(L). Thus L and £F(L) are complements of each other in JtF. To complete the proof that £F is an orthocomplementation, it remains to prove that £F is involutive. For LeJtF, £F{L)<^LL so that L<^ijF(L)L. This shows that L£gF(L)Ln F=€F(€F(L)) for all LeJtF. Let L' = t;F(L). Then, by what we proved just now, £F(L') is a complement of L' in JtF. But L is also a complement of L' = £F(L) and we have seen that L^£F(L'). This shows that L must
LOGICS ASSOCIATED
WITH HILBERT
SPACES
75
coincide with £F(L'). We see thus t h a t (jF is an orthocomplementation ofJ?F. We shall fix a nonzero vector w0 e W. Let !F be the class of all finite dimensional linear manifolds Fs^W such t h a t (i) w0 e F, (ii) dim F>3. By theorem 2.7 there exists, for any Fe^, an involutive anti-automorphism dF of K and a nonsingular symmetric 0 F -bilinear form <. , . } F on Fx F such t h a t < • , - X F induces £ F ; i.e., for Z, (5)
&•(-£) = {u : ue F, <jx,u)F = 0
for
allXG!}
and (6)
Oo,w 0 > = 1.
Moreover, (6) determines 0 F and <. , . } F uniquely. Suppose now t h a t Fl9 F2e3r with Fx c JF2. Then, for any L e JtFi,
fc1(I) = ^ n f 1 = { 6 > ( i ) } n ^ , This shows t h a t for any Z e ^ # F l , (7)
£Fl (L) = {u:ue
Flt (x,u}Fz
= 0
for all
x e £}.
In other words, the restriction of <. , . } F 2 to F1 x Fxis a, 6F2 -bilinear form which also induces the orthocomplementation f F . I n view of the normalizing condition (6), we m a y conclude at once t h a t
(8)
*'•
=
°F-
<*,y>F2 = <*,y>Fl
(x> y e ^ i ) -
Since any two elements of & are contained in a third element of J^, we easily deduce from (8) t h a t there exists an involutive anti-automorphism 6 of K, and a symmetric 0-bilinear form <. , . > on W x W such t h a t (i) 0=6F for all F e J ^ and (ii) the restriction of <. , .> to Fx F is ( . , . )F for all F e^. Lemma 4.2 follows a t once from this. We now resume our discussion of the lattice ££ of all closed linear manifolds of the Banach space V over D. We take V= W in lemma 4.2 and obtain an involutive anti-automorphism 6 of D and a definite symmetric 0-bilinear form <. , . > on V x V such t h a t (x,y) = 0 if and only if x±_y. We fix a nonzero vector v0 e V; then 6 and <\ , .> are uniquely determined by the condition (v0,voy = l. Let V* be t h e set of all continuous linear maps of V into D. Then V* is a vector space over D°, the division ring dual to D. For ueV, let j8tt be the linear m a p of V into D defined by
(9)
pu(x) = (x,u)
(xeV).
76
GEOMETRY
OF QUANTUM
THEORY
We shall also use the conjugate norm on F*; this is defined by (10)
||AII = sup \\{u)\
(A e V*).
\\u\\
I t is well known t h a t V*, under this norm, is a Banach space over the division ring D°. Lemma 4.3. pu e V* for any u e V. Moreover, u —>• fiu is an additive isomorphism of the abelian group V onto the abelian group V*. Proof. I t is a well known fact (cf. Bourbaki [2]) t h a t a linear m a p A of V into D is continuous if and only if the linear manifold (11)
ZA = {x:xeV,
X(x) = 0}
is closed in V. Let now A = j8u, and write Zu for ZA, if and only if
ueV.liyeV,yeZu
Zu = (D-tO 1 .
(12)
B u t (D-u)1 e ££ and is therefore closed by assumption. This proves t h a t fiu e V*. The m a p u-> flu is clearly additive. If, for ue V, fiu = 0, then (uius)=pu(u)=0; this implies t h a t u = 0. I n order to complete the proof of t h e lemma it remains to prove only t h a t u^/3u maps V onto F*. Suppose A e F * and # 0 . Let ZA be defined by (11). Then ZA e & and ZAT^ V. Therefore ZAL^0. Let u be a nonzero vector in ZKL. Since (x,uy = 0 for all x e ZA, f$u vanishes over ZA; and since u^O, j8 M ^0. Since ZA and Zu are both of codimension 1 and ZAciZu, we infer t h a t ZA = ZU. Let xQ e V be such t h a t X(x0) = 1. Since x0 <£ ZA, xQ $ Zu and hence Pu(xo) = <^o^> ¥* 0. Let c = (xQ,u)~1
and let us write u0 = ceu. Then for x e V, <X,UQ)
=
<X,U)C.
This shows t h a t PUo(xQ) = 1 = X(x0) and implies t h a t p = A. The main question at this stage is whether 6 is continuous. The nontrivial case is when D = C and we handle it first. I t is at this point t h a t essential use is made of the infinite dimensionality of V. Lemma 4.4. Let W be a complex Banach space of infinite dimension. Then one can find a sequence of vectors xl9 x2, • • • e W with the property that, for any bounded infinite sequence {cn} of complex numbers, there exists a continuous linear functional A on W such that (13)
X(xn)=cn
(W=l,2,.-.).
LOGICS ASSOCIATED
WITH HILBERT
SPACES
77
Proof. Let || • || be the norm on W. Let W* be the space of all complex valued continuous linear functionals on W. We define x± e W to be an arbitrary vector of norm 1 and choose xx* e W* such that
The celebrated Hahn-Banach theorem ensures that such an xx* e W* exists. We shall now define by induction two sequences {xn} and {xn*} in W and W*, respectively. Suppose xlt x2, • • •, xn e W and xx*, • • •, xn* e W* have been already constructed so that xi*(xj) = 8 y (Kronecker delta) (14)
(i 1)
1^*11 = 2" "
(i,j = 1, 2, • • •, n),
(i = 1,2,...,»).
Since W is infinite dimensional, there is a nonzero vector xn + 1 in W linearly independent ofxlf---,xn such that
Using the Hahn-Banach theorem once again, we can find x*+ x e W* such that x*+ i{xn + 1)-=^0, x*+ 1(a;1) = 0, i <w. Replacing #n + x and x*+ ± by suitable multiples, we may assume that \\x*+ ± || = 2 ~ n, a;*+ 1(xn + x) = 1. The sequences xl9 x2, - - -9 xn + 1 and xx*, • •, x*+1 then satisfy (14) with n replaced by n + 1 . By induction it follows that there exist infinite sequences {xn} and {xn*} such that xne W for all n, xn* e W* for all n, and *i*(*y) = 8 tf
(*,j =
|*i*| = 2 " « - »
1,2,..-),
(* = 1,2, • • •).
We claim that the sequence {xn} has the properties asserted in the lemma. In fact, let {cn} be any bounded sequence of complex numbers. For x e W, let X(x) be defined by (16)
\(x) = 2
onxn*(x).
n= l
The inequality |a;n*(a?)| < ||rrn*|| • \\x\\ = 2-(n""1)||a?|| shows that A is well defined and that x -> X(x) is a continuous linear functional on W, i.e., A e If*. Clearly, \{xn) = cn for all w. We now resume our discussion of V and 0. Lemma 4.5. If D = R, 6 is the identity; if D = C, 6 is the conjugation; and ijD = H, 6 is the canonical conjugation. Proof. Let F be any linear manifold of finite dimension > 3 containing v0. The map L -^L^ n F is, as we saw in the proof of lemma 4.2, an orthocomplementation in the geometry of subspaces of F and is, moreover,
GEOMETRY OF QUANTUM
78
THEORY
induced by <. , . )F (the restriction of <. , . ) to F x F). From theorem 2.8 we deduce t h a t 0 is the identity if D = R and is the canonical conjugation if D = H. Suppose now t h a t D = C. Clearly, as (X,X)F = 0 for am XEF implies t h a t # = 0, 6 is not the identity. 6 will therefore have to be the conjugation in this case, provided we show t h a t it is continuous. Since 6 is additive, this reduces to showing t h a t if cn - > 0 in D, {cne} is a bounded sequence. Suppose then t h a t D = C and t h a t 6 is not the conjugation. Then there exists a sequence {cn} in D such t h a t
Since dim F = oo, we can select a sequence {xn} in V such t h a t it has the property described in lemma 4.4. Let j8n be the element x->{x,xn) of F * (cf. lemma 4.3). Since \cne\ ^ o o , we may, by passing to a subsequence if necessary, assume in addition t h a t for all n,
(18)
K'| >H|j8 n ||.
By lemma 4.4 we can find a A e F * such t h a t X(xn) = cn for all n. Evidently, A7^0. Moreover, it follows from lemma 4.3 t h a t there is a u^O in V such t h a t X = f3u. Then (u,u)^0 and (xn,u} = cn for all n. If (19)
Zn =
(UyUy-^-Cn'^,
then A(Zn) = <2n>W> =
0
for all n. But then, as <. , . ) is symmetric,
(u,zny = o for all n, i.e.,
(20)
«u,uye) - \u,uy - ( O - \u,xny = o.
Since K O " 1 ^ , ^ ) ! < \cn°\ - ^ I f t J \\u\\, (18) implies t h a t (cne)-\u,xn}->0
(n->co) e
and hence we infer from (20) t h a t (u,uy((u,uy )~1=0, which is impossible as u^0. This contradiction proves t h a t 0 must be the conjugation and completes the proof of lemma 4.5. Lemma 4.6. f$(u -> f$u) is a continuous map of V onto V*. Proof. Since we may obviously consider R as a subfield of D as well as D°, V and F * m a y both be regarded as Banach spaces over R, denoted by ^ R a n d ^ R * J respectively. From lemma 4.5 we conclude t h a t fi(u —> j8u) is an R-linear map. As /3 is denned on the whole of F R , we m a y use the closed graph, theorem to reduce the proof of the continuity of j3 to the proof t h a t j8 is a closed map. Let {xn} be a sequence in F R such t h a t
LOGICS ASSOCIATED
WITH HILBERT
SPACES
79
xn -> x and px ~> A. We must prove that \ = f$x. Since /? is onto, we may assume that \=f$u for some ue V. From the formula (10) for the norm in V*, it follows that fi Xn(v)-> fiu(v) for all veV. Since, by lemma 4.5, 6 is continuous, we have, for v e V, (x,v) = lim (xn,v} n
= lim «v,xnye) n
= (v,u}e = (u,v). This proves that x = u or /3X = X. We have thus proved lemma 4.6. Lemma 4.5 shows that (. , .} is an inner product on V xV. Thus, on defining ||2J|02 = <(z,z>, we obtain a norm on V. V is a pre-Hilbert space with respect to <. , . >. Lemma 4.7. V is a Hilbert space under <. , . >. Proof. We must prove that V is complete under || • ||0. Let {xn} be a sequence in V such that \\xn — xm\\02 ~> 0, n, m -> oo. Then {y,xn — xm} -> 0 as n, m -> oo for each y e V. Let A(?/) = limn (y,xn}. Then A is a linear map of V into D. Since f$x e F R * for all n, we ma}^ apply the well known uniform boundedness principle to F R , a Banach space over R, to conclude that Ae V*. By lemma 4.3 there exists a ze V such that A=&. We claim that \\xn — z\\0 -> 0. Let e > 0 be arbitrary and N. be such that \\xn-xm\\0
<e
for n, m>N. Then, for y e V, \
(w>m ^
N
)
and hence \
(n> N),
which shows that \\xn — z\\0<e for n>N. Lemma 4.8. The topologies induced on V by || • || (the original norm) and || • || 0 coincide. Proof. Let y, xlf x2, • • • be elements in V. We must show that as n -> oo, \\xn~y\\o -** ^ ^ a n ( i o n ly ^ ll^n —2/|| -^ 0- Since F is a Banach space over R under both || • || and || • ||0, it is sufficient, in view of the open mapping theorem, to show the implication in one direction only. Let then \\xn — y\\ -> 0. Since the map j8 from V to F* is continuous, fiXn — f$y -> 0 in F* so that (21)
sup II 2 II £ 1
\(z,xn-y)\-+0
(n^co).
80
GEOMETRY
OF QUANTUM
THEORY
Now, \xn — y\ < 1 for all sufficiently large n and hence we infer from (21) that
\\xn-y\\o2 =
which is the relation required. We can now complete the proof of theorem 4.1. V is a Hilbert space under <(.,.> and the topology induced by the corresponding norm has been shown to coincide with the original topology for V. Consequently, the same linear manifolds are closed in both topologies. Let |(Jf -> £(M)) be the orthocomplementation of the Hilbert space V. Now, (y,x} = 0 if and only if x _L y, and hence for any M e&,ye M1 if and only if y e i(M). This proves that JL and £ coincide. The proof of theorem 4.1 is complete. Theorem 4.1 shows that a logic may be asserted to be standard as soon as it is identified as a member of an apparently more general class of lattices. We have proved theorem 4.1 to emphasize our point of view that the important role played by standard logics is not due to accident but stems from deeper mathematical reasons, and that theorems such as the present one lead to a clarification of some of these reasons. This said, the stage is set for the detailed analysis of the standard logics and we proceed to do it now. 2. THE STANDARD LOGICS: OBSERVABLES AND STATES We now proceed to describe the structure of the set of observables and the set of states associated with a standard logic and examine the manner in which the general concepts of Chapter III specialize. Throughout this section we write D to denote one of R, C, or H. Let 3f be a separable infinite dimensional Hilbert space over D. We write ££ for the logic o$f(^f ,D) of all closed linear manifolds of J f (with its canonical orthocomplementation). For any closed linear manifold M of ^ , let us write PM for the orthogonal projection on M. If f(E ->f(E)) is any observable associated with the logic j£f, then the map E -^ PnE) is a projection valued measure^ based on ^(R 1 ). Conversely, it is obvious that to any projection valued measure P(E -> PE) based on J^R 1 ) there exists a unique observable/(2£ -+f(E)) such that PE = PnE) for all real Borel sets E. •f The concept of a projection valued measure is nothing new; it is a standard tool of spectral theory. For complex Hilbert spaces we refer the reader to sections 35-38 of Halmos [2]. For the real case very few changes have to be made. For the quaternionic case and for projection valued measures defined on the real line, the theory is essentially the same, since the real field is the center of H . We shall assume the basic facts concerning projection valued measures as known.
LOGICS ASSOCIATED
WITH HILBERT
SPACES
81
Lemma 4.9. Let Ml9 M2 e J^. Then M1 <—> M2 if and only if the projections PMi and PM2 commute. Proof. Let Mlf M2 e j£f. Assume first that Mx <-» M2. Then there exists mutually orthogonal elements Nl9 N2, N of 3? such that Mi = Ni-\-N (i = l, 2). Clearly, PMi = PNi+PN, and one is led immediately to the equation PMiPM2 = PMiPMi. Conversely, let PMi and PM2 commute and let P = PMiPM2. Then P is a projection. If we write Q{ = PMi-P, then it is easily verified that Qt is a projection (£ = 1,2) and PQi = QiP = 0, Q1Q2 = #2$i = 0.1{N l9 N2, and N are the closed linear manifolds which are the ranges of Ql9 Q2, and P, respectively, then it follows easily that Nly N2, and N are mutually orthogonal and Mi^=NiyN (i = l, 2). This proves that M1 <-> M2. Lemma 4.10. Let X be a set and & a o-algebra of subsets of X. Let P(E -> PE) be a projection valued measure in £F based on & and !F the algebra (over R) of all real valued bounded measurable functions on X. For ge^ let ||^||=sup{|gr(a;)| :xeX}. Then, for any ge^, there exists a unique, bounded, self-adjoint operator Bg on #F such that
g(x)dvUtV(x)
(u, v e Jt?)9
(22) J Bg =
gdP,
in symbols I
where vUfV is the J)-valued measure on & defined by vUtV{E) = (PEu,vy. The spectral measure of Bg is the projection valued measure E -> Pg~ iiE). One has: (23)
\\Bg\\ < \\g\\
(gelF).
The correspondence g -> Bg is a homomorphism from the algebra 3F into the algebra (over R) of all bounded operators on Jf. In particular, B9l and Bg2 commute for all gl9 g2e &'. Proof. The proofs are similar to the ones in section 37 of Halmos [2] and we omit them. In the quaternionic case we have to use the fact that the real field is the center of H. Theorem 4.11. Let D be one of R, C, or H and ££ the logic of all closed linear manifolds of #f\ For any observable f(E->f(E)) associated with
82
GEOMETRY
OF QUANTUM
THEORY
££\ let Af be the self-adjoint {not necessarily bounded) linear operator with spectral measure E - > P / ( E ) . Then (24)
f->At
is a one-one correspondence between the set of all observables associated with ££ and the set of all self-adjoint operators on Jf. The observable f is bounded if and only if the operator Af is bounded. Two bounded observables fx and f2 are simultaneously observable if and only if the (bounded) operators Afl and Af2 commute. Iff is any bounded observable and p is a polynomial with real coefficients (in one variable), then the operator corresponding under (24) to p o f is p(Af). More generally, if fl9 • • • , / r are bounded observables, any two of which are simultaneously observable, and if p is a real polynomial of r real variables, then the observable ^ ° (A, ...,/r) (cf. theorem 3.17) has the corresponding operator p(Afi,..., Afr). Proof. The first assertion concerning the correspondence / - > A f is an immediate consequence of the spectral theorem. Next, an observable / is bounded if and only if for some compact set K £ R 1 , f(K) = 3tfp. This is possible if and only if PnK) = / , which in t u r n is well known to be equivalent t o the requirement t h a t Af be bounded. Let now fx and / 2 be bounded observables. Then fx and f2 are simultaneously observable if and only if fi(Ei) *~* 12(^2) f ° r an* pairs of Borel sets E1 and E2. From lemma 4.9 and standard spectral theory, it now follows t h a t A <->/2 if and only if Af and Af2 commute. We now" come to the last pair of assertions. Since the first of these is a special case of the second for r = 1, we shall confine ourselves t o the proof of the second. Let A , • • • ,fr be bounded observables such t h a t fi<->ft for all i, j = l, 2, • • •, r. By theorem 3.17 there exists a a-homomorphism F ->f(F) of &(Rr) into 3? such t h a t for any i (l t{ of R r onto R 1 . Let PF = Pf{F). Then F -> PF is a projection valued measure on «^(R r ). Since the/,- are bounded observables, there exists a compact set K c Ri such t h a t fi(K) = Jt? for i=l, 2, • • -, r. We now apply lemma 4.10 by taking X = Kr, @ = &(Kr). For any E e @(K), Pni-\E) = PftE\ and so, writing VuAF)
=
vlv(E)
=
(P^u,v),
we get (Bniu,vy
=
TT^!, t2, • • ., tjdv^fa, JKr
= J tdvljt) =
•••,*,)
LOGICS ASSOCIATED
WITH HILBERT
SPACES
83
Thus, Bni = Aft, i = 1, 2, • • •, r. Since g -> Bg is a homomorphism, we have, for any real polynomial p of ^, • • •, tr, Bp=p(BXl,...,BMr)=p(Afl,...,Afr). But the self-adjoint operator Bp has the spectral measure E -> Pp- i(E) by lemma 4.10. Bp therefore coincides with the operator corresponding to the observable E ->f(p~1(E)) which is precisely p o (flt • • -,/ r ). This proves the theorem completely. Remark. That f1 <—>/2 is equivalent to the equation AflAf2 =Af2Afl is a celebrated result of von Neumann. The fact that p o (fl9 • • -,/ r ) corresponds to the operator p{Afl, • • •, ^t /r ) is remarkable since it links the algebra of functions of fl9 • • • , / r with the algebra of operators on Jf. We now take up the problem of determining the states of the logic j£f. Our basic result is the theorem of Gleason [1] which asserts that every state of ££ can be described in a canonical way by what is usually known as a density matrix. We shall now proceed to a detailed discussion of this theorem. To any vector uetf? with \u\ = 1, we associate the mapping pu of J&? into the set of nonnegative real numbers by defining (25)
pu{M) = (PMu,u} = ||PMi*||2
(Me&).
2
If M=Jf, pu(M)=\\u\\ = l, while pu(0)=0. If Ml9 M2, • • • is a sequence of elements of ££ such that Mi\_Mj (i=£j) then, writing M = Vt Mu we have : pu{M) = \\PMu\\2
= 2 11****11* i
i
In other words, pu is a state of ££. It is clear that if c e D and \c\ = 1 , Pcu=Pu> The general description of states depends on the concept of an operator of trace class. On complex Hilbert spaces they are defined and treated in many books on Hilbert space theory (cf. Dunford and Schwartz [2], for example). The treatment in the real and quaternionic cases is similar. A bounded operator A of $f is of trace class if, for any orthonormal basis {ej, the series 2 |<4«f,«f>| < oo; i
then the sum tr(^) = ^ i
<MA>
84
GEOMETRY
OF QUANTUM
THEORY
exists for any orthonormal basis {et} and is independent of the basis used; it is called the trace of A. If A is of trace class and B is any bounded operator, AB and BA are of trace class; and moreover, tr(AB) = tr(BA). If A and B are of trace class, so are a A +/3B where a, j8 are in the center of the division ring over which J f is defined. We shall assume the reader to be familiar with the basic properties of these operators. The states of the form pu(u e Jf, ||w||=l) can also be described in another way. Let U be the projection on the one-dimensional linear manifold D u. Then, for any bounded operator A, AU is a bounded operator, which is of trace class, since U is (trivially) of trace class. Let {en} be an orthonormal basis for Jf7 with u = e±. Then tr(AU) = tr(UA) (26)
=^(AUeiA> i
=
(Au,u).
In particular, (27)
pu(M) = tr(PMU).
Suppose that %, u2i • • • is any sequence of vectors in #F with H^H = 1 for all i. Let alf a2, • • • be a sequence of nonnegative numbers with «i +«2 + • * • = L Let Ut be the projection on the one-dimensional manifold D-iv Then
(28)
^ =
2 ^ i
is a well-defined, bounded, nonnegative, self-adjoint operator of trace class and (29)
tr(*7) = 1.
If M e <£?, we have tr(PMU) = ^at
tT(PMUt)
i
= 2a^twi
Thus the function pv defined by (30)
pu(M) = tr(PMU)
is a state of j£? and in fact is the state 2i ciPut' Conversely, let U be a bounded, self-adjoint operator such that (i) U is nonnegative, i.e., (Ux,x}>0 for all xeJf, and (ii) U is of trace class and tr(U) = l. Then,
LOGICS ASSOCIATED
WITH HILBERT
SPACES
85
the function pv defined by (30) is a state of ££\ We call the operators U which are bounded, self-adjoint, nonnegative, and of trace class, von Neumann operators. The question naturally arises whether every state of J&? is of the form pv for a suitable von Neumann operator U of trace 1. It was Gleason [1] who first proved that this is indeed so. His result, which is a cornerstone of the mathematical foundations of quantum mechanics, is undoubtedly one of the most profound theorems in this subject. Gleason's theorem is quite complicated to prove and we need to develop a number of technical lemmas. We begin with the concept of a frame function. Let $ be the unit sphere of Jtf, i.e., (31)
g = {x '.xeJtf, ||*|| = 1}.
A real valued function/(x ->/(#)) defined on & is a frame junction if (i) J (ex) = f(x) for xei and c e D with \c\ = 1, (ii) there exists a constant W such that 2n/( e n) — W f° r a n v orthonormal basis {en} of 3tf (recall that J f is separable). The constant W is called the weight of/. Suppose that U is a bounded, self-adjoint operator of trace class. Let us define the function fv by (33)
fu(x) =
(xet).
Then/^ is obviously a frame function whose weight is tr U. A frame function / is called regular if there exists a bounded, self-adjoint operator U such that (34)
f=U
Clearly, if/ is regular, there exists only one bounded, self-adjoint operator U such that (34) holds; for, if U and V satisfy (34), <(*7- V)x, x} = 0 for SbllxeJf? and hence U— V = 0. Suppose that /x is a state of ££\ For any vector x e $ let us define (35)
Ux) = pQt-x).
Then/^ is a nonnegative frame function of weight 1. If U is a von Neumann operator of trace 1 and p=Pu t n e state defined by (30), a simple calculation shows (36)
f, = fu,
i.e., fu is regular. Conversely, if fi is a state of ££ and if we assume that
86
GEOMETRY
OF QUANTUM
THEORY
fu =fu for a bounded, self-adjoint operator, then it can be shown t h a t U is a von Neumann operator of trace 1, and fi=pv. I n fact, as
2
n
so t h a t U is of trace class and has unit trace. This shows t h a t the states fM and pv coincide for all one-dimensional subspaces of Jt? and hence coincide for all elements of ££\ I n other words, Gleason's theorem is entirely equivalent to proving t h a t every nonnegative frame function is regular. We shall first examine t h e special case when D = R and when J f is finite dimensional. If #? is the real line, then the only frame functions are constants. If #? is t h e plane R 2 , then it is obvious t h a t a frame function can be arbitrarily defined on any quadrant of the unit circle and hence need have no special properties. Lemma 4.12. Let &? — R2, the plane (D = the real field) and let $ be described as (37)
£ = {(cos 0, sin 0) : 0 < 0 < 2n}.
Then the function f : (cos 6, sin 6) - ^ cos nQ, n being an integer, is a frame function if and only if either n = 0 or n = 2 (mod 4). Proof. I n order t h a t / be a frame function we must have, for some constant W, (38)
cos7i0 + cos ( ™ ( 0 + | ) ) = W
for all 6. Differentiating with respect to 6, we have: M sin for all 0, i.e., ...
-./_
n7r\
.
nir
.,
n\ (sin nd)[ 1 + c o s — I + s m — 2 cos ^ 0 ^ = 0. This is possible if and only if either n = 0 or cos(mr/2)= — 1 , i.e., if and only if n = 0 or n = 2 modulo 4. Lemma 4.13. Let J f = R 3 be the three dimensional real Euclidean space. Then every continuous frame function on the unit sphere
LOGICS ASSOCIATED
WITH HILBERT
SPACES
87
Proof. The proof makes use of harmonic analysis on $ (cf. Weyl [1], pp. 60-63,142-146). Let # be the set of all real valued continuous functions on $ and G the group of all orthogonal transformations of R3 which have determinant + 1 . For any element h of ^ and any a e G, we define ha by ha(x) = hia-^)
(39)
(xeg).
A linear manifold Ji c <$ is said to be invariant if ha e J£ for all a e 6r whenever & / e ^#. Then, according to classical spherical harmonic analysis, one has the following description of <S\ (1) for each integer n = 0, 1, • • • there is a 2n + 1 dimensional subspace Fn of ^ which is invariant and no proper linear submanifold of which is invariant, i.e., Fn is irreducible under the action (39) of the group G; (2) an element h e <€ belongs to !Fn if and only if it is the restriction to $ of a polynomial p of x, y, z (the coordinates on R3) with the properties (i) p is homogeneous in x, y, z of degree n, (40)
,..v / a 2 a 2
(u)
d2\
W 2 + ^ 2 + ai 2 F = 0;
(3) every continuous function on $ is the uniform limit of a sequence of functions each of which is a finite linear combination of elements of the «^"n; and finally (4) if IF is any invariant linear manifold, closed under the uniform convergence topology of # , and nlt %, • • • (0
(41)
^= (2^)~>
where 2 denotes the algebraic sum, and the bar denotes the closure. Let now IF be the set of all continuous frame functions on $. Evidently G-transforms (cf. (39)) and uniform limits of continuous frame functions are continuous frame functions. Thus the set F is a closed linear manifold in ^ which is invariant. Therefore, there exist integers nx, n2, • • • with the properties described in (41). We now claim that if n is any integer other than 0 and 2, then, IF^IF. In fact, let n be any integer. We use cylindrical coordinates p, 6, z on $ denned by x = p cos 0, y = p sin 6, z = z. Then it is well known that d2
d2
dx^^df^d?
d2 _ d2
Id2
~
2+
Id 2
+
d2 +
Wp 'p d¥ }Tp d?'
88
GEOMETRY
OF QUANTUM
THEORY
One can easily check that the functions n P
cos n6, [pn-2(n-l)z2Pn-2]
cos {n-2)d
are homogeneous polynomials of degree n which satisfy the equation /a2
a2
d2\
Thus if J*"ncJ*-, then pncosnd and [pn-2(n-l)z2pn-2]-cos(n-2) d are frame functions. For any 0, (cos 0, sin 0, 0), ( — sin 0, cos 0, 0) and (0,0,1) form a frame of B3 so that the functions obtained by putting z = 0 are frame functions on the unit circle of the x, y plane. Thus cos nd and cos(w — 2)6 are both frame functions. By lemma 4.12 this is not possible unless w = 0 or 2. Thus we must have ^ ^ ( J ^ + J ^ ) - i . e . , J ^ c J ^ o + J ^ , since ^0 and ^ 2 are finite dimensional. We now claim that JF = ^" 0 +^2 and that every element of ^0 +^2 *s a regular frame function. In fact, ^0 consists of the constants which are regular. On the other hand, J^2 consists of the quadratic forms q(x,y,z) which satisfy id2
2+
d2\
d2
2+
W W ^)q= ' i.e., q(x,y,z) = ax2 + by2+cz2+ 2fyz + 2gzx + 2hxy with a+b+c = 0. Thus a function of ^ lies in ^0 +^2
^
an(
* only if it is of the form
x —> (Ax,x} (x is a vector in
LOGICS ASSOCIATED
WITH HILBERT
SPACES
89
which are orthogonal to q. If the latitude of q is 6, it can be shown that this great circle passes through only those points whose latitudes 9? satisfy — 6<(p<0. It is therefore appropriate to call this the EW great circle through q. The essence of our argument is contained in the next lemma. To motivate this lemma, let us consider in the usual coordinates x9 y, z the function (x,
y,z)->x2+y2.
This is a frame function which is nonnegative and vanishes at the north pole p = (0,0,1). It is constant on points with the same latitude. For any point q of the northern hemisphere, the points of Np other than q on the EW great circle through q, have latitudes strictly less than the latitude of qy and consequently this frame function reaches its minimum on the EW great circle through q, at q. The next lemma asserts that every frame function which is nonnegative has this property approximately. Lemma 4.14. Let p e $ and g be a nonnegative frame function which is constant on the equator opposite to p. Then for any q in Np we have: (42)
g(q) < g(x)+g(p)
for every point x on C, the EW great circle through q. Proof. We first assert that if r is any point of Np, then (43)
g(r) < k+g(p),
where k is the constant value which g takes on the equator. To see this, consider the EW great circle through r. If r' is one of the points at which this meets the equator, then r_[_r' and hence r, r' are two points of a possible frame. The frame function g being nonnegative, one has: g(r)+g(r') < W, where W is the weight of g. But if we take any two points a, a' on the equator which are orthogonal, a, a', p is a frame and, consequently, 2k+g(p) = W. Thus g(r) < W-g(r') < 2k+g(p)-k
< k+g(p).
Suppose now q is an arbitrary point in the northern hemisphere. The inequality just now established proves the lemma if x lies on the equator. For x below the equator, — x lies above the equator, and, since g(u) =zg( — u) for all u, it is enough to prove the inequality (42) when x lies on the EW great circle C through q and is also situated in the northern hemisphere.
90
GEOMETRY
OF QUANTUM
THEORY
Let y be a point on C, orthogonal to x, and lying in the northern hemisphere. We then have, if q' is a point where C meets the equator, g{x)+g{y) = g(q)+g(q') = g(q)+k and, since g(y)
by (43), one has: g{*) =
g(q)+k-g(y)
> g(q)-g{p)> which establishes the lemma. Lemma 4.15. With p and g as in lemma 4.14, let z be a fixed point in Nv. Let Mz be the set of points x on Np which have the property that for some point y of Np the EW great circle through x contains y and the EW great circle through y contains z. Then the interior of Mz is nonempty. Moreover, for any x of MZi one has: (44)
g(x)
Proof. Since g(x)
A ={(f,i ? ,0:0<J
0(^,£) <0},
then A is a nonempty open subset of Np. Suppose x lies in A. Then ip(x)<0 while the EW great circle through x meets the equator at a point at which ijj is positive. Since 0 is continuous, there must be a point y on the EW great circle through x at which 0 vanishes; thus x e Mz. In other words, i c M z . This proves the lemma. Lemma 4.16. With p and g as in lemma 4:A4:iletrj>0bea positive number such that g(p) < r/. Then there exists a point q of Nv and an open set M around q such that (45)
0 < sup g(u)- inf g(u) < 3^. ueM
ueM
LOGICS ASSOCIATED
WITH HILBERT
SPACES
91
Proof. Define b = miueN g(u). Then b>0. Choose zeNp such t h a t b
ueM
This proves t h e lemma. Lemma 4.17. Let f be any nonnegative frame function such that for some open neighborhood M of r we have: (46)
0 < s u p , / » - w£f(u) ueM
and r any point
< a.
ueM
Then for every point s of $ there is an open neighborhood Ms of s such that (47)
0 < s u p , / » - inf f{u) < 4a. ueMs
ueMs
Proof. Use coordinates so t h a t r appears as the north pole. Let t be any point on t h e equator opposite r. Since M is an open neighborhood of r, we can assume t h a t every point whose latitude is > (TTJ2) — d0 belongs to M, 60 being a suitable angle in ]0,7r/2[. Consider now a point u due south of t and latitude — J0 O . Then by continuity it is clear t h a t there exists an open set Mt around t with the following property: if k is a n y point of Mt, then, on t h e great circle through u and k, there are points orthogonal to u a n d k, respectively, which are situated in M. Let k, k' be two points of Mt, and C, C the respective great circles through u, k and u, k'. Let x, y and x', y' be pairs of points, respectively, on these two great circles such t h a t x, y, x\ y' e M, % A-U,
y _1_ &> x' _\_u,
y' J_ k'.
We then have: f(x)+f(u)=f(k)+f(y), f(x')+f(u)=f(k')+f(y'). Subtracting and rearranging, we have:
\f(k)-f{k')\ <
\f(x)-f(x')\+\f(y)-f(y')\
< 2a, so t h a t 0 < sup f{u)— inf f(u) < 2a. ueMt
ueMt
92
GEOMETRY
OF QUANTUM
THEORY
lis is any point of $, there exists a point t of $ such that t is on the equator opposite r and s is on the equator opposite t. Applying the preceding argument twice, we see that there is an open set Ms around s such that 0 < sup f(u) — inf f(u) < 4a. ueMs
ueMs
This proves the lemma. Lemma 4.18. Let f be a nonnegative frame function on
inf/fa) = 0. Let 7} > 0 and p a point such that f(p) < rj/2. Let or be the rotation around the z-axis with^) as north pole through an angle TT/2. Let the frame function g be denned by 9(x)=f(x)+f(ax). g is nonnegative, and at p we have: 9(P) < V> moreover, g is constant on the equator. By lemma 4.16 there exists a point q of the northern hemisphere and some neighborhood Mq of q such that 0 < sup g(u)— inf g(u) < 3^. ueMq
ueMq
By lemma 4.17 there exists an open set Mp around p such that 0 < sup g(u)— inf g(u) < I2rj. ueMp
ueMp
Since g(p)
13TJ.
ueMp
Since 0
for all u, we have: 0 < sup f(u) < 13^ ueMp
and hence we have: 0 < s u p / ( u ) - inf f(u) < 137?. ueMp
ueMp
LOGICS ASSOCIATED
WITH HILBERT
SPACES
93
Applying lemma 4.17 again, we see that every point of $ has a neighborhood over which the oscillation of / is at most 52^. If we choose 7/ = e/52, we are through. Lemma 4.19. Let ^ be a real separable Hilbert space of dimension at least 3. Then every nonnegative frame function f on the unit sphere $ of 34? is regular, Proof. Since every two-dimensional subspace L of J f can be imbedded in a three-dimensional subspace, it follows from lemma 4.18 that the restriction of/ to the unit sphere S'(L) of L is regular, and consequently there is a unique symmetric bilinear form on Lx L, say BL, such that (48)
BL(u,u) = f(u)
for all u G ${L). We shall now define a real valued function B on 3f x ^ as follows. Let x and y be two vectors of #F and L a two-dimensional subspace containing x and y. We define (49)
B(z,y) = BL{x9y).
Clearly, B(x,y) is well defined if x and y are linearly independent, since in this case L is unique; or if one or both of them is zero, since we then have BL(x,y) = 0 for any such L. If x and y were dependent and neither is zero, then they span a one-dimensional subspace. If Lx and L2 are two two-dimensional subspaces containing x and y, then both Lx and L2 are contained in a three-dimensional subspace, say M. Let C be a symmetric bilinear form on M x M such that
G(a,a)=f(a) for all vectors a in $(M). The restrictions of C to Lx x L± and L2x L2 are symmetric bilinear forms whose quadratic forms coincide with / on the unit spheres of L1 and L2, respectively. Since BLl and BL2 also have this property, we conclude that BLl{x9y) = C(x9y), BL2{x,y) = C(x,y), which proves that B(x,y) is well defined. Moreover, we have (50)
B(x,x)
=f(x)
for all unit vectors x in 3tf. We now claim that B is bilinear. The homogeneity of B as well as its symmetry involves only two vectors and so follows from the homogeneity and symmetry of BL. To complete the proof of the bilinearity (in view of the symmetry of B)> it is enough to show that B(x,y+z)
=
B(x,y)+B(x,z)
94
GEOMETRY
OF QUANTUM
THEORY
for any three vectors x, y, z. Let M be a three-dimensional subspace containing x, y and z. Let Lly L2, L3 be two-dimensional subspaces of M containing, respectively, x9 y; x, z; and x, y+z. Let C be a symmetric bilinear form on M x M such that C(u,u) =f(u) for all unit vectors u in M. Then we see at once that C(x,y) = .BLl(*,y),
O ( ^ ) = BL2(x,y)
and C(x,y+z)
=
BL3(x,y+z).
Since C is bilinear, we have: B(x,y+z)
= S L a (»,y+2) = C(x,y)+C(x,z) = =
BLl(xiy)+BL2(x,z) B{x,y)+B{x,z),
which proves the bilinearity of B. Since 0 < -B(a?,a;) < 1 for all x with ||a;|| = 1, x, y-> B(x,y) is a bounded bilinear form. Thus there exists a bounded, self-adjoint operator U such that for all x, y in J f -B(a,2/) = <JJx,y>, f(x) = <J7o?,a?>. Thus we have proved that / is regular. We now proceed to the case of Hilbert spaces over D, where D may either be the complex number field or the quaternion division ring. Let 3tf be a Hilbert space over D. A closed R-linear manifold S of Jtf* is called completely real if the inner product <(.,.> in J f x MP takes real values on S xS; equivalently, if and only if there is an orthonormal set {ej) such that 8 is the closure of the real linear combinations of the ey. Lemma 4.20. Let 01 be a real separable Hilbert space, and f0 a regular nonnegative frame function of weight W. Then for all a,be& with
HI = l*I=i, (51) |/o(«)-/o(6)| * 2JF||o-&|. Proof. There is an operator T, of trace class and self-adjoint such that (Tc,c} =f0(c) for all unit vectors c. Then |/o(«)-/o(6)| = \
LOGICS ASSOCIATED
WITH HILBERT
SPACES
95
But 0<(Tc,c}< W for all unit vectors c so that \\T\\< W. Thus \fo(a)-f0(b)\
<
2W\\a-b\\.
Lemma 4.21. Let J#* be a Hilbert space over D of finite dimension n. If f is a nonnegative frame function which is regular on each completely real subspace, then f is regular. Proof. We prove this by induction on n. For n = 1, the lemma is trivial. Let n>\. Let d = suj)m = 1f(u) and {xn} a sequence of points on $, the unit sphere of ^ , such that f(xn) -> d. We may assume, in view of the compactness of the unit sphere, that xn->yy \y\ = 1. We first show that f{y) = d. Let An=||-1<2/,^n>. Since xn-*y, A„->1. The inner product
\f(y)-d\ < \f(y)-AKxn)\ + \f(Kxn)-d\ <2W\\y-Xnxn\\+\f(xn)-d\, since /, by hypothesis, is regular when restricted to the completely real subspace spanned by \nxn, y, and f(xn)—f(Xnxn) (|An| = l). This shows that f(y) = d. Let us now extend the function / to a function F defined on all of 3^ by setting (52)
F(v) = 1
lH2/(IN"H
v*o
Clearly for c e D, (53)
F(cv) = \c\2F{v).
We now assert that if u is any vector of 3tf orthogonal to y, then (54) F(cy+u) = \c\2d + F(u) for all c E D. Let L be the two-dimensional real linear manifold spanned by y and u. This is completely real, / i s regular on its unit sphere so t h a t / i s the restriction to L of a nonnegative definite quadratic form Q. But y is the point at which the maximum of the quadratic form is reached and u is orthogonal to y in L. Thus the matrix of the quadratic form Q is diagonal with respect to the basis {y,u} of L, and consequently we have, whenever r and r' are two real numbers, F(ry+r'u) = \r\2d + \r'\2F(u). It is to be noted that u here is an arbitrary vector orthogonal to y.
GEOMETRY
96
OF QUANTUM
THEORY
Suppose now that v is any vector orthogonal to y and c^O an element from D, not necessarily real. We have: F(cy+v) =
F^y+c-H)) a f
= |c| J (y+c- 1 v) = \c\2F{y +
\c\-i\c\c^v),
x
but u= \c\c~ v is a vector orthogonal to ?/. Thus j ^ + c " 1 * ; ) = d + \c\-2F(\c\c-1v) = d + \c\~2F(v), proving that
F(cy+v) = \c\2d + F(v).
The restriction of / to the orthogonal complement Jf* of y is a frame function, which, on Jf, has the same properties as /. Suppose we assume the present lemma to be true for all Hilbert spaces of dimension k < n. Then F restricted to J f is regular by the induction hypothesis and hence there is an orthonormal basis {el5 • • •, en_x} for J f and real numbers ^i> * * • 9 dn_! > 0 such that F(ciei + • • • +c tt _ 1 e n _ 1 ) = {c^d, + • • • + |cn_1|2<Jn_1 for all cl9 • • •, c n _! G D. By the proof given in the previous paragraph, this means that F{cy+c1e1 + > • • +c n _ 1 e n _ 1 ) = |c|2d + |c 1 | 2 ^ 1 + • • • +|c n _ 1 | 2 d n _ 1 , which shows that / is the restriction to the unit sphere of the regular frame function fUt where U is the operator denned by Uy = dy, Uej = dfi^ j = 1, 2, • • •, n — 1. This proves the lemma. Lemma 4.22. Let 3f be a separable Hilbert space over D which is one of R, C, or H of dimension at least 3 and f a nonnegative frame function on the unit sphere $ of 3f. Then f is regular. Proof. Since the dimension of 3tf is at least 3, it follows that the restriction of / to any completely real subspace is regular. Only the case dim «^ = oo remains. Lemma 4.21 is applicable to any finite dimensional subspace of 3tf\ Thus for any finite dimensional subspace 8 of Jf7 there is an operator Us which is an endomorphism of 89 bounded, self-adjoint, and > 0 such that f(x) = (TJsx,x)
(xe8,
|| a; || = 1).
The uniqueness of Us implies that if Sx S^82i( USi x,y} — < US2x,y} if x,y e #!. This implies in the usual way that there exists a symmetric semi-bilinear form on 3/F x 3tf\ say, B such that
B(z,x)=f(z)
LOGICS ASSOCIATED
WITH
HILBERT
SPACES
97
for all x with ||a|| = l. Since 0 < B(x,x) < 1, there exists a bounded, nonnegative, self-adjoint operator U such that <JJx& = B(x,x) = f(x). This proves the lemma. Theorem 4.23 (Gleason [1]). Let Wbe the convex set of all von Neumann operators of unit trace on a separable Hilbert space #F of dimension at least 3 (possibly infinite) over D. Then the map (55)
U-^pu
(UelT)
(where pv is defined by (30)) is a convex^ isomorphism of iV onto the convex set £f of all states of the logic 3?. A state p e £f is pure if and only if p=pu for a unit vector u of Jf, pu being defined as in (25). pu=pv if and only if for some c e D with \c\ = 1, v = cu. In particular, the pure states of 3? are in natural one-one correspondence with the rays of 3^. Proof. That U ->pv is an isomorphism of iV onto Sf is immediate from lemma 4.22. Next we shall obtain a description of the pure states of ££. For any vector uetff with \u\ — 1, let pu be the state of ££ denned by pu(M) = \\PMu\\2 (Meg). We assert that pu is a pure state. Let U0 be the projection on the onedimensional space D-w. Clearly pu(M) — tr(PMU0) so that Pu = Pu0To prove that pu is pure, let pl9 p2 be states of JSP such that (56)
pu = apl + (\-a)p2,
where 0 < a < l . Since U -> PV is a one-one convex map of W onto the set of all states, there are (unique) elements Ulf U2e^r such th.dX pj=pUj (j = l,2) and hence (57)
U0 =
aU1+(l-a)U2.
7
Since <£7ya;,a;>>0 for all x e Jf , j = 0, 1, 2, and since (Uox,xy = 0 if x is orthogonal to u> we conclude from (57) that for xj_u, (Ujx,x}=0 (j== 1, 2). Now Vj is > 0 and hence there is a bounded self-adjoint Yi such that Vj2=Uj and we have <^Ujx,xs)= ||F;a:||2. This implies that Fy# = 0 for all x orthogonal to u and, consequently, that UjX = 0 for all x orthogonal to u. Since Uj is self-adjoint, this implies that Uj = cjU0 for some constant t If A and B are convex subsets of two vector spaces over the real number field and if u is a map of A into B, we say that u is convex if u (ax + by) = au(x) -f bu(y) for x, y eA, 0 < a , 6 < 1 and a -f 6 = 1.
98
GEOMETRY
OF QUANTUM
THEORY
c,(j = l , 2 ) , and c, = l as tr(C/i) = l . Therefore U1=U2=U0, which proves t h a t pu is a pure state. Conversely let UQ e iV be such t h a t ^C7o is a pure state. If al9 • • •, an, • • • ( > 0 , a x -fa 2 + • • • = ] . ) are the eigenvalues of U0, and ux, u2, • • •, un, • • • corresponding eigenvectors of norm 1, then (58)
3>c0 =
^anpUn. n
As pv is pure, (58) implies t h a t all but one of the an must be 1, the rest being 0. This shows t h a t pUo =pUj for some j . Remark. From theorem 4.23 we can obtain at once an important characterization of the principle of superposition of pure states. Let {ua} be a family of unit vectors of J f and let us write pa for pUa. Let p=pu be a pure state and let M be the smallest closed linear manifold containing all the ua. If p is a superposition of the collection {pa}, then p(M1) = 0 as pa(M1) = 0 for all a. Hence ue M. On the other hand, if u e M, then for any N e ££ having the property t h a t pa(]^) — ^ for all a, p(N) = 0, since a n y such N is contained in M1. In other words, pu is a superposition of the family {puJ if and only if u lies in the smallest closed linear manifold of J f containing all the ua. Thus, if we make the assumption t h a t the logic of the physical system © is standard, the pure states can be identified with the points of an infinite dimensional projective geometry with the principle of superposition translating into the usual concept of linear dependence of points. Our next object is.to discuss the nature of the statistics of one or more observables associated with the standard logic «£?. We begin with t h e following theorem: Theorem 4.24. Let x(E - > x(E)) be an observable associated with J?, U a von Neumann operator on 3f of trace 1 and pU9 the state M - > tr(P M U) of ££\ Let Ax be the self-adjoint operator on $F which corresponds to x. If f is any real Borel function bounded on the spectrum of x, f o x is a bounded observable and (59)
StfoX\Pu)=tT(f(Ax)U).
Ifx is nonnegative, so is A x and then $(x\ pv) exists if and only if B = AX12 U112 is an everywhere defined, bounded operator of Hilbert-Schmidt class; in this case (60)
*(z\Pu)
=
tr(B*B).
If x is arbitrary, then x has finite variance in the state pv if and only if AXU112 is an everywhere defined, bounded operator of Hilbert-Schmidt class. In this case, (60) reduces to (61)
*(x\pa)
=
tv{Vll2AxU112).
LOGICS ASSOCIATED
WITH HILBERT
SPACES
99 M
In particular, if u is a unit vector in Jf and pu is the state M -> || P u ||2 of ££\ x has a variance in the state pu if and only if ue @(AX) ;*(* in this case, (62)
£(x\pu) =
Proof. If S is the spectrum of x and K = (/[$]) ~, the closure of/[$], then (/ o x)(K) = 3f; and as K is compact, this proves that y=f o x is bounded. The spectral measure of the operator Ay is the measure E -> Qf- iiE) where Q is the spectral measure of Ax. Thus Ay=f(Ax) in the usual sense of the functional calculus of self-adjoint operator on 3tf. This proves (59) at once. We next take up the more delicate questions involving possibly unbounded x. Let x be any nonnegative observable i.e., a;([0,oo[)=Jf7, associated with ££\ Then Ax is a nonnegative, self-adjoint operator. We denote by Ax12 the unique, nonnegative, self-adjoint operator C such that C2 = AX. Let U be any von Neumann operator of unit trace on Jf such that S(x\pv) < oo. Let Q(E-+QE) be the spectral measure of Ax and k(E) = tr(QEU). Clearly i([0,oo[) = 1 and
*(*\Pu) = J" tdk(t). If cpetf is any unit vector, (QEUll2q>,Ull2(Py
= k(Eyi
(t^fdk^t) < oo, /Jo; where k^ is the measure E ~> . This implies that ifs lies in the domain of Ax12. In other words, AXI2U112 is everywhere defined. Since Ax12 is closed and £71/2 is bounded, it is easily seen that Al.l2U112 is closed. Consequently, B — A]}2XJ112 is a bounded operator. To prove that B is of Hilbert-Schmidt class it is enough to prove that 2n II-^Pn II2 < °° f° r some orthonormal basis of 3tf. Let {
jn^nii^s^n^vir Jo
n
V' 2 ) 2 «M*)
= T an\ n
tdkn(t) JO
p oo
= < oo.
tdk(t)
\ &{L) is the domain of the linear transformation L.
100
GEOMETRY
OF QUANTUM
I2
THEORY
112
For the converse, if B = AX U is defined everywhere and is a bounded Hilbert-Schmidt operator, we have, using the above notation, tr(B*B) = ^ I I i W = 2 «n f" Mkn(t) < oo. n
n
Jo
But then, by Fatou's lemma,
i:
tdh(t) = g(x\pv)
< oo.
Jo
At the same time, the above argument proves that *(z\Pu) = tr(B*B). Next we consider the case of finite variance. We use the same notation. x has a finite variance in the state pv if and only if f "^ t2dk(t) < oo. Proceeding exactly as before, we conclude that if this condition is satisfied, D = AXU112 is an everywhere defined operator of Hilbert-Schmidt class. Note that, if a n > 0 ,
a
n(Ax
2 n:an >0
n
J - oo
= f
tdk(t).
J — 00
This proves (61). For proving the converse, let AXU1I2 = D be everywhere defined and of Hilbert-Schmidt class. We want to prove that f°° t2dk(t)
< oo.
J — oo
Write, for any Borel set
E^R1,
K(E) =
{QEu^
Then tTiU^QsU1'2)
k(E) =
k
= 2
nW
n
and
if
' " t*dkn(t) = 2 IMx^'Vnl 2 n <
00.
I t follows once again by Fatou's lemma that P° t2dk(t) < oo.
LOGICS ASSOCIATED
WITH HILBERT
SPACES
101
For the last part, let U be the projection on the one-dimensional space spanned by
q*(c;E)=
\\Q(c; E)
<7p(c; •) is a probability measure on ^(R 1 ). We say that the observables {x±, • • •, xk} have joint statistical distributions if for each unit vector q> eJtf', there exists a probability measure q^ on &(Rk) such that for all vectors c = (c1? • • •, ck) of Rfc and all E e^(R 1 ), (64)
q9(E(c)) = q9(c; E),
where (65)
E(e) = {(«!, • • •, y : cA + • • • + c ^
G #}.
fc
The sets E(c) are the "half-spaces" in R . Since any D-valued measure on &(Rk) is uniquely determined by its values on the sets of the form E(c), it follows that q9, if it exists at all, is unique. Obviously, the above sense is a quite conservative one in which we may speak of the statistics of the observables {xly • • •, xk}\ the probability measure q0 will then be the joint distribution of {x±, • • •, xk} under the state of J? determined by 9. We shall now prove the following theorem (Varadarajan [1]). Theorem 4.25. With the notations described above, {xx, • • •, xk} have joint statistical distributions if and only if the operators Aj(l<j
(66)
&«r\m = *m 1
for allEe&iR ) andj = l, 2, • • •, k, TTJ being the projection (tl9 • • •, tk) -> tt of Rfc into R1. If we define for a unit vector yetf, and E e@(Rk), (67)
q9(E) = ||P'<*V||9,
then (63) and (64) are satisfied. Thus {xu • • •, xk} possess joint distributions.
102
GEOMETRY
OF QUANTUM
THEORY
We now proceed to prove the converse. If {x±, • • •, xk} possess joint distributions, it is obvious that for i,j
(D = R)>
ft*.*- = Ifa+o'-q*-*') (68)
q9%9. = Ifa + v'-qv-v'+iq-iv
+ v'-iq-to-o')
(D = C),
Qw = i(2*+ *'-&>-*'+ 2 to-Jr
( D = Q)
We assert that for each M e^(R 2 ), the map ,> and hence from (68) we infer that q9.AE(*)) = < « c ; E)
< ||9||2,
0<^ w
it is clear that 99, q>' -> q^^M) is a bounded form. We can now use a well known theorem to construct a unique bounded self-adjoint operator PM such that fr^W
=
Clearly, 0
<^M9>9>'>
= 2
We note also, as qOt0(M)>O, that for M±^M2, PMI
LOGICS ASSOCIATED
WITH
HILBEBT
1
SPACES
103
L
B vanishes on Y and hence B = 0 on (X n Y) -B, being self-adjoint, therefore leaves X C\ Y invariant and for x e X n Y, (Bx,x}<\\x\\2. X Y Thus B
From lemma 4.26 we m a y conclude t h a t PE1*E2
< G I A Q2.
Replacing E2 by R 1 —E 2 we obtain PEIXR'-E2
< Qi A ( / - & ) .
The last two relations yield, on addition, (69)
Q,
A Q2+e! A (/-&).
However, Qi AQ2 and Qi A (I — Q2) are orthogonal projections which are both
A
(I-Q2).
This last equation leads quite simply to the conclusion t h a t Q1 and Q2 commute (cf. lemma 3.7). This theorem, taken together with theorem 3.9 may be regarded as giving a complete description of the circumstances under which the statistics of several observables may be regarded as arising from observables with joint distributions. We end these remarks with a mention of the concept of the quasiprobability distributions, introduced by Wigner [2] and studied extensively by Moyal [1]. Let xl9 x2, • • •, xk be bounded observables with Al9 • • •, Ak as their corresponding operators. We use the same notation as in theorem 4.25. For fixed
J R j f c exp{(-l) 1 / 2 (c 1 ^ + - - - + c A ) } ^ ( c ; - ) .
I t is easy to prove t h a t <$( •; 93) is a continuous function of c±, • • •, ck for each 99 e J f and |*(
104
GEOMETRY
OF QUANTUM
THEORY
We may therefore introduce its Fourier transform, in the sense of distributions of Schwartz, say (-;
3. T H E S T A N D A R D LOGICS: S Y M M E T R I E S At the focal point of the axiomatic development of modern physics lies the concept of automorphism or symmetry. We have already pointed out t h a t the momentum observables of a physical system are intimately related to one-parameter groups of automorphisms of the configuration manifold of the system. Moreover, we had described in Chapter I I I how the dynamical group of a physical system can be described b y a one-parameter group of convex automorphisms of the set of states. I t is therefore extremely important to examine the concept of an automorphism when the logic in question is a standard one. The main purpose of this section is to obtain a complete description of the automorphisms associated with a standard logic. We shall introduce, for this purpose, the notion of a symmetry of a Hilbert space Jf. As always, J f is separable and the field of scalars D is one of R, C, or H. A mapping T(x-+Tx) of Jtf* into itself is said to be a symmetry (of 3tf) if (i) T is additive, one-one, and maps Jf7 onto itself, and (ii) there exists a continuous automorphism 0 of D such t h a t T is 0-linear and, for all x, y e Jf7, (71)
=
<x,yy.
If D = R, then 6 is the identity and the symmetries are none other t h a n t h e unitary operators of J f . If D = C, then 6 is either the identity or t h e conjugation; in the former case, T is unitary, whereas in the latter case T is anti-unitary. If D = H, then there exists a c e D such t h a t |c| = l and de = cdc ~1 for all d e D; in this case, T(dx) =
(cdc-^Tx
and (Tx,Ty}
=
c<x,yyc-\
Equation (71) implies t h a t
\(Tx,Ty)\
=
\(x,y}\
I n the general case of (71) we shall call T 6-unitary. The symmetries of 3f form a group; if T} is 0 ; -unitary( j = 1,2), TX T2 is dx 0 2 -unitary and T5"1 is 0 ; - 1 -unitary. The unitary (linear) operators of Jf7 form a normal subgroup
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
105
%{£?) of the group of all symmetries. We shall write 8(Jf) for the group of symmetries. The map which associates with any 0-unitary T, the automorphism 6 of D, is a homomorphism of S(J4?) onto the group Aut(D) of all continuous automorphisms of D. Since the kernel of this map is ^ ( J f ) , we have: (72)
S(Jtr)l<%{Jf) ~ Aut(D).
Note finally that if T is a symmetry and c e D is such that \c\ = 1, cT is a symmetry. Theorem 4.27. Let J^f be a separable infinite dimensional Hilbert space over D. Let a(M -> Ma) be an automorphism of the logic ££\ Then, there exists a symmetry T of ^ such that (73)
Ma = T[M]
(M e &).
If T' is a symmetry such that (73) is satisfied for all M, then there exists a c e D with \c\ = 1 such that T' = cT. Finally if T is any symmetry of Jf7, M -> T[M] is an automorphism of the logic ££\ Proof. It is quite trivial to check that if T is a symmetry of ^f7, M -> T[M] is an automorphism of the logic j£f. Suppose now that a(M -> Ma) is an automorphism of the logic J£. Let F be a linear manifold, chosen once for all, such that dim F = 3. By theorem 2.1 there exists an automorphism 6 of D and a 0-linear isomorphism L0, ofF onto Fa, such that for any linear manifold M^F, Ma = L0[M]. Let us define & to be the collection of all finite dimensional linear manifolds G which contain F. We claim that for each G e & there exists a unique 0-linear isomorphism LG of G onto Ga such that (74)
LGu = L0u
(u e F)
and (75)
Ma = LG[M]
(M c G).
In fact, by theorem 2.1 there exists an automorphism 0' of D and a fl'-linear isomorphism L' of G onto Ga such that Ma = L'[M] for all linear manifolds M<^G. Since Ma = L'[M] = L0[M] for all linear manifolds M^F, we conclude from the same theorem that there exists a nonzero C G D such that (i) L,u = cL0u for all ueF, and (ii) ae'=ca9c~1 for all aeD. If we write LG = c~1L/, then LG is a 0-linear isomorphism of G onto Ga satisfying (74) and (75). The uniqueness of LG follows from (74) and (75).
106
GEOMETRY
OF QUANTUM
THEORY
Since LG is uniquely determined by (74) and (75), it follows quite easily t h a t there exists a 0-linear m a p L of J$? into itself such t h a t (76)
Lu = LGu
(ueG,
GeF).
We claim t h a t L is one-one and maps 2tf onto itself. To prove t h a t L is one-one, let x e «*f and I># = 0. Evidently, x e G for some G e ^ and Z/# = LGx = 0. Therefore # = 0 as LG is an isomorphism. If y e 3f there is a 6r e IF such t h a t y e Ga and once again, as XG maps G onto 6ra, there will exist &n x e G such t h a t Lx — y. Since any M e £f with finite dimension is contained in some element of F, we see t h a t L is a 0-linear isomorphism of J f onto itself such t h a t (77)
Ma = £ [ i f ]
(dim I f < oo).
Let now G e JF. Since a preserves orthogonal complementation, it follows t h a t for x, y e G, (x,y} = 0 if and only if (Lx,Ly} = 0. Thus, the semi-bilinear forms x, y -> (x, y) and x, y -> (Lx,Ly}w~ 1} of GxG induce the same polarity of the geometry of subspaces of G, and hence there exists, by virtue of theorem 2.3, a unique dG^Q in D such t h a t (Lx,Ly>
= (x,y}ddG
(x, y e G).
Since G e
= 'd
(*, ^ G J f )
for a nonzero d e D . We now show t h a t 6 is continuous. This is obvious for D = R or H, since 6 is the identity in the former case and is an inner automorphism in the latter. Consider now the case D = C. If 9 is not continuous, there exists a sequence {cn} in D such t h a t cn —>• 0, but \cne\ - > oo. Choose an orthonormal infinite sequence {en : n — 1, 2, • • •} in 3tf and let xn = nen. By replacing cn by a subsequence if necessary, we may assume t h a t (79)
||Z*n||/|cn'|-*0
(n-*oo).
Write
Since c n * - > 0, z is well defined in J f . Clearly, z ^ O . Then, <#n,z> = c n and hence, for all n, un = (z,z}~1z — cn~1xn is orthogonal to z. Consequently, from (78) we obtain, for all n, (Lun,Lz>
= 0.
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
107
This implies, on simplification, that (80)
Kz,z>9) - \Lz,Lz>
= (cne) -
\Lxn,Lz>
for all n. But
KO-'iKM,,!,*)! < ic/i-^inizzji, which tends to zero by virtue of our assumption (79). Hence (80) implies that (Lz,Lz} = 0 or 2 = 0. This contradiction shows that 6 is continuous. We are now ready to complete the proof of theorem 4.27. Since 6 is continuous, it follows that for any real number r, re = r. Hence we see, by putting x — y in (78), that d is a positive real number. Write T — d~ll2L. Then T is a 0-linear isomorphism of Jti? onto itself such that (81)
(x,yy
and Ma = T[M]
{M e &, dim M < oo).
Since ||T#|| = ||#||, T is an isometry and is hence continuous. It follows therefore that M* = T[M] (M E &). The remaining statements are easy consequences of theorem 2.3. Therefore, we omit their easy verifications. The proof of theorem 4.27 is complete. We shall say that T induces a if (71) is satisfied. Corollary 4.28. / / D = R or H, then any automorphism a of J£ is induced by a unitary operator of Jf. 7/D = C, then a2 is induced by a unitary operator. Proof. For D = R there is nothing to prove. If D = H and T± is a 0-unitary operator inducing a where 6 is the automorphism d-^cdc-Wc\
= 1),
T = c~1T1 is a unitary operator inducing a. If D = C and a is induced by a *-unitary Tlt Tx2 is unitary and induces a2. In the physical literature it is customary to prove the above theorem in a somewhat different formulation (cf. Wigner [3]). To motivate Wigner's formulation we must introduce the notion of transition probabilities. Let r, r' be rays in 3? and
108
GEOMETRY
OF QUANTUM
THEORY
decide whether it is in the state p' or not. We shall define a physical symmetry associated with the logic J? to be any one-one m a p a(r -> r a ) of the set of all rays of Jf onto itself with the property t h a t [r a ,r' a ] = [r,r']
(82)
for all rays r, r'. The following theorem describes the natural connection between physical symmetries and symmetries of Jt?. Theorem 4.29. (Wigner [3]). Let $P be a separable infinite dimensional Hilbert space over D and T a symmetry of Jf\ Then r -> T[r] is a physical symmetry associated with ££'. Conversely, let a(r -> r a ) be any one-one map of the set 0k of all the rays of 3F onto itself with the property that r_[_r' if and only if r a JLr' a . Then, a is a physical symmetry associated with <&, and there exists a symmetry T of J f such that r a = T[r] for all r e@£. T is determined by a up to multiplication by a number c e D with \c\ = 1. Proof. If T is a symmetry of JtT, \(Tx,Ty}\ = \(x,y}\ for all x,yeJtf. From this we conclude easily t h a t r - > T[r] is a physical symmetry associated with ££. We now examine the converse. Let a(r —>• r a ) be any one-one map of @t onto itself which preserves orthogonality. For any subset A of 0t let (83)
Aa = {r01 :re
(84)
[A] = {x : x e Jf, x e r
A}, for some r in A},
and (85)
i ^ ^ r l r '
for all
r' e A}.
We shall first prove t h a t if i c j is such t h a t M = [A] is a closed linear manifold of ^f, i.e., if A is the set of all rays of a closed linear manifold M of Jf\ then [Aa] is also a closed linear manifold. To prove this, we note t h a t as M = [A] is a closed linear manifold, a ray r belongs to A if and only if r e (A1)1. But as a preserves orthogonality, this means t h a t a ray r' e Aa if and only if r'e (A1*)1, i.e., Acc = (Ala)1. But for any subset S^Zft, the set [S1] is a closed linear manifold. Consequently [Aa] is a closed linear manifold. We now set up a correspondence, also denoted by a, between elements of <£. For any M e 5£, let A be the set of all rays r such t h a t r c M; we set: (86)
Ma = [^«].
The argument of the preceding paragraph has shown t h a t Ma e j£?. By very definition it is clear t h a t a(M - > Ma) is an order-preserving m a p of ££. Since r c i f 1 if and only if r is orthogonal to all rays contained in M, it follows t h a t (Jf-L)a = ( i f a ) 1 . Thus a preserves orthogonality. These
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
1
109 ia x)
results applied to the m a p a " of 0t show t h a t the m a p M->M ~ defined b y (86) is the inverse of the map M - > Ma of J£. Consequently, we m a y conclude t h a t M - > Ma is a lattice automorphism of <£ which preserves orthogonality. B y theorem 4.27 there exists a symmetry T of J f such t h a t M<* = T[M] for all M e 3?. B u t then r a = T[r] for all rays r showing t h a t r -> r a is a physical symmetry associated with ££. Since any automorphism of ££ is determined by its restriction to the set of all one-dimensional subspaces of #F, it is clear t h a t T is determined by the m a p a of 3% u p to multiplication by a number c e D with \c\ = 1. This completes the proof of the theorem. I n Chapter I I I we introduced the concept of a convex automorphism of the set of all states. We shall describe all such automorphisms of the set of states of a standard logic. Let T be a symmetry of the Hilbert space Jf, and iV the convex set of all bounded, nonnegative, self-adjoint operators U of trace class with tr(Z7) = l. For any U e HT let aT(U) be defined by (87)
aT(U) =
TUT-1.
Lemma 4.30. Let U eiT and let T be a symmetry of J#*. Then aT( U) eW* and U -> aT(U) is a convex automorphism of W'. Moreover, T —> aT is a homomorphism of the group of symmetries of J4? into the group of convex automorphisms of # " . The proof is a routine verification. We leave it to the reader. If T is a symmetry of 3f and if we set, for U e W, (88)
<*T(pu) = p t t T ( U ) ,
where pu is the state M -> tr(PMU) of J&, then the map pv - > aT(Pu) is a convex automorphism of £f. We shall prove t h a t any convex automorphism of £f is of this form. Let iT + denote the set of all von Neumann r operators on 3tf. Clearly, iT^iT to + . If Tlt T2 e if + , we write T1
a(U) = a~(U) a~(c1A1+c2A2)
=
(UeiT), c1a~{A1)-{-c2a^{A2) (Al9 A2 e i^ + , cl9 c2 real numbers > 0 ) ;
a ~ has then the property: [
}
(hi) (iv)
tr(a~(,4)) = tv(A) (AeiT + ), TX
GEOMETRY
110
OF QUANTUM
Proof. Define a~(0) = 0. If T eiT t~1T eiT\ we then put (90)
a~(T) =
THEORY
and T^O,
+
t = tr(T)>0,
and
ta^T).
It is trivial to check that a~ has the properties described in the lemma. The uniqueness of a" follows from the fact that the property (ii) of a~ implies (90). Equation (90) implies that tr(a~(T))=tr(T), proving (iii). Finally, let Tx < T2. Then a~(T2)
= a-iTJ+a-iTt-TJ
> a~(T x )
as a~{T2-T1)eiT+ and is hence >0. Conversely, let a^(Tx) TX. Lemma 4.32. Let
(i)
^J-^2* (ii) For any Aeitr trA>2.
+
, the relations P1
imply that
Proof. Suppose that cp1±_cp2 and suppose that i e # + satisfies the relations P10, 00
tr(^) = ^ n= 1 > (Acp^cp^
+(Acp2,cp2)
^ <^l9l>9>l>+<^29>2>9>2>-
Consequently, tr(A) > 2. This shows that (i) implies (ii). Conversely, let (ii) be satisfied. Let us assume that cp1 is not orthogonal to
0 2 = lala-^&l"^-^'.
LOGICS ASSOCIATED
WITH HILBERT
SPACES
111
S an
Then {
P
1
(l °\ \0
p i
\<
2
OA
HI6I\ |6| 2 /
\|o||6|
Since |a| 2 + |6| 2 = l, the eigenvalues of the matrix of P2 — P1 are ±\b\, and hence there exists a unitary operator V such t h a t the operator V(P2 — P1)V~l has t h e matrix given by
/\b\ ° \ -\b\J
\0
1
Let M be the unique operator such t h a t VMV'
has the matrix given by
h
l\ \ °\ \0
0/
Obviously M is self-adjiont, > 0 and tr(Jf) = |6| < 1. Further, 7if7-i_7(P2_Pl)F-i > o and hence M>P2-P1. Let A = P±+M. tr(^4) = 1 + \b\ <2. This proves the lemma.
Then P1
P2
but
Theorem 4.33. Let Jf be a separable infinite dimensional Hilbert space over D and T a symmetry of ffl. For any U eW let pv be the state M -> tr(PMU) of & and UT=TUT~1. If (92)
<*T(PU) =
PUT,
s a
then aT{pu —>PuT) ^ convex automorphism of Sf. Conversely, let a be a convex automorphism of £f. Then there exists a symmetry T of Jt? such that (93)
a = aT.
T is determined up to multiplication
by a number c e D with \c\ = 1.
Proof. The first p a r t of the theorem is rssentially the content of lemma 4.30. To prove the converse, let a be a convex automorphism of £f. By theorem 4.23, a induces a convex automorphism of # " , denoted once again by a. Let a~ be the m a p of # " + , determined by a (cf. lemma 4.31). Since the extreme points of iT are the orthogonal projections on the onedimensional subspaces of $F and since a is convex, a induces a one-one correspondence of the set ^ , of all rays of Jf7, onto itself. Let a(r - > ra) be this correspondence. Since a~ preserves the trace and the partial ordering < of # " + (by (iv) of lemma 4.31), and since the orthogonality of two rays has been completely characterized in terms of the trace
112
GEOMETRY
OF QUANTUM
THEORY
function and the partial ordering < in lemma 4.32, it follows a t once t h a t r - > r a is orthogonality-preserving. By theorem 4.29 there exists a symm e t r y T such t h a t r a = T[r] for all rays r. If aT is the convex automorphism of £f associated with Ty this means t h a t aT and a coincide on the set of extreme points of / #^. From this it follows t h a t a — aT. The fact t h a t T is determined up to multiplication by an element C G D with |c| = l now follows from theorem 4.29. The theorem is completely proved.
4. LOGICS ASSOCIATED W I T H VON N E U M A N N ALGEBRAS If the logic J^7 of a quantum mechanical system is not isomorphic to a standard one, then the question of describing its observables, states, and symmetries becomes a more difficult one. There are no general results in this connection. I t is our aim to give a few examples which illustrate the wide possibilities as well as the difficulties involved in the construction of a n y general theory. A large class of logics which are sublogics of standard logics m a y be constructed using the theory of von Neumann algebras. We confine ourselves to the case of complex scalars. Let D = C be the complex number field and 3^ a separable infinite dimensional Hilbert space over C. We write ty(J^) for the algebra of all operators on Jf. For u, v e Jtf, let XUtV : A-> (Au,v}
(Aety(Jf)).
Then the Xuv are linear functionals on ty(34f). The weak topology on ty{3^) is the smallest one with respect to which all the AMfU are continuous. A von Neumann algebra ty is a subalgebra (containing / ) of ty(3#*) such t h a t (i) if A e ty, then A* e ty, and (ii) ty is a closed subset of ty(J4f) in the weak topology. lity' is the set of all elements of ty(Jf) commuting with all t h e elements of ty, ty' is also a von Neumann algebra and (ty')' = ty (cf. Dixmier [1] for the general theory of von Neumann algebras). Let ty be a von Neumann algebra. We define (94)
Se% = {M : M e Jg%?f), PM e ty}.
I t is then easy to verify t h a t J ^ is a logic. We shall call it the logic associated with ty. A closed linear manifold M lies in S£^ if and only if every element of ty' leaves M invariant. Thus the logic ££% may also be introduced as the logic of closed invariant linear manifolds of some von Neumann algebra. The one-one correspondence x - > Ax between observables of ££% and self-adjoint operators on 34?, which was discussed in section 2 of this chapter, clearly persists even in this case. However, not every selfadjoint operator is the operator of an observable associated with «#V
LOGICS ASSOCIATED
WITH HILBERT
SPACES
113
Let A be a self-adjoint operator, not necessarily bounded. Then A=AX for an observable x of jSf9 if and only if the spectral projections of A lie in 51. If A is bounded, t h e condition will then be t h a t A e 21. If A is unbounded, t h e condition will have to be stated more carefully. I n this case, A — Ax for an ^ - o b s e r v a b l e if and only if UAU~1 = A for all unitary operators U e 21' (cf. Dixmier [1]; such ^ ' s are said t o be affiliated to 2t).
p(M) = \\PM
(Jfe-SP.).
I n other words, the pure states of S£% are still in one-one correspondence with t h e rays of Jf7; however, not every ray can be used. Only the rays which belong to some Mn correspond to pure states. Physicists describe this situation by saying t h a t no ray of ^f7, which is a nontrivial superposition of rays in distinct Mn, can describe a (pure) state of J ^ . Each Mn is
114
GEOMETRY
OF QUANTUM
THEORY
called a coherent subspace of (pure) states. The conditions which demand t h a t the rays describing the pure states should lie in Mn for some n are usually referred to as super selection rules, (cf. Wick, Wightman, and Wigner [1]).
5. ISOMORPHISM AND I M B E D D I N G T H E O R E M S I t is very important in the foundations of quantum mechanics to examine the extent of generality of the concept of an abstract logic. I n particular, the question arises whether every logic can be regarded as a sublogic of a standard logic. Definitive answers to such questions are not known. We shall prove in this section some theorems which analyze t h e structure of a somewhat restricted class of logics. The main theorem of this section was proved by Piron [1]. A related but somewhat weaker result was obtained by Zierler [1]. The main tool of our analysis is the coordinatization theorem of Chapter I I . I n order to be able to apply t h a t theorem we must impose a sharp restriction on the class of logics to be analyzed. We introduce accordingly t h e following definition. A logic ££ is said to be projective if the following conditions are satisfied:
(96)
(i) given a # 0 in iff, there is a point x
For any a e = ^ , let (97)
$ ( a ) = {x : x a point, x < a}.
Note t h a t if S£ is projective and its lattice is complete, then every element a of j£? is the lattice sum of the points it contains. I n fact, if b is the sum of these points and b^= a, c = 6 i A a ^ 0 and there would exist a point x < c by (i) of (96) which contradicts the definition of b. Lemma 4.34. Let ££ be a projective logic whose underlying lattice is complete. If a and bj (j eJ) are elements of ££ and if a\bjfor allj, a_J_V; ^'* We have the usual identities:
LOGICS ASSOCIATED If a,be^,
then a
WITH HILBERT
SPACES
115
if and only if $(a)c=*g(&); a\_b if and only if
Proof. The first two assertions are proved exactly like the corresponding assertions in lemma 3.1. We therefore come to the proof of the last assertion. If a be a 0-bilinear symmetric form on V x V. Let us assume t h a t <. , .> is definite; i.e., (x,x} = 0 if and only if x = 0. For any set M of vectors we write (98)
Mx = {u : (x,u}
= 0
for all
x e M}.
Obviously, M1 is a linear manifold in F, M n M1 = 0, and M^M11. If M is a linear manifold and M = M11, we shall say t h a t M is <. , . >-closed. Note that no topology is involved in this definition. The pair (V, <. , .>) is said to be Hilbertian if for any closed linear manifold M, one has: (99)
V =
M+ML
algebraically. For any set i f c f w e write M~ = i f 1 1
(100)
If A, B^ V we write A^_B whenever = 0 for all a e A and b e B. Clearly 0 " = 0 , V~ = F, and 0L= V, V1 = 0. If M and JV are linear manifolds and M^N, then NL<^ML. If {if ; } is a collection of linear manifolds, ilf their algebraic sum, and Nj_Mj for all j , then i^ J_if. All this is trivial. Lemma 4.35. The following statements on a set M^V (i) M is a <. , .^-closed linear manifold. (ii) i f = i f " . (iii) M = N1 for some set N c F. Moreover, for any set (101) (102)
M^V, if111
=
if1,
( i f - ) 1 = if 1 ,
and (103)
(if-)" =
if".
are equivalent.
116
GEOMETRY
OF QUANTUM
THEORY
Finally, if {M^ is a collection of <\ , . >-closed linear manifolds, also a <. , .>-closed linear manifold.
P | ; Mj is
Proof. Clearly for any set i f c V, M^MLL. Thus ML<^(ML)LL. On the LL 1L L L other hand, as M^M , (M ) <^M . Therefore M1 = M111, which is 11 (101). Since M~ = if , (102) follows from (101). Further, (M-)~
= ( i f 1 1 ) 1 1 = (M111)1
=
M11
b y (101), so t h a t ( i f - ) - = i f " . This proves (103). We now come to the proof of the equivalence of (i) through (iii). Suppose M is a <. , . >-closed linear manifold. Then M = M1L, so t h a t M = M~. Thus (i) => (ii). If M = M~, then M = i f 1 1 = ( i f 1 ) 1 , proving t h a t (ii) => (iii). Suppose now t h a t M = NL. Then, by (98), M11 = N111 = N1 = M, proving (iii) => (i). Finally, let M = f|y M,. As M^Mj9 M11^Mj11 = Mj. Thus
M11
c Q if. = M.
This proves t h a t M is a <. , . >-closed linear manifold. Corollary 4.36. If M^V, M~ is the smallest <. , . >-closed linear containing M. In particular, if M^N, then M~^N~.
manifold
Proof. M" is <. , . >-closed b y (103). I t is therefore enough to prove t h a t if iV is a <. , .>-closed linear manifold and M^N, then M~^N. But M-=MxlcN11 = N. Finally if J f c t f , M^N~, d as iV" is <. , .>an closed, M~ <^N~. Lemma 4.37. If M is a finite dimensional subspace of V, then M is <. , . >closed. Proof. I t is enough to prove t h a t M1L^M. Let x e if 1 1 . Let N be the LL subspace spanned by M and x. N^M . Since < . , . > - is definite, its restriction to NxN is definite and hence nonsingular. B u t N is finite dimensional and hence it follows from (13) of Chapter I I and the definiteness of <\ , .> t h a t N is the direct sum of M and (M1 n N). Therefore, x = x' +x" where x' e M and x" e M1 n N. Since N^M11, x" e i f 1 1 n ML and hence x"= 0. Thus x e M. Lemma 4.38. If {if ; } is a family of linear manifolds of V, and if 2 / Mj denotes the {algebraic) linear span of the Mj, (104)
g
if,)-
= (c\Mj-y.
Proof. Let M = %jMj. Since Mj^M, ML^MjL so t h a t i f 1 ^ . if/. 11 1 1 This shows t h a t M~ =M ^(f)j Mj ) = N, say. On the other hand, as H ; Mj1^Mj1, Mj^Mj- <^N for all j so t h a t M^N. Since N is <. , .>closed, it follows from corollary 4.36 t h a t M~ s^N. Therefore N = M~.
LOGICS ASSOCIATED
WITH HILBERT
SPACES
117
Lemma 4.39. If (V, (. , . >) is Hilbertian and Mx, M2, • • •, Ms is a finite set of mutually orthogonal <. , .)-closed linear manifolds, then 2? = i -^j ^s also a <. , . >-closed linear manifold. Proof. I t is enough to consider the case 5 = 2 since the general case follows at once by induction. We want to prove t h a t (M1+M2)'
=
M1+M2.
Write M = M1+M2. Let x e V. Since (V, <. , . » is Hilbertian, we have the unique decompositions x = x±+y y = x2+z Now, as J f 2 c MXL, x2 e M^
(x^M^yeM^), (x2 e M2, z e M2L). so t h a t z e MXL also. Therefore, z e M^
n
M2L,
and we have: x = x1 +x2 +z
(x± e M±, x2
G
M2, z e MXL n M21).
Suppose now t h a t x e (M1 + M2)~. Then the above equation shows t h a t ZE (M1+M2)~ n Mi1 n M2L, which implies t h a t z = 0 on using (104). Therefore (M1 + M2)~=M1 +M2. Theorem 4.40. Let D be a division ring, V a vector space over D with 4:
y*, = £*,)-, (105)
A*, = nM,j
o
Conversely, let 3? be any complete projective logic. Then there exists a division ring D, an involutive anti-automorphism 6 of D, a vector space V over D, and a definite symmetric 6-bilinear form <\ , . > on V x V such that (V, <. , . ) ) is Hilbertian and ££ is isomorphic to <^{V, <\ , . ) ) . Proof. Let D, 6, V, <. , .> be as given in the first half of the theorem. Let j£? = j£?(F, <\ , . >) be the partially ordered set of <. , .)-closed linear manifolds of V. Assume t h a t (V, <. , . >) is Hilbertian. For any linear manifold M, M ~ is the smallest element of ££ containing it. I t is clear from lemmas 4.35 through 4.38 t h a t ££ is a complete lattice with the lattice
118
GEOMETRY
OF QUANTUM
THEORY
operations given by (105). Next we observe t h a t if we define, for a n y i f e i f , M1 by (97), then j£? becomes a logic. To check this, the only thing not immediately obvious is the weak modularity (2) (ii) of Chapter I I I . Suppose then MvM2e^ and M±^M2. Let N=M1±nM2. By lemma 4.35, N is a ( . , .)-closed linear manifold so t h a t NeJ?. Clearly N^M2, N±MX. Since (V, <. , . » is Hilbertian, V = M1+M11 and for a n y x e V we have the decomposition x — x1 +x2, where x± e Mlf x2 e MXL. If now x e M2, then x1 and x2 e M2 also, so t h a t x2 e N. Therefore M2 = MX +N and a fortiori M2 = M1 v N. <£ is thus a complete logic. We claim t h a t it is projective. By lemma 4.37, i f contains all finite dimensional subspaces. As dim F > 4 , (i), (ii), and (iv) of (96) follow at once from this fact; (iii) follows from the algebraic decomposition V—M+M1 for any M e J*?. ££ is thus a complete projective logic. We now come to the converse. Let i f be a complete projective logic. Let us now define ££' by (106)
^ ' = { a : a e ^ ,
a finite}.
B y (ii) of (96), a e 3?' if and only if the underlying lattice of if[0,a] is a geometry. Under the induced partial ordering, JSP' is thus a generalized geometry. Therefore there exists by theorem 2.16, a division ring D, a vector space V of dimension at least 4 over D, and an isomorphism of ££' onto the generalized geometry of all finite dimensional subspaces of V. Let y denote any such isomorphism. Fix a nonzero vector v0 e V. Let W be any finite dimensional subspace of V containing v0 and of dimension at least 3. The isomorphism y transfers the orthocomplementation in o£P[0,a] into an orthocomplementation on the projective geometry of subspaces of W. Hence, by theorem 2.7 there exists an involutive anti-automorphism 9W of D and a symmetric, definite # w -bilinear form <. , . > w on W x W, inducing the orthocomplementation in question, with (v0,voyw = l. We now argue as in lemma 4.2 to conclude t h a t 6w=0 is independent of W and t h a t there exists a symmetric, definite 0-bilinear form ( . , . > on V x V such t h a t (i)
y~(a) = {v : v e V
and
v e y(x)
for some point x < a}.
We shall first show t h a t y~(a) is a linear manifold in V. This is obvious for a = 0 or a = 1, as then y ~ ( 0 ) = 0 and y ~ ( l ) = V. Let then a ^ O , ^ 1 . If v i> v2 e y~( a )> there are points x1,x2
LOGICS ASSOCIATED
WITH HILBERT
SPACES
119
dimensional subspaces of V, y{xx) and y(x2) are rays in y(x1 V x2). Thus if v = c1v1 + c2v2 where c1? c 2 e D, t; lies in y(x1vx2) and hence v e y(x) for some point a; < x± V # 2 < a' y ~ ( a ) is thus a linear manifold. We shall prove next t h a t y ~ (a) is <. , . >-closed. From lemma 4.34 it follows easily t h a t for a point x e J?, a; < a if and only if x_\_y for all 1/ 6 $ ( a x ) . Therefore, (108)
y ~(o)
= (y^a1))1,
which shows t h a t y~(a) is <(. , .>-closed. Moreover, as (108) is valid for all a (in particular, for a1), we have: (109)
y~(a±) = (y~(a))^.
(109) shows t h a t y~ preserves orthocomplementation. From (107) we conclude, on using lemma 4.34, t h a t for a, b e J^, y~(a)£y~(&) if and only if a
120
GEOMETRY
OF QUANTUM
THEORY
assert that, given any complete projective logic £? associated with D which is one of R, C, or H, there exists a pre-Hilbert space V, with inner product <. , .>, such t h a t (i) (V, < . , . > ) is Hilbertian and (ii) 3? is isomorphic to the complete projective logic of all <. , . >-closed linear manifolds of V. I t is natural to expect t h a t V is then actually a Hilbert space. This essentially leads to Piron's theorem. We shall obtain it as a consequence of a few lemmas. I t is an interesting problem to examine what natural assumptions on a geometry imply t h a t the associated division ring is isomorphic to one of R, C, H. This is a classical question and is intimately connected with t h e topologies on a geometry (cf. Kolmogorov [1], Weiss and Zierler [1]). Lemma 4.41. Let V be a pre-Hilbert space over D (one of R, C, or H) and let Jf be its completion. If x0, x0' e Jti? are orthogonal, there exists a pair of sequences xn and xn' such that /110x
(i) xn _L xm' fa, m = 0, 1, 2 , . . .), (ii) #n> *^m ^ ^ and xn —> x0, xm —> x0 •
Proof. We shall prove first t h a t if z e Jf? is a nonzero vector, t h e n {z}1 n V is dense in {z}1. I n fact, let y_[_z, and let {yn}, {zn} be sequences of elements of V such t h a t yn->y, zn-> z. Write (HI)
Vn =
Then yn' e V and (yn',z}
Vn-iyn^X^n^}'1^
= 0 for all n. As yn - > y and zn -> z,
showing t h a t yn' -> y. This proves our assertion. Suppose next t h a t y> Vii ' ' ' > Vsare mutually orthogonal nonzero vectors. Applying repeatedly t h e result proved just now, we conclude in succession t h a t { y j 1 n V is dense in {2/J 1 , {ys^s-i}1 n F is dense in {f/s^s-i} 1 and so on. I n other words, given any finite dimensional subspace Y of £? and a vector y\_ Y, there exists a sequence {zn} in V such t h a t zn_]_ Y for all n and zn - > y. This said, we come to the proof of the lemma. Let x0, x0' e J f be given with x0_\_x0'. We shall show by induction t h a t there are sequences {xn} and {xn'} in V such t h a t (U2)
K-*o|| ^2"n' \\*n'-*o\\ <2"», ^n J_ xm' (n,m = 0, 1,2, • • • ) .
Suppose ajj, z2> • • •, xk and x±', x2', • • •, xk' have been constructed so t h a t (112) is satisfied for all n, ra = 0, 1, 2, • • •, k. By the result proved in t h e previous paragraph, there is a vector x'k + 1e V such t h a t 4 + i _L*n
(n = 0, 1, •••,&),
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
121
Applying the same argument again, we can find xk + 1e V such that (n = °> 1, • • • , * + !.),
a* + i ±xn'
Thus {xn}0<.n£k +1 and {xn'}0^n^k + 1 satisfy (112) for n and m
xn-+x0,
zn-+z0,
xn±zm
(n, m = 0, 1,2, • • •)•
Let M c 7 be the lattice sum in & (F, <. , . » ofthe rays D-a n (n = 1,2, • • •)• Then, as (F, <.,.>) is Hilbertian, V =
M+M1.
Now zm_]_#n for all n so that zm e ML for all m. Hence #0 e Mcl and z0 e (M1 ) c l where cl denotes closure in Jf. Moreover, it is obvious that (ML)cl^{Mcl)'. Therefore, (114)
y = x0 +z0,
xQ e Mc\
z0 e (Mcl)'.
On the other hand, as V = M + MX9 we can also write (115)
y = x0' +z0',
x0f e M,
z0' e M1.
As M-t-^iM01)', a comparison of (114) and (115) shows that x0' = x0 and z0' = 20. In particular, x0 e V. This completes the proof of the lemma. Lemma 4.43. Let ££ be a projective logic with the property that any family of mutually orthogonal points of ££ is at most denumerable. Then ££ is complete. Proof, It is enough to prove that arbitrary lattice unions exist in 3?. Let {aj- : j e F} be a family of elements of ££\ For any countable subset
D^F, let c(D) = V «y jeD
f The argument for completeness is essentially that of I. Amemiya and H . Araki, Publ. Res. Inst. Math. Sci. Kyoto Univ. A2 (1966), p . 423. Their theorem is t h a t if V is pre-Hilbertian and &(V, < . , . » is orthomodular, then V is complete.
122
GEOMETRY
OF QUANTUM
THEORY
Let B be a family of mutually orthogonal points x such that (i) for each x e B there is a countable set D^F such that x
and for each n let Dn c: F be a countable set such that *n < c(Dn)
for all 7i. Write D = [Jn Dn. If we define x by a = V xn> n
then x
NOTES ON CHAPTER IV 1. Hidden variables: Is the quantum mechanical description of microphysics complete'1. The question of "hidden variables" has persisted in the foundations of Quantum Theory from its very birth. The basic assumption in Quantum Mechanics, namely that in any state of an atomic system a physical quantity can in general have no sharply defined value (even under idealized measuring conditions) but only a probability distribution of the possible values, appeared to many people to be in very sharp conflict with all classical experience, and hence unacceptable, unless it was interpreted as follows: one is not able to get a complete description of the physical state of the system; there are additional coordinates ("hidden variables") that cannot be measured; and therefore the statistical character of the results of the experiments is due to the averaging over these hidden variables. This view, which regards the Hilbert space description of quantum states as
LOGICS ASSOCIATED
WITH HILBERT
SPACES
123
fundamentally incomplete, is reinforced by the example of statistical mechanics where the situation is essentially of this character. Von Neumann was among the earliest to analyze this question, and his conclusion was that no mechanism of hidden variables could account for the statistical content of Quantum Mechanics as currently formulated. That he regarded this question, as well as his resolution of it, to be of great importance is clear from the fact that he raises the issue already in the preface of his book (pp. ix, x), returns to it briefly (pp. 209-210) before devoting a substantial effort (pp. 295-328) to a detailed mathematical and physical treatment (all references are to the 1955 English translation by Robert T. Beyer of von Neumann's book). We shall briefly summarize his argument but suggest strongly that the reader study von Neumann's brilliantly presented analysis. Von Neumann's reasoning, roughly speaking, has two parts to it, one physical, the other mathematical; and they reinforce each other. The physical side of his argumentation uses the technique of ensembles and begins by considering a statistical ensemble consisting of a large number of models of the physical system under consideration. If we reject the view of Quantum Mechanics, and suppose that the values of physical quantities in the ensemble have positive dispersions only because the ensemble has not been resolved in a sufficiently fine manner, we must admit that we should be able to carry out such a resolution into subensembles in which the dispersions are zero, or at least diminished from those of the original ensemble. The method of doing this is to measure the physical quantities in succession, replacing the ensemble each time to one of the subensembles where the quantity has a sharply defined value. But then, the Heisenberg uncertainty relations between complementary physical quantities operate in such a way that the precision achieved for a quantity in any stage is destroyed at the next stage when a complementary quantity is measured; we thus do not get ahead. However, one may take exception to this line of reasoning by pointing out that there could conceivably be other methods of penetrating to the dispersion free ensembles. To settle this point decisively one must therefore answer the following question: Given an ensemble in which certain physical quantities have positive dispersion, is it possible to resolve this ensemble into a superposition of two subensembles, different from the original one and distinct from each other? Ensembles which cannot be so resolved are the pure (homogeneous in von Neumann's terminology) ones, and the above question becomes the following: Is every pure ensemble dispersion free? If the answer is no, then the hidden variables interpretation must be given up. To answer this question in the conventional model of Quantum Mechanics von Neumann begins with the very interesting observation that since an ensemble is characterized (statistically) by the expectation values of all the
GEOMETRY
124
OF QUANTUM
THEORY
physical quantities in it, one may replace the ensemble in the mathematical argument by the corresponding expectation functional; we must remember that if A is any bounded observable, its probability distribution is completely determined by the knowledge of all the expected values Ex$(An),n
= 0,1,2,....
Thus, for von Neumann, an ensemble is the same as the corresponding real valued (expectation) functional Exp:J5
V^G^(Jf);
Exp is linear (over R);
(d) Exp is continuous in the strong operator topology. Nowadays, one works with the space B(Jf) of all bounded operators on &?\ the conditions (a)-(c) define a state, (b) being equivalent to Exp(^4J.t) ^ 0, yAeB(J^); with (d), we have a normal state. He showed that the functionals Exp are in one-one correspondence with bounded operators U which are self-adjoint, ^ 0 , and of trace 1, the correspondence being defined by (2)
Exp(A)=tT(UA).
The correspondence preserves convexity and so matches the pure states with the extreme points of the convex set of the U's. He verified that the latter are the one-dimensional projections, and showed by explicit calculation that these are never dispersion free. This concluded his proof. One must emphasize that his reasoning explicitly assumes that Quantum Mechanics is described by the conventional model. It is possible to modify von Neumann's argument so as to allow for the presence of superselection rules, but the essence of the reasoning does not change. Many subsequent criticisms of von Neumann's proof are unjust because they do not take into account the above assumption in his proof, although von Neumann states it quite unmistakably (see the first sentence of the first paragraph on p. 210, and lines 17 -h on p. 324; see also the remarks of L. van Hove, Von Neumann's contributions to Quantum Theory, pp. 95-99, in John von Neumann, 1903-1957, Bull. Amer. Math. Soc. 64 (1958), No. 3, Part 2). The mathematical content of von Neumann's argument may thus be summarized as follows: B(3tf) does not admit any dispersion free normal state.
LOGICS ASSOCIATED
WITH HILBERT
SPACES
125
The first question t h a t arises naturally now is whether von Neumann's definition of a state was unduly restrictive. Obviously the conditions (c), (d), of (1) need to be examined more closely. A conservative definition would replace (c) b y (3)
(c')
E x p ( ^ +B) = Exp(^4) + E x p ( £ )
when
AB =
BA.
Let us call maps A: BQ(J^) -> R satisfying (a), (b) of (1) and (c') of (3)physical states. I t then follows t h a t |A(-4)| <\\A\\ and t h a t the restriction of A to the projections defines a finitely additive probability measure on the logic J£?(e2f). Conversely, it follows from the spectral theorem t h a t any finitely additive probability measure on ^ ( J f 7 ) is the restriction to the projections of a unique physical state of B0(J^). I n other words, physical states m a y be identified with finitely additive probability measures on j £ ? ( ^ ) . Thus the strongest generalization (in this direction) of the von Neumann result would be to show t h a t there are no finitely additive dispersion free probability measures on J>?(Jf). The case when 3 ^ d i m p f )
126
GEOMETRY OF QUANTUM
THEORY
full generality o n l y recently, and the main motivation for reaching this result had always been the desire to generalize Gleason's theorem to more general von Neumann algebras t h a n B(Jf). A second question t h a t arises now is whether one can relax the assumption in the von Neumann-Mackey-Gleason result t h a t the logic of the physical system is <&(<#?). There is an extensive literature on this question and I shall limit myself to a brief discussion of two results, those of Neal Zierler and Michael Schlessinger (Boolean imbeddings of orthomodular sets and Quantum Logic, Duke Math. J., 32 (1965), pp. 251-262), and of Simon Kochen and E . P . S p e c k e r (The problem of hidden variables in Quantum Mechanics, J . Math. Mech., 17 (1967), pp. 331-348); for more detail and other references t h e reader should consult the lucid discussion of these problems in the book of Beltrametti and Cassinelli (of notes to Chapter I I I ) . I n both of these articles, the interpretation of what is meant by hidden variables is such t h a t it implies the existence of some sort of imbedding of the proposition system into a Boolean algebra. So the mathematical issue in both cases comes down to showing t h a t under suitable assumptions on the proposition systems such imbeddings do not exist. I n Z-S it is assumed t h a t the set i ? of experimentally verifiable quantum propositions is partially ordered, has 0 and 1, and an orthocomplementation ± : S£->£ (0 X = 1, l x = 0, a ^b=>b^ < a - \ a-L± = a, avaA- = 1, aAa± = 0) which is weakly modular (a^b=>a1Ab exists and b = a\/(a-LAb)); it is not assumed t h a t j£? is a lattice. As usual if a ^ 6 X we say a and b are orthogonal; sums of mutually orthogonal elements exist in «j£f. If x, ye<£f they are said to commute if we can write x = xxyz, y — yx\/z where xv yv and z are mutually orthogonal; this is possible if and only if x and y are contained in a Boolean subalgebra of & (recall t h a t 5 c i f is a Boolean subalgebra if 0, leB; if c,a,beB ^C^GB, a\jb, a Kb exist in ££ and belong to B; and if B becomes a Boolean algebra with respect to v, A, _L). If B is a Boolean algebra, a n imbedding of S£ into B is a m a p h (<& ->B) such t h a t
(4)
(a) (b) (c) (d)
A(0) = 0, h(l) = l; x ^yoh(x)
Actually it is enough to require (b), (c) and the single set of relations h(x+y) = h(x)\/h(y) in (d). Z - S prove t h a t if ££ contains a copy of j£f(R 3 ), then i f has no imbedding into a Boolean algebra; indeed, it follows from Stone's theorem t h a t any Boolean algebra has {0,l}-valued probability measures, so t h a t such imbeddings would give rise to a {0,l}-valued probability measure on i f ( R 3 ) . Z-S also proved t h a t for arbitrary ££\ if one
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
127
assumes t h a t there are imbeddings into Boolean algebras which satisfy (d) of (4) in the following stronger form (5)
(d')
h(x v y) = h(x) V h(y)
iixyy
exists in 3?,
then x\jy will exist in ££ only when x and y commute; thus if S£ is in addition a lattice (e.g., a logic), such an imbedding will exist only when J£ is already a Boolean algebra. The analysis of K - S is along similar lines but it is formulated in the language of partial Boolean algebras. A partial Boolean algebra is a set L with a family 93 ={B} of subsets of L with the following properties:
/m
(a)
each B e 93 has the structure of a Boolean algebra; if B, B' e 93 and B^Bf (as sets), B is a Boolean subalgebra of B';
(b)
given B , B ' e SB, S n B ' e S B ;
(c) (d)
each element of L is in some Be 35; if a l 5 ..., an e L and any two of them are contained in some member of 93, there is a member of SB t h a t contains all of them.
(6)
I t follows easily from these axioms t h a t L has the following properties: (a')
t h e unit and null elements of all members of 93 are the same; we write 0 and 1 for them;
(b')
if a e L there is a unique a±eL such t h a t a1- is the complement of a in any B e SB t h a t contains a;
(c')
write, for a, b e L, a ^ b to mean t h a t a and b are in some B e SB and satisfy this relation there; then ^ is a partial order on L;
(d')
with respect to (<,J_), L becomes orthocomplemented and weakly modular. Obviously, if we start with an orthocomplemented weakly modular L its Boolean subalgebras will give L the structure of a partial Boolean algebra if (d) of (6) is satisfied. A morphism of a partial Boolean algebra (L,SB) into another (2/,93') is a m a p h:L->Lf such t h a t for any 2?e93, there is a 22'eSB' such t h a t h(B)^B' and h:B->B' is a morphism of Boolean algebras. Imbeddings are injective morphisms. For K - S the basic assumption is t h a t the set 3? of experimental propositions is a partial Boolean algebra, and as we mentioned earlier, their interpretation of the existence of hidden variables will imply the existence of an imbedding of S£ into a Boolean algebra. They prove t h a t the existence of such an imbedding is equivalent requiring t h a t if a, b e ££ and a^b, there is a {0,1}-valued probability measure honJ? such t h a t h(a) ^h(b). They then proceed to construct explicitly a finite subset F of J^(R 3 ) with the property t h a t the partial Boolean subalgebra of J^(R 3 ) generated by F does not admit a single {0,1}-valued probability measure.
GEOMETRY
128
OF QUANTUM
THEORY
The case of the logics of the two-dimensional spaces has been completely excluded in the above discussion. The spin observables of an electron are an example of such a system. Here the Hilbert space <#?= C2. If
(8)
*!-£ I), *,-(<
""Q),
*,= (J _J)
are the Pauli spin matrices, then for a n y a = (ava2>a3) in the unit sphere S2 of R 3 , the self-adjoint operator a • = a1a1-f a2a2 + a3o-3 corresponds to t h e spin component along the a-direction; its spectral projections are i ( l ± a • a), corresponding to t h e eigenvalues + 1. The map a-> | ( 1 +a-a) is a bijection of S2 with the set of (Hermitian) one-dimensional projections of Jf, and gives rise to a bijection p->fP between the set of probability measures^? on JS?(JH and functions/: £ 2 - > [ 0 , l ] such t h a t
(9)
/(2)>0,
/(o)+/(-S) = l
(«e^2),
with_p(J(l+a-a))=/3,(a). The q u a n t u m states, i.e., the states in t h e conventional model of Quantum Mechanics are the functions ql (beS2) denned by (10)
qt(a) = i ( l + a -S)
(• = scalar product in R 3 ).
I t is clear t h a t the functions (10) form only a small part of the convex set of functions (9) (even if we assume they are Borel). The extreme points of the latter set are the functions t h a t take only the values 0 and 1, the dispersion free states; they are the characteristic functions 1A where A^S2 is a n y subset with Au ( — A) — 82i Ar\( — A)= 0. This suggests t h a t we can express the states (10) as mixtures of suitable families of dispersion free states and hence t h a t a classical imbedding may exist for t h e q u a n t u m system associated to Jf. This is in fact so, and the details were first given in articles of Kochen-Specker (loc. cit.) and J . S . B e l l (On the problem of hidden variables in Quantum Mechanics, Rev. Mod. Phys., 38 (1966), pp. 447-452). I n what follows we shall describe briefly the construction of Kochen-Specker; Bell's example is built along similar lines. The space of "hidden variables " is S2. I t s points w define dispersion free states frv\ for each self-adjoint operator AsB(J^), there corresponds a function FA on S2; and for each pure q u a n t u m state q, there corresponds a probability measure JJLQ on S2. The FA and /xfl are related by (i) (11) (ii)
FU(A) = U(FA) for a n y Borel function u, r q(A) = FAdnQ. Js*
The relation (i) asserts t h a t the functional relationships among q u a n t u m observables are preserved under the m a p A ->FA, and relation (ii) asserts
LOGICS ASSOCIATED
WITH HILBERT
SPACES
129
that the probability distribution of A in the state q is that of FA with respect to /zff. If A = |(A7+?• a) (A e R) (having eigenvalues J(A ± 1)), (12)
^
= *(A+1)1«.T>0+*(A-1)1^
where 1^> 0 is the characteristic function of the set of w's with w -"f ^ 0, and so on. It is immediate that the assignment (12) possesses the property (i) of (11). If qt is the quantum state (10), the associated probability measure fit=Pot is defined by (13)
dfit = 4(6-5?) l%fe0dw,
where did is the surface measure on $ 2 , given by dw = (l/4;r) sin Odd d
1_HMW= f
4(S-tS)diS.
Finally, for each w let/^(a ->/^a)) be the function which is 1 when w • a > 0, 0 when w-a<0, and which is defined so satisfy (9) when w-1z = 0. Then/^ defines a dispersion free state on J^(J^), and the formula (14) becomes (15)
qt== fadix%{w) fadptii
exhibiting qs as a mixture of dispersion free states; the absence of explicit definition off$(r) when w-T=0 does not matter, since this set has measure 0 in w. We now turn our attention to the analysis from the hidden variables point of view, due to Bell, of the Einstein-Podolsky-Rosen example (J.S.Bell, On the Einstein-Podolsky-Rosen paradox, Physics, 1 (1964), pp. 195-200); actually Bell considers the variant, due to Bohm and Aharanov, of the EPR example. The system consists of a pair of spin J particles and only their spin observables are of interest. Thus the Hilbert space is (cf. the remarks below in the discussion of composite systems) j T = C2(g)C2. The particles are in the "singlet" state defined by the state vector (16)
® = -jz{(p+®(p--(p-®
where 9^ are the spin states of eigenvalues + 1, say, in the z-direction. For instance, the spin parts of the states of the two electrons in the Helium atom or the H2-molecule are in this state (which is the unique rotation invariant
130
GEOMETRY
OF QUANTUM
THEORY
antisymmetric state); even if the electrons are stripped away by dipole radiation and become widely separated spatially, (16) describes to a good approximation the compound state of the two electrons (cf. the discussion of E.P. Wigner in Interpretation of Quantum Mechanics, in Quantum Theory and Measurement, edited by J.A.Wheeler and W.H.Zurek, Princeton University Press, Princeton, NJ, 1983, pp. 260-314, especially pp. 291-293). Let Xlt% and X2,t be the observables measuring the spin components of the first and second particle in the direction a. In Quantum Mechanics they are represented respectively by the operators a • 5
Pw(a,b)
= -a-b=
-cos0,
where 6 is the angle between a and b. The question studied in loc. cit. by Bell is whether the correlated statistics of the two particles in the singlet state that is summarized by (17) can be obtained from a hidden variables model that is local in the sense introduced by Einstein, Podolsky, and Rosen in their famous paper, namely, measurements on either particle do not disturb the other. Bell's analysis showed that the correlations arising from any local hidden variable model must satisfy certain limitations that conflict with (17). For nonlocal models, however, these inequalities may be violated (cf. M.Flato et al, Helv. Phys. Acta, 48 (1975), pp. 219-225). The model consists of a probability space X of points A, and a probability measure dX on X; moreover, the assumption of locality enables us to say that there are random variables A^aiX) and A2(b:X) corresponding to Xltt and X2.%, respectively, for a, beS2. The Aj(a:X) are ± 1 valued. The correlations are defined by P HV (a,6) = f A1(a:X)A2(b:X)dX, where the suffix stands for hidden variables. If a, b, a', b' are four points of S2 then (suppressing A)
\PBv(*,ty-PKv&h\^\JAi(a)AM
+I
U^AS'Jil+A^AMdxl
<(1±PHVK^)) + (1+PHV(^).
LOGICS ASSOCIATED
WITH HILBERT
SPACES
131
So (cf. M. Lamehi-Rachti and W. Mittig, Phys. Rev. D, 14 (1976), pp. 25432555) we obtain the inequality (18)
|P HV (a,S)-P H V (a,6')| + 1 i V («',&')+^Hv («'£)l < 2,
which must be satisfied by the correlations in any local hidden variables model. In Quantum Mechanics, P Q M ( ^ 3 ) = — 1; PHv satisfies this if and only if (19)
A^aiX) =
-A2(a:\).
If we assume (19), then (18) becomes (20)
|P HV (S,6) - P H V (a,c)| ^ 1+P H V (S,c).
This is PeW's inequality. It is trivial to check that (17) violates (20). Indeed, if 012, 023> #3i a r e the angles between a, b, b, c, and c, a, then for (17) to obey (20) we should have sin2 (0ia/2) + sin2 (023/2) ^ sin2(031/2), which is false if 612 = 023 — n/3, 631 = 2TT/Z. If we assume the model to have natural covariance properties with respect to rotations, then P H v (*W depends only on the angle 0 between a and b say P HV (a,b) = P H V (0). Obviously,
The inequalities £18) now imply bounds for Puy(6). Indeed, taking a', b' coplanar with a, b we get, writing P for P H V , (21)
\P(e)-P(d + a)\ + \P(0')+P(6' + a)\ < 2 ,
for all 9, 6', a from which one gets (cf. Lamehi-Rachi and Mittig, loc. cit.), (22)
|P(jr/6)|«f
\P(n/4)\
\3P(n/3) + P(0)\<2.
The article of Bell attracted the attention of physicists and led to experiments whether (22) or (17) is true (cf. Wigner loc. cit.). The experimental results so far support Quantum Mechanics (see some of the experimental evidence presented in the articles of the collection edited by Wheeler and Zurek, loc. cit.). Even though the basic Hilbert space has dimension 4 here, the above treatment of the hidden variables question cannot be subsumed under the von Neumann-Mackey-Gleason paradigm. The point is that for each A e X, the model gives only the values of Exp(Z1>^m-Z2,Sn) (m,n^Q); and hence each XeX can define a dispersion free state only on the maximal abelian subalgebras jtf(aj)) of J^ generated by a-a®l and l®b-<j. Hidden variable theories such as these are called contextual in the book of Beltrametti and Cassinelli (cf. notes on Chapter III). The reader who wants to
132
GEOMETRY
OF QUANTUM
THEORY
know more about these and get a sharper perspective on the entire question of hidden variables cannot do better t h a n consult the book of Beltrametti and Cassinelli, for its thorough t r e a t m e n t of related issues as well as t h e references it contains to other literature. For a question as deep as the one t h a t asks whether quantum mechanical description of the atomic world is complete, a mere mathematical analysis will hardly provide a completely satisfying answer. The argument must also be based on analyzing possible measurements and the limitations imposed on t h e m by the uncertainty relations. The paper of Einstein, Podolsky, and Rosen, and Bohr's refutation of it, offer perhaps the most profound discussion of this question from the natural philosopher's point of view (cf. the relevant articles in the collection edited by Wheeler and Zurek, loc. cit.). 2. Algebraic formulation of Quantum Mechanics. I n some sense the algebraic method of viewing the foundations of Quantum Mechanics is the oldest of all approaches, and arguably, the most flexible one. The early papers of Born, J o r d a n , Heisenberg, and Dirac repeatedly emphasized t h a t q u a n t u m observables are represented by matrices and so form a noncomm u t a t i v e algebra, in sharp contrast to the commutative algebra of classical observables (for a historical account as well as translations of the early papers, see Sources of Quantum Mechanics, B. L. van der Waerden, Dover, New York, 1967). Eventually this came to be formalized as the principle t h a t the quantum observables are in one-one correspondence with t h e selfadjoint operators on a (separable, complex) Hilbert space. I n spite of t h e overwhelming success of the models built out of this assumption it was always understood t h a t this was a very ad hoc solution to the mathematical, physical, and philosophical problems raised by atomic physics. Especially troublesome was the fact t h a t both the operations of addition and multiplication involved possibly noncommuting (hence nonsimultaneously observable) observables, t h u s making it very difficult to give a n y convincing " e x p l a n a t i o n " for them. (Perhaps one should not p u t too much faith in such " e x p l a n a t i o n s " . Most explanations make implicit or explicit use of classical perceptions and are not too relevant in the atomic realm.) There was also the fact t h a t every self-adjoint operator was supposed to represent an observable; in the presence of superposition rules one should modify this suitably, b u t even then this was gradually seen as a far-reaching enlargement of what can be effectively observed in atomic systems—see Interpretation of Quantum Mechanics, E . P . Wigner, in Quantum Theory and Measurement, edited b y J . A . W h e e l e r and W . H . Z u r e k , p . 298. Multiplication was felt to present the greater problem since it could be argued t h a t if A and B are observables, A+B can be effectively defined by prescribing t h a t its expected value in any state is the sum of the expected values of A and B in t h a t state. This, of course, is a little deceptive, for, as we have
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
133
already remarked above, the linearity of expectation values is a deep property of the propositional logic; its validity for the standard logics (in dimension ^ 3) is a consequence of Gleason's theorem while its truth for the more general logics of projections in von Neumann algebras has only recently been proved (see the remarks on Gleason's theorem, and also §14.5 of the book of Beltrametti and Cassinelli, cited in the notes to Chapter III). The first systematic investigations on the algebraic structure of the set of observables are due to Jordan, von Neumann, and Wigner; the initial papers are those of Jordan in Gott. Nachr. (1932), pp. 569-575, Gott. Nachr. (1933), pp. 209-217, and Z. Phys., 80 (1933), pp. 285-291, and that of Jordan, von Neumann, and Wigner in Ann. Math., 35 (1934), pp. 29-64. One of the motivations for these studies was the feeling that the generalizations of Quantum Mechanics to relativistic and nuclear physics might force such a more general view of observables. Jordan's basic observation was that although the set of bounded self-adjoint operators do not form an algebra under multiplication, it is nevertheless an algebra over R, even abelian, if we define a new product by A • B = \{AB + BA). This product is bilinear and satisfies the distributive laws A • (B + C) — A • B + A • C, although not the associative law; moreover, the identity A -B = \[{A +B)2 — A2 — B2] shows that this product can be obtained from the linear and power structures. Nowadays such algebras, emerging as they did out of the pioneering studies of Jordan, are known as Jordan algebras; see the book of Jacobson, The structure of representations of Jordan algebras, Amer. Math. Soc. Colloq. Publications, vol. XXXIX, 1968, Providence, R.I. Typically, one starts with an associative but not necessarily commutative algebra and define the new multiplication by a-b — \(ab + ba) to obtain a Jordan algebra. But it is not always possible to obtain a given Jordan algebra in this fashion. The fundamental result of Jordan, von Neumann, and Wigner in their 1934 paper is the complete classification of all finite dimensional Jordan algebras A over R which are irreducible and formally real, i.e., satisfy al9 ...,aneA,
ax2+...+ a2 = 0 => ax = ... =an = 0.
It follows from their classification that, with one exception (that of M38, the algebra of self-adjoint 3 x 3 matrices over the Cayley numbers), all of these are obtained in the above-mentioned manner from associative algebras; and further, that if some special cases are disregarded, one gets just the Jordan algebras of Hermitian matrices over R, C, or H. The finite dimensionality assumption is a very severe one and von Neumann began the study of infinite dimensional topological algebras (Mat. Sb., 1 (1936), pp. 415-484; Collected Works, Vol. I l l , No. 9, Pergamon, Oxford, 1961). His investigation in this paper contains a deep analysis of
134
GEOMETRY
OF QUANTUM
THEORY
t h e spectral theory of individual elements and its relationship to simultaneous observability. Although he had, in connection with his analysis of t h e possible existence of hidden variables (cf. his book), introduced the notion of the state of a system as a positive definite linear form on t h e observable algebra, this paper does not contain a study of the states. With the appearance, in the early 1940s, of the path-breaking papers of Gel'fand and Naimark on Banach algebras, the essential aspects of this approach became clearer. This theme was taken up by I . E . S e g a l in 1947 (Postulates for general Quantum Mechanics, Ann. Math., 48 (1947), pp. 9 3 0 948). Segal's postulates on the observable algebra were simpler and more direct t h a n von Neumann's, and were motivated by what is operationally possible; in addition, he studied in detail the structure of the states. The case when the (bounded) observables are the self-adjoint elements of a complex C*-algebra (a complex Banach algebra with involution * such t h a t \\x*x\\ = ||#||2 = ||#*||2, ||1|| = 1) is a possible system in t h e sense of Segal. Although special, it is an important one since it includes the conventional model of Quantum Mechanics. If U is the C*-algebra whose self-adjoint elements are in correspondence with the bounded observables of the system, a state of the system is a linear function / o n U such t h a t /(1) = 1 and f(x*x)^0 for all xeVL. The Gel'fand-Naimark-Segal representation theorem associates t o / a n essentially unique triple (J^f,nf,ifjf) consisting of a Hilbert space J^f, a unit vector i/jf in 2tfu and a *-representation 777 of U in J f f with ifjf as cyclic vector such t h a t f(x) = (7rf(x)ijjf,ifjf) (x e U); / is pure if and only if nf is irreducible. I n this way state vectors are once again unit vectors in Hilbert spaces; b u t the Hilbert space is not fixed as in conventional Quantum Mechanics, b u t varies with the state. The importance and usefulness of these more general types of observable algebras become clear in the theory of Second Quantization where one works with systems of identical particles (photons, electrons, He 3 , He 4 , etc.), b u t where the number of particles m a y not remain fixed. I n this case it t u r n s out in contrast to the case of systems of a fixed number of particles, t h a t t h e canonical commutation rules (or anti-commutation rules) have m a n y essentially different representations. Nevertheless, it was proved by Segal (Mat. Fys. Medd. Dan. Vid. Selsk, 31 (1959), No. 2) t h a t the C*-algebra, which is the norm-closure of the observables depending only on finitely m a n y particles, is canonically determined; therefore it m a y be regarded as t h e algebra of field observables. The possibility t h a t one m a y have to v a r y t h e Hilbert space with the state has also been discussed by Dirac (Lectures on Quantum Field Theory, Belfer Graduate School of Science Monograph Series, No. 3, Belfer Graduate School of Science, Yeshiwa University, New York, 1966). The study of infinite dimensional observable algebras has been pursued in recent years in another direction, closer to the original theme of J o r d a n ,
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
135
von Neumann, and Wigner. This is the study of what are called J B , J B W algebras, which are the J o r d a n algebra analogues of the self-adjoint parts of C*-algebras and von N e u m a n n algebras. The state spaces of these are compact convex sets with an interesting geometric structure, and they can be characterized in an intrinsic manner. This thus opens up another approach to the foundations in Quantum Mechanics. For a survey of many aspects of this theme we refer the reader to the article of E. M. Alfsen (On the state spaces of Jordan and C*-algebras, in Algebres d'operateurs et leurs applications en Physique Mathematique, CNRS Coll. Intern. No. 274, 1979, pp. 15-40). I n recent years people (Haag, Kastler, and others) have used algebras of operators for formulating some general properties of physical systems (for example, a quantized field) extended in space-time, t h a t are local in the sense t h a t the observables t h a t are localized in noncausally connected regions commute with each other. These ideas have clarified somewhat the nature of the high-energy scattering processes t h a t dominate elementary particle physics. The reader who wants to understand the algebraic approach m a y start with the book of G. G. E m c h (Algebraic Methods in Statistical Mechanics and Quantum Field Theory, Wiley, New York, 1972), and the references cited there. I t should be remarked, however, t h a t the mathematical questions arising in the a t t e m p t to construct rigorous models of scattering processes involving particle creation and annihilation (Quantum Electrodynamics, for example) are quite deep and are tied up with themes of an entirely different sort (cf. J . Glimm and A. Jaffe, Quantum Physics: A Functional Integral Point of View, Springer-Verlag, New York, 1981). The approach t o Quantum Mechanics via functional integrals was pioneered by R . P . F e y n m a n [2] (cf. also his book with A . R . H i b b s , Quantum Mechanics and Path Integrals, McGraw-Hill, New York, 1965) and is known as the path integral formalism. I t is widely used by physicists although mathematically difficult to justify. The attempts to understand this formalism a t a rigorous level have proved very beneficial and effective in the mathematical theory of elementary particles and their interactions; see the book of Glimm and Jaffe, loc. cit. Finally, for a view of the foundations of Quantum Theory t h a t is very close to phenomenology one should refer to the series of papers of J . Schwinger which are a p a r t of his book Quantum Kinematics and Dynamics, W. A. Benjamin, New York, 1970. 3. Statistics of mixed states. Consider a system governed by the standard logic ££(ffl) where &? is a separable complex Hilbert space, and let it be in t h e state pu where U is a von Neumann operator on #F. By the spectral theorem we have U = YinKPLn where \v\2,... are the distinct eigenvalues > 0 of U, Ln the eigensubspace corresponding t o An, and PLH
GEOMETRY
136
OF QUANTUM
THEORY
is the orthogonal projection ^->Ln. If dn = dim(Ln), wn = dn\n, Un=d^P^y then tr( C7W) = 1 , 2 nW„ = 1, and the formula (23)
Pu =
and
^wnpun n
exhibits pu as a canonical mixture of the states pun with weights wn, the supports of the p un being mutually orthogonal. If U has a simple spectrum, i.e., if dn = l for all n, t h e ^t/ n are all pure states; and (23) is the unique decomposition of p u as a convex combination of pure states defined by orthonormal vectors. If some of the dn are > 1, the right side of (23) can be split further but not uniquely; indeed, if (ttnx-)i^
PU^^Wn
2
dn^PunV
which exhibits p u as a convex combination of pure states defined by t h e orthonormal family of vectors (uni). Thus, as soon as some dn > 1 we cannot definitely say of which pure states p u is a mixture of; the phenomenon is clearly nonclassical and is another example of the sharp contrast between q u a n t u m and classical statistics. 4. Composite systems. One of the most widely used methods of analyzing complicated systems is t h a t of viewing them in terms of suitable subsystems and their interactions. For example, for m a n y purposes it is completely adequate to view an atom as a system of electrons moving in a centrally symmetric force field. We are thus led to the general problem of composition of systems. Let <5lf S 2 , • • • , S ^ be q u a n t u m mechanical systems; in conventional Quantum Mechanics we associate complex separable Hilbert spaces Jtf?1,...93tf,N9 to them with the usual prescriptions of states and observables. I t is then possible to form a composite system S containing S 3 as subsystems in a natural manner, and in fact to do this in a universal way. The logic of S is ££{#?) where Jf =
J$?1®...(x)3eN.
The m a p ia:A->l®...®A®...®l
(1 ^ a ^ N, A in the ath factor)
imbeds the algebra B(J^a) in B(3tf)\ by restricting the ^4's to the projection operators, each ia gives rise to an imbedding ia :&(Jt?a) c> jgf(jT)
(1 < a < N).
Thus the properties of the S a are faithfully reflected within S . If a is a state of S , i.e., a probability measure on j£Ppf), then (25)
aar=(TOia
LOGICS ASSOCIATED
WITH HILBERT
SPACES
137
is a probability measure on ^(J^a) and so defines a state of S a , the state that S a is in when S is in the state a. The map (26)
CT^(CT 1 J
...,GN)
is clearly surjective; if
138
GEOMETRY
OF QUANTUM
THEORY
t h e theory of composite systems in quaternion Quantum Mechanics becomes more difficult. F o r quaternionic Quantum Mechanics, t h e impossibility of forming tensor products is interpreted by Finkelstein et al.[l] to mean t h a t there are no truly independent systems, i.e., t h a t there are complementarity relations between observables of a n y two systems in quaternionic Quantum Mechanics. The issue becomes even more complicated when we treat it in t h e framework of general logics. The reader is referred t o t h e discussion in Chapter 24 of t h e book of Beltrametti a n d Cassinelli (cf. Notes on Chapter I I I ) and t o t h e literature cited there. The above considerations have t o be modified substantially when one wants t o treat systems of identical particles, because, in Quantum Mechanics there are n o experiments t h a t can keep track of individual particles throughout t h e course of some physical process. Thus, in such assemblies of identical particles, only the physical quantities t h a t involve t h e various particles symmetrically are permitted t o be observables. Mathematically, this means t h e following: if 34? is t h e Hilbert space of a single particle, t h e n for a system of N such particles, the self-adjoint operators on tfW) = 34?® ... & jtr (N times) t h a t represent physical observables must commute with t h e natural action of t h e permutation group SN of { 1 , . . . , N} on 34?{N); here we recall t h a t a n y GSSN acts on 3^{N) by sending u± (x)... ® uN t o ^o-i(D® ••• ® ^ - W ^ e ^ ) . Let U be t h e algebra of all operators on #£{N) commuting with this action on SN- T O analyze t h e structure of U we proceed as follows. For a n y irreducible character r of SN let E(T) be t h e projection operator (N! Vdim(r))- 1 £ T(*)«
on j f w
seSjf
t h a t projects onto t h e space of " r - s y m m e t r i c " tensors; if j f W ( r ) = E(T) (JirW), t h e 34?{N)(T) are mutually orthogonal, stable under SN and U with (28)
JfW = @ j f W ( T ) ; T
moreover, we can factorize each Jf{N\r) (29)
(but noncanonically), as
#W(T)KK{T)®L{T),
where U (resp. SN) acts only on t h e first (resp. second) factor in such a w a y t h a t SN acts irreducibly on L(r) with character r while U is isomorphic t o B(K(T))®1 (see Weyl [1]). Thus, if J ? is t h e logic of ^ - s t a b l e closed subspaces of ^N), then the J f iN\r) are the atoms of the center of J*?, t h e logics £?(T)=£? njfN\r) are standard (being isomorphic t o JP(K(T))9 a n d
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
139
J£? = @ T J > ? ( T ) is t h e central decomposition of «^f. Although (29) is not canonical, once we fix such a factorization, for each r, t h e pure states of S£ m a y be identified with the rays of Jf { N ) t h a t lie in some K(T), SO t h a t the K(T) are t h e coherent subspaces of pure states. The reader should view this as an example of noncommuting superselection rules. Two specific coherent spaces of pure states deserve special attention. These are t h e cases where T is one-dimensional so t h a t K(T) = ^N)(T). More precisely, we consider t h e two cases: (a) T = 1 , T(S) = 1 for all SESN; (b) T = sign, T(S) = signature of s for all s e # # . I n (a), 34?{N)(r) is t h e subspace of symmetric tensors of rank N over 34?, or the subspace of degree N, 8{N)(34?), in t h e symmetric algebra S(34?) over 34?. I n (b), 34?{N\T) is t h e subspace of skew-symmetric or alternating tensors of rank N over 34?, or t h e subspace of degree N, A{N)(34?), in t h e Grassmann or Exterior algebra A(34?) over 34?. I n atomic theory, in order to get agreement with experiment it is necessary to assume t h a t for systems of N electrons no states other t h a n those in Am{34?) occur in nature, i.e., the logic is the standard one associated with the Hilbert space Am(34?). If u{ (l^i^N) are orthonormal vectors in 34?, it is customary t o interpret U±A...AUN as " t h e state in which t h e electrons (in some order) occupy t h e states uv ...,M;v". The fact t h a t uiiA...AuiN = 0 if two of t h e ip's are identical is usually formulated as t h e Pauli exclusion principle, namely, " n o two electrons can be in t h e same state." I n radiation theory which studies photon assemblies it becomes necessary t o assume t h a t for a system of N photons only t h e symmetric tensor states occur in nature, i.e., t h e logic is t h e standard one associated with t h e Hilbert space S{N)(3f). I n either case it is necessary t o give additional prescriptions, typically by introducing specific operators, t o describe physical quantities a n d processes characteristic of the systems of particles in question. The systematic way of doing this is known as Second Quantization. I n particle physics it is assumed t h a t t h e above description in terms of A{N)(34?) or S{N)(34?) extends to assemblies of particles of a n y type whatsoever. Particles which require t h e exterior algebra description are known as Fermions a n d are said t o obey Fermi-Dirac statistics, while t h e others are known as Bosons and obey Bose-Einstein statistics. Experiment has shown t h a t particles with half integral spin (electrons, neutrons, protons, H e 3 atom, etc.) are Fermions while particles with integral spin (photons, H e 4 atom) are Bosons. All theoretical explanations of this remarkable connection between spin a n d statistics depend on t h e theory of relativistic quantum fields. 5. Measurement in Quantum Mechanics. The theory of measurement is a t the heart of Quantum Mechanics. Indeed it is the uncontrollable disturbance caused by measurements on t h e physical systems being observed t h a t leads to t h e statistical character of t h e quantum mechanical picture of t h e
140
GEOMETRY
OF QUANTUM
THEORY
physical world. It is natural to try to determine the extent to which one can understand the nature of the measurement process within the framework of Quantum Mechanics itself. Of course Quantum Mechanics gives unambiguous prescriptions for calculating the probability distribution of any physical quantity at a given instant of time, if the state of the physical system is given at an earlier time (Schrodinger's time-development equation), so that, no theory of measurement is needed to ensure the logical consistency of the theory. But the state vectors are not physical objects; they are idealizations, and one has conceptual access to them only through results of experiments. Thus, if the theory is to be confronted with experience, and is to be used as a guide for understanding existing phenomena as well as for predicting new ones, it has to contain instructions that tell us how to determine the states from physical data, i.e., how measurements on a system change its state. The fundamental work on the general theory of measurement is that of von Neumann, most of which is contained in Chapters V and VI of his book [1]. Since then, the subject has received a lot of attention and new themes have been introduced. We shall limit ourselves to a brief discussion of von Neumann's work and some of the later developments inspired by it. The reader who wants to get a historical perspective and a deeper insight into these questions should refer to the volume edited by Wheeler and Zurek (Quantum Theory and Measurement, Princeton University Press, Princeton, NJ, 1983) that collects together most of the crucial papers on the subject. For simplicity we shall restrict ourselves to the conventional model of Quantum Mechanics. Let a quantum mechanical system be in a pure state represented by the normalized vector cp (of the Hilbert space Jf7 underlying the system), and let A be a self-adjoint operator with a purely discrete simple spectrum that represents some observable. If a1,a2,... are the eigenvalues of A and
UA=Z
{U
LOGICS
ASSOCIATED
WITH
HILBERT
SPACES
141
where P[
MA:U-+UA
on the convex set of states. I t is clear from (30) t h a t MA is a convex m a p ; but it is a convex endomorphism and not a convex automorphism since in general it changes pure states into mixed ones. The essential content of von Neumann's work is a profound analysis of the mathematical a n d physical nature of the endomorphisms of the type MA. The first question is to what extent MA fails to be invertible. Von Neumann viewed this from the more general point of view of deciding whether (the effect of) MA is irreversible. To this end, he introduced also the automorphisms of the convex set of states of the form ajyiU^DUD-1;
(32)
these are induced b y unitary operators D of 3tf t h a t represent the dynamical transformations. H e then formulated the irreversibility of MA as the fact t h a t it is impossible to go from MA(U) to U by repeated applications of transformations of the form MB, OLD. For proving this he introduced, for any U, its entropy H(U): (33)
H(U) = — Ktr(Uln
U),
K > 0 being Boltzmann's constant.
Then he proved t h a t (34)
H(U)^0,
H{U) = 0oU
=p(p
for some cpeJf with |M| = 1,
and further t h a t (35) (36)
# K ( £ 7 ) ) = #(*7), H(MA(U))
> H(U)
unless
MA(U)
= U.
The irreversibility of the transition U^UA follows from (35), (36). As remarked by von Neumann himself, there are m a n y unitary invariants besides H(U) t h a t do not decrease under the measurement transition U->UA. For instance, the function t h a t assigns to U the value —A where A is the largest eigenvalue of U, has this property. But as we shall explain briefly below, the entropy function H has thermodynamic significance; and the relations (35)-(36) show t h a t the measurement transition U->UA is actually irreversible even at the thermodynamic level. Since it is easy to construct examples of changes U->U' where the entropy stays the same but the largest eigenvalue decreases strictly, we see t h a t there are thermodynamically reversible changes which cannot be brought about by repeated
142
GEOMETRY
OF QUANTUM
THEORY
applications of MB, az>. Thus the inequalities (36) go far beyond merely establishing the mathematical irreversibility of the transitions U->UA. Von Neumann was led to the definition (33) of entropy by his analysis of ensembles of noninteracting systems each of which is in the state represented by the (von Neumann) operator U. H e used the methods of phenomenological thermodynamics for his purpose. The only difference between t h e classical setting and the present one is t h a t the individual systems are now governed by the laws of Quantum Mechanics. The problem was to calculate, for an ensemble of N systems in the state U, the decrease in entropy t h a t should be achieved to bring the systems to a pure state p^. He found t h a t the pure states are transformable among themselves without any change of " h e a t energy" and so all ensembles where the individual systems are in a pure state p^ have the same entropy, which can be normalized to be 0. H e then showed, assuming the validity of the first and second laws of thermodynamics, t h a t the ensemble of N systems with state U requires a decrease in entropy equal to —Nf
-NKtr(UlnU).
The reversal of the processes used here shows t h a t two ensembles with the same entropy can be transformed into each other by thermodynamically admissible processes. This analysis, which led to the formula (37) for the entropy, also suggested the properties (33)-(36). The relation (35) asserts t h a t the entropy remains constant during time evolution if no measurements are carried out. This will appear paradoxical because classically the entropy always increases. The paradox disappears if we realize t h a t the entropy H(U) is the entropy of the microstate. The classical entropy is a macroscopic quantity and its time variations are due to the fact t h a t the observer does not know the microstate of the system. To define an analogue of the classical entropy one m a y introduce a Boolean cr-algebra 9ft of "macroscopic projections" t h a t form the logic of macroscopic quantities. If we assume t h a t 9JI is atomic and its atoms EVE2,... are finite dimensional projections where (38)
sn = dim(En)
> 1,
then it is reasonable to define the macroscopic entropy H^iJJ) £7 by (39)
-^tr(UEn)lntT{UEn).
H(U) = n
$n
Von N e u m a n n proved t h a t (40)
HniU)
> H(U),
of the state
LOGICS ASSOCIATED
WITH HILBERT
SPACES
US
with equality if and only if (41)
U =
j;(tT(UEn)/sn)En. n
Unlike the previous case, the macroscopic entropy is always > 0 for any U. Moreover, it will change under time evolution. Von Neumann investigated this as well as the associated ergodic theorems in his paper in Z. Physik, 57 (1929), pp. 30-70, to which we refer the reader for further details. The second aspect of the work of von Neumann is the investigation of the maps MA from the point of view of the Quantum Mechanics of the composite system formed by the system S under investigation and the system S ' consisting of the "measuring a p p a r a t u s . " Let 34? and 3tf" be the Hilbert spaces associated with them and let ^ = ^f?®Jf7/. Let A be as before a self-adjoint operator with a purely discrete spectrum, ava2,..., (pv 9? 2 ,... being the corresponding eigenvectors (normalized). The act of measurement of A in £> m a y be regarded as a " t e m p o r a r y " insertion of an energy coupling in the composite system t h a t allows S ' to interact with S . The time development of the composite system in the (short) time interval of measurement is then given by the dynamical group t -> exp( — itE) where E is the energy operator in 2tf. If £ is a unit vector in 3tf" representing the initial state of the measuring apparatus, and q> is a unit vector in £? representing the initial state of S , the state of the composite system after measurement is exp( — irE)((p(x)£;) where r > 0 is the duration of measurement. This state induces a state of S which m a y then be regarded as the state of (5 after measuring A. To make sure t h a t this scheme captures at least some of the key aspects of measurement it is natural to require t h a t the following conditions be satisfied. (i) (ii)
There is an orthonormal set (| n )n^i in &" such t h a t the state £ n of S ' " c o r r e s p o n d s " to the state (pn of (3. Forah>eJ^, exp( - irE)(
with M ? ) | = |(?Wn)|. The pairing (pn <-> f n is the mechanism by which the observer, recording the state f w of £>', recognizes t h a t S is in the state
2IW..)IX. = -aMP*)n
If for a given A we can find (£ w )n^i, f > ®> T s u c n t n a t W a n d (ii) are satisfied, then it would be justifiable to view the transitions U-+MA(U) of the states of S , which are irreversible within S , as the effects on S of the
144
GEOMETRY
OF QUANTUM
THEORY
perfectly reversible and continuous transitions provided in the composite system by its time development. I n other words, the causal and continuous change of states arising out of the dynamical evolution gives one a unified quantum mechanical way of looking a t the physical world and its interactions with the observer. Von Neumann proved t h a t this is possible; t h a t is, for any A, and for fixed r > 0, £, (£n)w^i> there is a suitable E satisfying (i) and (ii) above. This result of von Neumann inspired much subsequent work whose focus was on the following issue: To what extent do the requirements (i) and (ii) above as well as the operator E constructed by von Neumann give a faithful picture of measurement processes encountered in the physical world ? I t could be argued t h a t the description of S ' by the Quantum Mechanics of the logic ££{2tf") does not take into account the fact t h a t very often Q' is " v e r y big " compared to S and so the logic of S ' has a rich supply of superselection rules. The requirement t h a t measurement is of very short duration is also troublesome if these questions are to be treated relativistically. Finally, if there is additional structure in Jf, Jf", and J ^ , it is not clear t h a t the E constructed would respect them. The first example of nonmeasurability of this kind was exhibited by E.Wigner (Z. Physik, 131 (1952), p . 101). He showed t h a t if there are additive conservation laws for E, then A cannot be measured unless it commutes with the conserved quantity; and further, t h a t for approximate measurements of A with arbitrary accuracy to be possible, the measuring apparatus should be large enough t o admit states which are superpositions of sufficiently m a n y states with different quantum numbers of t h e conserved quantity. Actually for Wigner, A was the z-composnent of the spin of a spin i particle, the conserved quantity being the z-component of the angular momentum. These results were proved in full generality by H . Araki and M.Yanase (Phys. R e v . , 120 (1960), pp. 622-626). They proved t h a t if Lx (resp. L2) is a bounded self-adjoint operator on J4? (resp. ^ f ' ) , and if r > 0, vectors (£n)n^i t h a t are orthonormal in ^f', and a self-adjoint operator E on J f can all be found such t h a t U = exp( — irE) satisfies the von Neumann criteria for measurement of A for some initial state £ in 3tf\ then for U to commute with L = LX®1 + 1 ® £ 2 *t is necessary t h a t A commutes with Lv Their work also showed t h a t when Jf7 and &" are finite dimensional and the cardinality of the spectrum of L2 is allowed to be sufficiently large, then approximate measurements of A (in a certain sense) can be made with any desired accuracy. This raises the question, when A, g, and J*f' are fixed, of explicitly calculating the lower bound of the accuracy of approximate measurements of A in terms of the variance of L2 in the state £. F o r spin measurements of spin i particles (when the conserved quantity is the z- component of the angular momentum), M. Yanase (Phys. Rev., 123 (1961), p p . 666-668) has investigated this question.
LOGICS ASSOCIATED
WITH HILBERT
SPACES
145
6. Quantum Probability Theory. For a given logic j£? the notions of probability measures on ££ and observables associated to ££ define a noncommutative generalization of conventional probability theory. For standard logics one can develop such a generalization in great depth: see for instance the articles in Quantum Probability and Applications to the Quantum Theory of Irreversible Processes, Springer Lecture Notes in Mathematics, 1055 (1984), edited by L. Accardi, L. Frigerio and V. Gorini. One of the concepts that is needed for such a theory of Quantum Probability, and which goes beyond what we have treated in Chapters III and IV is that of conditional probability. Let J f be a complex separable Hilbert space and ££ = <££{3tf). Suppose £> is a probability measure on j£? and U the associated von Neumann operator. We identify elements of S£ with the orthogonal projections corresponding to them. If E is a projection and p(E) > 0, the conditional probability measure given that E has occurred is defined as the probability measure p (. \E) given by (42)
P(F\E) =
tr(EUEF)/tr(EUE).
It is not difficult to show that p(. \E) is the unique probability measure on j£f with the following property: to any F^E, it assigns the probability p(F)/p(E). Clearly p(. \E) is defined by the von Neumann operator V where (43)
(tr(UE))-1-EUE.
V=
If F is a projection commuting with E we have (44)
p(F\E)=p(EF)/p(E),
which is exactly as in the conventional probability theory; however, for arbitrary F we should use (42). Ifp=p(p where ye Jt and \\q>\\ = 1, we have (45)
P(-\E)=P+,
*P=\\E
Let E = E1 + E2+... be an orthogonal sum of projections EXiE2i... with p(En) > 0 for all n. If the projection F commutes with all the En, we have (46)
p(F\E) = 2
(p(F\En)p(En)/p(E)).
n
However, for general F, (46) is no longer true. If p =p(p (
(«, «m .? (g)>i*.>+j, f « K « . The first term in the right side of this formula is the same as that in the right side of (46). The additional terms in the right side of (47) may be viewed as arising out of the "interference effects " that are typical in quantum theory
146
GEOMETRY
OF QUANTUM
THEORY
when F does not commute with the En. For a very interesting discussion of the well-known two-slit experiments of Quantum Mechanics from the point of view of conditional probability see Chapter 26 of the book by E. Beltrametti and G. Cassinelli (cf. Notes on Chapter III). We refer to the same place for a treatment of conditional probabilities relative to a Boolean sub<7-algebra of «£?, and, more generally, relative to a sublogic of ££. 7. Measures on the projections of a von Neumann algebra: Generalizations of Gleason's theorem. Gleason's theorem may be viewed as a complete description of all countably additive probability measures on the standard logic J^(^f) when dim(Jf) ^ 3. It is natural to ask whether similar results can be proved for probability measures on the logic of projections of general von Neumann algebras. This question was first raised by G. W. Mackey (Amer. Math. Monthly, 64 (1957), pp. 45-57), and it appears that it has now been settled completely. Let ^ b e a complex separable Hilbert space and s/ a von Neumann algebra of bounded operators on £F. We w r i t e r p for the logic of projections in s/ and s/sa for the real vector space of all self-adjoint elements of s/. By a finitely additive probability measure ons/ p we mean a map \x: si p -> [0,1] such that /u(0)=0, /x(l) = l, fi(P + Q)=ti(P)+fji(Q) for P, Qesip with PQ = 0. Given such a JJL, one can associate to it the corresponding expectation functional Eu:jtfsa->R r
in the following way: if A es/8a and PA(E->PE) is the spectral measure of A on the a-algebra £% of Borel subsets of R, then (48)
E„(A) = r
\dv(\),
J — 00
where v is the finitely additive probability measure E-*n(PE) on £8. I t is easy to see that the correspondence
is bijective from the set of all finitely additive probability measures to the set of all functional
ons/p
E:jtfsa->R with the following properties: (49)
(i) E is linear (over R) on any abelian subalgebra of s/sa; (ii) E(A2) > 0 for any A e
Such functionals are called physical states of s/. The concept of a physical state is in principle weaker than that of a state ofs/ which demands linearity
LOGICS ASSOCIATED
WITH HILBERT
SPACES
147
in place of (i) while retaining (ii) and (iii). If <$/ = B(34?), Gleason's theorem implies that for any countably additive /z, E ^ is a state of B(J^). The problem of generalizing Gleason's theorem on s/p may then be split into two parts: (1) to determine for which s/ all physical states are states; (2) if a state of s# is normal, i.e., if it is countably additive on mutually orthogonal projections, to obtain representations of it in terms of the usual trace ons/ (whenever such a trace exists). The first problem focuses directly on the issues arising out of the noncommutativity of the algebra of observables, and was recognized from the beginning to have crucial significance from the foundational point of view. Since the linearity of physical states is no longer true when dim(^f) = 2 it is clear that we must assume that jrf has no central direct summand of type I 2 . Recent work seems to have now shown that for such an stf, all its physical states are actually states. This result, certainly one of the most fundamental and profound ones in the subject, is the culmination of the efforts and contributions of many people spread over several years. In addition to the famous paper of Gleason the following (partial) list of articles may be cited: J. F. Aaarnes, Trans. Amer. Math. Soc. 149, (1970), pp. 601-625; J. Gunson, Ann. Inst. H. Poincare, A 17 (1972), pp. 295-311; A.A.Lodkin, Funk. Anal. Priloz., 8 (1974), pp. 54-58; M.S.Matvieicuk, Funk. Anal. Priloz., 15 (1981), pp. 41-53, and Teor. Mat. Fiz., 48 (1981), pp. 261-265; E.Christensen, Comm. Math. Phys. 86 (1982), pp. 529-538; F. J. Yeadon, Bull. London Math. Soc, 15 (1983), pp. 139-145, and Bull. London Math. Soc, 16 (1984), pp. 145-150; and A. Paszkiewicz,.preprm£, to appear in J. Funct. Anal. For a survey of these and other relevant articles and a discussion of the ideas used in the proof we refer to the article of P. Kruszynski: Extensions of Gleason's theorem, in Quantum Probability and Applications to the Quantum Theory of Irreversible Processes, Springer Lecture Notes in Mathematics, 1055 (1984), pp. 210-227, edited by L. Accardi, A. Frigerio and V. Gorini. The second problem is really part of the study of normal states, and may be viewed as a type of Radon-Nikodym theorem. Such theorems go back to H. A. Dye (Trans. Amer. Math. Soc, 72 (1952), pp. 243-280), and the reader should consult the literature on von Neumann algebras for definitive formulations of such results.
CHAPTER V MEASURE THEORY ON G-SPACES 1. BOREL SPACES AND BOREL MAPS F r o m this point onwards we shall emphasize some of the more sophisticated and specialized aspects of q u a n t u m theory. We shall deal only with complex separable Hilbert spaces, referring to them as Hilbert spaces without a n y qualification. We pointed out in Chapter I I I the role played by representations of physical symmetry groups into the groups Aut(j£?) and A u t ( ^ ) , where =£? is the logic and £f is the state space of some q u a n t u m mechanical system. If we assume t h a t <£ is standard, then the results of Chapter I V describe rather completely the structure of 3? and Sf. I t would t h u s appear quite feasible to study the representations of the groups which are important from the physical point of view. Such a study would shed considerable light on various aspects of q u a n t u m theory. Now, t h e methods used in the theory of representations of groups are very sophisticated and depend heavily on analysis on homogeneous spaces. The object of this chapter is to present the basic mathematical theory of homogeneous spaces and the function spaces associated with them. The main reference to the theory of measure and integration on groups a n d homogeneous spaces is the book of A. Weil [1]. This however deals mostly with compact and abelian groups. The general theory which is presented in the following sections is to a very large extent the work of Mackey [2], [3], [4], [5], [6], [7]. We shall begin with the concept of a Borel space. A Borel structure on a set X is simply a o--algebra & of subsets of X\ t h e pair (X,$) will be referred to as a Borel space. Elements of & will be referred t o as t h e Borel subsets of X. We shall use this terminology only when there is no ambiguity about t h e CT-algebra involved. By the usual abuse of language we shall refer to X itself as a Borel space. If (X,3S) and (Y f€) are Borel spaces, and / i s a m a p of X into Y, / i s said to be Borel i f / _ 1 ( ^ ) ^ J*; if/ is one-one, maps X onto Y, and i f / _ 1 ( ^ ) = ^ , we shall say t h a t / is a Borel isomorphism, and speak of X and Y as being isomorphic. When X=Y and / is an isomorphism we shall call / an automorphism. 148
MEASURE
THEORY ON G-SPACES
149
The theory of Borel spaces is a very extensive one. We shall constantly use the main results of this theory in our work. For this, our references are to Halmos [1], Kuratowski [1], [2]. Articles of Mackey [5] and Blackwell [1] are also very close to our general point of view. I n the next few paragraphs we shall summarize what is essential for our purposes; proofs of these and other related results may be found in the sources mentioned above; cf. also Varadarajan [3], [4]. A Borel space (X,36) is said to be separable if (i) {x} e 36 for all x e X and (ii) there is a countable set @^3# such t h a t 36 is the smallest oalgebra of subsets of X containing Q)\ Q) is then said to be a generating class. A class Q)<^36 is said to be separating if, for each pair of points of X, there exists a set A EQ) which contains one b u t not both of them. I t is useful to note t h a t in any Borel space (X,38), if 3) generates 36 and {x} e 3$ for all x E X, then Ql is separating. I n fact, if this is not true, there will be a pair of points x, y e X {x^y) such t h a t for each A e 3, either {x,y}^A or {x,y} n A= 0. The set of all such A is a a-algebra containing 3d and hence every set in 36 has the above property. But this is impossible as {x} does not have the property. Let X be a topological space. The smallest Borel structure containing all the open subsets of X is called the natural Borel structure of X. When X is a metric space, the natural Borel structure coincides with the smallest Borel structure with respect to which all continuous functions on X are Borel. Suppose Xa (a E J) are Borel spaces and X any set. For each a e J let 7Ta be a m a p of X into Xa. Then there exists a unique smallest Borel structure on X with respect to which all the maps 7ra are Borel. If Y is a Borel space and / a mapping of Y into X, f is Borel if and only if 7ra o / is Borel for each a. If X is the product space of the Xa and ira the projection of X on Xa, X, equipped with t h e above mentioned Borel structure, will be called the product of the Borel spaces Xa. If each Xa is a topological space and its Borel structure is the natural one, then it is not in general true t h a t the product Borel structure coincides with the natural Borel structure associated with the product topology on X. This is so, however, if J is countable and each Xa has a second countable metrizable topology. If (X,36) is a Borel space and Y is a subset of X, then the class 3#Y defined by aY
= {BnY:
BeO}
is a Borel structure for Y. (7,36 Y ) is said to be the Borel subspace defined by Y. The natural injection of Y into X is Borel. Note t h a t Y itself need not be a Borel set in X. If X is a topological space with its natural Borel structure, then 36Y coincides with the natural Borel structure of Y when Y is considered as a topological space with the relative topology.
150
GEOMETRY
OF QUANTUM
THEORY
An important class of Borel spaces may be singled out in the following way. Let T be a complete separable metric space and X a Borel subset of T. Then the Borel subspace defined by X is said to be standard. The motivation for this terminology rests on the remarkable theorem t h a t two standard Borel spaces are isomorphic if and only if they have the same cardinal number and t h a t a standard Borel space is finite, countable, or has the power of the continuum. Standard Borel spaces are separable. If (X,&) is a standard Borel space, (Y f€) is separable, and if/ is any one-one Borel m a p of X into Y, then (i) the range Yf of / is a Borel set in Y, (ii) the Borel subspace defined by Yf is standard, and (iii) / is a Borel isomorphism of X with Yf. If (X^&i) (i = l,2,- • •) is a sequence of standard Borel spaces, then their product is standard. A Borel subspace of a standard Borel space is standard if and only if it is defined by a Borel set. Any uncountable standard Borel space is isomorphic to the unit interval. If X is any locally compact Hausdorff space satisfying the second axiom of countability, the Borel space obtained by equipping X with its natural Borel structure is standard. A measure on a Borel space is any nonnegative set function which is countably additive on its Borel structure; +00 is an allowed value. A measure p on (X,&8) is finite if p(X) < 00; is a-finite if X = Un -^n> where each Xn is a Borel set and for each n, p{Xn) < 00. p is said to be standard if there exists a Borel set X 0 such t h a t (i) p(X — X0) = 0 and (ii) the Borel subspace defined b y X0 is standard. Let X be a locally compact Hausdorff space satisfying the second axiom of countability. By a Borel measure on X we shall understand a measure with the property t h a t the measure of a n y compact set is finite. Any Borel measure is cr-finite. Let CC(X) be the linear space of all complex valued continuous functions on X with compact support and let \L be any Borel measure on X. Then
£:/->£/^
(feCc(X))
is a linear functional on CC(X) such t h a t fl(f)>0 whenever / is real and > 0 . Conversely, if A(/—>• A(/)) is a linear functional on CC(X) such t h a t A(/) > 0 w h e n e v e r / i s real and > 0 , there exists a unique Borel measure /x on X such t h a t A = /x. This is the well known Riesz theorem (cf. Halmos [1], pp. 216-249). Any Borel measure on X is regular, i.e., for any Borel set E^X, (1) fi(E)= sup p(C). C compact
Given a Borel space (X,&) and a a-finite measure p on it, we call a set A^X p-measurable if there exist Borel sets Bx and B2 such t h a t B±^A^B2 and p(B2 — B1) = 0. When there is no ambiguity about the measure t h a t is involved, we speak simply of measurable sets. The
MEASURE
THEORY
ON G-SPACES
151
collection of ^-measurable sets is a cr-algebra and hence defines a Borel structure on X, say &p, called the p-completion of &. The measure p has a unique extension to &v as a measure. The functions which are Borel with respect to the Borel structure &v are called ^-measurable. If / is any ^-measurable function on X, there exists a Borel function g such t h a t f=g almost everywhere, i.e., {x : f(x) y^g(x)} e &v and has ^-measure zero. Let (X,&) be a Borel space, let q be any cr-finite measure, and let p be a measure. We shall say t h a t p is absolutely continuous with respect to q, p«q in symbols, if p(E) = 0 for any Borel set E for which q(E) = 0. If / is a nonnegative Borel function, the function p \ E —> J fdq is a cr-finite measure which is « g ; p is finite if and only if/ is an element of ^?1(q). Conversely, if p is a cr-finite measure « g , there exists a nonnegative Borel function / on X such t h a t p(E)=\fdq for all Borel sets E. f is essentially uniquely determined by p in the sense t h a t if / ' is a Borel function and p(E)—\ fdq for all Borel sets E, then / = / ' g-almost everywhere. / is called a Radon-Nikodym derivative of p with respect to q and is denoted by dpjdq. If p, q, r are cr-finite measures, then p<^q and q«r implies p « r and the relation dp\dr = dpjdq. dqjdr holds r-almost everywhere. I n particular, if p<^q and q<£p, p and q have the same null sets and dp\dq-dq\dp = \ almost everywhere; both the derivatives are then positive almost everywhere. Let p be any cr-finite measure on X, and let us choose disjoint Borel sets X1,X2, • • • such t h a t X = U n Xn and p(Xn)
q(E) =2CnP(EnXn)
{E*&)
n
is finite, and p and q are mutually absolutely continuous. The relation of mutual absolute continuity is obviously an equivalence relation in the set of all cr-finite measures on X. The corresponding equivalence classes are called measure classes on X. If £ is a Borel automorphism of X and p is a cr-finite measure on X, we define the cr-finite measure pl by (2)
p\E)
= p(t-\E))
(Ee&)
p is said to be invariant if pl=p; p is said to be quasi invariant if pl and p are mutually absolutely continuous. Clearly p is quasi invariant if and only if the class of null sets of p is invariant under t. A measure class is said to be t-invariant if it contains a cr-finite measure quasi invariant with respect to t. We shall now prove two lemmas of a technical nature. They are of importance in the sequel.
152
GEOMETRY
OF QUANTUM
THEORY
Lemma 5.1 (Federer-Morse [1]). Let X and Y be two mectric spaces and let f be a map of X onto Y. Suppose that there exists a sequence {Kn} of compact subsets of X such that (i) K1^K2^ • • •, (ii) X = ^Jn Kn, and (iii) for each n, the restriction of f to Kn is a continuous map of Kn into Y. Then there exists a Borel set E^X such that (i) / maps E onto Y, (ii) for any yeY, E meets / _1 ({?/}) in exactly one point, and (iii) Enf-Hf[Kn])
{m(y):yeY}
and E = g[F]. B y its definition, F meets each set h'1^}) exactly once. Hence h is one-one and maps F onto Y. From this it follows immediately t h a t / is one-one on E and maps E onto Y. I n other words, E meets each set f'Hty}) ( y e Y) exactly once. I t remains to prove t h a t E is a Borel set. For any integer n>\, An<^C be defined by
let
An = {c : c e C, h(d) ^ h(c) for a n y d eC with d < c — l/n}. We claim t h a t An is open in C. Suppose this is not true. Then there exists a sequence cs in C such t h a t cs $ An for all s = 1,2, • • •, but cs - > c e ^4n as 5 - > oo. Since cs $ An there exists a c / < cs — (l/n) in (7 such t h a t h(cs') — h(cs). If s± < s2 < • • • is a sequence of integers such t h a t cSk' -> c' as k - > oo, then c' e C , c r < c —(1/TI), and A(c') = A(c), a contradiction. ^4n is therefore open. We show next t h a t F = (~}n An. If c e F, it follows from the definition of m(y) t h a t for any c'
MEASURE
THEORY ON G-SPACES
153
the right of c. Hence there exists a d > 0 such t h a t all points of T are at a distance >d away from c. If (l/n)
En = {Dn-Dn
n (f-VLKn-i)))}
u #„-!•
We claim t h a t the 2£n's have all the desired properties. Clearly E1s^E2— We claim t h a t for each n, / is one-one on En and / [ ^ n ] = = / [ J ^ n ] - This is proved by induction on n. Since E1 = D1, this is true for n = \. Suppose now t h a t n > 1 and t h a t / is one-one on En _ x with f[En _1] =f[Kn _ -J. Then the formula (3) implies easily t h a t / is one-one on En and maps En onto f[Kn]. Let E = {JnEn. Then i£ is a Borel set, / is one-one on E, and f[E]=Y. Since En^Kn, it is obvious t h a t i£ n / - ^ / ^ ] ) ^ ^ for all n. The lemma is completely proved. Corollary 5.2. Let X and Y be standard Borel spaces and f a Borel map of X into Y. Let p be a finite measure on X and q the measure E -> p(f ~ 1(E)) defined on Y. Then, the range f[X] of f is q-measurable and its complement has q-measure zero. Moreover, there exist Borel sets A and Z such that (i) i c l , Z^f[X], (ii) q(Y— Z) = 0, and (hi) A is a section for f over Z, i.e., f is one-one on A and maps A onto Z. Proof. The corollary is trivial if X is countable. Hence we consider the case when X is uncountable. We m a y assume t h a t X and Y are both identical with the Borel space associated with the unit interval [0,1]. B y a theorem of Lusin (Halmos [1], p. 243) we can find compact sets Kns^X such t h a t ( a j ^ c ^ c . . . , (b) p(X — {Jn Kn) = 0, and (c) / i s continuous on each Kn. Let Z = {Jn f[Kn]. Each f[Kn] is compact and hence Z is a Borel set. Since \JnKn<^f~1{Z), p{X-f~1(Z)) = 0 so t h a t q(Y-Z) = 0. This already shows t h a t / [ X ] is g-measurable and q(Y — f[X]) = 0. We can now apply the Federer-Morse lemma to [JnKn and construct a Borel set A^X such t h a t / is one-one on A and maps A onto Z. The proof of the corollary is complete. Remark. I t is actually true (though it is harder to prove) t h a t the r a n g e / [ X ] is a universally measurable set; i.e., for any finite measure a on Y, f[X] is an a-measurable set. Corollary 5.2 is valid even in this case,
154
GEOMETRY
OF QUANTUM
THEORY
with q replaced by a. We shall not prove these results but refer the reader to the paper of von Neumann [2] and Kuratowski [1], [2]. We refer to this stronger version of corollary 5.2 as von Neumann's cross-section lemma. Suppose X and Y are Borel spaces. Let v be a measure on Y, and for each x e X let us be given a ^-equivalence class of Borel functions on Y. The problem we now want to examine is whether we can select for each x a representative f(x,.) in the corresponding class such that the function x, y ~^f(x,y) is Borel on Xx Y. We need two auxiliary lemmas. Lemma 5.3. Let X be a Borel space, 8 a separable metric space, and f a Borel map of X into 8. Then there exists a sequence {fn} of Borel maps of X into 8 such that (i) each fn takes only countably many values, and (ii) fn(x) ->/(#) uniformly for x e X (as n -> oo). Proof. Fix the integer n> 1. Since 8 is a separable metric space, we can cover 8 by a countable family {Bm} of balls of radius 1/2%. Let the sets C m be defined by C1 = B1 and Cm = Bm — {Jj<m B} (m>l). Then the Cm are disjoint and ^ = l J m C r Since Cm^Bm, diameter of Cm< \\n for all m. For each m such that Cm is nonempty, let sm be some point of Cm. We define fn as the map
/»(*) = *« if
xef'HCJ.
Then/ n is Borel, has countably many values at most, and dist(/ n (#), f(x)) < \\n for all xe X. Lemma 5.4. Let X be a Borel space and 8 a separable Banach space. Let K be a subset of S*, the dual of 8, with the property that the linear combinations of elements of K are dense in 8*. Iff is a map from X to 8 such that for each he K, the function x -> k(f(x)) is Borel on X, then f is a Borel map of X into 8. Proof. Let ^ * be the smallest Borel structure on 8 with the property that all the elements of S* are Borel functions with respect to it. From our assumption about K it follows easily that for each A e&*, f~1(A) is a Borel subset of X. To prove that / is Borel, it is enough to prove that &* coincides with the natural Borel structure & of 8. Obviously, &*^&. Since £% is generated by the open sets, it suffices to prove that any open set belongs to «^*. Now, as 8 is separable, every open set is a union of countably many balls. Thus it is enough to prove that «^* contains all balls. Further, as x -> ex (c e C) and x -> x + a (a e 8) are easily seen to be automorphisms of the Borel space (8,&*), we are reduced to proving that the unit ball B = {x : \\x\\ < 1} lies in ^ * . Now, it is well known that under the weak *-topology for 8* ( = the smallest topology for 8* such that k -> lc(x) is a continuous function on S* for each xeS), the unit ball B* of
MEASURE
THEORY
ON G-SPACES
155
S* is a compact metric space (Dunford-Schwartz [1]), hence separable. Let { ^ ^ ^ ^ - - J b e a countable dense subset of t h e unit ball of S* in the weak *-topology. Then (by the Hahn-Banach theorem), B = {x : \x*(x)\ < 1 for all x* e B*} = 0 {x
:
\xn*(*)\
< 1 for all n).
n
This shows t h a t 5 e J * and completes t h e proof t h a t &* = &. This proves t h a t / is a Borel m a p . Now we are in a position to formulate a n d prove t h e lemma on the selection of representatives. Lemma 5.5. Let X and Y be Borel spaces, with Y separable. Let /x and v be a-finite measures on X and Y, respectively. Let f be a complex valued function on Xx Y and let Y be a union of Borel sets Y{ of finite v-measure such that (i) for each x e X, y ->f(x,y) is a Borel function on Y which is v-integrable on each Yt, (ii) for any i and any Borel set F^Y{, x -> J", f(x,y)dv(y) is a Borel function of x. Then, there exists a complex Borel function f* on XxY and a Borel set N^X of fx-measure zero such that for each xeX — N,
W
f*(*,y) = f(*,y)
for v-almost all y. Proof. We m a y clearly drop down on one of t h e Yt so t h a t we m a y assume t h a t v{ Y) < oo. Also since fi intervenes only in terms of its null sets, we m a y replace /x b y a finite measure mutually absolutely continuous with respect to JJL; hence we assume t h a t /x is also finite. Write S = J£>1(v), t h e Banach space of (equivalence classes of) v-integrable Borel functions on Y. 8 is a separable Banach space since the Borel structure of Y is countably generated. Since y->f(x,y) is a v-integrable Borel function on Y, it defines an element, s&y f(x), of 8. We claim t h a t x-+f(x) is a m a p of X into 8 which satisfies t h e conditions of lemma 5.4. Now the dual of ££x{v) is J£™{v) and every bounded Borel function on Y is a uniform limit of linear combinations of characteristic functions XF of Borel subsets F of Y. XF defines t h e linear functional tF : u-> [ u(y)dv{y) on ^(y). B y hypothesis, x -> tF(f(x)) is Borel for all F. We m a y therefore conclude from lemma 5.4 t h a t x->f(x) is a Borel m a p from X to 8. So X is the disjoint union of countably many Borel sets on each of which / is bounded. We m a y assume therefore t h a t / itself is bounded. By lemma 5.3 there exists a sequence {/„} of Borel maps of X into S such t h a t (i) each fn is bounded and takes only countably many values, a n d (ii)/ n (:r) ->f{x) uniformly in XEX. Since fn is bounded and takes
156
GEOMETRY
OF QUANTUM
THEORY
only countably m a n y values, it is obvious t h a t there exists a /x x vintegrable Borel function fn* on I x Y such t h a t for each xeX, the function y->fn*(%,y) defines the element fn{x) of S. We then have:
sup xeX
\fn*(x>y)-fm*(z>y)\My)-+o JY
as n, m —> oo and hence
J Jl/,*(*.y)-/»*(*.y)|% x ")(*,») -* o XxY
as n,m-> oo. As ^(fxxv) is complete, there is a Borel function / * on X x Y such t h a t / * is (p x y)-integrable a n d
JJ|/»*(*,y)-/*(*.y)|^xv)(a;,y)-»«0 Xxy
as n->co. Select a subsequence {/nfc*} such t h a t fnk*(x,y) -+f*(%,y) for (//, x y)-almost all x, y. Then there exists a Borel set N of ^-measure zero such t h a t for each x e X — N, fn*(x,y) ->f*(%,y) for v-almost all y e Y. I t is clear t h a t for each x e X — N, f*(x,y) =f(x,y) for v-almost all y. This proves the lemma. Corollary 5.6. Let X, Y, p, and v satisfy the same restrictions as in lemma 5.5. Suppose that Q(x->qx) is a map from X into the space of all finite measures on Y such that (i) qx«.v for all x e X, and (ii) for each Borel set Fs= Y', x -> qx{F) is a Borel function on X. Then, there exists a Borel function f* on XxY and a Borel set N^X of fx-measure zero such that for each x e X — N, y ->f*(x,y) is a version of dqx\dv. Proof. For each x e X select a Borel function fx on Y such t h a t fx = dqxjdv, and define / on 1 x 7 by f(x,y)=fx(y). The corollary follows from the lemma at once.
2. LOCALLY COMPACT G R O U P S . HAAR M E A S U R E The groups which are commonly encountered in physical problems, such as the Lorentz group or the rotation group, are topological groups (cf. Pontrjagin [1] for the theory of topological groups). Moreover t h e underlying topological spaces are locally compact and satisfy the second axiom of countability. Because of this and other technical reasons, we shall restrict ourselves in the sequel to locally compact groups satisfying the second axiom of countability (lcsc). The Borel structures associated with the underlying topologies of these groups are standard. We shall
MEASURE
THEORY
ON G-SPACES
157
describe in this section a few known facts about lcsc groups and H a a r measures. A detailed t r e a t m e n t can be found in Halmos [1] (pp. 266-289). Suppose t h a t G is a group which is at the same time a Borel space. We shall say t h a t G is a Borel group if the map x, y - > xy ~1 of G x G into G is Borel; G is called a separable (standard, etc.) Borel group if the underlying Borel structure is separable (standard, etc.). I n this book we shall deal only with separable Borel groups. If G is a separable Borel group, the maps x - > x'1, x —> xh, and x-^hx (h e G) are Borel automorphisms of the Borel space G. Any lcsc group is a standard Borel group. Suppose G is any lcsc group. I t is a classical fact t h a t there exists a nonzero cr-finite measure \L on G such t h a t fx is left invariant, i.e., p(E) = JJL(XE) for all Borel sets E and all x e G. /x is essentially unique in the sense t h a t if /x' is another cr-finite left invariant measure, then \x! = C-(JL for some constant c>0. /x is called a left Haar measure on G. /x( £7) > 0 for every open set U and fx is a Borel measure. Similarly there exists a nonzero Borel measure /xr which is right invariant, i.e., fjLr(E) = /jLr(Ex) for all Borel sets E and x e G. Again, the right invariance determines /xr u p to multiplication b y a positive constant. /xr is called a right Haar measure on G. I n general, it is not true t h a t a right H a a r measure is left invariant. When this is so we shall say t h a t G is a unimodular group. Suppose G is any lcsc group and t h a t /x and /xr are left and right H a a r measures on G. For a n y x e G, the measure E - > fi(Ex) is also a left H a a r measure and hence there exists a constant A(x) > 0 such t h a t (5)
^Ex)
=
A(xME)
for all E. Clearly A does not depend on which left H a a r measure we use in (5). The function A(x - > A(#)) is known to be a continuous homomorphism of G into the multiplicative group of positive real numbers, i.e., (i) (6)
A(e) = 1
(e the identity of G),
(ii)
A(xy) = A(x)A(y)
(iii)
A is continuous.
(x,y e G),
(i) and (ii) are easy consequences of (5); the third needs a little work (cf. Loomis [2], pp. 117-120). A is called the modular function on G. If/ is a continuous function vanishing outside a compact set of G, we have, from (5), (7)
f f(zx)dfi(z)
= A(a)-1 f
JG
f(zW(z).
JG
From this it follows t h a t (8)
f f(z)A(z)-^(z) JG
= f JG
f(zx)A(z)-^(z).
158
GEOMETRY
OF QUANTUM
THEORY
The equation (8) shows that the measure /xr defined by (9)
^(E) = f A(2" i)dM(z)
is a right Haar measure. It follows from (9) that \xr and \i are mutually absolutely continuous; the Radon-Nikodym derivatives d^jd^ and d[L\d[Lr are, respectively, the functions z -> A(2) -1 and 2 -> A(z). A simple calculation using (9) shows that (10)
^(stf) =
A(x)~^r(E)
for all Borel sets E and all ^ e G . It is clear from (9) that G is unimodular if and only if A is the trivial homomorphism, i.e., A(z) = 1 for all ze G. Since x -> log A(#) (# e G) is a continuous homomorphism of G into the additive group of real numbers which does not have any nontrivial compact subgroups, it follows that A is trivial on any compact subgroup of G. Compact groups are therefore unimodular; in this case /x and /xr are finite. It is customary, when G is compact, to reserve the term Haar measure to the unique left and right invariant measure giving measure 1 to the whole of G. If G is not compact, fji and jjir are not finite. Abelian groups are trivially unimodular. Since A is a homomorphism, A(z) = l whenever z is of the form xyx~xy~x for x, y G G. Let G' be the smallest closed subgroup of G containing all elements of the form xyx~xy~x (x, y e G); A is then identically 1 on G'. Hence a sufficient condition for G to be unimodular is that G = G'. There exist groups which are not unimodular. Suppose that G is the group of all matrices g,
ly
x
\0
l)
\
(x, y real, y > 0).
We identify g with the upper half-plane by the mapping g —> (x,y). If
^E)=jyhdxdy'
^E)=\\\dxdy*
E
E
then /x and \xr are, respectively, left and right Haar measures on G.
3. ^-SPACES Suppose that X is a Borel space and G is a separable Borel group. We shall say that X is a G-space if for each g e G, there exists a Borel automorphism tg(x -> g • x) of X such that (i) te is the identity, (11) **i*a = **Aa
fai'
£2 e 0).
MEASURE
THEORY
ON
G-SPACES
159
X is said t o be a Borel G-space if the map g,x->g-xofGxX into X is Borel. We shall say t h a t X is a standard Borel G-space if X is a Borel G-space and if the Borel structure of X is standard. For any G-space X and any x e X, the set (12)
Gx = {g:geO,
g-x = x)
is a subgroup of G. I t is called the stability subgroup a t x. The set (13)
G-x = {g-x: g e G}
is called the orbit of x. If y = h-x (he G), then Gy = h-Gx-h~1 as is easily verified. Thus the stability subgroups of the points on the orbit of x are conjugate to Gx and any subgroup conjugate to Gx is the stability subgroup of some point on the orbit of x. If X is a Borel G-space and Y is a subset of X, we shall say t h a t Y is G-invariant if Gx^Y for all x e Y; in this case, t h e Borel subspace defined by Y becomes in a natural and obvious fashion, a Borel G-space. I t is called the Borel G-subspace defined by Y. If Xx and X 2 are two Borel G-spaces, and / a Borel m a p of Xx into X 2 , then / is called a G-homomorphism if for all x±e Xl9 and all g e G, (14)
f(9-x1)
=
g-f(x1).
A G-isomorphism is a G-homomorphism / with the additional property t h a t / i s a Borel isomorphism of X1 onto X 2 . If there exists a G-isomorphism of X x onto X 2 , we shall call X x and X 2 isomorphic Borel G-spaces. Suppose t h a t X is a locally compact Hausdorff space satisfying t h e second axiom of countability and, further, t h a t G is a lcsc group. Let X be a G-space with the additional property t h a t the map g,x—>g-x of G x X into X is continuous. Then X is obviously a Borel G-space. We shall refer to it as a locally compact G-space. G is said to act continuously on X. I n this case, for each g e G, x -> g-x is & homeomorphism of X onto itself. Since G is a countable union of compact sets, G-x is also a countable union of compact sets for any x e X. Every orbit in a locally compact G-space is thus a Borel set. The stability subgroups are, moreover, closed subgroups of G. A m e a s u r e p o n a Borel G-space X is said to be invariant ifp(g- E) =p(E) for all Borel sets E^X and all g e G. I n m a n y problems it is necessary t o work with measures which are only quasi invariant. A measure p on X is said to be quasi invariant if p(E) = 0 if and only if p(g-E) = 0 for g G G, i.e., if the action of G on X transforms j9-null sets into ^-null sets. Notice t h a t if p is quasi invariant a,ndf1,f2 are two functions on X which are equal almost everywhere, their transforms f±9 and f29 also have this property for a n y g e G, where ft9(x) =fi{g~1-x) (x e X, £ = 1,2). For a n y measure p on X and any g e G, let us define the measure p9 b y (15)
p'(E)=p(g-1-E).
160
GEOMETRY
OF QUANTUM
THEORY
9
p is quasi invariant if and only if p and p are mutually absolutely continuous for all g e G. Therefore, if p is quasi invariant, so is any measure mutually absolutely continuous with respect to p. If X is a locally compact (r-space and p is a Borel measure, then so is p9. A simple example of a (r-space (0 any lcsc group) is obtained when we take X — G and define, for x e G and geG, g-x = Lg(x) = gx. Obviously G — Ge and the stability subgroups are all trivial. If yu is a left H a a r measure, /x is invariant. If /xr is a right H a a r measure, \xr is not in general invariant; but it is quasi invariant since it is mutually absolutely continuous with respect to fx (cf. (9)). A Borel (r-space X is said to be transitive or homogeneous if X = G- x for some x e X. I n this case, X — G-y for each y e X. The above example is a simple example of a transitive (r-space. If X is a standard Borel (r-space which is transitive and if the stability groups associated with the points of X are all trivial, then we shall say t h a t X is an affine space associated with G. If we assume G is standard and choose a point x e X, the m a p g -> g • x is a Borel isomorphism of G onto X which is a (r-isomorphism of the (r-space G (G acting on itself by left translations) on the (r-space X. B u t this isomorphism depends on the choice of x. We shall conclude this section with a theorem due to Varadarajan [2] which imbeds a standard Borel G-space in a compact metric (r-space, G being any lcsc group. Theorem 5.7. Let G be a lcsc group and X be any standard Borel G-space. Then there exists a compact metric G-space Y, and an invariant Borel set E^Y such that X is G-isomorphic to the Borel G-subspace of Y defined byE. Proof. Let /x be a n y left H a a r measure on G. Denote by 3?1 the Banach space of (equivalence classes of) complex Borel functions f on G such t h a t (16)
\\f\l, = J \f(x)\dp(z)
< oo.
UhsQ and fh(x)=f(h~1x) for x e G, fh e J^ 1 and | | / 1 i = |l/lli- I t is h well known t h a t f,h^f is a continuous map from J^f1 x G into J^f1 (Loomis [2], p . 118). Let !F be the complex vector space of all complex valued bounded Borel functions on X, and for u e !F let (17)
Halloo = sup
\u(x)\.
xeX
Since the m a p g,x->g~1-x of GxX into X is Borel, the function g,x->u(g~1-x) is Borel for each ue^. If feJ?1, the function x - > u 1,x jG f(9) (9~ )dv<(g) is Borel on X; we shall denote it b y / * u. I t is obvious t h a t / * ue^F and | | / * ^||oo< \\fWx' |M|a>. u-*f* u is a linear m a p of J ^ into itself.
MEASURE
THEORY ON OS PACES
161
For any nonempty set ^ c ^ w e write Y[Q)) for the set of all complex valued maps c{f,u->c(f,u)) on J ^ x i ^ such t h a t (i) for each ue@, / - > c(/,w) is linear on JS?1, and (ii) \c(f,u)\ < | | / | | r \\u\\«, for a l l / , u e &1 x 2. We equip Y(@) with the smallest topology with respect to which all the maps c->c(f,u)(fE^?1,ue^,ceY(2)) are continuous. From the defining properties (i) and (ii) it is obvious t h a t Y(@) is a compact Hausdorff space. If Q) is countable, it is clear t h a t Y(@) is even a compact metric space. We shall now convert Y(@) into a 6r-space. We define, for c e Y{Q)) and g e G, the element g • c by (18)
g-c(f,u)
c(f^1\u),
=
where, as usual, fh(x)=f(h-1-x) {x,heG). Since | | / f f _ 1 | | i = \\f\\lf it is clear t h a t g-ce Y(@), and a simple calculation shows t h a t Y(Q)) is a (r-space. We now claim t h a t the m a p g, c-> g-c of G x F(i^) into F ( S ) is continuous. I n view of the definition of the topology on Y(@) this reduces to proving t h a t for each (f,u) e ££x x 2, the map g, c - > c(/ g - 1 ,w) from ^ x 7 ( S ) into the complex numbers is continuous. However, an easy calculation shows t h a t the inequality
\c(f°,u)-c'(f\u)\
< l^p^-c'ip^l
+
Wp-f^WuW^
is valid for all c, c' e Y{Q)) and g,he G, and the right side tends to 0 when h->g and c' - > c. G thus acts continuously on Y(Q)). We shall now imbed X in F ( ^ ) for a suitably chosen 3i. Let AX,A2, • • • be a sequence of subsets of X which generate the Borel structure of X, and let un be the function which is 1 on An and 0 on X — An. Let 2 = {ul3u2,--'} and let Y—Y{2). Since 2 is countable, F is a compact metric space. For any x e X let us define cx by (19)
c x (/,») = ( / * « ) ( * )
( / e J2?1, « e # ) .
Obviously cxe Y. x -> cx is thus a map, say £, of X into F . Since the m a p # — > ( / * tt)(a;) is Borel for all fixed / , w, it follows t h a t f is a Borel m a p of X into F . Further, if <7 e G, we have: <W(./» = = =
(f*n)(g-x) jj(h)u((h-i.g).x)dn(h) jGf(gt)u(t-^x)dfM(t)
= (9'<>)(f,u). This proves t h a t £ is a Borel 6r-homomorphism from the 6r-space X into the 6r-space F . Let E be the range of £. 2£ is clearly invariant.
162
GEOMETRY
OF QUANTUM
THEORY
We assert t h a t £ is one-one. Suppose in fact t h a t x, y are two points of X such t h a t (/ * u)(x) = (f * u)(y) for all / , u e &1 x Q). Then, for each ue@, u(g~l •x) = u(g~1 -y) for almost all geG. As Q) is countable, there exists a Borel null set N^G such t h a t for any g e G — N, u(g~1-x) = u(g~1-y) for all ueQ). Let <70 be any point of G — N, and let # 0 — #o~ 1 *x> y0 = g0~1 -y. As the Borel space X is separable and as Al9 A2i- • • is a generating family, it is also a separating family (cf. Section 1). Hence, u(x0) = u(y0) for all u e Q) implies t h a t x0 = y0. Therefore x = y. Therefore f is a one-one Borel m a p of X into Y. Since X and Y are standard, this implies t h a t the range E of f is a Borel subset of Y and t h a t | is a Borel isomorphism of the Borel space X onto the Borel subspace of Y defined by E. Since we have already proved t h a t E is an invariant subset of Y and t h a t £ is a (r-homomorphism, the proof of theorem 5.7 is complete. Corollary 5.8. Let Gbea Icsc group and let X be a standard Borel G-space. Then, for any x e X, the stability subgroup Gx (of G) at x is closed. Moreover, the orbit G-xof any x in X is a Borel subset of X. I n fact, it is enough, in view of theorem 5.7, to prove these facts when X is an invariant Borel subset of a compact metric space on which G acts continuously. I n this case, both assertions are obvious. We shall end this section with two technical results which describe the manner in which the quasi-invariant measures on a Borel 6r-space transform under the action of G. For any cr-finite measure p on X we define P9 by (15). Lemma 5.9. Let X be any Borel G-space, G, a Icsc group. Let p be a a-finite quasi-invariant measure on X and let fp(g,.) be a version of the Badon-Nikodym derivative dpjdp{g~1S). Then for fixed glfg2e G, (20)
fp(9ig2,x)
= fp(9i,g2 • x)fp{g2tx)
for p-almost all x. Proof. We begin the proof with a formula which describes how a Radon-Nikodym derivative transforms under a Borel automorphism. Let /x, v be a-finite measures on X such t h a t IJ,«V. Let t(x->t-x) be a Borel automorphism of X. Define the measures yf and vl by setting fjLt(E) = fM(t-1(E)) and vt(E) = v(t~1(E)), for all Borel sets E. Then / x i «v t . If E is any Borel set, li\E)
= f Jt 1(E) = f JE
(dfildv)(x)dv(x)
(dfjildv)(t-1'x)dvt(x),
MEASURE
THEORY
ON OS PACES
163
showing t h a t (21)
drf/diS = (dfji/dvY.
This said, we come to the actual proof. Fix gl9 g2 e G. Then p{929i)~ (p i) 2 and dpldp(929i)=:(dpldp92)-(dpldp9i)92 from what we said above. This implies easily t h a t 9 9
fP(9i92>x) =fp(9i>92'V)fp{92>v) for ^-almost all x. Our next result tells us t h a t we can choose these Radon-Nikodym derivatives so they are " a l m o s t " Borel on GxX. Theorem 5.10. Let G be a Icsc group, \x a left Hoar measure on G, and let X be a separable Borel G-space. Let p be a a-finite quasi-invariant measure on X. Then, there exists a Borel function Fp on GxX which is positive everywhere such that for ^-almost all g e G, x -> Fp(g,x) is a version of the Radon-Nikodym derivative dp/dp^'^. In particular, the identity (22)
Fp(gig2,x)
=
Fp(gi,g2-x)Fp(g2,x)
is satisfied for JUXJLIX p-almost all (gi,g2,x) e GxGx
X.
Proof. We s t a r t with the special case when p is finite. We first claim t h a t for each Borel set E^X, g~^p(g-E) is Borel. I n fact, the map t : g, x—>g, g-x is a Borel automorphism of GxX and hence A = t[GxE] is a Borel subset of GxX. Let ^ be a left H a a r measure on G. Then we know t h a t for each g e G, the set Ag = {x : (g,x) e A} is a Borel subset of X and g - > p(A9) is a Borel function on G. Since Ag = gE, our claim is proved. We are now in a position to apply corollary 5.6. I t follows from this corollary t h a t there exists a Borel function Fx on GxX such t h a t for almost all g e G, x-> F^g^x) is a version of dp{9~^\dp. Since p is quasi invariant, for each g, F-^g.x) is j9-almost everywhere positive and hence, by the Fubini theorem, F1 is positive for almost all points of GxX. We m a y thus assume t h a t F1 is positive everywhere without altering the conclusions. P u t Fp = \jF1. Then F is a positive Borel function on GxX and for /x-almost all g,x-^ Fp(g,x) is a version of dp/dp(9~1}. Finally, the identities (22) follow from (20). This completes the proof of the theorem when p is finite. If p were a-finite, we choose a finite measure q such t h a t q<&p and p«q. Let a(x -> a(x)) be a positive Borel function which is a version of the Radon-Nikodym derivative dp/dq. Choose Fq as a Borel function on GxX corresponding to q, and define (23)
Fp(g,x)
=
a{x)-Fq(g,x)-a-1{g^x).
164
GEOMETRY
OF QUANTUM
THEORY
The identity dpjdp™ =
dpldq-dqldq™.(dqldp)™
implies t h a t Fp has all the required properties.
4. T R A N S I T I V E ^-SPACES We shall now examine the transitive G-spaces somewhat more closely. Let G be a lcsc group. Let GQ be a closed subgroup of G. Let X = GjGQ, the space of all left cosets gG0, g e G. Let ft be the canonical mapping of G onto X defined by (24)
p-.g-^gGo,
and for any element g of G and a; = hG0 of X let <7 • x be defined by (25)
g-x =
ghG0.
Clearly, g-x depends only on g and x and not on the element h used to define x. G t h u s acts on X and this action is transitive. Let X be equipped with the quotient topology, t h a t is, [ / c l is open if and only if ^~1{U) is open in G. A m a p c(x - > c(x)) of X into 6? is called a ItoreZ (continuous) section if it is a Borel (continuous) m a p and if ft(c(x)) — x for all x e X; t h e n its range meets each left G0 coset exactly once. I t is said to be regular if for each compact subset K of G, c[X] n /? _1 (j8[ir|) is a set with compact closure. Theorem 5.11. X, under the quotient topology, is a locally compact Hausdorff space satisfying the second axiom of countability and ft is an open continuous map. Moreover, X is a G-space and the map g,x->g-x of G x X into X is continuous. If K is any compact subset of X, there exists a compact subset Kx of G such that ft[K1] = K. Further, there exists a regular Borel section for G/G0. A map f of X into some Borel space Z is Borel if and only if the map f°=f o ft of G to Z is Borel. In particular, if E^X, E is a Borel set if and only if the set ft~1(E) is a Borel subset of G. Finally, let Y be a transitive standard Borel G-space, y e Y and let G0 be the stability subgroup at y. Then the map t : ft(g) —> g-y is a Borel isomorphism of X = GjGQ onto Y which is a G-isomorphism of the associated G-spaces. If Y is a locally compact G-space satisfying the second axiom of countability, t is a homeomorphism. Proof. That X is locally compact Hausdorff second countable and t h a t ft is open are well known (cf. Pontrjagin [1]). Now, G is clearly transitive on X. The m a p (g,gr) - > {g,fi(g')) of G x G onto G x X is continuous and it is easily shown t h a t the topology for GxX is the quotient topology relative to this m a p . The m a p g, x - > g-x is therefore continuous.
MEASURE
THEORY
ON G-SPACES
165
Let Wlt W2i- - • be a sequence of compact sets in G with nonempty interiors such t h a t ( J n (inter ior (W n)) = G. If then K is a compact set c j , then for some n,
K £ U /TO, and hence # = £ [ 7 ^ ] , where K^p-^K) n (\JJ=1 Wj). Clearly Z x is compact. We shall next prove the existence of a regular Borel section for G/G0. Let {Kn} be a sequence of compact subsets of G such t h a t (a) K1^K2^ •• • and U n ^ « = ft and (b) any compact subset of G is contained in some Kn. Using lemma 5.1 we can construct a Borel set E such t h a t (i) i£ n /S _1 (j8[iL n ])cir n for all n, and (ii) ft is one-one on i£ and maps 2£ onto X . ft is thus a Borel isomorphism of i£ onto X as E and X are both standard. If c denotes the inverse map, c is a Borel isomorphism of X onto E. c is a Borel section. I t is obvious from (i) and (b) above t h a t c is regular. Suppose now the set A c X is such t h a t 0~ 1(A) is a Borel set in 6?. Then E n P~1(A) is also Borel. Since j8 is one-one on E, and since the Borel subspace defined by E is standard, A = fi[E n j8 _1 (J[)] is a Borel subset of X. If / is a Borel m a p of X into Z, where Z is some Borel space, then / ° = / © ($ is a Borel m a p of G into Z. Conversely, let / ° be a Borel m a p of G into Z such t h a t f° =f o p for some m a p / of X into Z. If F is any Borel subset of Z, p-1{f-1{F))=f°-1{F) is Borel in 0 and hence / ^ ( - P ) is Borel in X . / is thus a Borel map. We finally come to the uniqueness of transitive G-spaces. Let 7 , y be as in the statement of the last assertion of theorem 5.11. Clearly the map t:P(g)^9'V is one-one, maps X onto Y, and is moreover such t h a t t(g • x) = g • t(x) for all xeX and g e G. If M^Y is open, t~1{M)=p[A], where A = 1 {9 '-g-ytM} so t h a t t' (M) is open, t is thus continuous. A classical category argument (Pontrjagin [1]) now implies t h a t t is a homeomorphism. Suppose we assume in the above proof t h a t Y is only a standard Borel G-space. We then use corollary 5.8 to conclude t h a t Gy is a closed subgroup of G. Thus t is well defined and the same argument as t h a t given above now shows t h a t t is Borel. Since both X and Y are standard, t is a Borel isomorphism. This completes the proof of the theorem. Since G acts on X , we m a y speak of invariant measure classes on X . We shall now proceed to study these. We use a "lifting" technique to reduce the problem to G itself. Now G acts on itself by left and right translations. We have accordingly left and right invariant measure classes. We write /x for a left H a a r measure and define fxr by (26) fjir is a right H a a r measure.
^(E)
= /x^-1).
166
GEOMETRY
OF QUANTUM
THEORY
Lemma 5.12. There exists a unique measure class on G which is left (or right) invariant, and it is the measure class determined by fx. In particular, it is both left and right invariant. Proof. Let a be a finite right quasi-invariant measure. We want to prove t h a t a(E) = 0 if and only if /JL(E) = 0. Let / be a Borel function on G such t h a t (i) 0 < f(x) < 1 for all xeG, (ii)
f(x)dfi(x)+ JG
f(x)dfir(x)
< oo.
JG
Such / are easily constructed. Suppose now for a Borel set A, a(A) = 0. Since a is right quasi invariant, a(Ax) = 0 for all x in G and hence a(Ax~1)f(x)dfji(x)
= 0,
/JG. i.e., writing XA f ° r the characteristic function of A, L ( 1 XA(yx)da(y)^f{x)d^x)
= 0.
By Fubini's theorem, this means
J (J XA(yx)fWdA*))My) = o and hence, as the inner integral is > 0 , XA(yx)f(x)Mx)
= o
JG
for a-almost all y. Thus, for some y0 e G, JG
XA(y
= 0
a n d hence, as f(x) > 0 for all x, XA(VOX) =
0
1
for /x-almost all a;, i.e., fM(y0' A) = 0 so t h a t /x(^4) = 0. Conversely, if fi(A) = 0, /x( y ~ XA) = 0 for each y and hence
J XA(yx)f{xW(x) = 0 for each i/, proving t h a t
Using Fubini's theorem, we get:
J.
a(Ax-1)f(x)dfx(x)
= 0
ME AS U BE THEORY
ON
G-SPACES
167
and hence, as the integrand is > 0 , a{AxQ-x)
= 0
for some x0, showing, as a is right quasi invariant, t h a t <x(A) = 0. Thus /xr, a, ft are all in the same measure class. If a is left quasi invariant, the measure a defined by a(A) = a(A~1) is right quasi invariant so t h a t /x, a, jjir are in the same measure class. This proves the lemma. Remark. Lemma 5.12 enables us to speak without ambiguity of null sets in G; a Borel set A c= 0 will be called null if it is a set of measure zero for the unique measure class singled out by lemma 5.12. We now return to the general transitive case. Let G0 be a closed subgroup of G and X = G/G0. We now observe t h a t , on X, quasi-invariant measures can be constructed easily. Let aQ be a finite measure on G which is quasi invariant, and let a be the measure on X defined by a(A) = aoip-^A)). Then a(E) = 0 if and only if p~1{E) is a null set in G. a is thus quasi invariant. Our aim is to show t h a t all quasi-invariant measures can be obtained in this manner, up to absolute continuity. Let CC(G) denote the class of all complex functions on G which are continuous and have compact support. Let CC(X) denote the analogously defined space of functions on X. For any / in CC(G) let us write Mf for the function on G defined by (27)
(Mf)(g)
= f
f(gh)dfi0(h);
J G0
here /x0 is a left H a a r measure on the lcsc group G0 chosen in some way. I t is then clear from the left invariance of/x 0 t h a t , for any h0 in G0, (28)
(Mf)(gh0)
=
(Mf)(g)
for all g e G. Hence there exists a function / ~ on X such t h a t (29)
f~W(9))
= (Mf)(g)
(geO).
We shall use the mapping / -> / ~ to lift measures from X to G. Lemma 5.13. Let P be the set of all real nonnegative elements of CC(X) and P' a subset of P having the following properties: (i) if fl9f2 e P' and c1? c2 are constants > 0 , c 1 / 1 + c 2 / 2 e P', (h) if f' £ P' and fx e P, fife P', and (iii) for any x e X, there is an f e P' with f(x)>0. Then P' = P . Proof. Let xeX. Then there exists an fxeP' such t h a t fx(x) = \. Therefore, for some open set Ux with compact closure and containing x, fx(y) ^ i f ° r a U V E Ux. Let fx be an element of P such t h a t f^y) =fx(y)~x E an< for all y e Ux. Then fx'=fifx P' ^ has the property t h a t fx'(y) — ^ for all y e Ux. Suppose now t h a t K is any compact subset of X. For each
168
GEOMETRY
OF QUANTUM
THEORY
x e K, let Ux be an open set containing x and fx an element of P ' such that fx = 1 on Ux. Choose points xl9 • • •, xk e K such that K^[Jf=1 UXj. Thenf'=f'Xl+• • +f'Xk belongs to P ' and f'{y)>\ for all y e K. Choose an f2 e P such that f2(y)=f'(y)~1 f° r a u y E %• Then f"=f2f belongs to P' and f"{y) = 1 for all y e K. Finally, let / e P and let K be a compact set such that f(y) = 0 for y $ K. Choose a n / " e P' such that f"(y) = 1 for all y e K. Then / = / / " e P'. This proves the lemma. We are now ready to formulate the main properties of the map / - * / - •
Lemma 5.14. For any feCc(G), f~ eCc(X). The map f->f~ has the following properties: (a) / > 0 implies f~ > 0 ; (b) / - » / ~ maps CC(G) onto CC(X) and maps the set of real nonnegative elements of CC(G) onto the set of real nonnegative elements of CC(X); (c) for any continuous function u on X, {u°f)~=uf~ (where u° = u op) (d)ifaeG, (fa)~ = (f~)a. Proof. We first prove that for a n y / in CC(G), / ~ eCc(X). Clearly if/ vanishes outside a compact set K^G,f~ vanishes outside f$[K]. Therefore there remains only the continuity of/". For this it suffices to prove that Mf is continuous on G since then the continuit}' of/" would follow from the fact that X has the quotient topology. Let g0 e G and W0 be a compact neighborhood of e. Now
\(Mf)(g)-(Mf)(g0)\ < f
\f(gh)-f(g0h)\d^0(h).
J G0
Since feCc(G), given e>0, there exists a compact neighborhood W^ W o around e such that \f(a)— f(a')\ <e whenever aa'~1eW. If/ vanishes outside the compact K, and gg0~x e W,
\{Mf){g)-{Mf)(g0)\ <
J
|/(^)-/(^)|^oW
(fifO -1 W-lK)nG 0
< e
dfM0(h), K'nG0
where K' = (g0 ~1W0 " XK) n G0. This proves that Mf is continuous at gr05 finishing the proof t h a t / " e CC(X). The map / - > / ~ is linear and it is obvious that / > 0 =>/~ >0. Next, let 16 be a continuous function on X and u° = u o p. Then
Jf(«°/)(0) = «•>((/) f f(gh)dp0(h). J G0
This shows that M(u°f) = u°Mf. Hence (u°f)~=uf~. Third, let aeff. Then (Mfa){g) = (Mf)(a~1g) from which we conclude that (fa)~=(f~)a. This proves that (a), (c), -and (d) are satisfied. It only remains to prove (b).
MEASURE
THEORY
169
ON G-SPACES
Let P' ={f~
:feCc{G),
/ r e a l and > 0 } ,
and let P be the set of all real nonnegative elements of CC(X). We claim t h a t P' satisfies the conditions of the lemma 5.13. (i) is obvious and (ii) follows from (c). To check t h a t (iii) is satisfied at each point of X, it is enough to do so at x0 = /3(e) because G acts transitively on X and P' satisfies (d). But then, if feCc(G)>0 and is 1 in a neighborhood of e, (Mf)(e) =f~(x0) > 0 obviously. This shows now t h a t P' = P . I n particular, CC(G) is mapped onto CC(X). The lemma is completely proved. For any Borel measure a on X, the map / - > [ f~da
is, by virtue of
lemma 5.14, a nonnegative linear functional on CC(G). Let us denote by a0 the Borel measure defined on G by this linear functional. We thus have a mapping a -> a0 from the set of Borel measures on X into the set of Borel measures on G. Since / - > / ~ has the whole of CC(X) for its range, it is obvious t h a t ax° = a2° implies a± = a 2 . a -> a0 is therefore a one-one map. Let us now consider the equation (30)
f /d«o
f f~da
=
JG
(feCc(G)).
JX
Our next aim is to extend (30) to certain Borel functions on G. Lemma 5.15, Let f be a Borel function on G such that
(31)
f
1/(^)1^0(4)
<
00
J Go G 0 '" ^
for each g e G. Then there exists a unique Borel function f~ on X such that for all g e G, f~(P(9))
= f
f(gh)dno(h)-
G00 JJ G
If f> 0, then f ~ > 0 <md we have (32)
f /rfa° == [ JG
f~da
Jx
in the sense that either both sides are infinite or both sides are finite and equal. Iff is bounded and vanishes outside a compact subset of G, thenf~ is bounded and vanishes outside a compact subset of X. Proof. Since g,h^f(gh)
is Borel on GxG0,
it is clear t h a t g ->
X
\G / ( ^ ) ^ / O ( ^ ) is Borel on G. Moreover, this function of g is constant on the left 6r 0 -cosets. Hence, by theorem 5.11, the existence and uniqueness of / ~ are immediate. A l s o , / > 0 i m p l i e s / ^ > 0 . Suppose /(g) = 0 for all g $ K,
170
GEOMETRY
OF QUANTUM
THEORY
K being a compact set, and let \f(g)\
= I f I < A
f(9h)di*o(h)
JG0
\ Jgg
<
dfji0(h)
1x
KnG KnG0
ApoiK-WnGo).
This proves t h a t / ~ is bounded. We now come to the proof of (32). We consider first the case when f=XK> the characteristic function of the compact set K^G. We can choose a sequence {/n} of nonnegative elements in CC(G) such t h a t fn j XK point wise on G. Then / n ~ \ f~ point wise on X and
f fda° = lim f fnda° JG
n^co JG
= lim
fn~da
n-*co Jx
= f da
L~ '
This proves (32) in this case. Next, for any Borel set A^K, J=XA is bounded and has compact support and hence %A~ *S bounded and has compact support. I t is therefore a-integrable and the m a p A - > j XA~dcc is easily verified to be a measure on the Borel sets of K. By what we have seen earlier, this measure coincides with a 0 on all compact A^K. Therefore it must coincide with a0 for all Borel sets A. But then this proves (32) for f=XA (A^K). At this stage we know t h a t (32) is true for all simple functions / vanishing outside compact sets. For the general case, l e t / > 0 and let {/n} be a sequence of nonnegative simple functions with compact supports such t h a t fn f / pointwise on G. T h e n / n ~ f / ~ pointwise on X. Hence by the Fatou-Lebesgue lemma, f fda* = lim f
JG
n->oo j
fnda°
G
= lim
fn~da
n->oo J x
L
f~d, a.
This completes the proof of the lemma. Corollary 5.16. For a ° ( / S - V ) ) = 0.
a Borel
set A^X,
a(^4) = 0, if
and
only
if
MEASURE
THEORY
171
ON G-SPACES
Proof. Suppose t h a t A is a Borel set c j and a(A) = 0. If K is a compact set ^P~1(A), then, XK~ = 0 outside ^4 and so, by (32),
«°(#) = j XK~d* = 0. 0
This proves, by regularity of a , t h a t a°(p~1(A)) = 0. Conversely, let a0(j3~1(A)) = 0. Let K^A be compact. Since ^~1(K) is closed, there exists an increasing sequence {Kn} of compact subsets of ^~1(K) whose union is ^~1(K). Then a°(Kn) = 0 for all n, so t h a t we can infer from (32) and the nonnegativity of XKU ~ t h a t X#n ~ = 0 almost everywhere for each n. Suppose now t h a t a(K) > 0. Then, there will exist an x1 e K such t h a t XKn~(x1) = 0 for all n. Choose a g± e ^~1(K) such t h a t fi(gi)=x1. Then /xo((^ri ~ 1Kn) n #o) = = 0 f ° r a U ^. This is absurd as (gx ~ 1Kn) n 6r0 f G0. Thus a(K) must vanish. By regularity of a, a(A) = 0. Corollary 5.17. / / a± and a2 are Borel measures on X, <x2<£cc1 if and only if a 2 ° « a 1 0 . In this case, (33)
dcc^/da^
Proof. Let a 2 « a i nonnegative ueCc(G), 8.15 and, moreover,
an
(da^dcc^0.
=
d ^ t = da2/da1. We m a y assume £ > 0 . For any the function ut° satisfies the conditions of lemma
(ut°)~ 1
1
=u~t.
0
Since u~t e J * ? ^ ) , ut° e J ^ ^ ) and we have: uPda^
=
JG
Jx
=
u~tda1 u~da2
=
uda2°. JG
This shows at once t h a t a 2 ° « a i ° a n d dcc2°lda1° = t0. On the other hand, if a 2 ° « a 1 ° , we obtain the relation a 2 « a 1 at once from the previous corollary. Corollary 5.18. / / G is compact, then for any Borel set A^X, cfi(p-*(A)).
a(A) =
Proof. I n this case, on normalizing fx and /x0 as usual, we see t h a t for any ueCc(X), corollary.
{uQ)~ =u.
Hence,
\ uda=\
u°da°.
This implies
the
172
GEOMETRY
OF QUANTUM
THEORY
From lemmas 5.14 and 5.15 the following theorem follows immediately. I t is a basic theorem in the theory of quasi-invariant measures on X. Theorem 5.19. There are quasi-invariant Borel measures on X = GjG0, where G is a Icsc group and G0 a closed subgroup. If A is a finite quasiinvariant measure on G, and A~ is the finite measure defined on X by the equation A~(^4) = X(^~1(A)), then A~ is quasi invariant. Any two quasiinvariant a-finite measures on X are mutually absolutely continuous. If E^X is any Borel set, the quasi-invariant a-finite measures on X vanish for E if and only if p~1(E) is a null set of G. If a is any Borel measure on X, a is quasi invariant (invariant) if and only if a0 is a quasi-invariant (left invariant) measure on G. Proof. If g E G and if a is a Borel measure on X , (a9)0 = (a0)9 as is easily seen from (d) of lemma 5.14. Moreover, any cr-finite measure is mutually absolutely continuous with respect to some finite measure. Lemmas 5.14 and 5.15 and corollaries 5.16 and 5.17 then lead quickly to the proofs of all the assertions of this theorem. Theorem 5.19 enables one to speak of null sets of X and G without specifying any quasi-invariant measure. If A is any quasi-invariant measure on X, the class of A-measurable sets is independent of A. I n other words, we m a y also speak of measurable sets and measurable functions on X without ambiguity. Corollary 5.20. Let A^X. Then A is measurable if and only if p~1(A) is measurable. If f is a map of X into some Borel space Z, f is measurable if and only iff0 =f o £ is a measurable map of G into Z. Proof. I t is enough to prove the first p a r t since the second p a r t follows immediately from the first. Suppose now t h a t A^X is measurable. Then there are Borel sets A1 and A2 such t h a t A±^A^A2 and A2 — A1 is a null set. Clearly p-1(A1)^p-1(A)^^1(A2) and p~1(A2)-p~1(A1) = ft~1(A2 — A1) is also null, ft'1 (A) is thus measurable. Conversely, let B = j3~1(A) be measurable. Let A be a finite quasiinvariant measure on G and let A~ be the induced measure on X. There exist sequences {Kn} and {Kn'} of compact subsets of G such t h a t (i) Kn^B and Kn'^G-B for all n, and (ii) X(B-\JnKn) = X{(0-B)U n Kn') = 0. Since j8 is continuous, the sets A1 and A2, where
A = U KKn] n n
are both Borel. I t is obvious t h a t AX^A^A2 A is t h u s measurable.
and A~(^42 — ^41) = 0.
MEASURE
THEORY
ON G-SPACES
173
Corollary 5.21. G acts ergodically on X, i.e., if f is a Borel map of X into some Borel space Z such that f(g-x)=f{x) for almost all (g,x) e GxX, then f is a constant almost everywhere. In particular, there exists, up to a constant multiplying factor, at most one invariant a-finite measure on X. Proof. By the Fubini theorem, there exists a point x± of X such that f(9'xi)=f(xi) f° r almost all g. We may clearly assume that x1 = x0 = f$(e). Select a Borel set A^G such that A is the union of countably many compact sets, G — A is of Haar measure zero, and f(g-x0)=f(x0) for geA. Let X0 = A-x0. Then X0 is a Borel set and ^~1(X — X0)^G-A is null, showing that X — X0 is null. Let a± and <x2 be nonzero invariant afinite measures. Then they are quasi invariant. Let f=da1jda2. Then, for each g e G, f(g-x)=f(x) for almost all x. f is thus a constant almost everywhere, proving that a± is a constant multiple of a2. Corollary 5.22. Let 8 be a standard Borel space and let f be a Borel map of G into 8. Suppose that for each h e G0 f(gh) =f(g) for almost all g. Then there exists a Borel map a of X into 8 such that f(g) = a{fi{g)) for almost all g. Proof. We may clearly assume that 8 is the unit interval. Let A be a quasi-invariant measure on G with A(6r) = l. Let A0 be a quasi-invariant measure on G0 with A0(6r0) = 1. Let A~ be the measure on X induced by A through the map /?. By the Fubini theorem, f(gh)=f(g) for almost all (g,h) GGXG0. Hence, for some Borel set N of G of measure zero, we can assert that for each g e G — N, f(gh) =f(g) for almost all h e G0. If we write fi(9) = f
f(9h)dXo(h)9
J G0
t h e n / x is Borel, and for geG — N, f1(g)=f{g)> Moreover, for any fixed g e G — N and h' e G0, the functions h ->f(gh'h) and h ^f(gh) are both equal to f(g) for almost all h. Consequently, f1{gh')=fx(g) for all {g,h')e{G-N)xGQ. Consider now the set X'=p[Q-N]. X' is a A~ measurable subset of X by corollary 5.2, and hence ^(X') is a A-measurable set which is a union of left 6r0-cosets, and, for all (g,h) e p~1(Xf) x G0, f1(gh)=f1(g). Moreover, X — X' is null and so G — ^~1(X') is also null. Define now the function f2 on G by
{M)
M9) =
\o
to^-W).
f2 is A-measurable and f2(g)=f(g) almost everywhere. Further, f2 is constant on the left 6r0-cosets, and therefore there exists, by corollary 5.20, a A~-measurable function/ 3 on X such t h a t / 2 = / 3 0 . Choose a Borel function a and X such that a(x)=f3(x) for almost all x. Then f(g) = a(f$(g)) for almost all g. This completes the proof.
174
GEOMETRY
OF QUANTUM
THEORY
5. COCYCLES A N D COHOMOLOGY The functional equations (22) are extremely significant in the theory of induced representations associated with 6r-spaces. We shall devote t h e present section to a study of these equations in an appropriately general context. Let X be a standard Borel 6r-space, G being a fixed lcsc group. We choose an invariant measure class ^ o n l . Let M be a standard Borel group, fixed throughout this section. We write 1 for the identity of M. Let / be a function on GxX with values in M. We shall say t h a t / is a (G,X,M)cocycle relative to ^ if the following properties are satisfied: (i) (35)
/ i s a Borel m a p of GxX
(ii)
f(e,x)
(iii)
= 1
for almost all
f(gi92>z) = f(gi,92'X)f{92>z)
into M, xel, for
almost all (9n92>x) E
GxGxX.
If there is no ambiguity concerning the group M, we shall refer to / simply as a (G,X)-cocycle relative to *$ or, even more briefly, as a cocycle relative to <€. Note t h a t the null set in (ii) is with respect to ^ while t h e null sets in (iii) are those of the measure class which is a product of t h e usual measure classes in G with %\ A c o c y c l e / is said to be a strict cocycle if it satisfies (35)(i) and further, (36) (ii') and (iii'), where (36)
(ii') (m')
f(e,x) = 1
J
for all
x e X,
f(9i92>x) = f(9i,92 • x)f(92,x)
for all
(91,92^)
eGxGxX.
Our decision to call the functions satisfying (35) and (36) cocycles is prompted by the fact t h a t these equations are generalizations of t h e identities which describe the cocycles in the cohomology theory of groups (cf. Eilenberg [1], pp. 3-37). We note here t h a t it was G. W. Mackey who first studied the cocycles in the context of an arbitrary transitive Borel G-space. I t was the detailed study and analysis of cocycles which enabled Mackey to formulate and prove the correct generalizations of the classical work of Frobenius on induced representations (of finite groups), to the case when the group in question was an arbitrary lcsc group. The aim of this section is to carry out a cohomology theory of (G,X,M)-cocycles. We begin by introducing equivalence relations in the set of all cocycles relative to a fixed invariant measure class # . Suppose t h a t f± and f2 are two cocycles relative to <€. We shall say t h a t fx and f2 are cohomologous, fi~f2 in symbols, if there exists a Borel m a p b(x -> b(x)) of X into M such t h a t
(37)
f2(g,x)
= %-a;)/i((7,a;)6(a;)
x
MEASURE
THEORY ON G-SPACES
175
for almost all (g,x) e GxX. B y taking b to be identically 1, we see t h a t if / i a n d / 2 are equal almost everywhere on Gx X, then fx~f2- I t is easy to verify t h a t ~ is a genuine equivalence relation in the set of all cocycles relative to %\ Each ~ equivalence class will be called a cohomology class relative to %>\ Suppose further t h a t f± and f2 are strict cocycles. We shall then say t h a t / x a n d / 2 are strictly cohomologous, fx xf2 in symbols, if there exists a Borel m a p b(x -> b(x)) of X into M such t h a t
for all (g,x) e GxX. # is also easily verified to be an equivalence relation in the set of all strict cocycles. Each #-equivalence class will be called a strict cohomology class. The m a p 1 : g,x -> 1 is a strict cocycle. Any cocycle cohomologous to it will be called a coboundary. If the cocycle is strictly cohomologous to 1, it will be called a strict coboundary. Obviously, a cocycle / is a coboundary (strict resp.) if and only if for some Borel function b on X with values in M, f(g,x) = b(g-x)b(x)~1 for almost all (all resp.) (g,x) eGxX. Suppose t h a t M is abelian. Then the set of Borel maps o f 6 r x I into M is an abelian group under pointwise multiplication; both the set of cocycles and the set of strict cocycles are subgroups of this group. The coboundaries form a subgroup of the group of cocycles and for two cocycles fi,f2,fi~f2 if and only if f±f2 ~x is a coboundary. Therefore the cohomology classes are in one-one correspondence with the elements of the quotient group of the group of cocycles mod the group of coboundaries. This quotient group is known as the (G,X,M)-cohomology group. From now on we shall assume that G is transitive on X. We choose a point x0 6 X arbitrarily and write G0 for the stability subgroup a t x0. G0 is closed, ft denotes the m a p g -> g • x0 of G onto X. /3 is a Borel map. For any m a p u of X into some space Z, u° denotes the m a p g - > u(P(g)) of G into Z. On the other hand, if v is a m a p of G into Z which is constant on the left 6r 0 -cosets, there exists a unique map u of X into Z such t h a t u° = v; we denote u by v~. Thus (v~)° = v and (u°)~ =u. We shall also extend this notation to two variables. Let Y, Z be Borel spaces and let u be a m a p of 7 x 1 into Z. We then write u° for the map y,9-+u(y£(g)) of Y x G into Z. We note t h a t u is Borel if and only if u° is Borel. At various points of our arguments in this section, we shall be concerned with functions on X, G, GxG, GxX, GxGxX, and so on. Since G and X possess unique invariant measure classes, there are canonically determined measure classes on each of the other spaces. For instance, if A and X are quasi-invariant a-finite measures on G and X, respectively, the measure Ax A' determines a measure class on GxX which depends only
176
GEOMETRY
OF QUANTUM
THEORY
on the measure classes of A and A'. By a (Borel) null subset of G x X we mean a set on which all the measures of this class vanish. Similar remarks apply to GxGxX, etc. In view of the uniqueness of the invariant measure classes, we are entitled to omit any reference to them and speak of cocycles and coboundaries. This we shall do. Suppose now t h a t / is a strict cocycle. Then it follows easily from (ii') and (iii') of (36) that h -^f(h,x) is a Borel homomorphism of the stability group Gx at x into M. We shall call it the homomorphism defined by fat the point x e X. If mx and m2 are two Borel homomorphisms of G0 into M, we shall say that mx and m2 are equivalent if there exists a k e M such that, for all heG0, (38)
m2(h) =
km^k-1.
The cocycles and their cohomology classes are difficult to handle because of the null sets involved. The strict cocycles and their classes are much better behaved in this respect since there are no null sets. Our main concern is to show that in every cohomology class of cocycles there is a strict cocycle and that the latter is determined uniquely up to strict cohomology. This means that the cohomology classes of cocycles are in canonical one-one correspondence with the strict cohomology classes of strict cocycles, which are easier to analyze. The simplest strict cocycles are the strict (G,G)-cocycles. Our method of examining strict (6r,X)-cocycles consists in "lifting" these to strict (6r,6r)-cocycles and analyzing the latter. Lemma 5.23. Let f be a strict (G,X)-cocycle. Then the function f° on GxG defined by
/ W ) = f(gMg')) is a strict (G,G)-cocycle. If F is any strict (G,G)-cocycle, there exists a unique Borel map b of G into M such that b(e) = l and (39)
F(g,g') = b(gg')b(gf)^
for all (g,gf) eGxG. Proof. The first assertion is trivial. For the second, we note that (36) gives F(g,g') = F(gg\e)F(g',e)-i for all (g,gf)eGxG. Equation (39) follows by putting b(g') = F(gf,e). If b± is another Borel map of G into M such that b1(e) = l and F(g,g') = biigg'Mg')'1, then HggT^igg'^Hg'r^g') for all (g9gf) EGXG. This 1 shows that for some keM, b(g)~ b1(g) = k for all g. Since b1(e) = b(e) = l, & = 1 and b = bx.
MEASURE
THEORY ON G-SPACES
177
In view of this lemma, with any strict (G,X) cocycle/, we can associate a Borel function on G. But f° is not the most general {G,G)~cocycle since it arises from a (G,X)-cocycle. This Borel function must therefore satisfy some special identities. The next lemma describes these identities. Lemma 5.24. Let m be a Borel homomorphism of G0 into M. Then there exists a Borel map b of G into M such that b(e) = 1 and (40)
b(gh) = b(g)m(h)
for all (g,h) eGxG0. Corresponding to any such map b, there is a unique strict (G,X)-cocycle f such that
f°(g,g') = Hgg'MgT1
(41)
for all (g,gf) e GxG; f defines m at x0. Conversely, if f is a strict (G,X)cocycle and b is the Borel map of G into M such that b(e) = 1 and b satisfies (41), then the restriction ofb to G0 coincides with the homomorphism m defined by f at x0 and b satisfies (40). Proof. Let m(h -> m(h)) be a Borel homomorphism of G0 into M. Let c(x -> c(x)) be a Borel section for G/G0. We may assume c so chosen that c(x0) = e. For any g e G, let a(g) = c(P(g))~1g. Since g and c(f$(g)) lie on the same left coset gGQ, a(g) e G0. We put (42)
b{g) = m(a(g)).
Since c and fi are Borel maps, g —> a(g) is a Borel map of G into G0 and hence g —> b(g) is a Borel map of G into M. Also b(e) = m(a(e)) = l. Moreover, for g e G and h e G0, a(gh) = a(g)h. This shows that b(gh) = b(g)m(h) for all (g,h) e GxG0. Let
* W ) = Hgg'W)-1. F is obviously a strict (G,G)-cocycle and (40) shows that F(g,g'h) = F(g,g') for all (g,gf,h) e GxGxG0. Hence, there exists a unique Borel map / of G x X into M such that
MP(g')) = * W ) for all (#, g') e Gx G. It is easily checked t h a t / i s a strict (6r,X)-cocycle, that f° = F, and that it is the only (6r,X)-cocycle with this property. Finally, for h e G0, f(h,x0) = F(h,e) = m(h). For the results in the converse direction, let / be a strict (G,X)-cocycle and let b be such that b(e) = 1 and (43) for all {g,g') eGxG.
f°(g,g') =
Hgg'MgT1
Then, for all (g,g\h) b(gg'h)b(g'h)-i =
eGxGxG0,
Hgg'MgV1,
178
GEOMETRY
OF QUANTUM
THEORY
from which it follows easily that for each h e G0, g' -» 6(g,)"16(gr/^) is a constant, say m(h). Thus b(g'h) = b(g')m(h) for all (g,h) e GxG0. Since b(e) = l, b(h) = m(h), while (43) implies that b{h)=f(h,x0), for all h e G0. This proves everything. Lemma 5.25. Let ft be a strict (G,X)-cocycle and let mi be the homomorphism defined by ft at x0 (i = 1, 2). Then the following statements are equivalent: (a)
(44)
mY and w,2 are equivalent,
(b)
f^f2,
(c)
/i~/2-
Proof, (b) => (c) is trivial. We shall prove that (a) => (b) and (c) => (a). (a) => (b). Suppose that m1 and m2 are equivalent homomorphisms. Now there exists by lemma 5.24, a Boral map bt of G into M such that (i) mt is the restriction of bt to G0, (ii) b{{gh) — b^m^h) for all (g,h) e GxG0 and (iii) fi0{g,g') = bi(gg')bi(g')-1 for all (g,g')eGxG. Let k e M be such that (45)
km^fyk-1
m2(h) =
for all h eG0. From (45) we obtain the equation b2(gh)kb1(gh)-' =
bMkbM-1
for all (g,h) e GxG0. This shows that the Borel map g -> b2(g)kb1(g)~1 is constant on the left 6r0-cosets. Therefore there exists a Borel map b(x —> &(#)) of X into Jtf with the property that (46)
ba{g)^i(9)-1
= W ) ) = b°(g)
for all g e G. An easy computation using (46) and the relation of bt to / f now gives us the relation f2°(9,9') = from which we obtain
b°(gg')UO(g,g'mg')-\
f2(g,x) = b{g^x)f1{g,x)b{x)-1 for all (g,x) eGxX. (c) => (a). We assume / i ^ / 2 - Therefore there exists a Borel map b of X into M such that
f2(g,x) = Hg-rfMg^Mx)-1 for almost all (<7,#) eGxX. (47)
/2 W )
This implies that =
bOigg'Wfag'mg')"1
for almost all (g,gf) EGXG. Let fy be the Borel map of Ginto M such that (i) mt is the restriction of b{ to # 0 , (ii) fi°(g,g,) = bi(ggf)bi(g,)~1 for all
MEASURE
THEORY
ON G-SPACES
179
(g,g')eGxG (lemma 5.24). We then obtain from (47), after an easy calculation, the equation (48)
^(gg'yWiggY'Ugg') = M ^ r w r 1 ^ ' )
for almost all (g,g')eGxG. This means (corollary 5.21 with X = G) that the function g ->b1(g)~1b°(g)~1b2(g) is almost everywhere equal to a constant, say k'1 e M. Thus (49)
b2(g) =
b^b^k-1
for almost all g e G. We now conclude from the connection between b{ and m{ that for each h e G0,
for almost all g. Thus (50)
m2(h) =
km^fyk-1.
This proves that (c) => (a) and completes the proof. Lemma 5.26. / / / is any (G,X)-cocycle, there exists a strict (G,X)-cocycle j 1 such that f(9>x) = fi(9>v) for almost all (g,x) s G x X; f1 is determined uniquely up to strict cohomology. Proof. L e t / b e a (6r,X)-cocycle and l e t / 0 be defined as usual, i.e., /W )
= f{0,P(9'))-
Then f°(gi92,93) = Z ^ i ^ s ) / 0 ^ , ^ ) for almost all (g1,g2>g3) eGxGxG.
Hence there exists a g0 e G such that
f°(gig2,go) = P(9i>9&o)fQ(92>9o) for almost all (g1,g2) £ GxG. If we put g=gi, g' = 9r20ro> w e that
Pigg'go-^goWg*-1^)-1
f°(9,g') = for almost all (g,gf) eGxG.
Let d0(9) =f°(99o~1>9o)>
Then d0 is a Borel map of G into M and
/ W ) = doigg'WoigV1
can
conclude
180
GEOMETRY
OF QUANTUM
THEORY
f
for almost all (g,g ) e GxG. Changing d0(g) to d(g) = d0(g)d0(e) ~1, we secure d(e) = 1 and at the same time
(51)
/W)
= digg'W)-1
for almost all (g,g') e GxG. Fix now an arbitrary h e G0. Then it follows from (51) that d(gg'h)d(g'h)-i = d(gg')d(g')^ for almost all (g,gf) e GxG, i.e., d{gg')-1d{gg'h) = d(g')'H(g'h) for almost all (g,gf)eGxG. Corollary 5.21 (with X = G) implies now that the function g' -> d{g')~1d(g'h) is almost everywhere equal to a constant, say m(h). In other words, (52)
d(gh) = d(g)m(h)
for almost all g. Clearly m(h) is uniquely determined by the requirement that the condition (52) be satisfied for almost all g. It now follows from (52) and the uniqueness of m(h) that (i) m(e) = 1, and (ii) for hlf h2 e G0, m{h1h2) = rn(h1)m{li2). We assert that mis a Borel map of G into M. Since M is standard, we may assume, in this argument, that M is the unit interval. Let
As cp is Borel on G x G0, this shows that m is Borel on G0. In other words, m is a Borel homomorphism of G0 into M. By lemma 5.24 there exists a Borel map d' of G into M such that d'(e) = l and d'(gh) = d'(g)m(h) for all (g,h) eGxG0. From (52) it now follows that for each h e GQ, d(g)d'(g)-1
=
d(gh)d'(gh)-i
for almost all 0. By corollary 5.22 we can find a Borel map k of X into ilf such that d(g) = k(p(g))d'(g) for almost all g. Write d1(g) = k(^(g))d'(g). Then d1(gh) = d1(g)m(h) for aZZ (g,h) e GxG0 and d1(g) — d(g) for almost all 0. By lemma 5.24, there exists a unique strict (6r,Z)-cocycle / x such that
A W ) = dAggld^g')-1 for all (g,gf) e GxG. Since dx~d almost everywhere, the last equation implies that f° =fx° almost everywhere on GxG. Thus (53) for almost all (<7,#)
fi(g,*) = fig,*) EGXX.
MEASURE
THEORY
ON G-S PAGES
181
If f2 were another strict cocycle such t h a t (53) is satisfied with / 2 in place o f / j , then f±=f2 almost everywhere on Gx X. In particular, / i ~ / 2 . From lemma 5.25 we now conclude t h a t / ^ / g . This completes the proof. Theorem 5.27 (Mackey [2], [6]). Let G be a Icsc group, X a standard transitive Borel G-space, and M a standard Borel group, x0, an arbitrary point of X, and G0, the stability group at x0. If y is any cohomology class of (G,X,M)-cocycles, the set yst of all strict (G,X,M)-cocycles in y is nonempty and is a strict cohomology class. The set y~ of the homomorphisms of G0 defined by the elements of yst at x0 is an equivalence class of Borel homomorphisms of G0 into M. y - > y~ is a one-one correspondence of the set of all cohomology classes onto the set of all equivalence classes of Borel homomorphisms of G0 into M. If M is abelian, the (G,X,M)-cohomology group is isomorphic to the group of all Borel homomorphisms of G0 into M. Finally, if X is an affine G-space (i.e., G0 is trivial), all cocycles are coboundaries. Proof. The proofs of all the statements are contained in lemmas 5.23 through 5.27. Remark. The theorem just proved is the cornerstone of the theory of induced representations. I t reduces the problem of construction and classification of cocycles to the problem of constructing the equivalence classes of Borel homomorphisms of G0 into M. Given any Borel homomorphism of G0 into M, lemma 5.24 tells us the method of constructing the strict cocycles which define this homomorphism. This construction depends however on the choice of a Borel section for the cosets gGQ. In general, there are no canonical choices for these sections. We shall see later on how, in at least some special cases, these difficulties can be overcome. Remark. I n m a n y problems, M is already a topological group with a second countable topology. I t is easy to indicate a sharpened form of theorem 5.27 which can be formulated in this context. We shall say t h a t a map u of a topological space A into a topological space B is locally bounded if for any compact set K^A, u[K] has compact closure in B. A mild sharpening of theorem 5.27 is contained in lemma 5.29. To give this we need a preparatory lemma, which is classical. Lemma 5.28. Let G be a Icsc group, and M a topological group satisfying the second axiom of countability. Suppose that M is given its natural Borel structure and that m is a homomorphism of G into M which, as a map G into M, is measurable. Then m is continuous. Proof. Since we can replace M by the range of m, we shall assume t h a t m maps 0 onto M. Let U be an open subset of M containing 1 and W an open set such t h a t I e W and WW'1^!!. Since the open sets sW (s e M) cover M, there is a countable family {sx, s2,- • •, sn,• • •} such t h a t
182
GEOMETRY
OF QUANTUM
THEORY
M = {JnsnW. Choose gneG such t h a t m{gn) — sn. If m~1(W) = A, then it is easily seen t h a t G = {Jn gnA. Let f t b e a left H a a r measure. By assumption, the sets gnA are all measurable and hence (JL(A) > 0. By a well known result, there exists an open set B such t h a t e e B^AA'1. Since W W'1^ U, it follows t h a t m[B]<^ U. This proves t h a t m is continuous at e. Since m is a homomorphism, m is continuous everywhere. Lemma 5.29, Let the notation be as in theorem 5.27. Suppose that there is a second countable topology on M which converts M into a topological group and whose associated Borel structure is the given one. Then, any cohomology class y contains a strict cocycle which is locally bounded as a map of GxX into M. Moreover, if there is a continuous section for GjG0, y contains a strict cocycle which is a continuous map of GxX into M. Proof. I t is enough to prove t h a t given a Borel homomorphism m of G0 into M, there are strict cocycles defining m at x0 and having the required properties. Now, ra, by lemma 5.28, is continuous. Let c(x->c(x)) be a regular Borel section for G/G0 (theorem 5.11) such t h a t c(x0) = e. For t e G, h(t) = c(P(t))'H. Then it follows from the regularity of c t h a t h is a locally bounded m a p from G into G0. If b(g) = m(h(g)) for all g e G, b is a locally bounded m a p of G into M. g, g' ~^b(gg')b(g')~1 defines a strict (6r,X)-cocycle/; / is clearly locally bounded and defines m at # 0 . If c is a continuous map of X into G, f is even continuous. Remark. The reader would have noticed t h a t the transitivity of the Borel G-space X has played a crucial role in the main theorems of this section. I t is a natural question to ask whether some of these results can be generalized to the case when X is not homogeneous. Suppose then t h a t X is an arbitrary standard Borel G-space and ^ an invariant measure class. The measure class will now have to play an explicit role since it is not unique unless X is transitive. With respect to a given measure class <€, the problem of describing the cohomology classes of cocycles is then well posed. When there is an orbit G-x0 such t h a t X — G • x0 is a null set for %\ this problem is essentially the problem for the transitive Borel 6r-space X1 = G-x0. Such measure classes on X we shall call transitive. However, not every invariant measure class is transitive. Let us call a measure class ^ which is invariant, an ergodic measure class if any Borel function / , having the property t h a t for each g e G, f(g-x)=f(x) for <€almost all x, is constant almost everywhere. Transitive measure classes are ergodic but not conversely. When X is not transitive under G, there will often exist ergodic invariant measure classes ^ which are not transitive. For any such ^ the problem of describing all the cohomology classes seems quite difficult.
MEASURE
THEORY ON G-SPACES
183
If ^ is transitive and G acts freely on X, i.e., for ^-almost all x e X the stability subgroups are trivial, i.e., consist only of {e}, then there exists obviously an xQ e X such t h a t X — G-x0 is ^ - n u l l and GXQ = e. I n this case, there are no cocycles other t h a n coboundaries. When ^ is assumed to be merely ergodic, this is not true any longer. We shall now construct examples where ^f is ergodic, G acts freely on X, and yet there are cocycles which are not coboundaries. Example 1: G is the discrete additive group of integers, X is the multiplicative group of complex numbers of modulus 1, and M = X. fx is the H a a r measure on X. We denote by z0 an element of X such t h a t the powers zQn (n=l,2,- •) are dense in X. For neG and zeX, n[z] = z0nz. The measure class of //, is ergodic, as m a y be verified easily. Suppose t h a t / is a Borel m a p of X into M. If we define, for (n,x) eGxX,
(m/M'-m-1*) rf(n,x)
(n> i),
= h
(n = 0),
l/^o-^)- 1 "-/^^)- 1
(*<<>),
then rf is a strict cocycle, as is easily verified, and
I t is obvious t h a t every cocycle is equal almost everywhere on G x X to some rf. We now ask when rf is a coboundary. Clearly, this is so if and only if there exists a Borel m a p b of X into M such t h a t rf(n,x)
b(z0nx)b(x)~1
=
for /x-almost all x. I n particular, b{x)f(x)
= b(z0x)
for almost all x. Now, not e v e r y / can satisfy this equation for suitable b. I n fact, we have, for/(#) = #, bjc + i
=
z
o
o fc ,
where & - > bk is the Fourier transform of b, i.e., Sfc = f b{x)xkdix{x),
(k = 0, ± 1 , ± 2 , - • •).
This implies t h a t |S fc | 2 is independent of k, and hence t h a t bk = 0 for all & since b e j£?2(X,/x). This means t h a t 6 = 0, a contradiction. Example 2: I n this example, G = R and ilf the multiplicative group of complex numbers of modulus 1; it is due to Helson and Lowdenslager [1].
184
GEOMETRY
OF QUANTUM
THEORY
We make frequent use of the Pontrjagin theory for locally compact abelian groups (Pontrjagin [1], Kaplansky [2], pp. 54-56). We write R for the additive group of reals in the usual topology and Rd for the additive group of the reals in its discrete topology. Let D be a countable subgroup of Rd and X the multiplicatively written character group of D. I n the (usual) topology of pointwise convergence, X is a compact metrizable group. Since no element of D is of finite order, X is connected. For X E D and x e X we write for the value of x at A. For a n y teR, A - > eitx (A e D) is an element of X. We write i for this element, t -> £ is a homomorphism of R into X. We note t h a t the image of R under this homomorphism is dense in X. Otherwise, by the duality there would exist a A ^ 0 in D such t h a t eitx
=
Y
for all teR, which is impossible. If D contains at least two elements which are rationally independent, t -> i is injective, as m a y be easily verified. The injection t - > t gives rise to an action of J? on X if we set t-x = tx
(t e R, x e X).
I t is obvious t h a t t, x - > t -x is a continuous map of R x X into X and t h a t X becomes an R-space. The H a a r measure /x on X is invariant and ergodic under this action. Our object is to exhibit (R,X,M )-cocycles which are not coboundaries. Suppose XeD. Then the function 0A : t, x—> (1 — eitx) is continuous on RxX and ifjx(t + u,x) = i/)A(t,u-x) + i/jA(u,x) for all t, ueR and x e X. However, I/JA is not real valued. Hence we pass on to
Cj-> +oo,
(b)
A; G D
(c)
and the A; are rationally independent,
V Xfj < oo. i=i
MEASURE
THEORY ON G-SPACES
185
Put c0 = 0, cj = c_j, X_j=— Xj. Then the series 2f= - »fy
(*)
f(t,x) = bfaMx)-1
for almost all (t,x) e RxX. Write at for the map x -> if>(t,x), and for yeX, let ^ be the map a?-> fe^a;)^)-1. Then y ~> ay is a continuous map of JL into the Hilbert space J^2(/x). This implies that (*) is valid for each t for almost all x. Suppose now that y e X is such that
£
cXV>(l-)(l-e«**)\.
Since s is fixed, the series under the exponential is uniformly convergent as x, y vary over X and hence x, y -» cy(#) is a continuous function on X x X. Hence y —> cy is a continuous map of X into J^2(/x). But if we put y = t for t e R, an easy calculation shows that
The continuity of the maps y -> cy and y -> ay shows that for each y e X, we must have for almost all x. If now y e X and (X^y) = 1 for all j , cy(x) = 1 and hence ay(s-x) = ay(x) for almost all x. As i? acts ergodically on X, ay must be a constant almost everywhere. This constant is obviously of modulus 1. The conclusion of the preceding paragraph can be sharpened. Suppose y e X is such that
186
GEOMETRY
OF QUANTUM
THEORY
We are now in a position to give the decisive argument. Since the A; are rationally independent, there exists, for each & = 1 , 2 , « - & ykeX such t h a t (
I
j ± k,
(j > 1),
Write ak for aVk. Since X is compact, there exists a subsequence {ykn} and a y0 e X such t h a t ykn -> y0. Then akn' -> aVQ. As
On the other hand, we claim t h a t , since ck -> +oo,
J aykdfx->0. I n fact, consider the integral JCfx =
exp{4ic Re}d/x(#)
for c > 0. If A ( # ) / 1 for some x e X, then x - » maps X onto a compact nontrivial subgroup of M, which is connected as X is connected, and so coincides with M. The measure JJL then goes over to H a a r measure on M and hence Jc,\ ==
exp{4ic Re m}dm JM
1 = — ^
f2*
exp(4ic cos 0)d0.
Jo
J c A is t h u s independent of A and - > 0 as c -> oo. This proves, as we asserted, t h a t
J aykdii-+0. This contradiction shows t h a t our cocycle I/J is not a coboundary. We shall end this section with an application to t h e theory of quasiinvariant a-finite measures on standard transitive Borel 0-spaces. For a direct derivation of the results which are about to be obtained now, t h e reader m a y consult Mackey's paper [3]. We write B+ for the multiplicative group of positive reals. Let G be a lcsc group, G0 a closed subgroup, X = G/G0 and ft the canonical m a p of G onto X. Let /3(e) = x0. Suppose t h a t F is a positive
MEASURE
THEORY
ON
G-SPACES
187
Borel function on Gx X and a is a quasi-invariant cr-finite measure on X. We write (54)
F ~ a
if for each g e G, x-> F(g,x) is a version of da/dct9'^. Theorem 5.30. Let A and A0 be the modular functions of G and G0, respectively. Write (55)
m(h) = A(A)A0(^)-1
{he G0).
Then, for any quasi-invariant a-finite measure a, there exists a strict (G,X,R + )-cocycle r such that r~a. Any such r defines the homomorphism m at x0. Conversely, if r is a strict cocycle defining m at x0, r~afor some quasiinvariant a-finite measure a; r determines a up to a constant multiplying factor. Finally, an invariant o-finite measure on X is necessarily a Borel measure, and there exists one such on X if and only if A(h) = A0(h) for all heG0. Proof. We begin our proof with a simple observation. Let S^G be a subgroup such that G—S is measurable and has measure zero. Then S = G. In fact, if S^G, and g e G—S, gS^G—S so that gS has measure zero. This implies that S is a null set, a contradiction. This said, let us consider a quasi-invariant a-finite measure a on X. Let F be a positive Borel function on G x X such that for almost all g, x->F(g,x) is a version of da/daf*-1* (theorem 5.10). Equation (22) shows that F is a (G,X,B+)-cocycle. Hence, by lemma 5.26 there exists a strict cocycle r such that r = F almost everywhere on GxX. For each g e G, let fg be a version of da/da (9-1) and let S = {g : r(g,x) = fg(x) for almost all x}. Since the fg satisfy (20) and r satisfies (36), it is easily seen that 8 is a subgroup of G. Further, G—S has measure zero. Therefore S = G. In other words, r ~ a . If r' is another strict cocycle such that r' ~ a, then r' = r almost everywhere on GxX. Hence / « / • by lemma 5.25. Thus, R+ being abelian, r' defines the same homomorphism of GQ as r does, at x0. Let ma denote this homomorphism. We claim that ma is independent of a. Suppose that a is another a-finite quasi-invariant measure. Let a be a positive Borel function which is a version of da/da. Then it is quickly verified that (with r ~ a ) (56)
r' :g,x^
a(x)r(g,x)a(g -x)'1
is a strict cocycle such that r ' ~ a \ But then, r'&r, so that ma> = ma. Let m0 denote this homomorphism of G0, determined by all the measures a.
188
GEOMETRY
OF QUANTUM
THEORY
Suppose now t h a t r is a strict cocycle defining the homomorphism m 0 a t x0. Let ax be a a-finite quasi-invariant measure and rx a strict cocycle such t h a t r1^a1. Since rx also defines m0 at x0, r1^r. Hence there exists a Borel m a p b of X into R + such t h a t (57)
r(g,x) = ^aOr^aOfcter-a;)- 1
for all (g,x) e G x X. If a is the a-finite quasi-invariant measure such t h a t da/da1 = b, then (57) shows t h a t r~a. If a is another a-finite quasiinvariant measure such t h a t r~a also, then a is a constant multiple of a. I n fact, let a be a positive Borel function such t h a t a is a version of da jd<x. Then for each g, a(x) = a(g-x) for almost all x, showing t h a t a is a constant almost everywhere. Suppose now t h a t X admits an invariant a-finite measure, say y. Then the strict cocycle 1 : g, x -> 1 is such t h a t l ~ y . Since 1 defines the trivial homomorphism at trivial. Conversely, let m 0 be trivial and let a be a a-finite quasi-invariant measure. Let r be a strict cocycle defining ra0 a t x0 such t h a t r~a. Since m 0 is trivial, r&l, and so there exists a Borel m a p b of X into R + such t h a t (58)
r(g,x) =
b(g-x)b(x)^
for all (g,x) e Gx X. The equation (58) shows t h a t the cr-finite measure y defined by the equation dyjda — b is invariant. I n other words, X admits a a-finite invariant measure if and only if m 0 is trivial. I t remains to prove t h a t m0 = m. To this end we use the results and notations of lemmas 5.14 and 5.15 and theorem 5.19. Using the argument of lemma 5.29, we construct a Borel m a p d of G into R+, such t h a t (i) d(e) = 1 and d is a locally bounded m a p of G into R +, (ii) there exists a strict cocycle r such t h a t r°(g,g,) = d(gg')d(g,)~1 for all (g,g') e GxG, and (iii) r defines the homomorphism m of G0 at x0. Let y be the Borel measure on G defined by the linear functional (also written y)
y.f^^J{g)d(g)-H^g),
(feCc(0))
Ii being a left H a a r measure on G. We claim t h a t y vanishes on the kernel of the m a p / - ^ / ~. I n fact, suppose / e CC(G) is such t h a t
f f(gh)dp0(h) = 0 J G0
MEASURE
THEORY 1
for all g in G. Multiplying by ud' respect t o /A, we get
ON
189
G-SPACES
where u e CC(G) and integrating with
L (L f^h)d^o(h)y(g)d(g)-^(g) = 0. We interchange t h e order of integration and change g t o gh while integrating with respect t o JX. We then obtain, on using t h e fact t h a t d defines m at; XQ) j
( A O ( A ) - 1 ^u(gh-^)d(g)-^(g)d^g)y^(h)
= 0.
If we now change t h e order of integration once again we obtain (59)
^u~(p(g))d(g)-y(g)dn(g)
= 0.
Since t h e m a p u->u~ maps CC(G) onto CC(X), we can choose u so t h a t u~ is identically 1 on t h e set P[K], where K is a compact subset of G outside of which / is zero. Then (59) leads t o t h e equation (60)
jaf(g)d(g)-1dix(g)
= o.
I n other words, there exists a linear functional y~ on CC(X) such t h a t y ( / ) = y~(/~) for all/eO c (6r). Moreover, we know from lemma 5.15 t h a t given a n y veCc(X), which is real and > 0 , there is a ueCc(G) such t h a t u~—v a n d u>0. This shows t h a t y~ is a nonnegative linear functional. Therefore there exists a Borel measure a on X, such t h a t y~ corresponds t o a. Then, obviously y = a°. Since y is quasi invariant, so is a (corollary 5.17). Fix geG. We shall compute da/da^-v. Let t be a positive Borel version of it on X. Then, by corollary 5.17, t° is a version oidy/dy(a~1\ But, since dy/dfM = d~1i we obtain * V ) = digg'W)-1
= r°(g,g')
for almost all g'. I n other words, for each g e G, x -> r(g,x) is a version of da/dcS9'^. This proves t h a t r~a. B y our definition of m0, m = m0. The above argument shows actually t h a t if r is a locally bounded strict cocycle defining m, there exists a quasi-invariant Borel measure a such t h a t r~a. Suppose now t h a t X admits a n invariant cr-finite measure a. m is then trivial a n d our remark a t t h e beginning of this paragraph shows t h a t there exists a Borel measure y such t h a t 1 ~ y. B u t this implies t h a t a is a constant multiple of y b y corollary 5.21. a is t h u s a Borel measure. The proof of t h e theorem is complete. Corollary 5.31. / / G is unimodular, G/G0 admits an invariant Borel measure if and only if G0 is also unimodular. If G is arbitrary but GQ is compact, then GjG0 admits an invariant Borel measure.
190
GEOMETRY
OF QUANTUM
THEORY
We end this section with a n example of a unimodular G a n d a nonunimodular G0. Let G=SL(2,C), the group of 2 x 2 matrices g (61)
la g= {
b\ \
(ad-bc
= 1)
of determinant 1. g - > a,b, c, d gives a n injective homeomorphism of G into C4. The group G, equipped with t h e relative topology obtained b y imbedding it as above in C4, is a lcsc group. Let G0 be the subgroup of all elements tatZ, where (62)
ta,z = ^
Z
\
(a,zeC,a^0).
G0 is closed. We claim t h a t G is unimodular and G0 is not, so t h a t X = G/G0 does not admit any CT-finite invariant measure. For the unimodularity of G we shall be more general and prove t h a t G admits no nontrivial continuous homomorphism into R + . Suppose t h a t m is one such. Since m(x) = m(yxy~1) for all x, y in G, it follows t h a t for any matrix x in G with distinct eigenvalues, m(x) = m(y) for all matrices y with t h e same eigenvalues as x. Let G' be t h e set of all matrices x in G with distinct eigenvalues and H the subgroup of diagonal matrices. Then, t h e restriction of m t o H is a continuous homomorphism of H into R + with the property t h a t m(x) = m(y) whenever the diagonal entries of x and y are permutations of one another. I t is easy t o deduce from this t h a t m is identically 1 on ^ . Since for a n y g e G' there is a n a e G such t h a t aga~x e H, m(g) = 1 for all g e G'. Now G' is a dense subset of G and so m is identically 1 on G. I n particular, G is unimodular. GQ is not unimodular. To see this, let (63)
77 = {tat0 : a e C*}.
Then H is a n abelian closed subgroup of G0 and is hence unimodular. If G0 were unimodular, then there would exist a n invariant Borel measure oh X0 = G0IH. N O W in each left #-coset in GQ there exists exactly one element of the form tlt2 (z e C); if we use the complex number z t o parametrize the coset tlfZH, t h e action of G0 on X0 becomes (64)
tCt2:l~>c2Ucz.
I t is obvious t h a t t h e space of complex numbers does n o t admit a n y Borel measure invariant under the maps (64). Thus GQ is not unimodular. I t is of interest t o note in this case t h a t X = GjG0 is actually compact. I n fact, if K denotes the subgroup of unitary matrices in G, it is a n easy consequence of t h e existence of the J o r d a n canonical form of an arbitrary matrix in G, t h a t G = KG0. Hence X is the image of K under the canonical mapping of G. I n particular, X is compact.
MEASURE
THEORY ON
OSPACES
191
6. B O R E L G R O U P S AND T H E W E I L TOPOLOGY Our main concern in this section is to examine how the topology of a lcsc group can be recovered from the measure-theoretic structure. Suppose t h a t G is a lcsc group and t h a t /x is a left H a a r measure. Then it is classical t h a t for a n y Borel set E with /UL(E)>0, the set EE'1 contains a neighborhood of e. This result hints t h a t the topology of G is completely determined by JJL. That this is indeed so was proved by Weil [1] (cf. Halmos [1], pp. 257-289 for a detailed discussion of Weil's results). For our purposes, this result of Weil is somewhat inadequate. The main theorem of this section is a sharpening of Weil's theorem, first proved by Mackey [5]. We shall use this sharpened form of Weil's result in a significant way in Chapter V I I . This sharpening essentially comes out of the following observations. The first point is t h a t starting with only a left invariant measure class on a separable Borel group G one can show t h a t the group possesses a left invariant measure. The second point is t h a t the Weil topology for G is second countable and generates the original Borel structure on G. The third observation now consists in imbedding ( J a s a dense subgroup of a lcsc group G*. The fact t h a t G is standard now implies t h a t G is a Borel set in G*. This leads easily to the identity G = G*. In other words, every standard Borel group admitting a left invariant measure class is actually a lcsc group under its Weil topology. This is the sharpening of Weil's theorem we referred to. I t is due to Mackey [5]. Since the main aim of our analysis is to obtain topological results from measure-theoretic assumptions, it is convenient to start with groups without a n y topology. Throughout this section, G will denote a separable Borel group, i.e., a group G with a Borel structure which converts G into a separable Borel space such t h a t the m a p x, y -> xy ~~1 of G x G into G is Borel. Since the left translations x -> zx are Borel automorphisms, we m a y consider measure classes invariant under these. We call them left invariant. I n an analogous fashion we define right invariant measure classes. A measure class which is both left and right invariant is called bi-invariant. Lemma 5.32. Let Gbea separable Borel group on which there exists a left (or right) invariant measure class. Then G admits a bi-invariant measure class. Proof. Let ^ be a left invariant measure class and A a finite measure in ¥>. We shall prove first t h a t for any Borel set E^G, s -> X(Es) is a Borel function of s. In fact, the map T : s, t -> s, ts is a Borel automorphism of the Borel space G x G and hence T[GxE] is a Borel set in GxG. I t is then obvious t h a t Es = {t: (s,t) e T[G x E]}, and hence the function s -> X(Es) is Borel by Fubini's theorem.
192
GEOMETRY
OF QUANTUM
THEORY
Define the measure Ax b y (65)
X^E) = f
\(Es)d\(s).
Xx is easily seen to be a measure and A1(G) = \(G)2>0. We claim t h a t the measure class determined by X± is bi-invariant. I t is enough to prove t h a t if X1(E) = 0, then X1(Ez) = X1(zE) = 0 for all ZEG. Suppose \x(E) = 0. Then, as X{Es) > 0, X(Es) = 0 for A-almost all s. Since Xetf and <€ is left invariant X(zEs) = 0 for A-almost all s, showing t h a t X1(zE) = 0. On the other hand, if N is a Borel set such t h a t X(N) = 0 and X(Es) = 0 for SEG-N, then X(z~1N) = 0 and A(#zs)=0 for seG-z^N. This proves t h a t X±(Ez) = 0 also. A similar argument can be given starting with a right invariant measure class. The lemma is proved. Lemma 5.33. Let G be a separable Borel group and let ^ be a bi-invariant measure class. Then & contains left invariant and right invariant a-finite measures. Proof. Let A be a finite measure in # , such t h a t X(G) — l. Define As b y XS(E) = A ^ " 1 ^ ) . Then the measures {As} are all mutually absolutely continuous. Since s->Xs(E) is a Borel function on G, corollary 5.6 is applicable and therefore there exists a Borel function r on GxG such t h a t for A-almost all s, t -> r(s,t) is a version of dX/dX^'^. Since A* and A are mutually absolutely continuous, r(s,t) > 0 for almost all t for A-almost all s. Hence r is positive (A x A)-almost everywhere on GxG. We m a y thus assume t h a t r is positive on GxG. The arguments of theorem 5.10 can now be used to conclude t h a t for (Ax Ax A)-almost all (s,t,u) e GxGxG, r(st,u) =
r(s,tu)r(t,u).
Then there exists a positive Borel function b on G such t h a t (66)
r(s,t) = b(st)/b(t)
for (A x A)-almost all (s,t) e GxG. I n fact, all we have to do is to imitate t h e argument of the first p a r t of the proof of lemma 5.26; the only fact used there is the bi-invariance of the class of null sets involved which is certainly satisfied in our present setup. The measure ^ such t h a t d^dX = b is then cr-finite and belongs to %\ From (66) we conclude t h a t for A-almost all s, d/jildfxis ~1} is equal to 1 almost everywhere. Thus /x = /x(s~1} for A-almost all s. Now the set S of all s E G such t h a t /x = /x (s_1) is a subgroup. Since
MEASURE
THEORY
ON G-SPACES
193
this subgroup has a complement of A-measure zero, it must coincide with G (cf. remark in proof of theorem 5.30). This proves t h a t /x is left invariant. A similar argument yields a right invariant measure in the measure class <€. The proof is complete. Let now ffl be a separable Hilbert space and °ll the group of all unitary operators on Jf. We shall give the strong topology for °ll\ i.e., the smallest topology which makes all the maps JJ-^Uf (f e J^) continuous. If { Un} is a sequence of elements in °U, Un^ V e °ll if and only if Unf —> Vf for all / belonging to some set D such t h a t the linear combinations of vectors of D are dense in J^. Lemma 5.34. °tt is a metrizable topological group satisfying axiom of countability and its Borel structure is standard.
the second
Proof. Since 3f is separable, there exists a countable set D dense in Jf7, and the topology for 9/ is clearly the smallest one which makes all the maps U - > Uf (fe D) continuous. As there are only denumerably many maps, it follows from standard topological arguments t h a t the topology for °ll is metrizable and second countable. To prove t h a t °ll is a topological group we must show t h a t the m a p U, V -> UV'1 is continuous. But, as the elements of °ll are unitary, we have, for a n y / e JfJ || Z 7 F - Y - C70 F 0 " Y|| < || F( F0" Y) - F 0 ( F 0 " y ) || + || ?7( F 0 - Y) - t70( F Q - V)||. This inequality proves at once t h a t °ll is a topological group. To prove t h a t the associated Borel structure of ^ is standard, we must exhibit a Borel isomorphism of ^/ with a Borel space which is known to be standard. Let X be the topological product of countably m a n y copies of Jj? and let ^ = {/i,/ 2 5* * •} be a countable orthonormal basis for Jtf*. For each U e %, let £(£/) be the element {Ufl9 Uf2, • • •) of X ; (67)
£:U-+(Ufl9Uf29...).
£ is clearly a homeomorphism of °2l into X. We want to show t h a t the range of £ is a Borel set in X. Clearly, (zl9 z2, • • •) e X lies in the range of £ if and only if {^l3 z 2 , • • •} is an orthonormal basis, i.e., if and only if (zm,zny (68)
= Smn 2
2 KWi)! = 1
(Kronecker delta)
for all i.
m
These are countably m a n y equations and each one of them defines a Borel subset of X. Thus the range of £ is Borel. £ is thus an isomorphism of % with the Borel subspace of X defined by (68). As this is a Borel subset of the standard space X, we can conclude t h a t °ll is standard.
194
GEOMETRY
OF QUANTUM
THEORY
Let now G be a separable Borel group and let /x be a left invariant (j-finite measure on G. Let Jfu be the Hilbert space of complex valued (equivalence classes of) Borel functions / on G such t h a t ||/|| 2 = £ \f(g)\2d(*(g) < co.
(69)
For a n y g e G, let Lg be the operator of Jfu defined by (70)
(Lgf)(t)=f(g~H)
(teG).
Lg is clearly a unitary operator of Jt?u. We have Le = 1, L
9192 =
L
the identity operator,
9lL92
f
°
r a
U
01> 02 G
G
'
Denote by ^ M the group of all unitary operators on J^u. Lemma 5.35. Let G be a separable Borel group which has a left invariant a-finite measure p. Then the map L(g —> Lg) of G into ^u is a one-one Borel map. If G is standard, its range L[G] is a Borel set in %u and L is a Borel isomorphism of G with L[G]. Proof. I n view of the definition of the Borel structure on ^ , we must, in order to show t h a t L is Borel, prove t h a t for each / e J^, g -> Lgf is a Borel m a p of G into ^ . Since ^ is separable, we can use lemma 5.4 to reduce this t o showing t h a t for each / , / ' e J^, g -> (Lgf,f'y is a Borel function on G. B u t (* denotes complex conjugation)
= f
f(g-H)f'(t)*dp(t),
JG
which is obviously Borel, as g,t-^f(g H)fr(t)* is a Borel function on OxG. We show next t h a t L is one-one. As L is a homomorphism of G into ^u, this reduces t o showing t h a t if g^e, Lgf^f for s o m e / e J f r Suppose t h a t Lgf=f for all f e Jtf^. Since the Borel structure of G is countably generated and /x is a-finite, there exists a sequence {En} of sets of finite /x-measure such t h a t {En} generates the Borel structure of G. Let fn = ^En. Then fn e J^u and the sequence {/n} separates the points of G (since each single point set of G is a Borel set). Let N be a set of ^-measure zero such t h a t for x £ N, (72)
/ . ( r 1 * ) =/„(*)
for all n. The fact t h a t {/n} is a separating sequence now implies t h a t g - 1X = x for all x e G — N. The t r u t h of this for even one xeG — N implies g—e. L is thus one-one. The remaining statements follow from the fact t h a t /„ is standard.
MEASURE
THEORY
ON G-S PACES
195
We shall now introduce the Weil topology. Let G be a separable Borel group admitting a left invariant a-finite measure /x. Consider for each g e G, the unitary operator Lg in Jf?u. The Weil topology for G is then defined to be the smallest one which makes the maps (73)
g-+LJ
of G into J ^ continuous for all / e J^. I n view of the definition of the strong topology for ^u, it is clear t h a t the Weil topology is t h a t topology for G which makes the m a p g - » Lg of G into %u a homeomorphism. Thus 6r, under the Weil topology, is a topological group satisfying the second axiom of countability. Let E and F be Borel sets of finite measure. We write (74)
P(E,F)
= fi((E-F) u (F-E)).
Then from the relations P(E,F)
= f
\XE(t)~xF(t)\2d^t),
JG P(E,F)
=
P(xE,xF),
it is easy to deduce t h a t # n -> a; in the Weil topology for G if and only if p(xnE,xE) -> 0 for all Borel sets E of finite measure; it is even sufficient to restrict the E9s to a Boolean ring of sets of finite measure which generate the Borel structure of G. Lemma 5.36. If G is a standard Borel group and \i a left invariant a-finite measure on it, the Borel structure generated by the sets which are open in the Weil topology coincides with the original Borel structure. Moreover, if G is a Icsc group, the Weil topology coincides with the original topology. Proof. Let G± be the range of L(g -> Lg). Let &* be the Borel structure for G generated b y the Weil topology and let & be the original Borel structure. Then, L being a homeomorphism of G onto Gl9 L is a Borel isomorphism of (G,&*) with the Borel subspace of ^ defined b y Gx. By lemma 5.35, L is also a Borel isomorphism of (G,&), with the Borel subspace of ^ M defined by Gx. Hence ^ = ^ * . Let us now assume t h a t G is a lcsc group. We want to prove t h a t its topology coincides with the Weil topology. Since both topologies are metrizable, it is obviously sufficient to prove t h a t for a n y sequence {gn} in G, gn-> e in the given topology of G if and only if p(gnE,E) -> 0 for each Borel set E of finite measure. All topological statements, in the argument to follow, refer t o the given topology on G. Let then gn -> e. If {Vs} is a decreasing sequence of compact symmetric (VS=VS~1) neighborhoods of e which form a base for G at e, then for a n y compact set K, it is easy to see
196
GEOMETRY
OF QUANTUM
THEORY
t h a t p | s {VSK) — K. Fix a Borel set E of finite measure, and let s > 0. Choose a compact K^E so t h a t fji(E — K)<e and choose s so t h a t /x(F s i£ — i f ) < £ . If n0 is chosen so t h a t gne Vs for n>n0, then for all such ?i, P f a n ^ ) < P(gnK,K) + 2P(E,K) < 2fi(VsK-K)
+ 2e
< 4c.
I n the converse direction, suppose ^ n 4> e. By passing to a suitable subsequence we m a y assume t h a t there exists a compact neighborhood V of the identity such t h a t gn $ V for all n. Choose a compact neighborhood W of e such t h a t WW"xc V. Then gnW nW=0 for all w, so t h a t P(gnW, W) = 2/n(W) > 0 . This shows t h a t grn cannot converge to e in the Weil topology. This completes the proof of the lemma. Consider now a standard Borel group G with a nonzero left invariant Borel measure /x. We equip G with the Weil topology. Following standard terminology, we shall say t h a t a subset A c; G is bounded if for each open set £7 in 6r one can find elements x1, • • •, xn e G such t h a t
-4 <= 0 xtU; i=l
I t is obviously enough to require this only for open sets containing e. We shall say t h a t G is locally bounded if some open set containing e is bounded. Notice t h a t open sets in G are Borel (lemma 5.36); if U is open a n d nonempty, there are elements G such t h a t G = \Jn xnU, from which it follows at once t h a t 0 < /JL( U) < oo. Lemma 5.37. G is locally bounded. Every bounded Borel set is of finite measure. Proof. Before we take up the proof of this lemma we make a simple observation. Let E be a Borel set with E = E~X and 0 < fi(E) < oo. Then there is a Borel subset F^E with F = F~* and 0 0 for all x e E and f fd/jL
< e}.
Since a; -> p(xE,E) is a continuous function on 6r, vanishing a t e, JV(^ : e) is an open set containing e. We claim t h a t 0 < fi(N(E
: e)) < oo.
MEASURE
THEORY
ON G-SPACES
197
We need only to prove t h a t fi(N(E : e)) < oo. Now, we have XE(t)XE(t-1X)dfl(x)dH.(t)
jj
= JXE(t)fM(tE)dfJL(t)
GxG
G
= M(^) 2 On the other hand, as E = E~1, jj
Xs^XEit-^d^dfiix)
= jj
=
XEMxEix-tydpWdpix)
fi(xE n E)dfji(x). G
Therefore,
fji(E)2 >
f n(xE n E)dfi(x). N(E:e)
But,
if XGN(E
:e), P(xE,E)
= 2p(E)-2ii(xE
n E) < e
so t h a t /x(^> This proves our claim t h a t 0
(r(E)-el2)p(N(E:e)). : e))
At this stage, we know t h a t sufficiently small open subsets of G are of positive finite measure. Let U be any open set containing 1 with 0 < fji(U)
(n = 1 , 2 , . . . ) -
i= l
Choose an open set L containing e such t h a t LL_1clf n V. Then it is easy to see t h a t the sets x1L, x2L, • • • are disjoint and contained in VV'1. 1 Since VV~ ^U,0
(75)
8(»,y) = S / ( ! f n / ^ >
2-"
is easily verified to induce the Weil topology; 8 is left invariant.
GEOMETRY
198
OF QUANTUM
THEORY
Lemma 5.38. Let Gbea second countable topological group and let it satisfy the two conditions: (a) (b)
there is a left invariant metric 8 for the topology of G, if {xn} is a 8-Cauchy sequence, then {# n - 1 } contains a 8-Cauchy subsequence.
Let G* be the completion of the metric space (G,8) with G considered as imbedded in 6r*. Then there exists exactly one way to convert G* into a topological group such that G becomes a dense topological subgroup of it. Proof. As 8(xn,xm) = 8(xn~1xm,e), it follows t h a t {xn} is a S-Cauchy sequence if and only if xn~ 1xm -> e as n, m -> GO. Let G* be the completion of G (under 8); we regard G as imbedded in G*. We denote also by 8 the extension of the metric 8 to 6r*, which is a second countable topological space in its metric topology. We shall now show t h a t if {xn} and {yn} are Cauchy sequences in G, so is {xnyn}. Let V be any open neighborhood of e and let us choose an open neighborhood W of e such t h a t W3^ V. Let k be an integer such t h a t yn~1yrn£ W for n,m>k. Choose an open neighborhood W of e such t h a t Wf^W and yk~1W'yk<^W. Finally, select an integer kx>k such t h a t xn~1xm e W for n, m>k±. Then, if n, m>k±, yn
Xn
'%mym
= (yn
Vkf'Vk
\Xn
x
vx)yk'Vk
Urn
eW{yk^W'yk)W <= V. This proves t h a t {xnyn} is Cauchy. From this it follows easily t h a t there is a unique continuous m a p p* of 6r* x 6r* into G* such t h a t p*(x,y)=xy (x,y e G). We define, for x,y e G*, (77)
xy =
p*(x,y).
This product operation is clearly associative and xe = ex = x for all XEG*. If x E G* and {xn} is any sequence in G converging to x, then, for any limit point x' of the sequence {x n _ 1 } we have e. Condition (76b) now implies t h a t x'1 exists and t h a t x^1 converges to x"1. The rest of lemma 5.38 now follows trivially. Corollary 5.39. Let G be a locally bounded second countable topological group whose topology is induced by a left invariant metric 8. Then G satisfies (b) of (76). In this case G* is a Icsc group and for any compact set K of G*, K n G* is a bounded set in G. Proof. Suppose V is a bounded set in G. We claim t h a t any sequence in V has a Cauchy subsequence. To prove this we proceed as follows. Let
MEASURE
THEORY
ON G-SPACES
199
{Un} be a sequence of open subsets of G forming a base for the topology of G at e such t h a t U~+1Un + 1^ Un for all n. If {xn} is a sequence in V having infinitely m a n y elements, it follows from the boundedness of V t h a t , for each k, some translate akUk contains infinitely m a n y of the xn. This means t h a t for some infinite set Zk of the integers, xn ~ 1xm e U^1!]^ Uk_x whenever n, m e Zk. By choosing the Z's to decrease with k and using the Cantor diagonal procedure we m a y assure ourselves t h a t {xn} has a subsequence which is Cauchy. We shall now prove t h a t if {xn} is any Cauchy sequence, { # n _ i } has a Cauchy subsequence. I t is enough to exhibit a bounded set containing all but finitely m a n y of the elements £ n - 1 . Let U be a bounded neighborhood of e. Now there is a & such t h a t xn~1xke U for n>k. Hence xn ~1 e Uxk ~x for all n > k; as Uxk ~x is bounded, our assertion is proved. G* can thus be regarded as a second countable topological group and G a dense subgroup. To prove t h a t G* is locally compact, it is enough to exhibit some compact subset of G* with nonnull interior. Let V be any subset of G, bounded and open in G and containing e. Since any sequence in V contains a Cauchy subsequence, V, the closure of V in G*, is compact. Since G is second countable, there is a sequence {xn} in G such t h a t = Un *nV- We claim t h a t G* = {Jn xn(VV). I n fact, if y* e G* and {yn} is a sequence in G such t h a t yn-> y*, there exists an m such t h a t ym ~ 1yn e V for all n>m. Thus yne ymV for all n>m. Choose an integer k such t h a t ym E xkV. Then yne xkVV for all n>m, showing t h a t y* e xkVV. By the category theorem of Baire, some xkVV contains an interior point. This implies t h a t V V is a compact set with nonempty interior and shows t h a t G* is locally compact. For the last assertion, let K be compact in G* and U an open set in G containing e. Let C7* be an open subset of G* such t h a t U=U* n G. Select an open set F * in G* such t h a t e e V* and 7 * 7 * ^ U*. There exist #i*> #2*5* * * > xk* e £r* such t h a t K^\^Jf=1 xfV*. We choose xi e G
(l
such that xr^eV*.
Then
K^\JtsslxiV*V*^\J?=1ziU*
and so K n ffc(J{c=1 ^£7. We come to our final lemma. Let G be a separable Borel group with a left invariant a-finite measure /x. Lemma 5.40. / / G is standard, then G is already locally compact under its Weil topology. Proof. Under the Weil topology, G is locally bounded and second countable. We also know t h a t the Weil topology is induced by a left invariant metric. Corollary 5.39 is thus applicable and hence we m a y regard G as a dense subgroup of a lcsc group G*. For any Borel set ^ 4 * c £ * 5 A* n G is Borel in G and hence, (78)
p* : A* -> fi{A* n G)
200
GEOMETRY
OF QUANTUM
THEORY
is a measure on G*. Now \x is finite for bounded sets and so for any compact K^G*, IJL*(K) = IJL(K n G)
JG*
JG*
and hence
f f*(t)dn*(t) = f f™\t)dn*(t). JG*
JG*
Since this is valid for a l l / * e CC(G*), it is clear t h a t ix* is invariant under left translation by g*. I n other words, /x* is a left H a a r measure. xt* is evidently nonzero. Now G is standard and the identity m a p of G into G* is Borel and oneone. Hence G is a Borel set in G*. B u t then
H*(G*-G) =
fi((G*-G)nG)
= 0. Thus 6r is a subgroup which is also a Borel set with a complement of zero measure in G*. Therefore, the observation we made at the beginning of theorem 5.30 is applicable and enables us to conclude t h a t G = G*. I n other words, G is already locally compact in its Weil topology. Combining lemmas 5.32 through 5.40 we obtain the following theorem. Theorem 5.41. (Mackey [5]). Let G be a standard Borel group. If G admits a left (or right) invariant measure class, then G admits a left or right invariant o-finite measure. Moreover, there exists exactly one topology for G which converts G into a Icsc group and whose corresponding Borel structure is the original one on G. N O T E S ON C H A P T E R V The measure theoretic study of the orbit spaces of locally compact groups has been a very active subject in recent years because of possible applications to representation theory, number theory, statistical mechanics, and so on. For a survey treating m a n y of these aspects see R. Zimmer, Ergodic Theory, group representations, and rigidity, Bull. Amer. Math. S o c , 6 (1982), p p . 383-416.
CHAPTER VI SYSTEMS OF IMPRIMITIVITY
1. D E F I N I T I O N S This chapter discusses the notion of a system of imprimitivity for a group and, in some cases, describes all such systems to within unitary equivalence. These problems were first solved by Frobenius for the case of finite groups. For infinite groups their study is more recent. Their first appearance seems to be in a series of examples constructed by Murray and von Neumann [1], [2] in their theory of rings of operators. Anticipating terminology, we can say t h a t the examples of Murray and von Neumann form a class of ergodic systems of imprimitivity. On the other hand, the first work on transitive systems appeared in 1939. I n his great 1939 paper Wigner underscored their importance for building up the theory of representations of the inhomogeneous Lorentz group and studied, although in quite an implicit fashion, the systems of imprimitivity associated with this group. Also Gel'fand and Naimark [1] based their solution to the problem of describing the unitary representations of the "ax + b" group on an analysis of certain natural systems of imprimitivity. The systematic development of this concept is due to Mackey. I n a series of papers (cf. Mackey [7] for all references) he obtained the main general results of the subject and pointed out the far-reaching scope of the ideas involved therein. The variety of applications of these results is quite impressive. They enable one, among other things, to obtain the equivalence of matrix and wave mechanics; they lead to a concise description and classification of the relativistic wave equations; they make possible a unified approach to the theory of representations of the commutation rules; also, on the purely mathematical side, they lead to the notion of induced representations of locally compact groups which has played an important role in the theory of group representations; and finally, they have very interesting implications in the theory of invariant subspaces and factorization of analytic functions (cf. Helson [1]). We begin with the notion of a representation. Let G be a lcsc group and J$? a Hilbert space (always separable). A (unitary) representation of G in J^ is a homomorphism (1)
U:g-+U„ 201
202
GEOMETRY
OF QUANTUM
THEORY
of G into the group °ti of all unitary operators on J f such t h a t for each / e ^f7, the m a p (2)
9-+UJ 7
is continuous from G into Jf . Since °ll is second countable in its strong topology (lemma 5.34), lemmas 5.28 a n d 5.4 show t h a t for the homomorphism g -> Ug to be a representation it is sufficient t h a t (3)
9-*
(f.f'eJT)
are measurable on G for a l l / , f'eJf.A representation U in ^ f is said to be equivalent to a representation £7' in £?' if there exists a unitary isomorphism (4)
W : Jf - > JT
of 2tf onto J f ' such t h a t (5)
Ug' =
WU.W-1
for all g e G. The representation U on J f is said to be irreducible if the only closed linear manifolds of J f which are invariant under all t h e linear operators Ug are 0 and Jf. If G is a locally compact abelian group, it is well known t h a t its irreducible representations are none other t h a n its characters; i.e., continuous homomorphisms of G into the multiplicative group of complex numbers of modulus 1. Two characters are equivalent if and only if they are identical. The theory of characters was developed in its definitive form b y Pontrjagin [1]; this development was used by Weil [1] to obtain the Plancherel formula for locally compact abelian groups. On the other hand, when G is not necessarily abelian, b u t compact, a definitive theory can again be developed. The beginnings of this theory can be traced a t least to the theory of spherical harmonics. B u t in its modern form, the theory is undoubtedly the creation of H e r m a n n Weyl. Using powerful and original (transcendental) methods he obtained the representations of arbitrary compact groups, his version of the Plancherel formula for such groups (the Peter-Weyl theorem), and rounded off the theory with a succinct description of the representations of compact semisimple Lie groups (cf. his papers [3], [4], [5] in Mathematische Zeitschrift and his expositions [1] (to an audience of physicists) and [2]). The Lorentz group is neither compact nor abelian. The study of its representations must therefore be based on the theory of representations of lcsc groups which are neither compact nor abelian. This theory is vast a n d quite recent. A survey of some of this recent work is given in Mackey's address [7]. Suppose t h a t G is a lcsc group. Let X be a standard Borel space. We relate G and X by assuming t h a t X is a Borel 6r-space. Let / be a
SYSTEMS
OF
IMPRIMITIVITY
203
separable Hilbert space. A system of imprimitivity for G acting in ^ f is a pair (U,P), where P(E - > PE) is a projection valued measure on t h e class of Borel subsets of X (the projections being defined in Jf), and U(g - > Ug) is a representation of G in J f such t h a t the relations (6)
UgPEUg~i
= Pg.E
are satisfied for all g e G and all Borel sets E^X. The pair (U,P) is said to be based on X. The notions of equivalence and irreducibility are defined for systems of imprimitivity in a fashion similar to the corresponding notions for representations. Two systems (U,P) (in J f ) and (U',Pr) (in Jf') based on the same 6r-space X are equivalent if there exists a unitary isomorphism W of £F onto J f ' such t h a t
P/
WPEW'1
=
for all e G and all Borel sets E^X. The system (J7,P) is said to be irreducible if there is no closed linear manifold of <#?, other t h a n 0 and ^ which is invariant under all PE and all Ug. Note t h a t if U is an irreducible representation, (U,P) is an irreducible system of imprimitivity, but the converse is generally false. If (U^P1) (i = 1, 2, • • •) is a sequence of systems of imprimitivity all based on X, with (U^P1) acting on J^\ then it is natural to define their direct sum as the system (U,P) acting in Jf?, where
3tf = © ^ i
and where Ug:(f1,fa,---)-+{U91f1,U,%,---). I n symbols: @{U\Pi).
(U,P) = i
I n order to be able to handle the concepts of irreducibility, equivalence, and so on, it is convenient to transform these geometric concepts into algebraic ones. I t is customary to do this with the help of the notion of the commuting ring. We work with representations, b u t all of our remarks apply also to systems of imprimitivity. Let U be a representation of G in Jf. The commuting ring of U is the set 21 of all bounded operators A in J f such t h a t (8)
UgA
=AUg
for all g e G. Obviously % is an algebra and, if A e % then A* e %. Moreover it is easily seen t h a t 21 is closed under the weak topology. 21 is thus a von Neumann algebra (cf. Section 4, Chapter V I I , Volume I). The
GEOMETRY
204
OF QUANTUM
THEORY
structure of 2t determines very completely the geometry of the action of U. For example, a closed linear manifold L is invariant under U if and only if the projection PL on L belongs to 21. I n this case, by restricting each Ug to L we obtain a representation of G in L, called the subrepresentation of U defined by L. Thus, every projection in 21 defines a unique subrepresentation of U. This subrepresentation is irreducible if and only if the associated projection, say Q, in 21, is minimal, i.e., if Q'
(9)
=
f{x)dvUtV(x),
where vUtV is, as usual, the complex valued measure on X defined b y
vuJE)
= (PEu,v}
(u,veJf).
vUtU is a finite measure for any u e Jf7 and vu/u(X) = \\u\\2. Moreover,
M/«« 2 = J \m\'dvu.u(*)
{uetf).
The m a p p i n g / -> Af from the algebra of all bounded Borel functions into the algebra of all bounded operators in Jf7 has the properties described below; their verifications are easy: (i) (ii)
(10)
(iii)
/—> Af is an algebra homomorphism, if/ is real, Af is self-adjoint; more generally Af* = Af*,
11,4/11 < sup
\f(x)\.
oceX
Moreover, using standard arguments, we obtain, from (6), (11)
U„AfUa-i 9
w h e r e / ' is the function
= 1
x->f(g~ -x).
A^,
SYSTEMS
OF
IMPRIMITIVITY
205
An important special case arises when X is a locally compact Hausdorff space satisfying the second axiom of countability and the action of G on X is continuous. I n this case, it is perhaps more natural to replace the projection valued measure P by the homomorphism / —> Af (f eCc(X)). We have: Lemma 6.1. Let f -> Af be a homomorphism of the algebra CC(X) into the algebra of all bounded operators on a separable Hilbert space Jf7 satisfying the relations (10) and such that the linear manifolds spanned by the ranges of all the operators Af is dense in 3f. Suppose further that g —> Ug is a unitary representation of G in J f such that (U,A) satisfies (11). Then, there exists exactly one system of imprimitivity (U,P) based on X such that, for each feCc(X), Af is related to P by (9). Proof. For a n y / in CC(X) which is real and > 0 , we observe t h a t Af>0; in fact, / = / i 2 , where f1 is also real and > 0 and Af = Afl2>0. Consequently, the m a p / - ^ (Afu,u) defines a Borel measure vu>u on X. If K is any compact set a n d / i s an element of CC(X) such t h a t 0 < / < 1 and / = 1 on K, the inequality (hi) of (10) shows t h a t
"».«(*) ^ Ml 2 vUtU is thus a finite measure and vutU(X) < \\u\\2. At the same time, if u, v e J^ there exists a unique complex valued measure vUtV such t h a t (12)
fdvu,v
(feCc(X)).
For any Borel set E, the map u, v -> vUiV{E) is *-bilinear and 0 < vUtU(E) < \u\2 for all %£«#. Therefore, there exists a self-adjoint operator PE with 0 < PE < 1 such t h a t (13)
vUtV(E) =
for all u, v E Jf. We want to show t h a t E -> PE is a projection valued measure. Suppose K is a compact set and {fn} is a sequence of real nonnegative functions in CC(X) such t h a t / n j XK pointwise. Then (12) shows t h a t (14)
for all u, v e J f ; i.e., Afn converges weakly to PK (=> will denote weak convergence). I f / is any element of GC(X), AfnAf=>PKAf and AfAfn=> AfPK so t h a t PK commutes with all Af. Moreover, if f>0, then ffn \fxK and hence
= J ->
ffndvUtV fdvUtV.
206
GEOMETRY
OF QUANTUM
THEORY
I n other words, if/ e CC(X) is real and > 0, (15)
= f /»„„,„.
If we choose another compact set £ and a sequence {/n'} of real nonnegative elements oiCc(X) such t h a t / n ' j XL> w e obtain, from (15) and t h e convergence Af . => PL, the equation (PLu,PKv}
=
(PKnLu,v>.
Now, for fixed K, E - > (PEu,PKv} and E -> (PEnKu,v} are valued measures which coincide for all compact sets L. Hence incide for all E. For arbitrary b u t fixed E, we now use the coincidence for all K, and conclude t h a t for all Borel sets E and
complex t h e y coresulting F,
(PEu,pFvy = (PEnFu,vy for all u,veJ$f, i.e., PE^F = PEPF' For E = F this shows t h a t PE is a projection. The m a p E - > P £ has now all the properties of a projection valued measure except perhaps t h a t Px m a y not be 1. B u t Px = 1; for, if P x / 1 , there will be a nonzero u e Jt? such t h a t P x w = 0. This would imply t h a t vUtU = 0 and hence t h a t Af*u = 0 for a l l / e O c ( X ) . B u t then u would be orthogonal to the ranges of all the Af, a contradiction. Finally, by a routine transport of structure, the equation UgAfU9-x
= A<„
is seen t o lead t o t h e equation
ugpBua->-
= pg.E.
This proves the lemma. We shall now give some examples of systems of imprimitivity. X a n d G have the usual meanings, i.e., X is a standard Borel (r-space, G is lcsc. Example 1: Let a be a cr-finite invariant measure on X, Jf = J£?2(X,a); Ug and PE are defined by /io\
E
J = XEf> (UJ)(x)=f(g-i-x).
The identities Ugig2=U9iUg2, Ue = l as well as the relations (6) are t h e results of routine calculations. Further, the invariance of a implies t h a t Ug is unitary. For / , / ' e ^ the m a p g, x -> f(g~1-%)f'(x)* is Borel on G x X and hence the m a p
9^
=
jj(g-1-x)f'(x)*da(x)
SYSTEMS
OF
IMPBIMITIVITY
is a Borel map. U is thus a representation and (U,P) imprimitivity.
207
a system of
Example 2: Our second example is a variation on the theme of the first one. Let J f be a separable Hilbert space. We then define, for any a-finite measure a on X, J^ = ^2(X,J^,a), to be the vector space of aequivalence classes of Borel maps / of X into JT such t h a t I/I2 = j
(17)
\f(x)\Ha{x)
< oo;
here | • | denotes the norm in CtT. If we define, for/, / ' G ^ (18)
: /'> = j
where <. , . > denotes the inner product in Jf, then 3tf becomes a Hilbert space under < . : . > . If X is a separable Borel space, Mf is separable. We shall now select an arbitrary invariant measure class on X and a G-finite measure a in it. a need not be invariant. For each g e G we select a positive Borel function rg which is a version oidajda^9~1). We then define, for fEje Pjs
f = XE*' Ugf={(rgy)1,2f9>
(19) i.e.,
(20)
(uaf)(x) = W r ^ H ( r ^ ) .
The quasi invariance of a implies t h a t everything in sight is well defined. The identities (20) of Chapter V, which are atisfied by t h e rg, now lead to the conclusion t h a t g -> Ug is a homomorphism of G and t h a t (U,P) satisfies (6). The calculations are straightforward. Moreover, as each Ug is invertible and
\UJV = j W\f9\aMx)
= J r.l/Wl2^'-1^) it follows t h a t each Ug is a unitary operator. I n order to assure ourselves t h a t (U,P) is a system of imprimitivity, it remains only to check t h a t for each fjf'eJJ?, g - » (JJgf : / ' ) is a measurable function on G. Now, by theorem 5.10, there exists a positive Borel function r o n ^ x l such t h a t for almost all g, rg(x) = r(g,x) for a-almost all x. Therefore, for almost allgr,
jx{r(g,g-^x)Y'\f(g-^x),r(x)yda(x).
GEOMETRY
208
OF QUANTUM
THEORY
The right side is a Borel function of g since the integrand is Borel onGxX, so t h a t g - > < Ugf : / ' ) is a measurable function. But, as noted above, this is sufficient to assure us t h a t U is a representation. When X is the phase space of a classical mechanical system, G = R1, and the action of 0 on X is given by the dynamical group of the system, the representation U in example 1 was considered by Koopman [1] (a is t h e Liouville measure). Prompted by this, we shall call the system of example 2 (of which the system in example 1 is a special case) a Koopman system of imprimitivity. I n this example, the fact t h a t {U,P) satisfies (6) depends decisively on the identities satisfied by the rg. I t is natural to suspect t h a t one m a y be able t o construct other systems of imprimitivity, by using more general cocycles. This is substantially the case. We shall develop this idea in the following sections.
2. H I L B E R T SPACES O F VECTOR V A L U E D FUNCTIONS The Hilbert space in which the system of imprimitivity of example 2 of Section 1 acts is a space of J f - v a l u e d functions. This situation is very typical. We devote the present section to a detailed study of certain algebras of operators in such Hilbert spaces. Let J f be a separable Hilbert space. We use the notation | • | to denote the norm in J f a n d <. , . > to denote t h e inner product in Jf] For a n y operator A of JT we write \A\ for its norm. Let 8 be the set of all bounded operators in Jf! The unitary group M of J f is a subset of S. We shall consider S as a Borel space b y giving it the smallest Borel structure which makes all the maps A - > (Au,v > (u, v e J f ) Borel. When we speak of Borel maps from or into S, this is the Borel structure we have in mind even when it is not made explicit. Lemma 6.2. S, equipped with the above Borel structure, becomes a standard Borel space and M is a Borel subset of it. The maps A, B —>• AB and A, B->A±B are Borel from SxS into S, A,u->Au is Borel from SxX* into J f , while A - > A* is a Borel automorphism of S. The induced Borel structure on M coincides with the Borel structure associated with the strong topology on M. Proof. Choose an orthonormal basis {en} in Jf. Let T be the space of all complex matrices (a iy ). We m a y identify T canonically with a product of copies of the complex plane. T thus becomes a standard Borel space. For A e S, we define (21)
ait(A) =
(Ae^y.
SYSTEMS
OF
IMPRIMITIVITY
209
Obviously, the Borel structure on S is the smallest one which makes all the maps atj(A -> au(A)) Borel. Thus, if t(A) denotes the element {a{j{A)) of T, t(A -> t(A)) is a one-one Borel m a p of 8 onto a set, say S' contained in T, and the Borel structure on 8 is t h a t one for which i i s a Borel isomorphism with the Borel subspace of T defined by 8'. To prove t h a t 8 is standard, it is sufficient to prove t h a t S' is a Borel subset of T. Only the case dim t>T = oc is of interest. Now, if (aXj) is an element of T, it is easily seen to belong to 8' if and only if there is an integer L such t h a t for any integer N, and any N complex numbers cx. . ., cN with rational real and imaginary parts,
2 J2= I a»ct
< L(|Cl|2+...+|cw|a).
Since the conditions involved are only denumerably m a n y and since each condition defines a Borel subset of T, it follows t h a t S' is a Borel set. This proves t h a t 8 is standard. The equations atj(AB)
=
J^alk(A)akj(B), k
a{j{A*) = atJ(A±B) (Au,eky
aji(A)*9
=
aij(A)±aij(B),
= 2
J
tell us immediately t h a t A, u -> Au, A, B -> AB, and A -> A* are Borel maps with appropriate domains and ranges. Since M = {A : A eS, AA* = A*A
= 1},
it follows from these results t h a t M is a Borel set in 8. Finally, the induced Borel structure on M is the smallest one which makes all the maps A —> Au of M into J T (u e J f ) Borel, and is thus clearly the Borel structure associated with the strong topology on M. Corollary 6.3. Suppose Z is a Borel space and z^> A{z), z-> B(z), and z ~^u(z) are Borel mappings of Z into 8, 8, and JfJ respectively. Then z -> A*(z) and z -> A(z)B(z) are Borel maps of Z into 8 and z - > A(z)u(z) is a Borel map of Z into Jf. Consider now a standard Borel space X, a a-finite measure a on X and the Hilbert space je = ^2(X,Jf,a). For each Borel set fcj, (22)
PE'.f-*XEf
(fe#)
GEOMETRY
210
OF QUANTUM
THEORY
is a projection operator, and P(E - > PE) is a projection valued measure. The main concern of this section is with the algebra of bounded operators of J f which commute with all PE. We consider bounded Borel maps b: (23)
b : x -> 6(a), C = sup |6(a)| < oo
of X into S. If / is any Borel function on X with values in J f such t h a t f \f(x)\2da(x)
< oo, then x -> b(x)f(x)
is also Borel and
f |6(a:)/(z)| 2 da(z) < 0 f
|/(*)| 2 tfa(z).
Thus there exists a unique bounded operator, denoted by b~, such t h a t for each / e Jf for a-almost all x, (24)
(&~/)(s) = &(*)/(*).
6~ is an operator in JfJ its norm \\b~ \\
b~PE =
PEb~.
I t is obvious t h a t if b± is another bounded Borel m a p of X into 8 such t h a t b1(x) = b(x) for a-almost all #, b±~ =b~. We regard the collection of bounded Borel maps from X into 8 as an algebra with involution by defining (a1b1 + a2b2)(x) = a1b1(x) + a2b2(x) (26)
(«i,a 2 complex numbers)
(bc)(x) = b(x)c(x), b*(x) = b(x)*,
for all x e X. Lemma 6.4. For any bounded Borel map b of X into 8, b~ is a bounded operator in 34? commuting with P. Conversely, if B is an operator in M* which commutes with P , there exists a bounded Borel map b of X into 8 such that b~ =B. bx~ =b2~ if and only if bx(x) = b2(x) for a-almost all x. Finally, (&i&2)~ =°i~b2~, (&i*)~ = (&i~)*. In particular, b~ is a unitary operator {projection, normal, etc.) if and only if b(x) is unitary (projection, normal, etc.) in CtC for a-almost all x. Proof. We assume t h a t a ( X ) < o o ; the general case is obtained b y an obvious patching up over sets of finite measure which we leave to the reader. For any u e J f we write, for the purpose of this proof, u for the constant function x - > u of X into J f ; u G J f as a(X) < oo. We observe t h a t if B± and B2 are two bounded operators of & which commute with P, and if i ^ u = B2u for all u e Jf, then B1 = B2. I n fact, for
SYSTEMS
OF
IMPRIMITIVITY
211
U
any Borel set E^X, B1(XEU) = B2(XE )From this it follows t h a t Bxf= B2f whenever / is a Borel function on X (with values in J f ) taking only finitely m a n y values. The general case now follows from the fact t h a t such functions are dense in J^. Thus B1 = B2. This done, let B be a bounded operator in 3^ such t h a t B commutes with P . We m a y assume t h a t ||2?|| < 1 . Let {en} be an orthonormal basis for Jfl Let bn be a Borel m a p of X into JT such t h a t bn = Ben. Now | | J 5 | | < 1 ; so, for a n y N complex numbers cx,-'',cN, as B commutes with P , • • +cNeN)\\2
\\BPE{c1e1+'
= ||x s ( ci £e 1 + . •. + cNBeN)\\2 \\xE(ciei+--+cNeN)\\2,
< i.e., WxsiCiBe!*.
• • + cNBeN)\\2
< ( | C l | 2 + • • • + \cN\2)a(E)
for all E, i.e., f \clb1(x)+
• • • +cNbN(x)\2da(x)
< ( | C l | 2 + • • • +\cN\2)a(E)
for all E. For fixed JV and clf • • •, cN, these inequalities (for all E) imply that |c 1 6i(*)+ • • • +cNbN(x)\2
< | C l | 2 + • • • + \cN\2
for a-almost all x. B y varying N and the c's over complex numbers with rational real and imaginary parts, we can assure ourselves t h a t for some a-null set Z we have t h e inequalities
(27)
|CIM*)+---+%M*)|2
< hl2+--- + M 2
valid for all x E X — Z, all integers JV, and all sets (c1? • • •, cN) of jY-tuples of complex numbers with rational real and imaginary parts. From (27) we conclude t h a t for each xeX — Z there exists an operator b(x) in J f with \b(x)\ < 1 such t h a t (28)
br(x) = b(x)er
for all r. We define b(x) = 0 for xe Z. Equation (28) shows a t once t h a t x - > b(x) is a Borel m a p from X into S such t h a t \b(x)\ < 1 for all x and t h a t for the operator b~ (29)
6~e r = Ber
for all r. From (29) and from the observation made at the beginning of the proof, we deduce t h a t B = b~. To prove t h a t b is unique u p to null sets, let &! be such t h a t 6X~ = 6 ~ . Then for each r, b1(x)er = b(x)er for a-almost all x. This shows a t once t h a t b±(x) = b(x) for a-almost all x.
212
GEOMETRY
OF QUANTUM
THEORY
Routine calculations show t h a t (be)~ = 6 ~ c ~, (& + c ) ~ = 6 ~ + c ~ , and (b*)~ =b~*. If b(x) is unitary for almost all x, then it is obvious t h a t 6~ is unitary. Conversely, if&~6~* = &~*6~ = l, then b(x)b(x)* = b(x)*b(x) = l for almost all x, by the essential one-one nature of the map b -^»b~. Thus b(x) is unitary for almost all x. The assertions involving projection and normal operators are proved similarly. The proof of the lemma is complete. Our next lemma deals with families of operators commuting with P . Lemma 6.5. With notation above, let Y be a standard Borel space and (30)
b:y,x-+b(y,x)
a bounded Borel map of YxX into S. For each y e Y, let b(y,-) denote the map x - > b(y,x) of X into 8. Then the map y ->b(y,-)~ is a bounded Borel map of Y into the Borel space of all bounded operators of <#?. Conversely, if y -^ By is a bounded Borel map of Y into the Borel space of all bounded operators in #? such that By commutes with P for all y, and if v is a a-finite measure on Y, there exists a bounded Borel map b of Y x X into S such that (31)
By =
b(y,.)~
for v-almost all y. b is determined uniquely (v x a)-almost everywhere. If each By is unitary, we can arrange matters so that b(y,x) is unitary for all y, x. Proof. We m a y clearly assume t h a t a and v are finite. Let b be a bounded Borel m a p of 7 x 1 into S, say |6(i/,#)|
which is Borel in y as the integrand is Borel in y, x. We also have || By || < C. Conversely, let y -> By be a bounded Borel m a p of Y into the Borel space of all bounded operators of J f and let By commute with P for all y. We m a y assume t h a t | | ^ | | < 1 for all y. For each y e Y we choose a Borel map by of X into S such t h a t \by(x)\ < 1 for all x and By — by~. Let {en} be an orthonormal basis for J f and for any u e J f let us write u for the constant m a p x -^ u on X. Then, for E^X, {ByXEtm : O
=
(by(x)em,en>da(x),
JE
so t h a t for each Borel set y->\
E^X,
is a Borel function of y. Lemma 5.5 is now applicable and hence we can deduce the existence of complex-valued Borel functions rm n on 7 x 1 such t h a t for each m and n, for v-almost all y, (32)
=
rnfTn(y,x)
SYSTEMS
OF
213
IMPRIMITIVITY
for a-almost all x. There is a set N^ Y of v-measure zero such that for each y e Y — N and all m, n, (32) holds for a-almost all x. Since |6y(a;)| < 1, we deduce from (32) and the Fubini theorem that for (v x a)-almost all y, x,
2 Vn.m(y>*)\2 < °°
(33)
n
for each m, and
(34)
2
2
r
».»( ?>*)«•
Z |ci|2+---+|ck|a
for all Jc and all ^-tuples of complex numbers (cl5 • • •, ck) whose real and imaginary parts are rational. Let Z^Y xX be the Borel set of all (y,x) such that (33) and (34) are satisfied for the indicated ranges of m, k, ct. The complement of Z is obviously v x a-null. Let r* m be defined by (35)
r*m(y,x) =
n,m(y>x)
(y>x)G
z
>
[0 (y,x) $ Z. Then the r* m are Borel functions and the inequalities (33) and (34) are satisfied by the r* m for all y, x. Hence for each y, x there exists an operator b(y,x) such that \b(y,x)\ < 1 and (36)
(b(y,x)em,en> =
r*n(y,x).
Equation (36) shows that (y,x) -> b(y,x) is a Borel map of Y x X into S. Now, rnfm = r*fm (^ x a)-almost everywhere on 7 x 1 , so that we easily deduce from (32) that for ^-almost all y, (by(x)em,eny = (b(yfx)em,eny for a-almost all x, i.e., for y-almost all y, By = b(y,-)~. If b' is another bounded Borel map of 7 x 1 into S satisfying (31) for v-almost all y, it follows from lemma 6.4 and Fubini's theorem that b(y,x) = b'(y,x) for (v x a)-almost all y, x. For the last statement, note that if By is unitary for all y, Z = {(y,x) : b(y,x) is unitary} is a Borel set whose complement is of (v x a)-measure zero. All we have to do is to redefine b(y,x) to be 1 for (y,x) $ Z. The proof of the lemma is complete.
3. FROM COCYCLES TO SYSTEMS OF IMPRIMITIVITY We remarked in our example 2 of Section 1 on the connection between the cocycle identities (35) of Chapter V and the identities (6). The main aim of this section is to show that associated with a cocycle taking values
214
GEOMETRY
OF QUANTUM
THEORY
in a unitary group, there is a system of imprimitivity, and t h a t this system is determined u p to equivalence by the cohomology class of the cocycle. This is precisely the reason for our detailed study of cocycles in the previous chapter. We begin with an elementary lemma. Lemma 6.6. Let G be a Icsc group and Jt? a separable Hilbert space. Suppose that L(g -> Lg) is a mapping of G into the set of bounded operators in J$? such that (i) for all f,ffeJ^, g -> (Lgf,f'y is a measurable function, (ii) Lg is unitary for almost all g, and (hi) Lgig2 = LgiL92 for almost all (91^2) tGxG. Then there exists exactly one representation U of G in £? such that Lg = Ug for almost all g. Proof. Let /x be a left Haar measure on G and ££1 the Banach space of /x-summable functions. F o r / ^ / 2 e J£x we introduce their convolution
(/i*/«)w = | / i ( « ) / > r w ) . fi * / 2 is almost everywhere finite and defines an element of J*?1. Define the operator Lf, for f e J^?1, by setting, for u, v e J^, (37)
(Lfu,v}
= j
f(g)(Lgu,v)dn(g).
f->Lf is linear and \\Lf\\ < | | / | | i . A routine use of the Fubini theorem, coupled with (hi) of the lemma, gives us the equation Lf *f = LflLf2. Moreover, the sum of the ranges of the Lf is dense in ^f; for, if v e 3f and
j /(gKL^vyd^g) = 0 for all u e J4? and / e J?1, then, for each u e Jf, (Lgu,v} = 0 for almost all g. Taking a countable dense set of u's, we obtain the equation Lg*v = 0 for almost all g. Hence v = 0. We now apply a known theorem (Loomis [2], p . 128) and obtain a unique homomorphism g -> Ug of G into the group of invertible operators in J f such t h a t || Ug\\ < 1 for each g, g -> Ug is Borel, and for e a c h / e J?1, f f(gKUgu,v>dp(g)
= Lf
(u, v e
jf).
JG
This implies rather easily t h a t Ug = Lg for almost all g. Ug is therefore unitary for almost all g. Now, the set of g, for which Ug is unitary, is a subgroup. Hence Ug is unitary for all g, i.e., U is a representation. The uniqueness of U is obvious. Let X be a standard Borel G-space. We choose an invariant measure class ^ o n l and fix it throughout the rest of this section. Let a be a measure in this class. We introduce a separable Hilbert space J f and
SYSTEMS
OF
IMPRIMITIVITY
215
2
denote its unitary group by M. Let J^ = S£ (X,$f,a). By a cocycle we mean a (6r,X,M)-cocycle relative to a, i.e., a Borel map (g,x) - »
(38)
PEf =
XEf,
and for almost all g, (39)
(UJ)(x)
{r3(g-i-x)yl2cp(g,g-i-x)f(g-i-z)
=
for each f e Jf? for almost all x. Moreover, the equivalence class of depends only on <€ and y^, where y0 is the cohomology class of
(U,P)
Proof. Equation (38) defines PE and P(E -> PE) is a projection valued measure in Jf. Let us now define, for each g e G and / e <#?, the Borel function Lgf on X by (40)
(Lgf)(x)
{rg(g-'-x)Y^(9,9-1-x)f(g-1-x).
=
Since cp takes values in M, we have: \(Lgf)(x)\2
=
r9(g-i-x)\f(g-i-x)\*,
from which we conclude t h a t Lgf e J#* and | | i ^ / | | = ||/||. Lg is thus an isometry. I t is actually unitary because its range is the whole of Jf. In fact, if/' e J f and (Lgf : / ' ) = 0 for a l l / 6 Jtf.\ a brief calculation shows t h a t cp(g,x)*-f'(g-x) = 0 for almost all x which gives us the r e l a t i o n / ' = 0 . Next,
for almost all (gi,g2,%) e GxGxX. From this we conclude by a routine calculation t h a t Lgi92 = L9iLg2 for almost all gl9 g2 (for the rg we use identities (20) of Chapter V). We now verify t h a t g->Lg is a measurable map. We choose a positive Borel function r on Gx X such t h a t , for almost all g, r(g,x) = rg(x) for almost all x. Then, for/, f'e^f, <£„/:/'>=
f
{r(g,g-^x)Y'\
JG
for almost all g. Since the integrand on the right is a Borel function on GxX, the integral is a Borel function on G and hence g -> (Lgf : / ' > is a measurable function on G. We are therefore in a position to apply lemma 6.6. Thus there exists a unique representation U of G in J f such t h a t Ug = Lg for almost all g.
216
GEOMETRY
OF QUANTUM
THEORY
We calculate from (40) t h a t for each g e G, LgPE = PgELg for all Borel sets E^X. Therefore, if N is a Borel set of measure zero such t h a t Ug = Lg for g e G — N, then for each g e G — N, (41)
U9PEUa-i
= Pt.s
for all Borel sets E. Now, as U is a representation, the set of all g e G for which (41) holds for all Borel ^ i s a subgroup of G. As the complement of this subgroup has measure zero, we conclude t h a t (41) must be valid for all g e G and all Borel sets E. (U,P) is thus a system of imprimitivity based on X. Its uniqueness is obvious. To show t h a t the equivalence class of (U,P) depends only on the measure class of a and the cohomology class of
(42)
is a unitary isomorphism of J?2(X,Jf,a) (39) shows t h a t pE'
onto J*? 2 (X,J^a') and the formula WPEW-1
=
for all E, and for almost all g, Ug' =
WUgW~\
Once again, as the set of all g such t h a t Ug'= WUgW'1 is a subgroup, these equations are valid for all g. Thus (U,P) and (U\Pr) are equivalent. I t remains to examine what happens when we fix a but change 99. Let
b : x -> 6(a)
of X into i f such t h a t 9'(gr,a;) =
b(g-x)(P(gyx)b(x)-1
for almost all (gr,a;) e GxX. For any f e Jf> \f(x)\2
=
\b(x)f(x)\2,
and hence there exists a unitary operator i? of ££2(X,Ctif,<x) such t h a t (cf. Section 2) (44)
B = 6~.
From (44) we obtain (45)
£P£ = P££
SYSTEMS
OF IMPRIMITIVITY 1
for all Borel sets E. Next we compute BUgB' . almost all g e G, (BUgB-if)(x)
=
217
We have, from (39), for
{rg{g-^x)Y'%(x)9(g,g^.x)b(g-^x)-^(g-^x)
=
{rg{g-^x)y'Y(g,g-i.x)f{g^.x)
= (U„'f)(x) for each / e J4? for almost all x. Therefore (46)
Ug' =
BU.B-1
for almost all g. Using our usual arguments, we conclude t h a t (46) is valid for all g. Equations (46) and (45) show t h a t (U,P) and (U',Pf) are equivalent. This completes the proof of the theorem. Corollary 6.8. If
for almost all x, then for each g e G and each f (UJ)(x)
=
{rt{g-1-x)}»!>v{g,g-i-x)f(g-i-x)
for almost all x. I n this case, g -> Lg itself is a homomorphism so t h a t Lg= Ug for all g. Given
4. P R O J E C T I O N V A L U E D M E A S U R E S I n this book we have come across projection valued measures from time to time, but such uses as we have made of them have been quite superficial. The aim of this section is to describe some of the deeper aspects of the theory of projection valued measures. These results are all known and form the Hahn-Hellinger multiplicity theory. We recommend the reader to the accounts of Stone [2], Nakano [1], and Halmos [2]. We shall confine ourselves to a concise description of the main features of this theory. The main problem in this theory is t h a t of determining when two projection valued measures are equivalent. More precisely, let X be a standard Borel space, J f * a separable Hilbert space, and P{ a projection valued measure in MP1 based on X. We say P1 and P 2 are equivalent, P1~P2 in symbols, when there is a unitary isomorphism W of J^1 onto Jf72 such t h a t PE2=WPE1W-1 for all E. The main problem is, given X, to construct canonical forms for the projection valued measures based onl. We begin by choosing a a-finite measure a on X and a separable Hilbert
218
GEOMETRY
OF QUANTUM
space Jf! Let Jtfp = J?2(X,Jjr,a) E -> PE, where (47)
THEORY
and P be the projection valued measure
PEf=XEf
(/e^T).
We write P=P(Jf,a). The first result of the theory is t h a t P(Jf,a)~ P ( J f ' , a ' ) if and only if (i) a and a' define the same measure class, and (ii) dim J f = dim Jf'. If these conditions are satisfied, if J f = Jf', and if a is a positive Borel function which is a version of da /da, the m a p
W:f->a~ll2f is a unitary isomorphism of ^2(X,Jf,a) onto 3?2(X,Cf,a) such t h a t P(Xcc') = WP(Jf,a)W-\ This result motivates us to introduce the following definition. A projection valued measure P (acting in a separable Hilbert space) is said to be homogeneous if it is equivalent to P(Jf^a) for suitable C/f and a. The definition (47) of P(Jf,'<x) shows t h a t for a Borel set E^X, a(E) = 0 if and only if PE = 0. The measure class of a is thus determined by P in a very direct fashion. I t is called the measure class of P. The dimension of J f is said to be the multiplicity of P. Given any measure class ^ on X and any integer n (1
Jf = © JfB>
be the direct sum of the ^ > a n , and for E^X (49)
let PE be defined by
PE : ( / . . A , / * - • •) -+ ( P f ' ' . / . . ^ * ' / i . " ')
")*CT-finitemeasures a and j3 on X are called mutually singular if we can write X = A U B, where A and B are Borel sets with i n 5 = 0 , f}(A ) = a(B) = 0.
SYSTEMS
OF
IMPRIMITIVITY
219
(when a n = 0 for some n, J^n = 0 ) . Then P is a projection valued measure. We write P = P({J^}, {an}). The main theorem of spectral multiplicity theoryasserts t h a t every projection valued measure on X in a separable Hilbert space is equivalent to P({J^n}, {an}) for a suitable sequence {an}. If {an} and {an'} are two sequences of mutually singular measures on X, the theorem further asserts t h a t P(VQ,
{«„» - P{{JQ,
K'»
if and only if for each n, an and an' are in the same measure class. Given P, we therefore have for each n = oo, 1,2,••• a canonically determined measure class, say ^fn. ^n is called the measure class corresponding to the multiplicity n. For a given n, c€n m a y be 0. P is homogeneous if and only if c€n = 0 for all n except a single value nQ. The construction of the ^ n from P is delicate. We do not need this for our purposes. On the other hand, consider the set of all cr-finite measures a with the property t h a t a(E) = 0 if and only if E is ^ n - n u l l for all n (1
5. FROM SYSTEMS OF IMPRIMITIVITY TO COCYCLES We now examine the problem converse to t h a t studied in Section 3. We resume our concern with a standard Borel Cr-space X and a system of imprimitivity (U,P) based on X, acting in 3f. These will be fixed throughout this section. The main result is t h a t if P is homogeneous, one can construct a cocycle 9? relative to the measure class ^ of P such t h a t (U,P) is equivalent to the system of imprimitivity associated with ^ and (p. We shall use the notations and results described in Section 4 concerning projection valued measures. I n particular, for each n(co, 1, 2, • • •), J^n is a fixed separable Hilbert space of dimension n. We write Mn for the unitary group of J ^ . The norm and inner product in Ctfn are denoted by |. | and <.,.>, while the norm and inner product in 2tf and , # ^ a [a a a-finite measure on X) are denoted by ||. || and < . : . > . Lemma 6.9. With the notation described above, the measure classes ^n corresponding to the various multiplicities n are all invariant under the action of G. In particular, the measure class of P is invariant. Proof. Choose measures an e ^n. Then, P~P({Jfn}, {an}). We must show t h a t each ^n is invariant under G. Let g e G be arbitrary. Let us
220
GEOMETRY
OF QUANTUM
THEORY
write QE = Pg.E. Then Q(E->QE) is a projection valued measure in 3tf. The equation UgPBUg~i = Pg.E shows t h a t Q~P. B u t it is obvious t h a t Q ^ P ( { J ^ } , { a / } ) (via the m a p f-^f9) and t h a t the an9 are mutually singular. Hence for each n, an and 9 an define the same measure class. This proves the lemma. We recall t h a t a measure class ? on 1 is said to be ergodic if a n y Borel function / , with the property t h a t f9 =f ^-almost everywhere for each g e G, is constant ^ - a l m o s t everywhere. Lemma 6.10. If the measure class ^ of P is ergodic, then P is homogeneous. If (U,P) is irreducible, <€ is ergodic and so P is homogeneous. Proof. Choose mutually singular finite measures an (n = oo, 1, 2,• • •) such t h a t we then have P~P({3Tn},{an}). Let %\ be the measure class defined by an. Since the am are mutually singular, we can find mutually disjoint Borel sets En (n = cc, 1, 2,- • •) such t h a t X = [J1^n^o0 En and an(X — En) = 0 for all n. Suppose t h a t the measure class ^ of P is ergodic, b u t t h a t for two distinct values of n, ^ n # 0 . Let an be distinct real numbers and l e t / be the Borel function which takes the value an on En. For each n, f is a constant ^ n - a l m o s t everywhere. Since ^ n / 0 implies an(En)y^0, En is not ^ - n u l l for a n y n for which c€n is nonzero. Thus / is not a constant ^-almost everywhere. B u t (by lemma 6.9) as each ^ \ is invariant, f9=f, ^-almost everywhere for each g e G. This is a contradiction. We m u s t therefore have ^ n = 0 for all b u t a single value of n, i.e., P is homogeneous. The second statement would follow if we show t h a t ^ is ergodic whenever (U,P) is irreducible. Suppose now t h a t (U,P) is irreducible b u t ^ is not ergodic. Then there exists a real Borel f u n c t i o n / such t h a t / i s not a constant ^-almost everywhere but, for each geG, fg=f, ^-almost everywhere. Let c be a real number such t h a t , if E0 = {x :f(x)
y*^Z{y)
SYSTEMS
OF IMPRIMITIVITY
221
is a one-one correspondence between the set of all cohomology classes of (G,X\Mn)-cocycles relative to %?P and the set of all equivalence classes of systems of imprimitivity of the form (U\Pn,a). Proof. There exists a unitary isomorphism W of J4? onto Jfn>a such t h a t pn,a
WPEW-^
=
for all Borel sets E. We set Ug'=WUgW~\ Then (U,P)~ (U\Pn>a). na Consider now a system (U',P ' ). Let us choose, for each g in G, a positive Borel function rg such t h a t rg is a version of da/da^9'^. For simplicity of notation we write je'=J^nta, P' = Pn>a. Let (V,P') be the Koopman system of imprimitivity, i.e., for each g e G a n d / e <#", (51)
(Vgf)(x)
/(flT 1 -*).
= {rg{g-^x)Y*
Write (52)
Wg=
V
1
^ ' .
Since both (Uf,P') and (V,P') are systems of imprimitivity, it follows t h a t Wg commutes with PE for all E. Further, g -> Wg is a Borel map of G into the unitary group of 3tf". We now use lemmas 6.4 and 6.5. We select, for each g e G, a Borel m a p wg of X onto Mn such t h a t (53)
Wg = wg~,
and a Borel m a p 99' of G x X into Mn, (54)
such t h a t for almost all g, (55) For gr1; gr2 (56)
W„ = e
@>
we
9'{g,-)~.
obtain from (52), Wgig2 = ( J V ^ g i F g 2 ) J F 9 2 .
A quick calculation shows t h a t
Vg^WgiVg2 = K T 1 ' ) - ,
(57)
where, as usual, for x e X, wggf1}(x)
=
wgi{grx).
From (56) and (57) and lemma 6.4, it follows t h a t for each (gvg2)eGx (58)
w9ig2(x)
=
wgi(g2'X)w92(x)
for almost all x. Then (59)
9'(gL>g2-x)
G,
222
GEOMETRY
OF QUANTUM
THEORY
for almost all (glyg2,x) e GxGx X. This proves t h a t cp is a (G,X,Mn)cocycle relative to a. Moreover, as Ug' =VgWg, we have, for almost all g, Ug = Vg
{rg(g-1-x)}^cPf(g,g^-x)f(g-^x)
(EV/)(s) =
for / e J f 7 ' . I n other words, (U',Pn'a) is the system of imprimitivity associated with cp' and a. If 9 / is another cocycle satisfying (60), then, for almost all g,
(61)
WU/W-1
= Ug\
for all g, E. There exists (lemma 6.4) a Borel map w, (62) w : x - > w(x) of X into Mn such t h a t W = w~. But then the equation WUg"W~1= Ugf, together with the relations between
=
cPf(g,g-1-x),
i.e., w{g-x)^p"{g,x)w{x)-1 for almost all (g,x) eGxX. proof of the theorem.
=
This shows t h a t cp'~cp". This completes the
Remark. I t is easy to see t h a t if the cocycle cp' is a coboundary, the corresponding systems of imprimitivity are equivalent to the K o o p m a n systems. Thus, the examples of Section 5 in Chapter V give rise to systems (U,P) which are not equivalent to the Koopman systems.
6. T R A N S I T I V E SYSTEMS I n this section X continues to be a standard Borel (x-space and (U,P) a system of imprimitivity based on X. We shall completely describe all the transitive systems. Our success is due to the detailed knowledge of cocycles on transitive G-spaces which was accumulated in Section 5 of Chapter V. We recall t h a t a measure class ^ is said to be transitive if there exists x0 e X such t h a t X — G-x0 is a ^-null set. I n this case the
SYSTEMS
OF
IMPRIMITIVITY
223
orbit G • x0 is uniquely determined by <% and we shall say t h a t ^ lives on G-x0. A system of imprimitivity (U,P) is said to be transitive if the measure class ^ of P is such; if X ' is the orbit on which ^ lives, we say t h a t the system lives on X'. Note t h a t if (U9P) is transitive, the measure class of P is ergodic, and hence P is homogeneous (lemma 6.10). If xeX and X' = Gx, X' is an invariant Borel set and is hence a standard transitive Borel 6r-space in its own right. If a is a a-finite quasi-invariant measure on X', the measure a : E -> a'(E C\ X') defines a measure class on X which is invariant and which lives on X'. Obviously it is the only measure class living on X' with this property. Let x0 e X and X' = G-x0. Suppose t h a t GXQ = G0 is the stability subgroup at xQ. Let m be a representation of G0 in the Hilbert space J ^ (which we have chosen once for all) of dimension n (l
224
GEOMETRY
OF QUANTUM
THEORY
obviously. Theorem 6.11 now tells us that there exists some n and a (6r,X,ifn)-eoeyele 9 relative to a such that (U,P) is equivalent to the system associated with
(63)
Ef=XEf,
(ugf)(x) = K(r W'Wr W(r **)• We use the results and notation of Section 5, Chapter V. We select a Borel map d(g -> d(g)) of G into Mn such that d{gh)—d{g)m{h) for all (g,h)eGxG0, d(e) = l, and
t:x-^
t(x)
of X into the Borel space Sn of all operators in Jfn such that |£(a;)| < ||T|| for all x and (65)
T = r
If we use the fact that 2T commutes with all Ug, we obtain, after a brief calculation, the identity (66)
x
• x) = t(x)
for each g for almost all x, i.e., for each g e G (67) for almost all xeX.
9>(gr,a?)*(a;) = t{g-x)
ngg'Wigtf)
for almost all (<7,#') EGXG. Substituting the expression for
t°(g) = d(g)rd(g)-i
SYSTEMS
OF
IMPRIMITIVITY
225
for almost all g. Replacing g by gh (h e G0) and simplifying, we obtain (69)
mihjrmih)-1
= r.
This shows t h a t r is in the commuting ring of m. On the other hand, if we start from a r in the commuting ring of m, the m a p g -> d(g)rd(g~1) is a bounded Borel m a p of X into Sn which is constant on the left 6r 0 -cosets. Hence there is a bounded Borel map t from X into Sn such t h a t t° is given by (68) for all g. Routine calculation leads us from (68) to (67) to (66). At this stage, we know t h a t T = t~ is in the commuting ring of (U,P). The correspondence T ^± r is obviously one-one and is an algebra isomorphism as is easily seen from the formulas (68) and (65) connecting T and T. Moreover, as each d(g) is unitary, (68) also implies t h a t (70)
P(g)* =
d(g)r*d(g)-\
so t h a t T +t r is adjoint-preserving. Since the commuting ring of a representation (or a system of imprimitivity) determines completely the circumstances under which it is irreducible or a direct sum of irreducibles, the proof of the theorem is complete. Corollary 6.13. Let X be a standard transitive Borel G-space, x0 e X, and G0 the stability group at x0. Then every system of imprimitivity based on X is homogeneous and its measure class is the unique invariant measure class on X. There is a canonical one-one correspondence between the equivalence classes of systems based on X and equivalence classes of representations of G0. This correspondence preserves irreducibility as well as the property of being a direct sum of irreducibles. Remark. Theorem 6.12 gives a complete analysis of the transitive homogeneous systems of imprimitivity based on X. I t is natural to ask what happens when we give up the assumption of transitivity. According to lemma 6.10 if the measure class ^ of P is ergodic, then P is homogeneous. Theorem 6.11 then tells us t h a t we have only to study the cohomology classes relative to %\ However, the examples of Section 5 in Chapter V show t h a t the cohomology classes relative to an ergodic invariant measure class are incredibly more complicated t h a n in the transitive case. In view of this it has not been possible to carry out the analysis of ergodic systems of imprimitivity further t h a n t h a t of theorem 6.11. We note here t h a t if there is any ergodic invariant measure class ^ on X, then one can construct irreducible systems of imprimitivity (U,P) based on X such t h a t ^ is the measure class of P . I n fact, let a be a afinite measure in ^ and let cp be any ( ( r ^ l f ^ - c o c y c l e relative to # . Let 3tf = $f\ta and let (U,P) be the system of imprimitivity associated
220
GEOMETRY
OF QUANTUM
THEORY
with op and a. We m a y take Jf± to be the complex number space so t h a t Jf = J?2(X,a). We claim t h a t (U,P) is actually irreducible. I n fact, let T be an operator commuting with (U,P). By our lemma 9.4 there exists a complex valued, bounded Borel function t such t h a t
(Tf)(x) = t(x)f(x) for all / e 3tf. The condition t h a t T commutes with all Ug gives us (67). But, as t(x) is a complex number, it commutes with
a~(E)
=
a(G-E).
Now the fact t h a t D meets each orbit exactly once implies t h a t , for disjoint E, the sets GE are also disjoint. Hence, a~, as defined by (71), is a measure on the Borel space defined by D and a~(D) = 1. We claim t h a t for a n y Borel set E^D, a~(E) is either 0 or 1. If this were not so, we can choose a Borel set E^D such t h a t 0
SYSTEMS
OF
IMPRIMITIVITY
227
D — En whenever necessary, we may assume that a~(En) = l for all n. Put (72)
E0 = f| En; n=l
a~(E0) = l. Since the sequence {En} separates the points of D, EQ must consist of a single point, say x0. But then a(G-x0) = l, showing that a is transitive. A variation of the argument given just now leads to: Lemma 6.15. Suppose that there exists a sequence {fn} of invariant real valued Borel functions on X which separate the orbits of X, i.e., two points x,y of X lie on the same orbit if and only iffn(x) =fn(y) for all n. Then every ergodic invariant measure class on X is transitive. Proof. Let a be a measure such that a(X) = 1 and let a be quasi invariant and ergodic. Let &' be the smallest Borel structure on X which makes all t h e / n Borel. Every set in £%' is a Borel subset of X which is also invariant. Hence each such set has a-measure either 0 or 1. Let {En} be a sequence of sets which generates &' such that a(En) = 1 for all n. If we put (73)
E0 = H En,
then a(E0) = 1. Now the class of Borel sets which either contain E0 or are disjoint from it is a a-algebra containing all the En. Hence every set in 08' has this property. This implies that E0 is a single orbit; for, if x0 e E0 and X' = G'X0i the fact that t h e / n separate the orbits implies that (74)
X' = C){x :/»(*) =/„(*«>)}, n
and consequently that X'e&'\ X' = E0.
since X' C\Eo^0,
EQc:X', so that
Corollary 6.16. If 0 is compact, then every ergodic invariant measure class on X is transitive. Proof. Using theorem 5.7 it is easy to come down to the case when X is a compact metric space and G acts continuously on X. We shall verify that the condition of lemma 6.15 is satisfied. By the Stone-Weierstrass theorem, C(X) is separable. Let {un} be a countable set dense in C(X), and for u e C{X) let (75)
u(x) =
u(g-x)dp(g),
where /x is the (normalized) Haar measure on G. u is continuous on X. Clearly each u is invariant. We claim that {un} separates the orbits in X.
228
GEOMETRY
OF QUANTUM
THEORY
Let x, y e X such t h a t un(x) = un(y) for all n. Since {un} is dense in C(X), we deduce from (75) t h a t u(x) — u(y) for all u e C(X). Suppose now x and y were not in the same orbit. Then Gx and G-y are closed disjoint subsets of X, and hence there is a u e C(X) with 0
7. E X A M P L E S A N D R E M A R K S We shall now discuss a number of examples. The simplest example of a transitive space is G itself. If (U,P) is a system based on G, it is transitive and its measure class is the invariant measure class of G. We now apply the results of Section 6. The stability group at e e G is {e} whose only representations are trivial. Any trivial representation is a direct sum of trivial one-dimensional representations. We thus obtain Theorem 6.17. Let G act on itself by left translations. Let 3^ — where /x is a left Haar measure on G, and let P and U be defined by:
^2{^),
(UJ)(x) = f(g-ix). Then (U,P) is an irreducible system of imprimitivity based on G. Any irreducible system based on G is equivalent to this one. An arbitrary system of imprimitivity is a direct sum of irreducible systems. Rotations in the Plane. Let X = C, the complex plane, and let G be the multiplicative group of complex numbers of modulus 1. We put, for z e G and x e C, z-x — zx. X is a Borel 6r-space. For each
d>0,
Ed = {x : \x\ = d} is an orbit and these are all the orbits. Since G is compact, all ergodic measure classes are transitive. For d>0, the stability groups associated with Ed are trivial and hence, corresponding to each d>0, there exists, u p to equivalence, exactly one irreducible system living on Ed. We shall now describe this. Note t h a t Ex has an invariant measure, say a, such t h a t a(E1) = l. If we define the measure ad by f(x)dad(x) J Ed
=
f(dx)da(x), JE±
SYSTEMS
OF
IMPRIMITIVITY
229
then ad(Ed) = l and it is the unique invariant measure living on Ed with this property. Let J^d be the Hilbert space if?2(crd). Then, if we define (Ud,Pd) by PB*
(U/f)(x) d
= XEL =
f(z~ix),
d
(U ,P ) is the unique irreducible system living on Ed. When d = 0, Eo = {0} and the stability group is G itself. I n this case, P is trivial and U can be an arbitrary representation of G. If (U,P) is to be irreducible, U will have to be a character of G. The example discussed just now can be generalized. Let us assume t h a t X — Rn and G=SO(n), the group of orientation preserving rotations. The orbits in X are the spheres Sd of radius d > 0 and the origin 0. Let us choose d > 0 and consider the point xd = ( 0 , 0 , . . . , 0 , d ) . Then the stability subgroup xd is canonically isomorphic to SO(n—l). Since G is compact, we know from corollary 6.16 that there are no ergodic systems other t h a n the transitive ones. For each d>0, there are denumerably m a n y inequivalent irreducible transitive systems living on Sd; these correspond one-one to the equivalence classes of irreducible representations of 80(n — 1). However, when we proceed to describe explicitly the irreducible systems living on SXi for example, we run into a difficulty. The theory developed in Section 5 of Chapter V and in the present chapter directs us to start with an irreducible representation, say m, of SO(n — l) and construct a strict (GfSx)-cocycle which defines the representation m at the point (0, • • •, 0, 1). Such cocycles can be constructed, of course, but they depend on the choice of a section for the left coset space SO(n)ISO(n—l). Since canonical sections do not exist, our theory does not yield any simple geometric description of the unique irreducible system of imprimitivity living on S± which corresponds to m. This difficulty, although of no theoretical importance, can be quite irritating when we want to make any calculation. We shall next proceed to discuss a geometric method of writing down systems of imprimitivity which does not rely on the construction of sections. This method is of course not very general, but it covers most of the cases which occur in physics. Special Descriptions. Let I be a standard transitive Borel (r-space, x0 e X. Let G0 be the stability group at x0. We assume t h a t X possesses an invariant a-finite measure, say a. Let m be any representation of G0 in a separable Hilbert space Jfi We assume that there exists a separable Hilbert
230
GEOMETRY
OF QUANTUM
space Jf and a Borel homomorphism operators of C/f' such that
THEORY
m' of G into the group of invertible
(a)
Jf is a closed linear manifold of JT',
(b)
m'(h)u
J
(77)
= m(h)u
J
>
for all h e G0
and
u e Crff.
Let | • | and <(.,. > denote the norm and inner product in Jf'. Note t h a t the operators m'(g) for g e G0 need not be unitary. Let us consider now any point XEX. Since X is transitive, there exists a g e G such t h a t g-x0 = x. We now define the closed linear manifold Jf^c; Jf' by: (78)
Xx =
ro'(0)[Jf].
Equation (78) defines CtiTx without ambiguity. We now introduce an inner product < . , . > x in 3JTX, by choosing g e G such t h a t g • x0 = x and putting (79)
(u,v}x
= (m'ig-^m'ig-^v}
(u9veJTx).
Once again, (79) is well defined. The fact t h a t C/fx and <.>•>* defined and t h a t m is a homomorphism of G implies t h a t (80)
Xg.x
(81)
=
are
we
ll
m\g\Xx\
(m'(g)u,m'(g)v)g.x
=
for all (g,x)eGxX. I n particular, m'(g) is a unitary isomorphism of Jf^ onto J^.^. We now introduce the vector space y of a-equivalence classes of Borel maps f: (82)
f:X-+X",
such t h a t (83)
/(x)eJTx
for all x e X. If / , / ' e y (as usual we make no distinction between functions satisfying (83) and the equivalence classes which they define), we observe t h a t x —> (f(x),f'(x)}x is a Borel function. To prove this, we choose a section c, i.e., a Borel m a p c of X into 6r, (84)
c :x->
c(x)
c(x)-x0
— x
such t h a t c(x0) = e and (85) for all x. Then, (86)
(*),/'(*)>* = <m'(c(s)"*)/ (*),m'(c(z) - i)/'(*)>
and, as the m a p g -> m'(^) is Borel from G into the Borel space of operators on Jf', the Borelian nature of x->{f(x)J'(x))x follows from corollary 6.3.
SYSTEMS
OF IMPRIMITIVITY
231
For fe'f", we p u t ll/ll 2 = f Jx
(87)
Let
•** = {/: I/II2 < ^ } -
(88) For (89)
fJ'eJf,
= £(*),/»>xd«(ar)
is clearly finite. < . : . > is an inner product for Jf. We define for g e G and E^X, P
(90)
= XEL = m'(g)f(g-i.x),
EI
(Ugf)(x)
for all / e V. I t is clear t h a t PEf and Ugf e ^ t h a t PE and £7g are linear transformations of y^ and t h a t both J7ff and P £ leave Jf7 invariant; it is also clear t h a t UgPE = PgEUg for all g e (x, and Borel sets E^X. Theorem 6.18. Jf7 is a separable Hilbert space, and (U,P), as defined by (90), is a system of imprimitivity in Jf'. Moreover, the equivalence class of (U.P) is the one induced by the representation m of G0. Proof. We shall use the section c : x -> c(x) satisfying (85) for all x e X. With the help of c we m a p 3tf* on the Hilbert space S£2{X^a). We define for/e^J/by (91)
(Jf)(x)
=
m'(c(x)~i)f(x).
Since x->f(x), g-^m'(g), and x^c{x)~x are Borel maps with appropriate domains and ranges, it follows t h a t Jf is a Borel m a p of X into Jf. Moreover, (86) implies t h a t (92)
=
<(Jf)(x),(Jf)(x)>.
J is also linear and one-one. Therefore we conclude t h a t J is a linear isomorphism of J4? onto ^f2(X,J^,a) which preserves the inner products, i.e.,
(/£•#)•
This proves t h a t J f is a separable Hilbert space and t h a t J is a unitary isomorphism of 34? onto ^2(X,Jf,a). We then p u t Vg (93)
=JUgJ'\
9
QE
9
=
J *EJ
232
GEOMETRY
OF QUANTUM
THEORY
(V,Q) is a system of imprimitivity based on X, acting in j£?2(X,Jf^a). Using (90) and (91), an easy computation reveals t h a t f o r / 0 e j£? 2 (X,J^a), (94)
QEf0 = XEU
and (95)
(Vgf0)(x)
=
m'(c(x)-igc(g-i.x))f0(g-i-x).
Now, the element c(g • x) ~ xgc(x) lies in G0 because it sends x0 to . (g. x) = x0. Therefore, if we define (96)
c(g-x)~1
m(c(g-x)-1gc(x)),
then g, x ->
Therefore
(Vgf0)(x)
=
Equations (97) a n d (98) show t h a t (V,Q) is the system associated with m. Since (U,P)~(V,Q)} the proof of the theorem is complete. We shall discuss a number of situations where the conditions (77) connecting G0, G, and m are satisfied for every irreducible representation m of Go. (i) G is compact. I n this case, let m be an irreducible representation of G0. If we decompose the representation of G, induced by m, into irreducible components, and select any irreducible, m! say, which occurs in the decomposition, it follows quite readily from the Frobenius reciprocity theorem (cf. Weil [1]) t h a t the restriction of m! to G0 contains m as a subrepresentation. (ii) G is a connected, simply connected complex semisimple Lie group, GQ compact. Here, G is known to be unimodular so t h a t X admits an invariant Borel measure. From the general theory of the semisimple groups (cf. Helgason [1]), we know t h a t there exists a maximal compact subgroup K containing G0. Let m be an irreducible (finite dimensional) representation of G0 acting in Jf! By (i) above, there exists a finite dimensional Hilbert space J f ' containing J f and an irreducible representation m! of K in J T ' such t h a t m'(h) = m(h) for all h e G0. Now, it is well known (cf. Weyl [2], p. 267) t h a t there exists a homomorphism m" of G into t h e group of invertible operators of X' itself such t h a t m"(k) = m'(k) for ke K, and t h a t m" is even unique if we require t h a t the m a p g -> m"(g) is complex analytic on G. Thus the conditions of theorem 6.18 are satisfied
SYSTEMS
OF
IMPRIMITIVITY
233
iii this case. In particular, if G0 itself is maximal compact in G, the systems of imprimitivity based on GjGQ can be described in the geometric form of theorem 6.18. We shall do this in Chapter I X , when G = SL(2,C) and G0 = SU(2,€), in connection with the equations for the Dirac electron. We thus have: Theorem 6.19. Let G be a connected, simply connected complex semisimple Lie group, and G0 a maximal compact subgroup of G. Let a be an invariant Borel measure on X = GjG0. Let m be an irreducible representation of G0 in a Hilbert space Jf, and let m'(g -> m'{g)) be the unique complex-analytic homomorphism of G into the group of invertible linear endomorphisms of J>f such that m'(h) = m(h) for all h e G0. Then there exists exactly one map x —> <•>•>* °f X into the set of positive definite inner products on J f x J f such that < . , . }XQ = < . , . } and (u,v)x
=
(m'(g)u,m'(g)v\.x
for all (g,x) e G xX. such that
If J#* is the Hilbert space of Borel maps f of X into Jf
(99)
|!/1| 2 = j
(f{x),f(x))xda(x)
< oo,
and we define P and U by (100) then (U,P) by m.
PsfJ = XXef, J (UJ)(x)
=
m'(g)f(g-i-x),
is a system of imprimitivity
equivalent to the system
induced
Representations in Vector Bundles. The reader might have noticed t h a t the essential point of the proof of theorem 6.18 is contained in the fact t h a t the family of subspaces x -> C/fx transform " covariantly " under the action of G. I t is natural to suspect t h a t a more general formulation of theorem 6.18 can be obtained in the context of such covariant families. This is indeed so, and the natural setup for formulating this is that of a vector bundle, more precisely, a Hilbert space bundle. Let X and B be both standard Borel G-spaces and let X be transitive. We assume t h a t there is a Borel map (101)
TT:B->X
of B onto X with the property t h a t for each g e G, the diagram B-^->
(102)
J X
—
B
234
GEOMETRY
OF QUANTUM
THEORY
is commutative. If we write, for each g e G, D(g) for the automorphism of B induced by g, and x —> g • x for the automorphism of X induced by g, (102) translates to (103)
*(/%)(&)) = g-TT{b)
for each be B and each g e G. We shall assume also for simplicity that X admits an invariant a-finite measure, say a. We shall say that B is a G-Hilbert space bundle if, for each x e X, (104)
Bx = w-1({*})
is a separable Hilbert space whose natural Borel structure is the one induced on Bx by B and if, for each g e G and x e X, (105)
b -> D(g)b
(b e Bx)
is a unitary isomorphism of Bx onto Bg.x. If we write | - j ^ and <.,. ) a . for the norm and inner product on 2^, (105) means (106)
<6,6'>x =
(b, b' e Bx).
In particular, if Gx is the stability subgroup at x, D(g) leaves Bx invariant for g eGx and g -> D(g) defines a representation of Gx in Bx. We shall now consider sections of the bundle. A section of B is a Borel map/, (107)
f'x^f{x)
of X into 2? such that f(x) e Bx for all x e X. To handle the measurability problems involving these sections, we transform them into Borel maps of X into a single Hilbert space. Choose a point x0 e X and a Borel map c from X into G such that c(#0) = e and C\Xj '
XQ
==
3/
for all x e X. Since the map g,b -> D(#)& of Gx B into 2? is Borel, it follows quite simply that for any section f of B, x-> D(c(x))~1f{x) is a Borel map of X into I?Xo. Conversely, if x ~>f0(x) is a Borel map of X into l?Xo, then x -> Z>(c(#))/0(#) is a section of the bundle B. We denote by ^ the vector space of a-equivalence classes of sections. For any x e X, and any section /, (*),/(*)>* =
This shows that x-> |/(#)|£ is a Borel function on X. We shall say t h a t / is a square integrable section if
(108)
ll/l2 = J |/(*)|§ia(*) < 00.
SYSTEMS
OF
IMPRIMITIVITY
235
Yitff denotes the subset of y consisting of square integrable sections, then £F is a pre-Hilbert space under the inner product (109)
: /'> = £
The isomorphism, which associates with each section / , the m a p x —> D(c(x))~1f(x), induces a unitary isomorphism W of Jtf* onto ^2(X,BXQ,a). £F is thus a separable Hilbert space. For any Borel set Es^X we write (HO)
PE!=XEJ
(/6JT).
For each g e G, we define the operator Ug by (111)
(UJ)(x)
= D(g)f(g-i-x)
(fetf).
Computations similar to those of theorem 6.18 then give rise to the following theorem. Theorem 6.20. (U,P) is a system of imprimitivity. / / D° is the representation of the stability subgroup GXQ (of G at x0) in the Hilbert space BXQ, then (U,P) is equivalent to the system induced by D°. As an interesting example of representations in such bundles we mention the following. Let X be a O00 Riemannian manifold, and let each g E G be an isometry of X. We assume t h a t G is a Lie group acting transitively on X in such a fashion t h a t g, x —> g x is a O^-map of G x X into X. Let B be the tangent bundle^ of X. For each g, let D(g) be the differential of the isometry x-^gx. Then all the conditions of the preceding theorem are satisfied. The system of imprimitivity thus acts on the Hilbert space of square integrable vector fields and is equivalent to the system induced by the representation of the stability group G0 at x0 e X defined in the tangent s p a c e | to X at x0. I n Chapter I X , we shall see how the representations associated with the photon can be formulated, in the spirit of theorem 6.20, as acting in the Hilbert space of sections of a certain vector bundle on the light cone, leading thereby in a very simple way to the equations of Maxwell. Ergodic Intransitive Systems. The last example is simply t h a t of a standard Borel 6r-space which has ergodic nontransitive (invariant) measure classes. We take real numbers Al5 A2,- • • which are rationally independent and take X to be a torus having its dimension equal to the number of A's (finite or infinite). Let 6 r = R \ and for teR1 and (112) f Complexified.
* = (£i,£2,"-)
(\Q = 1 for all j )
236
GEOMETRY
OF QUANTUM
THEORY
in X, let (113)
(eUKiU,euH2,--).
tx =
Then, it is classical (Kronecker) that each orbit is dense and that the Haar measure on X is invariant and ergodic but not transitive.
8. SEMIDIRECT PRODUCTS We shall now use the theory of systems of imprimitivity to obtain the description of all irreducible representations of an interesting class of lcsc groups which are neither compact nor abelian. This class of groups includes the inhomogeneous Lorentz group and the group of Euclidean motions as members. For the special case of the Lorentz group, the analysis which we shall give now was carried out by Wigner in his 1939 paper. The general study is due to Mackey and comes out as a simple consequence of the work in Section 6. We begin by introducing the concept of semidirect products. Let A and H be two groups and for each h e H let (114)
th:a-^h[a]
be an automorphism of the group A. We shall assume that h -> th is a homomorphism of H into the group of automorphisms of A so that h[a] — a for all a e A if h = eH, the identity of H, (115)
thlh2 =thlth2
(h^eH).
Equation (115) converts A into an H-space. We shall now make G=
HxA
into a group by defining the multiplication in G by (116)
(h,a)(h',af) = {Kh\ath[a'}).
It is easy to verify that this definition converts G into a group with e = (eH,eA) as its identity (eA being the identity of A). Further, (117)
(M"1 =
(h-^h-^a-1]).
G is called the semidirect product of H and A relative to t and we put (118)
G=
HxtA.
When no confusion can arise as to what t is, we omit it, and write G = Hx'A. But it must be remembered that (i) A and H play very unlike roles in the formation of G, in contrast to the construction of direct products, and (ii) G can be formed only after specifying t, i.e., specifying how / / " a c t s " on A.
SYSTEMS
OF
IMPRIMITIVITY
237
Suppose now that A and H are lcsc groups and that the map (119)
h,a->h[a]
is continuous from H xA into A. Then, if we equip G with the product topology, it follows quite easily from (116) and (117) that G becomes a topological group. Clearly, G is a lcsc group also. We shall denote this topological group also by HxtA. We shall fix A, H, and t throughout this section. A quick calculation shows that (120)
(h,a)(h,,a,)(h,a)-1 =
{hh'h-1,ah[a']hh'h-1[a-1]).
It follows from (120) that (121)
A-
={(eH,a')}
is a closed normal subgroup of G and that
(122)
(MXetf^'XM - 1 = ( e H } # > _ 1 ) .
In particular, the inner automorphism of G induced by (h,eA) coincides on A with the automorphism (eH,a') -> (eH,h[a']). We put (123)
H~ = {(h,eA)};
then # ~ is a closed subgroup of G. Identifying A with A ~ and H with H~ we find G = AH, (124)
{c} = An
H, 1
h[a] = hah' . The inhomogeneous linear groups are typical examples of semidirect products. In these examples, A is a real (or complex) finite dimensional vector space and a —> h[a] is an invertible linear endomorphism of A. H is thus identified with a subgroup of the group of all invertible (linear) endomorphisms of A, which is usually closed. The Euclidean group and the inhomogeneous Lorentz group fit into this general framework. For the Euclidean group, ^4=R n and H is the group 80(n). For the inhomogeneous Lorentz group, J = R 4 and H is the group of all invertible linear transformations of R4 which preserve the quadratic form coordinates on A). It undoubtedly would be of great interest therefore to develop a theory of the representations of semidirect products. In complete generality this has not been done. However, under certain restrictive hypotheses, we shall construct a theory of representations of semidirect products which goes far enough to yield all physically important results. From now on, through the rest of
238
GEOMETRY
OF QUANTUM
THEORY
the section, we shall assume that A is a closed, abelian, normal subgroup of G, H a closed subgroup of G such that the conditions (124) are satisfied. The first step of our theory is to obtain the representations of A. The irreducible representations are the characters of A, i.e., continuous homomorphisms of A into the group of complex numbers of modulus 1. Let A be the set of all characters-of A. Under pointwise multiplication, A becomes a group also. We shall equip A with the topology of uniform convergence on compacta. Then it can be shown t h a t A becomes a topological group and is in fact a locally compact Abelian group satisfying the second axiom of countability (cf. Pontrjagin [1]). Let us then take a o--finite measure a on A and consider the Hilbert space J#p = J£?2(A,a). For each a e A a n d / e Jf let (Uaf)(x) = x(a)f(x)(x e A), where x(a) is the value of the character x at a. Then a -> Ua is a representation of A. These representations do not exhaust the representations of A. To obtain more general representations we replace a by a projection valued measure P based on A. Let J f be a (separable) Hilbert space and P a projection valued measure based on A acting in 3tf. For each a e A let Ua be the operator in Jf defined by
vu.(E) =
Ua = f
x(a)dP(x).
I t is easy to show t h a t Ua is unitary and t h a t a —> Ua is a representation of A in Jtf. Conversely, it is a known theorem t h a t if U is any representation of A in Jf, there exists a unique projection valued measure P based on A such t h a t (126)
Ua=
f
x(a)dP(x)
for all a e A. P is called the projection valued measure corresponding to U. The formula (126) shows how U can be constructed from P. Thus, for instance, it is quite clear t h a t if T is an operator commuting with all PE, then T commutes with all Ua. Conversely, let T commute with all Ua. Then T commutes with all the spectral projections of all the Ua. Now, if r]a is the continuous m a p x - > x(a) of A into the group M of complex
SYSTEMS
OF
239
IMPRIMITIVITY a
numbers of modulus 1, it follows from (126) t h a t Q {F -> Pv-i(F)) projection valued measure based on M and
is a
Ua = f zdQa(z). JM
This shows t h a t the Pv-i(F) are the spectral projections of Ua and hence t h a t T commutes with all of them. Since the sets of the form rja~1(F) generate the Borel structure of A, it follows t h a t T commutes with P. The reader who wants to study the (Fourier) analysis of representations of A m a y consult Loomis [2, Chapter V I I ] . Let us now consider the lcsc group G and a representation U of G in a separable Hilbert space Jtf. Our method of analysis of U is essentially the Fourier analysis of the restriction of U to A. We begin this analysis by first constructing the " a d j o i n t " action of H on A. Lemma 6.21. Let h e H. Then, for each x e A there exists one and only one y e A such that (127)
y(a) =
x(h~\a\)
for all aeA. If we write y = h[x], then h, x -> h[x] is continuous H x A into A and A becomes a H-space.
from
Proof. That y is unique and well defined is obvious. The remaining assertions are easy to prove. We omit the proofs. Not every representation of A in 30? can be enlarged to a representation of G in ^f. Our next question is: when can this be done? Lemma 6.22. Let U and V be representations of A and H, respectively, in a separable Hilbert space Jf and let P be the projection valued measure on A corresponding to U. Then, a necessary and sufficient condition that there should exist a representation W of G in ^ whose restrictions to A and H are U and V, respectively, is that (V,P) is a system of imprimitivity for H, based on A. In this case, W is unique. Proof. Let W be a representation of G in J f and let Ua= Wa (aeA), Vh= Wh (h e H). U and V are representations of A and H, respectively. Now hah'1 = h[a] so t h a t (128)
VhUaVh-i
=
UhM
for all (h,a) e H xA. Let P be the projection valued measure on A corresponding to U. Now, a routine calculation based on transport of structure shows t h a t the projection valued measure of the representation a - > VfrUaVh'1 of A is E -> VhPEVh~x, and t h a t corresponding to the
240
GEOMETRY
OF QUANTUM
THEORY
n
representation a -> Uh[a^ is E -> Ph[Ev l view of the uniqueness of the projection valued measures which correspond to representations of A we infer from (128) t h a t (129)
VhPEVh~i
=
Ph[E].
(V,P) is t h u s a system of imprimitivity for H based on A. Conversely, let us start with U, V, and P such t h a t P and V satisfy (129). Then U and V satisfy (128). Define W on G by (130)
Wah = UaVh.
The relations (128) are now enough to secure the fact t h a t W is a homomorphism. Since a^ Ua and h -> V h are Borel, x -> Wx is Borel. W is t h u s a representation of G and its restrictions to H and ^4 are V and £/, respectively. Since G = AH, W is unique. The lemma is proved. The lemma just proved enables us to establish a one-one correspondence between representations of G and systems of imprimitivity of H based on A. Lemma 6.23. A representation W of G is irreducible if and only if the corresponding system of imprimitivity for H based on A is irreducible. Two representations of G are equivalent if and only if the corresponding systems of imprimitivity are equivalent. Proof. Let If be a representation of G in JF and let (V,P) be the corresponding system of imprimitivity, also acting in J^. We use t h e notation of lemma 6.22. If T is an operator in J f , t h e n T commutes with all the Ua (a e A) if and only if T commutes with all PE. From this it follows t h a t T lies in the commuting ring of W if and only if T lies in t h e commuting ring of (V,P). This proves the first assertion. For the second, let Wl be a representation of G in Jf ( , and ( V \ P l ) the corresponding system of imprimitivity (i = l, 2). If T is a unitary isomorphism of J f 1 onto ^f2, it follows from the uniqueness of the correspondence Ul <± Pl t h a t Ua2 = TUa1T-1 for all a e A if and only if P £ 2 = T P / T " 1 for all Borel sets E^A. From this t h e second assertion of the lemma follows quickly. The preceding lemmas and the theory of Sections 5 and 6 reduce t h e problem of describing the irreducible representations of G to the problem of describing the irreducible representations of the various stability subgroups of H with respect to its action on A. I n view of t h e applications t o the study of relativistic equations we proceed to spell out t h e details of the relevant constructions. We choose a point x0 e A and a cr-finite measure a which is quasi invariant on t h e J^-space A and which lives on t h e orbit H[x0] = X. Let HXQ = H0 be t h e stability group a t x0, i.e., (131)
H0 = {h:heH,
h[x0] = x0},
SYSTEMS
OF
IMPRIMITIVITY
241
and let m be a n irreducible representation of H0 in a separable Hilbert space Jfl L e t us define Jtf* = JP2(A,Jf,a) a n d let (V,P) be a system of imprimitivity for H based on A, living on the orbit X, and induced b y the representation m. W e p u t , for ah e G (a e A, he H) and / e <#?, (132)
(Wahf)(x)
= x(a)(VJ)(x)
(x e A).
Since PEf=%Ef for all Borel sets E^A, it follows easily t h a t , for t h e representation a - > Ua of A corresponding t o P, (133)
(Uaf)(x)
= x(a)f(x)
(xeA),
so t h a t TF a/l = UaVh. Thus TF is a representation of G. From our work in Sections 5 a n d 6 we know t h a t t h e equivalence class of the system (V>P) depends only on t h e measure class of a a n d t h e equivalence class of m. From lemmas 6.22 a n d 6.23 it follows t h a t W is irreducible. W e shall say t h a t W is associated with x0 a n d m. The following theorem is now an immediate consequence of our work in Sections 5 a n d 6 of t h e present chapter. Theorem 6.24 (Mackey [2] [3]). Let us choose, for each H-orbit (o in A, one point x^ on o>, and an irreducible representation m, of the stability subgroup of H(0 at the point xco. Then the representation ~Wm,l° of G associated with x^ and m is irreducible. Wm,co and Wm',co' are equivalent if and only if the orbits a> and to coincide, and the representations m and m' are equivalent. If the H-orbit structure of A is smooth, then each irreducible representation of G is equivalent to some Wm,co.
N O T E S ON C H A P T E R V I The theory of induced representations is originally due t o Frobenius who studied it for finite groups. I t s generalization t o t h e category of locally compact second countable groups is due t o Mackey. Nowadays Mackey's theory is referred t o as t h e Mackey Machine; it is also known as t h e little group method among physicists. Significant contributions were also made b y I.M. GeFfand a n d M . A . N a i m a r k in their work on representations of complex classical groups (cf. Trudy Mat. Inst. Steklova, 36 (1950), pp. 1-288), a n d b y F . B r u h a t (Bull. Soc. Math. France, 84 (1956), pp. 9 7 205) who developed t h e theory of induced representations for Lie groups with t h e techniques of t h e theory of Schwartz distributions. Especially noteworthy are the various expositions of Mackey in: The Theory of unitary group representations in Physics, Probability, and Number Theory, Benjamin/ Cummings, Reading, Mass., 1978; Harmonic analysis as the exploitation of symmetry—ahistoricalsurvey,Bull. Amer.Math. S o c , 3 (1980),pp.543-698. Unitary representation theory is so vast and currently so active t h a t it is
242
GEOMETRY
OF QUANTUM
THEORY
not possible to begin to sketch its outlines. For some notable surveys see: I. M. Gel'fand, Proceedings of the International Congress of Mathematicians Stockholm, 1962, pp. 74-85; the articles in Harmonic Analysis on homogeneous spaces, Proceedings of Symposia in Pure Mathematics, Vol. XXVI, Amer. Math. Soc. Providence, R.I., (ed.) C.C.Moore, 1973; and V.S. Varadarajan, Harmonic analysis on real semisimple Lie groups, Proceedings of the International Congress of Mathematicians, Vancouver, 1974, pp. 121-127. The theory for semisimple groups is mainly the creation of Harish-Chandra; for this, see Harish-Chandra, Collected Papers, 4 Vols., Springer-Verlag, New York, 1984; see also Harmonic analysis and representations of semisimple Lie groups, Reidel, Dordrecht, 1980, edited by J. A. Wolf, M. Cahen, and M. de Wilde. Recently, interest has intensified in representation theory of infinite dimensional groups; for a view through the theory of infinite dimensional Lie algebras see the book of Kac and the references therein: V. G. Kac, Infinite dimensional Lie algebras, Birkhauser, Boston, 1983.
CHAPTER VII MULTIPLIERS 1. T H E P R O J E C T I V E G R O U P Earlier in Chapter I I I we showed t h a t in any mathematical description of a quantum mechanical system the requirement of covariance introduces a representation of t h e symmetry group 0 of the system into the group A u t ( ^ ) , where Sf is the set of all states of the system in question. If we assume t h a t the logic of the system is standard, this representation can be replaced by a representation of G into the group of all symmetries of the Hilbert space underlying the logic. If 0° is the subgroup of G consisting of all elements in the same connected component as the identity of G, then it can be shown t h a t when G is a Lie group the symmetry corresponding to each element of G° is a unitary rather t h a n an antiunitary operatorf. However, these unitary operators are not uniquely determined. Each of them can obviously be multiplied by a complex number of modulus 1 (called a phase factor) without changing the induced automorphism of the logic. I t is therefore not immediately obvious t h a t we have a unitary representation of G°, i.e., t h a t the phase factors involved are all removable. I n this chapter we shall examine this question. We remark t h a t it was first studied by Wigner [1]. He proved, among other things t h a t , if G is the inhomogeneous Lorentz group, any representation of G in A u t ( ^ ) can be induced by a unitary representation, at least of the universal covering group of G. This result was the starting point for his classification of relativistic wave equations. Wigner's discussion of this problem, however, was restricted to the Lorentz and rotation groups. I t was Bargmann [1] who first examined these problems systematically. He obtained m a n y general theorems which included Wigner's results as special cases. I n his work on group representations, Mackey was led to some of these f If x e G is of the form y1 for y e G, then the symmetry associated with x is necessarily unitary. Now, if G is a Lie group and G° the connected component of the identity, then for any x of the form exp X (X in the Lie algebra of G), x = y2, where 2/ = exp J X . Since the range of the map X —> exp X contains a neighborhood of the identity, it follows t h a t every element in a certain neighborhood N of the identity of G is mapped into a unitary operator in a given representation. Thus the elements of the group generated by the elements of N are also mapped into unitary operators. But G° coincides with this group. 243
244
GEOMETRY
OF QUANTUM
THEORY
problems from a purely mathematical point of view. I n his paper [6] he formulated the basic concepts of a theory of such phase factors for arbitrary lcsc groups and showed how the subject can be fitted into t h e general framework of the theory of group extensions. This is the point of view taken in this chapter. W i t h every phase factor, or multiplier as we call them, we shall, following Mackey, associate a group extension of the unit circle by the given group and reduce the solution of problems involving t h a t multiplier to problems involving this group extension. The subject then becomes essentially a p a r t of a cohomology theory of certain types of extensions. Our purpose is not to develop a general theory of group extensions. Such a general discussion m a y be found, in various contexts and a t widely differing levels, in the works of Hochschild [1], [2], Calabi [1], Moore [1], Mackey [7], [8], Kleppner [1], and a number of other authors. Our aim is t h e much more modest one of developing in some detail the basic concepts of group extensions, so as to be able to obtain the solutions to the problems arising in connection with the physical space-time symmetry groups. We might mention t h a t H e r m a n n Weyl (cf. [1]), has also emphasized the fact t h a t it is the automorphism of the logic and not the underlying symmetry t h a t is of significance in physics. Some of the facts about projective representations m a y be found in his book. Let J f be a complex separable Hilbert space. We denote by °ll the unitary group of 3f. If we equip °li with the strong topology, we have seen in Chapter V t h a t ^ becomes a metrizable second countable topological group and t h a t its Borel structure is standard. We denote by ^ the set of all elements of °ll which are multiples of the identity operator. 2£ is a closed, normal, subgroup of °U and is central, i.e., each element of 3£ commutes with all the elements of °ll. The group (i)
& =
#/#
is called the 'proactive group of 34?. We write IT for the canonical homomorphism (2)
ir:U-*ir{U)
(UeW)
of °U onto 8P. Theorem 4.27 now tells us t h a t the subgroup of the group of automorphisms of the logic of 3PP, consisting of automorphisms which can be induced by unitary operators, is isomorphic to ^ . We shall equip SP with the quotient topology. Thus, a set A<^0> is defined to be open if and only if TT~1(A) is an open subset of °li. I t is easy t o show t h a t SP becomes a metrizable topological group satisfying the second axiom of countability and t h a t IT is an open continuous m a p of °U onto ^? If we write (3)
T = {z : z e C,
\z\ = 1},
MULTIPLIERS
245
then the kernel of n is precisely the set of operators of the form z\ (z e T), and is hence a compact group. I n all statements involving the topology and Borel structure on &, we shall refer only to the above-described topology of & and its natural Borel structure. Lemma 7.1. The following statements concerning a sequence of elements in % are equivalent: W (ii) (hi)
{Un}n=s012...
\
Proof, (ii) => (iii) obviously. On the other hand, for fixed / , / ' e tf, U -> |<£7/,/'>| is a continuous function on °ll which is constant on each i^T-coset. Hence, by virtue of the fact t h a t 3P has the quotient topology, it defines a continuous function on &. The implication (iii) => (i) follows from the assertion t h a t this function is continuous on ^ We shall complete the proof of the lemma by showing t h a t (i) => (ii). We begin the proof with a simple remark. Suppose U is an operator of je such t h a t | < ^ / J ' > | = | >/'> I f o r a l l / , / ' £ . # : Then V = t\ for some t e T. I n fact, for a n y / , Ufe{f}11, so t h a t Uf is a multiple off. Since this is true for each / , U must be a multiple of the identity, U = tl. I t is obvious t h a t \t| = 1, and thus t e T. This said, we proceed to show t h a t (i) => (ii). Replacing Un by UJJQ'1 we m a y assume t h a t U0 = 1. Since || Un \\ < 1, there exists a n operator U with || U|| < 1 and a subsequence {Un]c} such t h a t
GEOMETRY
246
OF QUANTUM
THEORY
^ is hence Hausdorff. Let D be a countable dense subset of Jtf*. Then an easy argument shows t h a t 3^ is also the smallest topology with respect to which the maps dfJ, (f, / ' e D) are continuous. Hence y is metrizable and lemma 7.1 now implies the first statement. The second statement is now clear. Lemma 7.3. Let fe Jf7 be nonzero and let (4)
Qf = {U :Ue<%,
<£//,/> is real and > 0}.
Then Qf is a Borel set in % containing 1 which meets each ^-coset at most once. Moreover, 7r[Qf] is open in & and -n is a homeomorphism of Qf onto Proof. That Qf is a Borel set containing 1 is obvious. If (Ufjy > 0 and t e T, (tUf,fy > 0 if and only if t= 1. Thus Qf meets each iF-coset a t most once. On the other hand, if <£//,/> = 2 ^ 0 , then for V=\t\t~1U, one has < VfJ} = \t| > 0. This proves t h a t (5)
T T - V W / ] ) = {U:UeW,
Equation (5) shows a t once t h a t n[Qf] is open in 0*. I t remains to prove t h a t 7ris a homeomorphism onQ f . Let {Un}n = 0 1 . . . b e a sequence inQ f such t h a t 7T(Un)->7r(U0) in ^ . By lemma 7.1 there exists a sequence {tn} in T such t h a t tnUn^U0 in # . Since \
c : x - > c(x)
such that (i)
(7)
TT(C(X)) = x
for
all
xe&,
(ii)
c( w (l)) = l,
(iii)
there is an open set containing TT( 1) on which c is continuous.
The range of c is a Borel subset of % which meets each 3£-coset exactly once. Finally, iff is a map of & into some Borel space X, f is Borel if and only if the map IJ -^ f (TT(TJ)) of °ll into X is Borel. In particular, A<^& is Borel if and only if ' rr~1(A) is a Borel set in °ll. Proof. We begin our proof with the construction of c. Let {fl9 • • •, fn, • • •} be a countable dense set of nonzero vectors and let (cf. (4))
(8)
Qn = Qfn-
MULTIPLIERS
247
We now define sets E±, E2,• • •<^°llas follows: (9)
E, = Ql9
and for n > 1,
(10)
En = Qnn " h V : <£%/<> = 0}. t=l
Let 00
(ii)
^ = u tf». n=l
i£ is a Borel subset of ^ . We claim t h a t E meets each iF-coset exactly once. I n fact, let V e % and let m be the smallest of integers i such t h a t
2. M U L T I P L I E R S AND P R O J E C T I V E
REPRESENTATIONS
After these preliminaries involving the projective group, we are in a position to introduce the basic concepts centering around the so-called projective (or ray) representations of a group. We begin with the concept of a multiplier. Let G, K be lcsc groups with K abelian. By a K-multiplier for G we mean a Borel m a p m (12)
m : x, y -> m(x,y)
of G x G into K such t h a t (i)
m(x,yz)m(y,z)
(ii)
m(x,e) = m(e,x) = 1
= m(xy,z)m(x,y) for all
for all
x, y, z e G,
a; e 6r.
Note t h a t we are writing K multiplicatively and t h a t we write 1 for the identity of K. If K = T, the multiplicative group of complex numbers of modulus 1, we omit the prefix K and speak simply of multipliers. If K is additively written, the equations (13) become (i)
m(xy,z) + m(x,y) = m(x,yz) + m(y,z)
(14) (ii)
m(x,e) = m(e,x) = 0
for all
xeG.
for all
x,y,ze
G,
248
GEOMETRY
OF QUANTUM
THEORY
1
It follows from (13), on putting y — x' , z = x, that m(x,x~1) = m(x~1,x)
(15) for all x e G.
The set of all i£-multipliers for G is obviously a commutative group under pointwise multiplication. We denote this group by MK'(G). Two K-multipliers m1} m2 for G are said to be similar, mx~m2 in symbols, if there exists a Borel function a on G with values in K, a : # -» a(#) such that
(16)
^ ^ - s ^ " ^
for all x,y EG. Note that a(e) is necessarily equal to 1. If a ^-multiplier m is similar to the multiplier 1 : x, y —> 1, we shall say that m is exact. Thus m is exact if and only if for some Borel map a of G into K, (17)
m(*,*/) =
v U)
a(x)a(y)
for all x, y E G. From (17) it is easily calculated that the exact K-multipliers for G form a subgroup EK{G) of MK\G). We form the quotient group (18)
MK(G) = MK'(G)IEK(G)
and call it the K-multiplier group of G. When i£= T we write M(G) for MT(G) and call it the multiplier group of G. Suppose that m is a multiplier for G. A mapping U:g^Ui
9
of G into the unitary group % of a separable Hilbert space 3f is said to be an m-representation if (i) g-> Ug is Borel, i.e., g -> (Ugf,f'y (19)
is Borel for all
J J
(ii) ff. = l, (hi) ^ = m(x9y)UxUy
for all
x,yEG.
A Borel mapping -> UgofG into ^ is said to be a projective representation if there exists a multiplier m for # such that U is an m-representation; then m is obviously uniquely determined by U. We shall say that m is the multiplier of U. If £7 is a projective representation of (^ in ^ the map (20)
iroU:g-*7r(Ug),
MULTIPLIERS
249
of G into the projective group ^ is Borel and (iii) of (19) shows t h a t TT o U is a homomorphism. Conversely, we shall prove t h a t every Borel homomorphism of G into 3P arises from a projective representation in this fashion. Theorem 7.5. Let m be any multiplier for G and let U be any m-representation of G acting in £F. Then TT O U is a continuous homomorphism of G into 0*. If V is any projective representation of G in 3f such that TT o U = IT o V, then the multiplier of V is similar to m; and conversely, if m' is a multiplier similar to m, there is an m'-representation V with IT O Y — TT O U. Suppose u
:g->ug
is a Borel homomorphism of G into 0. Then there exists a projective representation U of G in #? such that (21)
u =
TT
o U;
moreover U can be chosen to be continuous in some open subset of G containing 1. Finally, if m is any multiplier for G, there exist m-representations of G. Proof. Let U be a projective representation of G in Jf. Then x-^ir( Ux) is a Borel map of G into 0. Since Uxy is a constant multiple of UxUy, 7r (Uxy) = 7T(Ux)7T(Uy). Thus TT o U is a homomorphism of G into ^ B y lemma 5.28,770 U is continuous. Suppose V is another projective representation of G in Jtf* and let m and m' be the multipliers of U and U', respectively. Then TT O U = TT O U' if and only if for each x e G there is an a(x) 6 T such t h a t Ux' = a(x)Ux. The equation TJX~XTJx' = a(x)l shows t h a t the function a(x -> a(x)) is Borel. If such an a exists, then a direct calculation shows t h a t
(22)
m>(x,y)
=
m(x,y)J^L
for all x, y e G. Thus m~mf. Conversely, suppose m~m''. Then (22) is satisfied for a suitable Borel function a on G with values in T and, by setting Ux =a(x)Ux, we obtain an m'-representation V with TT O U' = TT o
U.
Next, let u : g -> ug be a Borel homomorphism of G into ^ ; u is continuous b y lemma 5.28. We select a Borel m a p c of ^ into ^ satisfying the conditions (i) to (iii) of (7). Let (23)
Ug = c(ug).
250
GEOMETRY
OF QUANTUM
THEORY
Since c is Borel and U = c o u, U(g -> Ug) is a Borel map of G into /. Let A<^0> be an open set containing 7r(l) on which c is continuous. Then U is continuous on the open subset u~1(A) of G containing 1. If x, y e G, then it is obvious that there exists an m(x,y) e T such that (24)
Vxy = m(x,y)UxVv.
Since UxyUy~Wx^
= m(x,y)l,
it follows that x, y —> m(x,y) is a Borel map from GxG into T. The equation uxuyz = uxyuz, expressed in terms of the Ug% gives (i) of (13), while the relations uxe = uex = ux lead to (ii) of (13). m is, in other words, a multiplier for G, and U is an m-representation, continuous around e. Finally, let m be an arbitrary multiplier for G. We write J f = j£?2(6r,/xr), where /*r is a right Haar measure on G. For any a;e(? and / e <#?, we define (25)
(Uxf)(g) =
m(gix)-y(gx).
It is readily verified that £7e = l, Uxy = m(x,y)UxUy. The equation || /7 X /|| 2 = ||/1| 2 shows that each Ux is unitary. If/,/' e ^
<£W> = jGMg^)-1f(gx)r(grd^r(g), and as the integrand is a Borel function on G x X, # -> (Uxf,f'} is a Borel function on 6r. £7 is thus an m-representation. The proof of the theorem is complete. Corollary 7.6. Let m be any multiplier for G. Then m is similar to a multiplier which is continuous on some open subset of GxG containing (e,e). Proof. Let U be an m-representation of G, acting in a Hilbert space Jtif. From theorem 7.5 we know that there exists a multiplier m' similar to m and an m'-representation V acting in 3f such that (i) TT O U'—TT O U, (ii) for a suitable open subset N of G containing e, x —> Ux is continuous from N into °ll. Let iV\ be an open set containing e such that N^^N. As tf*y' = m'(x,y)Ux'Uy', and as 6T/ is continuous on N, it is clear that m' is continuous on N-^xN^ Corollary 7.7. Let m be a multiplier for G and U any m-representation of G in J^. Then m is exact if and only if there exists a representation U' of G in 3f such that TT O U' — TT O U. If TJ" is another representation of G in J f such that TT o U" = TT o U, there exists a continuous homomorphism x of G into T such that (26) for all g eG.
Ug" = XgVg'
251
MULTIPLIERS
Proof. The first assertion is an immediate consequence of theorem 7.5. For the second, we note that U'g~1Ug" = xg-l for each g e G, where X9 e T. Then g -> %g ™ Borel and multiplicative. Therefore x is a continuous homomorphism of G into T.
3. MULTIPLIERS AND GROUP EXTENSIONS The obvious problem now facing us is of course that of finding, for given lcsc groups G and K (K abelian), the if-multiplier group MK(G). If, for example, K=T and M(G) is trivial (as it may well be), then we know that every multiplier for G is exact. In this case there is no practical difference between projective and ordinary representations. Our aim is to examine this problem at least for special classes of G and K. We introduce in this section a very useful technique to study the multipliers for a given group, namely the method of group extensions (cf. Mackey [6]). Suppose G and K are two lcsc groups. We shall, as always, assume that K is abelian. By an extension of K by G we mean a triple (H,i,j), where H is a lcsc group, i is an isomorphism of K onto a closed normal subgroup of H, and j is a homomorphism of H onto G with kernel i[K]. We require that i be a homeomorphism and that j be continuous. {H,i,j) is said to be a central extension if i[K] is contained in the center of H. In the language of cohomology theory, (H,i, j) is an extension of K by G if and only if the sequence (27)
0
>K-^->H-^>G—>0
is exact. If (H,i,j) and (H',i',j') are two extensions, then we say that they are equivalent, if there is an isomorphism q of H onto H\ such that, the diagram
(28)
G
K H'
commutes, i.e., for all h e K, (29)
q(i(k)) = i'(k),
and for all h e H, (30)
j'(q(h)) = j(h).
Once again we require that q be a homeomorphism, i.e., q is an isomorphism of the lcsc group H onto the lcsc group H'. We recall that for two lcsc groups A and B, any continuous homomorphism of A into B, which is
252
GEOMETRY
OF QUANTUM
THEORY
one-one and maps A onto B, is necessarily a homeomorphism (cf. Pontrjagin [1]). We shall now associate with any central extension (H,i,j) of K by G, a K-multiplier for G. Since H is lcsc and i[K] is a closed subgroup, there is a Borel section c based on G, i.e., a Borel m a p c of G into # , (31)
c:G->H
such t h a t c(e) = i(l), the identity of H, and for all x e G, j(c(x))
= x
(cf. theorem 5.11). Therefore, the m a p (32)
x,k->
c(x)i(k)
is a one-one Borel m a p of the standard Borel space Gx K onto the standard Borel space H. Consequently, it is a Borel isomorphism. We denote by d, the inverse of this map, which is a Borel isomorphism of H onto GxK: (33)
d : c(x)i(k) - » x, k.
If x, y e G, c(xy) and c(x)c(y) belong to the same ^[TTj-coset of H, so t h a t (34)
i-1[c{xy)~1c(x)c(y)]
m(x,y) =
is an element of K. Clearly m is a Borel m a p of G x G into i£ and m(x,e) = m(e,x) = 1 for all # e 6r. If we now remember t h a t i[m(x,y)] commutes with all elements of H, we find, for x, y, z e G, i[m(x,yz)m(y,z)]
c(xyz)~1c(x)c(y)c(z)
=
c(xyz)-1c(xy){i[m(x,y)]}c(z)
=
= c(xyz) ~ 1c(xy)c{z)i[m(x,y)] = Thus m is a K-multiplier (35)
i[m(xy.z)m(x,y)].
for G. We note at the same time t h a t
c(x)i(k)c(y)i(k')
=
c(xy)i[m(x,y)kk']
for all x, y e G and k, k! e K. If we use the fact t h a t d is a Borel isomorphism between the Borel spaces H and G x K, then (35) tells us t h a t the definition of a product operation in Gx K given by (36)
(x,k)(y,k')
=
(xy,m(x,y)kk')
converts GxK into a Borel group, and t h a t d provides a Borel group isomorphism of this group with H. The converse to the result described just now is a much more difficult question. Is it possible to associate with each ^-multiplier for G a central extension of K by G\ The answer is affirmative as the following theorem, due to Mackey [6], shows.
MULTIPLIERS
253
Theorem 7.8. Let m be any K-multiplier. If we define for (x,k), (y,k') e GxK, (37) (x,k)(y,k') = (xy,m(xy)kkf), then GxK becomes a standard Borel group under this product, and there exists a unique Icsc topology for GxK such that (a) GxK becomes a Icsc group, denoted by G xmK, under this topology, and (b) the Borel structure of this topology coincides with the product Borel structure on GxK. If (38)
JoM = 9, then (G xmK, i0, j0) is a central extension of K by G. Every central extension of K by G is equivalent to one such. The extensions corresponding to the K-multipliers mx and m2 are equivalent if and only if' m1 and m2 are similar. Proof. We begin with a central extension (H,i,j )ofKhyG and associate with it a i£-multiplier m, using the development contained in the equations (32) through (36). Then d : c(x)i(k) -> (x,k) is an isomorphism of H onto G xmK, where G xmK is considered a group with (37) defining the product operation. G xmK is obviously a standard Borel group and is also Icsc if we equip it with the topology which makes d a homeomorphism. It is obvious that the extensions (H,i,j) and (G xmK, i0,j0) are equivalent. Conversely, let m be a if-multiplier for G. Then under (37), GxK becomes a standard Borel group, as is readily verified with the help of the relations (13). We write G xmK for this group. The relations m(x,e) = m(e,x) = l show that {(e,Jc) : k e K} is a central subgroup of G xmK. Clearly i0(k -> (e,k)) and j0((g,Jc) -> g) are Borel homomorphisms. We shall next prove that if A is a Haar measure on K and //, is a left Haar measure on G, JJL X A is a left invariant measure on G xmK. In fact, if/is a fx x A-summable Borel function on G x K and (g0,k0) e G x K, we have, by Fubini's theorem, j j f((go,h)(9,k))d(fix\)(g,k)
= Jj
GxK
f(g0g,m(g0,g)k0lc)d(fxx\)(g,k)
GxK
=
fa(jKftot0MW))dfito)
= £ ( £ /(flW*)d/*to))«*A(i) = jK ( £ f(g,kW{g))d\(k) = jj f(g,k)d(^xX)(g,k).
254
GEOMETRY
OF QUANTUM
THEORY
\i x A is thus left invariant. Theorem 5.41 now enables us to conclude t h a t GxmK becomes a lcsc group under its Weil topology. The Borel structure of this topology is by choice the product Borel structure on G x K, and it is the only lcsc topology with this property, as proved in t h a t theorem. The m a p j 0 : g, k - > g is a Borel homomorphism of the lcsc group G xmK onto the lcsc group G and is thus continuous. This proves t h a t i0[K] = kernel (j0) is closed in G xmK. A similar reasoning shows t h a t i0 is continuous also. Therefore (G xmK, i0,j0) is a central extension of K by G. The only thing t h a t remains to be proved is t h a t the extensions GxmiK and G xm2K are equivalent if and only if m1^m2. For a n y i£-multiplier ra, let us write (39) Suppose m1c^m2.
Hm = G
xnK.
Then there exists a Borel map a, a:g-+
a(g)
of G into K such t h a t /
/ = ml{x,y)
m2{x,y)
a(xy) ——
for all x, y e G x G. If we write (40)
q(g,k) = (g,a(9)k),
then it is trivial to check t h a t q is a Borel group isomorphism of Hmi onto Hm
q(g,l) =
(gAg))-
a is clearly a Borel m a p of G into K. Then, as (<7,l)(e,&) = (g,k) we find Q(g>k) = (g,a(9))(e,k)
= (gMg)k)-
(in Hm2)
MULTIPLIERS
255
If we now use the fact t h a t q preserves the group structures, we can conclude, after a brief computation, t h a t . m2(x,y)
.
. =
a(xy) mi{x,y)^-^)
for all x, y e Gx G. I n other words, m1~m2. proved.
The theorem is completely
Remark. The essential difficulty of the theorem proved above resides in the fact t h a t when the if-multiplier m is not continuous on GxG, the product topology would no longer serve to make G xmK lcsc. I n fact, if m is continuous on GxG, it is obvious t h a t under the product topology, Hm is a topological group and is even a lcsc group. Since the product topology generates the product Borel structure, it follows from the uniqueness of the Weil topology t h a t the two topologies coincide. Conversely, if the Weil topology for Hm coincides with the product topology, the m a p x - > (x,l) of G into Hm is continuous and hence x, y -> (x,l)(y,l) = (xy,m(x,y)) is also continuous, showing t h a t m is continuous. Note t h a t , in this case, c0 :#—>(#,1) gives a global continuous section for Hmji0[K]9 i.e., c 0 is continuous and Jo(c0(x)) = x for all xeG. This shows t h a t , by using continuous if-multipliers, we can construct only those central extensions (H,i,j) of K by G for which there exists a continuous section c from G into H. There are m a n y examples where such global sections do not exist. For instance, if K is finite but both H and G are compact and connected, no continuous section for Hji[K] can be defined on all of G. An example of this is obtained when we take II to be SU(n,€), the special unitary group in ^-dimensions, and K as the center of H. On the other hand, one can impose certain conditions on K under which any K-multiplier for G will be similar to one which is continuous in some open neighborhood of (e,e) in GxG. Corollary 7.6 tells us t h a t this is so when K=T. More generally, it can be shown t h a t this is the case when K is a compact Lie group, or when K and G are both Lie groups. We shall call a K-multiplier locally continuous if it continues on some open set containing (e,e) in Gx G. The assumption of local continuity for m is of a much less serious nature t h a n the assumption of global continuity. When m is locally continuous, it is almost obvious t h a t one can construct, using local mappings, a fiber space topology for Hm over G which is locally trivial and which coincides with the Weil topology. WTe shall derive this as a corollary of theorem 7.8. The result is stated in corollary 7.10. We need an auxiliary result. Corollary 7.9. Suppose m is locally continuous. Then there exists an open set N around e such that x-> (x,l) is a homeomorphism of N into Hm.
256
GEOMETRY
OF QUANTUM
THEORY
Proof. We select a symmetric open set Ax containing e with compact closure such that m is continuous on A± xA1 and select A2, A3,- • • such that each An is open, symmetric, contains e, and An + 1An + 1^An for all n. Let {xn} be a sequence in A3, let x0 e A3, and let xn-^ x0. We shall show that (xn,l) -> (#0>1) in the Weil topology of Hm. Write yn = x0~1xn. Then Vn e A2, yn -» e, and (42)
(^ 0 ,l)(i/ n ,l)(e,m(^ 0 ,i/ n )- 1 ) = (s n ,l).
Since the map k -> (e,k) is continuous from i£ into Hm and since m is continuous on A1xA1, (42) shows that (xn,l) will converge to (x0,l) provided (yn9l) -> (e,l). Consider now ^42 x iC =j 0 _1 (^4 2 ). This is an open subset of Hm containing (e,l) and (yn,l) belongs to it for all n. In order to prove that (yn,l) -> (e,l) it is enough to prove that \\L(yntl)f—f\\ —> 0 for all / in J?2(Hm) which vanish outside A2xK. Let D be the set of all functions on GxK, continuous with respect to the product topology and vanishing outside subsets of A2 x K which are compact in the product topology; it is enough to prove that L(Vntl)f-+f for a l l / e D. F i x / e D and let C c A2 and i£x c= i£ be compact sets such t h a t / vanishes outside CxK±. Now,
<W><*,*>=/(^*.^f If # £ (7, yn~1x $C for sufficiently large n and (LVntlf)(x,k) ==f(x,k) = 0 for sufficiently large w. If # e A2, the continuity of m on ^4X x ^ implies that (L(yntl)f){x,k) -+f(x,k) for all keK. Now, |/(x,Jfe)| < B for all (a;,ifc) e G x K and so, \(L(yrifl)f)(x,k)\
(43)
JJ|A,„,i)/-/l 2 ^xA)^0. GxK
We have thus proved that x-> (x,l) is continuous from A3 into Hm. As A3 is compact, it is a homeomorphism. This finishes the proof of the corollary. Corollary 7.10. i / m is continuous around (e,e), there exists an open set N containing e such that the Weil topology on the open set N xK coincides with the product topology. Proof. Let AT be an open set around e such that x -> (#,1) is a homeomorphism of N into Hm. Since (x,k) = (x,l)(e,k) and since x-+ (x,l) and k -» (e,k) are both homeomorphisms, the corollary follows at once.
MULTIPLIERS
257
Remark. Corollary 7.10 shows t h a t when m is locally continuous, Hm, as a fiber space over G, has a locally trivial fibration. For a more intensive study of the topology of Hm we refer the reader to the work of Calabi [1]. The main source of technical complications in our theory is the fact t h a t the topology on Hm in general bears no visible relation to t h a t of G and K. We now describe another corollary to theorem 7.8 which gives an example of the nature of the indirect arguments needed in analyzing Hm. Corollary 7.11. If K and G are both connected (resp. simply connected), so is Hm. If K and G are compact, so is Hm. Both the assertions follow from the fact t h a t i0[K] is a closed subgroup of Hm and G is isomorphic to Hmli0[K], Our first use of these group extensions is a description of the similarity classes of iC-multipliers. Let m be a iC-multiplier for G, and Hm the corresponding group extension. For any Borel m a p a of G into K with a(e) = 1, let ca(x) = (x,a(x))9 ma(x,y)
= m(x,y)
-
I t is easily seen t h a t ma = mfl2 if and only if there is a Borel homomorphism k of G into K such t h a t ax(x) = k(x)a2(x) for all x e G. The m a p ca - > ma establishes a correspondence between multipliers similar to m and Borel sections, and reduces the analysis of the similarity class of m to the analysis of the geometry of Hm as a fiber space over G. Theorem 7.12. ca is continuous (respectively continuous around e) if and only if ma is continuous (respectively continuous around (e,e)). m is exact if and only if there exists a Borel section c which is a homomorphism of G into Hm. Proof. Regarding continuity we shall prove only local assertions; the global ones follow from similar reasoning. Suppose now ca is continuous on an open set N around e. Choose an open set Nx around e such t h a t N^^N. Then the m a p (45)
q:x,y->
ca(xy)ca( y) ~ ^(x)
~x
is continuous from N± x Nx into Hm. A simple calculation yields (46)
q(x,y) =
(e.m^x^y)-1).
Since k - > (e,k) is a homeomorphism, (46) shows t h a t ma is continuous on N1xN1. Conversely, let ma be continuous on N x N for a suitable N. If we form the group Hma, we know from corollary 7.9 t h a t d:x->(xi\)
258
GEOMETRY
OF QUANTUM
THEORY
is a homeomorphism of an open set iVi^iV containing e into Hma. Now, from the relation between m and ma, it follows quickly t h a t p : x, k —> x, Jca(x) ~x is a Borel group isomorphism of Hm on Hma. Hence p is a homeomorphism. Therefore, the m a p p'1 °d is a homeomorphism of Nx into # m . B u t (p'1 o d)(x) = (x,a(x)) = ca(x). Therefore ca is a homeomorphism of N1 into i/ m . Now we come to the question of exactness. By very definition, m is exact if and only if for some a, ma(x,y) = 1 for all x, y e G, i.e.,
<«"
«•*> - S ^ S i
for all x, y e G. A trivial computation shows t h a t (47) is true if and only if x->(x,a(x)) is a homomorphism of (? into Hm. The proof of theorem 7.12 is complete. A trivial reformulation of the second half of this theorem gives the following corollary: Corollary 7.13. m is exact if and only if there exists a subgroup G' of Hm such that G' is a Borel set meeting each i0[K]-coset exactly once, i.e., G' O i0[K] = {(e,l)} and G' -i0[K] = Hm. We shall now make a brief digression by obtaining some more corollaries which m a y serve as illustrations of the use and applicability of the preceding theory. Corollary 7.14. Let G be connected and abelian. Then, for a K-multiplier m to be exact, it is necessary and sufficient that m(x,y) — m(y,x) in some neighborhood of (e,e). Proof. T h a t an exact multiplier is symmetric (when G is abelian) follows at once from (17). Conversely, let N be an open set containing e such t h a t m(x,y) = m(y,x) for all x,y e NxN. Then, the formula for multiplication in Hm shows t h a t (x,k) and (y,k') commute whenever x and y belong to N. Since G and K are connected, so is Hm and hence the elements of the form (x,k) (x e N) generate Hm. This shows t h a t Hm is itself abelian. Since K is isomorphic to Tr x Vs, it follows from the structure of locally compact abelian groups t h a t {e} x K is a direct summand of Hm. Corollary 7.13 now implies t h a t m is exact. Corollary 7.15. There exists a multiplier m' similar to m such that m' is locally bounded as a map ofGxG into K (cf. Section 5 of Chapter V). Proof. Let a be chosen so t h a t ca is a regular Borel section, i.e., x-> (x,a(x)) is a m a p of G into Hm which sends compact sets of G into sets with
MULTIPLIERS
259
compact closure (in Hm). Denning q as in (45), it is obvious t h a t q maps any compact subset of G x G into a subset of Hm which has compact closure. The formula (46) tells us now t h a t ma is locally bounded. We shall end this section with an application to the theory of projective representations. Theorem 7.16. Let m be any multiplier for G and let U(g -> Ug) be an m-representation of G in a (separable) Hilbert space Jf. Then (x,k) - > k~1Ux is a representation of Hm in Jtf*. Conversely, if (xjc) -> V(Xtk) is a representation of Hm such that V^^^k'1! for all k e T, and if we write Ux = V(Xfl), then U(x -> Ux) is an m-representation of G and F(a.tfc) = ^" 1 ?7' x . Proof. The proof needs only minor calculations and is omitted. Corollary 7.17. Let m be any multiplier for G. Then there exists an irreducible m-representation of G. Proof. Let V be a representation of Hm such t h a t F(e>fc) = A;~1l for all keT. Such V exist, by theorem 7.5, because G admits ra-representations. Let us decompose V into a direct integral of irreducible representations V,
(48)
V~ |V
(cf. Mautner [1], Mackey [5]). Then, from the theory of the direct integrals of representations, it is easy to see t h a t for some £ 0 , Ffe°ffc) = ^ ~ 1 l for all k e T (in fact ^-almost all £ will have this property). Write V°= V*o, and let U° be the corresponding m-representation of G. Since F°-r>fc) = ^~ 1 Z7 a , 0 , the irreducibility of V° implies t h a t of U°. Corollary 7.18. / / G is compact and m is a multiplier for G, then there exist finite dimensional m-representations of G. The irreducible m-representations of G are finite dimensional and every m-representation of G is a direct sum of irreducible m-representations. Proof. We observe t h a t Hm is compact (corollary 7.11). The present corollary now follows in exactly the same fashion as the previous one. The proof is even more elementary, since only direct sums, rather t h a n direct integrals, of representations, are involved.
4. M U L T I P L I E R S F O R L I E G R O U P S We shall now make a deeper study of the group extensions Hm under the assumption t h a t K and G are both connected Lie groups. Since K is abelian, it would be a direct product of a real vector group and a torus.
260
GEOMETRY
OF QUANTUM
THEORY
The first basic result of this section tells us t h a t Hm becomes a Lie group in a natural manner. The way is then opened for the application of results of Lie group theory. We shall use these to show t h a t any Kmultiplier for G is similar to one which is globally analytic on G x G (at least when G is simply connected), and t h a t the elements of the multiplier group are in natural one-one correspondence with equivalence classes of central extensions of the Lie algebra of K by the Lie algebra of G. For general information and terminology involving Lie groups and Lie algebras we refer the reader to the books of Che valley [1], [2], [3] and Jacobson [2]. Throughout this chapter K will denote a connected abelian Lie group and G a connected Lie group. The symbol Tr will denote a torus of dimension r and the symbol Vs will denote a real vector space of dimension s. Both are regarded as Lie groups in the natural sense. Our first aim is to introduce an analytic structure on Hm. The essential step in this is accomplished by the following lemma. Lemma 7.19, Let K be an additive vector group and m a K-multiplier for G. Then there exists a K multiplier m' similar to m such that m! is C°° on GxG. Proof. Note t h a t m satisfies the identities (14). I t is enough to prove this when K is the additive group of real numbers, since the general case will follow by applying the special case to each component. By virtue of corollary 7.15 we m a y assume t h a t m is bounded on each compact subset of GxG. Let /JL be a left H a a r measure and /xr a right H a a r measure forG. Let then K = B1. Choose a (700 function with compact support, s a y / , which is such t h a t (49)
$ f(9)d^(g)
= 1.
Let a be the function x -> a(x), where (50)
a(x) =
m(x,u)f(u)d[x(u). JG
Clearly, a is bounded on each compact set. Let (51)
™>i(x,y) = m(x,y) +
a(xy)-a(x)-a(y).
1
Then m1 is a R -multiplier similar to m and is bounded on compact subsets of GxG. Moreover, m x
i( ,y)
=
{m(x,y) + m(xy,u) — m(y,u) — m(x,u)}f(u)dfjL(u)
=
{m(x,yu) —
m(x,u)}f(u)dfji(u),
MULTIPLIERS
261
using the multiplier identities for m. Thus (52)
m^x.y)
m(x,u)f(y~1u)d^(u)
=
—
JG
m(x,u)f(u)dfji(u). JG
The uniform convergence of the integrals on the right of (52) shows that for each x e G, y —>rn^x^y)is a 0°° function. Define a1 to be the Borel function y-> ax(y), where (53)
a±(y) =
m1(u,y)f1{u)dfxr(u) JG
jfi being a C00 function with compact support and satisfying (54)
j fi(u)drr(u) = 1.
Let (55)
m2(x,y) = m 1 (s,y) + a 1 (a#)-a 1 (a;)--a 1 (y).
Equation (55) shows that m2 is a i?1-multiplier similar to mx and bounded on compact subsets of G x G. Moreover, a brief calculation shows that (56)
m2(x,y) =
m^u^f^ux-^d^u)JG
m1(u,y)f1(u)dfjLr(u). JG
From (56) it follows that m2 is a O00 function of each of its variables when the other is fixed. Finally, define a3 by (57)
a3(x) =
m2{x,u)f(u)dfi(u).
Write (58)
m'(x,y) =
m2(x,y)+a3(xy)-a3(x)-a3(y).
1
Then m' is a R -multiplier similar to m and (59)
m2(x,u)f(y-1u)dyL(u)-
m'(x,y) = JG
m2(y,u)f(u)dp(u). JG
Since x, y -> m2(x,u)f(y~1u) is C™ on GxG for each w, (59) shows that m' is O00 on G x 6r. The lemma is proved. We shall now examine the case when K is not simply connected. We cannot hope for global results but the local version of lemma 7.19 can be proved by the same technique. Lemma 7.20. Let K be a connected abelian Lie group and m0 any Kmultiplier. Then there exists a K-multiplier m 0 ' which is similar to m0 and which is (700 in an open neighborhood of (e,e).
262
GEOMETRY
OF QUANTUM
THEORY
Proof. Since lemma 7.19 takes care of the case K = RS, it is enough to prove the present result when K is a torus and thus when K=T. We m a y (and do) therefore assume t h a t K = T. By corollary 7.6 we m a y assume t h a t m 0 is locally continuous. I t is obvious t h a t we can choose a real bounded Borel function m on G x G such t h a t (60)
m0(x,y)
=
exip{im(x,y))
for all x, y e G, and such t h a t m satisfies the identities characteristic of an R1 -multiplier (cf. (14)) locally, i.e., for some symmetric open set M with compact closure containing e, (61)
m(x,e) = m(e,x) — 0
for all x in M, and (62)
m(xy,z)-\-m(x,y)
= m(x,yz) + m(y,z)
for all x, y, z E M. We now proceed exactly as we did in the previous lemma. We c h o o s e / t o be 0°°, and satisfying (49), b u t with the extra requirement now t h a t / be zero outside a compact subset of M. a is defined as in (50) and m1 as in (51). Since exp(im) = m 0 is a T-multiplier, (51) shows t h a t exp(^'m1) is a T-multiplier similar to m 0 . m1 is bounded o n ^ x ^ and satisfies (61) for x e M and (62) for x, y, z e M. Equation (52) is now valid for all x, y e M, from which we infer t h a t for each x e M, y -> mx(x,y) is C00 in M. We define a1 and m 2 as in (53) and (55), respectively, but after choosing / x such t h a t f1 also vanishes outside M. Once again, (56) is valid for all x, y E M. Proceeding as we did before, we infer finally t h a t m 3 is 0°° on MxM. This shows t h a t exp(im 3 ) = m' is (700 on MxM, and is a Tmultiplier similar to mQ. The lemma is proved. We are now in a position to introduce a differentiable structure on Hm. This will then enable us to treat Hm as a Lie group. From this, t h e introduction of analytic coordinates on Hm will follow in the usual fashion. Theorem 7.21. Let K be a connected abelian Lie group and G a connected Lie group. Let m be any K-multiplier for G. Then there exists a unique analytic structure for Hm (compatible with the Weil topology) which converts Hm into an analytic group. The maps k - > (e,k), of K into Hm, and (g,k) - > g, of Hm onto G, are analytic. Proof. Hm is connected by corollary 7.11. We know t h a t any topological group admits at most one compatible analytic structure which converts it into a Lie group (cf. Che valley [1], pp. 127-129). Thus we are concerned only with existence. Since we m a y pass to K-multipliers similar to m, we m a y assume, after lemma 7.20, t h a t m itself is 0°° around (e,e). F r o m
MULTIPLIERS
263
corollary 7.10 we know t h a t for some open set M containing e, the Weil topology on M x K coincides with the product topology. We choose M so small t h a t this condition is satisfied and t h a t m is O00 on M x M. We give to M x K the product of the O00 structures on M and K and choose a symmetric open set M± containing e such t h a t MXMX^M. For (x,k) and (y,k') e M1xK,
which shows t h a t the m a p (xjc), (y,k') -> (x,k)(y,k')_1 is (700 on (Mx xK)x (M-^xK). By a standard result in Lie group theory (cf. Che valley [1], pp. 134-135) there exists a O00 structure on Hm converting Hm into a Lie group, and such t h a t the induced C00 structure on MxK coincides with the product (700 structure around (e,l). I n this manner one converts Hm into a connected Lie group. Note, under the assumption t h a t m is O00 around (e,e), t h a t there exist open sets M2^G and K2<^K containing the respective identities, such t h a t M2 x K2 has the product (700 structure. Now, it is known (cf. Pontrjagin [1]) t h a t every C00 Lie group can be given a unique compatible analytic structure. Hm m a y be given this unique analytic structure. The final statement of the theorem follows from the known fact (cf. Chevalley [1], p p . 128-129) t h a t a continuous homomorphism of one analytic group into another is necessarily analytic. The proof of the theorem is complete. We can deduce several corollaries from this theorem. Corollary 7.22. The map z->(x,l) locally C™.
of G into Hm is C00 around e, m being
This follows from the fact t h a t the C°° structure of Hm is, around (e,l), the product structure. Corollary 7.23. Let m be an arbitrary K-multiplier for G. Then there exists a multiplier m! which is similar to m and is O00 {respectively analytic) on all of GxG if and only if there exists a 0°° (respectively analytic) map x->c(x) of G into Hm such that j0(c(x)) = x for all x, i.e., if and only if there exists a global O00 (respectively analytic) section. Proof. The proof is the same as for theorem 7.12; we need only replace " c o n t i n u o u s " by " ( 7 0 0 " and " a n a l y t i c " at the appropriate places. Corollary 7.24. Given any K-multiplier for G, there exists a similar to it which is analytic around (e,e).
multiplier
Proof. This requires the result that, for any analytic group, the coset space determined by a closed subgroup admits a local analytic section
264
GEOMETRY
OF QUANTUM
THEORY
(cf. Che valley [1], pp. 109-111). Applying this result to the analytic group Hm and the closed subgroup {e} x K —i0\K], we find a Borel section c, c : x -> (x,a(x)) such t h a t for some open set N around e, c is an analytic m a p on N • N. If we define
q(x,y) =
cixyMy)-1^)-1,
then q(x,y) = (e,ma(x,y) ~x) (cf. (46)) and q is an analytic m a p of N x N into Hm. Since k ~> (e,k) is an analytic isomorphism and since {e} x K is a closed analytic submanifold of Hm, we deduce t h a t ma is analytic on N xN. Obviously ma ~ m. This is as far as we can go with elementary methods. To obtain further, and in particular, global results, greater use must be made of Lie group theory. Our aim is to use some of the ideas of Hochschild [1], which in t u r n are based on a very interesting lemma of Malcev [1], to prove t h a t when G is connected and simply connected, there exists a canonical one-one correspondence between the Lie group extensions and the associated Lie algebra extensions. We shall moreover be able to show t h a t analytic multipliers exist in every similarity class and t h a t they are naturally induced by analytic K*-multipliers for G, K* being a covering group of K. We begin with some preliminary remarks. For any analytic homomorphism a of a Lie group A into a Lie group B, we write a for t h e induced homomorphism of the Lie algebras; a(exp J ) = e x p d ( Z ) for X e a, where a is the Lie algebra of A. Let b be the Lie algebra of B a n d let us assume t h a t A and B are both connected and simply connected. Consider now a homomorphism d(X —> d(X)) of the Lie algebra ft into t h e Lie algebra of all endomorphisms of the vector space a such t h a t d(X) is a derivation of a for all X, i.e., d(X)([Y,Z]) = [d(X)(Y),Z] + [Y,d(X)(Z)] for all Y, Z in a. Then there exists a unique analytic homomorphism b —> 8(b) of B into the Lie group of invertible linear transformations of the vector space a such t h a t S(exp X) = exp d(X) for all X eft. Since d(X) is a derivation of a, 8(exp X) is an automorphism of the Lie algebra a. Hence 8(6) is an automorphism of a for all b. The m a p 6, Y -> 8(b) Y is clearly analytic from B x a to a (a being a finite dimensional real vector space, can be considered in an obvious fashion as a real analytic manifold). Since A is simply connected, each 8(b) gives rise to a unique automorphism D(b) of A such t h a t D(b)(exp F) = exp{8(6)(7)} for all Yea. Since b - > 8(b) is a homomorphism, b -> D(b) is a homomorphism of B into the group of all analytic automorphisms of the Lie group A. We t h u s see t h a t A becomes a £-space. Moreover, the equation D(fe)(exp 7 ) = exp{8(6)F} shows t h a t the m a p b,a-^ D(b)a is an analytic m a p of BxA into A. Therefore if we form t h e semidirect product BxDA (cf. Chapter VI),
MULTIPLIERS
265
B xDA becomes a connected, simply connected Lie group. I t s analytic structure is just the product structure of B x A. As usual, we shall identify A and B with subgroups of B xDA; A will be a closed normal subgroup, and for b e B, aeA, bob'1 will coincide with D(b)a. Moreover, for Yea and Xeh, exp X exp Y exp( — Z ) = exp{exp(c?(Z))7}. This shows that, if we identify the Lie algebra of B xDA with the direct sum b-fa, then for Y e a and X e b, [X,Y] = d(X)Y. Lemma 7.25. Let Q be a Lie algebra over a field of characteristic zero and c^1an ideal in g. Suppose that either dim(a,l'g^ = 1 or g/gx is semisimple. Then gx is a semidirect summand of g, i.e., there exists a Lie subalgebra p such that g = gx + p and gx n p = 0. Proof. If dim(g/g 1 ) = l, then we may take for p any one-dimensional subspace not in g x . I n case g/gx is semisimple, the assertion is a refinement of the well known Levi-Malcev theorem and follows easily from it. Let r be the radical of g. Let 77- be the canonical m a p of g on g/g^ Since 7T[T] is a solvable ideal of g/g1? 77-[r] = 0, so t h a t r c i g ^ Let p' be a semisimple subalgebra of g such t h a t p' n r = 0 and p' + x = aj. Such a p' exists, by virtue of the Levi-Malcev theorem (cf. Chevalley [3], pp. 135-145). As p' n g! is an ideal in p', there exists a subalgebra p (which is even an ideal in p') such t h a t p + (p' n Q1) = p' and p n (p' n gi) = 0 (cf. Chevalley [3], pp. 64-65). p is semisimple and it is trivial to verify t h a t p n Qx=0 and £ + 9i = 9Lemma 7.26 (Malcev [1]). Let H be a connected and simply connected Lie group and K a closed connected normal subgroup of H. Let y be the canonical homomorphism of II onto II\K. Then there exists an analytic map c of IIIK into II such that (63) for all t
y(c(t))
E
= t
H/K.
Proof. We follow Hochschild's proof [1]. Let f) be the Lie algebra of H and ! the Lie subalgebra which corresponds to K. Since K is normal, ! is an ideal in t). Let l) = f)1^l)2^ • • • =5 f),=f be a composition series of f) over !, i.e., each f)i + 1 is an ideal in fy, and is maximal in f)t. We prove lemma 7.26 by induction on r, the length of the composition series. Since f)2 is maximal in t)1? either dim(^ 1 /t) 2 ) = 1 or fy1lt)2 ^ semisimple. Therefore, by lemma 7.25, there exists a subalgebra b of f)x such t h a t fyx = b + f)2 and b n I)2 = 0- For JC e b and 7 e | 2 , write (64)
d(Z)7 =
[X,Y].
Let i7 2 and B be connected simply connected groups corresponding to f)2 and b, respectively. I n view of our earlier remarks, the m a p X -> d(X)
266
GEOMETRY
OF QUANTUM
THEORY
gives rise to an analytic map b, a -> D(b)a of B x H2 into H2 such that H2 becomes a Z?-space and for each b e B, D(b) is an analytic automorphism of H2. We may therefore form the group B xDH2 and identify its Lie algebra with t}± in such a way that B x {eHJ and {eB} x /7 2 correspond to b and J)2, respectively (eB is the identity of B and eH2 that of H2). Now, i? xDH2 is simply connected and so is H. Hence we may identify H with B xDH2. In other words, we may assume that H2 and B are the analytic subgroups of H corresponding to f)2 and fc, respectively. Then H2 C\ B = {eH} (the identity of H) and H = H2B. Let H — HjK and let y be the canonical homomorphism of H onto IT. Put H2 = y(H2),B — y(B). Now, for each 6 G 5 , Z)(6) leaves K and i/ 2 invariant and so induces an automorphism D(b) o£H2\K. Clearly, 6, a -> jD(6)a is analytic from 2?x (H2jK) into # 2 /iL and so one can form BxD(H2/K). It is obvious that our identification of 11 with B xDH2, gives rise to a natural identification of H with B x^{H2jK). This proves that H2 and 5 are closed in H, that i/ 2 n 5 = {e5) and H = H2B. In particular, y is an isomorphism of B onto B. By the induction hypothesis, there is an analytic map c' (s -> c'(s)) of i/ 2 into # 2 such that y(c'(s)) = s for all s e H2. Define c on H by (65)
c(5y(6)) = c'(s)b
(s eH2,be
B).
Then c satisfies (63). Corollary 7.27. The map (66)
&, t - > c(*)ifc
*«5 aw analytic isomorphism of the manifold K x (H/K) onto the manifold H. Proof. Let £ denote this map. £ is analytic and one-one. For a e H, £ " » = (c(y(a))-*a,y(a)) which is clearly analytic. Corollary 7.28. Let f) be a Lie algebra over the reals, H a connected and simply connected Lie group corresponding to f). Let I be an ideal in f). Then the analytic subgroup K of H which corresponds to I is closed. Moreover both K and H/K are simply connected. Proof. That K is closed is well known (cf. Chevalley [1], pp. 125-127). Corollary 7.27 implies that H is (topologically) homeomorphic to K x H/K, and so the simple connectivity of H implies that of K and H/K. Remark. It is known that if K and H/K are simply connected, so is H (cf. Chevalley [1], pp. 59-60). Thus corollary 7.28, together with this result,
MULTIPLIERS completely solves the problem of simple connectivity for H, K, HjK, when K is closed, connected, and normal.
267
and
Lemma 7.29 (Hochschild [1]). Let G be a connected, simply connected Lie group and (H,i,j) an extension of the connected Lie group K by G. Then there exists an analytic map c of G into H such that (67)
j(c(g)) = g
for all g e G. Proof. Let H* be a connected and simply connected covering group of H and let 8 be the covering homomorphism of H* onto H. Then j o 8 is a homomorphism of H* onto G. Let Kx be the kernel of j o 8 and K^ the component of the identity of Kx. If K1°^K1, H^jK^ would provide a nontrivial covering for G, which is not possible since G is simply connected. Hence K1 = K1°, i.e., K1 itself is connected. From Malcev's lemma we know t h a t there exists an analytic map c* of G into H* such t h a t U ° S)(c%)) = g for all g e G. If we write c = 8 o c*, then c is analytic from G into # and satisfies (67) for all g e G Corollary 7.30. Let G be a connected, simply connected Lie group and K a connected abelian Lie group. If m is a K-multiplier for G, then there exists a K-multiplier m' similar to m which is analytic on GxG. Proof. Let Hm be the extension associated with m. From lemma 10.29 we know t h a t there exists a global analytic section from G into Hm. Then, corollary 7.23 implies t h a t there is a multiplier similar to m which is globally analytic over GxG. Remark. T h a t m' can be chosen to be C00 was proved b y Bargmann [1], We shall use the lemmas of Hochschild and Malcev to obtain a direct relation between central extensions of Lie groups and the corresponding central extensions of Lie algebras. Let !, g be Lie algebras over R with ! abelian. A central extension of ! by g is a triple (Jj,a,/?), where f) is a Lie algebra over R, a is an isomorphism of ! into the center of I), and j8 a homomorphism of f) onto q such t h a t a[f] = kernel (/3), i.e., (68)
[«[!],$] = 0
and (69)
0
> t - ^ f) - £ - • g — > 0
268
GEOMETRY
OF QUANTUM
THEORY
is an exact sequence. Central extensions (f),a,/5) and (f)'\a,/3') of f by g are said to be equivalent if there exists a (Lie) isomorphism £ of t) onto t)' such t h a t the diagram
(70)
f
_
^
^ _JL_> g
Y
>^
commutes. Suppose if is a connected abelian Lie group, G a connected Lie group, and (H,i,j) a central extension of K by G. Let !, f), and g be the Lie algebras of K, i / , and (7, respectively. Then it is obvious t h a t (fy,i,j) is a central extension of ! by g. We shall say t h a t (f),i,j) is the associated central extension of! by g. A routine transport of structure shows t h a t corresponding to equivalent central extensions of K by G, are associated equivalent central extensions of ! by g. Theorem 7.31. Let G, K be connected Lie groups and g, I the corresponding Lie algebras. Suppose that K is abelian and G is simply connected. For any equivalence class r of central extensions of K by G, let f denote the equivalence class of the associated central extensions of i by g. Then r -> f is a one-one correspondence and its range is the set of all equivalence classes of central extensions of I by g. Proof. We shall first give consideration to the case when K is also simply connected. Let (f),a,/3) be any central extension of f by g. Let H be a connected and simply connected Lie group which corresponds to f). Since K is simply connected, there exists an analytic homomorphism i of K into H such t h a t i = a; similarly there exists an analytic homomorphism j of H onto G such t h a t j = p. The range i[K] being an analytic subgroup of H having the ideal a[f] for its Lie algebra, corollary 7.28 implies t h a t i[K] is closed and simply connected. Hence i is also an isomorphism. On the other hand, i[K] is the connected component containing the identity of the kernel Kx of j . If i[K]^K1, H/i[K] will be a nontrivial covering of G^HjK^ which is impossible as G is simply connected. Therefore i[K] is precisely the kernel of j . I n other words, (H,i,j) is a central extension of K by G and (f),a,j3) is its associated extension. The uniqueness of i and j , once H is specified, implies at once t h a t two extensions of K by G, constructed in this manner, are equivalent if and only if their associated Lie algebra extensions are equivalent. Thus we have proved the lemma in the special case when K is simply connected. Suppose now t h a t K is connected but not simply connected. Let K* be a connected, simply connected Lie group covering K. Let e be the covering homomorphism. Let Z be the kernel of e. I n view of the result for
MULTIPLIERS
269
the simply connected case, it is enough to set up a natural correspondence between central extensions of K and K* (by G) which respects equivalence. Let (H*,i*,j*) be a central extension of K* by G. Then H* is simply connected. Let H' = H*'/i*Z and let 8 be the canonical homomorphism of H* onto H. Then 8 has discrete kernel. Since kernel (8)^i*[K*], there is a unique homomorphism j of H onto G such t h a t j * = ^ o 8. Since kernel (S) = {* [kernel (e)] there is a unique homomorphism i of K into i / such t h a t 8 o i* = i o £; i is then injective. I t is then obvious t h a t kernel (j) = i[K]. Thus the diagram H*
?
K* -^
\
• H
K
/
Q commutes. Moreover, (H,i,j) is a central extension of K by G. Suppose conversely t h a t (H,i,j) is a central extension of K by G. Let H* be a connected, simply connected Lie group which covers H. Let 8 be the covering homomorphism. Then there exists a unique homomorphism i* of K* into H* such t h a t 8 o i* = i o e. Write j * = j o 8. As G is simply connected, we conclude t h a t kernel (j*) is connected. From this it follows quickly t h a t kernel (j*) = i*[K*]. Hence i*[K*] is simply connected. Since K* is also simply connected, i* is injective. Hence (i/*,^*,j*) is a central extension of K* by G. Moreover, the diagram (71) commutes, and (72)
kernel (8) = i*[Z].
We have thus set up a natural correspondence between central extensions of K and K* by G. Routine diagram chasing now shows t h a t this correspondence respects equivalence. The proof of theorem 7.31 is complete. Let (H*,i*,j*) be a central central extension of K by G. onto H such t h a t (71) is a (H*,i*,j*) covers (through e)
extension of K* by G and let (H,i,j) be a Suppose there is a homomorphism 8 of H* commutative diagram. Then we say t h a t (H,i,j).
Corollary 7.32. Let G be connected and simply connected. Let K, K* and e be as above. Then the map m* -> e o m* maps the set of all analytic (respectively C°°) K*-multipliers onto the set of all analytic (respectively C°°)
270
GEOMETRY
OF QUANTUM
THEORY
K-multipliers, mx*~ m 2 * if and only if e o m x * ~ e o m2* so that m* —> e o m* induces an isomorphism of MK*(G) onto MK(G). Proof. Let m be an analytic if-multiplier for G. Let Hm be the associated extension. Then Hm has the product analytic structure. Since we can cover the extension Hm by an extension of K* by G, we m a y assume t h a t for some K*-multiplier m ^ , the extension Hm * covers Hm. I n view of corollary 7.30, we m a y assume t h a t rax* is analytic on Gx G so t h a t Hm* has the product analytic structure. Let $ = {((7,1) : g eG}^Hm. Let 8 be t h e covering homomorphism of Hm * onto Hm. Write 8* = 8~1(S). A standard topological argument shows t h a t 8 is a covering m a p of each connected component of 8* onto S. Let S0 be the component of 8* containing the identity (e,l) of Hmi*. T h e n $ 0 is a closed regular analytic submanifold of Hmi. Since 8 is simply connected, 8 is an analytic isomorphism of S0 onto Slm Hence there is an analytic m a p a of G into K* such t h a t for all g e G, (73)
8((gM9))) = (9,1).
Since 8 is a homomorphism, (73) gives (74)
8(xy,m1*(x,y)a(x)a(y))
Now, (xy,a(xy))~1(xy,a(x)a(y)m1*(x.y))
=
8((xy,a(xy)))(e,m(x,y)). is computed to be
so t h a t (74) leads to the equation
8 e,
( ^r mi ^ ,2/) ) = (e'w(a:'2/))'
and consequently,
(75)
Ho^T ^
for all x,y eG. If we write m*(x,y) applied, we have I7g\
/
=
for the element in (75) to which e is
m = e o m*
m* is clearly analytic. This proves the result in the analytic case. The O00 case can be disposed off similarly. The equation e o m* = m implies, on the other hand, t h a t the m a p x, k* -> x, e(k*) is a homomorphism oiHm. on i / m which sets up a covering of t h e extension Hm by Hm*. From theorem 7.31 we now conclude immediately t h a t m^ctm^ if and only if e om1*^.eom2:¥. From this it
MULTIPLIERS
271
follows at once t h a t m* -> e o m* induces an isomorphism of MK*(G) onto j r *(). Corollary 7.33. Let G be a connected, simply connected Lie group and K a connected abelian Lie group. If m is a K-multiplier, then m is exact if and only if, in the extension oftbyq which is associated with Hm, the image of I is a semidirect summand. Proof. Let (Ij,a,j8) be the extension of f by g which is associated with (Hm,i0,j0) (where i0(k) = (e,k) and j0(g,k) = g). If m is exact, there exists a Borel homomorphism c(x -> c(x)) of G into Hm such t h a t j0{c(x)) =x for all xeG (theorem 7.12). Since G and Hm are Lie groups, c is analytic. Hence the range B of c is an analytic subgroup of Hm. If b is the subalgebra of ^ corresponding to B, it is obvious t h a t b n a[t] = 0 and b + a[f] = 1). Conversely, suppose there exists a subalgebra b of I) such t h a t b + a[f] = f) and b n a[f] = 0. Then there exists obviously an isomorphism y of g into f) such t h a t y maps g onto b and /3 o y is the identity. Since G is simply connected, there is a homomorphism c of G into # m such t h a t c — y. Since /S o y is the identity, j 0 o c is locally the identity. Since G is connected, j 0 o c is the identity everywhere. This proves t h a t m is exact. The only problem which is still to be examined is t h a t of classifying the central extensions of ! by g. This can be done very simply as the problem is linear. We now proceed to indicate the standard solution to this problem (cf. Che valley-Eilenberg [1]). Let p be a bilinear map of g x g into ! such t h a t (i) p is skew symmetric, i.e., (77)
p(X,Y)+p(Y,X)
= 0
for all X, F e g , and (ii) p satisfies the identity (78)
p{[X9r\,Z)+p([Y,Z],X)+p([Z,XlY)
= 0
for all X, Y, Z e g. We shall associate with p a central extension of ! by g. Let *)P = 9 x 1 , and let us define, for (X,S) and (X',S') e fjp, (79)
[(X,S),(X',S')]
=
{[X,X'lp(X,X')).
From the identities (77) and (78) it follows t h a t f)p becomes a Lie algebra under the bracket operation (79). Moreover, (79) shows trivially t h a t (X,S) —> X is a homomorphism of t)p onto g, t h a t (0,$) lies in the center of fyp for all Set, and t h a t S -> (0,S) is an isomorphism of I into l) p . Write a0(8) = (0,S), (80) Po(X9S) = X. (f)p,a0iPo) is then a central extension of! by g.
GEOMETRY
272
OF QUANTUM
THEORY
Lemma 7.34. Every central extension ofI by g is equivalent to an extension I) j, for a suitable bilinear map pof§x§ into I satisfying (77) and (78). \)Pl and \)Vi are equivalent extensions if and only if there exists a linear map q of g into I such that (81) for all
p2(X,Y)-Pl(X,Y)
=
q([X,Y])
I J e g .
Proof. Let (f),a,j8) be a central extension of f by g. By identifying ! with <x[f] and g with a linear manifold supplementary to «[!], we m a y assume t h a t t) = g x f , a = a0> a n ( i P = Po- We have now only to determine the form of [.,.] for pairs of elements of f). Since (X,S)}(X',S') -> [(X,S),(X',S')] is bilinear and skew symmetric and since (X,8) - > X is a homomorphism of 1) onto g, it follows t h a t there exists a skew symmetric bilinear m a p of f) x fy into f, say r, such t h a t (82)
[(X,S),(X',S')]
= ([X,X'],r(X,S;
X',S'))
for all X, X' e g and S,S' et. The fact t h a t (0,#) lies in the center of ^ for all S in ! now implies t h a t (83)
r(0,£; X',S')
= 0
for all I ' e g and $, $ ' e !. The bilinear maps of f) x f) into I which are skew symmetric and satisfy (83) depend only on X and X'. Thus there exists a skew symmetric bilinear map p of g x g into ! such t h a t r(X,S; X',S') = p(X,X') for allJT, X' e g and £, £ ' e t. We have then (84)
[(X,S),(X',S')]
=
([X,X'],p(X,X')).
If we now use the fact t h a t [.,. ] satisfies the Jacobi identities, we see t h a t (78) is satisfied b}^ p. I n other words, the extension t) is equivalent to $p. Suppose now t h a t f)p and t)P2 are two extensions of ! by g corresponding to the bilinear maps p± and p2. I t is easily seen t h a t for any linear m a p q of g into !, the m a p (85)
Lq:(X,S)->(X,q(X)+S)
is a linear m a p of f)Pl into fyP2, sends (0,S) to (0,8), and is such t h a t the diagram *Pt — l J - +
*>»*
(86)
is commutative; and t h a t every linear m a p L of fyPl into fyP2, which sends (0,S) to (0,$), and which is such t h a t (86) is commutative (with L
MULTIPLIERS
273
instead of Lq), is of the form Lq for suitable q. I t is then clear t h a t Lq is a linear isomorphism of f)Pl onto f)P2. A trivial calculation shows t h a t Lq is a homomorphism of the Lie algebras t)Pl and fyP2 if and only if plt p2 and q are related as in (81). This completes the proof of the lemma. We shall call a skew symmetric bilinear m a p p of g x g into ! closed if it satisfies (78). We shall say t h a t p is exact if there exists a linear m a p q of g into ! such t h a t (87)
3J(Z,r)
= gr([Z,r])
(IJe
9
).
The vector space of all closed skew symmetric bilinear maps of g x g into ! is denoted by W{(Q), and the subspace of exact elements is denoted by @f(g). We write (88)
Wt(q) = W(0)/@f(g).
Corollary 7.35. For apeSDfy'(g), p is exact if and only if {0} x I is a semidirect summand off)v. Proof. If p is defined by (87) for some linear m a p q of g into f, it is immediate t h a t the range h of the m a p X -> (X,q(X)) is a subalgebra of f)p such t h a t a0[f] n b = 0 and a0[f] + b = f)p. Conversely, let b be such a Lie subalgebra. Then for each l e g , let us define the element q(X) e I by the condition t h a t (X,q(X)) eh. q is evidently a linear m a p of g into I, and the condition \(X,q(X)),(Y,q(Y))] = ([X,Y],q([X,Y])) leads to (87). The theory developed so far has established a canonical isomorphism between the groups MK(G) and MK>(G), and also a canonical one-one correspondence between MK*(G) and 9Qfy(g) (G simply connected). Now 5D^i(g) is a vector space. On the other hand, if we write K* additively, it is isomorphic to a vector group, so t h a t MK*(G) also becomes a vector space, with addition as the group operation in it. I t is indeed most natural to expect t h a t the canonical correspondence between the elements of MK*(G) and Wt{o) is actually a linear isomorphism. We shall now prove t h a t this is indeed so. We assume K* = Rn. The i? n -multipliers for G then satisfy (14) and form a vector space over R, and the exact ones form a subspace of it. MRn(G) is thus a vector space. Lemma 7.36. Let G be connected and simply connected. For any analytic En-multiplier m for G, let [m] be the similarity class of m. Let 6[m] be the element of %RR*(Q) which corresponds to the equivalence class of central extensions of Rn by G which are associated with [m]. Then (89) is a linear isomorphism
M-&Im] of MRn(G) onto $0lfl»(g).
274
GEOMETRY
OF QUANTUM
THEORY
Proof. Let m be an analytic i^-multiplier for G and let Hm be the associated group extension. For (xjc) and (y,k') e Hm ( = GxmRn in our earlier notation), (x,k)(y,kf) = (xy,k + k' + m(x,y)). Now (e,0) is the identity of II m, and Hm has the product analytic structure. We shall now determine the Lie algebra f) of Hm. Following Chevalley [1] we regard elements of f) as left invariant vector fields on Hm\ for each (x,k) e Hm, we identify the tangent space to Hm canonically with Gx x Rn, Gx being the tangent space to G at x. Of course we identify the tangent space to Rn at any of its points with Rn itself in the usual fashion. Let l e g b e the left invariant vector field x -> Xx on G and let u e Rn. We write (X,u) for the left invariant vector field of Hm which assigns to (e,0) the tangent vector (Xe,u). Since left translation by (x,k) is (y,k') —> (xy, Jc + k' + m(x,y)), we see that the tangent vector assigned by (X,u) at (x,k) is given by the expression Xxiu +1 ~r± m(x, exp tX) J Write now for any C00 function / on G x G and Ylt • • •, Yr, Z±, • • •, Zs e \ y,zeG, (90)
f{y; Y^-
Yr: z; Zv .. Zs)
J
= (a r+ v^i- • • d ^ r • .a^s)tl = ...=fr=Ui = ...=Ws=0,
where = Kv^VhYi'
• ex^trYr,zexi^u1Z1-
• -exp^ s Z s ).
Then (X,u) is given by (91)
(X,u) :x,k-+
(Xx, u + m(x : e; X)).
From (91) we find (92) [(X,u),(Y,v)] : x, k-> ([X,Y]X, m(x; X: e; Y)-m(x;
Y : e; X)).
From the identities (14) we obtain, on differentiation, m(x; X :e;Y)
= m{x : e; XY) + m(e; X : e; Y).
This then leads to (93)
\(X,u),(Y,v)]
:x,k->
([X,Y]X, m(x:e;[X,Y])
+ m(e;X : e; Y)-m(e;
Y : e; X))
so that finally, on comparing (93) with (91), we obtain (94)
[(X9u)9(Y,v)] = ([X,Y], m(e; X : e; Y)-m(e;
Y : e; X)).
MULTIPLIERS
275
I n other words, if we write (95)
Bm(X,Y)
= m(e; X : e; Y)-m(e;
Y : e; X)
then the Lie algebra I) of Hm can be identified with g x Rn in a canonical fashion and (96)
[(X,u),(Y,v)]
= ([Z, 7 ] , i U X 7)).
J5m is evidently a skew symmetric bilinear m a p of g x g into Rn. The Jacobi identity applied to (96) leads as before to the fact t h a t Bm is a closed form. We know already t h a t [ra] -> b[m] is a one-one correspondence of MRn(G) onto 9#fl"(g), and (96) shows t h a t 6[m] is the element of 9ftB*(g) defined by Bm. B u t a look at (95) shows t h a t Bm(X,Y) depends linearly on m, for fixed X, F e g . Consequently, [m] -> fe[m] is a linear isomorphism of MRn(G) onto $ft#»(g). This proves the lemma. Collecting all these results together we obtain the following theorem. We write G for the Lie group we have been concerned with. K is a connected abelian Lie group, K* a connected, simply connected covering group of K, and e is a covering homomorphism of K* onto K. Theorem 7.37. Let G be a connected and simply connected Lie group, and let K, K*, and e be defined as above. Then any K-multiplier for G is similar to one which is analytic on GxG. For any analytic (respectively (700) K*-multiplier m*, e o m* is an analytic (respectively C°°) K-multiplier for G, and every analytic (respectively O00) K-multiplier can be represented in this form; m x * and m 2 * are similar if and only if e o rax* and e o m 2 * are similar. If we associate with any equivalence class of central extensions of K by G the equivalence class of associated extensions of I by g, we obtain a one-one correspondence whose range is the set of all equivalence classes of central extensions of I by g. These latter classes are in canonical one-one correspondence with elements of the space Tlt(q). If we write K* additively and treat it as a real vector space, MK*(G) becomes a real vector space and the induced correspondence of MK*(G) with 9ttf(g) is a linear isomorphism. Finally, a K-multiplier G is exact if and only if, in the associated extension ofibyQ, the image of t is a semidirect summand.
5. E X A M P L E S We shall now discuss a number of examples which illuminate various aspects of the theory described in Sections 1 through 4. We restrict ourselves to the case of (ordinary) multipliers and R1-multipliers.
276
GEOMETRY
OF QUANTUM
THEORY
Covering Groups: There is no direct connection between simple connectedness and exactness. We shall give, further on, examples of connected b u t not simply connected Lie groups admitting only exact multipliers, and of simply connected Lie groups whose multiplier groups are not trivial. There are, however, some elementary relations between multipliers and representations of a group and its covering group t h a t m a y be worthwhile putting down. Let G be a connected Lie group, G* a connected and simply connected covering group of G, and 8 the covering homomorphism. Let x* - » Vx* be an irreducible unitary representation of G* in a Hilbert space Jf. By Schur's lemma, V maps every element in the center of 6* into a scalar multiple of the identity. Since the kernel Z of 8 is contained in the center of G*, it follows at once t h a t V2*x* = x(z*)Vx. f ° r an< z * e ^ a n d #* e G*, X being a character of Z. This implies t h a t for any g e G, all the operators Vg* with 8(g*) = g, define the same element of the projective group of Jf. I n this way, each irreducible representation of G* induces a projective representation of G. The relation between G and G* is also useful in another way. Let m be a multiplier for G and let m* be the multiplier x*, y* -> m(8(x*),8(y*)). It is then easily seen t h a t m -> m* induces a homomorphism, say h, of M(G) into M(G*). Since G* is simply connected, theorem 7.37 can be applied to the study of M(G*). In this way useful global insights into the structure of M(G) may be obtained. However, it must be remembered t h a t h is in general neither injective nor surjective, so t h a t , in this approach to the analysis of M(G) in terms of M(G*), the kernel and the range of h must both be determined. For example, if G is a torus, M(G) is trivial while M(G*) is nontrivial as soon as dim 6 r * > l ; if G is semisimple, M(G*) is trivial, but M(G) will not be (cf. supra). 6r = R5, a vector group in s-dimensions. I n this case we can explicitly describe the multiplier group (cf. also Weyl [1], pp. 273-274). We formulate our conclusions in the form of the following theorem. I t is an easy consequence of the results described in theorem 7.37. Theorem 7.38. For any real skew symmetric bilinear form p on Rs x R 8 , (97)
mp : x, y -> exp
ip(x,y)
is a multiplier for Rs. The mapping which assigns to p, the similarity class of mp, is an isomorphism of the additive group of all skew symmetric bilinear forms on RsxRs onto the multiplier group M(Rs). In particular, all multipliers of R 1 are exact. Proof. I t is trivial to check t h a t mp is a multiplier for any bilinear form p on R*xR* (skew symmetric or not) and t h a t m1>1+3,a = m J , l m 3 , a .
MULTIPLIERS
277
To prove the remainder of the theorem we pass to the covering group R 1 of T and use theorem 7.37. Now, for any bilinear form p,x,y-^p(xiy) is an R 1 -multiplier for Rs and it is exact if and only if it is symmetric (corollary 7.14). I n particular, px^p2 ^ a n d onty # P i ~ P 2 *s symmetric. If Px and p2 are skew symmetric, it is now obvious t h a t px ~p2, if and only ifPi —Pi- If P is a n y bilinear form, we can always write it as px + c, where px is skew symmetric and c is symmetric; then p~p±. These conclusions then carry over to the T-multipliers mv. In order to complete the proof we go back to (95). We write Bp = Bm and prove t h a t the mapping p -> Bp has the set of all skew symmetric bilinear maps of R s x Rs into R 1 for its range (we identify g with R 5 also in the obvious fashion so t h a t exp is the identity map). This is trivial, for, by (95), - iB9(X, Y) = 2p(X, Y)
(X,Ye
R-),
whenever p is a skew symmetric bilinear form on R s x R 8 . The last assertion follows trivially from the fact t h a t 0 is the only skew symmetric bilinear form on R1 x R1. Connected Semisimple Lie Groups: Suppose q is a semisimple Lie algebra over the reals and (fy,a,p) a central extension by it of the abelian Lie algebra !. Since f)/a[f] is semisimple, it follows at once t h a t a[f] is the radical of f). By the Levy-Malcev theorem we conclude t h a t there exists a subalgebra b of ^ such t h a t a[f] + b = I) and «[!] n h = 0. This observation also leads a t once t o the conclusion (cf. theorem 7.37) t h a t if 0 is a connected and simply connected semisimple Lie group, then every multiplier for G is exact. Since the classical groups are semisimple, this result applies to the universal covering groups of these groups. For example, we m a y take GQ to be the connected component of the group of all invertible linear transformations of an s-dimensional vector space leaving invariant a nonsingular symmetric bilinear form. When s>2, G0 is known to be a semisimple Lie group. If G denotes a universal covering group of G0, then all multipliers for G are exact. The examples of greatest interest from the physical point of view arise when 5 = 3 and the bilinear form has a positive definite quadratic form, and when 5 = 4 and the bilinear form has a Minkowskian quadratic form. I n other words, the universal covering groups of the rotation and Lorentz groups (connected) admit only exact multipliers. Inhomogeneous Groups: The symmetry groups of space time cannot be semisimple because they contain the space-time translations as a normal subgroup. Therefore, the results derived in the previous example are inadequate for physical applications and it is necessary to extend their scope. We shall give here one such extension.
278
GEOMETRY
OF QUANTUM
THEORY
Let G be a connected Lie group and let p(x -> p(x)) be a (not necessarily unitary) representation of G in a finite dimensional real vector space V. The m a p x, v -> p(x)v is analytic from G x V into V and converts F into a 6r-space. The semidirect product (98)
Gp
=
Gx0V
can then be formed. Gp (in the product analytic structure) is a connected Lie group; if G is simply connected, it is a simply connected Lie group. We identify V and G with closed subgroups of Gp. V is normal a n d for x G G, veV, xvx~1 — p(x)v. We shall call Gp the inhomogeneous group associated with G and p. Our purpose now is to prove a theorem from which it follows t h a t under certain conditions the covering group of the inhomogeneous group admits only exact multipliers. We shall call the representation p of G in a real vector space V admissible if no nonzero skew symmetric bilinear form on V x V is invariant under p. The fact t h a t m a n y representations of a group are admissible is contained in the following lemma. Since we work with real vector spaces it has to be formulated in the context of absolutely irreducible representations. A representation n(x -> 7T(X)) of G in a real vector space W is said to be absolutely irreducible if the only endomorphisms of W which commute with all the 7T(X) are scalar multiples of the identity. If G is semisimple, absolute irreducibility implies irreducibility in the usual geometric sense and absolute irreducibility of n is equivalent to the irreducibility of 77 in the complexification of W (G semisimple). The classical Schur's lemma can be formulated for absolutely irreducible representations as follows: if n and IT are absolutely irreducible representations of a semisimple group G, acting in W and W, then the vector space of linear maps L from W into W such t h a t LTT(X) = 7T'(X)L for all x e G is of dimension < 1, the dimension being 1 if and only if n and 7/ are equivalent. Lemma 7.39. Let p be a representation of a connected semisimple Lie group G in V. Suppose that V is a direct sum of p-invariant subspaces V1,- • •, Vs such that for each i, (i) the subrepresentation pt (of p) defined by V{ is absolutely irreducible and (ii) there exists a nonzero symmetric bilinear form on Vt x Vi which is invariant under p{. If the representations px, p2,- • •, ps are mutually inequivalent, then p is admissible. Proof. Let V* be the dual of V and for x e G and teV*, let p*(x)t be t h e element v -> t(p(x~1)v) of V*. Then p*(x)(t -> p*(x)t) is easily verified to be an invertible endomorphism of V*. I t is also easy to verify t h a t p* : x -> p*{x) is a representation of G in V*, the so-called contragredient representation acting in V*. Now any bilinear form B on V x V gives rise
MULTIPLIERS
279
to a linear transformation TB of V into F*, and B is invariant if and only if TB intertwines p and p*, i.e., (99)
TBP(x) = P*(x) TB
for all x e G. Let Fi*<= F * be the annihilator of 2y#i ^ r Then p* leaves each Vt* invariant. From the fact t h a t pt leaves a nonzero bilinear form on V{ x Vt invariant, it follows t h a t there exists a nonzero linear m a p Tu mapping V{ into F f * and F y into 0 for j / i , such t h a t T{ intertwines p and p*. From the standard form of Schur's lemma we conclude, using the mutually inequivalent nature of pl9- • •, ps, t h a t the operators JT 1; • • •, Ts span the vector space of operators which intertwine p and p*. This implies t h a t every operator which intertwines p and p* is a TB for a symmetric bilinear form B on V x V. Hence any bilinear form on F x F left invariant by p is necessarily symmetric. Consequently, p is admissible. Theorem 7.40. Let G be a connected semisimple Lie group and p an admissible representation of G in a real finite dimensional vector space. Then, for the universal covering group Gp* of Gp, every multiplier is exact. In particular, if a(x -> ax) is a representation of Gp into the group of automorphisms of the logic of a {complex separable) Hilbert space ^ there exists a representation U(x* - » Ux*) of G* in J4? such that for each x in G, <xx is induced by Ux* for any x* in Gp* lying above x. If p does not contain the trivial representation as a subrepresentation, U is the only representation of Gp* with this property. Proof. Let us first describe Gp* in a specific manner. connected and simply connected covering group of G and be the covering homomorphism. Then p o e = p* : x* -> representation of G* in V. p* is obviously admissible and group (100)
Gp* =
Let G* be a let e(x* -> x) p(e(x*)) is a we form the
G*xp.V.
I t is clear t h a t Gp* is connected and simply connected and the m a p x*, v -> e(#*), v is the covering map of Gp* on Gp (we identify V and G* with subgroups of 6rp* as usual). If a is the given representation of Gp in the group of automorphisms of the logic of ^ we know (cf. footnote in Section 1) t h a t there is a projective representation V1 of Gp in J f which induces a. We lift V1 to the projective representation V : x*9 v -> V\ix.)tV of Gp*. We now apply theorem 7.5. I n order t h a t a be induced by a representation of Gp* in the manner described in the theorem, it is necessary and sufficient t h a t there is a unitary representation of Gp* which gives rise to the same representation as V in the projective group of Jf; clearly this will happen if and only if the multiplier of V is exact. Also, the representation of Gp* which induces a is unique up to
280
GEOMETRY
OF QUANTUM
THEORY
multiplication by a continuous homomorphism of Gp* into the group T (corollary 7.7). Thus, in order to prove the assertions of the theorem, it is enough to prove two things: first, that all multipliers for Gp* are exact, and second, that Gp* admits no nontrivial continuous homomorphisms into T under the assumptions made on p. We take up the first point. Let g, fc, and gp be the Lie algebras of G*, V, and Gp*, respectively, t) is an ideal in §p, g a subalgebra in gp, and t> n g = 0, *> + g = 8,.
If p is the representation of g in V that corresponds to the representation p* of 6r*, we shall, by using the canonical isomorphism of t> on V, assume that p and p act also on b. Thus, for Z e t) and l e g , (101)
[X,Z] = p(Z)(Z);
this corresponds to the relation x*vx*~1 = p(e(x*))v (x* e 6?*, v e V). In order to prove that all multipliers for G0* are exact, it is sufficient to prove that iff is any abelian Lie algebra and (fy,p,q) a central extension of ! by gp, then p[l] is a semidirect summand of t). We shall now demonstrate this. We write t>' = g~1(t))
(102) and (103)
fl'
= g-i(8).
q maps g' onto g and the kernel of q is p[l]. Hence we conclude from lemma 7.25 that there exists a subalgebra a such that ar\p[f]=0 and ct+^[f] = g'. q is an isomorphism of a onto g, so that a is semisimple. Since b n g = 0, t>' n a^p[l] so that i>' n ct = 0. In other words, (104)
t ) ' n a = 0,
Now b' is an ideal in I), and so for any X e a, (ad X)(Z) = [X,Z] lies in t)' whenever Z e b ' . Thus Z -> ad Z defines a representation of a in t)'. As p[t] is central, ad X annihilates all the elements of p[t]. Therefore, as a is semisimple, there exists, by Weyl's theorem (cf. Chevalley [3], pp. 70-73) a subspace bx such that f)1 n p[t] = 0, h1 +p[l] = t)r, and h1 is invariant under all ad X, X e a. Let (105)
b = bi + a.
b is a linear manifold, b n p [ ! ] = 0, and &+#[!] = *). O u r claim that ^)[!] is a semidirect summand will be proved if we show that b is a subalgebra
281
MULTIPLIERS
of f). Since [X,7] e ^ whenever X e a and 7 e b 1 ? and since a is a subalgebra, we have only to show that [X, 7] e bx whenever X, 7 are in h±. We shall actually prove that [X,Y] = 0 for X, Y in h±. Now g maps hx onto t), so that the abelian nature of t> implies that, for X, Y e &1? [ X , 7 ] E ^ P ] - We write (106)
#i(X,7) = ^ ( [ X - H )
(X,
YehJ.
Kx is evidently a skew symmetric bilinear map of hx x bi into I. If Z e a and X, 7 e b l5 [X,[7,Z]] + [7,[Z,X]] = 0, since this element is nothing but — [Z,[X,7]], and hence 0 as [X, 7] belongs to the center of f). Therefore,
(107)
x1([z,:n,x)+ir1(r,[z,xi) = o.
Now q is an isomorphism of the Lie algebra a onto g, and the vector space bx onto t>. Therefore there exists a unique skew symmetric bilinear map K of t) x t> into ! such that (108)
i ^ X ' , r ) = K(q(X'),q( Y'))
(X', Y' e bx)
and (109)
K([Z97],X) + K(Y,[Z,X])
= 0
for all X, 7 in fc and Z in g. In other words, (110)
K{p(Z)T,X) + K{7,p(Z)X)
= 0.
Since (110) is valid for all Z in g, we conclude that K is invariant under the representation p. If we choose a basis {^4l3 • • •, An} for I and write
K(X9T) = J ^ ( i ^ M i , then each j£t is an invariant skew symmetric bilinear form on t)xk). Each K% is 0, as p is admissible. This implies that K = 0. Therefore, Kx = 0. This proves that ft = h1 + a is a subalgebra of I). In other words, ^>[f] is a semidirect summand of f). We finally come to the second point, namely, to the proof that under the conditions mentioned in the theorem Gp* admits no one dimensional representation other than the trivial one. If it does, gp = t> + g will admit a nontrivial one-dimensional representation. Therefore there will exist a (complex valued) nonzero linear function A on qp such that A([X,7])=0 for all X, 7 in g^. Since g is semisimple, every element of g is a linear combination of elements of the form [X,7] with X, 7 in g (cf. Chevalley [3], p. 67) so that A = 0 on g. Let A' be the restriction of A toto.Then the equation A([Z,X]) = 0, for X e t) and Z e $, shows that the null space of A' is invariant under p. Since p is semisimple, we can select a subspace complementary to the null space of A' and left invariant by p. This space is
282
GEOMETRY
OF QUANTUM
THEORY
one-dimensional and so, as g is semisimple, we obtain a nonzero vector fixed by p. This is a contradiction. The proof of the theorem is now complete. This is a basic theorem insofar as relativistic applications are concerned. I t tells us, for example, t h a t any representation of the (connected) inhomogeneous Lorentz group can be obtained from a unitary representation of the covering group. To see this, all we have to do is take V as 224, G as the connected component of the group of Lorentz matrices acting on V, and x -> p(x) as the action of G in V. I t is quite easy to check t h a t p is admissible. I n this special case, theorem 7.40 was obtained by Wigner [1] and, in his work, this theorem was the point of departure for his classification of relativistic wave equations. I n the special case when G is the group of all invertible linear transformations of V leaving a nondegenerate quadratic form on V invariant (dim V>2), theorem 7.40 was obtained b y Bargmann [1]. Compact Groups. Much is known about multipliers for compact groups (cf. Moore [1]). We shall content ourselves with the proof of a simple result. Theorem 7.41. Let G be any compact group satisfying the second axiom of countability. Then its multiplier group is a torsion group, i.e., every element of M(G) is of finite order. If G is a torus, or a connected and simply connected compact Lie group, G admits only exact multipliers. Proof. Let m be a multiplier for finite dimensional m-representation £ - > £ 1/n denote a Borel nth. root of which fixes 1 and whose Tith power being the dimension of V. Then
G. By p(x-> £, i.e., is the
p' : x -> (det
corollary 7.18 there exists a p(x)), acting in V say. Let a Borel m a p of T into itself coordinate function £ - > £, n
p(x))llnp(x)
is a projective representation of G; and p defines the same homomorphism of G into the projective group of V, as p. If m! is the multiplier of p , then m and m' are similar. But det p'{x) = 1 for all x so t h a t for all x, y in G, [m'(x,y)]n = l. Hence (m)n is exact, proving t h a t a n y element of M(G) has finite order. If G is a connected and simply connected compact Lie group, then G is semisimple by a theorem of H. Weyl (cf. Helgason [1]), so t h a t G admits only exact multipliers. Suppose finally t h a t G is a torus, say G=Tr. Let e : v - > e(v) be the covering homomorphism from R r onto Tr. If m is any multiplier for G, ms;v, v' -> m(e(v),e(v'))m a multiplier for R r . Since ms is exact from some integer s > 1, so is mes. B u t the multiplier group of R r is canonically iso-
MULTIPLIERS 1
283
r
morphic to the R -multiplier group of R which is a vector group. Hence mes cannot be exact unless me is itself exact. Therefore, there is a Borel function a from R r to T such th.&tme(v9v')=-a(v+v')/a(y)a(v'). Since e is a local group isomorphism, we can find a Borel function & on Tr, with values in T, and an open set N of Tr containing the identity, such t h a t (111)
m(x9y) =
b(Xy)
b(x)b(y)
for all x, y e N x N. Equation (111) implies t h a t m is exact (cf. corollary 7.14). Groups Which Are Not Semisimple. If the Lie group G is not semisimple, then it m a y admit nonexact multipliers. We shall now sketch the construction of the multiplier group for the inhomogeneous Galilean group. We closely follow Bargmann [1] in this discussion. If (x,t) (x = xl9 x2, x3) denotes the coordinates of space-time points, then an inhomogeneous Galilean transformation is a m a p (112)
r
:x,t->x',t'9
where x' = Wx + tv + u, (U3)
f ,4. t = t + rj, W being a rotation in 3-space (det W= + 1 ) , u, v vectors in 3-space, and iq a real number. The set of all such transformations forms, under composition, the so-called inhomogeneous Galilean group; we have, for the multiplication in this group, (114)
(W,ri,v9u)(W'9ri',v'9u')
= (WW9r] + v' , W + v9Wu' + u + r)'v).
If we associate with r the matrix
(
W
v
0
1
u\ 7)
0 0 1/ then r -> M(r) is a faithful representation. The Galilean group is a semidirect product; the subgroup A of all translations in space-time, (116)
A = {(MAtO}
is normal, the subgroup G0 of transformations (117)
G0 =
{{Wfl,v,0)}
284
GEOMETRY
OF QUANTUM
THEORY
is a closed subgroup; A and GQ generate the whole inhomogeneous group. In fact, (118)
(l,rj,0,u)(W,0,v,0) = (W,rj,v,u)
and (119)
(W,0,v,0)(l,rj,0,u)( Wfi,vfi) - 1 == (1,77,0,^ + ^ ) ,
which exhibits explicitly the semidirect product nature. We denote the inhomogeneous group by G. The subgroup GQ, unlike the Lorentz group, is not semisimple, and this accounts for some of the more subtle differences between the Galilean theory and Lorentzian theory. Let M and H denote the subgroups of G0 given by (120)
M = {(1,0,^,0)},
(121)
H = {(W ,0,0,0)}.
M is a closed normal subgroup of G0, H is a subgroup with G0 = MH; moreover, (122)
(lf,0,0,0)(l,0,v,0)(lf,0,0,0)- 1 =
{l,0,Wvfi),
so that G0 is once again a semidirect product, with respect to the action of H in M given by (122). Note also that the space motions of the form (W,0,0,u) constitute the subgroup S = UH, where U is the subgroup (of all translations in space) {(1,0,0,^)}. U is a normal subgroup of S and (123)
(PT,0,0,0)(l,0,0,^)(lf,0,0,0)-1 =
(l,0fi,Wu);
therefore once again we have a semidirect product. The first step in determining the multiplier group of G is to examine the central extensions at the Lie algebra level. Theorem 7.37 reduces this to the question of determining the closed skew symmetric bilinear forms on g x g, modulo the exact ones. We shall now do this. As is customary, we denote the subalgebras of g by the German letters which correspond to the Latin letters that denote the analytic subgroups of G. The relations (116) through (123) give considerable insight into the structure of g. a is an ideal in g, g = ct + g0, a n g0 = 0; m is an ideal in g0, g0 = tn + ^), m n f) = 0; g0 and fy are subalgebras of g. Also 3 = u-|-f) is a subalgebra, u is an ideal in §, u n 1) = 0. Since A, M, and U are abelian and since H acts on all of them via the inner automorphisms, H acts naturally on a, m, and u. The differential of this action is the usual adjoint action of f) on a, m, u. Thus, using the standard notation according to which [p,p'] is the linear space spanned by all elements of the form [X, Y] with X ep and Yep', (124)
[f),m] ^ m,
[i),u] c u,
MULTIPLIERS
285
and (125)
[5,tf = 0,
where p is the one-dimensional subalgebra of g which corresponds to the subgroup {(1,^,0,0)} of A. Moreover, from (122) and (123) we see t h a t the representations of H in m and u are equivalent and absolutely irreducible and moreover, leave the Euclidean inner product on these invariant; these are therefore admissible by lemma 7.39. By a direct calculation or by use of the irreducibihty of these representations of f), we find (126)
[%m\ = m,
[$,u] = u.
Moreover an easy computation shows t h a t (127)
|>,m] c u,
[u,m] = 0.
Let B be any closed skew symmetric bilinear form on g x g . Then, for X, Y e a and Z e g 0 , we have, as [X,Y] = 0, (128)
B([Z,X], Y) + B(X9[Z, Y]) = 0.
Equation (128) shows t h a t B is invariant under the representation of G0 in a. Since the action of G0 on A leaves no nonzero bilinear form on Ax A invariant, (128) implies at once t h a t B = 0 on a x a. Note t h a t this has been proved for any closed form B. Next, we observe t h a t g 0 = m + l) satisfies the conditions of theorem 7.40. I n fact, t h e group H is semisimple and its action on M (which is isomorphic to 3-space) as given by (122) is, as observed by us already, admissible. Hence the restriction of B to g 0 x g 0 is exact. Replacing B by a form which differs from B by an exact one, we m a y t h u s assume t h a t (129)
B = 0
on
ax a
and
g0xg0.
We shall study the form B on axq0. Now a + ^ is a Lie algebra and a is an ideal in it. Moreover, the action of H on A, as given b y (119), is admissible by lemma 7.39, since A decomposes into a direct sum of U and the one-dimensional space of time translations. Hence theorem 7.40 applies once again and we conclude t h a t the restriction of B to (ct + f)) x (ct + f)) is exact. Therefore, there exists a linear function q on g, vanishing on m, such t h a t B(X,Y) = q([X,Y]) for all X, Y in ct + I). Since t) is semisimple, [^,^)] = ^ so t h a t (129) implies t h a t q vanishes also on f). If we define B' by (130)
B'(X,Y)
= B(X,Y)-q([X,Y])
(X,
7Ggxg),
then B — B' is exact and (131)
B' = 0
on
ctxct, ctxf),
and
g0 x g0.
286
GEOMETRY
OF QUANTUM
THEORY
If X G p, Y em and Z efy, the identity B'{X,[Z,Y])
+ B'(Z,[Y,X])
+ B'(Y,[X,Z])
= 0
implies, in view of (131) and the relations [Y,X] e u, [X,Z] = 0 (cf. (127)), that (132)
B'(X,[Z,Y])
= 0.
From (126) and (132) we conclude t h a t (133)
B' = 0
on
pxm.
We are now ready to prove t h a t the set of closed forms, modulo the set of exact forms, is one dimensional. To this end, we must analyze the restriction of B' to it x m. Now, for Z e f), we have, as [u,m] = 0, (134)
B'([Z,X], Y) + B'(X,[Z, Y]) = 0
for all X, Y e it x m. Since H induces equivalent representations on it and rrt, the vector space of bilinear forms o n u x m has a one-dimensional subspace of bilinear forms invariant under the natural action of H. Let C0 be a basis for this space. Let C be defined as the unique symmetric bilinear form on g x g such t h a t C=C0 on u x m and (7 = 0 on all of ax a, Qox9o> ctxf), and pxm. Then (131), (133), and (134) show t h a t B' is a multiple of C. I n other words, the vector space of closed forms modulo t h e subspace of exact forms has dimension < 1. On the other hand, it is easily verified t h a t C is actually closed. If C were exact, and C(X, Y) = q([X,Y]) for all X and Y in g, then the equation [it,m] = 0 would imply t h a t (7 = 0 on it x tn, a contradiction. Thus the space WIRI(Q) has dimension exactly 1. Therefore there exists a nonexact ^ - m u l t i p l i e r m 0 * for the covering group G* of G such t h a t every multiplier for G* is similar to one of the form x*, y* —>• exp[trm 0 *(x*,i/*)] for a suitable real number T. Actually m 0 * can be chosen to induce a nonexact ^ - m u l t i p l i e r mQ for G itself. Bargmann [1] has indicated a simple method for constructing m 0 . Let us write i = (W,rj), ^ = (v,u). Then we can expand m 0 as a series 2n
287
MULTIPLIERS
form and write down the multiplier identities, we obtain equations which are quite easy to solve. We find (135)
m0(r,r') = <^, W > - < v , W > + T/
where <.,. > denotes the Euclidean inner product on 3-space and (136)
r = (W,rj,v,u),
r' = (W
J>'>')•
It is of course necessary to check that ra0, when lifted to a ^-multiplier m0* for G*, is not exact. But suppose m0* is exact. Then m0*(x*,y*) = m0*(y*,x*) whenever x* and y* commute, and hence m0(x,y) = m0(y,x) if x and y commute and are close to the identity. We contradict this by choosing xs = (l,0,sv,0) and ys = (l,0,0,su), with s near 0; xsys = ysxs but mo(xs>ys)-mo(ys>Xs)= -2s2(u,v}. Theorem 7.42. Let G be the inhomogeneous Galilean group. Then for any real number r (137)
raT
: r, r' -> exp i^ [
is a multiplier for G, where r = (W,r),v,u) and
r' = (W',rj',v',uf).
Every multiplier of G is similar to one of this form. The mT, for distinct r, are not similar. In particular, m% is nonexact for r / 0 .
NOTES ON CHAPTER VII For a closely related treatment of multipliers see K. R. Parthasarathy, Multipliers on locally compact groups, Lecture Notes in Mathematics, No. 93, Springer-Verlag, Berlin, 1969. For a systematic treatment of the cohomology of locally compact second countable groups see the papers of C.C.Moore: Trans. Amer. Math. Soc, 113 (1964), pp. 40-63, 113 (1964), pp. 64-86; 221 (1976), pp. 1-33; 221 (1976), pp. 35-58; see also his review article in Group representations in Mathematics and Physics, Lecture Notes in Physics, No. 6, Springer-Verlag, Berlin, 1970.
CHAPTER VIII KINEMATICS AND DYNAMICS 1. T H E ABSTRACT S C H R O D I N G E R E Q U A T I O N I n this chapter we shall use the main results of Chapters V through V I I to study the dynamical and kinematical properties of quantum mechanical systems. This will lead to the Schrodinger equation, the well known formulas for the position, momentum, and spin observables, and to t h e standard expressions for the Hamiltonians which enter the q u a n t u m theory of the atom. The first problem we study is t h a t of determining the dynamical group of a system, say (S. According to our general principles, we must exhibit the dynamical group as a one-parameter group of convex automorphisms of the convex set of all states of (3. I n this section we obtain the form of the most general dynamical group of a system whose logic is standard. Our analysis leads to the so-called abstract Schrodinger equation of motion. This equation, in its usual form, gives a set of directions for computing the possible values of the energy observable a n d describes the development of the state of the system in time. I n keeping with the spirit of our development, we can say t h a t the Schrodinger equation gives the infinitesimal form of the temporal development of an arbitrary quantum mechanical system. Let us then consider a system <&; the quantum mechanical character of (B is p u t in evidence by the assumption t h a t the logic of <S is standard. We shall actually assume t h a t the logic of (& is t h a t associated with the complex number field. We write Jf for a separable infinite dimensional Hilbert space over C whose logic ££ is isomorphic to t h a t of 3 . We identify j£? with the logic of <&. We write £f for the convex set of all states ofJK According to our earlier discussion in Chapter I I I , the dynamical group of <& is then to be described by a Borel one-parameter group t - > D(t) of convex automorphisms of S^ the Borel structure being such as to make t h e maps t-+(D(t)p)(a) Borel, for all p e S? and a tS£. Now, by Gleason's theorem (cf. Chapter IV), A -+PA is a convex isomorphism of the set of von Neumann operators of 288
KINEMATICS
AND
DYNAMICS
289
a
trace 1 onto ^ , PA being t h e state a->tr(P A) (aeJ?). Consequently, each D(t) induces an automorphism of the convex set of von Neumann operators of trace 1, so t h a t we can select, for each t, a symmetry Ut ofjj? such t h a t (1)
D(t)pA =
pUtAUt-i
(cf. theorem 4.33). As D(t) = D(t/2)2, it follows t h a t each Ut is a unitary operator. Now, t h e m a p t-+Ut need not have any measurability properties, as t h e choice of Ut, for each ty was made quite arbitrarily. B u t , if u and v are unit vectors of Jtif, and if we write A for t h e projection on t h e space spanned b y u, a n d L for t h e space spanned by v, PUtAUt-l(L)
= K^^>|2-
Consequently, t h e m a p t^\(Utu,v}\2 is Borel for all u, v e J4f. If we therefore write IT for t h e canonical homomorphism of t h e unitary group of c2f onto t h e projective group of ^f, this implies (corollary 7.2) t h a t t h e m a p t->7r(Ut) is a Borel homomorphism of R 1 into the projective group. Consequently it is induced b y some projective representation of R 1 in J f (cf. Chapter V I I ) . Now, according to theorem 7.38, all multipliers of R 1 are exact. Hence there exists a (unitary) representation V (2)
V:t-+Vt
of R 1 in J f , such t h a t for each t, and each von N e u m a n n operator A of trace 1, D(t) is given by (3)
D(t)-PA-*PvtAvr1'
The representation V is determined u p to multiplication b y a character of t h e real line, i.e., if V'(t - > Vt') is another representation satisfying (3), then there is a real number c such t h a t (4)
Vt' = e"°Vt
for all t. To t h e one-parameter group V we can apply Stone's theorem (cf. Riesz-Nagy [1]). According to this theorem, there exists a unique selfadjoint operator H such t h a t (5)
Vt =
ex$(-itH)
for all t; t h e domain <2)H of H is given by (6)
QJH = \v : lim - (Vtv — v) exists L I t->o t J
and, for v e Q)H, (7)
Hv = ilim- (Vtv — v). t~*o t
290
GEOMETRY
OF QUANTUM
THEORY
We can thus say, in view of (5) and (3), t h a t H induces the dynamical group t -> D(t). Since the representation V is determined only u p to multiplication b y a character of the real line, D determines H u p to an additive constant, i.e., H and H + c-1 (c real) induce the same dynamical group. Collecting these facts, we have the following theorem: Theorem 8.1. Given any self-adjoint operator H on the complex separable Hilbert space J f of infinite dimension, there exists a unique dynamical group D(t - > D(t)) of Sf such that, for each t, D{t) transforms the state corresponding to the von Neumann operator A into the state corresponding to the von Neumann operator exp( — itH)A exip(itH). H and H + cl (c a real number) induce the same dynamical group. Conversely, any dynamical group D(t —> D(t)) of convex automorphisms of £f is induced in the above fashion by a self-adjoint operator H, and D determines H up to an additive constant. Finally, for any t, D(t) transforms the pure state determined by the unit vector u into the pure state determined by the unit vector exp( — itH)u. Theorem 11.1 is the central result of elementary quantum theory. I t tells us t h a t the dynamical groups of physical systems correspond one-one to one-parameter unitary groups and even to self-adjoint operators, provided we ignore additive constants in the self-adjoint operators. Now, according to our basic principles, the observables of the system are in one-one correspondence with the self-adjoint operators of 3ff. Consequently, we reach the remarkable conclusion t h a t to every dynamical group there can be associated, in a canonical fashion, a class of observables, a n y two of which differ by an additive constant, and t h a t any such class will generate a dynamical group. This conclusion brings us to a very interesting formal analogy with classical mechanics. I n its Hamiltonian form (Chapter I, Volume I), the dynamical group of a classical system is determined also b y a function H on the phase space, i.e., a classical observable, and, since H enters the differential equations in terms of its first-order derivatives, this observable in t u r n is determined by t h e dynamical group only up to an additive constant. Prompted by this analogy, we shall call any self-adjoint operator H, which induces the dynamical group of a system, a Hamiltonian of the system. Given a physical system, the observable which corresponds to a Hamiltonian of the system is undoubtedly a very crucial one, since it describes the dynamics of the system completely. I n physics, this observable is usually identified with the energy of the system. We shall give, further on, some of the reasons t h a t make this identification a very reasonable one. Assuming this identification, we can now recover t h e classical prescription of Schrodinger for the determination of the energy levels of the system. If u is a unit vector of 3tf which describes a pure state of the system under
KINEMATICS
AND
DYNAMICS
291
consideration, and if P(E -> PE) is the projection valued measure giving the spectral decomposition of the operator H, then the probability distribution of energy, when the system is in the state corresponding to u, is the measure E^\\PEu\\2;
(8)
if H has a purely discrete simple spectrum, say {Al5 A 2 ? - } , and if ar u1,u2,-m' e orthonormal eigenvectors of H corresponding to the eigenvalues Al5 A 2 , - - , respectively, then Als A 2 ,--- are the possible values of the energy observable, say En, and, in the given state, (9)
Fr(En = Am) = \
Let us now consider t h e dynamical group induced by the unitary group £ - > e x p ( — itH). I t is well known, and easy to prove in the context of Hilbert spaces (cf. Gel'fand [1]), t h a t there are vectors u in Jf such t h a t t —> exp( — itH)u is an analytic function of t with values in Jtif; and t h a t , moreover, such vectors form a linear manifold which is dense in Jf, and on which H is essentially self-adjoint. If we take such a vector u and write ut = exj)( — itH)u, then we obtain the differential equation (10)
-~
= -iHut,
u0 = u.
The group t -> exp( — itH) can be clearly recovered from (10). The equation (10) is thus the differential equation which describes the motion of the pure states of our system. We shall call it the abstract Schrodinger equation (cf. D i r a c [ l ] , p p . 108-111). From the representation t-+Vt we can now determine how the probability distribution of any observable changes with time. Let £ be any observable and A the self-adjoint operator associated with it. Let E -> PE be the spectral measure of A. If the unit vector uQ represents the state at time t = 0, then the distribution of $ a t time t is the measure qt where (11)
qt(E) = <exip(itH)PE
exV(-itH)u0,u0).
I t is of course quite possible t h a t the distribution of £ a t time t = 0 remains unchanged for all time, no matter what the initial state is. From (11) we infer t h a t the necessary and sufficient condition for this is t h a t Vt~lPEVt
= PE
for all t and all Borel sets E, i.e., (12)
exp( — itH) exp( — isA) = exp( — isA) exp( — itH)
292
GEOMETRY
OF QUANTUM
THEORY
for all s a n d t. The equation (12) yields, on differentiation, (13)
[H,A] = 0,
where (13) is interpreted in t h e strong sense, i.e., t h e spectral projections of H a n d A commute. I n other words, t h e observable | must be simultaneously observable with energy. I n analogy with classical mechanics we shall call f an integral of the motion. The reader m a y notice t h a t (13) is also the condition for f t o be an integral of t h e classical motion determined by a classical Hamiltonian H, provided we interpret t h e bracket in (13) as t h e Poisson bracket. Later on, we shall come across further instances of this analogy. In particular, energy is always conserved during motion. We shall conclude this section with a few remarks on t h e "Heisenberg picture." Consider a system whose dynamical development is described b y t h e one-parameter group t - > Vt. The pure states of t h e system t h e n move in time along t h e parametrized curves t - > Vtu. We shall call these curves t h e motions of t h e system or, alternately, t h e dynamical states of the system. Clearly a n y such motion t —> ut is completely determined b y the vector u0. Hence t h e rays of t h e Hilbert space of t h e system can also be regarded as representing t h e dynamical states of t h e system. If u a n d v are two unit vectors which represent t h e dynamical states £ a n d 77, respectively, t h e number |<^,v>| 2 can evidently be interpreted as t h e probability of finding t h e system in t h e dynamical state 77 when a measurement is made on it to find out whether it is in t h e dynamical state £ or not. Thus \(u,v)>\2, which depends only on t h e rays determined b y u a n d v, has a n invariant physical significance. I t is possible t o reformulate t h e general prescriptions of q u a n t u m mechanics in terms of the concept of the dynamical state. We shall indicate how this is done b y computing t h e probability distributions of t h e observables as time changes. Suppose x is a n observable a n d Ax — A its operator. Suppose t h a t t h e system is in t h e dynamical state u0. Then, t h e probability distribution of x a t time t is t h e measure qt
:E-*(PEVtu0,Vtu0},
where E - > PE is t h e spectral measure of A. If we write At = then P* : E - > Vt'1PEVt is t h e spectral measure of At a n d
Vt~xAVt,
qt(E) = < P ^ 0 ^ o > , which describes qt entirely in terms of u0. We see, therefore, t h a t t h e probability distribution of x a t time t can be computed if we replace A by At, b u t evaluate t h e relevant inner products with u0 instead of ut. I n other words, we may regard the change in time as inducing a change in the operators which represent x; the dynamical state remains unchanged.
KINEMATICS
AND
DYNAMICS
293
I t must be realized, however, t h a t this is only an alternative language to describe the same state of affairs. In physical literature, this way of looking at things is called the Heisenberg picture and the dynamical states are called the Heisenberg states. Our original way of looking at things is referred to as the Schrddinger picture. If A is the operator which represents the observable x, then at time t, it will be represented by the operator At=Vt~1AVi in the Heisenberg picture. The (formal) time derivative of At at £ = 0 then becomes i[H,A], where we write Vt = exp( — itH): (14)
A =
i[H,A].
We thus obtain the celebrated principle of Heisenberg according to which the time derivative of an observable is i times the commutator of its operator (at time 0) with the energy. Appropriately interpreted, (14) retains its meaning in classical mechanics also. 2. COVARIANCE AND COMMUTATION R U L E S We shall now consider a system such as a particle. A characteristic property of such systems is t h a t one can associate with each of them, a certain 0°° manifold M, which is usually called its configuration space. If we are interested in a classical interpretation of such a system, M would actually be its configuration space. Suppose t h a t © is such a system and t h a t we are interested in a quantum mechanical description of ©. Suppose we still like to treat the configuration of the system as a physical observable. The mathematical way to p u t this in evidence is to assign, for each Borel set E^M, a closed linear manifold S(E) of the Hilbert space Jf of (&; S(E) will then represent the experimental statement t h a t the configuration of © belongs to E. I t is, moreover, natural to assume t h a t the assignment E -> S(E) is a cr-homomorphism of the Borel structure of M onto a Boolean sub -a- algebra of the logic ££ of Jf^ i.e., t h a t the mapping E —> PS(E) = PE, which assigns to E the projection on S(E), is a projection valued measure based on M. Let us now suppose t h a t we have in addition a Lie group G acting differentiably on M and t h a t we want a description of our system which is covariant with respect to G (for example, we may treat a particle in a spherically symmetric potential field in space; M will then be R3 and G will be the rotation group). We then associate, with each element g of G, an automorphism ag of the logic 3? of our system such t h a t a(g -> ag) is a representation of G in Aut(J£f); moreover, in view of the simple meaning of the elements S(E) of ££, we shall require t h a t a and S be related as follows: (15)
ag[S(E)] = S(g[E]).
294
GEOMETRY
OF QUANTUM
THEORY
We take the point of view t h a t (15) summarizes the physical assumptions we have made. If G is a connected Lie group, we can select a projective representation U(g —> Ug) of G in Jf7 which induces a. Then (15) can be written as (16)
U„PEUt-1
=
Pgm-
If U is a representation, then (16) expresses the fact t h a t (U,P) is a system of imprimitivity. If U were not a representation but its multiplier is exact, then we m a y change U to a representation so t h a t it still induces a. I n such a case, we m a y therefore say t h a t the existence of a configuration space plus the requirement of covariance leads to a system of imprimitivity, the representation and projection valued measure of which have direct physical significance. The exactness of the multipliers of the projective representations which induce a cannot of course be always taken for granted. B u t in m a n y cases this is the case (cf. theorem 7.40). The equations (16) can be written formally in their infinitesimal forms. Let U be a fixed representation satisfying (16). Let g be the Lie algebra of G. For any l e g and each t, x -> exip(tX) • x is a diffeomorphism of M; and as t varies, we obtain a one-parameter group. Let rx be the vector field of this one-parameter group. For any f eCcc(M) and x e M, we have (by definition), (17)
(rxf)(x)=jt(f(exV(tX)-x))t
= 0.
For any real (7°° f u n c t i o n / on M, let Af be the operator (18)
A, = \
fdP.
JM
Af is possibly unbounded b u t self-adjoint. Then, (16) leads to (19)
U9AfUg->
=
A(fSy
Now, t —> U(exptX) is a representation of E1, and hence there exists a selfadjoint operator Bx such t h a t exp( — itBx) = U{exptX) for all t. Bx is determined uniquely and we have, formally from (19), [A„Ar]
= 0,
[BX,BY]
=
iB[x>Y],
aBx + bBY = B[aX + bYWe shall call the equations (20) the Heisenberg commutation (10), Chapter I).
rules (cf.
KINEMATICS
AND DYNAMICS
295
The observables associated with the operators Af and Bx which enter the relations (20) m a y be given interesting physical interpretations. We consider first the operators Af. We recall t h a t for any set E, 8(E) represents the experimental statement t h a t the configuration of 3 belongs to E. Now, from the formula (18) for Af, we see t h a t the operator Af has the spectral measure J P - > Pf-i(F) (F a Borel set of the line). Hence, if we denote by £ the observable corresponding to Af, the statement t h a t the value of £ lies in F must be represented by the subspace 8(f~1(F)). I t is thus identical with the statement t h a t the configuration of 3> belongs to the set f~1(F). I n other words, if we observe, instead of the configuration x of 3 , the value of/ a t x, we obtain an observable whose associated selfadjoint operator is Af. We shall therefore call these observables configuration observables. Note t h a t the function / represents classically an observable with the same significance. Thus f-> Af is an algebra homomorphism which assigns to each classical configuration observable the "corresponding" quantum observable. For example, if M = Rn a n d / is the kih coordinate function (tl9- • •, tn) -> tk, then Af is the operator which represents the kih configuration coordinate. On the other hand, the vector field rx gives rise to the classical momentum observable corresponding to the symmetries t -> exp(L£) acting on M, and the analogies we have so far developed with classical mechanics lead us to define Bx as the quantum mechanical momentum observable corresponding to the symmetries generated by X. In other words, the covariance expressed by (20) can be formally regarded as describing certain commutation rules between the configuration observables and the momentum observables of the system. We have thus arrived at the real mathematical meaning of the celebrated commutation rules of physics; at the same time our extensive knowledge of systems of imprimitivity enables us to understand why they are so far-reaching.
3. T H E S C H R O D I N G E R
REPRESENTATION
Given M and G, the main problem of course is to determine the pairs (a,S) which satisfy the relations (15). This is, at least in the special case when a is induced by a representation of G, nothing but the problem of determining the systems of imprimitivity for G which are based on M. The analysis of Chapter V I shows t h a t such a determination is completely known when G acts transitively on M. Fortunately, in m a n y physical applications, G is transitive on M, so t h a t these results can be applied. On the other hand, whether or not G acts transitively on M, there always exist the Koopman systems of imprimitivity associated with the invariant measure classes on M. In this section we shall explicitly
296
GEOMETRY
OF QUANTUM
THEORY
describe these latter systems in the form in which they are most familiar in physics. We start with the O00 manifold M and a Borel measure v on it. We write Jf = J?2(M,i>). For any real C00 function / on M we define Af to be the operator of multiplication by / : (21)
A,:
For any C°° vector field Z on M we define Bz as the operator (22)
Bz:
-iZcp.
Both (21) and (22) are meaningful only when
(23)
Proof. By definition, (Z
f (Z
for all
£= jM
(Z
= f (Z{v-))(x)dv'(*), JM
KINEMATICS
AND DYNAMICS
297
so that (24) implies that u is a constant, giving the invariance of v . This said, we come to the proof proper. For 99, if; e Ccco(M), (p(x)i/j(t~1-x)dv(x),
JM
since v is invariant. Therefore, we have, on differentiation, f (Zcp)ijjdv = - f (p(Zifj)dv, JM
JM
and it follows that the operator Bz : 9 - ^ — iZy is symmetric on Ccco(M). We shall now prove that the closure of Bz in 34? is self-adjoint. It is known from standard spectral theory that a necessary and sufficient condition for this is that for any nonreal complex number £, the operator Bz + £ • 1 map Cc °° (M) onto a dense linear manifold of 34?,'. Suppose that f0e 3^ and -i
f (Z
JM
for all 99 in Cc°°(Jf). We must prove t h a t / 0 = 0. Let p = £/i. Then p is not purely imaginary and (25)
f (Z^/odv = p f
JM
for all 9 6 0c°°(M). Consider now the manifold M' = R1 x M and the one-parameter group t -> gt' of diffeomorphisms of M', where gt'(e,x) = (t + 89t-x). Write Z' for the corresponding vector field; for u e Ccco(R1) and
+ r
u{s)e"sds [ (Z
j - 00
JM
= 0,
as is easily seen by integrating by parts the integral with respect to ds in the first term and using (25). A standard approximation argument (cf. Schwartz [1], pp. 108-109) now leads to (((Z'w)(s,x)ep%(x)dsdv(x) M'
= 0
298
GEOMETRY
OF QUANTUM
THEORY
for all w eCcco(M'). In view of our observation at the beginning of the proof (cf. (24)) we conclude that the complex Borel measure A, where
A(^) = J]Wo (x)dsdv(x), is invariant with respect to the group t-> gt'. But dsdv is also an invariant measure. Hence for each t, we must have epis + t)f0(t-x) = epsf0(x) for almost all s, x, i.e.,
/ 0 ( M = e-«f0(x) for v-almost all x. Therefore, for each real t, jM \fo(x)\2dv(x) = ( J
|/0(s)|2efo(s))|
and since p is not purely imaginary, we conclude t h a t / 0 = 0. As mentioned before, this proves that the closure of the operator
(VJ)(x)=f(t->-x)
(feJf),
are unitary and Vt = exp( — itA) for a self-adjoint operator A. Further, A(p = Bz
*«> =
XB'9,
(Ug
(U,P) is a Koopman system of imprimitivity associated with v and G. If we take 3tf as the Hilbert space of a system with M as its configuration space and G as its group of symmetries, then we see that the configuration observables are represented by the multiplication operators Af (feCco(M)). For any element l e g , the Lie algebra of G, the momentum observable Bx corresponding to the symmetries x -> exp(L£) x of M is the differential operator —irx, where rx is the vector field on M defined by (rxcp)(x) = L-t
KINEMATICS
AND
DYNAMICS
299
the differential operator Bx = — irx is the self-adjoint operator which is the closure of the operator 99 —> —irx(p on CC™(M) and U(exv tX) = exp( — itBx) for all t. The Bx and Af satisfy the commutation rules: [A„Ar]
(27)
t^^/] [BX,BY]
= 0, iA =
w
=
iBlXtY]i
aBx + bBY = B(aX
+ bY),
w h e r e / , / ' eO°°(Jf), a, 6 are constants, and X, 7 e g ; the equations (27) are valid on the dense linear manifold CCQD(M). This particular representation of the equations (20) is known as the Schrodinger representation. From our point of view the relations (16) are the fundamental ones; the commutation rules (20) are to be regarded merely as their formal infinitesimal versions. The reason for the frequent appearance of the commutation rules in physical literature is simply due to the tradition in mechanics and physics according to which the laws governing physical systems were always expressed in the infinitesimal form whenever possible. Rigorous formulations of (20) can be given; but they get involved with the domain considerations inevitable in any considerations of unbounded operators, and the equations lose the global geometric flavor possessed by the " i n t e g r a t e d " relations (16). The mathematical importance of the integrated versions of (20) seems to have been first recognized by Weyl ([1], pp. 272-276). When G acts transitively and freely on M, we m a y use theorem 9.17 to conclude t h a t every irreducible system of imprimitivity for G based on M is equivalent to the Koopman system. We m a y state this loosely by saying t h a t every irreducible representation of the commutation rules (20) is equivalent to the Schrodinger representation. When M was Rn and G was the group of its translations, this was first proved by Stone [3] and von Neumann [3]. Our method of proof, based on theorem 6.17, is t h a t of Mackey [9]. This result means t h a t the apparently special choices made in the Schrodinger representation, for representing the states and the position and m o m e n t u m observables, are in fact the most general ones possible subject to the commutation rules (20), provided we assume irreducibility. We m a y thus regard this theorem as proving the isomorphism of Schrodinger's wave mechanics (based on the Schrodinger representation), and Heisenberg's matrix mechanics (based on the commutation rules (20) only and not on any specific choice of the Af and Bx), and thereby unifying the two great developments. The irreducibility of the (U,P) can be given an interesting interpretation. The system (U,P) is irreducible if and only if the elements in the commuting ring of (U,P) are the multiples of the identity or if and only if
300
GEOMETRY
OF QUANTUM
THEORY
the smallest von N e u m a n n algebra containing the spectral projections of the Bx and Af is the ring of all operators in *#?. Let us define a self-adjoint operator A t o be a function of the Bx and Af if A belongs t o the von Neumann algebra generated b y the Bx and the Af. Then every selfad joint operator is a function of the Bx and the Af. On the other hand, if we consider the classical system with configuration space M, it is true t h a t all the observables are functions of the configuration observables and the linear momenta. We thus see t h a t the assumption of reproducibility corresponds (roughly) to the assumption t h a t there are essentially no new observables other t h a n the configuration ones and the linear momenta.
4. A F F I N E C O N F I G U R A T I O N SPACES An affine space is a pair (M,V), where (i) V is a real finite dimensional vector space, and (ii) V acts freely and transitively on M. B y the usual abuse of language we shall refer to M itself as an affine space. V is said to be the vector space associated with M. For v e V we write tv for the transformation of M induced by v and refer to it as the translation by v. We write tv-x = x + v. Given x and y in M, we shall write y — x for the unique element v of V such t h a t tv-x = y. Since t_v and tv are inverse to each other, y — x=—(x — y)f while the relation tu + u, = tutu> leads to the equation (x — y) + (y — z) = x — z. A one-one m a p L of M onto itself is called an affine automorphism of M if there exists a linear automorphism L° of the vector space V such t h a t for all x and y in M, (28)
Lx-Ly = L°(x-y).
Since for fixed x, y -> (x — y) maps M onto V, it is clear t h a t L° is uniquely determined by L. I f L° is the identity, we shall call L a translation. I f v e V, it is trivial to check t h a t tv is a translation. Conversely, let L be a translation and let x and y be two points such t h a t y = Lx. I f we write y = x + v, then it follows from the equation Lz1 — Lz2 = z± — z2 t h a t L is t h e translation x - > x + v. If we choose a point x0e M and m a p M bijectively onto V by the m a p v - > x0 + v, then the affine automorphisms of M are easily seen t o be precisely those which go over t o the inhomogeneous transformations (29)
a->L°a
+b
of V. Thus, in the notion of the affine automorphisms of M, we have a n invariant description of the inhomogeneous linear transformations which are associated with vector spaces.
KINEMATICS
AND
DYNAMICS
301
If we choose a basis vl9 • • •, vn for V, and a point x0 e ilf, then the map
(30)
P i , " - , W - > * o + ( 2 *>*>) n
is one-one from R onto if, and its inverse maps M onto Rn bijectively. We shall call the inverse of (30) an affine coordinate system on M with origin in x0. Suppose x0, xx e M, {vl9- • •, vn}9 {wl9 • • •, wn} are two bases for V, and y, y the affine coordinate systems of M associated with x0, {Vi,"',vn} and xl9 {wl9--9wn}. Then there exists (au'-,an)eRn and an invertible matrix (d^)*i==1 such that, for any xeM, the images (*i>"m,tn) = y(#) and (£/,• • • 9tn') = y'(x) are related by
(31)
t{ = 2 dtfr + at. ;=i
We thus obtain the usual formulas which connect the two coordinate systems. From the formula (31) it follows that there is a unique C00 structure on M such that each affine coordinate system is a diffeomorphism of M onto Rn. With this C00 structure, the map v, x -> x + v is C°° from V x M into M. It is obvious that M admits an invariant Borel measure, say dm. dm goes over to a multiple of the Lebesgue measure in Rn under any affine coordinate system. From our analysis in Section 2 we conclude that the only irreducibly covariant systems with M as their configuration space are described by the Schrodinger representation. The Hilbert space of such systems can be taken to be J?2(dm). The configuration observables are then represented by the multiplication operators. If v e V, the momentum observable corresponding to the symmetries x->x + tv is represented by the operator Bv: (32) Bv = -id(v), where d(v) is the usual linear differential operator associated with v. We shall call this the linear momentum observable in the direction of the vector v. If we choose now a point x0 e M, a basis {vlt • • •, vn} of V, and map M onto Rn with the help of the corresponding affine coordinate system y, we can obtain concrete forms for these results. For simplicity of description we identify M with Rn and 2tf with j£?2(Rw). Then the &th coordinate function, say qk, is represented by the operator of multiplication by tk: (33) qk ~
pj -
302
GEOMETRY
OF QUANTUM
THEORY
The second of the commutation rules (20) then becomes (35)
[,] =
(-l)1/a8wl.
which is the form in which the rules were stated first by Heisenberg. If cp is a unit vector in ^f, i.e., (36)
[ \cp{x)\2dm(x) = 1, JM
then the probability distribution q^ of the configuration of the system in the (pure) state represented by cp is given by (37)
q0(E) = j
\cp(x)\Hm(x).
Since we are identifying M with R n using a coordinate system, the formula (37) becomes
(38)
q9(E) = j
\
The formulas (34) and (35) were first postulated by Schrodinger when M was three dimensional and the system to be analyzed was t h a t of a single particle. Formula (34) gave the forms of the position operators, (35) the forms of the components of the linear momenta. Formula (38) gave t h e probability t h a t in the state represented by cp, called the wave function of the particle, the particle is to be found in the region E and was first recognized by Max Born. From our point of view, we see t h a t the whole development falls out of the assumption of affine covariance and the irreducibility of the system (U,P). The special forms of the configuration and momentum observables can also be used to illustrate the phenomenon of complementarity so peculiar to q u a n t u m theory. Suppose Jf?, U, P are as above. Let us choose an affine coordinate system for M and identify M with Rn and 34? with j£?2(Rn) via this coordinate system. The expressions for the configuration observables show t h a t they commute among themselves. On the other hand, the translation operators Uv also commute among themselves, so t h a t the momentum operators Bv commute among themselves. Let us now take a function 9?ee£?2(Rn) of unit norm and compute the probability distributions of the vector observables q = (qx, • • •, qn) and p = (pl9 • • •, pn). For q, this is straightforward; it is
(39)
^ - > J M^-'-^Jl^i-'-^n-
To compute the distribution of p we must use Fourier transforms. Let (40)
& : I/J - > ^
KINEMATICS
AND DYNAMICS 2
so3
n
be the Fourier-Plancherel transform of j£? (R ) onto itself which is the unique unitary automorphism of j£?2(Rn) with the property t h a t for all ^eC.(R»),
(41) fi(tv ••-,*») = ^ j ^ i JRW e x P ~ * A + ' * • + *»*»] x 0(0?!, • • •, xn)dx1 • • • dxn ,,,
fl
for all fe, ,yeR (Bochner-Chandrasekharan [1]). I t is t h e n easy to show t h a t for v = (v1}- • •, vn), ^U^"1 is the operator of multiplication by exp [ — *(vi*i + • • • +vntn)]. From this it follows easily t h a t the spectral measure of Bv is given by F - > J r _ 1 P F u J ^ where (42)
Fv = {(tl9 • • • , g : « A + • • • + V n * n e F}
(F a Borel set on the line). From (42) we see t h a t the probability distribution of p in the state given by y is (43)
E ^
f
\
JE
Formulas (39) and (43) show the remarkable statistical connection between the configuration observables and momentum observables. If 99 is such t h a t cp vanishes outside of a small region S, the configuration of the system is localized in S statistically; formula (43) then tells us t h a t the distribution of the momenta is very diffuse. Conversely, states 99 for which 9? has a small region for its support localize the momenta; but the configuration observables will have diffuse probability distributions. No single state exists in which both the position and the momenta can be localized sharply and the same system exhibits both types of properties in appropriate states.
5. E U C L I D E A N SYSTEMS: S P I N We shall now discuss the situation when G is transitive on M b u t its action is no longer free. The different systems with M as their configuration space and G as their symmetry group will now be labeled by representations of the stability groups. I n the special case when i f is a threedimensional Euclidean affine space and G is the covering group of its group of motions, this leads to the kinematical classification of such systems in terms of their spin. A Euclidean affine space is a triple ( J f , F , < . , . » , where M is an affine space with associated vector space V, and <.,.> is a positive definite inner product on V x V. By the usual abuse of language we shall refer to M itself as a Euclidean space. An affine automorphism L of M is called
304
GEOMETRY
OF QUANTUM
THEORY
a motion of M if the associated linear transformation L° of V is orthogonal with respect to <(.,.) and has determinant + 1 . The set of all motions of M is a group, denoted by G. For each v e V, the translation tv : x -> x + v belongs to G and v -> tv is an isomorphism of V onto a normal subgroup T0 of 6r. For any x e M we write G^ for the stability subgroup of G at #. Given x0 e M and a basis {vlf • • •, vn} (w = dim V) of F, the affine coordinate system relative to these is called a Euclidean coordinate system, if {vir ' *> vn} is a n orthonormal basis for V. It is clear that if we map M onto Rw using a Euclidean coordinate system, G goes over to the inhomogeneous SO(n)-group in Rn and GXo goes over to the rotation group SO(n), where x0 is the origin of the coordinate system. From this observation we can obtain at once the structure of the group G. For any x e M, (44)
G = T0GX,
T0 n Gx = {identity}.
Moreover, there is a canonical way to convert G into a Lie group. T0 and Gx are closed, and Gx is compact. If n>2, G is connected. If n>3, Gx is semisimple and its universal covering group is also compact. We shall choose a universal covering group G* over G and denote by 8 the covering homomorphism of G* onto G. 8 is an isomorphism of T 0 *, the component of the identity of 8~1(T0), onto TQ\ on the other hand, for any xeM, 8~1(GX) = GX* is connected, simply connected, and covers Gx (through 8). G* is a semidirect product of T0* and Gx*. Suppose we consider a quantum mechanical system © for which a Euclidean affine space M is the configuration space. Assume the dimension of M to be > 3. For a description of (£> which is covariant with respect to the group G of motions of M, we must introduce a pair (U,P), where P(E -> PE) is a projection valued measure based on M and U a projective representation of G such that for all g and E, (45)
U9PEUg-i = Pgm.
Both U and P act in ^ the Hilbert space of (2. Now the group G satisfies the conditions of theorem 7.40 and hence we know that the universal covering group G* of G has only exact multipliers. We may thus assume that the projective representation U arises from a unique unitary representation V of G* in Jf. V and P will then satisfy the relation (46)
Vg.PEVg.~i = P6(g.hE
(g*eG*).
The action g*, x -> 8(g*)-x converts M into a transitive 6r*-space and (46) tells us that (V,P) is a system of imprimitivity for G*. (V,P) is irreducible if and only if the pair (U,P) is so. Conversely, suppose (V,P) is an irreducible system of imprimitivity for G* based on M, acting in #F. Then the equation (46) shows that for any element g* in the kernel of 8, Vg* commutes with all PE; but Vg* also commutes with all the Vx* (x* e G*),
KINEMATICS
AND
DYNAMICS
305
since g* lies in the center of G*. Hence the irreducibility of (V,P) will imply t h a t V maps the elements of the kernel of 8 into scalars. From this we at once infer the existence of a projective representation U of G such t h a t a;* —> Ud(x^ and x* -> Vx* will define the same homomorphism of G* in the projective group of Jf. Our problem is t h u s reduced to the determination of the irreducible systems of imprimitivity for G*, based on M. Unlike the situation involving only the translations, there exist now irreducible systems (V,P) which are not unitarily equivalent. From t h e general theory developed in Chapter V I we know t h a t the unitary equivalence classes of our pairs (V,P) are in canonical one-one correspondence with the equivalence classes of irreducible representations of the stability subgroup GXo* at some point x0 e M. However, the relation of unitary equivalence between the systems (V,P) is in general finer t h a n the relation of physical equivalence; we shall call two systems of imprimitivity for G* based on M, say (V,P) and (V',P'), acting in 2tf and 3tf", respectively, physically equivalent if there exists a unitary or antiunitary isomorphism W of J f onto Jtf" such t h a t (i) P= W o P' o W'1, and (ii) the representations V and W o V o W'1 of G* in 3%" give rise to the same homomorphism into the projective group of 3tf'. Notice t h a t as G* has no one-dimensional representations (cf. theorem 7.40), (ii) is equivalent to requiring t h a t V = W o V o W'1. If IF is unitary we are back to unitary equivalence. B u t if W is antiunitary, the relation of physical equivalence will in general not imply unitary equivalence. We have, in fact, the following elementary result which clarifies the connection between unitary and physical equivalence. Lemma 8,3, Let X be a transitive Borel H-space for a Icsc group H and let x0 e X be an arbitrary point. Let H0 be the stability subgroup at x0. Suppose that nij (j=l, 2) are two unitary representations of H0. Let (V3\P3) be the associated systems of imprimitivity acting in J^j, say, j = l, 2. Then, in order that there exist an antiunitary operator W of 34?1 onto Jf2 such that V2=W o V1 o W~x and P2= W o P 1 o W'1, it is necessary and sufficient that m1 and m2 be contragredient to each other. Proof. We choose a quasi-invariant measure a on X. Let J f ; be the Hilbert space in which m ; acts. Then, according to the general theory of Chapter VI, there is a strict (#,X)-cocycle y\ whose values are in the unitary group of Jf 7 with ( F ; , P ; ) ^ ( F ; , P ; ) > where (Vj,Pj) is associated with cpj and a and acts in ££2(X,3f3,a). We choose a fixed (but otherwise arbitrary) antiunitary automorphism 8 of J T 1 and write S~ for the mapping of j£? 2 (X,Jf 1 ,a) onto itself defined by (47)
(S~f)(x)
= Sf(z)
(xeX).
306
GEOMETRY
OF QUANTUM
THEORY
S~ is evidently a n antiunitary automorphism of o^f2(JC,jT1,a) a n d S~ P1S~~1 = P1, as m a y be easily calculated from t h e formula PE^—XEJWe write (48)
U1
=S~V1S~-\
(U1,?1) is also a system of imprimitivity for H based on X. Moreover, using t h e formula (39) of Chapter V I it can be easily calculated t h a t U1 is t h e system associated with t h e strict cocycle y1, where (49)
StpHgrfS-1.
Thus there exists an antiunitary isomorphism W of <£2(X,Xx,a) onto £>2(X,Jf2,a) such t h a t Wt1W~1= V2 a n d WP1W~1 = P2 if a n d only if there exists a unitary isomorphism W which sets u p a n equivalence between (U1,?1) a n d (V2,P2); in fact, W and ffl are related b y t h e equation W = WS~ '*. Hence a necessary and sufficient condition t h a t there should exist an antiunitary operator W such t h a t Wo V1 o W~~1=V2 and W o P1 o W~x = P2 is t h a t t h e cocycles cp1 a n d
KINEMATICS
AND
DYNAMICS
Since T0* is normal in G* and G* = T0*GXQ*, the unique decomposition (50)
g* = t(g*)r(g*)
(51)
we have, for a n y g* e G*,
(t(g*) e T 0 *, r(g*) e GXo*).
g1*g2* = t(gi*)[r(g1*)t(g2*)r(gl*)-1]r(g1*)r(g2*)
The equation once t h a t
r(gi*g2*)
307
= r(gi*)r(g2*)
shows
at
(g±*, g2* e GXQ*).
Moreover, for g* in GXQ*,
(52)
r[g*) = g*.
(51) and (52) show t h a t , if we define 99s by 9*(9*,x) = s(r(9*)), then 99s is a strict cocycle of the type we want. We are finally in a position to describe the irreducible pairs (V,P) explicitly. Let s(h - > s(h)) be an irreducible representation of GXo* in a (finite dimensional) Hilbert space J f and let J^s =
(53) For E^M
£>2(M,Jf,dm).
and g* e G* we define E
(54)
*
(V\4)(x)
=
XE
'*'
=
8(r(g*))Wig*)-i.x).
Then (Vs,Ps) is the system of imprimitivity corresponding t o the irreducible representation s of GXQ*. The most important special case of the preceding discussion arises when M is three dimensional. We shall now proceed t o consider this case in somewhat greater detail. We select a Euclidean coordinate system y with origin x0. Using 7, we shall identify M and V with R 3 ; t h e bilinear form on F x F becomes the usual one. G will then be identified with the group of all rigid motions of R 3 . GXo goes over t o t h e rotation group 80(3). We write G0 for GXQ and G0* for Gx*. We shall first construct an explicit covering of G0 b y G0* following the well known construction of Weyl ([1], p p . 143-146). We introduce t h e Hermitian matrices (Pauli spin matrices) (55)
CTl =
(i
0)' CT2 = (!
0)' CT3 = (o - 1 ) '
and the correspondence x -> A(x), where x e R 3 and (56)
A(x) =
x1G1-hx2a2-{-x3a3.
308
GEOMETRY
OF QUANTUM
THEORY
Let (57)
£ 0 * = SU(2,C).
Then for g* e G0*, A(x) -> (g*)A(x)(g*)-1 is a linear transformation of the vector space of all 2 x 2 Hermitian matrices of trace 0. Hence, via the correspondence (56), we obtain a linear transformation <%*) of R 3 : (58)
A(&(g*)x) = 2
2
g*A(x)g*-\
2
Since det A(x) = — (x1 -\-x2 i-x3 ) is preserved under unitary similarities, d(g*) is an orthogonal transformation of R3. It is well known that 8(g* -> S(g*)) is a covering homomorphism of G0* onto G0. Since the group G consists of elements (g,\) (v e R3, g e G0) with the composition (59)
(g,v)(g'S) - (gg',y+g-v')>
we may define G* to be G0* x R3 with the composition (60)
(g*,y)(g*',y') =
teV>+%*)•')•
The map (61)
8 : (*,v) -> (%*),v)
is a covering homomorphism from G* onto G. It is useful to describe the one-parameter subgroups of G0* which map onto the familiar rotations around the axes. An easy computation leads to the following. If t -> g^(t) are the three one-parameter subgroups of G0* defined by (eim (62)
0
"•W-(0
\
/ cos(/2 i sm tl2\
e-4
^ cos
•ft) = /
92 U
= {isJl2 ecJ/2)' *l
\-sin*/2
s i n tl2
\
coat ft)'
then t-> &(gr*(t)) = gr(t) is given by
(
cos t — sin t 0
sin t
0\
cos t
0
0
1/
(63) /cos £ 02W = ( 0
\sin £ The theory of the irreducible representations of 6r0* is well known (cf. Weyl [1], pp. 143-180). The equivalence classes of these representa-
KINEMATICS
AND
DYNAMICS
309
tions are labeled by a parameter j which takes the values 0, J, 1, • § , • • • . For a given j , the associated representation, say D\ is of dimension 2j + 1 , and there exists an orthonormal basis in the representation space of D3, say {vjv : v=j, J — 1, • • •, — j}, such t h a t (64)
&(g3*(t))vjv
= e""*v
The basic representation g* - > g* corresponds t o ^ = J. The representations D3 for integral^' m a p the kernel of 8 ( = { ± 1}) into the identity, and hence induce representations of G0. The other D3 induce only projective representations of G0. An explicit realization of the D3 can be obtained as follows. We regard GQ* as operating in the obvious manner on C2 = W and consider the tensor product W®N= W ® • • -(g) W (N factors, N = 2j). Then W®N is a Hilbert space of dimension 2^. If we define, for g* e G0*, ir(g*) by (65)
Tr(g*) =
g*
then 7r(g* - > 7r(g*)) is a representation of G* in W®N. -n leaves the space of symmetric tensors invariant. Let Wj be this subspace. Then W3 is irreducible under TT, and the representation defined by Wj is equivalent to Dj. If e± = I J, the reader can trivially check t h a t (66)
ir(03*(O)(ei ® • • • ® ei) = c«*(ei ® • • • ® ex).
Finally, as the representations D y have different dimensions for distinct values of j , and as mutually contragredient representations have the same dimension, we see t h a t each Dj is contragredient to itself. I n view of our earlier discussion, the physical equivalence classes of the pairs (a,S) satisfying (15) depend on the parameter^' ( j = 0, \, 1,- • •), and for any j , (a,S) is the pair obtained from the corresponding system of imprimitivity induced by the representation D1 of 6^0*. The Hilbert space of wave functions of a typical system which is irreducibly covariant under t h e Euclidean group of R 3 can t h u s be regarded as the space ffi of square integrable Borel functions on R 3 whose values are symmetric tensors of rank (N = 2j) over the twodimensional complex space W. Let W3 be the Hilbert space of symmetric tensors of rank 2j over W with norm | • | and inner product < . , . > induced naturally from W. Let D3 denote the representation of G0* in W3. Then, for the elements I/J of J^3, we have
(67)
f |^(x)|^ x < oo,
and for a n y element (g*,\) of G*, we have (68)
( F ( 9 . » ( x ) = D'(g*)4,(8(g*) - H - 8(g*)' M ,
310
GEOMETRY
OF QUANTUM
THEORY
while PE is, as usual, the operator ijj - > XE^- The operators of V induce the lattice automorphisms corresponding to the elements of G. The formulas (67) and (68) are the usual ones in physical literature which express the "transformation of wave functions when the frame is r o t a t e d . " Note t h a t if (g,\) is any Euclidean motion, (68) tells us t h a t we must first transform if; into the function x —> i/f(^_1x — # _ 1 v ) , and then multiply the result by the matrix Dj(g*), where g* lies above g. There is of course an ambiguity in the choice of g*; but this only changes Dj(g*) into + Dj(g*) and the resulting lattice automorphism depends only on g. For integral values of j , there is no need to go to G*, as Dj defines already a representation of G0 itself. From (68) we can obtain the expressions for the various momenta corresponding to t h e elements of t h e Lie algebra of G. If v e R 3 , a n d we consider t h e translations x -> x + 1 \ of R 3 , we obtain, for t h e corresponding linear momentum By, the expression (69)
£ v :<£->
-id(v)ijj,
where d(\) operates componentwise. This is exactly the same as in (32). On the other hand, we now have in addition the angular momenta, which correspond to the various one-parameter subgroups of 6r0*. Let Xl9 X2, and X3 be the elements of the Lie algebra of G0* which correspond to the one-parameter groups g±, g2, and g3, respectively. If Mlf M2, and M3 are the corresponding momentum operators, we have, from (62), (63), and (68), for r= 1,2, 3, (70)
Mri/; = i{dri/j + I)j(Xr)ilj},
where 0i =
r #2 o
^3 "o
ox2 (71)
52 =
d
x
3
ox3 x
"5
d l -7,— »
cxx d3 =
ox3
d dx±
d dx2
and
(72)
t)'{XT) =
{itD>(g*{t))Y_.
I t is usual to write Mr° for idr and Mrs for the operator I/J - > iDJ'(Xr)ifj. Mr is called the angular momentum around the The operators Mr° and Mrs are actually defined by the relations exp(-^ifr°) = ex$(-itMrs)
Atr,
= Btr,
KINEMATICS
AND
DYNAMICS
r
311 r
where t -> At is the one-parameter group defined by (At ip) • (x) = ijj(S(gr*(t))~1x), and t ~> Btr is the one-parameter group defined by (Btri/j)(x) = Dj(gr*{t))ilj(x). Since Atr and Btk commute, so do Mr° and (73)
[Mr°,Mks]
= 0
(r,k = 1 , 2 , 3 ) .
Mr° is called the orbital angular momentum around the ic^-axis, and Mrs is called the spin angular momentum around the # r -axis. By lemma 8.2, Mr° is essentially self-adjoint on C c °°(R 3 ,Tf); Mks is bounded. Moreover, Mr is also essentially self-adjoint on Oc°°(R3,TF). We note t h a t the group (r 0 *, as well as its one-parameter subgroups, is compact. Hence for any element X of the Lie algebra of G0*, the operator Bx (defined by exp( — itBx)= ViexptX) will have a discrete spectrum. From the formulas (62) it follows at once t h a t the angular momentum observables Mr have only the half integral numbers 0, ±\, ± 1 , + § , • • • as their possible values. This was one of the earliest facts recognized in quantum mechanics and emphasized by Niels Bohr (cf. Weyl [1], pp. 185-191). The formulas (54) show t h a t insofar as the translations of M are concerned, we have the usual system of imprimitivity for the translation group. However, if j^O, this restricted system is no longer irreducible. Therefore there will be observables which will be simultaneously measurable with respect to the linear momenta and the configuration observables. The operators M±s, M2S, and M3S correspond to such observables as can be directly verified from the equation Mrsifj =
iD3\Xr)ifj.
These are usually called the spin observables. Note t h a t each Mrs acts only on the components of I/J, and not on the argument x of ifj. The possible values of Mrs are just 2j + 1 in number; these are j , j — 1, j — 2, • • •, — j . The parameter j , which enters so basically in the description of the system, is called the spin of the system. The case j = | is of special interest. I n this case, t h e ^ ' s m a p R 3 into C2, so t h a t each iff can be written as I , ) , where I/J1 and I/J2 are square integrable on R 3 . A brief calculation t h e n shows t h a t t h e operators Mr8 are described by (74)
ifrs:^->-Jar^,
where al9 o-2> a n d ^3 are described by (55). The observables Mrs have only the two possible values ± \. W i t h this we have completed what might be called the kinematical description of covariant systems with a Euclidean configuration space. We have obtained the usual expressions for the familiar observables and
312
GEOMETRY
OF QUANTUM
THEORY
shown t h a t the introduction of the spin observables is an inevitable consequence of the geometry of the configuration space. This seems to be the mathematical meaning of spin.
6. P A R T I C L E S We shall now use the analysis given in Sections 1 through 5 to develop some of the formulas and expressions which are at the heart of nonrelativistic q u a n t u m mechanics, namely those dealing with the mechanics of one particle. We shall, however, not go into m a n y details. I n the next chapter we shall make a deeper examination of the relativistic nature of a particle. Our present remarks are t h u s intended to be brief and serve only the purpose of a formal description of a few standard facts of q u a n t u m mechanics. We shall first examine the usual description of a " f r e e " particle. Clearly one would like to associate with a free particle a configuration space M and therefore give meaning to statements about the position of the particle. At the same time, the basic nature of such a system requires a covariance with respect to the whole group of motions of M. We assume t h a t M is a three-dimensional Euclidean affine space. Then the results of Section 5 m a y be applied to yield the result t h a t these systems are labeled by the spin parameter j . To specify the system completely we must determine also its Hamiltonian. Now the Hamiltonian H cannot be an arbitrary self-adjoint operator since one would like our system to possess the usual rules of conservation of momenta. According to our principles (cf. Section 1), the requirement t h a t the momenta be integrals of motion leads at once to the condition t h a t the dynamical operators ex-p(-itH) commute with the operators Vg* which correspond to the various elements g* of G* (cf. (13), (14)). Let us now determine the most general self-adjoint operator H such t h a t exp( — itH) commutes with all the operators of the representation V. We consider the case when the spin is 0. We choose a point x0 e M and a Euclidean coordinate system with x0 as origin. We can then identify M with R 3 and t h e Hilbert space of t h e system with j£?2(R3). If (g*,Y)eG* and ^eJS? 2 (R 3 ), (F (ff ., v #)(x) = ^ ( % * ) - H x - v ) )
(xeR3).
A
Let &: ^r -> r/r be the Fourier-Plancherel transformation of J£?2(R3) onto itself (cf. (41)). Then a straightforward calculation shows t h a t (75)
( ^ F < 9 . , v ) t r - V ) W = exp[-i<x,v>] •?>(%*) " M -
KINEMATICS
AND
DYNAMICS
313
1
Let K = !FH.!F~ . Then the fact that K commutes with all the operators ^ ^ d . v ) ^ - 1 implies that K is the operator of multiplication by some Borel function k (possibly unbounded). If we now use the condition that K commutes with all the operators ^ F ( 9 % 0 ) ^ ' - 1 , we can conclude that for each rotation g, k(x) = k(g~1x) for almost all x. From this it follows quite simply that k must coincide almost everywhere with a Borel function of x12 + x22 + x32. We may therefore conclude at once that H must be a Borel function, in the sense of the functional calculus of spectral theory, of the Laplacian, V:
<76>
a2
d2
d2
V= ^ +— +- dx±2 dx225 dx3 2
A similar argument shows that, conversely, for any Borel function ^(V), the operators exp[ — itk±(V)] commute with all the operators of the representation V, so that in the system with Hamiltonian k±(V), all the linear and angular momenta will be conserved. Covariance with respect to the group of motions of M cannot therefore pinpoint the Hamiltonian H of a free particle other than up to a function of V. The usual form of H can be obtained only after we introduce the notion of relativity. If we demand that the system be covariant (in a suitable sense) under the full space-time Galilean group, then one can show that the only possible Hamiltonians are of the form
<77>
H
= -k v>
where m is a real number; we choose m > 0 so that the spectrum of H is bounded below (positivity of energy). We shall not go through the calculations here; we shall prove (77) in the next chapter, where we take care of the case of arbitrary spin also. The formula (77) is truly remarkable; for, if we notice that the operator pr=z —i(d/dxr) corresponds to the linear momentum along the #r-axis, then (77) can be rewritten as
(78)
H=
±(Pl*+p2z+p3*).
Formula (78) is also true if we interpret both sides classically; H is then the well known Hamiltonian for a free particle of mass m. Thus equation (78) further deepens the formal analogy between classical and quantum mechanics. If we are dealing not with a free particle but one moving in a potential field, then it is natural to assume that the expressions for the momentum and position observables retain their validity. But the expression for energy will now have to be different. It is natural to retain (78) for the
314
GEOMETRY
OF QUANTUM
THEORY
kinetic energy of the particle and add to it a term involving the potential energy. Now the potential energy must be regarded as a function of the position of the particle only and hence must be represented by an operator which is a function of ql9 q2, and q3. We thus arrive at the expression of the Hamiltonian for a particle moving in a potential field: (79)
Hs = 1- (Pl* +p2* +p3*)
+S(qi,q2,q3),
where S(q1}q2,q3) is t h e operator of multiplication by t h e function S on R 3 ; the operator S represents the potential energy of the particle. Strictly speaking, one must be rather careful about (79), since we are adding to the operator (ll2m)(p12-\-p22+p32) another self-adjoint operator of possibly unbounded character. However, in a large number of physically interesting cases there is no ambiguity about the definition of Hs as a self-adjoint operator. The formula (79) would then complete the description of the one-particle systems. Consider now a special case when S is a C00 function and of a sufficiently simple type so t h a t Hs is essentially self-adjoint on the space of 0 0 0 functions with compact supports. If we now use the Heisenberg picture and compute the time derivatives of the pr's and g r 's we find, from (79) and (14), on CC™(B3), (80)
p =
Fr
qr =
—dH*ldqr, ' 3Hsidprl
where dHsjdqr is the operator of multiplication by 3S/dqr, and dHsjdpr is the operator (l/m)pr. These are the analogues of the classical Hamiltonian equations in their operatorial form. They formed the point of departure for Heisenberg's classic approach to the problems of the atomic systems. On the other hand, (79) was the starting point of Schrodinger's investigations. These two approaches, together with the statistical interpretation of Max Born (cf. (39)), were the beginning of the epoch which led to the modern revolution in physics (cf. Weyl [1]). We conclude with two remarks bearing on our special assumptions on the configuration space. Our main assumption is t h a t M is a Euclidean affine space. This in fact was decisive since it determined the structure of its group of motions and cleared the way for the application of the main theorems dealing with systems of imprimitivity. However, there are other interesting classes of spaces on which the motions act transitively; for example, the Riemannian manifolds with constant negative curvature (cf. Helgason [1]). I n this case, it would be of interest to carry out an analysis analogous to what we have done, at least when the curvature is small. I n particular, it would be of interest to compute the spectra of
KINEMATICS
AND
DYNAMICS
315
various important Hamiltonians as functions of the curvature e, for small £<0.
Our second remark concerns the existence of configuration space. This led to our being able to calculate, for any state of our system, the probability that a particle is found in a specified region of the space M. However, there are interesting physical arguments which prohibit, for certain particles Hke the photon, the possibility of experimentally verifying whether they are in a specified region of space. Such particles cannot be described, even kinematically, by the considerations of this chapter. We shall examine these questions from a different and more unified point of view in the next chapter. NOTES ON CHAPTER VIII 1. Quantization. Roughly speaking, Quantization refers to a process by which one establishes a "correspondence" between certain classical and quantum systems. The mathematical requirements of this correspondence are, however, capable of being formulated in many ways and so there are several approaches to quantization. Historically it was Niels Bohr who first understood the importance, and even the necessity of constructing such correspondences. His celebrated "correspondence principle" was based on his discovery that for electron orbits with "large" quantum numbers, his new frequency energy relation approximated quite closely the classical relation; here large means large in relation to fi. Thus it became possible to view Quantum Mechanics as a generalization of classical mechanics whose results were approximated by those furnished by classical theories when fi, now treated as a parameter, tends to 0. With the discovery of Quantum Mechanics by Heisenberg, Schrodinger, and Dirac, the relationship between classical and quantum theories became clearer. In the Schrodinger representation, the classical state space is the phase space R2n = Rn x Rw while the quantum Hilbert space is L2(Rn), and one "associates" to classical observables which are functions of qlt...9qn9 Pv->Pn, the quantum observables obtained by the correspondence q.-^Q = multiplication by qh
However, since the P , and Qj do not commute, such correspondences are not precisely defined even at the formal level, and so one needs additional prescriptions. The first rigorous quantization appears to be that of Hermann Weyl [1] (pp. 272-280) who associated to the classical function a =
a(qv...,qn,pl9...9pn)
316
GEOMETRY
OF QUANTUM
THEORY
the operator A where A =
\a(u1,...,un,v1,...,vn) x e^p{ih"1(t^P1 + . . . + unPn + vxQx + . . . + vnQn)} duv. .dun dvx... dvn,
where a is the inverse Fourier transform of a and the exponential is interpreted via the unitary representation of the Heisenberg group. Since then this theme has been examined repeatedly and from many other points of view. It is natural to replace the classical phase space R2n by a symplectic manifold of dimension 2n, or at least by the cotangent bundle of a "configuration manifold" of dimension n. The questions involve the differential geometry and cohomology of these manifolds as well as the analytical aspects of symplectic geometry and Fourier transforms. The following is a partial list of sources. M. Flato and D. Sternheimer, in Harmonic Analysis and representations of semisimple Lie groups, edited by J. A. Wolf, M. Cahen, and M. de Wilde, Reidel, Dordrecht, 1980, pp. 385-448. V.Guillemin and S.Sternberg, Geometric Asymptotics, Mathematical Surveys, No. 14, Amer. Math. Soc, Providence, R.I., 1977. B . Kostant, in Lectures in Modern Analysis and Applications, III. Lecture Notes in Mathematics, No. 170, Springer-Verlag ,Berlin, 1978, pp. 87-208. J. Leray, Lagrangian Analysis and Quantum Mechanics, M.I.T Press, Cambridge, Mass., 1981. V. P. Maslov, Theorie des perturbations et methodes asymptotiques, suive de deux notes complementaires de V. I. Arnold et V. C. Buslaev, Paris, Dunod, 1972 (tr. by J.Lascoux and R.Senor). J. E.Moyal, Proc. Cambridge Phil. Soc, 45 (1949), pp. 99-124. J. M. Souriau, Construction explicite de Vindice de Maslov Applications, Fourth International Colloquium on Group Theoretical Methods in Physics, University of Nijmegen, 1975. 2. Second Quantization. This is a framework for treating systems of identical particles from the quantum mechanical point of view and studying their behavior when the number of particles is large. It goes back to P. A. M. Dirac's famous paper on radiation theory (Proc. Roy. Soc. London, A, 114 (1927), pp. 243-265) and is fundamental in quantum statistical physics and quantum field theory. If J f is the Hilbert space associated with a single particle, then second quantization associates to JT a Hilbert space £? in an invariant manner; 3F will be the space corresponding to an unlimited number of particles identical with the original one and will have additional structure determined by the existence of certain operators. The key features present in J f may be described as follows:
KINEMATICS
AND
DYNAMICS
317
(a) The operator N = the number of particles. There is a self-adjoint operator N with pure discrete spectrum and eigenvalues 0,1,2,...; if £Fn is the eigenspace for the eigenvalue n, states corresponding to the rays in Jfn are those in which the system has exactly n particles. Also <&C
Y
— ***' •
The vectors of Jfn may be denoted by \n)\ thus N\n) = n\n). (b) Vacuum state. The space J f 0 is one dimensional, dim(Jf70) = l. The state corresponding to this space is called the vacuum. (c) Occupation numbers: The operators N(y). If (
(i) A(
(F-D)
(i) A(
(In this brief discussion we do not insist on domain considerations in interpreting these rules.) The corresponding particles are respectively known as
318
GEOMETRY
OF QUANTUM
THEORY
Bosons and Fermions. It is a fundamental empirical fact that particles with integral spin are Bosons while those with half-integral spin are Fermions. (e) Cyclicity of vacuum. It is natural to assume that A(
{9ieX)
span a dense subspace of X. (f) Covariance. There is a strongly continuous unitary representation r of the unitary group of X in Jf, commuting with N. Thus T(U)Jf n = Xn for all n, Ue unitary group of Jf. It is not difficult to show that once the type of the particle is specified (that is, a Boson or Fermion), then there is essentially one X associated to JT. X is called the Fock space; it was first discovered by V. Fock. We give a brief description of it. Bosons. X = 8(Jf)i the symmetric algebra over Jf\ This is defined essentially in the same way as in the case when dim( JT) < oo. More precisely, let T(Jf) = T be the covariant tensor algebra of Jf, i.e.,
T = c i 0 j r e p f ®jf)e... = ©pr®m), where we take the infinite Hilbert space direct sum. Let Js be the closed linear span of all tensors of the form u ® (x ® y — y ® x) (x) v, where x,yeX,u,ve
2 W > 0 ^ ® m = T°- Then 8 = T/J8. dfn
As in the finite dimensional case, we have a linear Hilbert space isomorphism 8 — 3 7e yn im
where T**mm is the subspace of symmetric tensors:
8 is of course graded, the grading being inherited from T: 8 = (+) 8mi
Xm — 8m;
and the algebraic sum
s°= 2 sm dfn m>0
KINEMATICS
AND DYNAMICS
319
is an algebra. The operators A(cp) and A((pY are unbounded, they are first defined on 8° and then extended by formation of closures. On S° they are defined by A(q?)r : u h-> Vm -f- 1 qm
A(
(ueSm),
(ueSm),
where d((p) is "differentiation" along
= 1
show that the A(q>) and A(yY are bounded operators. For the standard model we choose J f = A, the exterior algebra of Jf. It is defined as T/Ja where Ja is the closed linear span containing all tensors of the form u ® (x 0 'J + y ® x) 0 v, where x, yeJf, u, VGT°. A is graded in the obvious way; Am = 3tf?m is the image of jT®m under T-»A. If we write A0 = Yi Am
(algebraic sum),
A0 is an algebra. As usual A denotes multiplication in A. The creation and annihilation operators are defined by A(
(ueAm),
A(
(weAJ,
where d(q>) is the unique endomorphism of A0 such that (i) (ii) (hi)
3(^)1=0, d(
If U is a unitary operator of JT, T( U) is the corresponding naturally induced unitary operator of A; T(U) is an algebra automorphism of A0. Finally |0> = 1.
GEOMETRY
320
OF QUANTUM
THEORY
In the Fock representation for a B-E system choose a real form JTR of J f and write P{9)
A(9) +A(9y
—vi—'
AW-AW)
m
~
*V2 •
Then we have
[P(9),pm=o,
[Q(9),Qmi=o
mvhpm]=KvS-
These are analogues of the Heisenberg commutation rules, and reduce to them when dim(JT) < oo. The uniqueness of the irreducible representation of the Heisenberg commutation rules (in the " Weyl" or "integrated " form) when dim(jT) < oo makes clear why the B-E system is unique. We can in fact use the Schrodinger representation to construct another model for the Fock representation. Let J^ — L^Jfg) (with respect to the Lebesgue measure) and let P((p) = -. d((p),
Q(
(multiplication by <
Then it is enough to take AW)^M-iQW)>
A(
y_m+iQ(i>)m
V2
V2
The vacuum vector is the function exp(-i|N| 2 ). This type of model can be constructed even when dim(Jf) = oo. It is, however, technically more complicated because there is no translation invariant (or translation quasi invariant) measure on Jfj^ when dim(Jf) = oo, and one has to use a more sophisticated type of integration theory, namely integration with respect to the isotropic Gaussian weak distribution dg on JTR (cf. I. E. Segal in Les Problemes mathematiques de la theorie quantique des champs, CNRS, Paris, 1959, pp. 57-103). The identification L2(JfRidg) « J f allows us to "quantize" any classical dynamical system whose configuration space is JfR and whose dynamics is linear and unitary on Jf. If (Ut) is the classical dynamical group, (T(Ut)) is the quantized dynamical group; moreover, the grading J f = 0 n Jf?n leads to a particle interpretation. This was how Dirac quantized the electromagnetic field and constructed the quantum theory of radiation, namely interaction of light with matter. For the Fermion case, we keep the real form JfR as before but now define
P(9) = A(9)+AW,
iQ(y) =
AW-A(V).
KINEMATICS
AND
DYNAMICS
321
Then P and Q are linear maps of JTR into the space of bounded self-adjoint operators on #? satisfying P(
= QM2 =
<
{P(
This suggests the introduction of the Clifford algebra, Cliff (j£?) associated to a real Hilbert space j£f, which is a complex algebra with unit 1, containing j£? and generated by it, such that u2 = (u,u)l
ue£?.
ClifF(J^) exists and is essentially uniquely determined by <£. If we put B(
and write j £ for the real Hilbert space underlying Jf, (<^>jr = Re<w,v» then (P,Q)»R is a bijection between F-D systems based on JfR and representations R of Cliff (JT) such that R(u) is bounded self-adjoint for all ue.JT. For a more systematic study of the framework of second quantization the reader may refer to F. A. Berezin, The Method of Second Quantization, Academic Press, New York, 1966, and J. M. Cook, The Mathematics of Second Quantization, Trans. Amer. Math. Soc, 74 (1953), pp. 222-245. In the original article of Dirac referred to earlier, he used this framework to study the interaction of an atom with the electromagnetic field and obtained the known formulas for probabilities of transitions between the energy levels. For a detailed treatment see Dirac, loc. cit., and von Neumann [1] (pp. 254294). Dirac's paper inaugurated the beginning of an entire era in which a major goal was to erect a theory of quantized fields obeying the principle of special relativity. To get a picture of what has been done, as well as what remains to be done, we refer the reader to the following books and to the articles referred to in them: J.Schwinger (Ed.), Selected Papers on Quantum Electrodynamics, Dover, New York, 1958. S.Schweber, An Introduction to Relativisitic Quantum Field Theory, Row, Peterson and Co., New York, 1961. J. M. Jauch and F. Rohrlich, The Theory of Photons and Electrons, SpringerVerlag, Berlin, 1980. N. N. Bogoliubov and D. V. Shirkov, Introduction to the Theory of Quantized Fields, Wiley, New York, 1980. J. Glimm and A. Jaffe, Quantum Physics: A Functional Integral Point of View, Springer-Verlag, Berlin, 1981.
CHAPTER IX RELATIVISTIC FREE PARTICLES 1. R E L A T I V I S T I C I N V A R I A N C E
The fundamental assumption in Newtonian mechanics is the absolute nature of space and time. That our empirical knowledge of the physical world points up to the contrary was one of the cornerstones of Einstein's great critique of space and time. The analysis of Einstein and Lorentz of t h e empirical and mathematical nature of the physical phenomena established t h a t only space-time, as a four-dimensional manifold, has an invariant physical significance, and t h a t its familiar splitting into space and time is essentially dependent on the observer. Each observer was t h u s seen to describe only the points of the space-time manifold in his vicinity by four coordinates; the formulas describing the transformation of coordinates of the same space-time point by two different observers involved smooth functions. I n other words, space-time became an abstract C™ manifold and each observer was seen to describe a neighborhood of this manifold in terms of local coordinates. Only statements which retain their meaning in every coordinate system were asserted to have any physical significance, i.e., the global differential geometric statements. Einstein's theory of gravitation is, for example, a formulation of t h e theory of gravitational phenomena involving only intrinsic geometric objects like connections and tensor fields. I n problems dealing with the q u a n t u m aspects of atomic systems, it is usual to assume t h a t gravitational fields have no influence on the physical phenomena (cf. Dirac [1], p. 253). As a consequence of this assumption, the group of transformations which connect the different observers is drastically reduced. We shall make the assumption t h a t the observers are what are known as inertial, and t h a t the relevant group of transformations is the so-called group of special relativity. We shall now proceed to make a few explanatory remarks on these concepts. I n special relativity it is assumed t h a t each observer is able to establish coordinates for the whole of the space-time manifold and, moreover, t h a t he uses a rather special coordinate system to achieve this. One makes the assumption t h a t the notion of a classical free particle, not acted upon b y external forces, is meaningful for the observer, and t h a t the trajectories 322
RELATIVISTIC
FREE
PARTICLES
323
of such a particle are described by linear functions of the time coordinate in his coordinate system. I t is customary t o call such coordinate systems and observers inertial. I t m a y be noted t h a t an inertial coordinate system is simply one in which the first law of Newton is valid—every particle is a t rest or continues in its state of uniform motion in a straight line, unless it is acted upon by external forces to do otherwise. Given two inertial coordinate systems, say 0, 0', each of which maps the space-time manifold in a one-one fashion onto R 4 , a correspondence T0t0, is established between t h e points of R 4 ; (x0ixvx2,xz) and (x0*\x± ,x2>3') correspond under T0>0, if and only if they represent t h e same space-time point, in 0 and 0', respectively, j W h a t can one say about the mapping (1)
TOQ,,
1 QQ, : (x0,x1,x2>XQ) —> (x0 ,X± ,X2 ,X3 ){
I t is natural to expect TQ Q. to be an inhomogeneous linear transformation; for, as both coordinate systems are inertial, the trajectories of a free particle must be represented by straight lines in both systems, i.e., T00, must m a p straight lines into straight lines. Hence it m u s t preserve coplanarity and parallelism. Consequently, if we pass (through t h e use of homogeneous coordinates) to the associated projective geometry of R5, T00, will induce an automorphism of this projective geometry and is therefore described in R 5 by a unique linear automorphism (cf. t h e results of Chapter I I ) ; back in R 4 , this would lead to t h e affine nature of T0i 0,. I t is roughly along these lines t h a t one is led to the basic hypothesis of special relativity: space-time is an affine space%; each inertial observer describes it with an affine coordinate system, and the transformations connecting the coordinates of the same point in two different inertial systems are inhomogeneous linear: 3
(2)
x- = ^
ciijXj + Ut.
For three inertial observers 0, 0'', and 0", we have: =
(3)
-* o,o"
-L o',o"To,o'i
W
J- o',o — •*• o,o'-
Clearly the set of all TQO, forms a group. This group is called the spacetime relativity group. I t has been obtained explicitly as a group of inhomogeneous linear automorphisms of R 4 . I n t h e abstract affine space-time it gives rise to a group of affine automorphisms. t xn and XQ sir© the time coordinates: x^ x/ (j= 1, 2, 3) the space coordinates. J Cf. Chapter V I I I , Section 4.
324
GEOMETRY
OF QUANTUM
THEORY
However, not every inhomogeneous m a p of the form (2) can arise as a transformation between inertial coordinate systems. Further examination of the empirical nature of space-time phenomena yields certain restrictions on the T00>. Before Einstein and Lorentz, it used to be assumed implicitly t h a t the notion of simultaneity of two events was an invariant one, i.e., if 0, 0' are two inertial coordinate systems, if ( ) and (yo,yi,y2,2/3) are the points of i£ 4 which correspond under 0 to two spacetime points, and (x^x^x^x^) and (yo',yi ^2^3) represent the same pair of points under 0 ' , then x0' =y0' if and only if x0 = y0. I t follows from this t h a t in the formula (2) for T00>, we must have (5)
x0
=
(IQQXO-{-UQ.
If we assume t h a t the unit of time measurement and the sense of time propagation is the same in both systems, we must have a 00 = l, i.e., (6)
x0' = xQ + u0.
Furthermore it was assumed t h a t the spatial distance of pairs of spacetime points with the same time coordinate also remained unchanged, i.e.,
(7)
2 (Xi-yd2 = f «-2/iTi=l
i=l
From this it follows t h a t the matrix of order 3 x 3 in (2), (8)
W = (at,)la:Us3
is orthogonal. Consequently, if we write x for the column vector with components x1} x2, and x3, x' = Wx + x0v + u, (9) , _ ° XQ
XQ-{-UQ,
where W is an orthogonal 3 x 3 matrix, v , u e R 3 , and WQGR 1 . The collection of all transformations (x 0 ,x) -> (x0',xf) of the form (9) is easily verified to be a group. We now assert t h a t every transformation of the form (9) is a possible transformation between two inertial coordinate systems. Since the possible transformations constitute a group, it is enough to exhibit a class of transformations which actually arises in practice and which also generates the group of all transformations of the form (9). If we p u t in (9) ^ 0 = 0, u = v = 0, we obtain a transformation which connects two coordinate systems with different coordinate axes but with the same space and time origin, and the same sense of time flow, and which are mutually a t rest; evidently all such transformations are possible. If we p u t W = l and v = 0, we obtain transformations which connect two coordinate systems with different space and time origin, have the same sense of time flow,
RELATIVISTIC
FREE
PARTICLES
325
have coordinate axes in the same orientation, and are mutually at rest. Clearly all such have to be included in the relativity group. Finally, if we put W=l, ^ 0 = u = 0, we obtain the transformations x' — x-r x0v,
(10)
The formula (10) represents the transformation which connects two coordinate systems which have the same time origin and sense of time flow, which are moving uniformly with respect to each other with velocities v and —v, and whose frames coincide for XQ — XQ ' = 0. These also have to be admitted. An easy calculation shows that these three classes of transformations generate the group of all transformations of the form (9). In other words, the physical assumptions we have made lead one to the group described by (9) as the relativity group. It is called the orthochronous inhomogeneous Galilean group, and the corresponding physical principles go under the name of Galilean relativity. If we assume det W — -f1 in (9), we obtain the usual inhomogeneous Galilean group. The transformation x' = x,
(ID XQ
—
XQ,
represents time inversion. If we add this to the orthochronous group determined by (9), we generate a larger group called the complete inhomogeneous Galilean group. When dealing with bodies and particles that move with very high velocities, these principles are violated quite decisively. The well known Michelson-Morley experiments led to the remarkable conclusion that in every inertial coordinate system, the velocity of propagation of electromagnetic and light signals in vacuum remained constant. It was Einstein who embodied this as a basic principle and deduced from it the consequence that the affine transformation (2) must be such that the same numerical value is obtained in both coordinate systems for the velocity of these signals. Choosing this numerical value to be the unit of time measurement, this means that for corresponding points of R4, the relation
(12)
(y0-x0)2-
1 (2/y~*y)2 = (
(y/-x/)a
;• = l
must be satisfied. If we now substitute (2) in (12), we find, for the 4 x 4 matrix, (13)
A = ( a , X = o,
the equation (14)
A* FA = F,
326
GEOMETRY
OF QUANTUM
THEORY
where F is the diagonal matrix: (15)
i^ = d i a g { l , - 1 , - 1 , - 1 } ;
equivalently, t h e linear transformation induced in R 4 by A preserves t h e Minkowskian form (16)
:x02-x12-x22-x32.
&
The set of all A satisfying (14) is called the Lorentz group, and the set of all affine transformations (2), with A Lorentz is called the complete inhomogeneous Lorentz group. I t can be shown (cf. Section 2) t h a t |«oo| ^ 1 a n d t h a t the set of all transformations with aQ0 > 1 is a subgroup, t h e orihochronous subgroup. Exactly as we showed in the Galilean case, it is possible to argue t h a t the entire orthochronous Lorentz group must be admitted as the group of special relativity. To prove this, it is again sufficient to exhibit a generating class of transformations which actually arise. The transformations x' = Wx + u, (17) all arise explicitly; they connect inertial coordinate systems which are mutually at rest and which differ in their space and time origins and the orientations of their spatial coordinate axis but have the same sense of time flow. Consider, on the other hand, the transformations X0' = (l-V2)-ll2Z0 Xl'
= vil-V^-^Xo
x
= x2,
+ +
v(l-V2)-ll2Xl9 il-V2)-11^,
(18) 2 XQ
=
X3,
for values v with — 1 < ^ < 1 . Equation (18) represents, for fixed v, a homogeneous Lorentz transformation. The space-time point which corresponds to the point (r,0,0,0) in the O'-system is represented by (T(1-V2)-112, -rv(l-v2)-112, 0, 0) in the O-system. This shows t h a t (18) represents the transformation which connects 0 and 0' when 0 is moving relative to 0' with uniform velocity v along the a^-axis. Transformations of the form (18) must therefore be included in the space-time group. B y symmetry, the transformations in which x2 and x3 play the role of x1 are also admissible. I t can be shown t h a t these three classes of transformations, together with those defined by (17) and the time reversal (11), generate the complete inhomogeneous Lorentz group (cf. Section 2). I n other words, in the special relativity of Einstein and Lorentz, the space-time group in question is the complete inhomogeneous Lorentz group.
BELATIVISTIC
FREE
PARTICLES
327
I t is usual in physical literature to do all this, not in natural units, b u t in the usual, say, C.G.S. units. If c denotes the velocity of light in vacuum as computed b y inertial observers, then x0 = ct, where t is t h e time coordinate, and (18) will appear in the form
*(19)
.
l 2
2
Vl-v /c
Xl'
= — =2 =2 t Vl-v /c
Xo
=
*• +
vlc2
Vl-v2/c2 1 Vl-v2/c2
-
Xn,
When v/c —> 0, this formula becomes t' = t, (20)
XX + tVy
which are the formulas expressing the same relation between the two frames in Galilean relativity. I n other words, we m a y view t h e application of Galilean relativity as satisfactory when t h e velocities of t h e particles and bodies involved are small compared to c. For high-speed particles, the formulas (19) must be used, i.e., only Lorentz relativity is to be postulated. These remarks conclude our extremely brief discussion of special relativity. The reader who is interested in a more detailed analysis of space and time is referred to Weyl's treatise [6]. The main problems of relativistic particle physics concern the various high-speed particles and their interactions. A general theory of such phenomena must combine the principles of q u a n t u m theory with those of special relativity to obtain the relevant equations of motion and to draw the necessary physical conclusions. I n such generality these problems are still unsolved. I n this chapter we shall be concerned only with the free particles and their equations. Before entering into a discussion we shall formulate mathematically t h e requirement t h a t the physical description be relativistically invariant, i.e., the physical content of the systems under consideration be the same for all inertial coordinate systems. Consider now a system 0> (2) which connect pairs of inertial systems. Each observer 0 associates with (& a Hilbert space Jt?0 whose rays are identified by him with the dynamical or Heisenberg states (cf. Section 1 of Chapter VIII) of S . I n the coordinate system 0, S is described by the determination of the set of all its dynamical states as the rays of
328
GEOMETRY
OF QUANTUM
THEORY
J^Q. I n general, 3f0 will vary essentially with 0. We shall formulate the relativistic invariance of (the description of) <& by the requirement t h a t J4?o be independent of 0, i.e., the set of motions of © is the same in all inertial coordinate systems. We shall write J f for this Hilbert space. Even though J^0 is the same in all inertial systems, the same physical state of © m a y be described b y two entirely different rays of Jf by two different observers, 0, 0'. Hence we assume the existence of a one-one mapping t00> of the set of all rays of J f onto itself, (21)
t0t0,:r-+r',
such t h a t r and t00>[r] represent (in 0 and 0') the same state of
RELATIVISTIC
FREE
PARTICLES
329
where (22)
XQ = x0-a,
x- = Xj
(1 < j < 3).
Then the m a p (23)
a - > UT_a
( - 0 0 < a < oo)
gives the dynamical group of the system 8 . We have already seen t h a t the group R 1 has only exact multipliers and so we can replace (23) by a representation a -> Va = exip( — iaH). I n view of our analysis in Chapter V I I I it is natural to assume t h a t H represents the total energy of the system. If U itself is a representation, then, (24)
UTa =
exp(iaH).
Equation (24) can be generalized to describe the total linear momenta. Let A G R 4 , let a = (a 0 ,a), and let Ta be t h e transformation x - > # ' , where
x' = x —a. Then there is a unitary representation s -> Ws of R 1 such t h a t s -> W8 induces the same homomorphism into the projective group as s - » UTsa where a = (0,a). We write Ws = exip(isBa) and call Ba the total linear momentum in the direction a. We may, of course, consider all the spacetime translations at the same time by considering the restriction of U to the group of all space-time translations Ta. If we assume for the moment t h a t the multipliers of the relativity group have exact restrictions to the translation subgroup, we can obtain a representation a^Wa (a = (a 0 ,a)) which can replace the restriction of U to the translation subgroup. By the general theory of representations of abelian groups, there exists a unique projection valued measure Q on t h e dual P 4 of R 4 such t h a t (25)
Wa=
f
exp
i(a,p}dQ(p)
« . , . > is t h e duality between R 4 and P 4 ). The spectrum of Q is called t h e spectrum of the system. Notice t h a t the energy and the m o m e n t u m operators commute with one another. This leads, in the usual fashion (cf. Chapter VIII) to the law of conservation of the linear momenta. I t is clear t h a t those forms of the representation U in which Q appears in its canonical form will be especially simple to analyze mathematically. Such descriptions are called momentum space representations. Experiments involving the violation of parity have shown t h a t one does not always have invariance of the physical content of theories with respect to the Lorentz transformations which do not belong to the connected
330
GEOMETRY
OF QUANTUM
THEORY
component of t h e identity of G. Also it is well to remember t h a t only elements of the connected component of the identity of 0 are represented by unitary operators. For t h e other elements it is necessary to use both unit a r y and anti-unitary operators. We shall examine this question later on. We shall call a system elementary if the associated projective representation of the connected component of the relativity group is irreducible. I n intuitive physical terms we can describe this by saying t h a t starting from any state we m a y obtain all the states by Lorentz transformations and the principle of superposition. Two elementary systems m a y be said to be physically equivalent if there exists a unitary or anti-unitary isomorphism of the underlying Hilbert spaces which intertwines the associated projective representations. An elementary particle in its free state or a free particle is defined to be a physical equivalence class of elementary systems. One of the main results of this chapter is the classification of the elementary particles in terms of their mass and spin. I n classical and also in nonrelativistic quantum mechanics, the free particles do not play a very important role. However, the knowledge of the structure of the transformation laws for the states of free particles is very basic in most problems of particle theory. The reason for this is, roughly speaking, t h a t most of the experiments involve scattering of various kinds of particles, and the incoming and outgoing are clearly those of free particles. I t is because of this t h a t a study of the transformation laws of free particles is quite fundamental.
2. T H E L O R E N T Z G R O U P I n this section we collect a number of facts involving the Lorentz group. The notations t h a t will be used throughout this chapter are also introduced in this section. We begin with the homogeneous Lorentz group, denoted by 3. I t is the group of all 4 x 4 matrices (26)
L = (ay)?.,_ 0
such t h a t (27)
VFL
=• F,
where t denotes matrix transposition, and F is given by (28)
J = diag{l,-1,-1,-1}.
We write e0, e l5 e2, e2 for the usual basis vectors so t h a t (x0,x1,x2,x3) = x0e0+x1e1-\-x2e2-\-xze3. E a c h L induces a linear transformation of R 4 and (27) is the condition t h a t it preserve the quadratic form: (29)
&
:x02-x12-x22-x32.
RELATIVISTIC
FREE
PARTICLES
331
The group 3 is a closed subgroup of the group of all invertible linear transformations of R4, and as such is a lcsc group, even a Lie group. We shall examine its topology. Consider an L e 3. In view of (27), (30)
detL = ± 1 .
Further, L is an orthogonal matrix if and only if L preserves both the forms xx2 + x22 + x32 and x02. Hence such an L is of the form
(31)
L =
where W is a 3 x 3 orthogonal matrix. We consider next a positive definite symmetric matrix L. Let L112 denote the unique positive definite symmetric matrix whose square is L. We contend that if L is in 3, so is L112. In fact, from the equation LFL = F we obtain (L1I2)2 = (FL~1I2F)2, which shows, in view of the uniqueness of L112, that L1I2 = F(L1,2)~1F or that L112 e 3. By a trivial induction we conclude that L(ml2n) e 3 for all integers m and n. From this observation we can determine the structure of the set of all positive definite Lorentz matrices. Consider, to this end, the mapping
8->es of the vector space of 4 x 4 real symmetric matrices into the space of symmetric positive definite matrices. It is well known (cf. Che valley [1], pp. 14-15) that this is a homeomorphism. From what we saw just now, it can be asserted that whenever es e 3, e(ml2n)S e S for all m and n and hence that eaS e S for all real a. If we differentiate the equation (eaS)F(eaS) = F with respect to a, at a = 0, we obtain (32)
SF+FS
= 0
(8 symmetric).
Conversely, suppose that S is a 4 x 4 symmetric matrix satisfying (32). If we write
332
GEOMETRY
OF QUANTUM
THEORY
which satisfy (32) is a three-dimensional linear space p with a basis given by 1 M01 = |
0
0\
_
/0 I,
0
1
0\
i f 02 =
(33) 0 ^03
0
=
If we write (34)
Hp = {L : L e 3, L positive definite symmetric},
then we m a y conclude t h a t Hp is homeomorphic to the three-dimensional Euclidean space; S - > es is a homeomorphism of p onto Hp. I n particular, 1 e Hp and / P is connected so t h a t HP^H, where H is the component of the identity of H. Consider now an arbitrary L e S. Since F = F'1, the equation L~1F(Lt)-1 = F shows t h a t (L'^EH and hence D itself belongs to H, i.e., (35)
H* =
3.
We now examine the polar decomposition of L, i.e., L = L1(LtL)112, where L± is orthogonal. From (35) and our earlier discussion we know t h a t (LtL)ll2eHp and hence L± e 3. Hence Lx must be of the form (31). I n other words, every LeB can be uniquely written as 0 (36)
0
0\
L = e'W
where W e S0(3), e= ± 1, and e = + 1, and Sep, i.e., S is symmetric a n d satisfies (32). Since the 0-0th entry of es is seen to be > 1 by direct computation, it follows from (36) t h a t for L, Ko1
(37)
* *'
£d00 > 0. 3
Moreover, det L = e' so t h a t (38)
e'(det L) = 1.
RELATIVISTIC
FREE
PARTICLES
333
Thus e = 1 and e = 1 if and only if a00 > 1 and det L — + 1 . Collecting all these facts, we have Theorem 9.1. The component H of the identity of S is given by (39)
H = {L:a00
> 0,
det L = + 1 } .
The map 0
0
0\
W,8-
w is a homeomorphism of 80(3) x p onto H. For any Lorentz L, \a0Q\ > 1 and det L= + 1. H has, besides H, exactly three connected components and these are the H-cosets of H determined by the three elements Is, It, and Ist, where Is, It, and Ist are obtained from (36) by putting 8 = 0, W=l, and for e and e the respective pairs of values ( 1 , - 1 ) , ( — 1 , - 1 ) , and ( — 1,1). We write K for the subgroup of H corresponding to $0(3). The Lie Algebra. Since H is a linear group, we shall canonically identify its Lie algebra f) with a Lie subalgebra of matrices or endomorphisms of R*, the bracket being the usual commutator A, B ->[A,B] = AB — BA. The exponential m a p then becomes the usual one, 8 -> es. A m a t r i x 8 G I) if and only if eaS e 3 for all real a. Arguing as in (32), we find t h a t (40)
f>={S:StF
+ FS = 0}.
1) is six-dimensional. If we write I for the set of matrices (41)
I = {S :Set),
Se1 = 0},
then the elements 8 of X are precisely those of the form ^0
0
0
0\
0 (42)
8 = " 0
S±
Ko where 8X is skew symmetric: S^ =
-8X.
f is the subalgebra of f) corresponding to 80(3) imbedded in H. f) is the direct sum o f ! and p: 1) = ! + p, t n p = 0.
334
GEOMETRY
OF QUANTUM
THEORY
A useful basis for t) m a y be constructed in t h e following manner. L e t etj denote t h e 4 x 4 matrix whose only nonzero entry is t h e i-jth, which is 1 (i,j = 0, 1,2,3). Let (43)
MXj = -Mj{
(44)
= ei;-e;,
(i,j = 1, 2, 3, i # j),
Moi = Moi = eoi + eio
(i = 1, 2, 3).
Then M12, M23, M31, M01, M02, a n d M03 span t). The commutation rules are given b y :
(45)
[MwMjk]
= Mik
( i ^ j ^ k ^
[M 0i,M0i]
= Mtj
( i # j , i , j = 1,2,3),
[Moi,Mjk]
= 0
(< =^ j # £ # i,
[if w ,Jf o t ] = -M0J
(i # j ,
i,
i,j, i = 1, 2, 3),
i , j , * = 1, 2, 3),
t, j = 1, 2, 3).
The Covering Group. We shall introduce in a n explicit fashion t h e universal covering group of H. We consider t h e real four-dimensional space of 2 x 2 Hermitian complex matrices a n d t h e linear isomorphism of R 4 with it given b y (46)
A : (x0,xl9x2,x3)
-> X0G0 + X1G1+X2(J2
+ X3G3,
where
H* = 8L{29C)9
and for a n y m e H* t h e linear transformation (48)
S(m): $ - > m£m*
of t h e space of 2 x 2 Hermitian matrices. I n view of (46), 8(m) induces a linear transformation in R 4 . L e t us also write 8(m) for this transformation. From (48) we see t h a t m -> 8(m) is a homomorphism of H* into t h e group of invertible linear transformations of R 4 . Since d e t m = l , d e t £ = det m£m*, so t h a t t h e transformation 8(m) must preserve t h e quadratic form x02 — x±2 — x22 — x32. Thus S(m) e H a n d m - > 8(ra) is a continuous homomorphism of H* into 3, hence into H. From (48) it follows easily that (49)
kernel (8) = { ± 1 } .
8 is thus a local isomorphism a n d therefore its differential 8 induces a n isomorphism of the Lie algebra of H* into t h a t of H. B u t t h e Lie algebra of H*, being canonically isomorphic t o t h e Lie algebra of 2 x 2 complex matrices of trace zero, is, as a real vector space, six-dimensional. Conse-
RELATIYISTIC
FREE
PARTICLES
335
quently, 8 is surjective. From this we infer a t once t h a t 8 maps H* onto the connected Lorentz group H: (50)
S[H*] = H.
The topology of H* m a y be analyzed in t h e same way as t h e topology of H. The technical details are similar. Using polar decomposition of elements of H*9 we conclude t h a t H* is homeomorphic t o t h e Cartesian product of K*, its unitary subgroup, and the space of positive definite complex Hermitian matrices of determinant 1. Since S ±+ es sets u p a homeomorphism of t h e latter space with t h e three-dimensional real vector space of 2 x 2 Hermitian matrices of trace 0, we find t h a t H* is homeomorphic t o K* x R 3 . I n particular, H* is simply connected. 8 is t h u s a covering homomorphism of H* onto H. We note t h a t for an m e H*, 8(m) is orthogonal if a n d only if the m a p £ - > m£m* preserves tr(£) as well as det(£). This can happen if and only if m is unitary. Thus (51)
8[K*] = K,
K* =
8~1(K).
The Lie algebra representation 8 corresponding to 8 is easily computed. Let f)* be the Lie algebra of 2 x 2 complex matrices of trace 0, (52)
if = {X : t r X = 0,
X complex}.
Then (53)
S(X)£ = X | + fX*.
We take t h e basis {ia1,ia2,i<j3,a1,(j2>(T3} f ° r *)*• An easy calculation gives: 8(i°i) = 2if 2 3 , 8(0,) = 2M019
8(ia2) = 2M31, 8(a 2 ) = 2M02,
&(ia3) = 2Jf 1 2 , 8(a3) = 2 ^ 0 3 .
From (54) we see t h a t 8 maps Hermitian elements of f)* into symmetric elements of f). Hence 8 maps the set of positive definite elements of H* onto Hp. Further, for a unitary keK*, 8(k) e K. Hence, for a general element m of H* of t h e form rn — hp, k unitary, p positive-definite Hermitian, we have 8(m*) = 8(p)8(k-1) = 8(p)8(k)t = (8(Jc)8(p))t so t h a t (55)
8(m*) = 8(mY
(m e H*).
On t h e other hand, let £ be t h e reflection in J?4 given by
£ corresponds under t h e m a p (46), to the conjugation £ —> f of Hermitian matrices. Since mgm* = (mgm*) ~, we have (57)
8(m) = J8(m)£.
Combining (55) and (57), we find (58)
8(ml) = £S(m<)£.
336
GEOMETRY
OF QUANTUM
THEORY
Representations of SL(2, C) in Finite Dimensional Vector Spaces. The group H* has no finite dimensional unitary representations other t h a n t h e trivial representation, but it has a number of nonunitary representations in finite dimensional vector spaces. We shall now describe these. We write W = C2 and W3 for the subspace of symmetric tensors of rank 2j (j = 0, \, 1, • • •) in the tensor product W (g) W (g) • • • (g) W (2j factors). The representation m - > m ( g ) r a ( g ) - - - ( g ) m leaves W3 invariant. We write D3 for the representation of H* defined by W\ D3 is already irreducible when restricted to the unitary group K*. If {e1,e2} is the standard basis for W, the matrix entries of the representation of H* in W (g) • • • (g) W, relative to the basis {eri (g)- • -(g) erN} (N = 2j), are (59)
f{rl9 • • •, rN, sl9 • • •, sN; m) = ari8l • • -a r „ s „,
where m e H* is the matrix {ars)?tS = 1. The / ' s are complex analytic functions on H*. D3 is thus a complex analytic representation of H* whose restriction to X * is equivalent to the irreducible representation of K* associated with j . For any j , we m a y consider the representation D3 conjugate to D3\ i.e., we take a basis for W3 and conjugate the matrix representation obtained. D3 is also an irreducible representation of H*. Its restriction to K* is once again equivalent to the irreducible representations of K* corresponding to j . No Dk is equivalent to any D3 except when k = j = 0. I t is known t h a t for arbitrary j , f (j,j' = 0, i , 1,. • •) the representations (60)
D*'r : D3
are all irreducible and t h a t these exhaust, up to equivalence, the irreducible finite dimensional representations of H*. For a proof the reader m a y refer to Weyl ([2], pp. 267-268). The Inhomogeneous Group. The Lorentz group H acts in R 4 as a linear transformation group and hence one may form the associated inhomogeneous group, (61)
^ = J?x'R4
where we use the notation x ' to denote the formation of the semidirect product. For (L,x) and (L',x') belonging to G, (L,x){L',xf)
= (LL', Lx' + x).
The group H* m a y also be considered as acting on R 4 through t h e action (62)
mx = S(m)x
(m eH*,xe
R 4 ).
We write 0* for the associated semidirect product (63)
G* = H* x ' R 4 .
RELATIVISTIC
FREE
PARTICLES
337
The subgroup of G given by (64)
G = Hx'
R4
is the component of the identity in G, and the m a p (65)
8 : m, x -> 8(m), x
is the covering homomorphism of G* onto G; G* is obviously simply connected. The group G is generated by G and the "inversions" (Is,0), (It,0), and (Ist,0), where Is is space inversion, It is time inversion, and Ist is spacetime inversion: writing (x0,x) for x0e0 + x1e1 + x2e2 + x3e3 Is : (66)
{x0,x)->(x0,-x),
It : ( : z 0 , x ) - > ( - # 0 , x ) , Ist : ( z 0 , x ) - > ( - : r 0 , - x ) .
The elements Ix (r = 0, s, f, 5f; IQ is the identity) form an abelian group having four elements, the group of inversions, which is denoted by Hinv: (67)
Him
=
{I0,Is,It,IJ.
The group Hinv is a discrete subgroup of the homogeneous Lorentz group H, and H is the semidirect product of Hinv and H: (68)
H = HinyH,
HnHinv
= {l};
the action of # i n v on H is of course through inner automorphisms. For any r, L-> IXLIX~X is an involutive automorphism of H, and the assignment which sends r to this automorphism is a homomorphism. As H* is the universal covering group of H, there exists for each T, a unique automorphism h ->hT of H* such t h a t (69)
IMh)^-1
8(hT) =
for all h. h->hx is involutive and the assignment s, which sends r to the automorphism h->ht, is necessarily a homomorphism. We m a y thus form the semidirect product (70)
H* =
HinvxsH*.
We continue to write 8 for the m a p IT,h^8(h)IT; 8 is a homomorphism of Hinw xs H* onto H. The complete inhomogeneous group G is the semidirect product of G and # i n v ; each Ix gives rise to the automorphism (L,x) ->
(ItLIT,Ixx)
338
GEOMETRY
OF QUANTUM
THEORY
of G. The corresponding m a p (h,x)->(hl9I^) is seen to be an automorphism of G*, and the assignment s, which sends r t o this automorphism, is once again a homomorphism. We m a y thus form the semidirect product G*, G* = HlDV xsG*. The group G* has a natural homomorphism 8 onto G: (71)
8
:IT,h,z-+(8(h)I%,x).
I t is useful to determine explicitly the automorphisms h->hr. r = st, IstLIst~1 = L so t h a t (72)
For
hst = h.
Suppose now t h a t r = t so t h a t Ix is time inversion. I t is clear t h a t It = — F, F being the matrix of the fundamental form. Hence 8(h~1*) = (8(h)~1)t (cf. (55)) =F8(h)F so t h a t (73)
h'1*.
ht =
As IsIt = Ist, we have (74)
hs = K
Orbits in Momentum Space. The space R 4 is a locally compact abelian group under addition and hence possesses a dual group. We identify this group with the vector space dual of R 4 . To keep things explicitly separated we write P 4 for t h e vector space dual of R 4 . Elements of P 4 will be written (75)
p = (POIP^PKPS)
= (Po>P),
and we shall define the canonical duality between R 4 and P 4 b y (76)
{x,p} =
x0p0-x1p1-x2p2-x3p3.
W e associate with each # e P 4 the character p : x -+ exp i{x,p} of R 4 . The mapping p->p maps P 4 isomorphically onto t h e character group of R 4 . We t h u s identify P 4 with the character group of R 4 . Since each element L of H is a matrix, it acts in R 4 as well as in P 4 . I n view of the definition of {.,.} and p, we have: (77) (78)
{Lx,Lp} p(L-^x)
= =
{x,p}, (LPr(x).
RELATIVISTIC
FREE
PARTICLES
339
We now compare (77) and (78) with (127) of Chapter VI, and conclude t h a t p -> Lp gives t h e adjoint action of L on P 4 . I t is customary t o refer t o P 4 as momentum space. The group H acts on P 4 and converts P 4 into a 17-space. P 4 then breaks u p into //-orbits and it is necessary to compute these. I t is simplest to deal with t h e //-orbits. W e introduce t h e following subsets of P 4 : Xm+ = {P ' P2~P2-p2-P2
= ™2,Po > 0},
= {P • P2-P2-P2~P2
Xm~
= ™2,Po < 0},*
where m is real and > 0. We write *oo = {0}. Also, for a n y m > 0, we write Y
m = {P '-P2-P2-P2-P2
= ~^2}.
Xm + U Xm~ and Ym are obviously invariant under the whole homogeneous Lorentz group / / , in particular, under H. We claim t h a t Xm + and Xm~ ( m > 0 ) are invariant under H. If m > 0 , the consideration of the mapping {p1,p2>Pz) -> ± {pi2 + P22 + Pz2)112 shows t h a t Xm+ and Xm~ are the connected components of Xm+ U Xm~. Since H is connected, it must leave each of them invariant. For m > 0 , Xm+ and Xm~ are homeomorphic to R 3 ; for m = 0, t o R 3 -{(0,0,0)}. Thus X w ± ( m > 0 ) , Ym ( m > 0 ) , a n d X 0 0 are all if-invariant, and P 4 is t h e union of these.
We shall next prove that H acts transitively on each of these. For real x, y let IV X 0 (80)
Lx,v =
y
0
0
1
0 0
°\
1
If y>0 and i/2 — x2 = l, it is clear t h a t LXtV belongs to H. Consider now Xm+ with m > 0 and a point #> on it. The point (m,0,0,0) lies on Xm +. Choose now x = (p02 — m2)ll2jm, y=pQjm in (80); since ^ 0 > ^ a n d 2 2 2 2 2 p0 — m —p1 -j-p2 +p3 , x and y are.real and y>0. Then L = Lxy sends (m,0,0,0) to a point of the form (Po,Pi,0,0). Now P i , 2 = ^ i 2 + i > 2 2 + ^ 3 2 so t h a t we can find a rotation R e K sending (Po,Pi,0,0) to (Po,p1,p2^P3)' This shows t h a t EL sends (m,0,0,0) to ^). H is thus transitive on Xm +. For X m ~ we consider ( —m,0,0,0) and use a similar argument. Consider next X0+ and a point peX0 + . (1,1,0,0) e X0+ and, if we choose x = (p02-l)l2p0 and y = (p02 + l)l2po in (80), L = LXtV sends (1,1,0,0) into (Po>Pofifi)' Once again, there is a rotation ReK such t h a t RL sends (1,1,0,0) to p. A similar argument, using ( — 1,-1,0,0) instead of (1,1,0,0)
GEOMETRY
340
OF QUANTUM
THEORY
establishes t h e transitivity of H on X0~. The transitivity of H on X00 is trivial. We now take u p t h e Ym ( m > 0 ) . Let p = (p0,Pi,p2>P3) £ Ym; p0 need not be > 0 . If we choose x=pQjm a n d y = (p02 + m2)ll2lm, then L = LXtV GH and sends t h e point (0,ra,0,0) of Ym into (p0,my,0,0); as m2y2 = Pi2+P22+P32> there is a rotation R e K such t h a t RL sends (0,ra,0,0) to p. This proves t h a t Ym is transitively acted on by H. The stability subgroup of H a t (ra,0,0,0) or ( —ra,0,0,0) (ra>0) is t h e rotation subgroup K of H. The stability subgroup a t (0,ra,0,0) is t h e subgroup of H consisting of all matrices L of t h e form
L =
0
1 0
0
^20
0
<X22
^23
\a30
0
#32
#33/
where > 0 and t h e 3 x 3 matrix (a i; ) i>j= 0,2,3 form Zf
.
XQ
nas
determinant 1 and preserves t h e ^2
^3 •
I n our usual notation (cf. Chapter V), we denote these by Hi±m000) and f/(0>m o>0). The first two are compact and connected. The last is not compact. All are semisimple. For the action of the covering group H* on P 4 , t h e orbits are t h e same. For t h e stability groups, we h a v e : (81)
#<*±m,0,0,0> =
K*.
±
For X0 t h e nature of t h e stability subgroups changes radically. I t is most convenient t o construct these corresponding to t h e two points ( ± 1,0,0, ± 1 ) . We consider X0+ a n d t h e pre-image in H* of t h e stability group a t (1,0,0,1): #8.0.0.1) = 8- 1 (#a.o.o.i))-
(82)
Using (46) we find t h a t m e #* t o,o,i) if a n d only if
-u--u-
//<*,o,o,i> is thus t h e group E* of all matrices m of the form (84)
mz,a=
(*
"_\
(a,zeC,\z\
= 1).
RELATIVISTIC
FREE
PARTICLES
341
m
{ i,a : « e C} is a normal subgroup isomorphic to the plane, {mz 0 : |z| = 1} is isomorphic to T, and E* is the semidirect product of these two where the action of z on a is given by a—>z2a. Therefore E*/{± 1} (1 in the denominator is the 2 x 2 unit matrix) is isomorphic canonically to the Euclidean group in the plane. In other words H(1001) is isomorphic to the group of Euclidean motions of the plane R2. The groups //* >0,O,D a n d #(i,o,o,i) a r e n ° t semisimple. They are, in fact, solvable. The calculations are similar for the case of X0~; the stability subgroups are the same. We note that our argument showing that H is transitive on Xm + established something more. It actually showed that given any p e Xm + there is a rotation ReK and a Lorentz transformation Lxy (x,y>0, cf. (80)), such that RLxy sends (m,0,0,0) to p. As the stability subgroup at (m,0,0,0) is K, we conclude from this that any L e H can be written as (85)
L = RZS,
where R and S are rotations, and Z — Lxy y2 — x2 = 1. If we define (86)
for suitable x>0, y>0 with
v = x/y,
then LXtV becomes the transformation connecting two inertial frames, the second of which is moving uniformly with respect to the first with velocity v (cf. (18)). This shows that the general Lorentz transformation is a product of the rotations and transformations to moving frames of the form (18). We have already used this fact in Section 1. The Invariant Measures on the Orbits. All the H-orbits admit invariant Borel measures. This is obvious for X m ± and the Ym (m>0) because both the group H and the stability groups are semisimple and hence unimodular. For X0 ± we note that the Euclidean group is unimodular so that each of X0± also admits an invariant measure. We shall now calculate these measures explicitly. Choose a real number a>0. For any £>a let D% be the //-invariant open set D^{p : p0>0, a
[ f(P)dP
(/eO c (P*)).
JDZ L
Since det L= 1 for LeH, F(f ,£) = F(f,£) for all £ > a and L e H. Moreover, i f / > 0 , F(f^) increases with £. Hence, on writing
J,(f) = ^ F(f,i)
(I > a),
we see that / -> J^(f) is an //-invariant nonnegative linear functional on CC(P4). It therefore determines an //-invariant Borel measure on P 4 . On the other hand, the transformation S
'• (P0>Pl>P2>P3) ">
(P02~Pl2-P22-P32>Pl,P2,P3)
342
GEOMETRY
OF QUANTUM
THEORY
is one-one on D% so that we also have f^f/((M+IP12)1,2,P),W
F(fn
F(M) 2
2
2
la.
2(U+IPIT2
dp
r
2
(|p| =^ 1 +^ 2 +^3 ). From this we obtain Jj(/)
- J P . 2(f+| P |T 2
p
'
This formula shows that the measure defined by Jm2 is actually supported by the set Xm + and is thus the invariant measure on the orbit Xm + . We write am+ for this measure. Thus (87)
k
^
"Jp. 2(»«+|p|T»
#
-
Note that as m -> 0 + , the integral on the right remains finite, since the function p -> |p| _ 1 is integrable around the origin in the space P 3 . Hence the same formula defines the invariant measure on X0 + also. Note that X0+ is not closed in P 4 and yet the invariant measure on X0+ is actually a Borel measure on P 4 . A similar computation leads to the following formula for the invariant measure am ~ on Xm ~ : (88)
f
fifa
" - f /(-(^ 2 +lPl 2 ) 1 / 2 .P) r f p
We may follow an analogous method for determining the invariant measures j8m on the orbits Ym. We choose a and £ such that a < f < 0 and define D{ as the open set {p : a
We may now proceed as in the earlier case and obtain the formula
[ m -[ /((-m^lpl^ ; p)+/(-(-m 2 +lp| 2 ) 1 / 2 ? P), a (89) J y / ^ " J | p | > m * 2 (-m-+|p|T (M)
forall/GQP4). The foregoing observations are collected together in the following theorem. Theorem 9.2. The group H acts transitively on Xm+, Xm~ (ra^O), X 00 , and Ym (m > 0) which, as m varies, exhaust the orbits of H in momentum space. The stability subgroup of H* at (+ m,0,0,0) is the unitary subgroup
RELATIVIST1C
FREE
PARTICLES
343
K*, while the stability subgroup at (± 1,0,0, ±1) is the subgroup E* (85) of H*. Each orbit admits invariant measures. If we use pl9 p2, p3 as global coordinates on X m ± and Ym, the invariant measures a m ± and ftm are given by dam±
r
„
dp,dp2dp3 2(m2+p12+p22+p32)112 2( — m2-\-p12+p2
v
-
''
+P3 )
3. THE REPRESENTATIONS OF THE INHOMOGENEOUS LORENTZ GROUP The inhomogeneous Lorentz group G as well as its covering group 0* are semidirect products. Moreover, from theorem 7.40 we conclude that all the multipliers of 0* are exact and also that every irreducible projective representation of G is induced canonically from a unique irreducible representation of G*. The group G* is a semidirect product and to it we apply the theory of Chapter VI, especially theorem 6.24. The orbits of H* in P 4 have been calculated in Section 2 and it is obvious that the orbit structure is smooth. Furthermore, the stability subgroups of H* associated with these orbits are described in (81) and (84). Theorem 6.24 therefore enables us to describe the irreducible unitary representations of G* completely. In what follows we shall carry out such a description, arranged according to the various orbits. The Orbit X 00 . Since XQ0 — {p :p = 0}, the stability subgroup is the whole group H*. The irreducible representations of G* associated with X00 are obtained by selecting an irreducible unitary representation n of H* and defining the representation U00,71 ofG* by: (90)
U00>* : h, x -> 7r(h)
(heH*).
The Orbits X m ± ( m > 0 ) . We select ( + m,0,0,0) as representative points of Xm ±. The stability subgroup of H* is the unitary subgroup K* whose irreducible unitary representations are the representations D} 0' = 0,4,1,...) described in Chapter VI. Let us choose a strict (ZT*,XTO±)cocycle
jf».±.> =
^ftMfV^).
For #eR 4 and heH* we define the operator D ^ - ' of Jf7™-^ by (92)
(UZ;±-'f)(p) = exp ^ , i ) V ( / . 5 S ( ^ ) - ^ ) / ( 8 ( ^ ) - 1 p ) ;
344
GEOMETRY
OF QUANTUM
THEORY
here {.,.} is defined as in (76) and 8(h) is the Lorentz matrix corresponding to heH*. In view of our discussion in Chapter VI we know that Un'±->:K,x^U%;±->
(93)
is the irreducible representation of G* corresponding to the orbit X m ± and the representation D3 of K*. The Orbits XQ±. The stability group at the points (±1,0,0, ±1) is the set of all matrices (z, a e C, \z\ = 1) m •-a~ (o
-1)'
2
We have written E* for this group. Our first objective is to determine the irreducible unitary representations of E*. Now, E* is itself a semidirect product so that this can be done with the help of the theory of Chapter VI. We write A ={m1>a}, T' ={mZi0}. Clearly v
l,azmz,0
(94) ™z,0™>l,a™
™l,zV
Thus, with respect to the action z, a -» z2a of T' in A, E* becomes identified with the semidirect product of A and T'. For each complex number b, Vb ' w>i.a - * e x P * Re(6a*) is a character of A, and b —> yb is an isomorphism of the additive complex number group onto A. The adjoint action (cf. (127) of Chapter VI) of T' on A is given by
and the T -orbits in A are O 0 = {2/o}
and
c„ =
{yb : \b\ = P}
(P > 0).
The stability subgroup at y0 is T' itself; at any point of Cp it is {m:tlt0}. The representations of E* associated with y0 are simply those of T' lifted to E*. Thus, for each integer n, (95)
*n:™2ta->3n
(" = 0, ± 1 , ± 2 , - . . )
is a unitary irreducible representation of E*. On the other hand, if p>0, we note that the stability group associated with Gp has two one-dimensional irreducible representations—the trivial one, and the one which sends m ± i,o t ° ± 1 • Corresponding to these, there are two inequivalent irreducible
EELATIVISTIC
FREE
PARTICLES
345
representations 7TP and -np of E*. Since TTP is associated with the trivial representation of the stability group corresponding to Cp, its description is immediate. For describing TTP we must first determine a strict (T',CP)cocycle with values in T which defines the homomorphism m ± l f 0 —> ± 1 a t a point of Cp. I t is obvious t h a t the cocycle (96)
mz,o,y->z
(yeCp)
has the required properties. We thus obtain the following description of 7TP and TTP . Both of them act in (97)
Jf =
&2(T,dt),
and (98) (99)
K ( m M ) / ) ( f ) = exV[iP
Re^a*)]/^"2*),
« ( m * . « ) / ) M = exp[» P Re(fc" V*)]z/(z- 2 *);
we have simply taken into account the formulas (94) and applied (132) of Chapter VI. Having determined the irreducible representations of E* we can now obtain the representations of G* which are associated with the orbits X0 ± . We select for n = 0, + 1 , ± 2 , - • • a strict (H*,X0±) cocycle with values in T, say
at ( ± 1,0,0, ± 1 ) . Then, in t h e Hilbert space .*"*•» = J ^ 2 ( X 0 ± , a 0 ± ) ,
(100)
we have the representation U±,n of G* given by: (101)
(U£$f){P)
= exp
i{x,p}
Furthermore, choose for each p, strict ( J fl r *,X 0 ± )-cocycles cpp a n d cpp'f which take values in the unitary group o£jf = J&2(T,dt)i and which define, at ( ± 1,0,0, ± 1 ) , the representations irp and TTP of E*. The corresponding representations U±tP, U±,p,r of G* act in (102)
X±"
= J^2(Z0±,Jf;a0±)
and (103)
(104)
m.xpJ)(p) p/
= exp * { » , i i V ( A , 8 ( A ) - W ( 8 ( * ) - ^ ) . l
m.x f)(v) = ^p ^y-'(*,8(i)-W(«(*)" rt'
The Orbits Ym ( m > 0). The stability subgroup a t (0,m,0,0) is a semisimple Lie group b u t is not compact. I t s unitary representations other t h a n the trivial representation are all infinite dimensional. The general theory
346
GEOMETRY
OF QUANTUM
THEORY
associates with any one of these, say 77, a representation Uim,n of G*. We shall not bother to write these down since, as we shall see presently, they do not seem to be useful in describing physical particles. We shall now take u p the question of classifying the representations obtained into physical equivalence classes. I t is clear t h a t two representations of G* are physically equivalent if and only if-there exists a unitary or anti-unitary isomorphism which intertwines them. Now we have seen in the course of the proof of lemma 8.3 t h a t , given an arbitrary representation 77 of a group, there is a canonical anti-unitary isomorphism of the Hilbert space of the representation 77 onto its dual which intertwines 77 with its contragredient. I t follows, therefore, t h a t two representations of G* are physically equivalent if and only if they are either equivalent or mutually contragredient. The case of contragredience is settled by the following lemma. A
Lemma 9.3. Let F x ' B be a semidirect product and letO^B be an F-orbit under the adjoint action. Let Fy be the stability subgroup of F at yeO. Then Fyo-i = FyQ, and if O - 1 is the orbit of yf1, then for any representation TT of FVo, the representations of F x' B associated with (0,TT) and (O - 1 ,^ 0 ) are mutually contragredient, TT° being the representation contragredient to TT. Proof. Let TT act in a Hilbert space Jf! We select a finite measure a on 0, quasi invariant under the F action, and define, for he F, rh = dajda{h~1). Then, the representation U of F x' B which is associated with (0,TT) can be written as {UhJ)(y)={rh(h-i-y)yi*y(b)
(yeO),
where
W)(y)=sf(y-1)
(j/eO"1)-
J is clearly an anti-unitary isomorphism of 3/P onto e ^ ? 0 = o ^ 72 (0~ 1 ,J^jS). We compute J o U oJ~1=V. A brief calculation shows t h a t , f o r / 0 e Jf?°. we have
(Vh,bf°)(y) =
{rh(K-i.y-i)Y^(b)S9^
I t is obvious t h a t z - > r^z'1) is a version of dfijdfth~r) and t h a t h, z-> 1 S^hyZ'^S' is a strict (i^,0 _ 1 )-cocycle defining the representation 1 S OTTOS' a t 2/o_1- Hence V is the representation associated with O'1 and S o TT o ^ - 1 . Since S © 77 O A 9 _ 1 ~ 7 7 ° and J o U oJ-^-^U0 (the contragredient of U), the lemma is proved.
RELATIVISTIC
FREE
PARTICLES
347
Applying this lemma we see t h a t t h e physical equivalence classes corresponding t o positive m are {Um, + 'j,Um'~J}. For the representations associated with X0± it is clear in the first place t h a t U + ,n and U~'~n are physically equivalent. On the other hand, lemma 9.3 applied t o E* shows t h a t the representations up and 77-/ are both self-contragredient, so t h a t we conclude t h a t U + tp and U~>p as well as U + ,p,/ and U~'p'f are physically equivalent. W e can now formulate t h e main result of this chapter. I t was obtained b y Wigner in his 1939 paper. Theorem 9.4 (Wigner [1]). Any irreducible representation of G* induces a projective representation of G which is irreducible and, conversely, given an irreducible projective representation of G there is exactly one irreducible representation of G* which induces it. An arbitrary irreducible representation of G* is equivalent to one associated with some H*-orbit in P 4 and some irreducible representation of the stability subgroup of H* at some point of this orbit. The irreducible representations Um, + ,j (j = 0, J , • • • ) , jj + ,±n ( ^ = Q? l ? 2,• • •), U + 'p and U + tP'f (p>0) are physically inequivalent and an irreducible representation of G* which is associated with one of the orbits X m ± ( m > 0 ) is physically equivalent to exactly one of these. Not all irreducible representations of 6r* can be used t o represent free particles. Consider first of all the representations U00,n associated with the trivial orbit X00. I n every one of these, the space-time translations are represented b y the identity operator. This implies t h a t the energy and momentum observables vanish identically. Consequently, all these representations have t o be given up, with t h e exception of the trivial one-dimensional representation of G*, in which every Lorentz transformation goes over to the identity. I t represents the vacuum. We shall next examine an irreducible representation U of G* associated with one of the nontrivial orbits X m ± and Ym. Let Q0 be the energy and Qi, Q25 Q3 the linear momenta along the three axes in a given coordinate system. Since t h e space-time translation b y # e R 4 goes over into the operator of multiplication by the function p -> exp i{x,p}, it follows t h a t the operators Qj are multiplication operators. A direct calculation shows t h a t in the Hilbert space of U, Qo2-Qi2-Q22-Q32
(105) 2
= A/, ±
where A= + m if U is associated with X m a n d — m2 if it is associated with Ym. If we now consider a function (scalar or vector valued) which belongs t o t h e Hilbert space of U a n d vanishes outside a very small compact set, it follows t h a t in the corresponding dynamical state, t h e momenta Ql9 Q2, Q3 a n d energy Q0 have distributions with extremely small variance and hence we may approximately regard the particle as having sharply defined energy and momentum values. Now for a classical
348
GEOMETRY
OF QUANTUM 2
THEORY 2
free relativistie particle, the quantity Q0 — Qi — Q22 — Q32 represents the square of its rest mass. Consequently, the constant A must also be interpreted as the square of the rest mass. This rules out the alternative A = — m2 as otherwise it would seem t h a t we have to admit particles with imaginary mass. This leaves us with the representations associated with X m ± ( m > 0 ) a n d the trivial representation. The representations associated with X m ± represent free particles of rest mass m. The form of the representations suggests t h a t at least for m>0 we interpret Um, + fj as representing a particle with spin j . We shall see much later (Section 7) t h a t this is indeed the case. I n the meantime we shall define the parameter j for the Um' + 'J' and the parameter \n for the U + ,±n to be the spin of the particle. The representation Um' + '112 describes the electron; the representations Um' + t0 represent the mesons. The representations U + '±n, U + 'p and U + 'Pt' represent the massless particles. We shall see later t h a t U + ,±1 represent the neutrinos and U + » ± 2 the photons. The representations U + ,p and U + 'p>' correspond to infinite dimensional representations of the stability group E*, and hence the states of the corresponding particles are represented by infinite dimensional vector valued functions on XQ ±. The physical interpretation of these particles is somewhat unclear, as they would have to describe particles whose spin observables have infinite spectra (cf. Section 5).
4. C L I F F O R D A L G E B R A S Clifford Algebras. The customary representation of the states of a relativistie electron is in terms of spinor fields in space-time. Such descriptions depend heavily on the properties of the Clifford algebra and its representations. We devote the present section to this topic. Let V be an n-dimensional vector space over the complex field C. We shall assume n>4. Let B be a nonsingular symmetric bilinear form on V x V. I t is known t h a t there exist bases {v1, v2, • • •, vn} for V such t h a t (106)
r
S r* g,
B(vr,vs) = ^
corresponding to such a basis we shall construct an associative algebra with unit element 1, which we shall call the concrete Clifford algebra F. We do this as follows. For each subset A^N = {1, 2,- • •, n} we choose an abstract element eA with e0 = l, and define F to be the vector space over C with the eA as a basis; dim(i^) = 2 n . We shall first define a product operation between elements of the set eA. To do this formally, we define, for A,
B^N,
AAB = AKJ B-AnB;
for A^N,
jeN,
let p(A,j)
be the
RELATIVISTIC
FREE
349
PARTICLES
number of elements ie A with i >j; and for A, BczN, p(A,B) = 2 ; e B p(A, Let us then define (107)
j).
(-irA-*eAAB.
eAeB =
I t is easy to prove t h a t this operation is associative and t h a t leA = eAl=eA for all A. To prove the associativity, let A±, A2, A3^N. I t is easily seen t h a t (A1 AA2) AA3 = A ± A (A2 AA3). On the other hand, p(A1,A2AA3)
=
jeA2A
p(AxAA2iA3)
P(Ai>3) = p{Al9A2)+p(Al9A3)
2
= 2
(mod 2),
A3
P(AxAA2,j)
= p{Al9A3)+p(A29A3)
(mod 2).
JeA3
From these relations we obtain the relation e
A1(eA2eA3)
—
(eA1eA2)eA3
a t once. Since the eA span F, it follows t h a t F becomes an associative algebra under the product operation defined by (108)
2 p A e A 2 qAeA = I
(-\rA-B)pAqBeAhB.
The mapping which sends vr to e{r} extends to a linear isomorphism of V into F. Let us write T for this linear m a p and write er = r(vr). From (107) we then obtain (109)
e r 2 = l,eres + eser = 2S rs ,
and for A = {i1, i2i • • •, ir} with 1 < ix < i2 < • • • < ir < n, (HO)
eA =
ehei2-.-eir.
Thus F is generated by the er and the generators satisfy the relations (109). From (109) we obtain the following equation: (111)
T(V)2 = B(v,v)l
(veV).
The construction of F which we have given here is not an invariant one since we have used the basis {vl9- • •, vn} explicitly. We shall now indicate a more abstract construction. Let % be the tensor algebra over V and let $ be the two-sided ideal in % generated by the set of all elements of the form (112)
av = v
®v-B(v,v)l.
We observe t h a t Qf is proper. I n fact, consider the concrete Clifford algebra F associated with a basis {^i,•••,#„} satisfying (106). By the basic property of the tensor algebra %, there exists a homomorphism, f say, of % into F, such t h a t f = r o n V. Since T[V] generates F, f maps % onto F. Since T(V)2 = B(V,V)1 in F, f(av) = Q so t h a t $ ^ k e r n e l (f). This proves t h a t 3 is proper and, indeed, t h a t dim %/!$ > 2 n .
350
GEOMETRY
OF QUANTUM
THEORY
We shall now demonstrate t h a t f has precisely $ as its kernel. This will be done if we show t h a t dim Xj^<2n. Consider t h e basis {vl9- • •, vn}. E v e r y element in % is a linear combination of elements of t h e form v vt modulo $ , a n d u ®' '' ® vi • Now, vt (g) vt = 1 a n d v% hence it is obvious t h a t vi± ® • • • (g) vir = ± w modulo $ , where w is of t h e form vjx ® • • • ® vh with 1 <jx <j2< • • • <js
(113)
commutes; for there exists a homomorphism f' of % into c€' such t h a t f' = r on V, a n d as r ' [ F ] generates c€', ¥ is surjective. B y (ii) above, av e kernel (f') so t h a t $ c kernel (f') a n d hence f' induces a unique homomorphism r~ of ^ onto # " such t h a t (113) commutes. Furthermore, this universal property determines t h e pair (^,7r) u p t o isomorphism. I n fact, let ^f' be a n associative algebra with unit 1' and v' a linear mapping of V into c€' such t h a t (i) TT'\V\ generates (€' a n d 7T'(V)2 = B(V,V)V
for all
v E V, (ii) &',TT') has the universal property mentioned above. Then there are homomorphisms 6 a n d 6', such t h a t 6 maps ^ onto ^ ' , a n d 6' maps ^f' onto ^ , a n d t h e diagram
(114)
commutes. Diagram (114) shows t h a t 6 o 6' a n d 0' o 6 are both identities and hence t h a t 6 and 0' are isomorphisms.
RELATIVISTIC
FREE
PARTICLES
351
The invariant construction of ^ enables us to associate automorphisms of ^ with 2?-orthogonal transformations of V. More precisely, let L be an invertible endomorphism of V such t h a t B(Lv,Lv)
= B(v,v)
for all v. Clearly, there exists an algebra automorphism L~ of % which extends L. I t is obvious from (112) t h a t aLv = L~(av) for all v e V from which we easily conclude t h a t L~ ($) = !$. This proves t h a t L~ induces an automorphism of ^ . We write this automorphism as x - > xL (x e *€). Since L -> L~ is a homomorphism, we conclude t h a t the mapping, which assigns to L the automorphism x -> xL of ^ , is a homomorphism of t h e orthogonal group of B into the group of automorphisms of %\ Note t h a t (115)
TT{LV) =
TT(V)L.
Simplicity. The central fact in the theory of Clifford algebras is t h a t when dim V — n is even, <% is simple. We shall now proceed to prove this result. Suppose t h a t 3 is an arbitrary nonzero two-sided ideal in <€. Any nonzero element u e 3 can be written uniquely as
(116)
u = p0l+
Piu-",iMvh)' ' ' W K)-
2
We shall choose u to be a nonzero element of 3 with the smallest number of nonzero P\lt... ,tr in its expansion (116). Now, a direct calculation shows t h a t when r is odd, Trfa) • • -7r(vir) and n(Vj) commute ifj is one of il9 • • •, ir and anticommute if j is none of ilt- • •, ^r; whereas, when r is even, the situation is exactly the reverse. Consequently, when n itself is even and there is at least one ptl . . . > i r ^ 0 in (116) above, one can find s, (klt' • m, ks), and some j , such t h a t pklt... ,ks^® a n d ^ ( ^ J * ' #7r(vfcs) anticommutes with TT(^.). AS T T ^ M ^ J - • • 7r(vir)]7r(vj) = ± T T ^ J • • • 7r(v
irivJuniVj) =p0l+
Ptlt-.iMvii)"'7r(vir)
2 (i 1 ,--.f r )^(fci,---,fc s )
wherepj >4 = + p t h a t w' e 3,
i f i r
. W r i t i n g ^ ' = |(w -f 7r(vy)wTr(Vy)), we get, onnoting
u' = p0l +
2
Pt1,...Arfvh)' ' "*(%)>
(il,--',ir)^(fcl,-",/C s )
where 2>J'1,...,ir = i(.Pi 1 ,... > i r +Pi lf ...,*,)• Since the number of nonzero Pilt...tir is strictly less t h a n the number of nonzero Pilt...,ir, we have a contradiction to the minimal nature of u. This shows t h a t a l l ^ i l t . . . i r must vanish in (116). But then p0^0 and hence l e ^ , showing t h a t 3} = *%. We have therefore proved t h a t ^ is simple when n is even.
352
GEOMETRY
OF QUANTUM
THEORY
The Spin Representation. By the classical Wedderburn structure theorem (cf., for example, Weyl [2], pp. 79-96, 280-290), there exists, up to equivalence, exactly one finite dimensional irreducible representation of the simple associative algebra ^ when n is even. Let us select a representation of this unique equivalence class and denote it by p. It is known that in the left regular representation of ^f, p occurs exactly as many times (as a direct summand) as its dimension. Hence (118)
dimP = 2nl2
(n even).
Let us write S for the vector space in which p acts. Let GB be the orthogonal group of V and, for any L e GB, let x -> xL be the automorphism of ^ induced by L. Then x -> p(xL) is also an irreducible representation of ^ in 2 . Consequently, there exists an invertible linear transformation Sx of S such that L 1 (119) P(x ) = s^wsr for all x e ^ ; note that this is guaranteed as soon as we have
S^WWr1
p(ir(Lv)) =
for all v e V. We may even choose $ x such that det S1 = l. Now, by Sehur's lemma, 8X is determined by (119) up to a constant multiplying factor. Hence, if L, L' e GB and we choose 81,81f, and 8Q such that (119) is satisfied for L, L', and LU with Sl9 $ / , and 80, respectively, we obtain the relation (120)
S 0 = a/SW
(creC);
the determinant condition on 80, Sx, and $ / implies that \a\ = 1. Let G^ be the lcsc group of all invertible linear transformations of S of determinant 1. The set of all (L,SX) e GQXG^, such that p(ir{Lv))8l
=
SlP(n(v))
for all v e V, is obviously a closed subset of GBxG^, and the map (L,8±) -> L is continuous and maps this set onto GB. Hence, by the Federer-Morse lemma (cf. Chapter V), there exists a Borel map L-^S^L) of GB into Gz such that (119) is satisfied by SX{L) for all x e *€. But then (120) implies that L-^SX(L) is a projective (nonunitary) representation of GB in S. We have S±(LL') = m(L,L,)S1(i>)S1(iv,) for all L, U e GB. The associative law L(L'L") = (LL')L" shows at once that m is a multiplier for GB. Since GB is semisimple, we may use the theory of multipliers (Chapter VII) to lift 8X to a representation 8' of the covering group GB* of GB. As GB* does not have any one-dimensional representa-
RELATIVISTIC
FREE
PARTICLES
353
tions other t h a n t h e trivial one, S' is uniquely determined b y S±. S'(/-> S'(/)) is t h u s t h e unique representation of GB* in 2 such t h a t (121)
p(n(I/v)) =
S'WpiirmSV)-1
for all v e V, £ -> U being t h e covering homomorphism of GB* onto GB. The elements of 2 will be called spinors a n d t h e representation $ ' ( / - > $ ' ( / ) ) a n d its irreducible components are called t h e spin representations. We treat the elements of 2 as determining a new type of quantity, say q, such t h a t when V undergoes t h e orthogonal transformation I/, q undergoes t h e transformation S'(f). I t was E . Cartan who discovered, in infinitesimal form, t h e spin representations; t h e global version was constructed b y H . Weyl. Cartan also proved t h a t t h e spin representations of GB* cannot be obtained b y decomposing t h e tensor algebra over V for n > 4 with respect t o the natural action of GB*. T h e reader who is interested in a deeper study of these matters should consult Cartan's book [1], Chevalley's account [4], a n d t h e article of Brauer a n d Weyl [1]. We shall now extend these results t o t h e case of real vector spaces a n d quadratic forms. L e t V° be a real vector space of dimension n, n being even. L e t B° be a nonsingular symmetric bilinear form on V° x V°. W e write V for t h e complexification of V° a n d B for t h e bilinear form on V xV which extends B°. As n > 4, the group of invertible endomorphisms of V° of determinant 1 which leave B° invariant, is semisimple. We write GBo for t h e connected component of the identity of this group. Identifying each endomorphism of V° with its natural extension t o V (as a complex endomorphism), we have GBo<^GB. L e t GBo* be t h e universal covering group of GBo. Then, denoting b y 2 t h e space of all spinors, we have a projective representation
of GBo in 2 such t h a t p(7r(Lv))=S1(L)p(7T(v))S1(L)~1 for all L e GBo a n d v E V°. Since GBo is semisimple we m a y argue as before t o deduce t h e existence of a unique representation
of GBo* in 2 such t h a t (122)
PMI/V))
=
SQ(S)p(n(v))S°(Syi
for all v e V° a n d i e GBo*, £-> U being t h e covering map. I t is usual t o call S° a n d its components also as spin representations. I t is usual t o exhibit S° in a particular way involving t h e so-called Dirac y matrices. We shall select a basis {w0, • • •, w n _i} for V° such t h a t (123)
B°(wr,Ws) = E^
354
GEOMETRY OF QUANTUM
THEORY
where (124)
+1 -r
- i
(0 < r <
p-l),
,
l-l (p < r < n-l). If F is the matrix diag{e0, e±, • • •, e n _i}, then 6rBo can be identified with the connected component of the identity of the group of all matrices L such that DFL = F. Since F~x — F, we can argue exactly as we did in the Lorentz case to conclude that Ll e GBo, whenever L e GBo. Since GBo* is the universal covering group of GBo, there exists an anti-automorphism (125)
(->{*
of GBo* such that it induces the anti-automorphism L-> Ll of GBo (cf. (55)). If we write (126)
£(/) = ^ V * ) - 1
(SeGBo *\
then it follows from (122) that £ ^> S(£) is the unique representation of 6rBo* in 2 such that (127)
8ifi)-ip(ir(v))S(t)
=
P(rr((I/yv))
for all v e V° and / e GBo*. Let us now write (128)
yr = p(7r(wr)).
Then, as p(n(v)2) = B(v,v)l for all v e V°, we have (129)
yr2 = erl
lyrys+ysyr = o
(0 < r <
n-l),
(r ^ s).
Conversely, given endomorphisms y0,- • •, y n _ 1 of some vector space 2 ' satisfying (129), there exists obviously a unique representation p of the Clifford algebra ^ in 2 ' such that //(7r(wr)) = yr; in fact, we need only to define p by this equation on V and extend it to ^ using the universal property (113). p will then be equivalent to a multiple of p. Hence dim 2 ' will be a multiple of 2 n/2 . If dim 2 ' = 2 n/2 , then p will itself be irreducible, leading to the irreducibility of the set {y0,« • •, y n _i}. In this case, the equation (127) becomes
(130)
#(0-VS(0 = n 2
where ars(L) is the r-sth. matrix entry of L with respect to the basis K'-'^n-l}' We summarize our discussion so far in this section in the form of the following theorem. Theorem 9.5. Let V be a complex n-space, n>4:, and even Let B be a nonsingular symmetric bilinear form on V xV. Then, the Clifford algebra
RELATIVISTIC
FREE
PARTICLES
355
over V associated with B is simple, and has, up to (linear) equivalence, exactly one irreducible finite dimensional representation p; dimp = 2 n/2 . Suppose now that V° is a real linear subspace of V of real dimension n such that the restriction B°, of B to V° x V°, is real. Let {w0,- - - ,wn_1\ be a basis for V° such that the relations (123) are satisfied. Then there exists a set {y0, • • •, y n _i} of 2 n/2 x 2 n/2 complex matrices such that the relations (129) are satisfied. {y0, y l3 • • •, y n _i} is necessarily irreducible and any tivo sets of such ys are connected by a similarity transformation yy -> AyjA'1. Finally, if n>4, there exists a unique representation S of the universal covering group of the connected component of the identity of GBo such that the relations (130) are satisfied. The matrices yr are usually known as Dime's y matrices. Consider now the case when n = 4c and V° has a basis {w0,w1,w2,w3} such that B(w0,w0) = 1, B(wr,wr) = — 1 (r = l , 2 , 3). We are then dealing with a Minkowskian form. We shall now exhibit a particular choice of the y's. Let ax, o-2, <73 be the matrices ((55) of Chapter VIII) defined by
ai =
(i
o)'CT2= C ~o)'
CT3=
C -i)'
and let y0, y l5 y2, y 3 be the 4 x 4 matrices defined by
(13D
ro=(?
J), * = £ 7 )
(-1,2,3).
It is then a routine matter to check that the relations (129) are satisfied. Since the y's are 4 x 4 matrices and their irreducible realization is also four-dimensional, {y0>yi>y2>y3} must be irreducible. In this case, GBo is the connected homogeneous Lorentz group H, GB* is the group SL(2,C) = H*} and the covering map is what we have been denoting by m -» S(m). The relations (130) become: 3
(132)
Sim^YrSim)
= ^
(r = 0, 1, 2, 3).
s=0
y0 is a Hermitian matrix; yi, y2, and y3 are skew Hermitian. We may use (132) to analyze the representation S of H* in greater detail. Let I)* (cf. (52)) be the Lie algebra of H*. Then, we have, from (132), (133)
[S(X),Yr1 = - j [ «rS(3(X))ys
(r = 0, 1, 2, 3)
s=0
(for 8, see (54)). We shall now determine the matrices S(X) in terms of the y's. Since the y's are irreducible, we can write S(X) as a linear combination of the 16 elements 1, yr, yrys, yrysyi> yo7i72>/3- The commutation rules
GEOMETRY
356
OF QUANTUM
THEORY
(129) and (133) imply easily t h a t each S(X) must be a linear combination of 1 and the yrys (r<s). Since the matrices of any representation of a semisimple Lie algebra are of trace 0,f and since tr(yrys) = — tr(y s y r ) = 0 for r^s, S(X) must be a combination of only the yrys. Let
S(X) =
2
c
rS(X)yrys.
0
Substituting this into (133) we obtain, after a brief calculation, crs(X) = |« s r (S(Z)) so t h a t
(134)
S(X) = |
2
0
If we now use the formula (54), we find, after some calculation, (135)
[X * ( X , = (0
0 _
z +
\ ) ,
from which we obtain the formula (m (136)
*<»> = (<>
0
\
<»->•);
here the asterisk denotes complex adjunction. Formulas (135) and (136) display the structure of the representation $ explicitly. The notion of a spinor field on any manifold on which the inhomogeneous Lorentz group acts can now be introduced. Let X be the space on which the inhomogeneous group G = Hx'R* acts. Let 2 be the space of four complex dimensions on which the spin representation S of H* acts. A spinor field is then a function if; from X into 2 with the following transformation property: if we transform X into itself by the element (8(h),x) where x is the translation and 8(h) is the homogeneous component, then I/J transforms into the function if)', where (137)
0'(y) = S(hM8(h)tx)-^y)
(y e X).
5. R E P R E S E N T A T I O N S I N VECTOR B U N D L E S AND W A V E EQUATIONS The descriptions obtained in Section 3 for the representations of t h e inhomogeneous Lorentz group are not quite suitable for making explicit some of the physical features of free relativistic particles. Also explicit f Since any element of the Lie algebra is a sum of elements of the form [X, Y] (cf. Chevalley [3], p . 67).
RELATIVISTIG
FREE
PARTICLES
357
calculations become difficult since t h e formulas of Section 3 involve certain cocycles whose existence was only proved abstractly. Our first concern in this section is t o give more geometric descriptions of these representations which will be free of the disadvantages mentioned above. I n each case we shall construct a n appropriate vector bundle on the orbit in question, a n d exhibit t h e states of t h e particle as square integrable sections of this bundle. These results were first rigorously obtained b y Bargmann a n d Wigner [1]. W e shall follow their discussion without a n y essential modifications. We shall next prove t h a t these sections can be regarded as tempered distributions (in t h e sense of L. Schwartz) in momentum space P 4 . This enables us t o take Fourier transforms a n d obtain t h e Hilbert spaces as spaces of (tempered) distributions on R 4 . T h e differential equations satisfied b y these distributions will then be the wave equations. We shall see a t this stage t h e famous Dirac a n d Maxwell equations making their appearance. The distributions in space-time which describe the states of the particles are actually functions a t least for a dense linear manifold of states. These functions are then seen t o transform under the Lorentz group exactly like spinor fields, t h u s tying u p t h e theory with t h e usual principles of relativistic physics according t o which the states of the free particles are to be represented b y spin fields in space-time. The Representations U171'+ ,j Corresponding to Mass m , Spin j . We shall fix m > 0 a n d consider t h e orbit Xm+ with invariant measure am +. We introduce the complex 4-space C4 in which the Dirac matrices y 0 , y 1? y 2 , y 3 act as linear transformations, and treat t h e elements of C4 as spinors. The spin representation h-^S(h) (h e H*) is, of course, determined b y the equations (132). W e take for t h e y's t h e matrices defined b y t h e relation (131). W e write {e0,e1,e2,e3} for the standard basis vectors of C 4 ; e ; has t h e components 8j0, S ;1 , Sj2, Sj3. We write < . , . > for t h e usual Hermitian inner product for C 4 ; <er,es> = 8rs. We begin b y defining t h e bundle B+Al2 as (138)
B+-w = i(p9v) I
:p G Xm + , v e C4, J 3W = /c = o
4 J
with the projection (139)
7r:(p,v)->p.
An easy calculation based on (132) and (55) shows t h a t if (p,v) e B+'112, then so does ( 8 ( % > , £ ( & * " » . Thus (140)
h,(p,v) - > (p,v)h = ( S ( % v S ^ * ~ »
358
GEOMETRY
OF QUANTUM
THEORY
112
converts B + ' into a i/*-space. The action is smooth and projects to the usual i/*-action on Xm +. For any p e Xm+ the fiber a t p is the linear subspace of C4 given by
(141)
B£>™(p) = i(p,v) : 2/*r«v
mv
=
\
For the point (ra,0,0,0), (141) reduces to the equation y0v = v. From our formula (131) for y0 we see the fiber at (m,0,0,0) is the twodimensional subspace spanned by ex + e3 and e2 + e4- Hence (141) also defines a two-dimensional subspace. We shall construct next a family of covariant inner products on t h e fibers of the bundle. From the formulas (136) for 8(h) we obtain (142)
S(h)*y0S(h)
= y0
after a trivial calculation. Hence the representation 8 leaves invariant t h e Hermitian form v - > m~1(yQv,vy in C4. This form is actually positive definite on each fiber of the bundle B + '112. To see this, we consider a (p,v)eB+>112. We have: 3 fc=0
Now, (y0v,v} is real as y0 is Hermitian, and (ykv,v} is purely imaginary, as yk is skew Hermitian. Hence, as the right side of the above equation is real, we must have 2 ? = I Pk^Yk0^ = 0- We m a y therefore write Po
m-\y0v,v>
= pQ-\v,v>
(v e Bm
+
^(p)).
Equation (143) exhibits explicitly the positive definite character of the form v -+ m~\y0v,vy on the fibers of B^'112We now introduce the Hilbert space J f + ' 1 / 2 of Borel sections of the bundle B^'112- The value of ||
Now, using the formula in theorem 9.2 for a m + , we can rewrite this as
(144)
IMI2 = f
(
RELATIVISTIC
FREE
PARTICLES
359
where dp±dp2dp3 1 dp*®r = 2(p H ^ -22+p ^ 2+p^ 22 T — r + m22T ) 1 2 3
(145) v
The representation U of G* in ^ + -1/2 is given by (146)
(Uh,xcp)(p) = exp ^}cp(S(/*)->)>\
where
(IK:)This shows t h a t the representation of K* on the fiber at (m,0,0,0) is just D112. I n view of theorem 6.20, we m a y conclude t h a t the representation defined by (146) is equivalent to JJm' + '112. We now proceed to the case of higher spin. We consider a half integer j ( > \) and write 2j = N. For each v = 1, 2, • • •, N we define y / (r = 0, 1, 2, 3) to be the operator in the iV-fold tensor product C4 (g) • • • (g) C4 = C4«)iV which sends u± ® • • • (x) uN to u± ®- • • ® uN', where ut'= ut for i^v and uv' = yruv: in the obvious notation, (147)
y r v = 1 ® . . . ® y r ® 1 (g) • • • (g) 1 -X^m+
(1 < v < tf).
J
For 2? e we define B + (p) as the set of all (p,t), where t belongs to the subspace of C4<8)iV consisting of all symmetric tensors t such t h a t 3
(148)
^pryrvt
= mt
(v = 1,2,-
,N).
r=0
We define B + -* by (149)
B£->=
U
B + ->(p).
peXm +
It is easy to verify that B + '3(p) is the subspace of the symmetric tensors which belong to B + >1,2(p) (g)- • -(g) B+-ll2(p). For (p,t) e B + 'j and h e H*, we define (150)
(pjtf =
WhfaSsih*-1)^
where (151)
SN(h) =
S(h)(g).--(g)S(h).
360
GEOMETRY
OF QUANTUM
1
THEORY
fll2
S^h*" ) maps the space B + (p) (g)• • -(g) B + ,ll2(p) onto the space Bm'll2(&(h)p) ®- • *® B^AI2{h{h)p), and leaves the space of symmetric tensors invariant. Hence (150) converts B^,j into a i/*-space. Furthermore, SN leaves invariant the Hermitian form
where <.,.># denotes the natural Hermitian inner product in C40Ar; (143) shows that for all t e B + >ll2(p)
Ml 2 = f ,
(p)MP)>NdK-N(v),
JR3
where (153)
dfi+• "(p)
2(
^a
+
^
dp1dp2dp3 2 + ^ 2 + m2)(iV +1)/2
The representation U is defined by (154)
(UKxcp)(p) = exp ^ ^ ( S t A ) - ^ ) * .
C7~ j/m.+./ by theorem 6.20. The reader might have noticed that we have not treated the spinless case in terms of the spinor calculus. This can also be done provided we consider, instead of the bundle (149), the bundle where the fibers are made up of skew symmetric tensors. More precisely, let (155)
B+>° = {(p,t) : p e Xm + , t e C4 ® C4, * skew symmetric, (2pryrv)t = mt,v = 1,2},
with the norm defined for any Borel section
\\
The representation of G* is then defined in the same way as (154). For a given p e Xm +, the fiber of B + >° at p consists of all skew symmetric elements of B£,ll2(p) (g) B + 'll2(p) which form an one-dimensional space. As the unitary group K * has no nontrivial one-dimensional representations,
RELATIVISTIC
FREE
361
PARTICLES 0
it induces the trivial representation on the fiber of B + ' at (m,0,0,0). Hence the representation of 0* defined in the Hilbert space of sections of B + >° is equivalent to Um> + >°. The Representations U+'±n. These representations may be obtained by passing to the limit in Um'+ J'as m -> 0 + . We write (157)
B0 + = i(p,v) :peX0\veV,
j^ pryrv = oj.
B0+ is a bundle over X0
+
and the action h,(p,v) -> (p,v)h, where
(158)
{p,v)h = (8(A)#,/S(A*"»
converts it into a i/*-space. For any ra>0, the fiber of B + '112 at
w
= ime1
+
i(l
+
(l+m2)i/2)e3?
2 1/2
v2< > = i ( l + (l+m )
)e 2 + |me 4 ,
which converge, as m - ^ 0 + , to e3 and e2, respectively. Now e3 and e2 span the fiber of B0+ at (1,0,0,1). Since ZT* is transitive on Xm+ and X0 + , the same is true at any other point: if p e X0 +, there are points p(m) e Xm + which converge to p as m —>• 0 + , with the property that any vector v in the fiber of B0+ atp can be expressed as lim v(m), where v(m) belongs to the fiber of Bm ,112 at p(m \ However, all this could be done with — m instead of m, showing that the bundles Btm12 also collapse to B0 +. On the other hand, the endomorphism (159)
r = ^0717273 ll2
transforms B + ' (p) into B±fil2(p) for each p e Xm+ for any m > 0 , as is easily seen from the fact that T anticommutes with all the yr. In the limit, r leaves the fibers of B0 + invariant, leading to further degeneracies. We calculate quickly that
(160)
r = ( J _J).
Hence T commutes with all S(h) (cf. (136)). Thus, if {p,v) e B0 +, (p,Tv) e B0 +. On the fiber at (1,0,0,1) T has the eigenvalues ± 1. Hence the same is true on all the fibers. We may now define the bundles B£f±1'2 by (161)
£ 0 + ' ± 1 / 2 = {(p,v) : (p,v) e B0 + , Tv = +v).
H* leaves both B$'+1/2 and B£ - ~1/2 invariant. A simple calculation shows that for the stability group E* at (1,0,0,1) the representation induced on the fibers at (1,0,0,1) becomes (162)
m^a-^z*1.
362
GEOMETRY
OF QUANTUM
THEORY
We have already seen that the Hermitian form v-^Po~\v,v> is left invariant by the S(h) and is positive definite on the fibers of B + ,1/2 . Therefore, letting ra->0-f, we may conclude that the form is still invariant. At (1,0,0,1), the fiber of B0 + is spanned by e3 and e2 on which the form considered is even positive definite. Hence, it is positive definite on each fiber. We are thus in a position to apply theorem 6.20. J#*£'±l12 is the Hilbert space of Borel sections
(163) where
(164)
dft-Hr) =
dp^pzdpg 2(Pi2+P22+P32)
The representation U, which is equivalent to U + ,±1, is defined by (165)
(Uh,x
For higher n, we proceed as in the case of nonzero mass. We write for n>l,
(166) B^±nl2
= Up,t) :peX0\teW®\
(% prYAt
= 0,
vn = +t, v = 1,2,..-,^1, where
(167)
iy0vny2y3v.
r* =
The action of H* on these bundles is the obvious one. The norm of a section
(168)
|M| 2 = f (
with (169)
dj80+-»(p)
dpxdp2dp3 2(Pi +p22+P32Yn 2
+ 1)l2
We shall end this discussion with a description of the representation U + t2 0 U + '~2 which lends itself more readily to physical interpretation. We consider the bundle (170)
B = Up,v) :peX0\veC\
{p,v} = p0v0-
£ prvr = o | .
RELATIVISTIC
FREE
PARTICLES
363
H* acts on B as follows: (p,v)h =
(171)
(8(h)p,8(h)v).
4
For v, v' e C we introduce the Lor entz-Her mite form Q(v,vf) = — {v,v'*} = —v0v0'* + v1v1'* + v2V2* + v3v3'*. We m a y regard the fiber B(p) of B at p as the complexified tangent space to X0 at p so t h a t B is simply the complexified tangent bundle, and (171) gives the usual action on such bundles. Since 8(h) is real for each h, 8(h) preserves this Hermitian form also. We claim t h a t v —> — {v,v*} is nonnegative on each fiber of B. I t is enough to prove this at (1,0,0,1), where the fiber is spanned by w1 = (1,0,0,1), w2 = (0,1,0,0), and w3 = (0,0,1,0). The form, expressed in the basis just described, becomes a1w1+a2w2 + a3w3 -> \a2\2+ \a3\2. Let Jj?~ be the space of equivalence classes of Borel sections cp of B for which (172)
\\
-{
+
(p) < oo.
£F~ is complete, b u t || • || as defined in (172) is not a norm, but only a seminorm. I n order t h a t ||
Pm-Pm
= 0
(0 < j < k < 3)
+
be satisfied for a 0 -almost all p. We write J f for the Hilbert space obtained by going over to the quotient space modulo (173). The representation U~ is first defined in Jf?~ by setting (Uh.x
This proves t h a t
U~U+-2@U+>-2.
364
GEOMETRY
OF QUANTUM
+ ,p
THEORY
+ p
The Representations U and U ' ''. These cannot be obtained by any limiting process from the representations corresponding to particles of nonzero rest mass. Since the inducing representations are infinite dimensional, these are not of much physical interest as the usual interpretation forces the spin observables to have infinite spectra. We do not discuss these in this book. The reader who is interested may refer to the articles of Bargmann-Wigner [1] and Wigner [5]. The Infinitesimal Operators. One of the advantages of describing the representations of the group G* in Hilbert spaces of sections of various vector bundles is t h a t the infinitesimal operators, which are in general unbounded, are all applicable to the smooth sections of the bundle. We can thus study the infinitesimal representation in a very explicit fashion. If X is any element of the Lie algebra of G* and U is a representation of 0*, we write X~ for the self-adjoint operator such t h a t (175)
exV(itX~)
= Uexptx.
The calculations involved in computing the various operators X are very simple, and we confine ourselves to the determination of the angular momenta Mrs, 1 < r, s < 3 (cf. Section 2 for the definition of the elements Mrs). Mass m > 0, Spin j . We obtain from (134) the equation
v=l
K<£
for the differential of the representation (cf. (151)) SN. Choosing X in t h e Lie algebra of H* so t h a t h(X) = Mkg (1 < & < / < 3 ) , we find, from (154), (176)
Mkf9
= MMff + Mfcotp,
where (99 is a smooth section with compact support)
(1")
MZ& = (it 2
7^9
and
(178)
M^
=
i(p(±-Pk±y
I n formulas (176) through (178), 1 < k < £< 3. The operators M^0 represent the orbital angular momenta, and the operators M^.s represent the spin angular momenta. Note t h a t the operators M^.s (k<£) are bounded, whereas the operators M^.0 are essentially self-adjoint when restricted to the linear manifold of the C00 sections with compact support of the bundle, as is easily seen by applying van Hove's lemma (cf. Chapter VIII) to the manifold Xm +.
RELATIVISTIC
FREE PARTICLES
365
The case of the zero mass particles is discussed similarly. The formulas (176) through (178) remain valid without any change. The Wave Equations. The traditional way of describing the states of a free particle is to p u t down a linear differential equation in space-time whose solutions will form the linear space, the rays of which are to describe the dynamical states of the particles. The equation will in general be invariant under the relativity group so t h a t there will be a natural representation of the relativity group in the space of solutions. Such equations are known as wave equations. We shall now obtain wave equations corresponding to the various particles. There are a few technical points to be attended to in the formulation of these equations. I n the first place, the solutions in general will be Schwartz distributions in space-time. I n the second place, not every solution will represent a state; the correct way to single out the relevant solutions is to define a norm on the space of solutions and take only those whose norm is finite. These norms, which are of course Hilbertian, are in general quite difficult to describe, because the general element is not a function but a distribution. The physical literature in general does not pay attention to these details. An a priori approach to the theory of such Hilbert spaces of distributions on manifolds has been given by Schwartz [3]. From our point of view, we overcome these difficulties quite simply because the Hilbert spaces are already precisely defined; the states are certain square integrable functions (scalar or vector valued) on the orbits. We then treat these as measures on P 4 and take their Fourier transforms. The transformed space will consist of distributions, and we can obtain the form of the equation satisfied by them by a look at the measures on momentum space. I t might be mentioned t h a t when there is no momentum space (for example, when space-time is curved), the above approach will fail. The problem of rigorous descriptions of the wave equations in such cases must be considered open. Given a distribution in P 4 , one cannot define its Fourier transform unless it is tempered. We recall, therefore, the definition of tempered distributions (cf. Schwartz [2]). Let V be a real ^-dimensional Euclidean space with the positive definite product < . , . > and Lebesgue measure dv. For any translation invariant differential operator D and any complex polynomial q on V, the function
is a seminorm on CC°°(V) and the collection of these seminorms induces a locally convex topology for C C °°(F); its completion may be identified with S?(V), the space of O00 functions on V for which sup \q(v)(Dcp)(v)\ < co veV
366
GEOMETRY
OF QUANTUM
THEORY
for all q and D. If {v±, • • •, vn} is an orthonormal basis for V, the operators D V i . .. Vn = (^/^x 1 ) v i • • • (dldxn)vn linearly span the algebra of all translation invariant differential operators, being the linear coordinate functions on V associated with the chosen basis. The topology can t h e n be also induced by the collection of the seminorms sup
1 ( 1 + ^ 2 + . . . +xn2)k(DVu...tVn
. ., xn)\
X\,'".Xn
for various k = 0, 1, 2,- • • and vl9- • •, vn>0. A tempered distribution E is a complex valued linear functional on Oc°°( V) which is continuous in t h e topology induced by these collections of seminorms. By extending these to £f(V) we m a y also regard a tempered distribution as a continuous linear functional on S?(V). For any
(x e V)
and
is a n automorphism of £?(V). Given a n y tempered distribution %->%),9ey(F)), we m a y then define its Fourier transform $ by the equation (179)
E(9)
=
E($).
If L is any linear automorphism of V of determinant + 1, L induces, in a natural fashion, an automorphism of the space £f{V), and (by duality) of the space of tempered distributions. We write (180)
m(
=
EWX),
where
(181)
=
£(L°\
where L° is defined by (182)
(L°v,v'y
=
{v, v' e V).
The m a p
& \E->£ is then a linear automorphism of the space of tempered distributions. Let {vl9- - •, vn} be an orthonormal basis, P and Q two complex polynomials of n real variables, and write q(x1v1 + • • • +xnvn)
= Q(xl9• • •, s n ),
Dp =
P(dldxir-,dldxn).
RELATIVISTIC
FREE
PARTICLES
367
Then, for any tempered distribution E, (qDPE)~ = DQ*P*E, where P* is the polynomial P*{x1v1 +
Vxnvn) = P{ixl9 • • •, ixn)
and
"--«- -'£)• All this is well known and may be found in Schwartz's book [2]. The possibility of describing the free particles by relativistically invariant wave equations in R4 depends on the following lemma. Lemma 9.6. Let s be a real number > 0 and f a complex valued Borel function on P4 such that f \f(p)\2p0-sdam + Jxm Then the distribution^
+
(183)
Tf--
(p) < oo
(TO > 0).
f(p)
(feC™^))
is well defined, tempered, and its Fourier transform Tf satisfies the "wave equation" (\3 + m2)Tf = 0,
(184) where D
d2 dx02
d2 dxx2
d2 dx22
d2 dx32
Proof. We shall show that Tf is a tempered distribution. Consider first the case m > 0 . Then Xm+ is a closed submanifold in P 4 and the above integral obviously exists for all cpeCJJ?*). To prove that Tf is a tempered distribution, we take a k > %s + 2 and rewrite Tf(
T,W) = \ £ [f(Po,?W+mMf)f(Po,9)dp, where G(P) =
P(os-1)l2(l+Po2+Pi2+P22+P32rkl2,
and p0 =
(m2+p12+p22+p32)112.
f Actually a complex Borel measure.
368
GEOMETRY
OF QUANTUM
THEORY (s + 1)/2
Since both G and the function p ^f(p0,v)Po for a suitable constant 0, and all 9? e 0C°°(P4),
are in j£?2(dp), we have,
|7V(
This proves at once that Tf is a tempered distribution. We proceed to the case m = 0. We shall first show that Tf(
+ 1)/2
p -> 9 (^O,P)P ( O S - 1)/2
and
are both square integrable in P 3 , the assertion follows. We now argue as before to conclude that
\Tf(
for suitable constants C and 1c. This completes the proof that Tf is a tempered distribution. We may therefore form the Fourier transform Tf of Tf. Since Tf(cp) = 0 whenever
= 0
or (185)
(Po2-Pi2-P22-Ps2-m2)Tf 2
2
2
2
= 0. 2
The distribution (p0 — Pi — Pi — P3 — m )Tf is tempered and its Fourier transform is (— • — m2)Tf. Hence (185) leads to the equation (U + m2)Tf = 0. Corollary 9.7. / / / moreover vanishes outside a compact subset of P 4 and is also bounded, then Tf is a function on R4—in fact, an entire function and the equation (184) is satisfied in the usual sense. Proof. Tf is now a finite complex measure onP 4 with compact support. It is well known that Tf is then an entire function. Equation (184) then implies the same relation in the usual classical sense. We are now in a position to describe the wave equations in R4. Let us first consider the case of mass m and spin \. The Hilbert space of the representation ZJm' + '112 is the space of sections / of the bundle B^'112 with the norm defined by (144). Writing /
=
(/0>/l>./2\/3)>
RELATIVISTIC
FREE
PARTICLES
369
we have (186)
^
YrPrf(p)
=
mf(p)
r=0
for am + -almost all p; the norm is given by
From lemma 9.6 we m a y now conclude t h a t Tfr is a tempered distribution on P 4 for all r = 0, 1, 2, 3 . Let <X> be t h e Fourier transform of Tf, considered as a (C 4 -valued) vectorial distribution on R 4 ,
O=
(TfJfl,ff2,ff3).
Then O is tempered, and we have, as a consequence of (186),
(187)
J i * 737 0 = »»*• r= 0
^ r
We observe t h a t the equation
(•+m2)
(188)
^ r=0
o
iyrv — <& = m®
(v = l , 2 , . . . , t f ) .
"Xr
If we require the distribution <& to take the values in the space of symmetric tensors of rank N over C4, then (188) corresponds to spin %N; as before, (188) implies the wave equation
(n+™2)# = o.
370
GEOMETRY
OF QUANTUM
THEORY
If N = 2 and the values of O are skew symmetric tensors, then (188) defines the spinless particle of nonzero mass. The zero mass equations are described similarly. The wave equations are
<189)
2 ^ / ^ = 0 r=0
(v = l,2,...,n)
OX
r
and (190)
tyoViVV* = " *
(v = 1, 2, • • •, n)
for the spin \n case; for spin — \n, the equation (190) is replaced by (191)
t y o V i V a W * = <&
(v = 1, 2, • • •, »).
I n all cases, therefore, we have obtained linear differential equations of the first order to describe the states of the particles. The early a t t e m p t s to formulate relativistic one-particle theories led to the second-order equation ( • + m 2 )O = 0 and were therefore regarded with serious misgiving until Dirac discovered the equation (187). I t is sometimes asserted t h a t the concept of spin enters inevitably in a n y relativistic description of a free particle as soon as one a t t e m p t s to formulate Lorentz invariant first-order wave equations to describe the states of the particle. This is a misnomer inasmuch as even the spinless particles are described by such equations (cf. (155)). There is only one mathematically meaningful way to look at spin—namely, as a consequence of the method by which one differentiates between different systems which are covariant with respect to the group of motions of configuration space. We have been somewhat vague as to which solutions of these wave equations are to be singled out for describing the q u a n t u m mechanical states of the particle. These solutions will have to be tempered, b u t this alone is not sufficient. I n fact, consider a tempered vectorial solution O of (187). We m a y then write 0 = l F , where T is a tempered distribution on P 4 . Since ( D + m 2 ) O = 0, (p02-p12-pi2-pA2-m2)xP = 0, so t h a t T vanishes for all elements of CC°°(P4) of t h e form (Po2 -Pi2 -P22 - P 3 2 " m*)
I t can be shown easily t h a t an element of O ^ P 4 ) has this form if a n d only if it vanishes on Xm + u Xm~. Therefore, we can construct distributions x¥+ and *P" living on the submanifolds Xm+ and Xm~, respectively, such t h a t *p = ^ + - | _ Y ~ . Clearly, T " = 0 is a necessary condition (the so-called restriction to positive energy solutions). B u t the actual conditions are very complicated; these imply t h a t ^ " = 0 , ^ " ^ is a measure on Xm+ absolutely continuous with respect to am +, and (dx¥ + ldam + )-p0~112 e
RELATIVISTIC 2
+
4
FREE
PARTICLES
371
+
J? (Xm ,C ,a m ). I t seems difficult to exhibit natural conditions on O itself which are equivalent to these. We shall conclude this section with a few remarks on t h e representation U + t2 0 U + ,~2 which was realized in t h e space of sections of t h e complexified tangent bundle to X0 + (modulo the subspace of sections whose values a t each p e XQ + are proportional to p). If we write (192)
/ = (/ 0 ,/ 1 ,/ a ,/ a )
for a section of the complexified tangent bundle of X0 +, then (170) leads to
(193)
PofoiP)- 2 PMP) = 0. /c=l
Let O = Tf = (O 0 >*l,*2,*3), where Tf is t h e vectorial measure,
Then we have, as {p^—p^—p^—p^Tf
= 0,
• # = 0. Moreover, (193) gives us
(194)
to-*«-a^-^*"-aS*»
=0
-
However, different 0 will correspond in general to t h e same state; t h e condition t h a t <X> a n d O' m a y represent t h e same vector in t h e Hilbert space of states is the Fourier transform of (173), i.e., ^
—
^
= 0
(r, Jc = 0, 1, 2, 3)
(
or (195)
curl T = 0.
The equations Q O = 0 are t h e well known Maxwell equations; the condition (194) is t h e so-called Lorentz condition and (195) gives the gauge condition. I n particle physics, t h e representation which we are discussing now is assumed to represent t h e photon. A closer analysis reveals t h a t the two representations U + t±2 represent t h e photon in its two states of left and right circular polarization. We do not go into these ideas here.
GEOMETRY
372
OF QUANTUM
THEORY
6. I N V A R I A N C E U N D E R T H E I N V E R S I O N S We have so far restricted our attention to the transformation properties of the states of a free particle under the connected inhomogeneous Lorentz group. We shall now examine the behavior of the states under the inversions also. The main result of this section asserts t h a t the irreducible representation of the connected group G* associated with a particle of mass m > 0 can be enlarged in an essentially unique fashion to a representation of the complete inhomogeneous Lorentz group G into the group of automorphisms of the states. We shall also obtain explicit formulas for the operators which correspond to the inversions. We show further t h a t the irreducible representations U + ,±n corresponding to zero mass a n d spin \n do not possess this property of extension. Consider now a representation a : (L,x) -> a(L>x) of the complete inhomogeneous group G into the group of automorphisms of the logic of a Hilbert space. We have seen already t h a t the restriction of a to G is induced by a unique representation U of G*. Let us now take an inversion Ix (T = S, t, or st) and consider a symmetry Sx of the Hilbert space J f (on which a and U act) which induces au% oy Since a
UtLIx,ITx)
=
a
(/ T ,0) a (L,x) a (/ x ,0)5
it follows t h a t the representations (h,x)-^-S^U^^S^1 and (h,x) —>• Uih tI X) induce the same homomorphism into the projective group of 3tf (cf. (69) for definition of hx). This means, as G* has no nontrivial one dimensional representations, t h a t (196)
SiU^ySr1
= UihiJxX)
for all (h,x) EG*. Conversely, let U(h,x-+ U{h>x)) be a representation of G* in a Hilbert space Jf7 whose logic is denoted by £f, and let a(L,x - » aiLtX)) be a representation of G in Aut(j£?) such t h a t a is induced by U. Since t h e group G is the semidirect product of G and the group of inversions Hinv, it is obvious t h a t for a to extend to a representation of G into Aut(J^f), it is necessary and sufficient t h a t there exist elements a s , at, ast of Aut(j£?) such t h a t (197)
at2
is the identity
for r = s,t,st, (198)
ast = a s a t
and, such that, if ST is a symmetry which induces aT, (196) is satisfied for all (h}x)eG*. The conditions (197) and (198) are t h e conditions for
RELATIVISTIC
FREE
PARTICLES
373
Ix —> ax to be a representation of i/ i n v , and (196) is just the condition characteristic of semidirect products. If such az exist, then the m a p (199)
(LIr,x)-+a(LtX)ax
is a representation of G which extends a from G. We shall now analyze the conditions (196), (197), and (198). I n the first place, ST can be unitary or anti-unitary and is determined by aa 0) up to a constant multiplying factor. Since IT is a reflection, S2 must be a multiple of the identity, i.e., # z 2 = £l for a constant £ with |£| = 1. If ST is unitary, we may, by replacing St by £~ 1/2 $ T , assume t h a t Sx2 = l. On the other hand, for anti-unitary Sz, (cSx)(cSx)=ST2 for all complex c with absolute value 1, so t h a t in this case £ is an intrinsic invariant of aa 0) itself. Actually, £ = ± 1 , as the following lemma shows. Lemma 9.8. Let S be an anti-unitary operator in a Hilbert space Jf such that $ 2 = £1 for some complex constant £ of modulus 1. Then £ = ± 1. Proof. Let {e;} be an orthonormal basis for Jf and let e / = # e y . Then {e/} is an orthonormal basis also and there exists a unitary operator S± such t h a t S& = e/ for all j . Write 82=SS1~1. Then # 2 is anti-unitary and $ 2 2 = 1. From 2 S = £1 we obtain the equation S2S1S2 ~1 = ^>S1 ~x. If E - > PE is the spectral measure of Sl9 based on the unit circle, then the last equation leads to
for all Borel subsets E of the unit circle. Since # 2 2 = 1, we obtain PE = P {2E
for all E. This means t h a t £2 = 1. Let U be a representation of G* in a Hilbert space Jf. Then there exists a unique projection valued measure Q on P 4 such t h a t Ua,x) = J
4
exp
i{x,p}dQ(p)
for all x e R*. The support of Q is called the spectrum of U. Lemma 9.9. Let a be a representation of G, the complete inhomogeneous Lorentz group, into the group Aut <j£?, J£ being the logic of the Hilbert space ffi. Let U be the representation of G* which induces a on G. Suppose that (a) the restriction of U to the translation subgroup is not trivial, (b) there exists a constant c such that the spectrum of U is contained in the set {p :p0>c}. Then a(Itt0) is induced by an anti-unitary operator for r = t or st, and by a unitary operator if T = S.
374
GEOMETRY
OF QUANTUM
THEORY
Proof. Consider first time inversion. Let St be a symmetry of J^ which induces aatt0). Let M be the spectrum of U. Then i f is a closed subset of P 4 invariant under H. Now on any orbit Ym (m > 0), p0 can take arbitrary negative values. So M^{p : p0>0}. The condition (196) gives us
Suppose now t h a t 8t is unitary. Let Q be the projection valued measure of the representation x - > U(ltX). Then ex
Uajtz) =
JP4
P
exp
i{I&>p}dQ{p) i{x,Itp}dQ(p),
which shows t h a t x->UaJtX) h a s t h e associated projection valued measure E - > Qit[Ey Since St is unitary, these two projection valued measures would be equivalent. I n particular, their spectra will be the same. This implies t h a t the spectrum of Q is invariant under It. This implies t h a t M = {0}. This contradicts t h e assumption (a) of t h e lemma a n d shows t h a t (It,0) must be represented b y a n anti-unitary operator. T h e argument for (Ist,0) is similar. Since Is = ItIst, (Is,0) must be represented by a unitary operator. We are now in a position to prove the main result of this section. Theorem 9.10. Let U be a representation of G* in a Hilbert space 34? with the property that for some m > 0, QXm + is the identity operator of £?, Q being the projection valued measure of the representation x —>• UltX. Suppose U induces the representation a of the group G into Aut(J^f), j£? being the logic of Jf. Then, in order that a may extend to a representation of the full inhomogeneous group G into Aut(j£P), it is necessary that the representation of the stability group at some point of Xm +, which is associated with U, be self-contragredient. If U is equivalent to one of Um, + J (m > 0, j = 0, | , 1, • • •), U + '°, or U + ,n 0 C7 + ' ~ n (n = l, 2,• • •), then U induces a representation a of G in Aut(j£f), and a has an extension to a representation of G in Aut(J^P); in the first two cases, the extension is even unique. Proof. Suppose t h a t U induces a representation a of G in Aut(«£?), and t h a t a has an extension to G as a representation into Aut(j£f). By lemma 12.9, (/T,0) must be represented b y a n anti-unitary operator 8Z when r = st. We have:
for all (h,x) e G*. Let p° be any point of X m + and let -n- be a representation of the stability group H° at p° such t h a t U is associated with Xm + and n. Then, a direct calculation shows t h a t t h e representation h, x —> Uht_x
RELATIVISTIC
FREE
PARTICLES
375
+
is associated with — Xm and 77. On the other hand, as S is anti-unitary, h, x-> Uh_xis contragredient to U, and hence, b y lemma 9.3, it is associated with - Xm+ and 77°, where 770 is contragredient to 77. Therefore 77 and 77° must be equivalent. Let now U be equivalent to one of Um> + >}' (ra>0), C7 + '°, U + >n © U + >~n. I t is obvious t h a t U maps t h e elements ( + 1 , 0 ) of G* into multiples of t h e identity. Hence U induces a representation a of G into Aut(J*f). We shall first prove t h a t if U~Um' + 'j ( m > 0 ) or if U~U + t0, a has at most one extension to G as a representation into Aut(J^f). I t is enough to prove t h a t for each r, t h e element a ( 7 i 0 ) of Aut(^f) is uniquely determined by U. Suppose t h a t $<1} and $<2) are two symmetries such t h a t
for all h,xeG* a n d r = l, 2. Then (S™)-^ commutes with all UhfX and, being a unitary operator, j* it must be a multiple of t h e identity, as U is irreducible. Hence t h e corresponding elements of Aut(j£?) coincide. We now come to t h e existence. A simple and direct argument can be given, b u t in view of t h e general usefulness of explicit formulas, we shall prove t h e existence p a r t of t h e theorem by actually writing down t h e relevant symmetries. F o r U=Um' + >° we take 3tf = £?2(Xm +,am + ) a n d write t h e representation U in t h e form (Un,x
(Sst
J being complex conjugation. I t is easily checked t h a t
and t h a t SzUn,x8t~
= Uh%jxX.
Note t h a t these formulas also remain valid when m = 0. For U=Um' + J ( j = i , 1 , . . ) we use t h e vector bundle description (154). Jtf* is now t h e Hilbert space of square integrable sections of t h e vector bundle B + J. Let N = 2j and let JN be t h e natural conjugation in t By lemma 12.9, St(r) is anti-unitary for r = t, st and unitary for r = s.
376
GEOMETRY
OF QUANTUM
THEORY
C (g)- • -(g) C (N factors). For any endomorphism A in C let A(N) be the endomorphism A ( g ) - - - ( g ^ 4 i n C ® - - - ® C . We now define St by (S*P)(P)
(201)
= (yoy^P^^),
(%>)(*>) = ( y ^ s ) ™ ^ - / * ! > ) , ( ^ ) ( P ) = Yom9(I*P).
Here, j9 e Xm + ; note that — Itp = Isp e Xm + . We observe that y0, y1? *y2> 73 a r e an* rea, l- Hence (cf. (147)) for v = 1, • • •, N, y0v, y±v, iy2v, y$v are also real. Therefore y0v, y / , and y3v commute with JN while y2v anticommutes with JN. It also follows from the same observation that yff\ (yiys){N\ and (y0yiy3)(AO commute with JN. This implies at once that
( 1 PrV/yYoymfV* = (yoyiy 3 r«/»(i P r y/), (202)
( 2 ^ / ) (yiy3)W)^ = ( y i y s r ^ ( | P,
v),
( | ^ v ) r r = yr(iKyrv). where 1
(203)
(T = t,st),
It follows from the equation S(h)y0 = y0iS'(^*-1) (cf. (136)) that Ss satisfies (196). We claim now that St also satisfies (196). This reduces to proving that (204)
(Ym)mJ»SH(h*-i)
=
Ss(h)(YlYar»Js.
(A
Since JN commutes with (yiy3) °, this reduces to proving that
(yiy3)™s(h*-y»> =
jNs(hr(yiY3rjN.
Now, JN is the natural conjugation and so we have
^ W " W ) w v * = (JS(h)(yiY3)jy»> where we write J for J±. A trivial calculation shows that (yiya)^*"1) = and this proves (204).
JS(h)(7ly3)J
RELATIVISTIC + N
FREE
PARTICLES
377
+t N
The case U~U ' @ U ~ still remains to be examined. The symmetries ST are again defined by (201). They still satisfy (203). This time, t h e ST operate on the space of square integrable sections of the sum B + -NI2@B + -~NI2. A trivial calculation (cf. (167)) shows t h a t T v anticommutes with both y(0iV) and {yoyiYz){N)JN while it commutes with {yiyz){N)JN {v~ 1, • • •, N). From this the results concerning U follow quickly. Note t h a t both Ss and Sst intertwine U + 'N and U + '~N while ^ l e a v e s the subspaces corresponding to U + ,±N invariant. Corollary 9.11. The representations U+'±N of G cannot be extended to representations of G into the corresponding groups of automorphisms of the underlying logics. Proof. The representations of the stability subgroup a t (1,0,0,1) are the characters mZtCL -> z±N which are not self-contragredient.
7. LOCALIZATION I n Chapter V I I I we described u p to physical equivalence all systems which are irreducibly covariant with respect to the group of all motions of a Euclidean configuration space. I n three dimensions this led to t h e classification of these systems through their spin and to the familiar expressions for the position and momentum observables. On the other hand, in this chapter we have described the Lorentz invariant free particles and it would be a natural question to ask whether we could write expressions for the position operators of these particles. To make this question precise, we select a particular inertial coordinate system, and ask, for x0 = 0 in this frame, whether a suitable projection valued measure E -> PE can be defined on the Borel sets of R3, the projections acting on t h e Hilbert space Jf of the particle, such t h a t the range of PE represents t h e experimental statement t h a t the particle at time £ = 0 belongs to E. We shall take the same point of view as we did in Chapter V I I I a n d require P(E —> PE) to be such t h a t the system is covariant with respect to the group of rigid motions of R3. This condition can be rephrased in a simple manner. Let O be an observer and let x0i xx, x2, x3 be his time and space coordinates. We identify the space of 0 with R3 and obtain the projection valued measure P on the a-algebra of Borel subsets of R3. If (h,a) e G* is such t h a t a0 = 0 and h e K*, then the new observer 0' determined by (h,a) differs from 0 only in space origin and space orientation. Since U(ha) represents the operator which expresses the transformations of states associated with the change from O to 0', the above mentioned condition of covariance can be expressed as follows: (205)
Uh,aPEU^a
= Pdm[E]
+a
(h e K*, a0 = 0).
GEOMETRY
378
OF QUANTUM
THEORY
The problem of describing the position operators can then be regarded as the problem of determining, for a given representation U of G*, all the projection valued measures P such that (205) is satisfied. Let us write V for the representation U when restricted to the subgroup of all (h,a) such that a0 = 0 and h e K*. Let us call this subgroup M. The pair (V, P) is a system of imprimitivity for M based on R3 and hence we can use the results of Chapter VIII to conclude that there exists a unitary isomorphism W of 3^ onto j£?2(R3,Jf,dx) and a representation g -> s(g) of the group K* in JT such that for (h,a) e M, (206)
(WVKaW-y)(x)
=
s(h)f(S(h)-Hx-z)),
and (WPEW~lf)(x)
(207)
=
X E (X)/(X).
In other words, a necessary and sufficient condition that there should exist at least one projection valued measure P satisfying (205) is that the representation V be unitarily equivalent to the representation, say Ls, of M described by the right side of (206). If W is a unitary operator effecting this equivalence, P may then be defined by (207). The representation appearing on the right side of (206) can be described in another, equivalent, form through the use of Fourier-Plancherel transforms. Let us consider the unique unitary automorphism (208) 2
^:/->/
of i? (R ,Jf;dx) onto itself such that for all
3
f f(x)(p(x)dx=
JR8
f / ( x ) ( f (2n)-*i*e-i<™>
WR3
/
Then a straightforward calculation shows that (210)
(J^U^/Hx) =
exV[-i(a,x}]s(h)f(8(h)^x),
where <.,. > is the usual Euclidean inner product. We may obviously change the measure dx to any rotation invariant measure in the same measure class without changing the form or the equivalence class of the representation defined by (210). We have therefore obtained the following lemma: Lemma 9.12. Let V be a representation of M in some Hilbert space &\ Let m be a real number > 0. Then, for the existence of a projection valued measure P based on R3 such that (V,P) is a system of imprimitivity, it is necessary and sufficient that there exist a representation s, s :h-> s(h)
RELATIVISTIC
FREE
PARTICLES
of K* in a Hilbert space Jf~ such that V is equivalent to the representation ofM in £>z(R*,jfyam) defined by (211)
(Vsh,af)(y)
=
e x p ^ a ^ M ^ ^ ) '
1
379
Vs
^ ,
am being the measure given by d
7_
(212)
2(m 2 +
Note t h a t m plays a completely harmless role in this lemma. We shall now discuss the problem of recognizing whether a given representation of M is equivalent to some Vs as defined by (211). To this end, we need to consider representations more general t h a n the representations Vs. Let X = R 3 - {0}, and let us regard X as a Z*-space with t h e action h, y -> &(h)y. Given a strict (iC*,X)-cocycle C with values in the unitary group of a Hilbert space J^l we consider the representation Vc of M in £>2(X,Xcxm) given by (213)
= exp[-i
(VU)(j)
The representation Vs is obtained if we take C(h,y) = 8(h). The strict cocycle C determines for each r > 0 , the strict (iT* ? ^ 1 )-cocycle Cr, (214)
Cr(h,u) =
C(h,ru),
where S1 is the unit sphere. Note t h a t i£* is transitive on S1. The next lemma determines the equivalence class of the representation Vc completely and it is our basic tool in the subsequent analysis of localizability. We make a few comments on (213). The projection valued measure Q(E -> QE), where QE is multiplication by XE> ^S determined in a direct fashion by the Vx,a'Vf.a = J* exp[-i]e*0(y). From this it follows t h a t the unitary equivalence class of Q is determined by the unitary equivalence class of Vc. I n particular, the dimension of JfJ which equals the multiplicity of Q, is determined by the unitary equivalence class of Vc. We recall t h a t T' is the subgroup of K* consisting of all diagonal matrices ml
,0 = (J \.)
(Id - I)-
T' is the stability subgroup of X * at the points (0,0,r) ( r > 0 ) .
380
GEOMETRY 1
OF QUANTUM 2
Lemma 9.13. Let C and C be two strict the unitary groups of Hilbert spaces J T 1 and Vcl and V°2 of M are equivalent if and only (K*^ycocycles (7/ and C2 define equivalent
THEORY
(K* ,X)-cocycles with values in JT 2 . Then, the representations if for almost all r>0, the strict representations of the subgroup
T'ofK*. Proof. Let R+={r :r>0} and Y = E + xS1. The m a p (r,u)-> ru is a homeomorphism of Y with X which sends (r,8(h)u) to 8(h)(ru). I t is easily shown t h a t the measure <xm on X goes over to a multiple of jS x a on Y under this map, where a is the normalized rotation invariant measure on S1, and dp = r2(m2 + r2)~ll2dr. Suppose now t h a t Vcl and Vc2 are equivalent representations of t h e group M. Then dim J T 1 = dim JT 2 so t h a t we may assume t h a t J T 1 = j T 2 = Jf^ Let D be a unitary operator which effects the equivalence. D commutes with all the multiplication operators A&, (AJ)(x)
= exp[-t]/(x).
From this it follows easily t h a t D commutes with all operators QE of the form (fe/)(x) = X*(x)/(x). From lemma 6.4 we conclude t h a t D = D~ for some Borel m a p of X into the unitary group of Jf, i.e., (Df)(x)
= D(x)/(x)
x->D(x)
(x G X).
The condition t h a t D intertwines the restrictions of Vcl and Vc2 t o Z * now leads to the equation C2(h,x) =
DiSihjxjC^h^Dix)-1
for each h e K* for almost all x. Going over to Y = R+ xS1 Fubini theorem, we obtain from this the equation
Cr2(h,u) =
and using
D^h^C^h^D^u)-1
for almost all triples (h,r,u), where Dr(u) = D(ru)
(ueS1).
I n other words, for j3-almost all r, the strict (iC*,AS1)-cocycles Crx and C2 are cohomologous. From theorem 5.27, as Z * is transitive on 8V it now follows t h a t for almost all r, the representations of the stability subgroup T' of K*, defined by Crx and Cr2, are equivalent. We now come to the converse. Suppose C1 and C2 satisfy the condition described above. We may assume t h a t J T 1 = JT 2 = J ^ Then the strict (Z*,^ 1 )-cocycles 0 / and C2 are cohomologous for /3-almost all r. Let X[
=
^2(S\X(J).
RELATIVISTIC
FREE PARTICLES 2
X
381 1,r
Let the representations of K* in ££ (S ,CtiC,a)—say A defined by
and
2,r
A —be
Let Q1 be the projection valued measure based on S1 defined by (E £ S1).
(e*7)(») = Jte(«)/(u)
Then, for any r for which (7/ and C2 are cohomologous, there exists a unitary operator Dr in Jfi commuting with all QEX and having the property (cf. theorem 6.12) DroA^oD,'1
= A2>\
Since Dr commutes with Q1, Dr commutes with the operator in Jf[ of multiplication by the function u -> exp *> for all a, r. If we then write Wja\rh for the operator in J^[ defined by (215)
(WtfJHu)
= exp[-fr](^7)(w),
then (216)
DroWl?aoDr-i
= Wl-:a
for all (h,a) e M. W1,r and W2,r are representations of M in Jf[ and their dependence on r is Borel; (216) is valid for /^-almost all r > 0. We shall now show that we may choose Dr so that r -> D r is a Borel function. To this end, we rewrite (216) as follows:
Let us now denote by °UX the unitary group of C^[ and consider the set 3f of all pairs (D,r) e(%1x B+ such that
for all (h,a) eM. Since TFy,r is a representation, it is a continuous map from M into (^/1 and hence it is clear that the set S is a Borel subset of f 1 x i ? + . Equation (216) shows that the map D, r ->r maps 3} onto almost all of R +. It follows now from von Neumann's cross-section lemma (cf. Chapter V) that there exists a Borel set NgzR+ of measure zero and a Borel map D' : r -> D'(r) of £ + - i V into ^ such that (D'(r),r)e@ for all reR+-N. be the unique unitary operator in J£2(R + ,J^,/3) defined by (D'~jF)(r) = D'(r)F(r) +
forreR -N.
Let D'~
382
GEOMETRY Ur
The definition of W in £>2(R +,X[fi) by
OF QUANTUM
THEORY
makes it obvious t h a t the operators WjhtQ> defined (W{,aF)(r)
=
WfcFir)
are unitary, t h a t Wj is a representation of M, and t h a t Df ~ intertwines W1 and W2. I n particular, W1 and W2 are equivalent. Now, the Hilbert space J?2(R +,J?l,f3) is canonically isomorphic to ^2(R + xS1, XI j8x o-), which in t u r n is canonically isomorphic to ^2(X,Jf,am) under the unitary isomorphism induced by the measure preserving transformation r, u - > ru of R + xS1 onto X. A little calculation now shows t h a t the representation Wj goes over to the representation V°3 in the Hilbert space J!^ 2 (X,J^a m ). Consequently, Vcl and V°2 are equivalent. This completes the proof of the lemma. Corollary 9.14. In order that there exist a projection-valued measure P based on R 3 such that (V c \ P) is a system of imprimitivity for the group M, it is necessary and sufficient that there exist a representation s s : h->
s(h)
of K* acting in X, such that for almost all r > 0, the restriction of s to the subgroup T' is equivalent to the representation h ^(7 1 (/^,(0,0,r)) of T'. Proof. Let s be a representation of K* and let Cs be the strict cocycle defined by Cs :h,x-+s(h).
(K*,X)-
Then the necessary and sufficient condition for the existence of a suitable P is simply t h a t , for a suitable s, C1 and Cs satisfy the condition described in lemma 9.13. The corollary now follows a t once if we observe t h a t 8(h) fixes (0,0,r) if and only if h = mc 0 for a suitable £ with |f | = 1. Corollary 9.15. Let (V,P°) be a system of imprimitivity based on R 3 for the group M and let P° have finite multiplicity.^ Then, an arbitrary projection valued measure P based on R 3 his the property that (V,P) is also a system of imprimitivity for M if and only if there exists a unitary operator A commuting with V such that (217)
P =
AoP°oA-\
Proof. If P is defined by (217) for a unitary A lying in the commuting ring of V, it is clear t h a t (V,P) is a system of imprimitivity for M. Conversely, let (V,P) be a system of imprimitivity based on R 3 for M. Then (V,P) and (V,P°) are equivalent to the systems induced by certain representations s and s° of K*. Since V ~ Vs and V~ Vs°, we have Vs ~ Vs°. f Since M is transitive on R3, P° is homogeneous (lemma 9.10). Hence it has a well-denned multiplicity.
RELATIVISTIC
FREE PARTICLES
383
From lemma 9.13 applied to the strict cocycles h, y->s(h), h, y-+s°(h), we conclude t h a t the restrictions of s and s° to the diagonal subgroup T' = {m^0} are equivalent. Now a look at the formula (211) shows t h a t t h e multiplicity of P° is the dimension of s°. Hence, s° is finite dimensional. Consequently, s is finite dimensional also. Now, it is clear t h a t two irreducible representations of K* are equivalent as soon as their restrictions to T' are equivalent. Using the decomposition into irreducible components, one can extend this result easily to finite dimensional representations which are not irreducible. This result implies t h a t s and 5° are equivalent. I n other words, the systems of imprimitivity (V,P) and (V,P°) are themselves equivalent. Consequently there exists a unitary operator A commuting with V which satisfies (217). This finishes the proof of the corollary. We are now in a position to prove our basic result on the existence of position operators. Let (S be a system to which is associated the representation U of the inhomogeneous Lorentz group in the Hilbert space J f . Then <& is said to be localizable at time x0 = 0 if there exists a projection valued measure P (in Jt) based on R 3 such t h a t
for all (h, a) e M. The following theorem is the main result of this section (cf. Wightman [1]). Theorem 9.16. Let the representation U associated with a system S be defined by the pair (Xm + ,7T), where m > 0 , and TT is a representation of the stability subgroup of L* at the point p(m), where we write p(m) for (m,0,0,0) when m > 0 and (1,0,0,1) when m = 0. If m>0, (3 is localizable. If m = 0, 0 and E* is the stability subgroup at p(m) for m = 0. Let us now select a strict (i/*,X m + )-cocycle C with values in the Hilbert space JT in which IT acts, such t h a t G defines the representation TT at p(m\ The representation U acts in 3tf> = £?2(Xm + ,X<Xm + ) and, for (h,a)eM, we have (Uh,af)(p)
= exp[-^•M^,S(^-»/(S(^)-1^).
We now use the bijection p -> ((m2 +
384
GEOMETRY
OF QUANTUM
THEORY
defined by C, has a restriction to the diagonal subgroup T' that is equivalent to the restriction of 7/ to T'. Consider now the point pr^m = ((m2 + r2)ll2,0,0,r). Let hm be the 2 x 2 diagonal matrix with entries A and 1/A, where rm- 1/2 [(m 2 + r 2 ) 1 / 2 - r ] 1 / 2 (m > 0), ~ \r-112 (m = 0). (m) It is easy to check that S(hm)prm=p . If we now write H*m for the stability group at pr?m, then T'= H*m n K* and its elements commute with hm. From the cocycle identities we now find, for h e T\ C{K,p™) = =
C(hmhhm-\p^) C(hm,Pr.ntMh,Pr.mMK-W*>)
so that C(h>Pr,m) =
RC(h,p(m))R-\
where R = C(hm,prtm)-\ In other words, the representations of T' defined by C at the points ((m2 + r2)1/2,0,0,r) are all mutually equivalent and are equivalent to the restriction of the representation rr to T'. We may now apply the criterion for localizability described in the previous paragraph to conclude that (& is localizable if and only if for some representation 7/ of K*, n, and IT' have equivalent restrictions to T'. For m > 0 , we may take 77 = 77-' always and so any system (& whose associated representation is defined by the orbit Xm+ with m>0 is localizable. For ra = 0 the condition obtained is exactly the one stated in the theorem. Corollary 9.17 (Newton- Wigner [1]). The particles corresponding to the representations Um' + 'J' (ra>0, j = 0, J, 1,- • •) and U + t0 are localizable, whereas those corresponding to the representations U + ,±n (n>l), are not. The system corresponding to the representation U + ,1@U + *~1is localizable. The photon is not localizable. Proof. The localizability of the systems corresponding to Um' + 'j (ra>0) and U+>° is immediate from theorem 9.16. Since K* has no nontrivial one-dimensional representations, the system corresponding to U +'± n is not localizable if n > 1. We now come to the systems whose corresponding representations are reducible. The representation U + 'n © U + '~n is associated with (X0 + ,7r(n)), where 7r(n) is the representation
of E* (cf. (84)). For 7i = l, TT(1) is obviously the restriction of D112 to T. Hence the system corresponding to U + tl © U + '~1 is localizable. If
RELATIVISTIC
FREE
PARTICLES
385
j
n>2, the representations D whose restrictions to T' contain the character m c 0 -> £n are precisely those for which 2j>n and 2j — n is an integer; the dimension of Dj will then have to be > 2 . Hence none of the systems corresponding to U + ,n 0 U + ,~n, n>2, is localizable. For n = 2, we obtain the result t h a t the photon is not localizable. Remark. The representation of K* of the lowest dimension whose restriction to T' contains the character mc 0 -> £2 is D1, namely the representation k - » 8(k) (k e K*). But the restriction of D1 to T' contains the characters m c 0 -> £ ± 2 a?i^ the trivial character mc 0 -> 1. I n other words, the failure of the photon to be localizable stems, in the last analysis, from the fact t h a t , for a given momentum vector, it has only two independent spin angular momentum states and that, in particular, it does not have any spin angular momentum parallel to its momentum. The particles with nonzero rest mass have been shown to be localizable, and for these there remains the problem of obtaining the expressions for the position operators. Such formulas were written by Newton and Wigner [1]. We shall now t u r n to this question. Instead of the position operators which are unbounded it is more convenient to work with the unitary representation of the additive group of R 3 they generate. To be more explicit, let U be the representation associated with a localizable system and let P be a projection valued measure based on R 3 such t h a t the relations (205) are satisfied. We define, for xeR 3 , B(x)=
JR 3
exp[-i<x,y>dP(y)] = e x p [ - i g ( x ) ] .
The B(x) are u n i t a r y and B(x^B(x)) is a representation of R 3 . The q(x) are the position operators. The correspondence (218) between B and P is one-one. I t is easily checked t h a t P satisfies (205) if and only if (218)
Uh,aB(x)U^a
=
exV[i<8(h)x,&>]B(8(h)x)
3
for all x e R and (hya)eM. Let us define Whx and .L(a) as follows: WhtX =
B(x)Uhi0,
i ( a ) = Uu.a. Here (h,a) e M. The set of all pairs (h,x) is identified also with M. I t follows after a brief calculation t h a t W:h,x-+Wh.* is a representation of M and t h a t (218) is equivalent to (219)
Wh,*L(*)WzX
= exp[»<8(A)a,x>]£(8(A)a).
386
GEOMETRY
OF QUANTUM
THEORY
We shall now use (219) to write by inspection the expressions for the B(x). Let m > 0 and let U be the representation Um' + ,° acting in Jf = o£?2(Xm + ,a m + ); noting t h a t when a = (0,a), {a,p} = — we find
(Uh.J)(p) =
exV[-Ka,V)]f(8(h)^p).
The m a p establishes a homeomorphism of Xm+ with P 3 . We shall use the notation r q t o denote the action in Xm+ obtained b y lifting t h e translation in P 3 v i a 7):
(220)
v(^(p))
= P + q-
Since the elements of K* leave pQ invariant, an easy calculation shows that 8(h)rQ8(h)-1
= r^)q
(hsK*).
From this it follows t h a t the map (h,q),p-+T_9(8(h)p) converts Xm + into an M-space. I n the Hilbert space ^ L(SL) is the operator of multiplication by the function p —> exp *"P> and (219) shows t h a t we have a system of imprimitivity for M based on Xm + considered as an M-space. The measure am+ is, however, not invariant under the " t r a n s l a t i o n s " r a and, on noting this, the following expression for the B(x) m a y be immediately written:
(22D
(*(,)/)<„) = { ^ p ^ ) ,
where p0 is, as usual, the function p -> (m 2 +
here plt p2, p3 are global coordinates on Xm+ and the d\dpk are t h e corresponding derivatives, w h i l e / i s a smooth function vanishing outside a compact set. The most general B satisfying (218) can now be determined by a n application of corollary 9.15. Let W be a n y unitary operator commuting with the representation U restricted to M. Commutativity with all
RELATIVISTIC
FREE
PARTICLES
387
a
Ui,a ( o = Q) leads at once to the fact t h a t W is multiplication by a Borel function, say u, such t h a t \u(p)\ = 1 for all p. The commutativity with all Uht0 (he K*) now forces u to be i£*-invariant and hence to be a function of p0. Hence the most general B such t h a t x -> B(x) is a unitary representation of R 3 satisfying (218) is of the form
^[i{k{Po)(p)-k(p0)(r^))}[^^l2f(^p),
(B(x)f)(p) =
where Jc is an arbitrary real valued Borel function on the positive reals. We shall now proceed to the case of nonzero spin. Let us take U=Um, + J in the form (154), acting in the Hilbert space J f of sections of the bundle B + J. Let us assume first t h a t j = \. I t is natural to assume for B(x) an expression of the form (B(x)f)(p)
{-!^L^ll2S(R(x,p))f(Txp),
=
where R(x,p) e H* and S(h -> S(h)) is the usual spin In order t h a t S(R(x,p)) m a y transform the fiber at it is enough to assume t h a t 8(R(x,p)*) maps p onto may be a representation satisfying (218), we should x, p —> R(x,p) satisfy the conditions: (i) /99ON
(")
(223)
hR(x,p)h~1
^ ( x + x ,^) =
B(0,p) (hi)
= R(8(h)x,8(h)p)
r
representation ofH*. r^p to the fiber a t p, r^p. I n order t h a t B require t h a t the m a p (heK*),
R(x,p)R(x',r^p),
= 1,
8(R(x,p)*)p
=
r^p.
An easy calculation using (152) shows t h a t each B(x) will be unitary. We shall now " solve " equations (223) for the R(xtp). R 3 acts freely and transitively on Xm+ by the action x, p -> r^p, and (ii) of (223) just asserts t h a t x, p-> R(xip)~1 is a strict (R 3 ,X w + )-cocycle. Therefore, by lemma 5.23 we m a y write R(x,p) = Lipy^L^p), where p —> L(p) is an arbitrary function. The condition (i) then becomes, on simplification, L(8(h)p)hL(p)~1
=
L(8(h)r^p)hL(r^p)-\
This shows t h a t the function p -> L(8(h)p)hL(p)~1 is r x -invariant for all x, and therefore must be a constant, possibly depending on h: L(8(h)p)hL(p)~i
= Ch
(heK*).
Equation (iii) of (223) can be analyzed similarly. We have, on substituting the expression for R in (iii), 8(L(p)*~i)p
=
8(L(T*p)*-i)(T^p).
388
GEOMETRY
OF QUANTUM
THEORY
-1
This shows that p -> S(X(p)* )^ must be a constant also. If we use the map p->A{p)
=
p0cr0+p1G1+p2a2+p3a3
4
which sends elements of P into 2 x 2 Hermitian matrices, and write the condition that p —> S(L(p*)~1)p has the same value for all p, we get L(p)*~1A(p)L(p)-1
mL(p^)*-1L{pim))-1,
=
where p(m) = (m,0,0,0). (hi) of (223) therefore can be replaced by m{L(p^)-1L(p)}*{L(pim))"1L{p)}'
A(p) =
It is now easy to determine a solution L. We define L(p(m)) = 1 and (224)
L(p) =
m-ll2A(p)112.
Note that for p e Xm+ (m>0), A(p) is positive definite so that L(p) is well defined and detZ(^) = l. We then obtain, for R, the following expression: (225)
S(x9p) =
Aipy^Ai^p)1'2.
We note that (i) of (223) is automatically satisfied (Ch = h). Collecting all this, we obtain (226)
i^M-\mS{A{p)-^A{r^)^)f{r^).
(B(x)f)(p) =
Actually we now use (226) to define B. It is not difficult to check that the B(x) are unitary and that x -> B(x) is a unitary representation of R3 satisfying (218). For higher spin, we replace (226) by (227)
^^^y2SN(A(p)-^A(^p)^)f(r^ph
(B(x)f)(p) =
where SN=S ® • • • ® 8 (N = 2j factors). We can now write the infinitesimal form of (226). To do this it is necessary, among other things, to compute
Obviously, if we write fH(x,P) =
A{p)-^\jtA(r^pY^^Q,
then ^(x,^) is a matrix of zero trace and
di(x,p) =
m(x,p)),
RELATIVISTIC
FREE
PARTICLES
389
where S is the differential of the representation S. From (i) of (223), we have: SWdtfapWh)-1
(228)
= 9t(8(h)x,8(h)p)
(heK*),
so that ffi is uniquely determined by its values for all x and all p of the form p = (p0,0fi,p3). For fixed x and such p, we find, on writing u =
jt{A(rtxPy^0
and differentiating the equation [A(rt^p)ll2]2 = A(Tt^p), that A(PY"u
+
uA(P?"
= h^o+P3)IPo \ x1 + ix2
Zi-ix, \ -x3(Po-P3)IPo/
Using the fact that u is Hermitian we can compute it from this equation. Then 9l(x,p) reduces, after some simplification, to m/
/
Zs/ZPo +m XiXi +ix2)IPo-Ps l;
*>p)x = I , , .
(Xi-ix2)/Po+P3 + ™\ -^3/ 2 Po /
Expressing %l(x,p) as a real linear combination of the ar and the iar, and using (54) and (134), we find, after some calculation, that 3 r=l
1 2m(p0 + m)
{^2^37273-^1^37371}-
This formula is valid for p = (p0fi,0,p3). A look at the requirement (228) of covariance now enables us to write the expression for $i(x,p) in the general case: m{X p)
'
= 2mMp!lm)I
PMt +
^ZxffoY,
(229)
The fact that dl(x,p) as defined above satisfies (228) for all unitary k is straightforward. It is also trivial that, for p1=p2 = 0 (229) reduces to the earlier expression. We omit the verifications of these. We finally obtain, from (229), the following expressions for the position operators q(x):
(230)
iq(x)f = - J xHdp-?- f+^i 2p fc= 1
k
2
Q
/-St(x,•)/
390
GEOMETRY
OF QUANTUM
THEORY
for a n y smooth section / vanishing outside a compact set. The generalizat i o n of (230) t o arbitrary j is trivial. If we define dlv(x,p) b y (229) with y / replacing y;- (cf. (147), for the meaning of y / ) , and write
m(x,P) = 2 9HX>:P)> v=l
t h e n (230) once again gives the expressions for the iq(x) when j = ^N is arbitrary. We shall now take u p t h e question of uniqueness. W e consider first j = i. I n view of corollary 9.15, we m a y replace the unitary representation BbyWoBoW'1, where IF is a unitary operator which commutes with all Uhf(l (h,a) e M, without destroying the imprimitivity equations (205). Now, for any such W, the fact t h a t t h e UltCL (a 0 = 0) commute with it shows t h a t there exists a Borel m a p w : p —> w(p) of Xm+ into the unitary group of C4 such t h a t , for each p, w(p) leaves the fiber of B + '112 a t p invariant, a n d (W
=
S(h)w(p)
for each h for almost all p. For p = (p0,0,0,p3) with p3>0, the S(h) for those h with 8(h)p=p are diagonal. I t is then not too difficult to determine t h e w which satisfy these equations. We choose a Borel function v, r —> v(r), of the positive reals into t h e group of 2 x 2 diagonal matrices of K* arbitrarily and define w by w(pr) = (V{^ ° ) \ 0 v(r)J w(8(h)pr)
(pr = ((m2 + r 2 ) 1/2 ,0,0,r), v(0) = 1),
= SihMpMh)-1
(heK*).
I t is clear t h a t w is well defined and meets the requirements. For a given w, the unitary operator W corresponding t o it is given b y (W9)(p)
= w(p)
(PeXm+).
The unitaries thus obtained exhaust t h e commuting ring in question. The case of higher spin is handled similarly. I t is noteworthy t h a t t h e position observables cannot be uniquely determined. Formulas (222) and (230), however, seem to give the simplest of the possible choices. For the choices represented b y (222) and (230), we can calculate the velocity observables b y differentiating the position
RELATIVISTIC
FREE
PARTICLES
391
observables with respect to time. Since (cf. Chapter XI) these are their commutators with energy, we find, using (230) and writing qk = q(ek), (231)
vk = qk =
pk/p0.
The velocities are conserved in time. Formulas (231) are the same as those in classical relativistic mechanics.
8. G A L I L E A N R E L A T I V I T Y Our concern so far has been exclusively with the relativity prescribed by the Lorentz group. Clearly the whole circle of problems can be analyzed under the over-all assumption t h a t the space-time group is the inhomogeneous group of Galilean transformations. We shall indicate briefly in this section the modifications t h a t are to be made in the foregoing analysis when we change the underlying group. To describe the free particles under t h e assumption of Galilean relativity, we must describe and classify u p to physical equivalence the irreducible (possibly projective) representations of the inhomogeneous Galilean group. We carry out our discussion along the same lines as we did for the Lorentz case. Let us now consider the universal covering group of the inhomogeneous Galilean group (cf. (112) through (123) of Chapter VII). This consists of t h e quadruples (h,r},Y,u) (JeP^eR^^ueR8), with the multiplication law (/&,77,v,u)(/&',7/,v',u') = (hh\ 77 + 77', 8(A)v' + v, S(/*)u' + u +77'v), w h e r e h , h'eK*, TJ, T/'GR 1 , U, v e R 3 . The space-time translations (l,r/,0,u) form an abelian normal subgroup; the group of the homogeneous Galilean transformations (h,0,v,0) acts on the translations (via inner automorphisms) as follows: (hfl,v,0) : (l,77,0,u)->(l,77,0,3(^)u + 77v). This gives, of course, the semidirect product structure of the inhomogeneous Galilean group. We introduce " m o m e n t u m space" P 4 of points (Po>P) a s the dual °f the vector space of space-time translations (l,77,0,u). Each (^o>P) gives rise to the character (l,i7,0,u) - > exp[^ 0 77 + < p , u » ] , where < . , . ) is t h e Euclidean inner product in R 3 . The adjoint action of (^,0,v,0) in P 4 can now be computed; an easy calculation shows this to be: (232)
(hfi,v,0)
: (p0,V) -> (2> 0 -
392
GEOMETRY
OF QUANTUM
THEORY
From (232) one can easily obtain the orbits for the homogeneous group in P 4 . We claim that the orbits are (233)
ZP0 = {(p0,0)}
(-00 < p0 < oo),
and (234)
Xr = {(^0,p) :
(r > 0).
That ZPo, which consists of a single point, is an orbit is trivial. On the other hand, (232) shows that Xr is invariant. We shall show that any point of Xr is obtained from (0,(0,0,r)) by applying a suitable element (&,0,v,0). In fact, (232) shows that any point (0,p) with (p,p> = r2 can be so obtained; to pass from (0,p) to (p0>P) w e choose v e R3 so that
-
(o °-i)
(|£| = 1)
and v = (vl9v2,0)
(vl9 v2 arbitrary).
The product of (m?0,0,v,0) and (ra^o,0,v',0) is (racc,o,0,v",0), where v" = S(raCf0)v' + v. On writing z(v) =
v1-iv2,
we obtain for z(v"): z(v")
=
PZ(V')+Z(Y).
In other words, the stability subgroup at (0,(0,0,r)) is isomorphic to the semidirect product of C and T under the mapping (mco,0,v,0) - ^ (£,z(v)), wherein the product in T x C is determined according to (235) (£,*)(£',*') = (K',£V + *). The orbit structure of P 4 is obviously smooth. For the stability subgroup at (0,(0,0,r)) we have already described the representations. These are the 7rn, "n : K, 0 ,0,v,0) -> {» (n = 0, ± 1, ± 2, • . .) and the TTP and TTP' (p>0); the latter act in ££2(T) and are described by (cf. (98) and (99)) K(m Ci0 ,0,v,0)/)(0 = e x p [ ^ R e ( ^ - M v ) * ) ] / ( r 2 ^ ) , (7r/(mc,0,0,v,0)/)(0 = ? e x p [ ^ R e ( ^ - ^ ( v ) * ) ] / ( r 2 0 .
RELATIVISTIC
FREE
PARTICLES
393
n
The representations L associated with any one of these irreducibles, say 77, m a y now be written explicitly. We only note t h a t (236)
(^?lf,f0,u)/)(l>o»P) = exp[^(7?^0 +
/ belonging to a Hilbert space of scalar or vector valued functions on Xr. The representations associated with ZPo are (237)
(/M?,V,U) - >
exp(ip0r))7r(h,0,v90)9
where -n is an arbitrary irreducible representation of the homogeneous group. I t is remarkable t h a t none of these representations can represent a n y physical particle. We shall see now why this is so. Formula (237) shows t h a t the time translations act as scalars and hence leave every r a y fixed— there is no temporal evolution. Hence it becomes necessary t o give u p these representations. We come next to t h e representations Ln associated with t h e orbits Xr ( r > 0 ) . Formula (236) shows t h a t t h e momentum operators Blt B2, B3 corresponding to t h e translations along t h e axes in space are the operators of multiplication by pl9 p2, and p3. B u t p±2 + p22+p32 — r2 and hence (238)
B±2 + B22 + B32 = r 2 l ,
which gives an unphysical relation between the three momenta. Moreover, it can be quickly seen t h a t the systems corresponding t o these representations are not localizable either. To see this, let us write V for t h e restriction of a n y one of these representations Ln t o t h e subgroup of translations in space only and let us assume t h a t there exists a projection valued measure P(E -> PE) based on t h e cr-algebra of Borel sets of R 3 acting in the Hilbert space 3%* of V such t h a t * w*E
' u
=
*E
+u
for all u e R 3 and E c R 3 . From theorem 6.17, we know t h a t there exists a unitary isomorphism W of 2tf with J5?2(R3,Jf,dx), where X is a Hilbert space, such t h a t
((WVuW-i)f)(x)=f(x-u) for all feJ?2(R3,Jf,dx). Now, u -> Vu is a representation of t h e abelian group R 3 a n d hence there exists a unique projection valued measure Q on the space of vectors p such t h a t Vu =
JP8
exp[i]iC(p).
394
GEOMETRY
OF QUANTUM
THEORY
From the uniqueness of Q and from formula (236) we see that Q vanishes for the complement of the set {p :
= exp[-;]/(p).
This implies that the measure class of the projection valued measure WQW'1 is that of Lebesgue measure, which is a contradiction, as we have already shown that this measure class is supported by the set {P :
mz : r, r' -> exp ^[-
where r = (A,iy,v,u),
r' = (/&',7/,v',u'),
and T
# 0
is a real number. Given raT, we can construct the associated extension of the inhomogeneous group in the usual way. The elements of this extension are pairs (r; £) with |£| = 1, and the group operation is given by (r; £)(r'; £') = K ; M'mt(r,r')). We denote this group by G1. Now mT(r,r') = l when v = v' = 0 and so
(MAii; £)(l,i/,0,u'; £') = ( l ^ + VAu'+u; «'),
RELATIVISTIC
FREE PARTICLES
395
so t h a t the set of all elements of the form (1,77,0,11; £) forms an abelian group. We denote this group by Ax and, when convenient, its typical element by (rj,u; £). Further, ml(r,r') = \ when u = u / = 0 and 77'= 0 so that (A,0,v,0; l)(A',0,v',0; 1) = (M',0,8(A)v' + v,0; l), showing t h a t the set of all elements (h,0,v,0; 1) forms a subgroup of Gx. We write Hx for this subgroup and, when convenient, denote its typical element by (h,v; 1). We have: (240)
(77,11; £ e x p - ^ < u , v > \\(h,v;
1) = (A,^,v,u; £)•
Gx is thus a semidirect product of A1 and Hx. The action of Hx on Ax via inner automorphisms is given by (241)
(A,v; 1) : (v,u; £) -> M W u + i ^ J ' ) ,
where (242)
r
=
C exp
- » I [2
We are interested in describing those irreducible representations of Gx which map the elements (1,0,0,0; £) into £ - 1 times the identity, for these are precisely the representations which give rise to the mx-representations of the inhomogeneous Galilean group (theorem 7.16). Fortunately, GT is itself a semidirect product of Ax and Hx, and so we can determine its irreducible representations using the theory of Section 8, Chapter VI. /\ To this end, we must determine the dual group A1 of Ax and classify the /\ /\ Hx orbits in A1. A typical element of Ax will be written as {p0,Y*; n) where p0 e R1, p e R3, and n is an integer; we shall associate with it the character (17,11; 0 -> £n e x p [ * ( ^ 0 +
(h,v; 1) : (p0,f9n)
->
(p0',v',n),
where /nAA^ (244)
Vo = ^ o - < v 5 S ( % > - i ^ r < v , v > , p' = 8(h)\) + nTV.
The reader m a y observe the difference between this action and the action (232).
396
GEOMETRY
OF QUANTUM
THEORY
I t is quite simple to determine the orbits and stability groups associated with the action (244). A brief calculation shows t h a t (245)
0
'.
I n other words, the function (p0,j>; n) - »
ZUtP = {(p0,v; n) :
0
=
p};
here n is an integer, p a real number, n and p are arbitrary except when n = 0, for which case p > 0 . The ZUtP are invariant under Hx and *
= U Zn.pn, p
When n = 0, (244) reduces to (232) a n d we find t h a t the sets Z0tP (p>0) are orbits. The set Z00 is not a n orbit; Z00 = {(p0,0; 0) i ^ e R 1 } , and each point of it is invariant under Hx. For n^O, and arbitrary p, Zn>p is an orbit. I n fact, the point (247)
an,p =
((2nr)-'p,0;n)
lies in ZntP, and one computes quickly from (243) t h a t t h e element ( l , ( w r ) - 1 p ; 1) of Hx sends an%p to the point ((2nr)~1(p —
RELATIVISTIC
FREE
PARTICLES
397
raT-representations of G are precisely those that are associated with the orbits Z_ltP, p arbitrary. We shall now describe these representations. Exactly as in the case of the Euclidean group (cf. Chapter VIII), we find that, if s : h~> s(h) is an irreducible representation of the unitary group K* in a (finite dimensional) Hilbert space Jf^ the map (h,v; 1), z-> s(h) (z e Z_lp) is a strict cocycle which defines the representation (h,0; I)-> s(h) of the stability group at a_lfP. If we denote by &? the Hilbert space (248)
je=&*(V\jr,dv), x,s p
then the representation U ' of Gx, associated with Z _x p and «s, is given by (249)
(Ul-rf)(V)
=
Vs(h)f(8(h)-i(v+Tv)),
where r = (^,7y,v,w) and (250)
Q>
=
(JeWPiD/a.) - 1 e x p J J" L p +T-y \ + JL
for this we have only to use the map £ identifying Z_x p with P3, the equation (240) which displays explicitly the semidirect product structure of Gx, the equation (244) which gives the adjoint action of H%, and substitute in the general formula (132) of Chapter VI. The map r -> U\\\p will then be the irreducible raT representation of the inhomogeneous Galilean group associated with UT,S'P. We now observe that the number p enters (250) only in the factor e~ipvl2x and hence may be dropped when we describe the corresponding projective representation of the inhomogeneous Galilean group. We shall write Vx,s for this irreducible, projective representation. It acts on Jj? and (251)
(JT/HP)
= ^ ( W ( 3 ( A ) - 1 ( p + rv))
with (252)
Q = exp»[/u,p + gv\+^:
In this form, the multiplier for Vx,s will be similar to raT. There remains the question of physical equivalence among these representations. We shall show first that Vx's and V~x,s are anti-unitarily equivalent. Consider any anti-unitary isomorphism J of the Hilbert space J f onto itself and define 0 by (253)
(0/)(p)=«//(-p).
398
GEOMETRY
OF QUANTUM 2
THEORY
3
0 is a n anti-unitary isomorphism of J^ (P ,JT,(Zp) onto itself. A quick calculation shows t h a t ®yx.s®-i
y-i,8'9
=
where 6*' is the representation h —>Js(h)J~1 of K*. Since dim(s') = dim(<s), s'~ s, so t h a t V~x>s'~ V~x,s. Herice F T , s a n d V~~x-S are physically equivalent. Suppose conversely t h a t F T i ,s i and FT2-S2 are physically equivalent. Let us first assume t h a t the equivalence is effected by a unitary operator. Restriction to the subgroup of Euclidean motions (17 = 0, v = 0) implies t h a t the corresponding representations are unitarily equivalent and enables us t o conclude, using lemma 9.13, t h a t S1^LS2. We m a y (and do), therefore, assume t h a t s1=s2 = s and t h a t F V S and F V S both act in J f = j£? 2 (P 3 ,Jf,dp). I t is then immediately obvious t h a t the multipliers of t h e two representations F V S and Vx*,s are similar. But we have observed t h a t the multiplier of VXyS is similar to mx. So our observation implies t h a t raT and ml2 are similar. The determination of the multiplier group of the Galilean group carried out in Chapter V I I now enables us to conclude t h a t rx and r 2 must be equal. Thus F V s i and FT2'S2 cannot be unitarily physically equivalent unless T1 = T2 and s1~s2. If, on the other hand, they are antiunitarily physically equivalent, then F _ T i , s i and FV S 2 must be unitarily physically equivalent, and we could deduce t h a t r 2 = — r 1? sx ~ s 2 - ^ n other words, the projective representations Vx's ( T > 0 ) are, for distinct r and inequivalent s, physically inequivalent, and any irreducible projective representation of the inhomogeneous Galilean group with nonexact multiplier is physically equivalent to one of these. The action of Vx,s can be described in another and perhaps more usual form by passing to Fourier-Plancherel transforms. Let (cf. (209))
be the Fourier-Plancherel isomorphism of i f 2 ( P 3 X ^ p ) onto J^ 2 (R 3 ,jr,^x) and let
Then (254) where (255)
Vrz>s = exp( - ^ -
a2
a2 2
dx±
2
dx2
v\r,
e2 dx32
and (256)
(IY)(x) = exp i T [i + < x - u ,
v}]s(h)f(8(h)-Hx-u)).
RELATIVISTIC
FREE
PARTICLES
399
V is self-adjoint with the usual domain. The expression (254) shows that the energy of the particle is the operator (257)
Ez = - l v .
Notice that we have thus proved the formula (77) of Chapter VIII. Further, under the (covering group of the) Euclidean group, the states transform according to the law described by (258)
fe,o,«>/)W
= s(h)f(8(h)-i(x-u)).
A look at (258) is enough to conclude that the system corresponding to Vx,s is localizable, and the projection valued measure which leads to the position operators is the usual one, E -> PE, where (259)
PE-I-^XEI T,S
In other words, V describes the Schrodinger particle of mass r and spinj where dim(,s) = 2 j + l . We have thus obtained the following theorem: Theorem 9.18. The irreducible representations of the covering group of the inhomogeneous Galilean group G cannot represent nontrivial particles and the corresponding systems are not localizable. On the other hand, for each T > 0 and each half-integer j = 0, J, 1,- • • there is an irreducible projective representation with a nonexact multiplier of the inhomogeneous Galilean group, say Vx,j and {with s = Dj) its action is given by (251) and (252). Every irreducible projective representation of G with a nonexact multiplier is physically equivalent to exactly one of the VT,j. The particle corresponding to Vz,j is the Schrodinger particle of mass r and spinj, and its energy operator is — (l/2r)V where V is the Laplacian (255). The analysis of the ordinary representations of the inhomogeneous Galilean group was first carried out by Inonu and Wigner [1], [2]. For the projective representations, the reader may consult the papers of Bargmann [1] and Wightman [1] (cf. also Pauli [1]). NOTES ON CHAPTER I X The mathematical and physical aspects of the theory of relativisitically invariant wave equation have spanned a vast literature; see, for instance, Invariant Wave Equations, Lecture Notes in Physics, No. 73, SpringerVerlag, Berlin, edited by G. Velo and A. S. Wightman, 1978. For an account of the early literature on this subject see E.M.Corson, Introduction to Tensors, Spinors and Relativistic Wave Equations, Chelsea reprint of the 1953 edition); see also Chapter IV of M. A. Nafmark, Linear Representations of the Lorentz Group (in Russian), Moscow, 1958.
BIBLIOGRAPHY B A E R , R.
[1] Linear Algebra and Projective Geometry, Academic Press Inc., New York, 1952. BARGMANN, V.
[1] On unitary ray representations of continuous groups, Annals of Mathematics, 59 (1954), pp. 1-46. [2] Note on Wigner's theorem on symmetry operations, Journal of Mathematical Physics, 5 (1964), pp. 862-868. BARGMANN, V., AND W I G N E R , E . P .
[1] Group theoretical discussion of relativistic wave equations, Proceedings of the National Academy of Sciences, U.S.A., 34 (1948), pp. 211-223. BlRKHOFF, G.
[1] Lattice Theory, American Mathematical Society Colloquium Publications, Vol. 25, 1960. BlRKHOFF, G., AND VON NEUMANN, J .
[1] The logic of quantum mechanics, Annals of Mathematics, 37 (1936), p p . 823-843. BLACKWELL, D.
[1] On a class of probability spaces, Proceedings of the Third Berkeley Symposium on Mathematical Statistics and Probability, Vol. I I (1956), pp. 1-6. BOCHNER, S., AND CHANDRASEKHARAN, K .
[1] Fourier Transforms, Princeton University Press, Princeton, 1949. BOURBAKI, N.
[1] Elements de Mathematique, I, Livre II, Algebre, Chs. 4 and 5, Hermann, Paris, 1959. [2] Elements de Mathematique, I, Livre V, Espaces vectoriels topologiques, Chs. 1 and 2, Hermann, Paris, 1953. [3] Elements de Mathematique, I, Livre III, Topologie Generate, Chs. 1 and 2, Hermann, Paris, 1961. B R A U E R , R., AND W E Y L , H .
[1] Spinors in n dimensions, 425-449.
American Journal of Mathematics, 57 (1935), p p .
CALABI, L.
[1] Sur les extensions des groupes topologiques, Annali de Matematica P u r a ed Applicata, 32 (1951), pp. 295-370. CARTAN, E .
[1] Lecons sur la theorie des spineurs, I, I I , Hermann, Paris, 1938. CHEVALLEY, C.
[1] [2] [3] [4]
Theory of Lie Groups, Princeton University Press, Princeton, 1946. Theorie des Groupes de Lie, Tome I I , Hermann, Paris, 1951. Theorie des Groupes de Lie, Tome I I I , Hermann, Paris, 1955. The Algebraic Theory of Spinors, Columbia University Press, New York, 1954.
CHEVALLEY, C , AND E I L E N B E R G , S.
[1] Cohomology theory of Lie groups and Lie algebras, Transactions of the American Mathematical Society, 63 (1948), pp. 85-124. 400
BIBLIOGRAPHY
401
D I R A C , P . A. M.
[1] The Principles of Quantum Mechanics, Oxford University Press, London, 1958. [2] The quantum theory of the electron, Proceedings of the Royal Society (London), A 117 (1928), pp. 610-624. DlXMIER, J .
[1] Les algebres d'operateurs 1957.
dans Vespace Hilbertien,
Gauthier-Villars, Paris,
DlJNFORD, N . , AND SCHWARTZ, J .
[1] Linear Operators, P a r t 1, Interscience Publishers, Inc., New York, 1958. [2] Linear Operators, P a r t 2, Interscience Publishers, Inc., New York, 1963. ElLENBERG, S.
[1] Topological methods in abstract algebra. Gohomology theory of groups, Bulletin of the American Mathematical Society, 55 (1949), pp. 3-37. F E D E R E R H., AND MORSE, A. P .
[1] Some properties of measurable functions, Bulletin of the American Mathematical Society, 49 (1943), pp. 270-277. FEYNMAN, R.
P.
[1] Concept of probability in quantum mechanics, Proceedings of the Second Berkeley Symposium on Mathematical Statistics and Probability, Berkeley, (1951), pp. 533-541. [2] Space-time approach to non-relativistic quantum mechanics, Reviews of Modern Physics, 20 (1948), pp. 367-387. F I N K E L S T E I N , D., J A U C H , J . M., SCHMINOVICH, S., AND SPEISER, D.
[1] Foundations of quaternion quantum Physics, 3 (1962), pp. 207-220.
mechanics,
Journal of Mathematical
G E L ' F A N D , I. M.
[1] On one-parametrical groups of operators in a normed space, Doklady Akademii Nauk, SSSR, 25 (1939), pp. 713-718. G E L ' F A N D , I. M., AND NAIMARK, M. A.
[1] Unitary representations of the group of linear transformations of the straight line, Doklady Akademii Nauk, SSSR, 55 (1947), pp. 567-570. GLEASON, A. M.
[1] Measures on the closed subspaces of a Hilbert space, Journal of Mathematics and Mechanics, 6 (1957), pp. 885-894. GUDDER, S. P .
[1] Spectral methods for a generalized probability theory, Transactions of the American Mathematical Society, 119 (1965), pp. 428-442. ILALMOS, P .
R.
[1] Measure Theory, D. van Nostrand Company, Princeton, 1950. [2] Introduction to the Theory of Hilbert Space and Spectral Multiplicity, Publishing Co., New York, 1957.
Chelsea
H E I S E N B E R G , W.
[1] The Physical York, 1930.
Principles
of Quantum Theory, Dover Publications, Inc., New
HELGASON, S.
[1] Differential Geometry and Symmetric Spaces, Academic Press Inc., New York, 1962. HELSON, H.
[1] Lectures on Invariant Subspaces, Academic Press Inc., New York, 1964. H E L S O N , H., AND LOWDENSLAGER, D.
[1] Invariant subspaces, Proceedings of the International Symposium on Linear Spaces, pp. 251-262 (Jerusalem, 1960), Pergamon Press, Inc., Oxford, 1961.
GEOMETRY
402
OF QUANTUM
THEORY
H E Y T I N G , A.
[1] Axiomatic Projective Geometry, Interscience Publishers, Inc., New York; P. Noordhoff N. V., Groningen, North Holland Publishing Company, Amsterdam, 1963. HOCHSCHILD, G.
[1] Group extensions of Lie Groups, I, Annals of Mathematics, 54 (1951), p p . 96-109. [2] Group extensions of Lie Groups, I I , Annals of Mathematics, 54 (1951), pp. 537-551. I N O N U , E., and W I G N E R , E.
P.
[1] Representations of the Galilei group, II Nuovo Cimento, 9 (1952), pp. 705-718. [2] On the contraction of groups and their representations, Proceedings of the National Academy of Sciences, U.S.A., 39 (1953), pp. 510-524. JACOBSON, N.
[1] Lectures in Abstract Algebra, Vol. I I , D. Van Nostrand Company, Inc., Princeton, 1953. [2] Lie Algebras, Interscience Publishers, Inc., New York, 1962. J O R D A N , P., VON N E U M A N N , J., AND W I G N E R , E.
P.
[1] On an algebraic generalization of the quantum mechanical formalism, Mathematics, 35 (1934), pp. 29-64. K A K U T A N I , S., AND MACKEY, G.
Annals of
W.
[1] Ring and lattice characterizations of complex Hilbert space, Bulletin of the American Mathematical Society, 52 (1946), pp. 727-733. K A P L A N S K Y , I.
[1] Any orthocomplemented complete modular lattice is a continuous geometry, Annals of Mathematics, 61 (1955), pp. 524-541. [2] Infinite Abelian Groups, University of Michigan Press, Ann Arbor, 1954. K E L L E Y , J.
L.
[1] General Topology, D. Van Nostrand Company, Princeton, 1955. K L E P P N E R , A.
[1] Multipliers KOLMOGOROV, A.
on abelian groups, Mathematische Annalen, 158 (1965), pp. 11—34. N.
[1] Zur Begrilndung der Projectiven Geometrie, Annals of Mathematics, 33 (1932), pp. 175-176. KOOPMAN, B.
O.
[1] Hamilton systems and transformations in Hilbert space, Proceedings of the National Academy of Sciences, U.S.A., 17 (1931), pp. 315-318. KURATOWSKI, C.
[1] Topologie, Vol. I, Warsaw, 1948. [2] Topologie, Vol. I I , Warsaw, 1950. LANDAU, L. D., AND LIFCHITZ, E.
[1] Quantum Mechanics, 1963. LOOMIS, L.
M.
non-relativistic
Theory, Gos. Izd. Phys. Mat. Moscow,
H.
[1] On the representation of a-complete Boolean algebras, Bulletin of the American Mathematical Society, 53 (1947), pp. 757-760. [2] An Introduction to Abstract Harmonic Analysis, D. Van Nostrand Company, Inc., Princeton, 1953. [3] The lattice theoretic background of the dimension theory of operator algebras, Memoirs of the American Mathematical Society, 18 (1955), pp. 1—36. MACKEY, G.
W.
[1] The Mathematical Foundations of Quantum Mechanics, W. A. Benjamin, Inc., New York, 1963. [2] Imprimitivity for representations of locally compact groups, Proceedings of the National Academy of Sciences, U.S.A., 35 (1949), pp. 537-545.
BIBLIOGRAPHY
403
[3] Induced representations of locally compact groups, I, Annals of Mathematics, 55 (1952), p p . 101-139. [4] Induced representations of locally compact groups, I I , Annals of Mathematics, 58 (1953), pp. 193-221. [5] Borel structure in groups and their duals, Transactions of the American Mathematical Society, 85 (1957), pp. 134-165. [6] Unitary representations of group extensions, I, Acta Mathematica, 99 (1958), pp. 265-311. [7] Infinite dimensional group representations, Bulletin of the American Mathematical Society, 69 (1963), pp. 628-686. [8] Les ensembles Boreliens et les extensions des groupes, Journal de Mathematiques Pures et Appliquees, 36 (1957), pp. 171-178. [9] A theorem of Stone and von Neumann, Duke Mathematical Journal, 16 (1949), pp. 313-326. MALCEV, A.
[1] On the simple connectedness of invariant subgroups of Lie groups, Doklady Akademii Nauk, SSSR, 34 (1942), pp. 10-13. MAUTNER, F . I.
[1] Unitary representations of locally compact groups. I, Annals of Mathematics, 51 (1950), pp. 1-25. [2] Unitary representations of locally compact groups. I I , Annals of Mathematics, 52 (1950), pp. 528-556. MONTGOMERY, D., AND Z I P P I N , L.
[1] Topological transformation 1955.
groups, Interscience Publishers, Inc., New York,
MOORE, C. C.
[1] Extensions and low dimensional cohomology theory of locally compact groups, I, Transactions of the American Mathematical Society, 113 (1964), pp. 40-63. MOYAL, J . E.
[1] Quantum Mechanics as a statistical theory, Proceedings of the Cambridge Philosophical Society, 45 (1949), p p . 99-124. MURRAY, F . J., AND VON NEUMANN, J .
[1] On rings of operators, Annals of Mathematics, 37 (1936), pp. 116-229. [2] On rings of operators, I I , Transactions of the American Mathematical Society, 41 (1937), pp. 208-248. [3] On rings of operators, IV, Annals of Mathematics, 44 (1943), pp. 716-808. NAKANO, H .
[1] Unitarinvariante hypermaximale 42 (1941), pp. 657-664.
normale operatoren, Annals of Mathematics,
N E W T O N , T. D., AND W I G N E R , E. P .
[1] Localized states for elementary systems, Reviews of Modern Physics, 21 (1949), pp. 400-406. P A U L I , W.
[1] Die allgemeinen Prinzipen der Wellenmechanik, Handbuch der Physik, 2 Aufl. Band 24, 1 Teil (Reprint, Edwards Brothers, Inc., Ann Arbor, Michigan), 1950. PlRON, C.
[1] Axiomatique Quantique, These, Universite de Lausanne, Faculte des Sciences, Bale, Imprimerie Birkhauser S. A. (1964), pp. 439-468. PONTRJAGIN, L . S.
[1] Topological Groups, Princeton University Press, Princeton, 1939. R I E S Z , F., AND N A G Y , B. SZ.
[1] Lecons d'analyse Fonctiohnelle, Akademiai Kiado, Budapest, 1952. SCHWARTZ, L.
[1] Theorie des distributions, Tome I, Hermann, Paris, 1957.
GEOMETRY
404
OF QUANTUM
THEORY
[2] Theorie des distributions, Tome I I , Hermann, Paris, 1959. [3] Application of distributions to the study of elementary particles in relativistic quantum mechanics, Technical Report No. 7, Department of Mathematics, University of California, Berkeley, 1961, pp. 1-207. SCHWEBER, S.
[1] An introduction to relativistic quantum field theory, Row, Peterson and Company, Evanston and New York, 1961. SEGAL, I.
E.
[1] Postulates for general Quantum Mechanics, Annals of Mathematics, 48 (1947), pp. 930-948. [2] Quantization of nonlinear systems, Journal of Mathematical Physics, 1 (1960), pp. 468-488. [3] Mathematical problems of relativistic physics, Lectures in Applied Mathematics, American Mathematical Society, Providence, R.I., 1963. SEIDENBERG, A.
[1] Lectures in Projective Geometry, D. Van Nostrand Company, Inc., Princeton, 1962. SlKORSKI, R .
[1] Boolean Algebras, Springer, Berlin, 1964. STERNBERG, S.
[1] Lectures on Differential Geometry, Prentice-Hall, Inc., Englewood Cliffs, N.J., 1964. STONE, M.
H.
[1] The theory of representation for Boolean algebras, Transactions of the American Mathematical Society, 40 (1936), pp. 37-111. [2] Linear transformations in Hilbert space and their applications to analysis, American Mathematical Society Colloquium Publications, Vol. 15, 1951. [3] Linear transformations in Hilbert space. I I I . Operational methods and group theory, Proceedings of the National Academy of Sciences, U.S.A., 16 (1930) pp. 172-175. V A N H O V E , L.
[1] Sur certaines representations unitaires d'un groupe infini de transformations, Academie Royale de Belgique Classe des Sciences, Memoires Collection in 8°, 29, No. 6 (1951), pp. 1-102. VARADARAJAN, V.
S.
[1] Probability in physics and a theorem on simultaneous observability, Communications in Pure and Applied Mathematics, 15 (1962), pp. 189-217; correction, loc. cit., 18 (1965). [2] Groups of automorphisms of Borel spaces, Transactions of the American Mathematical Society, 109 (1963), pp. 191-220. [3] Measures on topological spaces, Matematiceskii Sbornik, 55 (97) (1961), pp. 35-100. [4] Special topics in probability, mimeographed notes of lectures delivered at the Courant Institute of Mathematical Sciences, New York University, 1962. [5] Global theory of representations of locally compact groups, (Notes by Alan Solomon)—mimeographed notes of lectures delivered at the Courant Institute of Mathematical Sciences, New York University, 1961-62. V O N NEUMANN, J.
[1] Mathematical Foundations of Quantum Mechanics, Princeton University Press, Princeton, 1955 (translated by Robert T. Beyer from Mathematische Grundlagen der Quantenmechanik, Springer, Berlin, 1932). [2] On rings of operators. Reduction theory, Annals of Mathematics, 50 (1949), pp. 401-485. [3] Die Eindeutigkeit der Schrodingerschen Operatoren, Mathematische Annalen, 104 (1931), pp. 570-578.
BIBLIOGRAPHY
405
[4] On an algebraic generalisation of the quantum mechanical formalism (part I), Rec. Math. Matematiceskii Sbornik, N.S„ 1 (1936), pp. 415-484. [5] Thermodynamik quantenmechanischer Gesamtheiten, Gottinger Nachrichten (1927), pp. 273-291. [6] Beweis des Ergodensatzes und des H-Theorems in der neuen Mechanik, Zeitschrift fur Physik, 57 (1929), pp. 30-70. W E I L , A.
[1] L* integration dans les groupes topologiques et ses applications, Hermann, Paris, 1938. W E I S S , E., AND Z I E R L E R , N.
[1] Locally compact division pp. 369-371.
rings, Pacific Journal of Mathematics, 8 (1958),
W E Y L , H.
[1] The Theory of Groups and Quantum Mechanics, Dover Publications, Inc., New York, 1931 (translated by H. P . Robertson from Gruppentheorie und Quantenmechanik). [2] The Classical Groups, Princeton University Press, Princeton, 1946. [3] Theorie der Darstellung kontinuerlicher halb-einfacher Gruppen durch lineare Transformationen, Teil I, Mathematische Zeitschrift, 23 (1925), pp. 271-309. [4] , Teil I I , Mathematische Zeitschrift, 24 (1926), pp. 328-376. [5] , Teil I I I , Mathematische Zeitschrift, 24 (1926), pp. 377-395. [6] Space, Time, Matter, Dover Publications, Inc., 1950 (translated by Henry L. Brose from Raum, Zeit, Materie, 1921). W I C K , G. C , WIGHTMAN, A. S., AND W I G N E R , E. P .
[1] The intrinsic parity of elementary particles, The Physical Review, 88 (1952), pp. 101-105. WlGHTMAN, A . S.
[1] On the localizability of quantum mechanical systems, Reviews of Modern Physics, 34 (1962), pp. 845-872. [2] Relativistic Invariance and Quantum Mechanics, Supplemento al Vol. X I V , serie X , del Nuovo Cimento (1959), pp. 81-94 (Notes by A. Barut). W I G N E R , E. P .
[1] Unitary representations of the inhomogeneous Lorentz group, Annals of Mathematics, 40 (1939), p p . 149-204. [2] On the quantum correction for thermodynamic equilibrium, The Physical Review, 40 (1932), pp. 749-759. [3] Group theory and its applications to the quantum mechanics of atomic spectra, Academic Press Inc., New York, 1959. [4] Vber die Operation der Zeitumkehr in der Quantenmechanik, Gottinger Nachrichten Math-Physik Kl (1931), pp. 546-559. [5] Relativistische Wellengleichungen, Zeitschrift fur Physik, 124 (1947), pp. 665-684. [6] Relativistic invariance and quantum phenomena, Reviews of Modern Physics, 29 (1957), pp. 255-268. ZlERLER, N .
[1] Axioms for non-relativistic quantum mechanics, Pacific Journal of Mathematics, 11 (1961), pp. 1151-1169. [2] On the lattice of closed subspaces of Hilbert space (forthcoming).
INDEX
Borel group, 157 separable, 157 standard, 157 Borel isomorphism, 148 Borel mapping, 148 Borel section, 164 regular, 164 Borel space, 148 separable, 149 standard, 150 sub space of, 149 Borel structure, 148 natural (of a topological space), 149 Bose-Einstein statistics, 139, 317 Bosons, 139, 318 Bounded set (in a topological group), 196
Absolute continuity (of measures), 151 Affine coordinate system, 301 Affine space, 160, 300 automorphism of, 300 translation in, 300 vector space associated with, 300 Angular momentum around an axis, 310 orbital, 311, 364 spin, 311, 364 Annihilation operator, 317 Annihilator (of a linear manifold), 23 Anti-automorphism (of a division ring), 23
Banach space, over a division ring, 73 Basis (in a generalized geometry), 33 Bell's inequality (for local hidden variable models), 131 0-bilinear form, 23 definite, 26 right/left nonsingularity of, 23 symmetric, 24 Boolean algebra, 8 Boolean a-algebra, 8 Boolean subalgebra of a logic, 55 Boolean sub a-algebra of a logic, 55 Borel automorphism, 148 Borel G-space, 159 G-subspace of, 159 standard, 159
C*-algebra, 134 Center (of a complemented modular lattice), 19 Center (of a logic), 64 atoms of, 64 continuity of, 64 Chain (in a lattice), 18 length of, 18 Clifford algebra, 350 concrete, 348 Closed form, 273 Coboundary, 175 strict, 175 Cocycle relative to a measure class, 174 strict, 174 407
408
INDEX
Coherent subspace of states, 114 Cohomologous cocycles, 174 Cohomology class, 175 strict, 175 Cohomology group, 175 Collinearity (of points in a lattice), 20 Commutation rules, 5, 294 of Heisenberg, 294 Commuting ring (of a representation), 203 Complement (of an element in a lattice), 8 Complementarity, 302 Completion (of a a-algebra with respect to a measure), 151 Configuration observables, 5, 295 Configuration space, 4, 293 Continuous action (of a lcsc group), 159 Continuous geometries with a transition probability, 71 Continuous section, 164 Coordinates, in a generalized geometry, 28 affine, 34 homogeneous, 37 Coordinatization of a geometry, 20 of a generalized geometry, 38 of a complemented modular lattice, 40 Creation operator, 317 Cross-section lemma (of von Neumann), 154
Degrees of freedom (of a mechanical system), 1 Desarguesian generalized geometry, 29 Desarguesian planes, 28 Dimension (of an element), 19, 29 Dimension function (of a lattice), 19 Dirac equation (of the electron), 369 Dirac 7-matrices, 355 Dispersion free ensembles, 123
Dispersion free states, 124 Dual of a division ring, 22 Dual ideal, 9 Duality (between geometries), 23 Dynamical group of a classical system, 6 of a general quantum system, 68 Dynamical states, 292, 327
Einstein-Podolsky-Rosen paradox, 129, 137 Energy, 290 Entropy macroscopic, 142 of a quantum mechanical ensemble, 142 of a quantum mechanical state, 141 Ergodic measure class, 182 Euclidean affine space, 303 motion of, 304 Euclidean coordinate system, 304 Exact form, 273 Extension (of one lcsc group by another), 251 central, 251 equivalence of, 251 Extension, central, of one Lie algebra by another, 267 associated with a group extension, 268 covering of one by another, 269 equivalence of, 268 Exterior algebra over a Hilbert space, 139, 319
Fermi-Dirac statistics, 139, 317 Fermions, 139, 319 Frame (of a generalized geometry) at a point, 33 Frame function, 85 regular, 85 weight of, 85 Free action of a group, 183
INDEX
Free particle, 69 classical, 322 quantum, 330, 399
G-Hilbert space bundle, 234 section of, 234 square integrable section of, 234 G-homomorphism, 159 G-isomorphism, 159 G-space, 158 homogeneous, 160 locally compact, 159 transitive, 160 Galilean group, 283, 325 projective representation of, 394 Galilean relativity, 325, 391 Galilean transformation, 283 Gauge condition, 371 Generating sequence (for a Borel structure), 149 Geometry, 19 generalized, 29 projective, 18 Gleason's theorem, 97 generalizations of, to von Neumann algebras, 146
Haar measure, 157 Hamiltonian, 1, 3, 290 Heisenberg picture, 293 Heisenberg states, 293, 327 Hidden variables in Quantum Mechanics, 122, 131 Hilbert space over a division ring, 26 Hilbertian pairs, 115 Homomorphism between Boolean algebras, 9 defined by a cocycle at a point, 176 equivalence of, 176 a-homomorphism between Boolean a-algebras, 10
409
Independence (of points), 19, 32 Inertial coordinate system, 323 Inertial observer, 323 Infinitesimal operators (for the free particle representations), 364 Infinity element at, 33 points at, lines at, etc., 33 Inhomogeneous group, associated with a group and a representation of it, 278 Integral of motion, 292 Inversions, group of, 337 Involutive anti-automorphism, 25 Isomorphism (between geometries), 21 linearity of, 21
Jordan algebras, 133 formally real, 133
Kernel of a a-homomorphism, 11 Koopman system of imprimitivity, 208
lcsc group, 156 Lattice, 8 complemented, 8 distributive, 8 irreducible, 19 lines, planes, etc., of, 20 modular, 18 points of, 19 rank of, 19 Linear momentum in a given direction, 301 Localizable systems, 383 Locally bounded map (from one topological space to another), 181 Locally bounded topological group, 196 Logic, 42 associated with a division ring, 119
410
INDEX
Logic {cont'd.) associated with a von Neumann algebra, 112 of a classical system, 6 injection of, 43 irreducible, 64 isomorphism of, 43 projective, 114 standard, 44, 74, 80 sublogic of, 43 Lorentz condition, 371 Lorentz group, 326 Lorentz-Hermite form, 363
Maxwell equations, 371 Measure, 150 Borel, 150 finite, 150 a-finite, 150 invariant, 151, 159 quasi-invariant, 159 regular, 150 standard, 150 Measure class, 151 bi-in variant, 191 corresponding to the multiplicity n, 219 ergodic, 182 invariant, 151 living on an orbit, 223 of projection valued measure, 218, 219 transitive, 182 Measure on a logic, 50 Minkowskian form, 326 Modular function (of a lcsc group), 157 Momentum angular, 4 corresponding to symmetries, 4 linear, 4 Momentum observables, 5, 295 Momentum space, 339 representations, 329 Momentum vector, 1 Motions of a quantum mechanical system, 292
Multiplier (/C-multiplier), 247 exact, 248 locally continuous, 255 similar, 248 Multiplier group (^-multiplier group), 248 Multiplier of a projective representation, 248 Mutually singular measures, 218
Observables associated with a Boolean aalgebra, 12 associated with a logic, 46 associated with a standard logic, 82 bounded, 47 of a classical system, 2, 3 central, 64 discrete, 47 expected value of, 52 probability distribution of, 52, 63 spectrum of, 47 value of, 47 variance of, 52 Operators of trace class, 83 Orbit of a point, 159 Orthocomplementation in the geometry of a vector space, 26 in a lattice, 42 Orthogonal complement (of an element), 43
Partial Boolean algebra, 127 Partially ordered set, 8 null element of, 8 unit element of, 8 Phase space, 2 Physical states of von Neumann algebras, 125, 146 Polarity, 24 isotropic, 24 Position operators for nonrelativistic particles, 301, 302
INDEX
for relativistic particles, 377, 387, 390 Product of Borel spaces, 149 Projection valued measures, 80 corresponding to a representation of an abelian group, 238 cyclic vector for, 218 homogeneous, 218 measure class of, 218, 219 multiplicity of, 218 Projective group (of a Hilbert space), 244 Projective representation, 248 multiplier of, 248 Projectivity canonical fixed points of, 39 general, 39 special, 39 Pure state, 53
Quantization, 315 Quantum probability, 145 conditional, 145 Quasi invariant measure, 159 Quasi-probability distribution, 104 Quaternions, 21 conjugacy of, 22 conjugation of, 22 norm of, 21 unit, 22 Question observables, 50 Radon-Nikodym derivative, 151 Range of an observable, 60 Regular ring, 40 Representation absolutely irreducible, 278 admissible, 278 equivalence of, 202 of a group in the group of automorphisms of the convex set of states, 69 irreducible unitary, 202 physical equivalence of, 305 unitary, 201
411
Schrodinger equation (abstract), 291 Schrodinger particle, 399 Schrodinger picture, 293 Schrodinger representation (of commutation rules), 299 Second Quantization, 316, 321 Semidirect product, 236 Semilinear transformation, 20 Separable cr-algebras, 15, 53 Separating sequence for a Borel structure, 149 Simultaneous observability, 55 verifiability, 55 Smooth orbit structure, 228 Space inversion, 337 Space-time inversion, 337 Space-time relativity group, 323 Spectrum of a relativistic system, 329 Spin (of a free relativistic particle), 348 Spin matrices (of Pauli), 307 Spin observables, 311 Spin representation, 353 Spinor field, 356 Spinors, 353 Stability subgroup at a point, 159 State of a classical system, 1 of a logic, 52 of a standard logic, 85 State function (of a logic), 49 States, coherent subspace of, 114 Statistical interpretation of Born, 71 Strict coboundary, 175 Strict cocycle, 174 Strict cohomology class, 175 Strictly cohomologous strict cocycles, 175 Stone space, 9 Stone topology, 10 Strong topology (for the unitary group of a Hilbert space), 193 Subrepresentation, 204 Superposition of states classical, 53 of a logic, 53 Superposition principle, 54, 98
412
Superselection rules, 114 Symmetric algebra over a Hilbert space, 139, 318 Symmetry of a Hilbert space, 104 physical, 108 9-unitary, 104 Systems of imprimitivity for a group, 203 associated with a cocycle and a measure class, 217 direct sum of, 203 equivalence of, 203 induced by a representation of a subgroup, 223 irreducible, 203 living on an orbit, 223 physical equivalence of, 305 transitive, 223
INDEX
Time inversion, 325, 337 Total energy, 329 Transition probability (between states), 107 Uncertainty principle, 66, 71 Unimodular group, 157 Unitary transformation, 28 Vacuum, 317, 347 Velocity observables, 391 von Neumann algebras, 112 von Neumann operators, 85 Wave equations, 365, 399 of particles, 369, 370, 371 Wave function, 302 Weil topology, 195