+ All Categories
Home > Documents > Analytic Vectors

Analytic Vectors

Date post: 15-Dec-2016
Category:
Upload: edward-nelson
View: 224 times
Download: 0 times
Share this document with a friend
45
Annals of Mathematics Analytic Vectors Author(s): Edward Nelson Source: Annals of Mathematics, Second Series, Vol. 70, No. 3 (Nov., 1959), pp. 572-615 Published by: Annals of Mathematics Stable URL: http://www.jstor.org/stable/1970331 . Accessed: 06/09/2013 23:19 Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at . http://www.jstor.org/page/info/about/policies/terms.jsp . JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact [email protected]. . Annals of Mathematics is collaborating with JSTOR to digitize, preserve and extend access to Annals of Mathematics. http://www.jstor.org This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PM All use subject to JSTOR Terms and Conditions
Transcript

Annals of Mathematics

Analytic VectorsAuthor(s): Edward NelsonSource: Annals of Mathematics, Second Series, Vol. 70, No. 3 (Nov., 1959), pp. 572-615Published by: Annals of MathematicsStable URL: http://www.jstor.org/stable/1970331 .

Accessed: 06/09/2013 23:19

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .http://www.jstor.org/page/info/about/policies/terms.jsp

.JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range ofcontent in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new formsof scholarship. For more information about JSTOR, please contact [email protected].

.

Annals of Mathematics is collaborating with JSTOR to digitize, preserve and extend access to Annals ofMathematics.

http://www.jstor.org

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANNALS OF MATHNUMATICU

Vol. 70, No. 3, November, 1959 Printed in Japan

ANALYTIC VECTORS

By EDWARD NELSON*

(Received May 6, 1959) (Revised June 23, 1959)

1. Introduction

If A is an operator on a Banach space X, we will call an element x of X an analytic vector for A in case the series expansion of eAs x has a posi- tive radius of absolute convergence; that is, in case

11A nX !1n<c

for some s > 0. For example, if X = C([0, 1]) is the Banach space of con- tinuous functions on the closed unit interval, in the supremum norm, and if A is differentiation, an analytic vector for A is simply an analytic func- tion on [0, 11. If A is a bounded operator then every vector in X is an analytic vector for A, so that only unbounded operators will be of inter- est.

A number of problems in analysis may be regarded as the study of analytic vectors for an operator. It is a classical result that if A is an elliptic partial differential operator with analytic coefficients and if Au= 0, then u is an analytic function. In the last section we show more gener- ally that if u is an analytic vector for A then u is an analytic function. For example, if A is self-adjoint then any element of a spectral subspace corresponding to a bounded interval is an analytic function.

Most of our results concern representations of Lie groups and Lie al- gebras. In ? 7-8 we show that any (strongly continuous) representation of a Lie group on a Banach space has a dense set of analytic vectors (well- behaved vectors) and in ? 9 a simple criterion is found in order that a Lie algebra of unbounded skew-symmetric operators on a Hilbert space give rise to a unitary representation of a Lie group. Detailed descriptions of these problems will be found in the opening paragraphs of the mentioned sections. Here we shall discuss briefly the general method.

Suppose that A and X are two operators on a Banach space X. The main theorem in ? 3 gives a sufficient condition in order that every analyt- ic vector for A be an analytic vector for X. This involves estimat- ing 11 Xnx 11 in terms of 11 x 11, 11 Ax 11, ... , 1 Anx I1. Suppose that 11 Xx 1 j I Ax 11 for all x (leaving aside the question of domains of opera-

* National Science Foundation Fellow. 572

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 573

tors). If X and A commute, then I XIx n < 1 Anx I for all n. In the general case this is not so. However, it is true that II X2x 1 < II AXx 11= 11 XAx - (XA - AX)x 11 < 11 A2x 11 + 11 (XA - AX)x 11, so that we need some estimate for XA - AX = (ad X)A, and in general for the higher order commutators (ad X)nA. It turns out to be sufficient to require (for all x and n) II (ad X)nAx I I< ?n I Ax II, where the power series with coef- ficients cn/n! has a positive radius of convergence.

In all of the applications, A will be an elliptic operator (either in the ordinary sense as a partial differential operator or in an abstract sense as an element of the enveloping algebra of a Lie algebra) and X will be a first order operator. Then (ad X)nA is of order less than or equal to the order of A, and the ellipticity of A enables us to estimate (ad X)nA in terms of A.

The results on elliptic partial differential operators in the last section depend only on ? 2-4. Section 8, in which it is shown that a representa- tion of a Lie group on a Banach space has a dense set of analytic vectors, is based on somewhat different considerations, and is practically independ- ent of the rest of the paper. The treatment of unitary representations in ? 7 is simpler.

2. The calculus of absolute values

It will be convenient to develop some simple rules for calculating with relations of the form: Cx <1 < I Ax 1 + II Bx I for all x. We write this as I C l < ? A I + I B l and call the symbol lA l the absolute value of A. A formal definition is given below. We abbreviate

En= 11 A n by 11 elAIsx ||

Let X be a Banach space. By an operator A on X is meant a linear transformation defined on a linear subspace Z(A), called the domain of A, and taking values in X. By the absolute value I A I of A we shall mean the set consisting of A alone (for the sake of definiteness). We denote the set of all operators on X by 0(X). Let I 0(X) I be the free abelian semigroup with the set of all I A l, with A in 0(X), as generators. That is, an element a of I 0(X) I is a finite formal sum

(2.1) a =A1l+ . +lAi. if

(2.2) =lBl + - + B.l then a and ,8 are equal elements of 0(X) if they are identical except

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

574 EDWARD NELSON

possibly for the order of the summands. If a is a positive number we shall identify a with I aI 1, where I is the identity operator on X. We de- fine the product of a and ,3, given by (2.1) and (2.2), by

(2.3) a,8 = Li I

We recall that the product AB of two operators on X is the operator whose domain D(AB) consists of all vectors x in @(B) such that Bx is in @(A), in which case (AB)x = A(Bx). The sum A + B has as domain @(A) n @(B). With the definition (2.3), 10(X) 1 is a semiring and, because of the idetification of a with I aI 1, a semialgebra.

We shall adopt the convention of writing 11 Ax 11 = co to mean that the vector x is not in the domain of the operator A. With a in 1 0(X) I given by (2.1), we define 11 ax 11, for all x in X, by

(2.4) 11 ax = 11 Alx 11 + x ? * + 11 Alx

We define a relation < on 0(C) I by putting a < ,8 in case I Iax 11< ?,j/x11 for all x in X.

Notice that the following are true, for all operators A, B, and C on X:

(2.5) IA + B < AI + IB1

(2.6) if IjAI < IB, then AC < IBC.

In fact, for all x in @(A) n @(B), 11 (A + B)x 11 < 11 Ax 11 + 11 Bx 11 If x is not in i(A) n @(B), I Ax 11 I 11 Bx 11 = co, so that (2.5) is true. If I A I < 1 B I then in particular for a vector of the form Cx, 11 ACx 11 11 BCx 11, proving (2.6).

We shall have occasion to use elements 9 = En=0 CanSn of the semial- gebra of all power series in a variable s with coefficients in 1 0(X) 1. If also ir = on=0 sn, we define p < J in case an <? On for each n, and we define 11 Px 1, for x in X, by 11 px 11 = ? 11 alnX jjS. We define (d/ds)9(s)

and pq(t)dt by formal differentiation and integration.

A vector x in X is called an analytic vector for a in I O(:() I in case I eosx 11 < co for some s > 0. If a is given by (2.1), then

11emx1 = 1 -i 1j (1AI I + + I A, j)nx 11 sn

E00 0 At.. Aix l Inn

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 575

Thus II easx < oo if and only if the series expansion of e(Al+ +AZ)s x is ab- solutely convergent. If s < * s, .., s <_ s then eAl31+ +AtSnX is a fortiori absolutely convergent. Thus x is an analytic vector for a if and only if the series expansion of eAlsl+ +AZlSX converges absolutely for (si, *.., sz) sufficiently small.

Let t be an element of ((X),= l Xi I + * * * + I XJ 1. We define ad t on i O(l) i by

(2.7) (ad 4)a = X5I -AjXj I where a is given by (2.1). If X is in O(:), we define adX on O(l) by (adX)A XA - AX. The following combinatorial lemmas will be needed in ? 3-4. We recall that A D B mean that @(A) D @(B) and A agrees with B on @(B). The domain @(A) of an operator A is called invariant in case A maps @(A) into @(A).

LEMMA. 2.1. If X1, * *, X., A are operators on X then

(2.8) Xn * X1A D Lk o=aE(nfk) (ad *ad XX(,A)Xr(f**

where (n, k) denotes the set of all (n) permutations a of 1, -.., n such

that u(n) > u(n-1) > ... > a(k + 1) and a(k) > a(k-1) > -- >(1). If X1, *--, X", A have a common invariant domain, then equality holds in (2.8).

PROOF. For n = 0, (2.8) states that A D A and for n = 1 it states that X1A D AX1 + (ad X1)A, which is true. Suppose that (2.8) holds for n and let X,+1 be an operator on X. Then

Xn+lXn * X1ADEk =ZOae(n k) Xn+(ad Xr(k) * ad Xa(1)A)X,(n))... X,(k+

D Ik ?e(n(ic) (ad X.(k)... ad X(1)A)Xn+1X,,(n)... X(k+

+ (ad Xn+1 ad X(k) * ad X.(1)A)X,(n) -X* (k.+1).

Let z be a permutation in (n + 1, k). Then.

(ad X(k) ... ad X,()A)X-(n+l) ... X(k+l)

occurs as a term before the + sign in the braces (corresponding to a a in (n, k)) if z(n + 1) = n+1, and as a term after the + sign (correspond- ing to a a in (n, k - 1)) otherwise, for either z(n + 1) or z(k) must be

equal to n + 1, by the definition of (n, k). Since (n+?1) (n)+(nl) the correspondence is one-to-one, and (2.8) holds for n+ 1. The last state-

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

576 EDWARD NELSON

ment of the lemma is obvious, for then the two sides of (2.8) have the same domain.

LEMMA 2.2. If X1, ..., Xn, are operators on I then

(2.9) AXn ... X1 D Xn ... XjA -fin where

(2.10) f k = rle(lk) (ad Xa(k) ... ad Xa(1)A)Xr(n) ... (k+)

If X1i ., Xn have a common invariant domain, then equality holds in (2.9).

PROOF. The last statement is an immediate consequence of the preced- ing lemma. To prove (2.9), notice that it is true for n = 0 and n=1, as- sume it for n, and let Xn+j be an operator on X. Then (2.9) holds with Xn,* - - , X1 replaced by Xn+, * * , X2 and fin modified accordingly; calling fn the expression which replaces fn, we have

AXn~ X2X1 D Xn+j ... X2AX, - f X, D Xn+j X2X1A - Xn+j * X2(ad X1)A -f'X1.

Applying Lemma 2.1 to Xn+j * - X2 (ad X1) A (with (ad X1) A playing the r'le of A in that lemma), we see that X,+1 ... X2 (ad X1)A + f IX1D fn+1, each permutation z in (n + 1, k) with z(1) = 1 corresponding to a term in Xn+j ... X2 (ad X1)A and each z with z(1) # 1 corresponding to a term in fn+1.

LEMMA 2.3. Let e and a be in 1 0(X) 1. Then

(2.11) an < 4rna + Sk ( )((ad t)k a)4n-k

PROOF. Let=X1 + -j..-+-IXa, a=IA1l+...+IAj,and let A* denote the sum over all 1 < 1, 1 < g, < d, * 1 < gn < d. Then, us- ing Lemma 2.2,

an = * I A XgnA - XI E Xgn Xg1Aj I

+ E I Znlffe(n k(ad X(, ) Xad X(gl)Aj)X(9n) X(gk)

= e ? +E I *E (l)(ad Xgk * ad Xq1A,)X * Xqk1 I

since there are (n) permutations in (n, k), so that each term (ad Xq-

ad Xq1A)X * X, occurs (n) times. But this is < the right hand side

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 577

of (2.11), concluding the proof.

3. The main theorem

THEOREM 1. Let a be a Banach space, e and a in 10(X) 1. Let 4 < ca, (ad 4)na < cna, and

u(s) =

K((s) | 1 d t

Then e"s< ecK(S)

If c and Cn are such that c < oo and v(s) has a positive radius of conver- gence, v(s) < oo for some s >0, we shall say that a analytically dominates 4. If v has a positive radius of convergence, so does K. Therefore the following corollary follows at once from the theorem.

COROLLARY 3.1. Let X be a Banach space, e and a in I O(C) l. If a ana- lytically dominates 4 then every analytic vector for a is an analytic vec- tor for 4.

We state the next corollary, which is the consequence of Theorem 1 which will be used most often, without using the terminology of ? 2.

COROLLARY 3.2. Let X1, *- -, Xd, A be operators on a Banach space X. Let k and kn, n = 1, 2, ***, be such that for all x in the domain of A

IIXixII < k (11Ax 11 + II x 11) , 1 < i < d ad Xi,** ad Xi Ax II < kn(II Ax II + II x 11), 1 < t ? i Yin < do

Suppose that k < oo and that S1(kn/n!) Sn < 00 for some s < 0. If

there is an s > 0 such that

Ln-0 n !

then for (s1, ** , sd) sufficiently close to (0, *.*, 0)

Efl n!1 11** 1l Xi, Xi x 1Si, *si < cc En=O n! El{,-,ndA n l n

PROOF. Set e = I X1i + + I Xj I and a = I A I + I I 1. Notice that (ad 4)na = (ad 4)n I A l. Therefore a analytically dominates 4 with c given by dk and Cn given by dnkn. Corollary 3.2 now follows from Corollary 3.1.

We turn now to the proof of Theorem 1. Define by recursion

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

578 EDWARD NELSON

7a = II

(3.1) n= cw"a + k

Thus 1- = ca, w2 = c2a2 + c1ca, and each wrn is a polynomial in a. We will show that for all n (3.2) en <? w,

In fact, we will prove that for all n > 1

(3.3) aen4i it 1 w C

Since e < ca, en < ca 'n-i, so that (3.3) implies (3.2). For n = 1, (3.3) says that a < (1/c)ca. Suppose that (3.3) holds for all k < n. Then, by Lemma 2.3,

amn < sna + ELit )((ad _a) I

? Ea + E k

< ?na + Ek? I()ck I

< I

nl

proving (3.3). Let r(s) be the power series

w(s) = E 0 .

By (3.1) and the relation (n)=n!/k! (n - k)! we have

(n?+1) _" = C a +E1C(n + l-k) ~~' (n + 1 )! n! k! (it +-k)!

By definition of u(s), this says that d d (3.4) -r(S) = cr(s)a + u(s) )(S) . ds ds

That is, (d/ds)w(s) = caw(s)/(1 - i(s)). Using the definition of K(s), this implies that r(s) = eCcaK(s), as may be verified by differentiating both sides, giving (3.4), and observing that both sides have the same constant term i I lo By (3.2), efs < 7(s) = ec K(s) concluding the proof.

We have stated the theorem in terms of Banach spaces, but we have nowhere used the fact that X is complete. The results and proofs of ?2-3

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 579

remain valid if a is a vector space over the real or complex numbers with a "norm" 1111 satisfying 0 < 1I xjj < co, 1I x + y II< _I x 11 + 1I y 11,11ax 11= I a I II x II for all x, y in X and scalars a (with the convention that 0 00 =0). Also, we have nowhere used the fact that A1, *-., A, are linear. How- ever, only the theorem as stated will be used in the sequel.

4. Non-commuting vector fields

The theorem proved in this section will not be needed until ? 12, but it serves to clarify ? 7 and to illustrate Theorem 1.

Let U be an open neighborhood of the origin in RI (d-dimensional eu- clidean space). A function u defined in U and taking values in a Banach space X is said to be of class Co in case all mixed partial derivatives of u exist in the strong topology; e. g.,

1L(u(x, + h, x2, *Y** xZ) - U(xl, X2, *Y** XJ)) h

converges in X as h - 0, for all (x1, ..., Xn) in U. Let K be a compact subset of U and let

(4.1) || U 11K = SUPPE 11 U(X) 11

The function u is said to be analytic in U in case for each point x in U, there is an a > 0 such that if K is the closed sphere with center x and radius e, then u is an analytic vector for I 0/8x1 I + *-- + I 8/8xI I in the norm I IK; i. e., in case

En= 0 n! E 1!:::il ... lin:d || ax az IU K~ 1

0 in

is absolutely convergent for s1, , SI sufficiently small. The property of being analytic is, of course, invariant under analytic changes of coordi- nates. In some contexts, notably the study of Lie groups, it is more con- venient to use a family of d linearly independent analytic vector fields X1, -, XI which do not commute, rather than the vector fields (a/8x1), .-., (8/18x) of a coordinate system. (We recall that a vector field X= , a,(x)(0/8x) is called analytic in case the a, are analytic. If X and Y are two vector fields, so is XY - YX, but it is not 0 in general.) The next theorem asserts that the notion of analyticity with respect to X1, *-, XI is the same as the usual notion.

Let XK be the Banach space of all continuous functions from K to X in the norm (4.1), and let X be a vector field with continuous coefficients,

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

580 EDWARD NELSON

X = E a,(x)(O/8x,). If K is the closure of its interior, then we define an operator, also denoted X, with domain Z(X) consisting of all u in l which are in C (K); i. e., have extensions u to U of class Coo, with Xu being defined as the restriction of E aj(x)(8/8xj) it to K. If u = 0 on K then Xu = 0, since K is the closure of its interior. (If K were not the closure of its interior, we could take yK to be the space of functions defined on U of class Co with the " norm " (4.1). Then differentiation of X might take a function of " norm " 0 into a function with positive " norm ". By the remarks at the end of ? 3, the proof of the following theorem would apply to that case as well.)

THEOREM 2. Let U be an open set in R', Ka compact subset of U which is the closure of its interior. Let Y1, *., Yd be analytic vector fields on U which are linearly independent at each point of U, and let X1, ... , Xl be analytic vector fields on U. Then any analytic vector for

a = |YI ? ? + + I Ya I is an analytic vector for t = I + ? + i X? 1. In fact, Y) analytically dominates 4.

PROOF. Let Xi = EL ajj(x)(O/8xj), Y, = 57 1bj(x)(O/8xj), and 8 = I I/ax, I + + I 8/Oad|

Let a. be the maximum absolute value on K of all mixed partial deriva- tives of order n of the apj and bus. Notice that since the art, bus are a finite set of analytic functions and K is compact,

(4.2) a(s) = E s an Sn < 00

for some s > 0. Now I Xi < a08, so that e < la08. Let b be the maximum absolute value on K of elements of the matrix inverse to b,,(x), so that 8 < dbA. Since Y1, ... , Yd are everywhere linearly independent and K is compact, b < oo. Letting c = lao db, we have e < cr with c < oo.

Define by recursion

(4.3) ko(s) = a(s) kn+l(s) = d a(s)Dkn(s) + d.kn(s)Da(s)

where D = dids (and d is the dimension of Ra). We need to estimate (ad4)ny. Now (ad on)= f I ad Xin ... ad Xi, Y 1, where the summation is over all 1 ?< ins ., i1 < 1 and 1 < j < d, so that there are lnd terms in (ad 4)ny. Let

(4.4) ad Xi * ad X11YJ = -h1 p1l.. f1J(x)- n h=1 Phin 1~~~~~~X,

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 581

If p is a function on K and k is a power series with positive coefficients in one variable s, we say that p is majorized by k in case we always have I (8/x0m) ... (W/ax,) p(x) I < Dmk(O) for all x in K and m = 0, 1, 2, *.. It is clear that

(i ) if p is majorized by k then (8/8xg)p is majorized by Dk, and (ii) if p, p are majorized by k, k respectively then pk is majorized by ~~~~~~~~~~~~~~~A

kk and p ? p is majorized by k + k. Now we shall prove by induction that each plan .1 is majorized by

kn, for all n. For n = 0 this is clearly true, by definition of a(s). Suppose that it is true for n. Now

ad Xn+l ad Xn* ad XiYj = dg h~l +g( ax Phtn * fly as

- 9 ha1 Phin ...'j( ax an+10 ax

so that

(4.5) Phi= E aj ( aa Phi ...A

E- g=I P9gn *-- 1jX ain+1h)

The first term on the right hand side of (4.5) is majorized by the first term on the right hand side of (4.3), by (i) and (ii), and similarly for the second terms. This completes the induction. Consequently,

(ad t)ny < 1nd2kn(0)8 < 1nd3kn(O)by) It remains only to show that if Cn = 1'd3kn(O)b then E' 0(cn/n!) s' < cc for some s > 0, which is equivalent to showing that 57=0(kn(0)/n!)tn < oofor some t > 0.

We shall in fact show that ?E7?0(kn(s)/n!)tn is absolutely convergent for s and t sufficiently small. By (4.3), kn(S) = (D-a(s))na(s) where the opera- tor D-a(s) is defined by (D-a(s))f(s) = D(a(s)f(s)). But (D-a(s))na(s) = a(s)-'(a(s)D)na(s)2 and by the analytic change of variables given by r = - this is equal to c(r) (d)c(r)2, where c(r) is the analytic

oa~w wr function c(r) = a(s(r)). Since c(r)2 is an analytic function of r, the series ,n?=0(1/n!)c(r)-I(dn/drn)c(r)2tn is absolutely convergent for r and t suffi- ciently small. That is, E' O (kn(s)/n!)tn is absolutely convergent for s and t sufficiently small, concluding the proof. Notice that this argument

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

582 EDWARD NELSON

shows that E 0(k.(s)/n!)tn is absolutely convergent for all complex values of s and t sufficiently close to 0.

COROLLARY 4.1. Let t and C be as in Theorem 2 and let m be a positive integer. Then (ad t)nfl is the sum of lnd terms of the form |d=~lPh(X)(8Iaxh)1

such that if

(4.6) c. = SupP0J:flm, G * PA()

then the power series with coefficients cn/n! has a positive radius of con- vergence.

PROOF. We have seen that (ad t)ny is the sum of lnd terms each of which is the absolute value of a vector field of the form (4.4), where p, is majorized by kn. Since ph is majorized by kn, the right hand side of (4.6) is bounded by kn(O) +Dkn(O) + +D mkn(O). Since En 0(kn(s)/n!) tn is absolutely convergent for complex values of s and t sufficiently close to 0, E' 0(DJkn(s)/n!)tn is absolutely convergent for each j, if s and t are sufficiently small. This proves the corollary.

5. Extensions of operators

At various places in the following sections we shall show that the hypoth- eses of Corollary 3.2 are satisfied by operators X1, ..., Xa, A on a cer- tain domain and we shall also know that the closure A of A has a dense set of analytic vectors. (We recall that the closure A of an operator A, if it exists, is defined on the domain @(A) consisting of all x such that there is a sequence x, of vectors in Z(A) with xj -* x and Ax, -* y for some y in X, in which case Ax = y. In order for this to be well defined, it is necessary and sufficient that whenever x; are in Z(A) with xj 0 and Ax, -+ y, then y = 0.) We shall want to conclude that I Xi I + - + IXdI has a dense set of analytic vectors. I do not know whether this is true in general. In fact, let A and X be two operators having closures and such that I A I analytically dominates I X 1. Then it is not even clear that if x is an analytic vector for A then x is in the domain of X2. However, there are two ways of extending an operator A. Besides the closure, or strong extension, we may consider weak extensions. This theory is simplest for the case of a symmetric operator A on Hilbert space, in which case we mean A*. This case will suffice for the applications. In Lemma 5.2 we show that in case A= A*, so that the two extensions coincide, the desired inequalities do hold for the closures of the operators involved.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 583

See [30] for the theory of symmetric operators. Here we recall the fundamental definitions. An operator A is called densely defined in case Z(A) is dense. An operator A on a Hilbert space g is called symmetric in case A is densely defined and (Ax, y) = (x, Ay) for all x and y in Z(A), skew-symmetric in case it is densely defined and (Ax, y) = - (x, Ay) for all x and y in Z(A). The adjoint A* of a densely defined operator A is defined on the domain Z(A*) of all y in c' such that (Ax, y) (defined for x in @(A)) is a continuous linear functional of x. Since Z(A) is dense, there is a unique vector A*y such that (Ax, y) = (x, A*y) for all x in Z(A). An operator A is self-adjoint in case A = A*, skew-adjoint in case A = - A*, essentialy self-adjoint in case A = A*, and essentially skew-adjoint in case A =- A*.

The following lemma, and the last statements in Lemma 5.2, will not be used before ? 9.

LEMMA 5.1. Let X be a closed symmetric operator on a Hilbert space. Then X is self-adjoint if and only if it has a dense set of analytic vec- tors.

The lemma remains true if " symmetric " is replaced by " skew-sym- metric" and " self-adjoint " by " skew-adjoint ", by considering iX (which has the same analytic vectors as X).

PROOF. The necessity of the condition is a trivial consequence of the spectral theorem. In fact, if X is self-adjoint and p is a Baire function of a real variable such that I 9(X) I < ke-clxl for some c > 0, k < 0o, then any vector in the range of q(X) is an analytic vector for X. (Conversely, any analytic vector x for X is in the range of p,(A) for c sufficiently small, where qi(X) = ecIxl, since we may write x = p,(A)q-,(A)x for c suffi- ciently small.) For example, if X has resolution of the identity E(.), and 4) is a bounded Borel set, then any vector in the range of the spectral pro- jection E(1?) is an analytic vector for X. Such vectors are dense, by the spectral theorem.

We must prove that the condition is sufficient. For each s > 0 let s be the set of vectors x in D such that 11 elxlsx 11 < oo. Then s is a linear sub-space of i?. Let t U>0,. Our assumption is that &0 is dense in &.

If x is in Its and I t I < s define

(5.1) U(t)x = (iX)0 xtn

This converges absolutely in &. We will prove that if x is in Us and I t1 I + I t2 I < s then U(t1)x is in 't, and (5.2) U(t2) U(t1)x = U(t2 + t1)x

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

584 EDWARD NELSON

Let y= U(t1)x, y=} _{(iX)/fn!}xtn. Then y, is in @(X) and X(yj-y8)-+O as j, k -- cox since x is in. Since X is a closed operator, this mean that y is in Z(X) and Xyj-*Xy. Now Xy--* U(t1)Xx, so that XU(tl)x = U(t1)Xx. Consequently

(iXy,(1 (~n! xt ) t, - (iX)k+n Xtntk

converges absolutely as 1 -00o to

(iX)tM him 0 MiX!) x(t1 + t2)M = U(tl + t2)X

proving (5.2). Now if x is in v- and I t I < s then

d -(U(t)x, U(t)x) = 2 Re lim,0 U(t + h)-U(t)x,U(t)X d t lmo( +h-Ut)Utx

= 2 Re limft ( U(h) - 1 U(t)x, U(t)x) = 2 Re (iXU(t)x, U(t)x)= 0 h

since X is symmetric. Therefore for all x in T,

(5.3) 11 U(t)x 11 = 1x xI if I t I < s. Consequently

C. 11 XnU(t)x 11 = 11 U(t)XnX lTSn X E- n IIS < c En=O

n! S

n=O~~ n! n=On

if x is in Qand It i < s. That is, if It i < s then U(t)',c ts Now for an arbitrary real number t let the integer m be such that I t I < ms and define U(t)x, for x in t, by U(t)x= {U(t/m)}mx. Then U(t) is well defined for all real t on ?i, U(t)tt. c ?, and (5.2) and (5.3) are satisfied for all x in 1S * This may be done for each s > 0, and the resulting definitions of U(t)x are consistent if x is in 6?, and Us, so that U(t) is defined on k0, U(t)tt0 c Q, and (5.2), (5.3) are satisfied. Since ._o is dense in if and II U(t) 11=1 we may extend U(t) to all of A, preserving (5.2) and (5.3). In short, there is a one parameter unitary group U(t) on tp such that (5.1) holds if x is in and I t I < s.

By construction of U(t), for x in N, U(t)x is a continuous function of t. Since O is dense and U(t) is uniformly bounded (11 U(t) 11 = 1), U(t)x is a continuous function of t for all x in '?. That is, the one parameter unitary group U(t) is strongly continuous. By Stone's theorem [29], there is a unique self-adjoint operator Y such that U(t) = e"Y. We wish to show

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 585

that X = Y. Let X0 be the restriction of X to &. Then X0 c Y since for each x in ,

iYx = limbs U(h) 1 iXX. h

Now let Z be an arbitrary self -adjoint extension of X0 and let V(t) = eit. Then for all x in s and I t I < s we have by (5.1) and the fact that Z and X agree on

U(t)X = no(iZ)n t n!Xt

Since the series is absolutely convergent it must be V(t)x, by the spectral theorem. Thus for I t I _ s, U(t) and V(t) agree on S. Repeating the argument used in the definition of U(t) on S0 for all real t, we see that U(t) and V(t) agree on ?0 and consequently on .9. Therefore Z = Y, and X0 has a unique self-adjoint extension. The means [30, Chap. IX] that the closure of X0 is self-adjoint. Since X is a closed symmetric extension of X0, X is self-ad joint (in fact, X = XO = Y = Z), which concludes the proof.

LEMMA 5.2. Let X1, .*., Xa, A be symmetric operators on a Hilbert space Q with a common invariant domain X, and suppose that A is es- sentially self-adjoint.

LetA = IX11 + -- + IX, 1,a= IAI + I IK < ca and (ad 0)na < cana with c < oo and cn < 0o, for all n ? 1. For all finite sequences ii. 1 , in

(5.4) Z(An) C Z (X1... *X, )

Let Z =fnl(An) and let X1, **., X4, A be the restrictions of X1, ***, X,,A, respectively, to Z). Let a X I + + I Xa1, l I j + III. Then a < c, (ad t)na <_ Cnd for all n > 1.

If a analytically dominates a then there is an s > 0 such that the set of x in Z for which II e~sx II < oo is dense in i?, and each Xi is essentially self-ad joint.

In the applications the Xi will be skew-symmetric, but this does not al- ter the validity of the lemma (if " self-adjoint " is replaced by " skew- adjoint " in the last statement), by considering iX1, ..., iXd instead.

The reason for restricting the closures to Z is that Z(A) need not be included in (X1A - AX1), for example. By restricting to Z it suffices to consider X1A - AX1 instead. However, any analytic vector for A is an analytic vector for A, by definition of 0, so there is no loss in considering

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

586 EDWARD NELSON

A. PROOF. First we will prove (5.4), beginning with the case n = 1. If x

is in T(A) there is a sequence x, in Z with x, -* x and Ax, -* Ax. For all all i = 1, *.. I, d, || Xi (x - Xk) ?| C c(j| A(x, - xk) 11 + || x - xk 11) -O as j, k -+oo since c < oo. Therefore x is in Z(Xg). The same argument shows that if C = ad X,. ad Xi1 A then x is in @(C), since c, < oo.

Instead of proving (5.4) directly we shall prove

(5.5) Z(An) c Z(AXti-- Xn-) This implies (5.4) by the case n = 1 above. Suppose that (5.5) holds for n, and let x be in Z(An+l). Since A=A*, we need to show that X1.. *Xx is in ,(A*); i. e., that (Xn * * *Xi, Ay, x) is a continuous linear functional ofyin Z. Now X1n ..X,1A = AXgn .Xj1 + S, where AXn *-- Xi, is the term k = 0 in (2.8) and S is the remainder. By the induction hypothesis and Lemma 2.1, (Sy, x) is a continuous linear functional of y, and since x is in Z(An+'), A*x is in 4(An) and (AX * Xi1y, x) = (Xn * * Xi1 y, A*x) is by the induction hypothesis a continuous linear functional of y. This proves (5.5).

As a consequence of (5.4), the operators X1, ***, Xd (and A) leave Z invariant. Let x be in Z. Then x is in @(A), and there is a sequence xf in Z with xj -+ x, Ax, Ax. Then

11 Xix |I lim=b I| Xixj ||

? limxc(jI Ax, 11 + 1j x, 11) = c(jI Ax 11 + 11 x 11) so that i < cti, and similarly (ad j)njf = Cne.

Now suppose that a analytically dominates 0, so that the power series v(s) and /c(s) of Theorem 1 have a positive radius of convergence. Let E(-) be the resolution of the identity for the self-adjoint operator A, and let Q3 be the set of all vectors x such that for some bounded Borel set 1?, E (Q)x = x. Then Q3 c Z, Q3 is dense in S>, and II eltx II < oo for all x in 3 and 0 < t < co, by the spectral theorem. If s is small enough so that K(s) < oo, then by Theorem 1, 11 esx 1 < 11 eC'K(s)x 11 < co for all x in Q3. Any analytic vector for j is a fortiori an analytic vector for each X_, so that by Lemma 5.1 each Xi is essentially self-adjoint. Since X7cX cXj, each Xi is essentially self-adjoint, concluding the proof.

6. Domination in the enveloping algebra of a Lie algebra Lemma 6.2 below is the basic estimate used in ? 7 and ? 9. Throughout

this section we shall assume that ?& is a Hilbert space, Z a linear subspace of I, g a Lie algebra of skew-symmetric operators having Z as common

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 587

invariant domain. Notice that the Lie product [X, Y] = XY - YX of two skew-symmetric operators is skew-symmetric. Let a be the asso- ciative algebra, over the field of real numbers, of operators on i generat- ed by g. In other words, a (or strictly speaking, a with the identity operator on Z adjoined if necessary) is the enveloping algebra of g. It is a homomorphic image of the universal enveloping algebra of g. An ele- ment of a is said to be of order < n in case it is a real linear combination of operators of the form Y-1 * - Yk with k ? n and each Yi in g. The set of elements of a of order < n will be denoted a.. Let X1, *--, Xd be a basis for g. Then it is clear that (6.1) Xo, He XXi + XiX; 1< i, j C d constitutes a set of linear generators for A2. Let (6.2) A = X1 + *. +X.

LEMMA 6.1 If the operator B is in A2 then for some k < oo, I B I < kIA-It.

PROOF. It is clearly sufficient to prove this for the operators (6.1). For the Xi we have

E X _x 1 Edt (XiX, XiX) = (- AX, X)

( 1 -A + +)x, x) =(-(- I)2X, x) = 2 (A )X 2

By the Schwarz inequality for finite sequences,

(6.3) d=1 Xix 1 < ?/2d 11 (A - I)x 11

It remains to consider the He. We shall use the notation B+ for the restriction of B* to i, if B is in a. Thus (Xi,- - -Xin)+ = (-l)nX.n . Xi%. Let f be the set of elements in A of the form EQr+Qr (using finite sums only). Note that -A is in 2B. Now

-A + Hj = (Xi - Xj) (Xi - Xj) + ,Xt+X; i : j

(6.4) - 2A + Hii = 2 E,:6X Xt

- A - Hij = (Xi + Xj) (Xi + Xj) + S X X k

In all cases, therefore, if the a i are real and Be= ai jH, then there is an a ? 0 such that (6.5) -aA + B e ?.

Consider now 4A2 _ H2 = (2A - Hij)(2A + Hij) + B1 where B1 is in a3. Now -(2A - Hj) = EtY+ Yk with the Yk in g by (6.4), and similarly -(2A + Hij) = EZ+Z, with the Z, in g. Therefore

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

588 EDWARD NELSON

(6.6) 4A - H = k(YkZz) +(YkZz) + B,

where B2 is in a,3 (using the fact that if we permute the factors in the terms of highest order of an element of a the error committed is of lower order). Now a,3 is spanned by the operators (6.1) and those of the form Hijk = XiXjXt + XiXtXJ + XjXiXt + XjXkXi + XtXiXJ + XtXjX,. Therefore we may write B2 =E aij i +S, where S is a real linear com- bination of the Xi and the Hijk. Since the other terms in (6.6) are sym- metric, S must be symmetric. But the Xi and Hivj are skew-symmetric, so that S is also skew-symmetric, and consequently S= 0. Therefore, by (6.5), there is an a ? 0 such that -aA +B2 is in ?P. Hence 4A2 - H 2 -aA is in ??. Completing the square, we have for all x in ), II Hix jj2 <

((4A2 _ aA)x,x) < 11 (2A - (a/4))x 112. Letting k be the maximum of V 2 and V a /2, we have II H, x II < k Ij(A - I)x I1, which concludes the proof.

LEMMA 6.2. Let 4 = I X1 I + - - - + I Xd |andleta=IA-Ii. Thena analytically dominates 4. In fact, 4 ? V/d cta and there is a c < oo

such that for all n > 1, (ad 4)n a < cna. Also, I A I + I I I analytically dominates 4.

PROOF. The fact that 4 < d/2 ca is precisely (6.3). Now a2 is a finite dimensional vector space, of dimension at most d +d(d + 1)/2, since it is spanned by (6.1). If B is in (2 we define III B III to be the least num- ber k such that I B I = ka. By Lemma 6.1 this is always finite, and if III B III = 0 then B = 0, so that (a2 with this norm is a finite dimensional Banach space. For each i = 1, *- *, d, ad Xi is a linear transformation taking (a2 into itself, so there is a ci< oo such that III (ad XJB I IIciI iBIII. Let c = d max~ci. Now (ad 4)na is the sum of dn terms of the form lad Xi ad Xi A l, which is < c. ... cila. Therefore (ad t)na ? cnaa. The last statement holds a fortiori, concluding the proof.

In ? 11 we will need an extension of these lemmas.

LEMMA 6.3. For all positive integers m, if B is in 'J12m then for some k< oo,

(6.7) IBI kam where am (A - I)m j. If =Y Y1j ?- + Y I where Y, is in A-42m and ad Yj maps (a2m into itself, for j = 1, ***, 1, then atm analytically dominates ). For some c < oo, (ad r2)nam ? Cnam for all n ? 1.

PROOF. It suffices to prove (6.7) for operators of the form B=Xi,.* X. with s < 2m, since by definition of (a2M they span (a2m. For m = 1, (6.7) is true by Lemma 6.1. Suppose it to be true for m - 1. Making the con- vention that if s = 0 then Xi1 *-- Xis = I, (6.7) is true for s==0. Suppose

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 589

it to be true for S - 1. Then

l t t I1X3 xis < k I Xi3 .. *X. (A-I) I + k BP i

where B' (A - I)X3- Xi - xi Xi (A-I) is in A-1, so that IB'I C k'am (with k, k'< co). Now by the induction hypothesis on m, ji 3.** Xjsi <

atm-1, so that I X., *-- Xi I < k1am + kk'am, completing the induction on s and therefore on m. This proves (6.7). The last statements are proved just as in the proof of Lemma 6.2, since '-2m is finite dimensional.

It would be interesting to find a better method of studying order prop- erties of the enveloping algebra of a Lie algebra. For example, it was proved in [23] that in a unitary representation of a Lie group an elliptic element A of the enveloping algebra is, if symmetric, essentially self-ad- joint on the Ga.rding space. One would expect that if A has real coeffi- cients it would be semi-bounded, but the authors were unable to prove this. Even in a polynomial algebra (which is the universal enveloping algebra of an abelian Lie algebra) Hilbert showed that a positive element need not be a sum of squares, so that the crude computational method used above will not work except in special cases.

7. Analytic vectors for unitary representations of Lie groups

Let G be a Lie group and X a Banach space. By a representation T of G is meant a mapping of G into the set of bounded operators with domain X such that T(e) =I, where e is the identity element of G, T(az) = T(a) T(z) for all a and r in G, and such that for all x in X, a-* T(a)x is a continuous mapping of G into X (where X has the strong, i. e., norm, topology). The representation is called unitary in case X is a Hilbert space and each T(a) is a unitary operator. An element x of X is said to be an infinitely diffe- rentiable vector for T in case the mapping a -+ T(a)x of G into X is of class Coo. The set of all infinitely differentiable vectors will be denoted e. It is dense in X. In fact, Garding showed [12] that if the function 9 on G is of class C- and has compact support and if the operator T(p) on X is defined by

(7.1) T(p)x | T(a)x9(u)du G

where da is left-invariant Haar measure on G, then for all x in X, T(9)x is in e and the set of such vectors is dense in X. The set of all finite linear combinations of vectors of the form (7.1) is called the Garding space. Let g be the Lie algebra of G and let exp denote the exponential mapping (see [5]). If X is in a, then T(X)x is defined for all x in e by

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

590 EDWARD NELSON

(7.2) T(X)x = lim T(exp hX)x -x

Then T(X) is an operator with e as invariant domain. If T is unitary it follows from (7.2) that T(X) is skew-symmetric. Since E is invariant under each T(X) for X in g, T extends to a homomorphism of the universal en- veloping algebra CU(g) of g onto an algebra of operators with domain e.

The study of a representation on the enveloping algebra is a powerful tool in the study of Lie group representations. Harish-Chandra pointed out in [15] that for many purposes E is the wrong domain for the operators T(X). For example, a subspace of e may be invariant under all the T(X) without its closure being invariant under the T(a). For this reason Harish- Chandra introduced the notion of well-behaved vectors. Following Cartier and Dixmier [4], we will call them analytic vectors. A vector x in X is an analytic vector for T in case the mapping a -* T(a)x of G into X is an- alytic.

For this to be a useful notion it is necessary to show that the set of analytic vectors for T is dense in X. Harish-Chandra [15] showed this to be the case for certain representations of semi-simple Lie groups. Using similar methods, Cartier and Dixmier [4] showed that if T is either bound- ed or scalar-valued on a certain discrete central subgroup Z of G then the set of analytic vectors for T is dense. In particular, their result includes all unitary representations. This method involves proving the theorem first for certain Lie groups which as analytic manifolds are isomorphic to euclidean space and then using structure theory of Lie groups to extend the result.

In this section we shall give a new proof of the denseness of the analyt- ic vectors for a unitary representation. The proof makes no use of struc- ture theory. Since the proof is entirely infinitesimal, it may be applied to representations of Lie algebras (see ? 9). Besides showing that the analyt- ic vectors are dense, Theorem 3 describes a useful set of vectors which is a dense set of analytic vectors. An application of this additional inf or- mation is made in ?11. In ?8 it is shown that an arbitrary representation has a dense set of analytic vectors. This proof is by a different, though related, method.

First we need to establish the connection between analytic vectors for a group representation and analytic vectors in the sense of ? 2.

LEMMA 7.1. Let T be a representation of the Lie group G on the Banach space X. Let X1, -*, Xa be a basis for the Lie algebra g of G, and let d = I T(X1) I + + l T(X,) l. Then x in X is an analytic vector for T if

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 591

and only if x is an analytic vector for A. PROOF. Let x1, *--, xd be analytic coordinates in a neighborhood of

e. Since translations by elements of G are analytic isomorphisms, a--+ T(u)x is analytic on all of G if and only if it is analytic in a neighborhood of e (see [15, p. 209]). Let K = {v e G: lx1(a) ?I , *.., lx(u) I < e} Then x is an analytic vector if and only if for some E > 0, a -* T(u)x is an analytic vector in the norm (4.1) for I 8/8x1 I + *-- + I /8Xd I.

We shall identify g with the space of all vector fields on G which com- mute with all right translations by elements of G. If X is in g and x is in e then, by (7.2)

T(X)x = XT(u)x!0e. Therefore x is an analytic vector for 5 = I T(X1) I + * + I T(X,) I if and only if for some E > 0, a -* T(a)x is an analytic vector in the norm (4.1) for I Xi I + *-- + I XI l. By Theorem 2, the proof is complete.

The sufficiency of the condition, which is all that we shall need, may also be proved without appealing to Theorem 2. Since the exponential mapping is an analytical isomorphism of a neighborhood of 0 in g with a neighborhood of e in G, it is sufficient to prove that if x is an analytic vec- tor for 5 then X -* T(exp X)x is analytic in some neighborhood of 0 in g. Let X X1t, + ... + Xltd. Then this occurs if and only if

(7.3) E7OO 0 ...a~nI~aiaI I tal .

taad< c En=o n! Ea 1+...+a d-n ||1l1.a | l *atad 1 0

for t1, ..., td sufficiently small, where aka1...ad is the coefficient of ta.... ta din the expansion of exiti+ +XdtdX. Thus 11 *a, ...a ll is the norm of the sum of several terms whose norms occur in the expansion of 11 eOsx 11, if we take t1 = . = td s. Therefore the left hand side of (7.3) is < 11 efsx 11, which is < oo if x is an analytic vector for I.

The first statement in the conclusion of the following theorem is due to Cartier and Dixmier [4].

THEOREM 3. Let U be a unitary representation of the Lie group G on the Hilbert space &. Then the set of analytic vectors for U is dense in

Let X1, ..., Xd be a basis for the Lie algebra g of G, and let A= XI+ - - - +XI in the universal enveloping algebra of g. Then any analyt- ic vector for U(A) is an analytic vector for U, and the set of such vectors is dense in h.

PROOF. Let = i U(X1) 1 + + i U(Xd) I, a = U(A) - Is. By Lemma 6.2, a analytically dominates 5. (This lemma is applicable since the U(X), with X in g, are skew-symmetric and since U(A)= U(X1)2+ - - - + U(X,)2).

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

592 EDWARD NELSON

By [23], U(A) is essentially self-adjoint. By (5.4) in Lemma 5.2, every vector in n = fl=1 Z(U(A)n) is in 9(U(X1) ... U(Xn)) for all finite se- quences i1, * * in. Now if X is in g then Z( U(X)) is simply the set of all vectors x for which the limit in (7.2) (with T replaced by U) exists (this follows from Stone's theorem [29] and Segal's theorem [27] (or [23]) that U(X) is essentially skew-adjoint, for example). Therefore if x is in (, U(u)x has all partial derivatives at a = e. Now U(X) U(u)x = U(a) U( Y)x where Y= (Ad v-1)x, so that U(u)x has all partial derivatives at all a, and x is in e. That is, ( ct . Since the reverse inclusion is obvious, CY = C, and any analytic vector for U(A) is an analytic vector for U(A). Now U(A) is self-adjoint and so has a dense set of analytic vectors, by the spectral theorem (see Lemma 5.1). By Corollary 3.1 these are all analytic vectors for #, and by Lemma 7.1 they are analytic vectors for U. This concludes the proof.

We sketch here a brief proof, due to W.F. Stinespring and myself, that U(A) is essentially self-adjoint. Since it is semi-bounded from above by 0, it is enough to show that the range 9 of U(A) - I is dense. Sup- pose that x is orthogonal to 9. This implies (using (7.1)) that the numerical function (U(a)x, x) is a weak solution of the equation (A - 1)(U(a)x,x) = 0. By the regularity theorem for elliptic equations (e. g., [24]), it is a solu- tion in the ordinary sense. But (U(a)x, x) is a positive definite function, and so has a maximum at e. But this contradicts the maximum principle unless x = 0.

8. The heat equation on Lie groups

Garding showed [12] that a representation of a Lie group on a Banach space has a dense set of infinitely differentiable vectors by means of the integral T(a)xp(a)da where p is in CO(G) (i. e., of class C- on G and having compact support). If p is an analytic function on G such that it and all its mixed partial derivatives decrease rapidly enough at infinity, the same integral will give analytic vectors (see [15] and [4]). This argu- ment was used by Gelfand [14] in 1939 to show that a one-parameter group of operators has a dense set of analytic vectors. Gelfand considered only bounded representations (as he was investigating spectral properties which depend on boundedness) but his argument holds in general if we choose a suitable kernel. Let

(8.1) Ptx = 1 | T(a)xe#'2'4t'du

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 593

We must have 1I T(a) 11 _ klecilT' for some k1, c1 < co, so that the integral in (8.1) is absolutely convergent, and Ptx -* x as t -* 0. Furthermore, the integral remains absolutely convergent when the kernel is extended to the complex domain, and from this it follows that Ptx is an analytic vector for T. We will extend this argument to an arbitrary Lie group. The kernel in (8.1) is the fundamental solution of the heat equation au/at = f2u/Oa2 on the line. We shall use the fundamental solution of the heat equation au/at = Au on a Lie group.

Let G be a connected Lie group, X1, *--, X, a basis for its Lie algebra, U the left regular representation of G (given by U(a)f(t) = f(a-'z) for all f in 12(G), where 12(G) is formed with respect to left invariant Haar measure d: on G), A = X2 + ... +X2, and let A be the closure of U(A). Then (by [23]) A is a self-adjoint negative operator. Let Pt be the opera- tor etA, for 0 < t < oo. Any function in the range of Pt is in Z(An) for all n and, since A is elliptic, is equal a. e. to a continuous function (see [24]). By [1] or [20], this implies that Pt is an integral operator with a kernel of Carleman type.

That is, there is a function pt(a, z) such that for all a, pt(a, .) is in V22(G) and

Ptf(u) = |pz(a )f (z)dz G

for all f in 12 (G) and a in G. Since the operator PIt = e tA arises from the left regular representation of G, it commutes with all right translations. That is, for all u in G

Ptf (au) = |p z(a )f (zu)dz = |p z(a -')f (z)0(u-r)dz

where 0 is the modular function. Since we also have

Ptf(au) =| z(aq r)f (z)dz G

for all f in 22(G), pt(a, 2-1)0(u-1) = pt(au, z). Therefore if we denote the function of a single variable pt(a, e) by pt(a), we have that pt(a, ) =

pt(az-1)0(z-1). This gives us

Ptf (a) = | (-)f(z-')dz = pl*f(a)

Since A is real and self-adjoint, pt(a, z) = pt(z, a), so that letting a = en

pt(z-l)0(z l) = pt(z) and pt is in 22(G). Since ptps = pt+s pt*ps = pt+s so that pt is in C(G) (continuous on G and vanishing at infinity).

We call pt the fundamental solution of the heat equation on G. It de- pends on a choice of left versus right and a choice of basis for the Lie

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

594 EDWARD NELSON

algebra. The function pt has the property that pt(a) > 0 for all a in G and |pt(a)da = 1. This follows (e.g., by [16, p. 354, EJ]) from the fact that

s rot~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~- -IIad aife ,a for all X > 0 the resolvent kernel r,(a) = extpt(a)dt satisfies r^(a) ? 0 for all a in G and X r1(a)da=l, as follows easily by the maximum prin- ciple since convolution by r, inverts X- A. (Or we may observe that G.A. Hunt has shown [17] that there is a unique positivity preserving semigroup generated by some extension of U(A) and since U(A) on 02(G) is essentially self-adjoint, Pt must be that semigroup.)

Let p be any left invariant metric giving the topology of G and let N, = {a e G: p(a, e) < r}. For any Borel set B in G, we will write pt(B) for |pt(a)da.

B

LEMMA 8.1. For all t > 0 and c < oo there exists a k < oo such that

(8.2) p8(G - Nr) < ke-C7 for all r ? 0 and s < t.

The proof we give is based on the theory of Markoff processes. As this subject is not widely known, we include a large amout of explanatory material after the proof. A purely analytical proof of the lemma could presumably be constructed, perhaps showing first that the resolvent ker- nel r, decreases faster than any given exponential if X is sufficiently large, but the present proof is conceptually very simple.*

PROOF. We consider the diffusion process starting at e with sub-stochas- tic transition density function e-xtpt(a-lz) (so that Xdt is the probability of the particle being killed in the time interval dt, where X is a positive con- stant). With probability one the sample paths are continuous until the particle is killed, since A is a local operator. Now e-xtpt(G - Nr) is the probability that the particle is in G - N, at time t. Let q(X, t, r) be the probability that the particle is in G - N, at some time s with 0 < s < t9

so that for all s < t and r > 0O

(8.3) e Xsps(G - N,) < q(X, t, r)

Since the particle travels continuously before being killed, to go distance r + r' it must first travel distance r and then travel distance at least r' from that point, by the triangle inequality. Since both the metric and the transition probabilities are invariant under left translations, pt((ua) -1V) =

* Added in proof. Girding has found a simpler proof of the theorem of this section, to appear in Kungl. Fysiografiska Sllskapets i Lund F6rhandlingar, using the fact that Lemma 8.1 is essentially a special case of results proved by M. P. Gaffney in The con- servation property of the heat equation on Riemannian manifolds, Comm. Pure Appl. Math. 12 (1959), 1-11.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 595

pt(a-'z), this implies (using the strong Markoff property) that

(8.4) q(X, t, r + r') _ q(X, t, r)q(X, t, r') . In particular, for all ad 0 and positive integers n, q(X, t, na) _ q(X, t, a)n, and so for all r > 0,

(8.5) q(X, t, r) < q (X, t,[ r]a) < q(X, t, a)(rla)l

Choose a > 0 so that pt(G - Na) > 0 (actually, this is true for any a > 0 such that G-Na is non-empty, since G is assumed to be connected). Then by (8.3), q(X, t, a) > 0 for all X < oo. Now

(8.6) q(X, t, a)- 0 as X --oo

so that for any c <oo, if X is sufficiently large, q(X, t, a)la <e-c. By (8.5), q(X, t, r) < q(X, t, a)-le6 and by (8.3) if we let k = e tq(X, t, a)-1 then (8.2) holds, as was to be proved.

In greater detail, let G be the one point compactification of G by the point 0o, let

H = I!g t < G so that &2 is a compact Hausdorff space in the product topology, and let tt(o) = @(t) for all w in &2 and 0 ? t < oo. Then there is a unique regular Borel probability measure Pr on &2 (see [22]) such that the at are a Mar- koff process (in the sense of Doob [7, p. 80]) with stochastic transition function [7, p. 256]

pa(, E) = exAtpt(u1(E n G)) + (1 - ext)XE(oo) a # 00

pt (oo, E) = XE(oo)

(where E is a Borel set in G and XE is its characteristic function) and with initial probability distribution Pr ({wco: &(co) = e}) = 1. Let A,. be the set of all w) in &2 such that tt(w) is a continuous function of t for all 0 t < 00 except possibly for one value tQo) of t, and such that tt(w) e G for t < t,,(w)) and t(w)=o0 for t>t (w). Then Pr(A,) 1, as the proof of Theorem 3 in [21] shows, or by [25]. Let

F(t, r) = UO?<st {a): 8s(ow) e G - NJr

Since r(t, r) is a union (albeit an uncountable one) of open sets, F(t, r) is open, and q(X, t, r)=Pr(r(t, r)) is well defined. By definition,

etpt(G - Nr) = Pr({o): tt(w) e G - Nr}) and since {cw: t(w) e G - NrJ c P(t, r), (8.3) is clear.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

596 EDWARD NELSON

Let z-,(cw) = inf {t: 58(co) e G - N} if {t: s(w) e G-N } is non-empty, (o)) = 00 otherwise. Then the random variable zr, 0?< z- < c, is a stop-

ping time in the sense of R. M. Blumenthal [3]. Since our process satis- fies the regularity conditions in the hypothesis of Theorem 1.1 of [3], Z.r is a Markoff time, which means that tt(wo) = + (w) is also a Markof process with the same transition probabilities as the original process, and independent of events before the time z. in the sense defined in [3]. Now by the triangle inequality,

,A.. n F(t, r + r') c F(t, r) n n.. n r where F = U0?S?t {wt: 5 )(W) e G, +tsGr(o)) e G, p(Cs+rr(w)(w), 5r(w)(wj)) >r'}.

By the strong Markoff property (Theorem 1.1 of [3]), q(X, t, r + r') < q(X, t, r)Pr(F)/Pr( {i: z-j(w) < oo}) since Pr(F)/Pr({w: z-r(w) < co}) is the conditional probability, given that z-(w)) < oo, of F. (Notice that F c {o) mr() < 00 } since tswr(')(c) e G for all w in F.) By the invariance un- der left translations of the metric p and the stochastic transition function defining Pr, Pr(r) = Pr(r(t, r'))- Pr( {c: z-r(wo)) < ??} ) so that (8.4) holds.

The remaining point which may need clarification is (8.6). First, if s-+o then q(O, s, a) - 0 (in fact, it is shown in [21] that q(O, s, a) =o(s). Thus for any 5>O and t>O, there is an s, with O<s<t, such that q(O, s, a)<e, and since q(X, s, a) is a decreasing function of x, q(X, s, a) < E for all 0 < X< co. Now if co is in A,, and F(t, a) but not r(s, a), then 5(wc) e G. That is, q(X, t, a)<q(X, s, a)+exAsps(G) < s+e-^s. Therefore lim, ,,q(X, t, a)< E for all E > 0, and (8.6) holds.

The function pt(z-) satisfies a parabolic partial differential equation with analytic coefficients, namely (A -ia/at)pt(r) = 0. By results of Eidelman [9], [10, pp. 86-89], or Avner Friedman [11], pt(z-) is an analytic function of Z-, for each t > 0. (Also, for any t >0, p2t = pt * pt is in the range of P' and since A is a negative operator, is an analytic vector for A by the spectral theorem. By Theorem 3, p2t is analytic vector for the left regular representation U. Therefore

P41(,) = p2t * p2l(Z) P2\ prt(T-1)0(U-1)p2 (a)da

S ic p2ti(arl)0(y0(r-)p2t(i-s )0(i(-l)di G

= (ZT )(U(Z.-)p2, p l) = 0(Z- )(U(-)pt p t)

Since p2t is an analytic vector for U, p2t iS in 132(G), and 0(-r-') is an

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 597

analytic function of z-, p4t(z.) is an analytic function of z- for all t >0, i. e., pt(z) is an analytic function of z- for all t>O. This fact also follows from Theorem 8 in ? 12. Let x1, ..., xd be analytic coordinates at e, for E > 0 let Ve be the set of points in G whose coordinates satisfy I xi I < S, .*-, l xd I<e, and let V, be the set of n-tuples of complex numbersZk =Xk+iYk

such that I z, I a, ... , I zI I-< s. We identify Ve with the subset of Vs

consisting of all (z,, ...,Zd) in V, with y1 = -- = Yd= 0. Let t0>0. The estimates used in showing that pt(z-) is analytic are uniform in t > to (for example, in the above deduction of this fact from Theorem 3), so that there is an E > 0 and an M < 0o such that for each t _ to there exists a complex analytic function defined on V,, bounded in absolute value by M, and agreeing with pt on V1. We shall denote this function by the same symbol pt, so that pt(j) is its value for r in V,. Also, if E is sufficiently small then the mapping on Ve x V, given by (T. a)-+z-a extends to a com- plex analytic mapping on V. x V, taking values in Va for some 8 > s. We shall again denote this mapping by juxtaposition: (a, a) a , and shall assume E sufficiently small so that it is defined.

We have now derived a quantitative estimate on the decrease of pt(z) for z near infinity and the qualitative fact that it is an analytic func- tion. We need to combine these two and have quantitative information about the analyticity of pt(z) - the behavior of all its spatial derivatives at infinity. I am grateful to Professor Garding for showing me how to do this.

Let I be an open interval on the real axis, and let u be a function de- fined on I x V, and satisfying the heat equation

(8.7) (A - 2 )u(t, Z) = 0

there. (A subscript on A indicates the variable on which it operates.) We wish to obtain a complex analytic extension of u and obtain a bound for it. The fact that we have a fundamental solution of the heat equation which is analytic in z enables us to do this, as follows.

Let us define ps(a) = 0 if s < 0. Then on R x G, (A, -8/s)ps(a)=8 in the sense of distributions, where 8 is the distribution p -+4(0, e). Therefore convolution on R x G by ps(a) inverts A, - 8/8s. Convolution by ps(a) is integration with respect to the kernel ps- (az(-l)0(z-l). Now the adjoint operator to A, - a/ls is A, + 8/&s, since A, is symmetric. There- fore A, + 8/&s is inverted by integration with the kernel pt-s(Z-w)O(U-'). That is, if p is in CO(R x G) then

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

598 EDWARD NELSON

(8.8) (Acr + a p tS(z-r)0(au-)9(t, z)dtdz = p(a, s)

Let I' be an open subinterval of I whose closure is contained in I, let 0 <as' < a, and let k be in CO ( I x Ve) and be identically 1 on I' x V,. Then for all (t, z) in I' x 1,8, if u satisfies (8.7),

(8.9) u(t, r) = | |u(s, a)( A, + a ){Pt (z1)O(a)(1- (s, a))} dads.

This is an instance of Garding's formula (3) in [13]. It is proved by com- puting the inner product of both sides of (8.9) with an arbitrary function p in CO (I' x V1,), using (8.8). Now the integrand of (8.9) is 0 except on I x V-I' X 18 V because 1 - *(s, a) = 0 on I' x V1,. (Garding's for- mula is a " smeared-out " version of the Poisson formula, expressing a solution of the partial differential equation in terms of its values in the neighborbood of the boundary.) Let I" consists of all t in I' whose dis- tance from the complement of I' is at least to. Then if t is in I" only values of s with t - s > to contribute to (8.9). Hence there is a constant M' such that for all t in I" and ? in Ve,

(az. + 8 ){pta-S(-)0(a-1)(1 - *(S, a))} < M'.

If u is a solution of (8.7) then (8.9) with z replaced by , defines a complex analytic extension of u(t, *) to V, for all t in I ", such that

Ju(t, ')l <j MI u(s, a) Ida ds.

Any right translate pt(za) of pt(z) satifies (8.7) since A, commutes with all right translations. Therefore for any fixed t > 0 if 0 < t1 < t there is an M' < co and an s > 0 such that for all a in G, pt(za) has a complex analytic extension pt(?a) defined for r in V1 and satisfying

(8.10) 1 PI (ia) ? I ps(aau)dal ds )

(Notice that we have not defined the expression ?a for arbitrary a in G, but merely use the notation pt(?a) for the extension of pt(za) to T in V8.)

LEMMA 8.2. Let t > 0 and c < oo. Then there exists an e > 0 and a k < co such that for all a in G, ptQ(ia) is a complex analytic function of

i in V, agreeing with pt(zau) for T in 17 and satisfying

I pt(Ta) I da < ke-cr G -Nr

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 599

for all T in Ve and r > 0.

PROOF. Let E be sufficiently small so that (8.10) holds, Ve1 c N1, and the volume of VI is less than 1. Then a2 = u^u is in G - N_ if a, is in V. and a is in G-Nr, since r< p(e, a)< p(e, a1 )+p(a7', a) < 1 + p(e, a, a). We have

i pt(;T) I du < M' ps(ula)dql ds du G-Nr G-N r ti Vs

< M' | P | (a2)da2 < M'(t-t1)kececr ti G-N -

by Lemma 8.1, which proves the lemma. We shall now assume that the metric p is the geodesic metric of some

left-invariant Riemannian metric on G. Such a metric has the property that N1 is compact and if a is in Nr then there exist a1, .2.... 96[r+1] in N1 with 162 ... [r+1] = ?

LEMMA 8.3. Let T be a representation of a Lie group G on a Banach space X. Then there exist k1, cl < oo such that for all r>0 and a in Nr,

(8.11) 11 T(a) 11 < klecir

PROOF. We have II T(a) II < I T(a1) T(a2) ... T(U[r+l) 11 < a[r+l] where a = SUPEN, 11 T(a) II. Since T is strongly continuous and N1 is compact,

supEN, 1I T(a)x 11 < for all x in X. The principle of uniform bounded-

ness [16, p. 26] states that a< oo. Letting k,= a, el=log a we have (8.11).

THEOREM 4. Let T be a representation of a Lie group G on a Banach space X. Then T has a dense set of analytic vectors in X.

Let pt(a) be the fundamental solution of the heat equation on the con- nected component of the identity Go, pt(a) = 0 for a not in G,. For all x in I and t > 0

(8.12) P'tx= Gpt(u)T(u)xdu

exists and is an analytic vector for T. As t -O 0, PtX - x. PROOF. To show that the integral in (8.12) exists, observe that by Lem-

mas 8.1 and 8.3,

Aupt() II T(a)x 11 da < E = pt(a) 1 T(a) II 11 x II da

-< E n ke-c(n+')kle cln 1

which is finite if we choose c > cl. Since this also holds for all s < t, by Lemma 8.1, we have by the Lebesgue dominated convergence theorem

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

600 EDWARD NELSON

that for all r > 0, limpe ps(a) 11 T(a)x 11 du = 0, so that P3 x -* x as G-Nr

s -+ 0. In the same way, if t>0 then by Lemmas 8.2 and 8.3, for all r in a neighborhood Ke, and for all x in X, the integral

(8.13) |\'G )T(u)xdu G

is absolutely convergent. Since pt(jTu) is complex analytic in A, (8.13) is complex analytic in r. But for = zr in G, (8.13) is equal to T(-r')Ptx, so that z-* T(zr-)Ptx is analytic for r in a neigborhood Ve of e. Since

-* zr-1 is analytic, z -* T(zc)Ptx is analytic in a neighborhood of e, and so Ptx is an analytic vector for T, concluding the proof.

Our method of proving that a group of operators has a dense set of analytic vectors has used as an intermediary a one-parameter semigroup of operators, the one associated with the heat equation. E. Hille and R.S. Phillips raise the question in [16, p. 310, (p. 229 in the first edition)] as to whether every strongly continuous one-parameter semigroup of operators has a dense set of analytic vectors. The following example answers this question in the negative (cf., Lemma 5.1).

Define Tt for0 < t <oo, on 2(0, oo)by

Ttu(x) = u(x - t), x > t

Ttu(x) = 0, 0 <?x < t.

Thus Tt is simply translation to the right by t, and is a strongly continu- ous semigroup. Suppose that u is an analytic vector for Tt and let v be in 22(0, cI). Then (Ttu, v) is an analytic function of t for 0 < t < cA. If v has compact support then (Ttu, v) = 0 for t sufficiently large, and so (Ttu, v) = 0 for all t, (u, v) = 0, and so u must be 0. That is, the only analytic vector for this semigroup is 0.

9. Lie algebras of operators

Suppose we have a representation of a Lie algebra g by skew-symmet- ric operators defined on a common invariant domain in a Hilbert space if, and let G be the simply connected Lie group with Lie algebra q. In this section we shall give an answer to the question: When does the represen- tation of g come from a unitary representation of G ?

This question first arose in the special case of skew-symmetric opera- tors P, Q satisfying the commutation relation PQ - QP c i (or for a finite number n of degrees of freedom PjQk - GkPj c ak ia, for 1 < j, k < n). Both the matrices of Heisenberg and the Schroedinger operators P= d/dx and Q = multiplication by i x on 532(R) satisfy this relation, and

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 601

Dirac showed their equivalence. The example of d/dx and ix on 22(0, 1) shows that some condition on the operators is required. Weyl [31] ex- pressed the commutation relation in a bounded form equivalent to postu- lating a unitary representation of the Lie group G whose Lie algebra has a basis X, Y, Z satisfying [X, Y] = Z, [X, Z] = O [Y, Z] = 0, and von Neumann showed that there is only one irreducible unitary representation (up to unitary equivalence) of G in which Z goes into i. The problem of finding purely infinitesimal conditions of P and Q was solved by Rellich [26] who showed that if the energy operator P2 + Q2 is essentially self- adjoint then P and Q are what they should be. Theorem 5 is in a way an extension of Rellich's result to arbitrary Lie algebras, but the method of proof is quite different. Also, Rellich made the stronger assumption that the range of every E(QI), where E(.) is the solution of the identity for P2 + Q2 and 4) is a bounded interval, is contained in the (invariant) domain of P and Q.*

First we describe the connection with analytic vectors.

LEMMA 9.1. Let g be a Lie algebra of skew-symmetric operators on a Hilbert space & having a common invariant domain Z. Let X1, ... X, be a basis for g, = IX I+ - + IXi. Ifforsomes >0 the set of vectors x in Z such that 11 eO8x ii < co is dense in A, then there is on & a unique unitary representation U of the simply connected Lie group G having g as its Lie algebra such that for all X in g, U(X) =X.

PROOF. By Lemma 5.1, if X is in g then iX is self-adjoint, since any analytic vector for t is an analytic vector for X. Let exp be the exponential mapping in the sense of the theory of Lie groups [5] and let N be a neigh- borhood of e in G such that exp is one-to-one from a neighborhood of 0 in g to N. For a = exp X in N, define U(a) to be the unitary operator en.

Let X, Y, and Z be in g and suppose that (exp X) (exp Y) = exp Z in G. Then [5, p. 121]the two power series E' , (1/n!)Zn and Ek , (1/k!l!)XkY1 are formally equal. Consequently if x is a vector such that IleIxI +IYIxII < cc,

Ie 1ZIX 11 < oc then U(exp X) U(exp Y)x = U(exp Z)x. Now for X, Y. and Z sufficiently close to 0 in g (such that the absolute values of their coordinates with respect to the basis X1,... ,X, are less than (1/2)s) there is by hypothesis a dense set of vectors x such that 1I eIxI+{Y Ix I < cc,

I1 eIx 1 < co. Therefore, if a and r are sufficiently close to e in G, U(a)U(z) = U(az). That is, U in a neighborhood of e defines a unitary representation of the local group. (Strong continuity follows from the

* Added in proof. This restriction was removed by J. Dixmier, Sur la relation i(PQ -QP) = 1, Comp. Math. 13 (1958), 263-269.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

602 EDWARD NELSON

fact that if Ox < oo then exx - x 0 as X-* 0 in g, and it is suf- ficient to verify strong continuity on a dense set, since U(a) is uniformly bounded, 11 U(a) 11 = 1.) Since G is simply connected, there is a unique extension of U to be a unitary representation of G on ?&, concluding the proof.

This lemma is not a strict analogue of Lemma 5.1, since in this lemma we assume that for some fixed s > 0, - {x:1I eIOx j < co } is dense. In Lemma 5.1 we did not need to assume this because each .1 x: e I ex I sx 11 < co I was invariant under U(t), using the fact that the various powers of X all commute. I do not know whether Lemma 9.1 remains true if we merely assume the existence of a dense set of analytic vectors for 0, but we shall not need such a result.

THEOREM 5. Let g be a Lie algebra of skew-symmetric operators on a Hilbert space ? having a common invariant domain Z. Let X1, X, ,

be a basis for g, \ = Xl +... * +X2. If A is essentially self-adjoint then there is on & a unique unitary representation U of the simply connected Lie group G having gas its Lie algebra such that for all X in g, U(X) = X.

PROOF. Let e = IX1 I + ?-- + IX, l. By Lemma 6.2, 1A I+II Ianalyt- ically dominates I. The theorem follows by Lemma 5.2 and Lemma 9.1.

In Theorem 5 we assumed that Z was invariant under g and construct- ed a group representation by means of analytic vectors. Roughly speak- ing, the theorem concerns the passage from Co to C". The following refinement of the theorem makes only C2 assumptions, so to speak.

COROLLARY 9.1. Let g be a real Lie algebra, ?Q a Hilbert space. For each X in a let p(X) be a skew-symmetric operator on &. Let Z be a dense linear subspace of S& such that for all X, Y in g, Z is contained in the domain of p(X)p( Y). Suppose that for all X, Y in g, x in Z, and real numbers a and b,

p(aX + b Y)x = ap(X)x + bp( Y)x p([X, Y])x = (p(X)p(Y) - p(Y)p(X))x

Let X1, - ,X, be a basis for g. If the restriction A of p(X1)2 + -* + p(Xd)2

to Z is essentially self-adjoint, then there is on & a unique unitary rep- resentation U of the simply connected Lie group G having g as its Lie algebra such that for all X in g, U(X) = p(X).

PROOF. First, for each n, Z(An) c ..(p(X.) p(Xn)). The proof of this exactly parallels the proof of (5.4) in Lemma 5.2 except that in- stead of considering (ad p(X))A, which might have only 0 in its domain, we consider p((ad X)A), where A = Xf + * +X2 in the universal envel-

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 603

oping algebra of g. This is well-defined on A, since (ad X)A is a linear combination of the X{X. Define n = Z9(An), A the restriction of A to X, and X, the restriction of p(X;) to Z. Then we are in the situation of Theorem 5, and the corollary follows.

Two unbounded self-adjoint operators are said to commute (or be per- mutable [30]) in case their spectral resolutions commute. An operator C = A + iB with A and B self-adjoint is called normal in case A and B commute. The next corollary is a criterion for normality.

COROLLARY 9.2. Let A and B be symmetric operators on a Hilbert space 9 and let Z be a dense linear subspace of & such that Z is contained in the domain of A, B, A2, AB, BA, and B2, and such that ABx = BAx for all x in Z. If the restricton of A2 + B2 to Z is essentially self-ad- joint then A and B are essentially self-adjoint and A and B commute.

PROOF. This follows from Corollary 9.1 with g the two-dimensional abelian Lie algebra with basis X, Y and p(aX + b Y) = i(aA + bB).

That some such condition is required is shown by the counter-example in the following section. Also, Theorem 2 of [23] is a converse of Theorem 5, except that it is necessary to assume that the operators have a domain invariant under the group representation. In fact, the following corollary gives necessary and sufficient conditions for a representation of a Lie al- gebra by skew-symmetric operators to be the infinitesimal representation associated with a group representation.

COROLLARY 9.3. Let g be a real Lie algebra with a basis X1, ...* X,

G the simply connected Lie group with Lie algebra g, $) a Hilbert space, @ a dense linear subspace of ?. Let p be a representation of g by skew- symmetric operators with domain e. Then there is a unitary represen- tation U of G such that (Y is the space of infinitely differentiable vectors for U and U(X) = p(X) for all X in g if and only if

A = p(X1)2 + . . . + p(X,,),

is essentially self-adjoint and (Y = nflf 5(A )

PROOF. By Theorem 5 and [23], we need only show that if U is a unitary representation of G and A U(X1)2 + * + U(X,)2 then cW= nflf (An). The inclusion c is obvious, since e is invariant under A, and the reverse inclusion follows by (5.4) of Lemma 5.2.

10. A counter-example

Let g be the non-abelian three dimensional nilpotent Lie algebra, with

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

604 EDWARD NELSON

basis X, Y, Z such that [X, Y] = Z. Let '? = 22(0, 1). Then a represen- tation of g by skew-symmetric operators is given by

X ,d. dx

(10.1) Y ,x Z it

all of these oserators having C (0, 1) as domain. (By ix we mean the oper- ator u(x) -+ ixu(x).) It is clear that this representation does not give rise to a group representation. One shortcoming of this as a counter ex- ample is that d/dx is not essentially skew-adjoint on C (0, 1). Let us now consider a different representation of g, also given by (10.1) but with do- main for all of the operators being now the set of Co functions u on [0, 1] with u(0) = u(1). Then d/dx on this domain is essentially skew adjoint but, of course, the representation still does not give rise to a group rep- resentation. The shortcoming of this as a counter-example is that the domain is not invariant. In fact, the set of functions u for which the commutation relation (d/dx)(ixu) - ix(d/dx)u = iu has meaning is so small that d/dx is no longer essentially skew-adjoint on it. Other examples of representations of Lie algebras which do not generate group representa- tions are given in [28].

In this section we give a counter-example having both the features of essential skew-adjointness and invariance of the domain. It appears to be the first example of this nature. The Lie algebra in question is the two dimensional abelian Lie algebra, so it is also an example of two essential- ly self-adjoint operators defined on a common invariant domain which commute on that domain but whose special resolutions do not commute.

Suppose we have a Co- vector field X on a Co- manifold M. Then X gives rise to a flow x -* Ttx on M obtained by integrating X, except that each T, may be only partially defined, since the trajectories obtained by integrating X may lead outside of M. For each t, let Et denote the closure of the set where T, is not defined.

LEMMA 10.1. Let M be a Co- manifold, X a C? -vector field on M. Let 22(M) be formed with respect to a non-vanishing Co exterior d-form (where d is the dimension of M) and suppose that X is skew-symmetric on the domain CO(M). Let Et be defined as above. If for some t > 0 the sets Et and Ea have measure 0 then X is essentially skew-adjoint.

The differentiability assumptions could, of course, be relaxed. Also, it may be shown that the converse of the lemma is true (if X is essentially

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 605

skew-adjoint, then for all real t, Et has measure 0). PROOF. For all real t, define U(t) by

U(t)u(x) = u(TM(x)), x 0 Et U(t)u(x) = ?, x e Et

Since X is a skew-symmetric vector field, T, is a measure-preserving map of M - Et into M, and so for all t

(10.2) 11 U(t)u 11 < 11 U 11, u e 22(M)

Any function in 22(M) which is orthogonal to C-(M - E) must vanish a. e. on M - Et, so that if Et has measure 0 then C-(M - E) is dense in 22(M).

To show that X is essentially skew-adjoint it suffices to show that if X*u = X or X*ui= -u then u=0 [30, Chap. IX]. Suppose that X*u =u and let t > 0 be such that Et has measure 0. Notice that for 0 < s < t, U(s) maps C0 (M - E) into C0 (M), since X is a Co- vector field. Now if X*u = u, q is in C-(M - E), and 0 < s < t then

d (U(s)p, u) = (XU(s)p, u) = (U(s)p, X*u) = (U(s)p, U) ds

Therefore (U(s)p, u) = (p, u)es for 0 < s < t. Since C-(M - E) is dense in 22(M), 11 U(s)*u I = e 1 u 11, and u = 0 by (10.2). The analogous proof holds in case X*u =- u, by choosing t < 0. This concludes the proof.

Let M1 be the set of all points in the xy-plane such that 0 < x < 3 and 0<y<3 but not both 1<x<2andl<y<2. That is, M1 is the closed band between two concentric squares. Let M2 be the set M. minus the points with integral coordinates at least one of which is 1 or 2. That is, we remove the four vertices of the inside square and the eight points on the sides of the outside square facing them. Let C be the set of points in both M2 and the boundary of M2. Each point z in C except (0, 0), (0, 3), (3, 0), (3, 3) has a unique opposite point z' in C, namely the point z' in C such that the straight line segment joining z and z' is either horizontal or vertical and lies entirely in M2. Note that (z')' = z. Let M be the Co- manifold obtained by identifying opposite points of C in M2 and by identifying (0, 0), (0, 3), (3, 0), (3, 3). (It may be verified that M is the closed surface of genus 4 with one point removed, which we have given a flat Riemannian metric.)

Now let g consist of all operators a(&/8x) + b(&/8y), where a and b are real constants, with domain C-(M). It is easily seen that each X in g is a Co vector field on M. Let 22(M) be formed with respect to Lebesgue measure. Then g is a Lie algebra (abelian) of skew-symmetric operators

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

606 EDWARD NELSON

having CO (M) as common invariant domain. For each X = a(&/8x) + b(8/8y) in g and real t, Et consists of a finite number of line segments (in fact seven, if t is sufficiently small, of slope b/a and leading into points of M- M2). Therefore Et has measure 0 and by Lemma 10.1 each X in g is essentially skew-adjoint. However, g clearly does not give rise to a group representation, for let X be the closure of 8/8x, Y the closure of 8/8y. If u has support in the lower left hand square 0 < x < 1, 0 ?y? 1 then exeYu has support in the square 0 < x < 1, 1 < y < 2 above it whereas eYexu has support in the square 1 < x < 2, 0 < y < 1 to its right, so that eley # eyex.

We have proved the following result: There exist two symmetric operators A and B on a Hilbert space &

having a common invariant domain Z such that for all real a and b, aA + bB is essentially self-adjoint and such that for all x in Z, ABx = BAx, but such that the spectral resolutions of A and B do not commute. The set of all i(aA + bB) is a Lie algebra of skew-symmetric operators having a common invariant domain, such that each operator is essential- ly skew-adjoint but not giving rise to a unitary group representation.

11. An application to a paper of Dixmier

The purpose of this section is to show that our method yields some use- ful information other than the mere existence of a dense set of analytic vectors for a unitary representation. This will be done by giving a more direct proof of a lemma of Dixmier [6]. The application is based on the connection between analytic vectors and the operator A.

THEOREM 6. Let G be a simply connected Lie group with Lie algebra g. Let g' be an ideal in a, G' the connected subgroup of G with Lie algebra g'. Let U' be a unitary representation of G' on a Hilbert space &. Let r be a homomorphism of the universal enveloping algebra CU(g) of g onto the image under U' of the universal enveloping algebra cU(g,) of g', such that r(B') = U'(B ') for all B' in cU(g') (in the natural identification of LU(g') as a subalgebra of C7U(g)) and such that r(X) is skew-symmetric for each X in g. Then there is on & a unique unitary representation U of G such that U(u') = U'(W') for all a' in G'.

PROOF. Let X1, ..., X, be a basis for g such that X,,---, X, is a basis for W'. Let A' = X1+... +XX in cU(g'). Let (i' be the space of infinitely differentiable vectors for the representation U'. Let Y = I r(X1) I+ * + I r(X,) I and a = I U'(A') - II. Notice that r(X,) = U'(X,) for 1 < j < d'. Let m be a positive integer such that r(X,) is of order < 2m for 1 < j < d.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 607

Since q' is an ideal in g, cU(g') is invariant under ad X for all. X in g. Therefore for all X in g and B' in cU(g') - in particular for B' = (A' - -M

we have r(X)U'(B') - U'(B')r(X) = r(XB' - B'X) = U'(XB' - X), and this in an element of order not greater than the order of B'. There- fore Lemma 6.3 applies, and atm analytically dominates Y). By [23], (U'(A') -I)m is essentially self-adjoint and by Lemma 5.2 and Lemma 9.1 there is a unique unitary representation U of G such that U(X) = r(X), the representation U agreeing with U' on G'. This concludes the proof. Since e' = nl j(A' - I)mn), by Lemma 5.2, this shows also that E' is the set of infinitely differentiable vectors for U.

Precisely the situation described in Theorem 6 arises in a paper [6] (es- pecially pp. 347-352) of Dixmier. In this case G is a simply connected nilpotent Lie group, g' is an ideal of codimension 1 in g, and, letting X be in g but not g', there is a non-zero element A1 in the center 2(g') of cU(gt) and an element A2 in cU(gP) such that each of A1, A2, XA1 + A2 is either symmetric or skew-symmetric and XA1+A2 is in the center B(g) of U(g). By a Hermitian character of 2(g) is meant a homomorphism of 2(g) into the complex numbers sending symmetric elements into real numbers and skew-symmetric elements into purely imaginary numbers. If U is an ir- reducible unitary representation then A -+ U(A), for A in 2(g), is a Her- mitian character. By A is meant the set of Hermitian characters of B(g') corresponding to a unique (up to unitary equivalence) irreducible unitary representation of G'. It is required to prove that every Hermitian charac- ter of 2(g) whose restriction to B(g') is in A and which is non-zero on A, corresponds to a unique (up to unitary equivalence) irreducible represen- tation of G, and that its restriction to G' is also irreducible (see [6, Lem- ma 21]). To see this, let X be the character, X' its restriction to 2(g') (which is contained in 2(g) in the case under consideration), and U' an irreducible unitary representation of GI with character X'. This exists since X' is in A. Define a homomorphism r of CU(g) by letting it be U' on U(g') and setting

r(X) = X(XA1 + A2) -U'(A2) X'(A1)

By Theorem 6 there is a unique unitary representation U of G extending U'. The mapping r is so defined that U has the character X (since the field of fractions of the ring generated by XA,+A2 and 2(g') is the same as the field of fractions of 2(g); see [6]). Since U' is irreducible, U is a fortiori irreducible. Now let U be any irreducible unitary representation of G, U' its restriction to G'. If U' is not irreducible, let &1 and &: be two

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

608 EDWARD NELSON

closed invariant subspaces which are orthogonal complements to each other, and consider the representations U' and U' of G' on ,1_ and , Again by Theorem 6, each of these extends to a representation U1 and U2 of G, with character X. Thus U and the direct sum of U1 and U2 are both ex- tensions of U', and by uniqueness they must be equal. Therefore the ir- reducibility of U implies the irreducibility of U'.

12. Elliptic partial differential operators

Suppose that A is an elliptic partial differential operator with analytic coefficients, and let u be a solution of the equation Au = 0. Then it is a classical result that u is analytic (see F. John's book [18] and the referenc- es there). The most general result of this type is due to Morrey and Nirenberg [19, ? 5], who show that a solution of a general elliptic system (in the sense of Douglis and Nirenberg [8]) with analytic coefficients is analytic. Here we shall show more generally that if u is an analytic vector for A then u is analytic. For simplicity we consider only single operators rather than systems. This result is then applied to self-adjoint A to show that if f decreases exponentially fast then the operator f(A) has an analytic kernel.

Let U be an open set in RI, K a compact subset of U which is the clos- ure of its interior. Let A be a partial differential operator on U. We write A as A =-a,(x)DP, where p ranges over finite sequences P1i,..P of integers between 1 and d, DP = 8/8xp, ... 8/8xp , and the coefficients ap are invariant under all permutations of the indices in p. Letting Ip = ik be the length of p, the operator A is of order < m in case aps 0 whenever I p I >m. The characteristic polynomial a(x, a) of A, if A has order m, is defined for each x in Uby a(x, t)= IPI_ ajx)&P where 4==(4, ..., Y) is a d-tuple of real numbers and dP - = .- - pk. The operator A is called elliptic in case a(x, a) 0 0 for 0 + o.

We use 11 11 for the norm in 22(U), 1 XI 112 uI (x) 12 dx, and I1 I1 for U

the norm in 2(K), 11 U IK u(x) 2dx. Let A be an elliptic operator of

order m with C- coefficients on U. As the domain of A we take all func- tions u in 22(U) such that the distribution Au is also in 22(U). In other words, if A+ is the operator with domain C-(U) given by the formal ad- joint E Ip em (-1)IPIDP(aup(x) * ) of A, then A = (A+)*. It is known (see [24, Theorem 1 of ? 4]) that if u is in the domain of A then 11 DPu II < co for all I p I < m. By the closed graph theorem, therefore, there is a k < 0o such that for all u in the domain of A,

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 609

(12.1) 1 DPu IlK < k (11 Au 11 + 11 u

(Actually, the reference to the closed graph theorem is not necessary, since the value of k may be derived from the proof that II DPu II< o? for I p i < m.) If B is an operator of order < m, EIpi~m bp(x)DP, let

(12.2) III B III = supxeK pi b,(x) Then it follows at once from (12.1) that

(12.3) 11 BuIIK< k III B IBII(11 Au 11 + 11 u 11) . Let Z8Dm be the Banach space of all partial differential operators of order

< m with coefficients in C(K), with the norm (12.2). We say that a func- tion is in C-(K) if it has a Co extension to a neighborhood of K, and similarly that it is in C"(K) in case it has an analytic extension to a neigh- borhood of K. If X is a vector field with coefficients in C -(K) we define the operator adX on the Banach space Zm to have as domain all elements B of Z8m which have coefficients in C-(K), with (ad X)B = XB - BX. Notice that ad Xad Y - ad Yad X = ad(XY - YX) (the Jacobi identity).

LEMMA 12.1. Let X1, ... , Xl be vectorfields with coefficients in C@(K),

let 4= I ad X1 1+ - * + I+ad X, 1, and let 8=Iad (a/8x1) 1 +* - * +I ad (8/8Ix ). Then 8 + I I analytically dominates a. If A has coefficients in C-(K)

A

then A is an analytic vector for a. PROOF. The operator (ad (a/8x)) A is merely the operator obtained from

A by applying 8/8xi to its coefficients. (This is not true for the general vector field X, since derivatives of the coefficients of X occur in (ad X)A.) Since K is compact, to say that A has coefficients in C"(K) is equivalent to saying that A is an analytic vector for 8 + I I Therefore the last statement of the lemma follows from the preceding one by Theorem 1, and we need only show that 8 + I I I analytically dominates t.

Let c, = (SlnX, 111 so that Cn is the sum from i = 1 to i = 1 of the largest absolute value on K of derivatives of order n of the coefficients of Xi. Since the Xi have coefficients in C@1(K), the power series with coef- ficients Cn/n! has a positive radius of convergence.

For a certain integer r (depending only on m and d) we have for all vector fields X with coefficients in C-(K),

(12.4) lad X < r 80X 11 (8 + I I)

since the coefficients of (ad X)B (for any B in ZDm with coefficients in C-(K)) involve only the coefficients of X times first order derivatives of the coefficients of B and the coefficients of B times derivatives of order

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

610 EDWARD NELSON

< m (since B has order m) of the coefficients of X. By (12.4), A A

(12.5) < r(co + * +c.f) (8 + iIi

By Corollary 4.1, letting C = 8, (ad );)n8 is the sum of lnd terms of the form I X I with E III 8X III < k', where the power series with coeffi- cients k./n! has a positive radius of convergence. By (12.4)(and the Ja- cobi identity) this gives

(12.6) (ad t)n 8 ?< lndrkn(8 + I I 1) By (12.5) and (12.6), 8 + I I I analytically dominates 0, concluding the proof.

THEOREM 7. Let U be an open set in Rd, A an elliptic partial differen- tial operator with analytic coefficients in U. Let u be a function of class Co on U such that

(12.7) En50 sIA!1 n < CO

for some s > 0, where 11 V 112 = v(x) 12 dx. Then u is an analytic func- tion in U.

For each s > 0 there is an open set U, in the complex space Cd con- taining U and such that all functions u of class Co on U satisfying (12.7) have a complex analytic extension to U,.

We may remark that the assumption that u is Co is unnecessary, since by the regularity theorem any function in the domain of An for all n is automatically a Co function.

PROOF. Let x0 be a point in U. By an affine change of coordinates we may assume that x0 = 0 and that the sphere of radius 2 and center 0 is contained in U. We choose the closed unit sphere for the compact set K contained in U. Let m be the order of A and let

(1 (X2 + *-+X2))M~l$ K (12.8) q(x) =

0, x 0 K.

Although there is no such thing as a non-zero analytic function with compact support, the function q will behave like one for all our purposes. Notice that the restriction of p to K is in C@(K), so that if we define X, = p(x) (/8x,), --**, XI =p(x)(&/8x,) then Lemma 12.1 applies to them.

Let t = I X1 + *** -+ l XI 1. We shall show that any u satisfying (12.7) is an analytic vector for t. To do this it is enough, by Theorem 1, to show that a = A + I analytically dominates I. By (12.3) there is a

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 611

c < o0 (c = k ~lll Xi IIl) such that

(12.9) < ? ca. By Lemma 12.1 there is a sequence c,,, such that the power series with coefficients cu/n! has a positive radius of convergence and such that IInA II < cn.

Let CV be the set of all partial differential operators B = lpl<S, bp(x)DP of order < m defined on U such that the coefficients bp are continuous, vanish outside K, and such that their restrictions to K are in C (K). No- tice that

(adX) A =lipjim (- apDP - ap(DP -

is in CV since ( vanishes to the order m + 1 on the boundary of K. For the same reason, if B is in CV then (ad XJ)B is in CV, i = 1, - - *, d. By induction, therefore, each ad Xin * * * ad X%1A is in CV. Consequently I Iad X, ad X1 Au IlK =1 ad X *,* * ad XjAu I1. By (12.3), there- fore

11 ad Xi * ... ad XjAu 11 < k Ill ad Xin ... ad Xi1Alll(ll Au 11 + 11 u 11)

so that (12.10) (ad t)na < kcna

By (12.9) and (12.10), a analytically dominates #, and any u satisfying (12.7) is an analytic vector for 4.

Let K, be the closed sphere of radius 1/2 and center 0. Since K, is con- tained in the interior of K it is obvious (and also a special case of Theorem 2) that any analytic vector for # is an analytic vector for

iRoxl i~l --+Ex,z

with respect to V2(Ko). That is, letting 11 V2 I V(X) 12 dx,

(12.11) II11 D u 1,K sn <co

for some s0>0. By a theorem of Sobolev (see [24, p. 655]) there is a con- stant ko < oo such that

(12.12) SUPXEK01 v(x) I - koE P1 p[d/2]+1 1 D~v IIKO Applying (12.12) to v = DPu in (12.11), we have that

E supeKlI D u(x) I (12.13) SpxEO 1 s I!c

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

612 EDWARD NELSON

for some s, > 0, so that u is indeed an analytic function. The operator A being fixed, the value of s, in (12.13) depends only on the s in (12.7). However, any u satisfying (12.13) has a complex analytic extension to the set of all z in Cd such that the distance from z to Ko is less than s,, prov- ing the last statement of the theorem. This concludes the proof.

THEOREM 8. Let M be a paracompact real analytic manifold, a a measure on M which in each analytic local coordinate system has a non- vanishing analytic density. Let A be a self-adjoint elliptic partial dif- ferential operator on 22(M, pa) with analytic coefficients. Let f be a Baire function of a real variable such that for some k < co and c >0, If(X) IV ke-"I, for all X in the spectrum of A. Then there is a unique analytic function F on Mx M such that for all x in M, F(x, *) is in 22(M, pa) and for all u in 22(M, ,a)

(12.14) f(A)u(x) = 5F(x, y)u(y)dpa(y) M

Any element of 22(M, a) in the range of f(A) is equal a. e. to an analyt- ic function. The mapping x-+F(x, *) is analytic from M into V2(M, ia). There is a neighborhood M of M in a complex analytic manifold contain- ing M and an extension F(', P) of F to M x M which is complex analyt- ic in x and i, and such that the mapping x -+ F(x, *) is complex an-

?2(M alytic from M into 22(M, ia). PROOF. The existence and uniqueness of a continuous function F hav-

ing these properties is proved, for example, in [20], using the regularity theorem for elliptic operators (see [24]). We must prove that F is analyt- ic.

By the spectral theorem, any element of 22(M, [) in the range of f(A) is an analytic vector for A. By the regularity theorem (see [24]) any ele- ment of 22(M, p) in the domain of An for all n (in particular, any element in the range of f(A)) is of class Coo. (We identify an element of 22(M, s), which is an equivalence class of functions equal a.e., with a continuous representative of it, which must be unique if it exists.)

Let U be an analytic coordinate neighborhood in M whose closure is contained in an analytic coordinate neighborhood. Then di on U is given by dye- r(x)dx1 *-- dxn where * is bounded away from 0. Therefore '2( U, a) and 2 2(U)(formed with respect to Lebesgue measure) are the same spaces, with equivalent norms. Consequently, any u which is an analytic vector for A in ??2(M, [) is an analytic vector for A in 22( U). By Theorem 7, therefore, such a u is an analytic function. In particular, any element of 22(M, A) in the range of f(A) is an analytic function.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 613

By the last statement of Theorem 7, if M is a sufficiently small neighbor- hood of M in a complex analytic manifold containing M (which exists by [32, p. 133]) then any element of 22(M, M) in the range of f(A) has a com- plex analytic extension to M. If x is in M then u-+ f(A)u(x) is an every- where defined linear functional on 22(M, s), and it is easily seen to be con- tinuous (either by the closed graph theorem as in [1] or by the fact that all our estimates depend only on 11 u 21.2). Therefore there is a unique ele- ment F(x, *) of 22(M, 4a) such that

(12.15) f(A)u(x) = | F , y)u(y)di(y). M

Comparing (12.14) with (12.15) we see that F(x, y) is (as the notation suggests) an extension of F(x, y). Since each f(A)u(x) is a complex analyt- ic function, the map x * F(x, - ) from M to 22(M, a) is complex analytic (by [16, p. 96]). Since F(y, x) is the kernel of f(A), the same argument shows that there is a complex analytic mapping y F(-, a) of M into 22(M, ~A) extending y F(., y).

Let g(X) = f(X)2 and let G be the kernel of g(A), so that

G(x, y) == |F(x, z)F(z, y)dpa(z) = (F(x, *), F( -, y)) M

Let us define G(x, y) = (F(x, * ), F(., ii)). Then G(x, ji) is an extension of G(x, y) to M x M which is complex analytic in x and P separately, and hence [2, p. 33] jointly. Consequently G(x, y) is a real analytic function of x and y. But the function f is itself the square of a Baire function which is exponentially decreasing on the spectrum of A, so that the same argument applies to F(x, y), so that F(x, y) is a real analytic function of x and y. This concludes the proof.

INSTITUTE FOR ADVANCED STUDY

REFERENCES

1. W. G. BADE AND J.T. SCHWARTZ, On Mautner's eigenfunction expansions, Proc. Nat. Acad. Sci. USA, 42 (1956), 519-525.

2. SALOMON BOCHNER AND WILLIAM TED MARTIN, Several Complex Variables, Princeton

University Press, 1948. 3. R. M. BLUMENTHAL, An extended Markov property, Trans. Amer. Math. Soc. 85

(1957), 52-72. 4. P. CARTIER ET J. DIXMIER, Vecteurs analytiques dans les representations de groupes

de Lie, Amer. J. Math. 80 (1958), 131-145. 5. CLAUDE CHEVALLEY, Theory of Lie Groups, I, Princeton University Press, 1946. 6. J. DIXMIER, Sur les representations unitaires des groupes de Lie nilpotents, II, Bull.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

614 EDWARD NELSON

Soc. Math. France 85 (1957), 325-388. 7. J. L. DOOB, Stochastic Processes, Wiley, New York, 1953. 8. AVRON DOUGLIS AND LouIs NIRENBERG, Interior estimates for elliptic systems of

partial differential equations, Comm. Pure Appl. Math. 8 (1955), 503-538.

9. S. D. EIDELMAN, On analyticity of solutions of parabolic systems, Dokl. Akad. Nauk SSSR, N. S. 103 (1955), 27-30.

10. , On fundamental solutions of parabolic systems, Mat. Sb. N. S. 38 (80), (1956) 51-92.

11. AVNER FRIEDMAN, Classes of solutions of linear systems of partial differential equa- tions of parabolic type, Duke Math. J. 24 (1957), 433-442.

12. L. GIRDING, Note on continuous representations of Lie groups, Proc. Nat. Acad. Sci. USA 33 (1947), 331-332.

13. , On a lemma by H. Weyl, Kungl. Fysiografiska Stillskapets i Lund Fdrhand- lingar (Proceedings of the Royal Physiographic Society at Lund) 20 (1950), 25)- 253.

14. I. GELFAND, On one-parametrical groups of operators in a normed space, Dokl. Akad. Nauk SSSR (C. R. Acad. Sci. URSS), N. S. 25 (1939), 713-718.

15. HARISH-CHANDRA, Representations of a semisimple Lie group on a Banach space. I, Trans. Amer. Math. Soc. 75 (1953), 185-243.

16. EINAR HILLE AND RALPH S. PHILLIPS, Functional Analysis and Semi-Groups, Amer. Math. Soc. Colloquium Publications 31 2nded, Providence, 1957.

17. G. A. HUNT, Semi-groups of measures on Lie groups, Trans. Amer. Math. Soc. 81 (1956), 264-293.

18. F. JOHN, Plane Waves and Spherical Means Applied to Partial Differential Equations, Interscience, New York, 1955.

19. C. B. MORREY, JR. AND L. NIRENBERG, On the analyticity of the solutions of linear elliptic systems of partial differential equations, Comm. Pure Appl. Math. 10 (1957), 271-290.

20. EDWARD NELSON, Kernel functions and eigenfunction expansions, Duke Math. J. 25 (1958), 15-28.

21. , An existence theorem for second order parabolic equations, Trans. Amer. Math. Soc. 88 (1958), 414-429.

22. , Regular probability measures on function space, Ann. of Math. 69 (1959), 630-643.

23. AND W. FORREST STINESPRING, Representation of elliptic operators in an enveloping algebra, to appear in Amer. J. Math.

24. LOUIS NIRENBERG, Remarks on strongly elliptic partial differential equations, Comm. Pure Appl. Math. 8 (1955) 649-675.

25. D. RAY, Stationary Markov processes with continuous paths, Trans. Amer. Math. Soc. 82 (1956), 452-493.

26. F. RELLICH, Der Eindeutigkeitssatz fur die Lasungen der quanten-mechanischen Vertauschungsrelation, Nachr. Akad. Wiss. Gdttingen Math.-Phys. K1. Math.- Phys.-Chem. Abt. 11A (1946), 107-115.

27. I. E. SEGAL, A class of operator algebras which are determined by groups, Duke Math. J. 18 (1951), 221-265.

28. I. M. SINGER, Lie algebras of unbounded operators, Thesis, The University of Chi- cago, 1950 (unpublished).

29. M. H. STONE, On one parameter unitary groups in Hilbert space, Ann. Math. 33 (1932), 643-648.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions

ANALYTIC VECTORS 615

30. , Linear Transformations in Hilbert Space and Their Applications to Anal- ysis, Amer. Math. Soc. Colloquium Publications 15, New York, 1932.

31. HERMANN WEYL, The Theory of Groups and Quantum Mechanics (translated by H. P. Robertson), Methuen, London, 1931.

32. H. WHITNEY AND F. BRUHAT, Quelques propridtds fondamentales des ensembles ana- lytiques-reels, Comment. Math. Helv. 33 (1959), 132-160.

This content downloaded from 161.45.205.103 on Fri, 6 Sep 2013 23:19:16 PMAll use subject to JSTOR Terms and Conditions


Recommended