Schwartz functions, Hermite functions, and the Hermite operator

Jordan Bell
July 17, 2015

1 Schwartz functions

For ϕC(,) and p0, let

|ϕ|p=sup0kpsupu(1+u2)p/2|ϕ(k)(u)|.

We define 𝒮 to be the set of those ϕC(,) such that |ϕ|p< for all p0. 𝒮 is a complex vector space and each ||p is a norm, and because each ||p is a norm, a fortiori {||p:p0} is a separating family of seminorms. With the topology induced by this family of seminorms, 𝒮 is a Fréchet space.11 1 Walter Rudin, Functional Analysis, second ed., p. 184, Theorem 7.4. Furthermore, D:𝒮𝒮 defined by

(Dϕ)(x)=ϕ(x),x

and M:𝒮𝒮 defined by

(Mϕ)(x)=xϕ(x),x

are continuous linear maps.

Let 𝒮 be the collection of continuous linear maps 𝒮. For ϕ𝒮, define eϕ:𝒮 by

eϕ(ω)=ω(ϕ),ω𝒮.

The initial topology for the collection {eϕ:ϕ𝒮} is called the weak-* topology on 𝒮. With this topology, 𝒮 is a locally convex space whose dual space is {eϕ:ϕ𝒮}.

2 L2 norms

For p0 and ϕ,ψ𝒮, let

[ϕ,ψ]p=k=0p(1+u2)pϕ(k)(u)ψ(k)(u)¯𝑑u,

and let

[ϕ]p2=[ϕ,ϕ]p=k=0p(1+u2)p|ϕ(k)(u)|2𝑑u.

Because (1+u2)p(1+u2)q when pq, it is immediate that [ϕ]p[ϕ]q when pq.

We relate the norms ||p and the norms []p.22 2 Takeyuki Hida, Brownian Motion, p. 305, Lemma A.1.

Lemma 1.

For each p1, for all ϕ𝒮,

1pπ|ϕ|p-1[ϕ]p(p+1)π|ϕ|p+1.
Proof.

For 0kp,

(1+u2)p|ϕ(k)(u)|2𝑑u supu((1+u2)p+1|ϕ(k)(u)|2)(1+u2)-1𝑑u
=supu((1+u2)p+1|ϕ(k)(u)|2)π
π|ϕ|p+12,

hence

[ϕ]p2 =k=0p(1+u2)p|ϕ(k)(u)|2𝑑u
k=0pπ|ϕ|p+12
=(p+1)π|ϕ|p+12.

For 0kp-1 and u, using the fundamental theorem of calculus and the Cauchy-Schwarz inequality,

|(1+u2)(p-1)/2ϕ(k)(u)| =|-u((1+t2)(p-1)/2ϕ(k)(t))𝑑t|
|(p-1)t(1+t2)(p-1)/2-1ϕ(k)(t)|𝑑t
+|(1+t2)(p-1)/2ϕ(k+1)(t)|𝑑t
(p-1)(1+t2)-1/2(1+t2)(p-1)/2|ϕ(k)(t)|𝑑t
+(1+t2)-1/2(1+t2)p/2|ϕ(k+1)(t)|𝑑t
(p-1)((1+t2)-1𝑑t)1/2((1+t2)p-1|ϕ(k)(t)|2𝑑t)1/2
+((1+t2)-1𝑑t)1/2((1+t2)p|ϕ(k+1)(t)|2𝑑t)1/2
(p-1)π[ϕ]p-1+π[ϕ]p
pπ[ϕ]p,

which shows that

|ϕ|p-1pπ[ϕ]p.

3 Hermite functions

Let λ be Lebesgue measure on , and let

(f,g)L2=fg¯𝑑λ.

L2(λ) with the inner product (,)L2 is a separable Hilbert space. For n0, let

hn(x)=(-1)n(2nn!π)-1/2ex2/2Dne-x2,

the Hermite functions, the set of which is an orthonormal basis for L2(λ). We remark that the Hermite functions belong to 𝒮. For n<0 we define

hn=0,

to write some expressions in a uniform way.

We calculate that for n0,

Dhn=n2hn-1-n+12hn+1.

We define the Hermite operator A:𝒮𝒮 by

A=-D2+M2+1.

A is a densely defined operator in L2(λ) that is symmetric and positive, and satisfies

Ahn=(2n+2)hn.

There is a unique bounded linear operator T:L2(λ)L2(λ) satisfying

Thn=A-1hn=(2n+2)-1hn,n0.

The operator norm of T is T=12, and T is self-adjoint. For p1, Tp is a Hilbert-Schmidt operator with Hilbert-Schmidt norm TpHS=2-pζ(2p).

We define the creation operator B:𝒮𝒮 by

B=D+M

and we define the annihilation operator C:𝒮𝒮 by

C=-D+M,

which are continuous linear maps. They satisfy, for n0,

Bhn=(2n)1/2hn-1,Chn=(2n+2)1/2hn+1.

(We remind ourselves that we have defined h-1=0.) It is immediate that BC=A and that B-C=2D. Using the creation operator, we can write the Hermite functions as

hn=(2nn!)-1/2Cnh0=π-1/4(2nn!)-1/2Cn(e-x2/2).

For ϕ,ψ𝒮, using integration by parts,

(Dϕ,ψ)L2=ϕ(x)ψ(x)¯𝑑x=-ϕ(x)ψ(x)¯𝑑x=(ϕ,(-D)ψ)L2,

and

(Mϕ,ψ)L2=xϕ(x)ψ(x)¯𝑑x=(ϕ,Mψ)L2.

Thus,

(Bϕ,ψ)L2 =(Dϕ,ψ)L2+(Mϕ,ψ)L2
=(ϕ,(-D)ψ)L2+(ϕ,Mψ)L2
=(ϕ,Cψ)L2

and

(Cϕ,ψ)L2=(ϕ,Bψ)L2.

We shall use these calculations to obtain the following lemma.

Lemma 2.

For p0 and for ϕ𝒮,

Bpϕ=2p/2n=0((n+p)!n!)1/2(ϕ,hn+p)L2hn

and

Cpϕ=2p/2n=0(ϕ,hn-p)L2(n!(n-p)!)1/2hn.
Proof.

Because Chn=(2n+2)1/2hn+1,

(ϕ,Cphn)L2=(ϕ,hn+p)L2j=nn+p-1(2j+2)1/2=(ϕ,hn+p)L22p/2((n+p)!n!)1/2.

With

ϕ=n=0(ϕ,hn)L2hn,

and because (Bϕ,ψ)L2=(ϕ,Cψ)L2, we have

Bpϕ =n=0(Bpϕ,hn)L2hn
=n=0(ϕ,Cphn)L2hn
=n=0(ϕ,hn+p)L22p/2((n+p)!n!)1/2hn.

Because Bhn=(2n)1/2hn-1, and reminding ourselves that we define hn=0 for n<0,

(ϕ,Bphn)L2=(ϕ,hn-p)L2j=n-p+1n(2j)1/2=(ϕ,hn-p)L22p/2(n!(n-p)!)1/2.

Because (Cϕ,ψ)L2=(ϕ,Bψ)L2, we have

Cpϕ =n=0(Cpϕ,ψ)L2hn
=n=0(ϕ,Bpψ)L2hn
=n=0(ϕ,hn-p)L22p/2(n!(n-p)!)1/2hn.

We define the Fourier transform :𝒮𝒮 by

(ϕ)(ξ)=ϕ(x)e-iξxdx(2π)1/2,ξ.

:𝒮𝒮 is a continuous linear map, and satisfies

M=iD,D=iM.

From these we obtain

A=A,B=iB,C=-iC,

and one proves the following using the above.

Lemma 3.

For n0,

hn=(-i)nhn.

We further remark that for ϕ𝒮,

ϕ2-1/2(ϕL22+ϕL22). (1)

Finally, there is a unique Hilbert space isomorphism :L2(λ)L2(λ) whose restriction to 𝒮 is equal to as already defined. Thus for fL2(λ), as

f=n=0(f,hn)L2hn,

we get

f=n=0(f,hn)L2(-i)nhn.

4 Hermite operator

For p0 and fL2(λ), we define

fp2=n=0(2n+2)2p|(f,hn)L2|2.

We define

𝒮p={fL2(λ):fp<},

and for f,g𝒮p we define

(f,g)p=n=0(2n+2)2p(f,hn)L2(g,hn)L2¯,

for which

fp2=(f,f)p.
Lemma 4.

For ϕ𝒮, for each p0, ϕ𝒮p, and

ϕp=ApϕL2.
Proof.

Apϕ𝒮, so ApϕL2<. Because A is a symmetric operator and as Ahn=(2n+2)hn,

ApϕL22 =n=0|(Apϕ,hn)L2|2
=n=0|(ϕ,Aphn)L2|2
=n=0(2n+2)2p|(ϕ,hn)L2|2
=ϕp2.

For f,gL2(λ), because T is self-adjoint,

(Tpf,Tpg)p =n=0(2n+2)2p(Tpf,hn)L2(Tpf,hn)L2¯
=n=0(2n+2)2p(f,Tphn)L2(g,Tphn)L2¯
=n=0(2n+2)2p(f,(2n+2)-phn)L2(g,(2n+2)-phn)L2¯
=n=0(f,hn)L2(g,hn)L2¯
=(f,g)L2,

and so Tpfp=fL2, which shows that

TpL2(λ)=𝒮p.

If fi𝒮p is a Cauchy sequence in the norm p, then as T-pfi-T-pfjL2=fi-fjp, T-pfi is a Cauchy sequence in the norm L2 and so there is some gL2(λ) for which T-pfi-gL20. We have Tpg𝒮p, and

fi-Tpgp=T-pfi-gL20,

thus fiTpg in the norm p, showing that (𝒮p,(,)p) is a Hilbert space. Furthermore, Tp:L2(λ)𝒮p is an isomorphism of Hilbert spaces, and thus {Tphn:n0} is an orthonormal basis for (𝒮p,(,)p).

For pq,

fpfq,

so 𝒮q𝒮p. For pq, let iq,p:𝒮q𝒮p be the inclusion map.33 3 Hui-Hsiung Kuo, White Noise Distribution Theory, p. 18, Lemma 3.3.

Theorem 5.

For p<q, the inclusion map iq,p:𝒮q𝒮p is a Hilbert-Schmidt operator, with Hilbert-Schmidt norm

iq,pHS=2-q+pζ(2q-2p).
Proof.

{Tqhn:n0} is an orthonormal basis for (𝒮q,(,)q), and

iq,pHS2 =n=0iq,pTqhnp2
=n=0Tqhnp2
=n=0(2n+2)-qhnp2
=n=0(2n+2)-2q(2n+2)2p
=2-2q+2pζ(2q-2p).

5 The Hilbert spaces Sp

For fL2(λ),

f=n=0(f,hn)L2hn,

and for N0 we define fN: by

fN(x)=n=0N(f,hn)L2hn(x),x,

which belongs to 𝒮.

For k0, we define Cbk() to be the set of those functions that are k-times differentiable and such that for each 0jk, f(j) is continuous and bounded. With the norm

fCbk=j=0kf(j)

this is a Banach space. Because the Hermite functions belong to 𝒮, for fL2(λ) and for any k and N, the function fN belongs to Cbk().

Lemma 6.

If p1 and f𝒮p, then there is some FCbp-1() such that f is equal almost everywhere to F.

Proof.

Cramér’s inequality states that there is a constant K0 such that for all n, hnK0. For M<N, using this and the Cauchy-Schwarz inequality,

fN-fMCb0 =n=M+1N(f,hn)L2hn
K0n=M+1N|(f,hn)L2|
=K0n=M+1N(2n+2)-1(2n+2)|(f,hn)L2|
(n=M+1N(2n+2)-2)1/2(n=M+1N(2n+2)2|(f,hn)L2|2)1/2
=(n=M+1N(2n+2)-2)1/2fN-fM1.

Because f𝒮p𝒮1, fN is a Cauchy sequence in 𝒮1, hence fN is a Cauchy sequence in Cb0(), so there is some FCb0() such that fN converges to F in Cb0(). We assert that F=f as elements of L2(λ).

Using

Dhn=n2hn-1-n+12hn+1,

we calculate

fN =-N2(f,hN-1)L2hN-N+12(f,hN)L2hN+1
+n=0N-1(n+12(f,hn+1)L2-n2(f,hn-1)L2)hn,

hence for M<N,

fN-fM =-N2(f,hN-1)L2hN-N+12(f,hN)L2hN+1
+M2(f,hM-1)L2hM+M+12(f,hM)L2hM+1
+n=MN-1(n+12(f,hn+1)L2-n2(f,hn-1)L2)hn,

and for NM+2,

fN-fM1 =(2N+2)2N2|(f,hN-1)|L22+(2N+4)2N+12|(f,hN-1)|L22
(2M+2)2M+12|(f,hM+1)|L22+(2M+4)2M+22|(f,hM+1)|L22
+n=M+2N-1(2n+2)2|n+12(f,hn+1)L2-n2(f,hn-1)L2|2
=O(fN-fM2),

whence fN is a Cauchy sequence in Cb0(), and so fN is a Cauchy sequence in Cb1(). ∎

We prove that for p1 the derivatives of the partial sums fN are a Cauchy sequence in L2(λ).44 4 Jeremy J. Becnel and Ambar N. Sengupta, The Schwartz space: a background to white noise analysis, https://www.math.lsu.edu/~preprint/2004/as20041.pdf, Lemma 7.1.

Lemma 7.

For p1 and f𝒮p, fN is a Cauchy sequence in L2(λ).

Proof.

Because fN𝒮,

fN=DfN=B-C2fN.

Then

fN-fML212BfN-BfML2+12CfN-CfML2.

For M<N, as Bhn=(2n)1/2hn-1,

BfN-BfML22 =Bn=M+1N(f,hn)L2hnL22
=n=M+1N(f,hn)L2(2n)1/2hn-1L22
=n=M+1N|(f,hn)L2|2(2n)
n=M+1N(2n+2)2|(f,hn)L2|2,

and as Chn=(2n+2)1/2hn+1,

CfN-CfML22 =Cn=M+1N(f,hn)L2hnL22
=n=M+1N(f,hn)L2(2n+2)1/2hn+1L22
=n=M+1N|(f,hn)L2|2(2n+2)
n=M+1N(2n+2)2|(f,hn)L2|2.

Thus

fN-fML212fN-fM1+12fN-fM1=fN-fM1.

Because f𝒮p and p1, the series n=0(2n+2)2|(f,hn)L2|2 converges, from which the claim follows. ∎

Now we establish that if p1 and f𝒮p then there is some FCb0() such that f is equal almost everywhere to F, F is differentiable almost everywhere, and F𝒮p-1.55 5 Jeremy J. Becnel and Ambar N. Sengupta, The Schwartz space: a background to white noise analysis, https://www.math.lsu.edu/~preprint/2004/as20041.pdf, Theorem 7.3.

Theorem 8.

For p1 and f𝒮p, there is some FCb0() such that f is equal almost everywhere to F, F is differentiable almost everywhere, fN converges to F in the norm L2, and F𝒮p-1.

Proof.

Lemma 7 tells us that fN is a Cauchy sequence in the norm L2, and hence there is some gL2(λ) to which fN converges in the norm L2. For xy, by the fundamental theorem of calculus,

fN(y)=fN(x)+01fN(x+t(y-x))(y-x)𝑑t.

By the Cauchy-Schwarz inequality,

01|fN(x+t(y-x))(y-x)-g(x+t(y-x))(y-x)|𝑑t=xy|fN(u)-g(u)|𝑑uy-xfN-gL2.

Because fN-gL20 as N,

01fN(x+t(y-x))(y-x)𝑑t01g(x+t(y-x))(y-x)𝑑t.

Then by Lemma 6, taking N, for any y>x we have

F(y)=F(x)+01g(x+t(y-x))(y-x)𝑑t=F(x)+1y-xxyg(s)𝑑s.

By the Lebesgue differentiation theorem, for almost all x,

1y-xxyg(s)𝑑sg(x),yx.

Therefore for almost all x,

F(x)=g(x).

Thus F=g in L2(λ), and as fNg in L2(λ),

F =limNfN
=limN(B-C2)n=0N(f,hn)L2hn
=12n=0(f,hn)L2((2n)1/2hn-1-(2n+2)1/2hn+1)
=12n=0((2n+2)1/2(f,hn)L2-(2n)1/2(f,hn-1)L2)hn,

for which

Fp-12 =14n=0(2n+2)2p-2|(2n+2)1/2(f,hn)L2-(2n)1/2(f,hn-1)L2|2
12n=0(2n+2)2p-2((2n+2)|(f,hn)L2|2+2n|(f,hn-1)L2|2),

which is finite because f𝒮p. Therefore F𝒮p-1. ∎