This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Rapidly decaying Wigner functions are Schwartz functions

Felipe Hernández1 1Department of Mathematics, Stanford University, 450 Jane Stanford Way, Building 380, Stanford, CA 94305 USA  and  C. Jess Riedel2 2Physics & Informatics Laboratories, NTT Research Inc., 940 Stewart Drive, Sunnyvale, CA 94085, USA [email protected]
Abstract.

We show that if the Wigner function of a (possibly mixed) quantum state decays toward infinity faster than any polynomial in the phase space variables xx and pp, then so do all of its derivatives, i.e., it is a Schwartz function on phase space. This is equivalent to the condition that the Husimi function is a Schwartz function, that the quantum state is a Schwartz operator in the sense of Keyl et al., and, in the case of a pure state, that the wavefunction is a Schwartz function on configuration space. We discuss the interpretation of this constraint on Wigner functions and provide explicit bounds on Schwartz seminorms.

1. Introduction

In quantum mechanics, quantum states of nn degrees of freedom can be represented by positive semidefinite trace-class operators on L2(𝐑n)L^{2}({\mathbf{R}}^{n}). Each quantum state ρ\rho is associated with a kernel 𝒦ρ{\mathcal{K}_{\rho}} through (ρϕ)(x)=𝒦ρ(x,y)ϕ(x)dx(\rho\phi)(x)=\int{\mathcal{K}_{\rho}}(x,y)\phi(x)\mathop{}\!\mathrm{d}x, ϕL2(𝐑n)\phi\in L^{2}({\mathbf{R}}^{n}), and the corresponding Wigner function 𝒲ρ{\mathcal{W}_{\rho}} is

𝒲ρ(x,p):=1(2π)neipy𝒦ρ(xy/2,x+y/2)dy.{\mathcal{W}_{\rho}}(x,p):=\frac{1}{(2\pi)^{n}}\int e^{ip\cdot y}{\mathcal{K}_{\rho}}(x-y/2,x+y/2)\mathop{}\!\mathrm{d}y.

We denote the set of all such Wigner function as 𝒱(𝐑2n){\mathcal{V}}({\mathbf{R}}^{2n}). Our main result is a relationship between the decay of such Wigner functions and their smoothness.

To quantify this we use the Schwartz-type seminorms |F|a,b|F|_{a,b} :=:= supx,p|xaxpapxbxpbxF(x,p)|\sup_{x,p}|x^{a_{\mathrm{x}}}p^{a_{\mathrm{p}}}\partial_{x}^{b_{\mathrm{x}}}\partial_{p}^{b_{\mathrm{x}}}F(x,p)| of a function F:𝐑2n𝐂F:{{\mathbf{R}}^{2n}}\to{\mathbf{C}} on phase space, with multi-indices a=(ax,ap),b=(bx,bp)({0})×2n.a=(a_{\mathrm{x}},a_{\mathrm{p}}),b=(b_{\mathrm{x}},b_{\mathrm{p}})\in(\mathbb{N}\cup\{0\})^{\times 2n}. With shorthand notation |F|a:=|F|a,0|F|_{a}:=|F|_{a,0} for the seminorms that only measure the decay of FF, we say a function is rapidly decaying when |F|a<|F|_{a}<\infty and is a Schwartz function when |F|a,b<|F|_{a,b}<\infty for all a,ba,b. We denote the sets of rapidly decaying and Schwartz function by 𝒟(𝐑2n){\mathcal{D}}({\mathbf{R}}^{2n}) and 𝒮(𝐑2n){\mathcal{S}}({\mathbf{R}}^{2n}), respectively. Our main result:

Theorem 1.1.

If ρ\rho is a positive semidefinite operator whose Wigner function 𝒲ρ{\mathcal{W}_{\rho}} exists and is rapidly decaying, then 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function.

The assumed rapid decay of 𝒲ρ{\mathcal{W}_{\rho}} implies >𝒲ρ(α)dα=tr[ρ]\infty>\int{\mathcal{W}_{\rho}}(\alpha)\mathop{}\!\mathrm{d}\alpha=\operatorname{tr}[\rho] and hence that ρ\rho is trace-class and so a quantum state. Thus the theorem can be rephrased as the set relation 𝒱(𝐑2n)𝒟(𝐑2n)𝒮(𝐑2n){\mathcal{V}}({\mathbf{R}}^{2n})\cap{\mathcal{D}}({\mathbf{R}}^{2n})\subset{\mathcal{S}}({\mathbf{R}}^{2n}).

In this paper we prove Theorem 1.1 in two different ways. The first proof is a bit more abstract, making use of the twisted convolution. The second proof is a bit more direct, using only basic objects, but requiring more computation. The second proof also results in an explicit bound on the Schwartz seminorms |𝒲ρ|a,b|{\mathcal{W}_{\rho}}|_{a,b} of a Wigner function in terms of only its decay seminorms |𝒲ρ|a|{\mathcal{W}_{\rho}}|_{a} (Theorem 3.9).

In the rest of this introduction, we informally sketch the direct (second) proof of Theorem 1.1 in order to give the reader intuition, but we stop short of completing the computation. In the Sec. 2, we recall some notation and basic properties around quantum mechanics in phase space, which can be skipped by experienced readers. In Sec. 3 we present our two proofs of our main result and exhibit explicit bounds on the Schwartz seminorms of a Wigner function in terms of its decay seminorms. In Section 4 we connect our results to the notion of Schwartz operators in the sense of Keyl et al. [keyl2016schwartz], and in particular prove the equivalence of a large set of equivalent decay and regularity conditions for various representations of the quantum state. In Sec. 5, we make some concluding remarks about the “overparameterization” of a quantum state by the Wigner function.

1.1. Motivation

Why might one think the decay of a Wigner function constrains its derivatives? Consider a pure state ρ=|ψψ|\rho=\mathinner{|{\psi}\rangle}\!\!\mathinner{\langle{\psi}|} with |ψL2(𝐑n)\mathinner{|{\psi}\rangle}\in L^{2}({\mathbf{R}}^{n}). We can see from the identity

|ψ^(p)|2=𝒲ρ(x,p)dx|\widehat{\psi}(p)|^{2}=\int{\mathcal{W}_{\rho}}(x,p)\mathop{}\!\mathrm{d}x (1.1)

that rapid decay (in both xx and pp) of the Wigner function implies decay (in pp) of the Fourier transform ψ^\widehat{\psi} of the wavefunction. This implies that the wavefunction ψ\psi is smooth: |ψ|0,b<|\psi|_{0,b}<\infty for all b({0})×nb\in(\mathbb{N}\cup\{0\})^{\times n}. Unfortunately, a bit of trial and error suggests that it is not easy to generalize (1.1) and obtain a bound on the mixed Schwartz seminorms |ψ|a,b|\psi|_{a,b} (that is, to show that all the derivatives of ψ\psi are not merely bounded but are also rapidly decaying).

A better way to approach Theorem 1.1 avoids privileging either the position or momentum variables by performing a wavepacket decomposition of the quantum state ρ\rho. Using Gaussian wavepackets (coherent states), Zurek argued [zurek2001sub-planck] that if the Wigner function 𝒲ρ{\mathcal{W}_{\rho}} of any quantum state is largely confined to a phase space region of volume Sx×pS\sim\ell_{\mathrm{x}}\times\ell_{\mathrm{p}}, then the smallest structure it will develop is on scales of volume Δs(/x)×(/p)2/S\Delta s\sim(\hbar/\ell_{\mathrm{x}})\times(\hbar/\ell_{\mathrm{p}})\sim\hbar^{2}/S. This argument was further supported by numerical studies of “typical” states generated by chaotic quantum dynamics [zurek2001sub-planck].

1.2. Sketch of direct proof

Consider a family of wavepackets χα\chi_{\alpha} of the form

χ(αx,αp)(x)=ei(xαx/2)αpχ(xαx),\chi_{(\alpha_{\mathrm{x}},\alpha_{\mathrm{p}})}(x)=e^{i(x-\alpha_{x}/2)\cdot\alpha_{p}}\chi(x-\alpha_{x}), (1.2)

for fixed smooth envelope function χ\chi concentrated near the origin. (For example, χ\chi can be chosen to be a Gaussian.) Given the spectral decomposition of a quantum state

ρ=jλj|ψjψj|,\rho=\sum_{j}\lambda_{j}\mathinner{|{\psi_{j}}\rangle}\!\!\mathinner{\langle{\psi_{j}}|}, (1.3)

we can use the decomposition ψj=(2π)nχα|ψjχαdα\psi_{j}=(2\pi)^{-n}\int\mathinner{\langle{\chi_{\alpha}|\psi_{j}}\rangle}\chi_{\alpha}\mathop{}\!\mathrm{d}\alpha for each eigenfunction as an integral over phase space, which is a standard calculation proven in Lemma 2.9. We can then express ρ\rho as

ρ=1(2π)njλj|χαχβ|χα|ψjψj|χβdαdβ,\rho=\frac{1}{(2\pi)^{n}}\sum_{j}\lambda_{j}\int\mathinner{|{\chi_{\alpha}}\rangle}\!\!\mathinner{\langle{\chi_{\beta}}|}\mathinner{\langle{\chi_{\alpha}|\psi_{j}}\rangle}\mathinner{\langle{\psi_{j}|\chi_{\beta}}\rangle}\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}\beta, (1.4)

Applying the Wigner transform to both sides, this yields a decomposition

𝒲ρ=1(2π)n𝒲|χαχβ|χα|ρ|χβdαdβ.{\mathcal{W}_{\rho}}=\frac{1}{(2\pi)^{n}}\int{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\beta}}\rangle}\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}\beta. (1.5)

in terms of the Wigner transform 𝒲|χαχβ|{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}} of the “off-diagonal” operator |χαχβ|\mathinner{|{\chi_{\alpha}}\rangle}\!\!\mathinner{\langle{\chi_{\beta}}|}. Although 𝒲|χαχβ|{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}} is not a Wigner function (because |χαχβ|\mathinner{|{\chi_{\alpha}}\rangle}\!\!\mathinner{\langle{\chi_{\beta}}|} is not positive semidefinite for αβ\alpha\neq\beta), it is known [zurek2001sub-planck, toscano2006sub-planck] to be localized near the phase space point (α+β)/2(\alpha+\beta)/2 and has an oscillation with frequency roughly |αβ||\alpha-\beta|.

Since χα|ρ|χα\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\alpha}}\rangle} is just a convolution of the Wigner function 𝒲ρ{\mathcal{W}_{\rho}}, the rapid decay of 𝒲ρ{\mathcal{W}_{\rho}} implies the rapid decay of χα|ρ|χα\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\alpha}}\rangle}, and then in turn one can show the rapid decay of χα|ρ|χβ\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\beta}}\rangle} using the Cauchy-Schwartz inequality:

χα|ρ|χβ2χα|ρ|χαχβ|ρ|χβ,\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\beta}}\rangle}^{2}\leq\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\alpha}}\rangle}\mathinner{\langle{\chi_{\beta}|\rho|\chi_{\beta}}\rangle}, (1.6)

which holds because ρ\rho is positive semidefinite. The assumed decay and smoothness properties of χ\chi additionally give an estimate of the form

|𝒲|χαχβ||a,bC(a,b)(1+|α|+|β|)D(a,b).|{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}|_{a,b}\leq C(a,b)(1+|\alpha|+|\beta|)^{D(a,b)}. (1.7)

When combined with decomposition (1.5) of 𝒲ρ{\mathcal{W}_{\rho}}, this is enough to show that all the Schwartz seminorms |𝒲ρ|a,b|{\mathcal{W}_{\rho}}|_{a,b} are finite.

Our other proof requires additional machinery but still rests heavily on wavepacket decompositions of the quantum state and on the Cauchy-Schwartz inequality (1.6).

2. Preliminaries

This section establishes our notation and reviews standard features of phase-space representations of quantum mechanics. (To keep this paper self-contained, we provide proofs of the lemmas in this section in the Appendix.) Throughout this paper, we take χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}) to be a fixed Schwartz function that is normalized, χL2(𝐑n)=|χ(y)|2dy=1\|\chi\|_{L^{2}({\mathbf{R}}^{n})}=\int|\chi(y)|^{2}\mathop{}\!\mathrm{d}y=1, but otherwise arbitrary.111A standard choice is to specialize to a Gaussian coherent state χ(y)=exp(x2/2)/(2π)n\chi(y)=\exp(-x^{2}/2)/\sqrt{(2\pi)^{n}} (especially when used as in Subsection 2.2 as the reference wavefunction with respect to which Husimi function is defined). However, this specialization is not necessary and one could instead take χ\chi to be, e.g., a smooth and compactly supported wavefunction.

Experienced readers may prefer to skip directly to Sec. 3 for the proof of our main result and only refer back to this section as necessary.

2.1. Notation

In what follows, a wavefunction of nn continuous quantum degrees of freedom is represented by a member of L2(𝐑n)L^{2}({\mathbf{R}}^{n}) and denoted by ψ\psi, ϕ\phi, or χ\chi. A quantum state is the possibly mixed generalization, represented by a positive semidefinite (and hence self-adjoint) trace-class operator on L2(𝐑n)L^{2}({\mathbf{R}}^{n}) and denoted by ρ\rho or η\eta. Vectors on phase space are α,β,γ,ξ𝐑2n\alpha,\beta,\gamma,\xi\in{{\mathbf{R}}^{2n}} with position and momentum components denoted by (for example) αx,ξp𝐑n\alpha_{\mathrm{x}},\xi_{\mathrm{p}}\in{\mathbf{R}}^{n}. Multi-indices are a,b,c,d({0})×2na,b,c,d\in(\mathbb{N}\cup\{0\})^{\times 2n} (or ({0})×n\mathbb{N}\cup\{0\})^{\times n} in Sec. 4) with αb=αxbxαpbp=i=12nαibi\alpha^{b}=\alpha_{\mathrm{x}}^{b_{\mathrm{x}}}\alpha_{\mathrm{p}}^{b_{\mathrm{p}}}=\prod_{i=1}^{2n}\alpha_{i}^{b_{i}}, |b|=|bx|+|bp|=i=12nbi|b|=|b_{\mathrm{x}}|+|b_{\mathrm{p}}|=\sum_{i=1}^{2n}b_{i}, b!=bx!bp!=i=12nbi!b!=b_{\mathrm{x}}!b_{\mathrm{p}}!=\prod_{i=1}^{2n}b_{i}!, and (ab)=a!/((ab)!b!)\binom{a}{b}=a!/((a-b)!b!). We use bcb\leq c to mean bicib_{i}\leq c_{i} for all i=1,2,2ni=1,2,\ldots 2n.

The symplectic form is αβ=αΩβ=αxβpαpβx\alpha\wedge\beta=\alpha\cdot\Omega\cdot\beta=\alpha_{\mathrm{x}}\cdot\beta_{\mathrm{p}}-\alpha_{\mathrm{p}}\cdot\beta_{\mathrm{x}}, with Ω=(0II0)\Omega=\left(\begin{smallmatrix}0&I\\ -I&0\end{smallmatrix}\right) an antisymmetric matrix on 𝐑2n{\mathbf{R}}^{2n}, II the identity matrix on 𝐑n{\mathbf{R}}^{n}, and “\cdot” the dot product on 𝐑n{\mathbf{R}}^{n} and 𝐑2n{\mathbf{R}}^{2n}. The position and momentum operators are X=(X1,,Xn)X=(X_{1},\ldots,X_{n}) and P=(P1,,Pn)P=(P_{1},\ldots,P_{n}), which are combined into the phase-space operator R=(X,P)R=(X,P). For a given quantum state ρ\rho and reference wavefunction χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}), some associated functions over phase space, doubled phase space, and doubled configuration space are 𝒲ρ{\mathcal{W}_{\rho}}, 𝒬ρχ{\mathcal{Q}^{\chi}_{\rho}}, ρχ{\mathcal{M}^{\chi}_{\rho}}, ρ{\mathcal{F}_{\rho}}, and 𝒦ρ{\mathcal{K}_{\rho}} (defined below). We use “*” to denote the convolution, (fg)(α)=f(αβ)g(β)dβ(f*g)(\alpha)=\int f(\alpha-\beta)g(\beta)\mathop{}\!\mathrm{d}\beta. Given a matrix form Ω\Omega^{\prime} we also define the twisted convolution

(fΩg)(α)=eiαΩβ/2f(αβ)g(β)dβ.(f\circledast_{\Omega^{\prime}}g)(\alpha)=\int e^{i\alpha\cdot\Omega^{\prime}\cdot\beta/2}f(\alpha-\beta)g(\beta)\mathop{}\!\mathrm{d}\beta. (2.1)

For any wavefunction ϕL2(𝐑n)\phi\in L^{2}({\mathbf{R}}^{n}), we denote the linear functional associated with it using bra notation, ϕ|=(ψϕ¯(x)ψ(x)dx)𝒮(𝐑n)\mathinner{\langle{\phi}|}=(\psi\mapsto\int\bar{\phi}(x)\psi(x)\mathop{}\!\mathrm{d}x)\in{\mathcal{S}}^{\prime}({\mathbf{R}}^{n}), and denote the scalar result with a bra-ket, ϕ|ψ=ϕ|(ψ)=ϕ¯(x)ψ(x)dx\mathinner{\langle{\phi|\psi}\rangle}=\mathinner{\langle{\phi}|}(\psi)=\int\bar{\phi}(x)\psi(x)\mathop{}\!\mathrm{d}x. More generally, with an operator EE we write ϕ|E|ψ=ϕ|(Eψ)=Eϕ|(ψ)\mathinner{\langle{\phi|E|\psi}\rangle}=\mathinner{\langle{\phi}|}(E\psi)=\mathinner{\langle{E^{\dagger}\phi}|}(\psi). For any two wavefunction ϕ1,ϕ2L2(𝐑n)\phi_{1},\phi_{2}\in L^{2}({\mathbf{R}}^{n}), we use |ϕ1ϕ2|\mathinner{|{\phi_{1}}\rangle}\!\!\mathinner{\langle{\phi_{2}}|} for the rank-1 operator ψϕ2|ψϕ1\psi\mapsto\mathinner{\langle{\phi_{2}|\psi}\rangle}\phi_{1}.

2.2. The displacement operator and phase-space functions

In this subsection we recall standard results about quantum mechanics in phase space (see, e.g., Chapter 1 of Ref. [folland1989harmonic]).

Definition 2.1.

For ξ𝐑2n\xi\in{{\mathbf{R}}^{2n}}, define the (Weyl generator) displacement operator

Dξ:=eiξR=ei(ξxPξpX).D_{\xi}:=e^{i\xi\wedge R}=e^{i(\xi_{\mathrm{x}}\cdot P-\xi_{\mathrm{p}}\cdot X)}. (2.2)

The following lemma describes the action of DξD_{\xi} on an arbitrary wavefunction.

Lemma 2.2.

For any ϕL2(𝐑n)\phi\in L^{2}({\mathbf{R}}^{n}),

Dξϕ(y)=ei(yξx/2)ξpϕ(yξx).D_{\xi}\phi(y)=e^{i(y-\xi_{\mathrm{x}}/2)\cdot\xi_{\mathrm{p}}}\phi(y-\xi_{\mathrm{x}}). (2.3)

It’s easy to check these basic properties: DαDβ=eiβα/2Dα+βD_{\alpha}D_{\beta}=e^{i\beta\wedge\alpha/2}D_{\alpha+\beta} and Dα=DαD_{\alpha}^{\dagger}=D_{-\alpha}.

Now we introduce the quasicharacteristic function, the Wigner function, and the Kernel.

Definition 2.3.

For a given quantum state ρ\rho, the quasicharacteristic function is

ρ(ξ):=tr[ρDξ].{\mathcal{F}_{\rho}}(\xi):=\operatorname{tr}[\rho D_{\xi}]. (2.4)

where the trace is well defined because ρ\rho is trace-class and DξD_{\xi} is a bounded operator on L2(𝐑n)L^{2}({\mathbf{R}}^{n}). Because a quantum state ρ\rho is necessarily compact, it has a spectral decomposition [gohberg2000trace]

(ρϕ)(x)=i=1ψi(x)ψi|ϕ\begin{split}(\rho\phi)(x)&=\sum_{i=1}^{\infty}\psi_{i}(x)\mathinner{\langle{\psi_{i}|\phi}\rangle}\end{split} (2.5)

with unnormalized eigenvectors ψiL2(𝐑n)\psi_{i}\in L^{2}({\mathbf{R}}^{n}) and associated kernel 𝒦ρ{\mathcal{K}_{\rho}} satisfying (ρϕ)(x)=𝒦ρ(x,y)ϕ(x)dx(\rho\phi)(x)=\int{\mathcal{K}_{\rho}}(x,y)\phi(x)\mathop{}\!\mathrm{d}x and

𝒦ρ(x,y)=i=1ψi(x)ψ¯i(y){\mathcal{K}_{\rho}}(x,y)=\sum_{i=1}^{\infty}\psi_{i}(x)\bar{\psi}_{i}(y) (2.6)

almost everywhere. Finally, we define the Wigner function of ρ\rho as

𝒲ρ(α):=1(2π)neiαpy𝒦ρ(αxy/2,αx+y/2)dy,{\mathcal{W}_{\rho}}(\alpha):=\frac{1}{(2\pi)^{n}}\int e^{i\alpha_{\mathrm{p}}\cdot y}{\mathcal{K}_{\rho}}(\alpha_{\mathrm{x}}-y/2,\alpha_{\mathrm{x}}+y/2)\mathop{}\!\mathrm{d}y, (2.7)

where 𝒲ρL2(𝐑2n){\mathcal{W}_{\rho}}\in L^{2}({\mathbf{R}}^{2n}) because it is a Fourier transform of 𝒦ρL2(𝐑2n){\mathcal{K}_{\rho}}\in L^{2}({\mathbf{R}}^{2n}) in one variable.

More generally, we call 𝒲E(α):=(2π)neiαpy𝒦E(αxy/2,αx+y/2)dy{\mathcal{W}_{E}}(\alpha):=(2\pi)^{-n}\int e^{i\alpha_{\mathrm{p}}\cdot y}{\mathcal{K}_{E}}(\alpha_{\mathrm{x}}-y/2,\alpha_{\mathrm{x}}+y/2)\mathop{}\!\mathrm{d}y and E(ξ):=tr[EDξ]{\mathcal{F}_{E}}(\xi):=\operatorname{tr}[ED_{\xi}] the Wigner transform and quasicharacteristic transform of any kernel operator EE, which in particular exists for any rank-1 operator E=|ϕψ|E=\mathinner{|{\phi}\rangle}\!\!\mathinner{\langle{\psi}|} since 𝒦|ϕψ|L2(𝐑2n){\mathcal{K}_{\mathinner{|{\phi}\rangle}\!\mathinner{\langle{\psi}|}}}\in L^{2}({\mathbf{R}}^{2n}).

Lemma 2.4.

For any trace-class kernel operator EE, the corresponding Wigner transform and quasicharacteristic transform are symplectic Fourier duals:

𝒲E(α)=1(2π)2neiαξE(ξ)dξ.{\mathcal{W}_{E}}(\alpha)=\frac{1}{(2\pi)^{2n}}\int e^{-i\alpha\wedge\xi}{\mathcal{F}_{E}}(\xi)\mathop{}\!\mathrm{d}\xi. (2.8)

The preceding expression is sometimes used as the definition of the Wigner transform, and it is notable for manifestly respecting the symplectic structure of phase space. The perhaps more traditional definition (2.7) relies on the kernel, and hence privileges position over momentum, but has the advantage of being more obviously well-defined.

Lemma 2.5.

The Wigner function 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function if and only if the kernel 𝒦ρ{\mathcal{K}_{\rho}} is a Schwartz function.

Roughly speaking, this is because 𝒲ρ{\mathcal{W}_{\rho}} and 𝒦ρ{\mathcal{K}_{\rho}} are Fourier transforms of each other in one of their two variables (after the linear change of variables (x,y)(x¯=(x+y)/2,Δx=xy)(x,y)\to(\bar{x}=(x+y)/2,\Delta x=x-y)).

Lemma 2.6.

The twisted convolution of a rapidly decaying function with a Schwartz function is itself a Schwartz function.

The proof is essentially the same as for the similar statement with the normal convolution.

Lemma 2.7.

For any two quantum states ρ\rho and η\eta,

tr[ρη]=(2π)n𝒲ρ(α)𝒲η(α)dα.\displaystyle\operatorname{tr}[\rho\eta]=(2\pi)^{n}\int{\mathcal{W}_{\rho}}(\alpha){\mathcal{W}_{\eta}}(\alpha)\mathop{}\!\mathrm{d}\alpha. (2.9)

Now we introduce the Husimi function and the so-called matrix element; these are most often defined with respect to a preferred Gaussian reference wavefunction, but we will allow more generality (see, e.g., Ref. [klauder2007generalized]).

Definition 2.8.

Fixing a reference wavefunction χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}) that is normalized (χL2(𝐑n)=|χ(y)|2dy=1\|\chi\|_{L^{2}({\mathbf{R}}^{n})}=\int|\chi(y)|^{2}\mathop{}\!\mathrm{d}y=1), and Schwartz-class but otherwise arbitrary, we define the matrix element

ρχ(α,β):=χα|ρ|χβ,{\mathcal{M}^{\chi}_{\rho}}(\alpha,\beta):=\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\beta}}\rangle}, (2.10)

and the Husimi function

𝒬ρχ(α):=χα|ρ|χα=ρχ(α,α).{\mathcal{Q}^{\chi}_{\rho}}(\alpha):=\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\alpha}}\rangle}={\mathcal{M}^{\chi}_{\rho}}(\alpha,\alpha). (2.11)

using the shorthand |χα:=Dα|χ\mathinner{|{\chi_{\alpha}}\rangle}:=D_{\alpha}\mathinner{|{\chi}\rangle}.

Lemma 2.9.

For any trace-class operator EE and any χL2(𝐑n)\chi\in L^{2}({\mathbf{R}}^{n}) satisfying χL2(𝐑n)=1\|\chi\|_{L^{2}({\mathbf{R}}^{n})}=1,

tr[E]=1(2π)nχα|E|χαdα\displaystyle\operatorname{tr}[E]=\frac{1}{(2\pi)^{n}}\int\mathinner{\langle{\chi_{\alpha}|E|\chi_{\alpha}}\rangle}\mathop{}\!\mathrm{d}\alpha (2.12)

In particular, for any ϕ,ψL2(𝐑n)\phi,\psi\in L^{2}({\mathbf{R}}^{n})

ϕ|ψ=1(2π)nϕ|χαχα|ψdα\displaystyle\mathinner{\langle{\phi|\psi}\rangle}=\frac{1}{(2\pi)^{n}}\int\mathinner{\langle{\phi|\chi_{\alpha}}\rangle}\mathinner{\langle{\chi_{\alpha}|\psi}\rangle}\mathop{}\!\mathrm{d}\alpha (2.13)
Lemma 2.10.

For any quantum state ρ\rho and reference wavefunction χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}),

𝒬ρχ(α)\displaystyle{\mathcal{Q}^{\chi}_{\rho}}(\alpha) =(2π)n(𝒲ρ𝒲χ)(α)=(2π)n𝒲ρ(β)𝒲χ(βα)dβ\displaystyle=(2\pi)^{n}({\mathcal{W}_{\rho}}\ast{\mathcal{W}_{\chi}^{-}})(\alpha)=(2\pi)^{n}\int{\mathcal{W}_{\rho}}(\beta){\mathcal{W}_{\chi}}(\beta-\alpha)\mathop{}\!\mathrm{d}\beta (2.14)

where 𝒲χ(α):=𝒲χ(α){{\mathcal{W}_{\chi}^{-}}}(\alpha):={\mathcal{W}_{\chi}}(-\alpha) is a Schwartz function.

3. Proof that rapidly decaying Wigner functions are Schwartz function

The first (more abstract) proof of our main result is given in subsection 3.1 below. The second (more direct) proof follows in subsection 3.2. These two subsections are independent of each other and can be read in either order.

Both proofs will make crucial use of the Cauchy-Schwartz inequality in the following form:

Lemma 3.1.

For any quantum state ρ\rho, the Husimi function bounds the matrix element:

|ρχ(α,β)|2𝒬ρχ(α)𝒬ρχ(β)|{\mathcal{M}^{\chi}_{\rho}}(\alpha,\beta)|^{2}\leq{\mathcal{Q}^{\chi}_{\rho}}(\alpha){\mathcal{Q}^{\chi}_{\rho}}(\beta) (3.1)
Proof.

We have

|ρχ(α,β)|2=|χα|ρ|χβ|2χα|ρ|χαχβ|ρ|χβ=𝒬ρχ(α)𝒬ρχ(β)|{\mathcal{M}^{\chi}_{\rho}}(\alpha,\beta)|^{2}=|\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\beta}}\rangle}|^{2}\leq\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\alpha}}\rangle}\mathinner{\langle{\chi_{\beta}|\rho|\chi_{\beta}}\rangle}={\mathcal{Q}^{\chi}_{\rho}}(\alpha){\mathcal{Q}^{\chi}_{\rho}}(\beta) (3.2)

where the inequality is just the Cauchy-Schwartz inequality applied to the inner product ϕ1,ϕ2ρ:=ϕ1|ρ|ϕ2\langle\phi_{1},\phi_{2}\rangle_{\rho}:=\mathinner{\langle{\phi_{1}|\rho|\phi_{2}}\rangle}. ∎

Corollary 3.2.

If the Wigner function 𝒲ρ{\mathcal{W}_{\rho}} of a quantum state ρ\rho is rapidly decaying, then the Husimi function 𝒬ρχ{\mathcal{Q}^{\chi}_{\rho}} and the matrix element ρχ{\mathcal{M}^{\chi}_{\rho}} are also rapidly decaying.

Proof.

By Lemma 2.10, the Husimi function 𝒬ρχ{\mathcal{Q}^{\chi}_{\rho}} is a convolution of the rapidly decaying 𝒲ρ{\mathcal{W}_{\rho}} by the Schwartz function 𝒲χ(α)=𝒲χ(α){\mathcal{W}_{\chi}^{-}}(\alpha)={\mathcal{W}_{\chi}}(-\alpha), so 𝒬ρχ{\mathcal{Q}^{\chi}_{\rho}} must also be rapidly decaying. We then get rapid decay of ρχ{\mathcal{M}^{\chi}_{\rho}} using Lemma 3.1. ∎

We now turn to the first strategy.

3.1. Abstract proof

Here is a sketch of our strategy: We obtain a reproducing formula expressing ρχ{\mathcal{M}^{\chi}_{\rho}} as a twisted convolution of itself with a Schwartz function constructed from χ\chi, showing that ρχ{\mathcal{M}^{\chi}_{\rho}} must itself be a Schwartz function. Then we find an integral expression for the Wigner function 𝒲ρ{\mathcal{W}_{\rho}} in terms of the matrix element ρχ{\mathcal{M}^{\chi}_{\rho}}, from which it follows that 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function.

Lemma 3.3.

Let Ω=(Ω00Ω)\Omega^{\prime}=\left(\begin{smallmatrix}\Omega&0\\ 0&-\Omega\end{smallmatrix}\right) be a symplectic form on 𝐑4n{\mathbf{R}}^{4n}. Then the matrix element ρχ{\mathcal{M}^{\chi}_{\rho}} satisfies the following reproducing formula

ρχ=1(2π)2n(χ¯χ)Ωρχ{\mathcal{M}^{\chi}_{\rho}}=\frac{1}{(2\pi)^{2n}}({\mathcal{F}_{\chi}}\bar{\otimes}{\mathcal{F}_{\chi}})\circledast_{\Omega^{\prime}}{\mathcal{M}^{\chi}_{\rho}} (3.3)

where χ=Tr[|χχ|Dξ]=χ|Dξ|χ=χξ/2|χξ/2{\mathcal{F}_{\chi}}=\operatorname{Tr}[\mathinner{|{\chi}\rangle}\!\!\mathinner{\langle{\chi}|}D_{\xi}]=\mathinner{\langle{\chi|D_{\xi}|\chi}\rangle}=\mathinner{\langle{\chi_{-\xi/2}|\chi_{\xi/2}}\rangle} is the quasicharacteristic function of the pure quantum state |χχ|\mathinner{|{\chi}\rangle}\!\!\mathinner{\langle{\chi}|} and where χ¯χ:𝐑2n×𝐑2n𝐂{\mathcal{F}_{\chi}}\bar{\otimes}{\mathcal{F}_{\chi}}:{\mathbf{R}}^{2n}\times{\mathbf{R}}^{2n}\to{\mathbf{C}} is defined by

(χ¯χ)(ξ,ω):=χ(ξ)χ(ω).({\mathcal{F}_{\chi}}\bar{\otimes}{\mathcal{F}_{\chi}})(\xi,\omega):={\mathcal{F}_{\chi}}(\xi){\mathcal{F}_{\chi}}(-\omega). (3.4)
Proof.

We have:

ρχ(α,β)=χα|ρ|χβ=1(2π)nχα|χαχα|ρ|χβdα=1(2π)2nχα|χαχα|ρ|χβχβ|χβdαdβ=1(2π)2neiαα/2χ(αα)ρχ(α,β)χ(ββ)eiββ/2dαdβ,=1(2π)2n((χ¯χ)Ωρχ)(α,β)\begin{split}{\mathcal{M}^{\chi}_{\rho}}(\alpha,\beta)&=\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\beta}}\rangle}\\ &=\frac{1}{(2\pi)^{n}}\int\mathinner{\langle{\chi_{\alpha}|\chi_{\alpha^{\prime}}}\rangle}\mathinner{\langle{\chi_{\alpha^{\prime}}|\rho|\chi_{\beta}}\rangle}\mathop{}\!\mathrm{d}\alpha^{\prime}\\ &=\frac{1}{(2\pi)^{2n}}\int\mathinner{\langle{\chi_{\alpha}|\chi_{\alpha^{\prime}}}\rangle}\mathinner{\langle{\chi_{\alpha^{\prime}}|\rho|\chi_{\beta^{\prime}}}\rangle}\mathinner{\langle{\chi_{\beta^{\prime}}|\chi_{\beta}}\rangle}\mathop{}\!\mathrm{d}\alpha^{\prime}\mathop{}\!\mathrm{d}\beta^{\prime}\\ &=\frac{1}{(2\pi)^{2n}}\int e^{i\alpha^{\prime}\wedge\alpha/2}{\mathcal{F}_{\chi}}(\alpha^{\prime}-\alpha){\mathcal{M}^{\chi}_{\rho}}(\alpha^{\prime},\beta^{\prime}){\mathcal{F}_{\chi}}(\beta-\beta^{\prime})e^{i\beta\wedge\beta^{\prime}/2}\mathop{}\!\mathrm{d}\alpha^{\prime}\mathop{}\!\mathrm{d}\beta^{\prime},\\ &=\frac{1}{(2\pi)^{2n}}(({\mathcal{F}_{\chi}}\bar{\otimes}{\mathcal{F}_{\chi}})\circledast_{\Omega^{\prime}}{\mathcal{M}^{\chi}_{\rho}})(\alpha,\beta)\end{split} (3.5)

where to get the second line we use Eq. (2.13) of Lemma 2.9 for the inner product of |χα,ρ|χβL2(𝐑n)\mathinner{|{\chi_{\alpha}}\rangle},\rho\mathinner{|{\chi_{\beta}}\rangle}\in L^{2}({\mathbf{R}}^{n}) and to get the third line we use the lemma again for the inner product of |χβL2(𝐑n),ρ|χαL2(𝐑n)\mathinner{|{\chi_{\beta}}\rangle}\in L^{2}({\mathbf{R}}^{n}),\rho\mathinner{|{\chi_{\alpha}^{\prime}}\rangle}\in L^{2}({\mathbf{R}}^{n}). The final line is just the definition of the twisted convolution with respect to the form Ω\Omega^{\prime}. ∎

Corollary 3.4.

If the matrix element ρχ{\mathcal{M}^{\chi}_{\rho}} is rapidly decaying, then it is a Schwartz function.

Proof.

First note that χ\chi and therefore 𝒦χ(x,y)=χ(x)χ¯(y){\mathcal{K}_{\chi}}(x,y)=\chi(x)\bar{\chi}(y) are Schwartz functions. By Lemma 2.5 this means 𝒲χ{\mathcal{W}_{\chi}} is a Schwartz function, so therefore χ{\mathcal{F}_{\chi}} is a Schwartz function by Lemma 2.4, meaning that χ¯χ{\mathcal{F}_{\chi}}\bar{\otimes}{\mathcal{F}_{\chi}} is a Schwartz function. Then by Lemma 3.3, ρχ{\mathcal{M}^{\chi}_{\rho}} is the twisted convolution of a Schwartz function (χ¯χ{\mathcal{F}_{\chi}}\bar{\otimes}{\mathcal{F}_{\chi}}) against a rapidly decaying function (ρχ{\mathcal{M}^{\chi}_{\rho}}, by assumption), and is therefore also Schwartz-class by Lemma 2.6. ∎

We now deploy Lemma 2.4 to recover the Wigner function 𝒲ρ{\mathcal{W}_{\rho}} from the matrix element ρχ{\mathcal{M}^{\chi}_{\rho}}.

Lemma 3.5.

For any quantum state ρ\rho,

𝒲ρ(α)=1(2π)3nei(αβ/2)ξρχ(βξ/2,β+ξ/2)dξdβ.{\mathcal{W}_{\rho}}(\alpha)=\frac{1}{(2\pi)^{3n}}\int e^{-i(\alpha-\beta/2)\wedge\xi}{\mathcal{M}^{\chi}_{\rho}}(\beta-\xi/2,\beta+\xi/2)\mathop{}\!\mathrm{d}\xi\mathop{}\!\mathrm{d}\beta. (3.6)
Proof.

We have:

𝒲ρ(α)=1(2π)2neiαξtr[Dξ/2ρDξ/2]dξ=1(2π)3neiαξχβ|Dξ/2ρDξ/2|χβdξdβ=1(2π)3neiαξχ|DβDξ/2ρDξ/2Dβ|χdξdβ=1(2π)3nei(αβ/2)ξχ|Dξ/2βρDξ/2+β|χdξdβ=1(2π)3nei(αβ/2)ξρχ(βξ/2,β+ξ/2)dξdβ,\begin{split}{\mathcal{W}_{\rho}}(\alpha)&=\frac{1}{(2\pi)^{2n}}\int e^{-i\alpha\wedge\xi}\operatorname{tr}[D_{\xi/2}\rho D_{\xi/2}]\mathop{}\!\mathrm{d}\xi\\ &=\frac{1}{(2\pi)^{3n}}\int e^{-i\alpha\wedge\xi}\mathinner{\langle{\chi_{\beta}|D_{\xi/2}\rho D_{\xi/2}|\chi_{\beta}}\rangle}\mathop{}\!\mathrm{d}\xi\mathop{}\!\mathrm{d}\beta\\ &=\frac{1}{(2\pi)^{3n}}\int e^{-i\alpha\wedge\xi}\mathinner{\langle{\chi|D_{-\beta}D_{\xi/2}\rho D_{\xi/2}D_{\beta}|\chi}\rangle}\mathop{}\!\mathrm{d}\xi\mathop{}\!\mathrm{d}\beta\\ &=\frac{1}{(2\pi)^{3n}}\int e^{-i(\alpha-\beta/2)\wedge\xi}\mathinner{\langle{\chi|D_{\xi/2-\beta}\rho D_{\xi/2+\beta}|\chi}\rangle}\mathop{}\!\mathrm{d}\xi\mathop{}\!\mathrm{d}\beta\\ &=\frac{1}{(2\pi)^{3n}}\int e^{-i(\alpha-\beta/2)\wedge\xi}{\mathcal{M}^{\chi}_{\rho}}(\beta-\xi/2,\beta+\xi/2)\mathop{}\!\mathrm{d}\xi\mathop{}\!\mathrm{d}\beta,\end{split} (3.7)

where for the first line we use Lemma 2.4, for the second line we use the trace formula from Lemma 2.9, and for the fourth line we use the composition identity DαDβ=eiβα/2Dα+βD_{\alpha}D_{\beta}=e^{i\beta\wedge\alpha/2}D_{\alpha+\beta} for the displacement operator. ∎

Corollary 3.6.

For any quantum state ρ\rho, if the matrix element ρχ{\mathcal{M}^{\chi}_{\rho}} is a Schwartz function, then 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function.

Proof.

Lemma 3.5 shows that 𝒲ρ{\mathcal{W}_{\rho}} can be obtained from ρχ{\mathcal{M}^{\chi}_{\rho}} by (a) multiplying by the phase function eiβξ/2e^{i\beta\wedge\xi/2} (quadratic in the variables β\beta and ξ\xi), (b) applying a symplectic Fourier transform (exchanging the variable ξ\xi for the variable α\alpha), and then (c) integrating over the variable β\beta. All three of these operations preserve Schwartz-class functions, so 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function. ∎

With all the hard work done, our main result follows easily.

Theorem 1.1.

If ρ\rho is a positive semidefinite operator whose Wigner function 𝒲ρ{\mathcal{W}_{\rho}} exists and is rapidly decaying, then 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function.

Proof.

The rapid decay of 𝒲ρ{\mathcal{W}_{\rho}} means >𝒲ρ(α)dα=tr[ρ]\infty>\int{\mathcal{W}_{\rho}}(\alpha)\mathop{}\!\mathrm{d}\alpha=\operatorname{tr}[\rho], so ρ\rho is trace-class and hence a quantum state. We then conclude that ρχ{\mathcal{M}^{\chi}_{\rho}} is rapidly decaying by Corollary 3.2, so ρχ{\mathcal{M}^{\chi}_{\rho}} is a Schwartz function by Corollary 3.4. Therefore, 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function by Corollary 3.6. ∎

This proof is constructive and so can in principle be used to derive effective bounds for the Schwartz seminorms for 𝒲ρ{\mathcal{W}_{\rho}} in terms of only the decay seminorms. However, computing the bounds through this method would be very laborious, so instead we use a more direct method in the next subsection.

3.2. Direct proof

The strategy rests on showing that the Schwartz seminorms of 𝒲|χαχβ|{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}} depend only polynomially on α\alpha and β\beta. In this section, we will use for convenience the Schwartz type norms

Fa,b:=aabb|F|a,b\|F\|_{a,b}:=\sum_{a^{\prime}\leq a}\sum_{b^{\prime}\leq b}|F|_{a^{\prime},b^{\prime}} (3.8)

and the corresponding shorthand Fa=Fa,0=aa|F|a,0\|F\|_{a}=\|F\|_{a,0}=\sum_{a^{\prime}\leq a}|F|_{a^{\prime},0}.

Lemma 3.7.

For any quantum state ρ\rho and any reference wavefunction χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}),

𝒲ρ(γ)=1(2π)2n𝒲|χαχβ|(γ)ρχ(α,β)dαdβ.{\mathcal{W}_{\rho}}(\gamma)=\frac{1}{(2\pi)^{2n}}\int{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}(\gamma){\mathcal{M}^{\chi}_{\rho}}(\alpha,\beta)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}\beta. (3.9)
Proof.

Given the spectral decomposition ρ=jλj|ψjψj|\rho=\sum_{j}\lambda_{j}\mathinner{|{\psi_{j}}\rangle}\!\!\mathinner{\langle{\psi_{j}}|} we use Lemma 2.4 to get

𝒲ρ(γ)=1(2π)2neiγξtr[ρDξ]dξ=1(2π)4neiγξχβ|Dξ|χαχα|ρ|χβdαdβdξ=1(2π)4neiγξ|χαχβ|(ξ)ρχ(α,β)dαdβdξ\begin{split}{\mathcal{W}_{\rho}}(\gamma)&=\frac{1}{(2\pi)^{2n}}\int e^{-i\gamma\wedge\xi}\operatorname{tr}[\rho D_{\xi}]\mathop{}\!\mathrm{d}\xi\\ &=\frac{1}{(2\pi)^{4n}}\int e^{-i\gamma\wedge\xi}\mathinner{\langle{\chi_{\beta}|D_{\xi}|\chi_{\alpha}}\rangle}\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\beta}}\rangle}\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}\beta\mathop{}\!\mathrm{d}\xi\\ &=\frac{1}{(2\pi)^{4n}}\int e^{-i\gamma\wedge\xi}{\mathcal{F}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}(\xi){\mathcal{M}^{\chi}_{\rho}}(\alpha,\beta)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}\beta\mathop{}\!\mathrm{d}\xi\end{split} (3.10)

where to get the second line we apply Lemma 2.9 to the trace-class operator ρDξ\rho D_{\xi} twice. Using Lemma 2.4 yields (3.9). ∎

Lemma 3.8.

The Schwartz seminorms of the Wigner transform of the off-diagonal operator |χαχβ|\mathinner{|{\chi_{\alpha}}\rangle}\!\!\mathinner{\langle{\chi_{\beta}}|} obey

|𝒲|χαχβ||a,b\displaystyle|{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}|_{a,b} cbdaecfd(bc)(ad)(ce)(df)2|d||𝒲χ|ad,bc|αe^+fβc^e^+df|\displaystyle\leq\sum_{c\leq b}\sum_{d\leq a}\sum_{e\leq c}\sum_{f\leq d}\binom{b}{c}\binom{a}{d}\binom{c}{e}\binom{d}{f}2^{-|d|}|{\mathcal{W}_{\chi}}|_{a-d,b-c}|\alpha^{\hat{e}+f}\beta^{\hat{c}-\hat{e}+d-f}| (3.11)
4|a|+|b|(1+|α|+|β|)|a|+|b|𝒲χa,b,\displaystyle\leq 4^{|a|+|b|}(1+|\alpha|+|\beta|)^{|a|+|b|}\|{\mathcal{W}_{\chi}}\|_{a,b}, (3.12)

where we use a hat to swap the position and momentum components of a multi-index: a^=(ax,ap)^:=(ap,ax)\hat{a}=\widehat{(a_{\mathrm{x}},a_{\mathrm{p}})}:=(a_{\mathrm{p}},a_{\mathrm{x}}).

Proof.

First note that

|χαχβ|(ξ)=tr[|χαχβ|Dξ]=χ|DβDξDα|χ=ei(α+β)ξ/2+iβα/2χ|Dξ+αβ|χ=eiα¯ξ+iα¯Δα/2χ(ξ+Δα)\begin{split}{\mathcal{F}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}(\xi)&=\operatorname{tr}[\mathinner{|{\chi_{\alpha}}\rangle}\!\!\mathinner{\langle{\chi_{\beta}}|}D_{\xi}]=\mathinner{\langle{\chi|D_{-\beta}D_{\xi}D_{\alpha}|\chi}\rangle}\\ &=e^{i(\alpha+\beta)\wedge\xi/2+i\beta\wedge\alpha/2}\mathinner{\langle{\chi|D_{\xi+\alpha-\beta}|\chi}\rangle}=e^{i\bar{\alpha}\wedge\xi+i\bar{\alpha}\wedge\Delta\alpha/2}{\mathcal{F}_{\chi}}(\xi+\Delta\alpha)\end{split} (3.13)

where in the last line we introduced the shorthand α¯=(α+β)/2\bar{\alpha}=(\alpha+\beta)/2 and Δα=αβ\Delta\alpha=\alpha-\beta. Then 𝒲|χαχβ|{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}} is related to 𝒲χ{\mathcal{W}_{\chi}} by

𝒲|χαχβ|(γ)=eiγξ|χαχβ|(ξ)dξ=ei(α¯γ)ξ+iα¯Δα/2χ(ξ+Δα)dξ=ei(α¯γ)(ξΔα)+iα¯Δα/2χ(ξ)dξ=ei(γα¯/2)Δα𝒲χ(γα¯),\begin{split}{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}(\gamma)&=\int e^{-i\gamma\wedge\xi}{\mathcal{F}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}(\xi)\mathop{}\!\mathrm{d}\xi\\ &=\int e^{i(\bar{\alpha}-\gamma)\wedge\xi+i\bar{\alpha}\wedge\Delta\alpha/2}{\mathcal{F}_{\chi}}(\xi+\Delta\alpha)\mathop{}\!\mathrm{d}\xi\\ &=\int e^{i(\bar{\alpha}-\gamma)\wedge(\xi-\Delta\alpha)+i\bar{\alpha}\wedge\Delta\alpha/2}{\mathcal{F}_{\chi}}(\xi)\mathop{}\!\mathrm{d}\xi\\ &=e^{i(\gamma-\bar{\alpha}/2)\wedge\Delta\alpha}{\mathcal{W}_{\chi}}(\gamma-\bar{\alpha}),\end{split} (3.14)

so

γaγb𝒲|χαχβ|(γ)=γaγbei(γα¯/2)Δα𝒲χ(γα¯)=cb(bc)ei(γα¯/2)Δα(iΩΔα)c((γα¯)+α¯)aγbc𝒲χ(γα¯)\begin{split}\gamma^{a}\partial_{\gamma}^{b}{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}(\gamma)&=\gamma^{a}\partial_{\gamma}^{b}e^{i(\gamma-\bar{\alpha}/2)\wedge\Delta\alpha}{\mathcal{W}_{\chi}}(\gamma-\bar{\alpha})\\ &=\sum_{c\leq b}\binom{b}{c}e^{i(\gamma-\bar{\alpha}/2)\wedge\Delta\alpha}(i\Omega\cdot\Delta\alpha)^{c}((\gamma-\bar{\alpha})+\bar{\alpha})^{a}\partial_{\gamma}^{b-c}{\mathcal{W}_{\chi}}(\gamma-\bar{\alpha})\end{split} (3.15)

where in the second line we used

γbeiγξ=γxbxγpbpeiγxξpiγpξx=(iξp)bx(iξx)bpeiγxξpiγpξx=(iΩξ)beiγξ\begin{split}\partial_{\gamma}^{b}e^{i\gamma\wedge\xi}=\partial_{\gamma_{\mathrm{x}}}^{b_{\mathrm{x}}}\partial_{\gamma_{\mathrm{p}}}^{b_{\mathrm{p}}}e^{i\gamma_{\mathrm{x}}\cdot\xi_{\mathrm{p}}-i\gamma_{\mathrm{p}}\cdot\xi_{\mathrm{x}}}=(i\xi_{\mathrm{p}})^{b_{\mathrm{x}}}(-i\xi_{\mathrm{x}})^{b_{\mathrm{p}}}e^{i\gamma_{\mathrm{x}}\cdot\xi_{\mathrm{p}}-i\gamma_{\mathrm{p}}\cdot\xi_{\mathrm{x}}}=(i\Omega\cdot\xi)^{b}e^{i\gamma\wedge\xi}\end{split} (3.16)

Then, using |(iΩξ)b|=|ξpbxξxbp|=|ξb^||(i\Omega\cdot\xi)^{b}|=|\xi_{\mathrm{p}}^{b_{\mathrm{x}}}\xi_{\mathrm{x}}^{b_{\mathrm{p}}}|=|\xi^{\hat{b}}|, it follows that

|𝒲|χαχβ||a,bcbda(bc)(ad)|𝒲χ|ad,bc|Δαc^α¯d|cbdaecfd(bc)(ad)(ce)(df)2|d||𝒲χ|ad,bc|αe^+fβc^e^+df|4|a|+|b|(1+|α|+|β|)|a|+|b|𝒲χa,b.\begin{split}|{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}|_{a,b}&\leq\sum_{c\leq b}\sum_{d\leq a}\binom{b}{c}\binom{a}{d}|{\mathcal{W}_{\chi}}|_{a-d,b-c}|\Delta\alpha^{\hat{c}}\bar{\alpha}^{d}|\\ &\leq\sum_{c\leq b}\sum_{d\leq a}\sum_{e\leq c}\sum_{f\leq d}\binom{b}{c}\binom{a}{d}\binom{c}{e}\binom{d}{f}2^{-|d|}|{\mathcal{W}_{\chi}}|_{a-d,b-c}|\alpha^{\hat{e}+f}\beta^{\hat{c}-\hat{e}+d-f}|\\ &\leq 4^{|a|+|b|}(1+|\alpha|+|\beta|)^{|a|+|b|}\|{\mathcal{W}_{\chi}}\|_{a,b}.\end{split} (3.17)

To get the third line, we bound the terms in the sum on the second line using 2|d|12^{-|d|}\leq 1, |𝒲χ|ad,bc𝒲χa,b|{\mathcal{W}_{\chi}}|_{a-d,b-c}\leq\|{\mathcal{W}_{\chi}}\|_{a,b}, and |αe^+fβc^e^+df|(|1+|α|+|β|)|a|+|b||\alpha^{\hat{e}+f}\beta^{\hat{c}-\hat{e}+d-f}|\leq(|1+|\alpha|+|\beta|)^{|a|+|b|} and then sum the binomial coefficients. ∎

Theorem 3.9.

For any quantum state ρ\rho and reference state χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}), the following inequality holds:

|𝒲ρ|a,b\displaystyle|{\mathcal{W}_{\rho}}|_{a,b} (2π)5n24(|a|+|b|+n)𝒲χa,b𝒲χ2(a+b^)+6𝒲ρ2(a+b^)+4.\displaystyle\leq(2\pi)^{5n}2^{4(|a|+|b|+n)}\|{\mathcal{W}_{\chi}}\|_{a,b}\|{\mathcal{W}_{\chi}}\|_{2(a+\hat{b})+6}\|{\mathcal{W}_{\rho}}\|_{2(a+\hat{b})+4}. (3.18)

Note that the right-hand side contains only the decay norms of 𝒲ρ{\mathcal{W}_{\rho}}, and the left-hand side contains an arbitrary Schwartz seminorm, so this implies Theorem 1.1. We also observe that, on the right-hand side, the position and momentum indices are flipped in the derivative multi-index b^=(bx,bp)^=(bp,bx)\hat{b}=\widehat{(b_{\mathrm{x}},b_{\mathrm{p}})}=(b_{\mathrm{p}},b_{\mathrm{x}}) when it contributes to a coordinate power (rather than a derivative power).

Proof.

We start with Eq. (3.9) of Lemma 3.7 and apply Lemma 3.1 followed by Eq. (3.11) from Lemma 3.8 to get

|γaγa𝒲ρ(γ)|1(2π)2n|γaγa𝒲|χαχβ|(γ)||𝒬ρχ(α)|1/2|𝒬ρχ(β)|1/2dαdβ1(2π)2ncbdaecfd(bc)(ad)(ce)(df)2|d||𝒲χ|da,bc×1j=12n(1+|αj|2)|αe^+f|j=12n(1+|αj|2)|𝒬ρχ(α)|1/2dα×1j=12n(1+|βj|2)|βc^e^+df|j=12n(1+|βj|2)|𝒬ρχ(β)|1/2dβ.\begin{split}|\gamma^{a}\partial_{\gamma}^{a}{\mathcal{W}_{\rho}}(\gamma)|&\leq\frac{1}{(2\pi)^{2n}}\int|\gamma^{a}\partial_{\gamma}^{a}{\mathcal{W}_{\mathinner{|{\chi_{\alpha}}\rangle}\!\mathinner{\langle{\chi_{\beta}}|}}}(\gamma)||{\mathcal{Q}^{\chi}_{\rho}}(\alpha)|^{1/2}|{\mathcal{Q}^{\chi}_{\rho}}(\beta)|^{1/2}\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}\beta\\ &\leq\frac{1}{(2\pi)^{2n}}\sum_{c\leq b}\sum_{d\leq a}\sum_{e\leq c}\sum_{f\leq d}\binom{b}{c}\binom{a}{d}\binom{c}{e}\binom{d}{f}2^{-|d|}|{\mathcal{W}_{\chi}}|_{d-a,b-c}\\ &\qquad\times\int\frac{1}{\prod_{j=1}^{2n}(1+|\alpha_{j}|^{2})}|\alpha^{\hat{e}+f}|\prod_{j=1}^{2n}(1+|\alpha_{j}|^{2})|{\mathcal{Q}^{\chi}_{\rho}}(\alpha)|^{1/2}\mathop{}\!\mathrm{d}\alpha\\ &\qquad\times\int\frac{1}{\prod_{j=1}^{2n}(1+|\beta_{j}|^{2})}|\beta^{\hat{c}-\hat{e}+d-f}|\prod_{j=1}^{2n}(1+|\beta_{j}|^{2})|{\mathcal{Q}^{\chi}_{\rho}}(\beta)|^{1/2}\mathop{}\!\mathrm{d}\beta.\end{split} (3.19)

To compute the integral over α\alpha, we use

|αe^+f|j=12n(1+|αj2|)|𝒬ρχ(α)|1/222n𝒬ρχ2e^+2f+41/222n𝒬ρχ2(a+b^)+41/2,\begin{split}|\alpha^{\hat{e}+f}|\prod_{j=1}^{2n}(1+|\alpha_{j}^{2}|)|{\mathcal{Q}^{\chi}_{\rho}}(\alpha)|^{1/2}&\leq 2^{2n}\|{\mathcal{Q}^{\chi}_{\rho}}\|_{2\hat{e}+2f+4}^{1/2}\leq 2^{2n}\|{\mathcal{Q}^{\chi}_{\rho}}\|_{2(a+\hat{b})+4}^{1/2},\end{split} (3.20)

and likewise for the integral over β\beta. Integrating over α\alpha with (1+|αj|2)1dαj=π\int(1+|\alpha_{j}|^{2})^{-1}\mathop{}\!\mathrm{d}\alpha_{j}=\pi and likewise for β\beta, we obtain

|𝒲ρ|a,b\displaystyle|{\mathcal{W}_{\rho}}|_{a,b} (π2)2n𝒬ρχ2(a+b^)+4cbdaecfd(bc)(ad)(ce)(df)2|d||𝒲χ|da,bc\displaystyle\leq\left(\frac{\pi}{2}\right)^{2n}\|{\mathcal{Q}^{\chi}_{\rho}}\|_{2(a+\hat{b})+4}\sum_{c\leq b}\sum_{d\leq a}\sum_{e\leq c}\sum_{f\leq d}\binom{b}{c}\binom{a}{d}\binom{c}{e}\binom{d}{f}2^{-|d|}|{\mathcal{W}_{\chi}}|_{d-a,b-c}
(π2)2n22(|a|+|b|)𝒲χa,b𝒬ρχ2(a+b^)+4\displaystyle\leq\left(\frac{\pi}{2}\right)^{2n}2^{2(|a|+|b|)}\|{\mathcal{W}_{\chi}}\|_{a,b}\|{\mathcal{Q}^{\chi}_{\rho}}\|_{2(a+\hat{b})+4} (3.21)

using 2|d|12^{-|d|}\leq 1 and |𝒲χ|da,bc𝒲χa,b|{\mathcal{W}_{\chi}}|_{d-a,b-c}\leq\|{\mathcal{W}_{\chi}}\|_{a,b}. We then bound the decay seminorms of 𝒬ρχ{\mathcal{Q}^{\chi}_{\rho}} using Lemma 2.10:

αa𝒬ρχ(α)=(2π)n𝒲ρ(αβ)((αβ)+β)a𝒲χ(β)dβ\begin{split}\alpha^{a}{\mathcal{Q}^{\chi}_{\rho}}(\alpha)=(2\pi)^{n}\int{\mathcal{W}_{\rho}}(\alpha-\beta)((\alpha-\beta)+\beta)^{a}{\mathcal{W}_{\chi}^{-}}(\beta)\mathop{}\!\mathrm{d}\beta\end{split} (3.22)

so

|αa𝒬ρχ(α)|(2π)nba(ab)1j=12n(1+|βj|2)|(αβ)b𝒲ρ(αβ)|×|βab|j=12m(1+|βj|2)|𝒲χ(β)|dβ(2π)n2|a|π2n𝒲ρa𝒲χa+2\begin{split}|\alpha^{a}{\mathcal{Q}^{\chi}_{\rho}}(\alpha)|&\leq(2\pi)^{n}\sum_{b\leq a}\binom{a}{b}\int\frac{1}{\prod_{j=1}^{2n}(1+|\beta_{j}|^{2})}|(\alpha-\beta)^{b}{\mathcal{W}_{\rho}}(\alpha-\beta)|\\ &\qquad\qquad\qquad\qquad\qquad\times|\beta^{a-b}|\prod_{j=1}^{2m}(1+|\beta_{j}|^{2})|{\mathcal{W}_{\chi}^{-}}(\beta)|\mathop{}\!\mathrm{d}\beta\\ &\leq(2\pi)^{n}2^{|a|}\pi^{2n}\|{\mathcal{W}_{\rho}}\|_{a}\|{\mathcal{W}_{\chi}}\|_{a+2}\end{split} (3.23)

using 𝒲χa+2=𝒲χa+2\|{\mathcal{W}_{\chi}}\|_{a+2}=\|{\mathcal{W}_{\chi}^{-}}\|_{a+2}. Summing over the seminorms in the norm,

𝒬ρχa=ba|𝒬ρχ|b\displaystyle\|{\mathcal{Q}^{\chi}_{\rho}}\|_{a}=\sum_{b\leq a}|{\mathcal{Q}^{\chi}_{\rho}}|_{b} 2nπ3nba2|b|𝒲ρb𝒲χb+2\displaystyle\leq 2^{n}\pi^{3n}\sum_{b\leq a}2^{|b|}\|{\mathcal{W}_{\rho}}\|_{b}\|{\mathcal{W}_{\chi}}\|_{b+2} (3.24)
23n+|a|π3n𝒲ρa𝒲χa+2,\displaystyle\leq 2^{3n+|a|}\pi^{3n}\|{\mathcal{W}_{\rho}}\|_{a}\|{\mathcal{W}_{\chi}}\|_{a+2}, (3.25)

and then inserting into (3.21) yields (3.18). ∎

4. Schwartz states

In this section, we extend our main result by establishing an equivalence between the Schwartz-class and rapid-decay properties of many different representations of the quantum state. First, we will give a notion of Schwartz class to a set of orthogonal wavefunction {ψi}\{\psi_{i}\} appearing in a spectral decomposition 𝒦ρ(x,y)=iψi(x)ψ¯i(y){\mathcal{K}_{\rho}}(x,y)=\sum_{i}\psi_{i}(x)\bar{\psi}_{i}(y). Then, we recall the definition of a Schwartz operator as identified by Keyl et al. [keyl2016schwartz]. Finally, we will prove our large equivalence theorem.

To guarantee that 𝒦ρ{\mathcal{K}_{\rho}} is a Schwartz function, it is, of course, not sufficient for each ψi\psi_{i} to be a Schwartz function. For example, if each ψi\psi_{i} is a Gaussian wavepacket with increasing variance σi2i\sigma^{2}_{i}\propto i centered on the origin, then the overall variance X2=Tr[X2ρ]=x2𝒦ρ(x,x)dx=ix2|ψi(x)|2dx=ipiσi2\langle X^{2}\rangle=\operatorname{Tr}[X^{2}\rho]=\int x^{2}{\mathcal{K}_{\rho}}(x,x)\mathop{}\!\mathrm{d}x=\sum_{i}\int x^{2}|\psi_{i}(x)|^{2}\mathop{}\!\mathrm{d}x=\sum_{i}p_{i}\sigma_{i}^{2} can diverge if the norms pi=ψi|ψip_{i}=\mathinner{\langle{\psi_{i}|\psi_{i}}\rangle} are decreasing slowly, so that 𝒦ρ{\mathcal{K}_{\rho}} is not a Schwartz function even though each ψi\psi_{i} is. Instead, we consider the following definition.222Note that the multi-indices a,b,c,da,b,c,d in this section are nn-dimensional rather than 2n2n-dimensional because the wavefunction ψ\psi and the kernel 𝒦ρ{\mathcal{K}_{\rho}} take arguments in position space rather than phase space.

Definition 4.1.

A set {ψi}\{\psi_{i}\} of unnormalized wavefunctions (ψiL2(𝐑n)\psi_{i}\in L^{2}({\mathbf{R}}^{n}) for all ii) is jointly Schwartz when |{ψi}|a,b<|\{\psi_{i}\}|_{a,b}<\infty for all a,b({0})×na,b\in({\mathbb{N}}\cup\{0\})^{\times n}, where the Schwartz seminorms of such a set are defined by

|{ψi}|a,b2:=supxi|xaxbψi(x)|2.|\{\psi_{i}\}|_{a,b}^{2}:=\sup_{x}\sum_{i}\left|x^{a}\partial_{x}^{b}\psi_{i}(x)\right|^{2}. (4.1)

This is denoted {ψi}𝒮j(𝐑n)\{\psi_{i}\}\in{\mathcal{S}}_{\mathrm{j}}({\mathbf{R}}^{n}).

Note that this seminorm (4.1) is not simply a function of the seminorm |𝒦ρ|(a,c),(b,d)|{\mathcal{K}_{\rho}}|_{(a,c),(b,d)} of the kernel, nor is it simply a function of the individual seminorms |ψi|a,b:=supx|xaxbψi(x)||\psi_{i}|_{a,b}:=\sup_{x}|x^{a}\partial_{x}^{b}\psi_{i}(x)| of the wavefunctions ψi\psi_{i}. However, one can check that when the set {ψi}\{\psi_{i}\} is finite, the jointly Schwartz property is equivalent to the condition that all the wavefunctions are Schwartz functions individually, {ψi}𝒮(𝐑n)\{\psi_{i}\}\subset{\mathcal{S}}({\mathbf{R}}^{n}).

Lemma 4.2.

For any quantum state ρ\rho, the set of unnormalized wavefunctions {ψi}\{\psi_{i}\} of the spectral decomposition is jointly Schwartz if and only if the kernel 𝒦ρ{\mathcal{K}_{\rho}} is a Schwartz function.

Proof.

First assume that {ψi}𝒮j(𝐑n)\{\psi_{i}\}\in{\mathcal{S}}_{\mathrm{j}}({\mathbf{R}}^{n}). Then

|𝒦ρ|(a,c),(b,d)=supx,y|xaycxbydK(x,y)|=supx,y|i(xaxbψi(x))(ycydψ¯i(y))|supx,y(i|xaxbψi(x)|2)1/2(i|ycydψi(y)|2)1/2=(supxi|xaxbψi(x)|2)1/2(supyi|ycydψi(y)|2)1/2=|{ψi}|a,b|{ψi}|c,d,\begin{split}|{\mathcal{K}_{\rho}}|_{(a,c),(b,d)}&=\sup_{x,y}\left|x^{a}y^{c}\partial_{x}^{b}\partial_{y}^{d}K(x,y)\right|\\ &=\sup_{x,y}\left|\sum_{i}\left(x^{a}\partial_{x}^{b}\psi_{i}(x)\right)\left(y^{c}\partial_{y}^{d}\bar{\psi}_{i}(y)\right)\right|\\ &\leq\sup_{x,y}\left(\sum_{i}\left|x^{a}\partial_{x}^{b}\psi_{i}(x)\right|^{2}\right)^{1/2}\left(\sum_{i}\left|y^{c}\partial_{y}^{d}\psi_{i}(y)\right|^{2}\right)^{1/2}\\ &=\left(\sup_{x}\sum_{i}\left|x^{a}\partial_{x}^{b}\psi_{i}(x)\right|^{2}\right)^{1/2}\left(\sup_{y}\sum_{i}\left|y^{c}\partial_{y}^{d}\psi_{i}(y)\right|^{2}\right)^{1/2}\\ &=|\{\psi_{i}\}|_{a,b}|\{\psi_{i}\}|_{c,d},\end{split} (4.2)

where the third line is the Cauchy-Schwartz inequality. Therefore, {ψi}𝒮j(𝐑n)𝒦ρ𝒮(𝐑2n)\{\psi_{i}\}\in{\mathcal{S}}_{\mathrm{j}}({\mathbf{R}}^{n})\Rightarrow{\mathcal{K}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n}). To see the inverse, note that

|𝒦ρ|(a,a),(b,b)=supx,y|xayaxbybK(x,y)|supx|(xayaxbybK)(x,x)|=supxi|xaxbψi|2\begin{split}|{\mathcal{K}_{\rho}}|_{(a,a),(b,b)}&=\sup_{x,y}\left|x^{a}y^{a}\partial_{x}^{b}\partial_{y}^{b}K(x,y)\right|\\ &\geq\sup_{x}\left|\left(x^{a}y^{a}\partial_{x}^{b}\partial_{y}^{b}K\right)(x,x)\right|\\ &=\sup_{x}\sum_{i}|x^{a}\partial_{x}^{b}\psi_{i}|^{2}\\ \end{split} (4.3)

where the inequality holds because 𝒦ρ{\mathcal{K}_{\rho}} is a Schwartz function. This quantity diverges by definition if {ψi}𝒮j(𝐑n)\{\psi_{i}\}\notin{\mathcal{S}}_{\mathrm{j}}({\mathbf{R}}^{n}), implying 𝒦ρ𝒮(𝐑2n){\mathcal{K}_{\rho}}\notin{\mathcal{S}}({\mathbf{R}}^{2n}). ∎

The natural way to characterize the Schwartz class of quantum states was suggested by Keyl et al. [keyl2016schwartz] as those quantum states ρ\rho with bounded expectation value for all symmetric polynomials in XX and PP (i.e., tr[XaPbρPbXa]<\operatorname{tr}[X^{a}P^{b}\rho P^{b}X^{a}]<\infty for all a,b({0})×na,b\in(\mathbb{N}\cup\{0\})^{\times n}). More generally, for arbitrary operators (i.e., not necessarily positive semidefinite, self-adjoint, or trace-class), they define:

Definition 4.3.

An operator EE is a Schwartz operator, denoted E𝒮(L2(𝐑n))E\in{\mathcal{S}}(L^{2}({\mathbf{R}}^{n})), when |E|a,b,c,d<|E|_{a,b,c,d}<\infty for all a,b,c,d({0})×na,b,c,d\in({\mathbb{N}}\cup\{0\})^{\times n}, where the Schwartz operator seminorms are

|E|a,b,c,d:=sup|ψ|,|ϕ|=1|ψ|XaPbEPcXd|ϕ|.|E|_{a,b,c,d}:=\sup_{|\psi|,|\phi|=1}\left|\left\langle\psi\left|X^{a}P^{b}EP^{c}X^{d}\right|\phi\right\rangle\right|. (4.4)

Here, the supremum is taken over all normalized wavefunctions ψ,ϕL2(𝐑n)\psi,\phi\in L^{2}({\mathbf{R}}^{n}).

Equipped with Definitions 4.1 and 4.3, we can state a theorem that subsumes the results of Sec.  3.

Theorem 4.4.

For any Schwartz-class reference wavefunction χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}) and for any quantum state (i.e., positive semidefinite trace-class operator on L2(𝐑n)L^{2}({\mathbf{R}}^{n})) ρ\rho, with spectral decomposition {ψi}\{\psi_{i}\}, quasicharacteristic function ρ{\mathcal{F}_{\rho}}, Wigner function 𝒲ρ{\mathcal{W}_{\rho}}, kernel 𝒦ρ{\mathcal{K}_{\rho}}, Husimi function 𝒬ρχ{\mathcal{Q}^{\chi}_{\rho}}, and matrix element ρχ{\mathcal{M}^{\chi}_{\rho}}, the following conditions are equivalent:

  • 𝒲ρ𝒮(𝐑2n){\mathcal{W}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})

  • 𝒲ρ𝒟(𝐑2n){\mathcal{W}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n})

  • 𝒬ρχ𝒮(𝐑2n){\mathcal{Q}^{\chi}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})

  • 𝒬ρχ𝒟(𝐑2n){\mathcal{Q}^{\chi}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n})

  • ρχ𝒮(𝐑4d){\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{4d})

  • ρχ𝒟(𝐑4d){\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{4d})

  • ρ𝒮(𝐑2n){\mathcal{F}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})

  • 𝒦ρ𝒮(𝐑2n){\mathcal{K}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})

  • {ψi}𝒮j(𝐑n)\{\psi_{i}\}\in{\mathcal{S}}_{\mathrm{j}}({\mathbf{R}}^{n})

  • ρ𝒮(L2(𝐑n))\rho\in{\mathcal{S}}(L^{2}({\mathbf{R}}^{n}))

Furthermore, if the set {ψi}\{\psi_{i}\} is finite (e.g., if the state is pure, ρ=|ψψ|\rho=|\psi\rangle\!\langle\psi|), then the condition {ψi}𝒮(𝐑n)\{\psi_{i}\}\subset{\mathcal{S}}({\mathbf{R}}^{n}) is also equivalent to the above.

Proof.

We have:

𝒲ρ𝒮(𝐑2n)𝒲ρ𝒟(𝐑2n){\mathcal{W}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})\Rightarrow{\mathcal{W}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n}) because 𝒮(𝐑2n)𝒟(𝐑2n){\mathcal{S}}({\mathbf{R}}^{2n})\subset{\mathcal{D}}({\mathbf{R}}^{2n}).

𝒲ρ𝒟(𝐑2n)𝒬ρχ,ρχ𝒟(𝐑2n){\mathcal{W}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n})\Rightarrow{\mathcal{Q}^{\chi}_{\rho}},{\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n}) by Corollary 3.2.

ρχ𝒟(𝐑2n)ρχ𝒮(𝐑2n){\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n})\Rightarrow{\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n}) by Corollary 3.4.

ρχ𝒮(𝐑2n)𝒲ρ𝒮(𝐑2n){\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})\Rightarrow{\mathcal{W}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n}) by Corollary 3.6.

𝒬ρχ𝒮(𝐑2n)𝒬ρχ𝒟(𝐑2n){\mathcal{Q}^{\chi}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})\Rightarrow{\mathcal{Q}^{\chi}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n}) because 𝒮(𝐑2n)𝒟(𝐑2n){\mathcal{S}}({\mathbf{R}}^{2n})\subset{\mathcal{D}}({\mathbf{R}}^{2n}).

ρχ𝒮(𝐑2n)ρχ𝒟(𝐑2n){\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})\Rightarrow{\mathcal{M}^{\chi}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n}) because 𝒮(𝐑2n)𝒟(𝐑2n){\mathcal{S}}({\mathbf{R}}^{2n})\subset{\mathcal{D}}({\mathbf{R}}^{2n}).

𝒲ρ𝒮(𝐑2n)ρ𝒮(𝐑2n){\mathcal{W}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})\Leftrightarrow{\mathcal{F}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n}) by Lemma 2.4.

𝒲ρ𝒮(𝐑2n)𝒦ρ𝒮(𝐑2n){\mathcal{W}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})\Leftrightarrow{\mathcal{K}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n}) by Lemma 2.5.

𝒦ρ𝒮(𝐑2n){ψi}𝒮j(𝐑2n){\mathcal{K}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n})\Leftrightarrow\{\psi_{i}\}\in{\mathcal{S}}_{\mathrm{j}}({\mathbf{R}}^{2n}) by Lemma 4.2.

ρ𝒮(L2(𝐑n))𝒲ρ𝒮(𝐑2n)\rho\in{\mathcal{S}}(L^{2}({\mathbf{R}}^{n}))\Leftrightarrow{\mathcal{W}_{\rho}}\in{\mathcal{S}}({\mathbf{R}}^{2n}) by Proposition 3.18 in Ref. [keyl2016schwartz]. ∎

We say a quantum state satisfying the above equivalent conditions is a Schwartz state.

Note that 𝒦ρ𝒟(𝐑2n){\mathcal{K}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n}) is not an equivalent condition, being strictly weaker than the other conditions above.333For instance, the plateau wavefunction ψ(y)={1if 0y1,0otherwise}\psi(y)=\{1\,\mathrm{if}\,0\leq y\leq 1,0\,\mathrm{otherwise}\} is compactly supported in position space but in momentum space decays to infinity only as a polynomial. This is essentially because 𝒦ρ{\mathcal{K}_{\rho}} is a spatial representation, so momentum information is encoded only in its derivatives, whereas 𝒲ρ{\mathcal{W}_{\rho}}, 𝒬ρχ{\mathcal{Q}^{\chi}_{\rho}}, and ρχ{\mathcal{M}^{\chi}_{\rho}} are phase-space representations whose decay constrains both space and momentum features. Similarly, ρ𝒟(𝐑2n){\mathcal{F}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n}) is not an equivalent condition because rapid decay of the derivatives of 𝒲ρ{\mathcal{W}_{\rho}} does not assure that 𝒲ρ{\mathcal{W}_{\rho}} has rapid decay.444Consider the n=1n=1 quantum state ρ=k=0|ψkψk|\rho=\sum_{k=0}^{\infty}|\psi_{k}\rangle\!\langle\psi_{k}| with ψk(y)=ψ0(yzk)=(6/π2)k2exp[(yzk)2/2]/2π\psi_{k}(y)=\psi_{0}(y-z_{k})=(6/\pi^{2})k^{-2}\exp[-(y-z_{k})^{2}/2]/\sqrt{2\pi} with zk=k3z_{k}=k^{3}. This is a mixture of Gaussians of equal variance, so the derivatives are all rapidly decreasing and ρ𝒟(𝐑2n){\mathcal{F}_{\rho}}\in{\mathcal{D}}({\mathbf{R}}^{2n}), but the mean tr[ρX]\operatorname{tr}[\rho X] diverges so 𝒲ρ{\mathcal{W}_{\rho}} is not a Schwartz function.

5. Discussion

Although the Wigner formalism provides a complete representation of quantum states and dynamics, it is often regarded as less fundamental. (It only really becomes uniquely preferred in the classical limit, and under, e.g., certain symmetry demands to distinguish it from other deformations of classical mechanics; see for instance the introduction of Ref. [degosson2016bornjordan].) One practical reason is that computations are often more difficult using the Moyal product555Of course, the peculiar features of the Moyal product are not just a matter of practicalities: Because the Moyal bracket has phase-space derivatives of arbitrarily high order, the dynamics of the Wigner function are non-local., the so-called “\star-genvalue equations”, and so on [curtright2014concise]. Another reason is that the state space is awkward to define.

Formally, the state space of valid Wigner functions can be delineated with the quantum generalization [srinivas1975nonclassical, broecker1995mixed] of Bochner’s theorem [bochner1933monotone]. (See also illuminating discussion and further generalizations to some discrete spaces in Ref. [dangniam2015quantum].) This definition is sufficiently opaque that most physicists simply think of the allowed pure-state Wigner functions as the image of the Wigner transform of the space of allowed quantum states, L2(𝐑n)L^{2}({\mathbf{R}}^{n}), if they think of it at all. In particular, many simple (even positive-valued) functions on L1(𝐑2n)L^{1}({\mathbf{R}}^{2n}) are not the Wigner functions of any quantum states.

In contrast, the L2(𝐑n)L^{2}({\mathbf{R}}^{n}) (pure) state space of the Schrödinger representation is relatively simple to understand, and the “parameterization” of that space is natural in the sense that all possible functions are allowed modulo only the single, easy-to-interpret constraint of normalization. This becomes even clearer in the case of a finite-dimensional quantum system, where there are no complications related to the continuum and where any complex-valued function over configuration space suffices as a (not necessarily normalized) state. Delineating the corresponding set of Wigner functions for finite-dimensional systems is much more subtle [dangniam2015quantum].

In this sense, the Wigner representation is “overparameterized”. One can think of our Theorem 1.1 as better characterizing this overparameterization: the regularity of the interior of the Wigner function in terms of its derivatives of any order is tightly controlled by the Wigner function’s decay toward infinity, a feature that is obviously not shared by all normalized functions over 𝐑2n{\mathbf{R}}^{2n}. With the seminorm bound of Theorem 3.9, one can also recover a version of the uncertainty principle. For example, because the supremum of the gradient of 𝒲ρ{\mathcal{W}_{\rho}} is bounded by the decay seminorms of 𝒲ρ{\mathcal{W}_{\rho}}, it is impossible for 𝒲ρ{\mathcal{W}_{\rho}} to be supported in a ball of too small a radius.

6. Acknowledgements

CJR thanks Jukka Kiukas and Reinhard Werner for helpful discussion. FH was supported by the Fannie and John Hertz Foundation Fellowship.

Appendix A

Here we here recall the proofs of some standard results referenced in the main body of this paper.

Lemma 2.2.

For any ϕL2(𝐑n)\phi\in L^{2}({\mathbf{R}}^{n}),

Dξϕ(y)=ei(yξx/2)ξpϕ(yξx).D_{\xi}\phi(y)=e^{i(y-\xi_{\mathrm{x}}/2)\cdot\xi_{\mathrm{p}}}\phi(y-\xi_{\mathrm{x}}). (A.1)
Proof.

First, let us prove the statement for some Schwartz function χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}). Consider the following differential equation:

tft=iRξft\partial_{t}f_{t}=iR\wedge\xi f_{t} (A.2)

Solutions ftf_{t} preserve the L2L^{2} norm (because the operator on the right is anti-Hermitian), so solutions are unique. Moreover, since Dtξf0D_{t\xi}f_{0} is a solution to (A.2), any solution to (A.2) with initial condition f0=χf_{0}=\chi satisfies f1=Dξχf_{1}=D_{\xi}\chi.

It remains to check that the function

ft(y)=ei(ytξx/2)(tξp)χ(ytξx)f_{t}(y)=e^{i(y-t\xi_{\mathrm{x}}/2)\cdot(t\xi_{\mathrm{p}})}\chi(y-t\xi_{\mathrm{x}}) (A.3)

solves (A.2). We check this by a direct computation,

tft(y)=i((ytξx)ξp)ft(y)ei(ytξx/2)(tξp)ξx(χ)(ytξx)=iξpyft(y)itξxξpft(y)ei(ytξx/2)(tξp)ξx(χ)(ytξx)=iξpXft(y)iξxPft(y).\begin{split}\partial_{t}f_{t}(y)&=i((y-t\xi_{\mathrm{x}})\cdot\xi_{\mathrm{p}})f_{t}(y)-e^{i(y-t\xi_{\mathrm{x}}/2)\cdot(t\xi_{\mathrm{p}})}\xi_{\mathrm{x}}\cdot(\nabla\chi)(y-t\xi_{\mathrm{x}})\\ &=i\xi_{\mathrm{p}}\cdot yf_{t}(y)-it\xi_{\mathrm{x}}\cdot\xi_{\mathrm{p}}f_{t}(y)-e^{i(y-t\xi_{\mathrm{x}}/2)\cdot(t\xi_{\mathrm{p}})}\xi_{\mathrm{x}}\cdot(\nabla\chi)(y-t\xi_{\mathrm{x}})\\ &=i\xi_{\mathrm{p}}\cdot Xf_{t}(y)-i\xi_{\mathrm{x}}\cdot Pf_{t}(y).\end{split} (A.4)

Having proven the claim for χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}), we can extend it to any ϕL2(𝐑n)\phi\in L^{2}({\mathbf{R}}^{n}) by using the density of the Schwartz function in L2(𝐑n)L^{2}({\mathbf{R}}^{n}), i.e., by approximating DξϕD_{\xi}\phi to accuracy ϵ\epsilon with some choice of Dξχ(ϵ)𝒮(𝐑n)D_{\xi}\chi^{(\epsilon)}\in{\mathcal{S}}({\mathbf{R}}^{n}) and taking ϵ0\epsilon\to 0. ∎

Lemma 2.4.

For any trace-class kernel operator EE, the corresponding Wigner transform and quasicharacteristic transform are symplectic Fourier duals:

𝒲E(α)=1(2π)2neiαξE(ξ)dξ.{\mathcal{W}_{E}}(\alpha)=\frac{1}{(2\pi)^{2n}}\int e^{-i\alpha\wedge\xi}{\mathcal{F}_{E}}(\xi)\mathop{}\!\mathrm{d}\xi. (A.5)
Proof.

By linearity it suffices to check the case that EE is a rank-1 state of the form E=|ψϕ|E=\mathinner{|{\psi}\rangle}\!\!\mathinner{\langle{\phi}|}. Moreover, by the continuity of the Wigner transform in L2(𝐑n)×L2(𝐑n)L^{2}({\mathbf{R}}^{n})\times L^{2}({\mathbf{R}}^{n}) and the density of Schwartz functions in L2(𝐑n)L^{2}({\mathbf{R}}^{n}), we may furthermore assume that ψ,ϕ𝒮(𝐑n)\psi,\phi\in{\mathcal{S}}({\mathbf{R}}^{n}). In this case,

tr[EDξ]=ϕ|Dξ|ψ=ϕ|Dξ/2Dξ/2|ψ=Dξ/2ϕ,Dξ/2ψ=ei(z+ξx/4)ξp/2ei(zξx/4)ξp/2ϕ¯(z+ξx/2)ψ(zξx/2)dz=eizξpϕ¯(z+ξx/2)ψ(zξx/2)dz\begin{split}\operatorname{tr}[ED_{\xi}]&=\mathinner{\langle{\phi|D_{\xi}|\psi}\rangle}=\mathinner{\langle{\phi|D_{\xi/2}D_{\xi/2}|\psi}\rangle}=\langle D_{-\xi/2}\phi,D_{\xi/2}\psi\rangle\\ &=\int e^{i(z+\xi_{\mathrm{x}}/4)\cdot\xi_{\mathrm{p}}/2}e^{i(z-\xi_{\mathrm{x}}/4)\cdot\xi_{\mathrm{p}}/2}\bar{\phi}(z+\xi_{\mathrm{x}}/2)\psi(z-\xi_{\mathrm{x}}/2)\mathop{}\!\mathrm{d}z\\ &=\int e^{iz\cdot\xi_{\mathrm{p}}}\bar{\phi}(z+\xi_{\mathrm{x}}/2)\psi(z-\xi_{\mathrm{x}}/2)\mathop{}\!\mathrm{d}z\end{split} (A.6)

where to get the second line we use Lemma 2.2. Applying this identity into the right-hand side of (A.5), we get

1(2π)2neiαξE(ξ)dξ=1(2π)2neiαxξp+iαpξxeizξpϕ¯(z+ξx/2)ψ(zξx/2)dzdξxdξp=1(2π)neiαpξxϕ¯(αx+ξx/2)ψ(αxξx/2)dξx,\begin{split}\frac{1}{(2\pi)^{2n}}\int e^{-i\alpha\wedge\xi}{\mathcal{F}_{E}}(\xi)\mathop{}\!\mathrm{d}\xi&=\frac{1}{(2\pi)^{2n}}\int e^{-i\alpha_{\mathrm{x}}\cdot\xi_{\mathrm{p}}+i\alpha_{\mathrm{p}}\cdot\xi_{\mathrm{x}}}e^{iz\cdot\xi_{\mathrm{p}}}\bar{\phi}(z+\xi_{\mathrm{x}}/2)\psi(z-\xi_{\mathrm{x}}/2)\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}\xi_{\mathrm{x}}\mathop{}\!\mathrm{d}\xi_{\mathrm{p}}\\ &=\frac{1}{(2\pi)^{n}}\int e^{i\alpha_{\mathrm{p}}\cdot\xi_{\mathrm{x}}}\bar{\phi}(\alpha_{\mathrm{x}}+\xi_{\mathrm{x}}/2)\psi(\alpha_{\mathrm{x}}-\xi_{\mathrm{x}}/2)\mathop{}\!\mathrm{d}\xi_{\mathrm{x}},\end{split} (A.7)

where the last line follows from the Fourier inversion formula ei(zαx)ξpf(αx)dzdξp=(2π)nf(z)\int e^{i(z-\alpha_{\mathrm{x}})\cdot\xi_{\mathrm{p}}}f(\alpha_{\mathrm{x}})\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}\xi_{\mathrm{p}}=(2\pi)^{n}f(z). This is the definition of the Wigner transform 𝒲E(α){\mathcal{W}_{E}}(\alpha) of E=|ψϕ|E=\mathinner{|{\psi}\rangle}\!\!\mathinner{\langle{\phi}|}. ∎

Lemma 2.5.

The Wigner function 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function if and only if the kernel 𝒦ρ{\mathcal{K}_{\rho}} is a Schwartz function.

Proof.

First assume the kernel 𝒦ρ(x,y){\mathcal{K}_{\rho}}(x,y) is a Schwartz function. Then the function g(z,Δz)=𝒦ρ(zΔz/2,z+Δz/2)g(z,\Delta z)={\mathcal{K}_{\rho}}(z-\Delta z/2,z+\Delta z/2) is a Schwartz function since it is related to 𝒦ρ{\mathcal{K}_{\rho}} merely by a linear change in variables (a 4545^{\circ} rotation). And of course we have 𝒲ρ(x,p)=(2π)neipΔzg(x,Δz)dΔz{\mathcal{W}_{\rho}}(x,p)=(2\pi)^{-n}\int e^{ip\cdot\Delta z}g(x,\Delta z)\mathop{}\!\mathrm{d}\Delta z, so 𝒲ρ{\mathcal{W}_{\rho}} must also be a Schwartz function since it is just the nn-dimensional Fourier transform of gg (exchanging the variable Δz\Delta z for pp but leaving the variable zz). The argument works the same in the opposite direction, so we conclude that 𝒦ρ{\mathcal{K}_{\rho}} is a Schwartz function if and only if 𝒲ρ{\mathcal{W}_{\rho}} is a Schwartz function. ∎

Lemma 2.6.

The twisted convolution of a rapidly decaying function with a Schwartz function is itself a Schwartz function.

Proof.

Let F𝒮(𝐑2n)F\in{\mathcal{S}}({\mathbf{R}}^{2n}) and G𝒟(𝐑2n)G\in{\mathcal{D}}({\mathbf{R}}^{2n}). Then recall the definition of the twisted convolution,

FΩG(α)=eiαΩα/2F(αα)G(α)dα.F\circledast_{\Omega^{\prime}}G(\alpha)=\int e^{i\alpha\cdot\Omega^{\prime}\cdot\alpha^{\prime}/2}F(\alpha-\alpha^{\prime})G(\alpha^{\prime})\mathop{}\!\mathrm{d}\alpha^{\prime}. (A.8)

We claim that for any multi-index a=(a1,,a2n)({0})×2na=(a_{1},\dots,a_{2n})\in(\mathbb{N}\cup\{0\})^{\times 2n}, there exist constants C(a,a)C(a,a^{\prime}) such that

a(FΩG)(α)=aaC(a,a)eiαΩα/2(aF)(αα)(iΩα)aaG(α)dα.\partial^{a}(F\circledast_{\Omega^{\prime}}G)(\alpha)=\sum_{a^{\prime}\leq a}C(a,a^{\prime})\int e^{i\alpha\cdot\Omega^{\prime}\cdot\alpha^{\prime}/2}(\partial^{a^{\prime}}F)(\alpha-\alpha^{\prime})(i\Omega^{\prime}\cdot\alpha^{\prime})^{a-a^{\prime}}G(\alpha^{\prime})\mathop{}\!\mathrm{d}\alpha^{\prime}. (A.9)

Equation (A.9) is easily checked by induction on |a||a|. Now applying the triangle inequality we can estimate

|a(FΩG)(α)|aaC(α,α)|aF(αα)||(Ωα)aaG(α)|dα,|\partial^{a}(F\circledast_{\Omega^{\prime}}G)(\alpha)|\leq\sum_{a^{\prime}\leq a}C(\alpha,\alpha^{\prime})\int|\partial^{a^{\prime}}F(\alpha-\alpha^{\prime})||(\Omega^{\prime}\cdot\alpha^{\prime})^{a-a^{\prime}}G(\alpha^{\prime})|\mathop{}\!\mathrm{d}\alpha^{\prime}, (A.10)

which is the finite sum of convolutions of the rapidly decaying functions |αaF(α)||\partial_{\alpha}^{a^{\prime}}F(\alpha)| and |(Ωα)aaG(α)||(\Omega^{\prime}\cdot\alpha)^{a-a^{\prime}}G(\alpha)|. Therefore a(FG)\partial^{a}(F\circledast G) is rapidly decaying. Since every partial derivative of FGF\circledast G is rapidly decaying, it follows that FG𝒮(𝐑2n)F\circledast G\in{\mathcal{S}}({\mathbf{R}}^{2n}). ∎

(Note that a similar statement for normal convolutions can be proven in almost exactly the same way.)

Lemma 2.7.

For any two quantum states ρ\rho and η\eta,

tr[ρη]=(2π)n𝒲ρ(α)𝒲η(α)dα.\displaystyle\operatorname{tr}[\rho\eta]=(2\pi)^{n}\int{\mathcal{W}_{\rho}}(\alpha){\mathcal{W}_{\eta}}(\alpha)\mathop{}\!\mathrm{d}\alpha. (A.11)
Proof.

We first decompose ρ\rho and η\eta using the spectral theorem,

ρ=jλj|ψjψj|,η=jμj|ϕjϕj|,\rho=\sum_{j}\lambda_{j}\mathinner{|{\psi_{j}}\rangle}\!\!\mathinner{\langle{\psi_{j}}|},\qquad\qquad\eta=\sum_{j}\mu_{j}\mathinner{|{\phi_{j}}\rangle}\!\!\mathinner{\langle{\phi_{j}}|}, (A.12)

with λj,μj0\lambda_{j},\mu_{j}\geq 0 and jλj=jμj=1\sum_{j}\lambda_{j}=\sum_{j}\mu_{j}=1. Then tr[ρη]=j,kλjμk|ψj|ϕk|2\operatorname{tr}[\rho\eta]=\sum_{j,k}\lambda_{j}\mu_{k}|\!\mathinner{\langle{\psi_{j}|\phi_{k}}\rangle}\!|^{2}. The result then follows from

𝒲ψ(α)𝒲ϕ(α)dα=1(2π)2neip(y+z)𝒦ρ(xy/2,x+y/2)×𝒦η(xz/2,x+z/2)dydzdxdp=1(2π)nψ(xy/2)ψ¯(x+y/2)ϕ(x+y/2)ϕ¯(xy/2)dydx=1(2π)nψ(x)ψ¯(x+)ϕ(x+)ϕ¯(x)dx+dx=1(2π)n|ψ|ϕ|2,\begin{split}\int{\mathcal{W}_{\psi}}(\alpha){\mathcal{W}_{\phi}}(\alpha)\mathop{}\!\mathrm{d}\alpha&=\frac{1}{(2\pi)^{2n}}\int e^{ip\cdot(y+z)}{\mathcal{K}_{\rho}}(x-y/2,x+y/2)\\ &\qquad\qquad\qquad\times{\mathcal{K}_{\eta}}(x-z/2,x+z/2)\mathop{}\!\mathrm{d}y\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}p\\ &=\frac{1}{(2\pi)^{n}}\int\psi(x-y/2)\bar{\psi}(x+y/2)\phi(x+y/2)\bar{\phi}(x-y/2)\mathop{}\!\mathrm{d}y\mathop{}\!\mathrm{d}x\\ &=\frac{1}{(2\pi)^{n}}\int\psi(x_{-})\bar{\psi}(x_{+})\phi(x_{+})\bar{\phi}(x_{-})\mathop{}\!\mathrm{d}x_{+}\mathop{}\!\mathrm{d}x_{-}\\ &=\frac{1}{(2\pi)^{n}}|\mathinner{\langle{\psi|\phi}\rangle}|^{2},\end{split} (A.13)

along with 𝒲ρ=jλj𝒲ψj{\mathcal{W}_{\rho}}=\sum_{j}\lambda_{j}{\mathcal{W}_{\psi_{j}}} and 𝒲μ=jμj𝒲ϕj{\mathcal{W}_{\mu}}=\sum_{j}\mu_{j}{\mathcal{W}_{\phi_{j}}}. The sums can be interchanged because everything converges absolutely. ∎

Lemma 2.9.

For any trace-class operator EE and any χL2(𝐑n)\chi\in L^{2}({\mathbf{R}}^{n}) satisfying χL2(𝐑n)=1\|\chi\|_{L^{2}({\mathbf{R}}^{n})}=1,

tr[E]=1(2π)nχα|E|χαdα.\displaystyle\operatorname{tr}[E]=\frac{1}{(2\pi)^{n}}\int\mathinner{\langle{\chi_{\alpha}|E|\chi_{\alpha}}\rangle}\mathop{}\!\mathrm{d}\alpha. (A.14)

In particular, for any ϕ,ψL2(𝐑n)\phi,\psi\in L^{2}({\mathbf{R}}^{n})

ϕ|ψ=1(2π)nϕ|χαχα|ψdα.\displaystyle\mathinner{\langle{\phi|\psi}\rangle}=\frac{1}{(2\pi)^{n}}\int\mathinner{\langle{\phi|\chi_{\alpha}}\rangle}\mathinner{\langle{\chi_{\alpha}|\psi}\rangle}\mathop{}\!\mathrm{d}\alpha. (A.15)
Proof.

We start by proving (A.15). For ψ,ϕL2(𝐑n)\psi,\phi\in L^{2}({\mathbf{R}}^{n}),

ϕ|χαχα|gdα=[(ϕ¯(z)ei(zαx/2)αpχ(zαx)dz)×(ψ(z)χ¯(zαx)ei(zαx/2)αpdz)]dα=ϕ¯(z)ψ(z)ei(zz)αpχ(zαx)χ¯(zαx)dzdzdαxdαp=(2π)nϕ¯(z)ψ(z)(|χ(zαx)|2dαx)dz=(2π)nϕ|ψ\begin{split}\int\mathinner{\langle{\phi|\chi_{\alpha}}\rangle}\mathinner{\langle{\chi_{\alpha}|g}\rangle}\mathop{}\!\mathrm{d}\alpha&=\int\Bigg{[}\left(\int\bar{\phi}(z)e^{i(z-\alpha_{\mathrm{x}}/2)\cdot\alpha_{\mathrm{p}}}\chi(z-\alpha_{\mathrm{x}})\mathop{}\!\mathrm{d}z\right)\\ &\qquad\quad\times\left(\int\psi(z)\bar{\chi}(z^{\prime}-\alpha_{\mathrm{x}})e^{-i(z^{\prime}-\alpha_{\mathrm{x}}/2)\cdot\alpha_{\mathrm{p}}}\mathop{}\!\mathrm{d}z^{\prime}\right)\Bigg{]}\mathop{}\!\mathrm{d}\alpha\\ &=\int\bar{\phi}(z)\psi(z^{\prime})e^{i(z-z^{\prime})\cdot\alpha_{\mathrm{p}}}\chi(z-\alpha_{\mathrm{x}})\bar{\chi}(z^{\prime}-\alpha_{\mathrm{x}})\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}z^{\prime}\mathop{}\!\mathrm{d}\alpha_{\mathrm{x}}\mathop{}\!\mathrm{d}\alpha_{\mathrm{p}}\\ &=(2\pi)^{n}\int\bar{\phi}(z)\psi(z)\left(\int|\chi(z-\alpha_{\mathrm{x}})|^{2}\mathop{}\!\mathrm{d}\alpha_{\mathrm{x}}\right)\mathop{}\!\mathrm{d}z\\ &=(2\pi)^{n}\mathinner{\langle{\phi|\psi}\rangle}\end{split} (A.16)

where to get from the second to the third line we use the Fourier inversion formula.

Then, if EE is any trace-class operator, we can write using the singular value decomposition (which one can obtain from the spectral theorem applied to the polar decomposition E=UEEE=U\sqrt{EE^{\dagger}})

E=jσj|ϕjψj|E=\sum_{j}\sigma_{j}\mathinner{|{\phi_{j}}\rangle}\!\!\mathinner{\langle{\psi_{j}}|} (A.17)

for some orthonormal bases ϕj\phi_{j}, ψj\psi_{j} on L2(𝐑n)L^{2}({\mathbf{R}}^{n}). Here the singular values σj\sigma_{j} are nonnegative and satisfy jσj=E1<\sum_{j}\sigma_{j}=\|E\|_{1}<\infty. In this case we can expand the trace using this sum and apply (A.16)

tr[E]=jσjψj|ϕj=jσj1(2π)nχα|ϕjψj|χαdα=1(2π)nχα|E|χαdα.\begin{split}\operatorname{tr}[E]&=\sum_{j}\sigma_{j}\mathinner{\langle{\psi_{j}|\phi_{j}}\rangle}\\ &=\sum_{j}\sigma_{j}\frac{1}{(2\pi)^{n}}\int\mathinner{\langle{\chi_{\alpha}|\phi_{j}}\rangle}\mathinner{\langle{\psi_{j}|\chi_{\alpha}}\rangle}\mathop{}\!\mathrm{d}\alpha\\ &=\frac{1}{(2\pi)^{n}}\int\mathinner{\langle{\chi_{\alpha}|E|\chi_{\alpha}}\rangle}\mathop{}\!\mathrm{d}\alpha.\end{split} (A.18)

The last line is obtained by swapping the integral and the sum, which can be done because the sum is absolutely convergent. ∎

Lemma 2.10.

For any quantum state ρ\rho and reference wavefunction χ𝒮(𝐑n)\chi\in{\mathcal{S}}({\mathbf{R}}^{n}),

𝒬ρχ(α)\displaystyle{\mathcal{Q}^{\chi}_{\rho}}(\alpha) =(2π)n(𝒲ρ𝒲χ)(α)=(2π)n𝒲ρ(β)𝒲χ(βα)dβ\displaystyle=(2\pi)^{n}({\mathcal{W}_{\rho}}\ast{\mathcal{W}_{\chi}^{-}})(\alpha)=(2\pi)^{n}\int{\mathcal{W}_{\rho}}(\beta){\mathcal{W}_{\chi}}(\beta-\alpha)\mathop{}\!\mathrm{d}\beta (A.19)

where 𝒲χ(α):=𝒲χ(α){{\mathcal{W}_{\chi}^{-}}}(\alpha):={\mathcal{W}_{\chi}}(-\alpha) is a Schwartz function.

Proof.

We have

𝒬ρχ(α)\displaystyle{\mathcal{Q}^{\chi}_{\rho}}(\alpha) =χα|ρ|χα=tr[ρ(|χαχα|)]\displaystyle=\mathinner{\langle{\chi_{\alpha}|\rho|\chi_{\alpha}}\rangle}=\operatorname{tr}[\rho(\mathinner{|{\chi_{\alpha}}\rangle}\!\!\mathinner{\langle{\chi_{\alpha}}|})] (A.20)
=(2π)n𝒲ρ(β)𝒲Dα|χχ|Dα(β)dβ\displaystyle=(2\pi)^{n}\int{\mathcal{W}_{\rho}}(\beta){\mathcal{W}_{D_{\alpha}\mathinner{|{\chi}\rangle}\!\mathinner{\langle{\chi}|}D_{\alpha}^{\dagger}}}(\beta)\mathop{}\!\mathrm{d}\beta (A.21)
=(2π)n𝒲ρ(β)𝒲χ(βα)dβ\displaystyle=(2\pi)^{n}\int{\mathcal{W}_{\rho}}(\beta){\mathcal{W}_{\chi}}(\beta-\alpha)\mathop{}\!\mathrm{d}\beta (A.22)

where we get the second line from Lemma 2.7 and the third line from the fact that the map ηDαηDα\eta\mapsto D_{\alpha}\eta D_{\alpha}^{\dagger} on quantum states corresponds to a displacement of the Wigner function by α\alpha:

𝒲DαηDα(β)\displaystyle{\mathcal{W}_{D_{\alpha}\eta D_{\alpha}^{\dagger}}}(\beta) =1(2π)2neiβξDαηDα(ξ)dξ\displaystyle=\frac{1}{(2\pi)^{2n}}\int e^{-i\beta\wedge\xi}{\mathcal{F}_{D_{\alpha}\eta D_{\alpha}^{\dagger}}}(\xi)\mathop{}\!\mathrm{d}\xi (A.23)
=1(2π)2neiβξtr[DαηDαDξ]dξ\displaystyle=\frac{1}{(2\pi)^{2n}}\int e^{-i\beta\wedge\xi}\operatorname{tr}[D_{\alpha}\eta D_{-\alpha}D_{\xi}]\mathop{}\!\mathrm{d}\xi (A.24)
=1(2π)2nei(βα)ξtr[ηDξ]dξ\displaystyle=\frac{1}{(2\pi)^{2n}}\int e^{-i(\beta-\alpha)\wedge\xi}\operatorname{tr}[\eta D_{\xi}]\mathop{}\!\mathrm{d}\xi (A.25)
=𝒲η(βα)\displaystyle={\mathcal{W}_{\eta}}(\beta-\alpha) (A.26)

Furthermore, since χ\chi is a Schwartz function, so is 𝒦χ(x,y)=χ¯(x)χ(y){\mathcal{K}_{\chi}}(x,y)=\bar{\chi}(x)\chi(y), and hence by Lemma 2.5 we have that 𝒲χ(α)=𝒲χ(α){\mathcal{W}_{\chi}^{-}}(\alpha)={\mathcal{W}_{\chi}}(-\alpha) is a Schwartz function. ∎

References