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

Segal-Bargmann Transforms Associated to
a Family of Coupled Supersymmetries

Cameron L. Williams Department of Mathematics
Embry-Riddle Aeronautical University
3700 Willow Creek Road
Prescott, AZ 86301
Abstract.

The Segal-Bargmann transform is a Lie algebra and Hilbert space isomorphism between real and complex representations of the oscillator algebra. The Segal-Bargmann transform is useful in time-frequency analysis as it is closely related to the short-time Fourier transform. The Segal-Bargmann space provides a useful example of a reproducing kernel Hilbert space. Coupled supersymmetries (coupled SUSYs) are generalizations of the quantum harmonic oscillator that have a built-in supersymmetric nature and enjoy similar properties to the quantum harmonic oscillator. In this paper, we will develop Segal-Bargmann transforms for a specific class of coupled SUSYs which includes the quantum harmonic oscillator as a special case. We will show that the associated Segal-Bargmann spaces are distinct from the usual Segal-Bargmann space: their associated weight functions are no longer Gaussian and are spanned by stricter subsets of the holomorphic polynomials. The coupled SUSY Segal-Bargmann spaces provide new examples of reproducing kernel Hilbert spaces.

1. Introduction

The harmonic oscillator is one of the most fundamental systems in quantum mechanics and quantum field theory due in part to its Lie algebraic structure and the fact that many potentials near their local minima may be well-approximated by a quadratic potential. The Lie algebra associated to the quantum harmonic oscillator is comprised of the Hamiltonian and creation and annihilation operators which have the interpretations of creating or destroying quanta of the quantum field. For our purposes, we define the quantum harmonic oscillator Hamiltonian defined on sufficiently nice functions in L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) to be the operator HOf(x)=12d2dx2f(x)+12x2f(x)\mathcal{H}_{\text{HO}}f(x)=-\frac{1}{2}\frac{\mathrm{d}^{2}}{\mathrm{d}x^{2}}f(x)+\frac{1}{2}x^{2}f(x). The quantum harmonic oscillator Hamiltonian can be formally factored as HO=12(ddx+x)(ddx+x)f(x)+12f(x)\mathcal{H}_{\text{HO}}=\frac{1}{2}\big{(}-\frac{\mathrm{d}}{\mathrm{d}x}+x\big{)}\big{(}\frac{\mathrm{d}}{\mathrm{d}x}+x\big{)}f(x)+\frac{1}{2}f(x). A standard exercise in quantum mechanics shows that the Hermite-Gauss functions, hlh_{l}, given by hl(x)=(1)l2ll!πex22dldxlex2h_{l}(x)=\frac{(-1)^{l}}{\sqrt{2^{l}l!\sqrt{\pi}}}e^{\frac{x^{2}}{2}}\frac{\mathrm{d}^{l}}{\mathrm{d}x^{l}}e^{-x^{2}}, are the orthonormal L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) eigenfunctions of the quantum harmonic oscillator Hamiltonian.

For fL2(,dx)AC(dx)f\in L^{2}(\mathbb{R},\mathrm{d}x)\cap AC(\mathbb{R}\,\mathrm{d}x) such that xfL2(,dx)xf\in L^{2}(\mathbb{R},\mathrm{d}x), and fL2(,dx)f^{\prime}\in L^{2}(\mathbb{R},\mathrm{d}x), define the operator aa by af(x)=12(f(x)+xf(x))af(x)=\frac{1}{\sqrt{2}}\big{(}f^{\prime}(x)+xf(x)\big{)}. Restricted to a similar set of functions, the adjoint aa^{*} has the form af(x)=12(f(x)+xf(x))a^{*}f(x)=\frac{1}{\sqrt{2}}\big{(}-f^{\prime}(x)+xf(x)\big{)} so that HO=aa+12\mathcal{H}_{\text{HO}}=a^{*}a+\frac{1}{2} on sufficiently nice functions. The operators aa, aa^{*}, and aaa^{*}a satisfy the following commutation relations

(1.1) [a,a]=1,[aa,a]=a,[aa,a]=a.[a,a^{*}]=1,\qquad[a^{*}a,a]=-a,\qquad[a^{*}a,a^{*}]=a^{*}.

A simple inspection shows that the span{1,a,a,aa}\operatorname{span}\{1,a,a^{*},a^{*}a\} is closed under commutators and thus forms the basis of a Lie algebra. This Lie algebra is sometimes referred to as the oscillator algebra [32]. It has as a Lie subalgebra the Heisenberg-Weyl algebra which is generated by the commutation relation [a,a]=1[a,a^{*}]=1—also called the canonical commutation relation—in (1.1).

The quantum harmonic oscillator Hamiltonian can be abstractly viewed as a Hamiltonian of the form aaa^{*}a where aa, aa^{*}, and aaa^{*}a satisfy the above commutation relations. Therefore one may view HO\mathcal{H}_{\text{HO}} and its raising and lowering operators as a concrete representation of the abstract quantum harmonic oscillator given by (1.1) on L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x). Another representation is in terms of number states: given a separable Hilbert space \mathfrak{H} and an orthonormal basis (en)n=0(e_{n})_{n=0}^{\infty}, define aen=nenae_{n}=\sqrt{n}e_{n} (and extend aa linearly), then aen=n+1ena^{*}e_{n}=\sqrt{n+1}e_{n} and aaen=nena^{*}ae_{n}=ne_{n}. Simple calculations show that these operators also satisfy the commutation relations in (1.1) and thus form another representation of the quantum harmonic oscillator. The number states representation also makes it easy to discuss the proper domains of definition of aa and aa^{*}.

Yet another representation exists on the Segal-Bargmann space [6, 30, 7], denoted 𝔉\mathfrak{F}—a Hilbert space of entire functions on \mathbb{C} with measure dρ(z,z¯)=1πezz¯dA(z)\mathrm{d}\rho(z,\bar{z})=\frac{1}{\pi}e^{-z\bar{z}}\,\mathrm{d}A(z) where dA(z)\mathrm{d}A(z) is the usual Lebesgue measure on \mathbb{C} given by dxdy\mathrm{d}x\wedge\mathrm{d}y or 14idzdz¯-\frac{1}{4i}\mathrm{d}z\wedge\mathrm{d}\bar{z}. The functions (en)n=0(e_{n})_{n=0}^{\infty} given by en(z)=1n!zne_{n}(z)=\frac{1}{\sqrt{n!}}z^{n} form an orthogonal basis for the Segal-Bargmann space. On the Segal-Bargmann space, the complex derivative operator 𝒶𝒻(𝓏)=𝓏𝒻(𝓏)\mathpzc{a}f(z)=\frac{\partial}{\partial z}f(z) has adjoint 𝒶𝒻(𝓏)=𝓏𝒻(𝓏)\mathpzc{a}^{*}f(z)=zf(z). It is easy to see that the operators 𝒶\mathpzc{a}, 𝒶\mathpzc{a}^{*}, and 𝒶𝒶\mathpzc{a}^{*}\mathpzc{a} on the Segal-Bargmann space also satisfy the conditions of the oscillator algebra in (1.1), allowing for the application of complex analytic methods to the study of the quantum harmonic oscillator.

The Segal-Bargmann transform is a Lie algebra and Hilbert space isomorphism between the L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) representation of the oscillator algebra and the 𝔉\mathfrak{F} representation of the oscillator algebra. The Segal-Bargmann transform, denoted \mathcal{B}, is a unitary map from L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) to 𝔉\mathfrak{F} given by

(1.2) f(z)=1π14ez22e2zxex22f(x)dx\mathcal{B}f(z)=\frac{1}{\pi^{\frac{1}{4}}}\int_{\mathbb{R}}e^{-\frac{z^{2}}{2}}e^{\sqrt{2}zx}e^{-\frac{x^{2}}{2}}f(x)\,\mathrm{d}x

with inverse given by

(1.3) 1f(x)=1π14ez¯22+2z¯xx22f(z)dρ(z,z¯)\mathcal{B}^{-1}f(x)=\frac{1}{\pi^{\frac{1}{4}}}\int_{\mathbb{C}}e^{-\frac{\bar{z}^{2}}{2}+\sqrt{2}\bar{z}x-\frac{x^{2}}{2}}f(z)\,\mathrm{d}\rho(z,\bar{z})

The Hilbert space and Lie algebra isomorphism properties that define the Segal-Bargmann transform can be summarized (for sufficiently nice fL2(,dx)f\in L^{2}(\mathbb{R},\mathrm{d}x)) by:

(1.4) 1π|f(z)|2ezz¯dA(z)\displaystyle\frac{1}{\pi}\int_{\mathbb{C}}|\mathcal{B}f(z)|^{2}e^{-z\bar{z}}\,\mathrm{d}A(z) =|f(x)|2dx\displaystyle=\int_{-\infty}^{\infty}|f(x)|^{2}\,\mathrm{d}x
(1.5) (af)(z)\displaystyle\mathcal{B}(af)(z) =𝒶𝒻(𝓏)\displaystyle=\mathpzc{a}\mathcal{B}f(z)
(1.6) (af)(z)\displaystyle\mathcal{B}(a^{*}f)(z) =𝒶𝒻(𝓏)\displaystyle=\mathpzc{a}^{*}\mathcal{B}f(z)
(1.7) (aaf)(z)\displaystyle\mathcal{B}(a^{*}af)(z) =𝒶𝒶𝒻(𝓏)\displaystyle=\mathpzc{a}^{*}\mathpzc{a}\mathcal{B}f(z)

The isomorphisms can be summarized via the following commutative diagrams.

L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x)𝔉\mathfrak{F}L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x)𝔉\mathfrak{F}L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x)𝔉\mathfrak{F}L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x)𝔉\mathfrak{F}\mathcal{B}aa𝒶\mathpzc{a}\mathcal{B}\mathcal{B}aa^{*}𝒶\mathpzc{a}^{*}\mathcal{B}
Figure 1. Commutative diagrams which define the usual Segal-Bargmann transform \mathcal{B}.

On the L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) representation of the oscillator algebra, a=12(ddx+x)a=\frac{1}{\sqrt{2}}\big{(}\frac{\mathrm{d}}{\mathrm{d}x}+x\big{)} is a lowering operator and similarly 𝒶=dd𝓏\mathpzc{a}=\frac{\mathrm{d}}{\mathrm{d}z} is a lowering operator in the Segal-Bargmann representation. Likewise, a=12(ddx+x)a^{*}=\frac{1}{\sqrt{2}}\big{(}-\frac{\mathrm{d}}{\mathrm{d}x}+x\big{)} is a raising operator on the L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) representation of the oscillator algebra and 𝒶=𝓏\mathpzc{a}^{*}=z is a raising operator in the Segal-Bargmann representation. A straightforward induction shows that the Hermite-Gauss functions hlh_{l} are mapped to 1l!zl\frac{1}{\sqrt{l!}}z^{l} under the Segal-Bargmann transform.

In [35], the notion of a coupled supersymmetry (coupled SUSY) was introduced as a generalization of the quantum harmonic oscillator with a built-in supersymmetric nature, informed by supersymmetric quantum mechanics (SUSY QM) [12, 13]. Whereas SUSY QM only has a Lie superalgebra structure that is not strong enough to fully identify the eigenstructure of the associated quantum systems, coupled SUSYs have associated to them additional 𝔰𝔲(1,1)\mathfrak{su}(1,1) Lie algebra structures beyond the Lie superalgebra structures which allow for the exact solvability of the eigenstructure.

Recently, there has been a wide range of Segal-Bargmann-like analysis beyond n\mathbb{R}^{n} to Lie groups via heat kernel and geometric quantization techniques [20, 21, 31, 23, 22, 16, 10, 24]. There have also been Segal-Bargmann transform applications to the study of superalgebras and supersymmetric quantum systems [33, 5, 4], as well as analogues developed in the quaternionic and Clifford algebra settings [2, 11, 14, 15, 17, 26], and qq-deformations and related generalizations [25, 29, 3].

In this paper, we will present new generalizations of the usual Segal-Bargmann space and the Segal-Bargmann transform that have a supersymmetric flavor that arise from a specific class of coupled SUSYs, leveraging the 𝔰𝔲(1,1)\mathfrak{su}(1,1) structure that come with coupled SUSYs.

The remainder of the paper is organized as follows. In Section 2, we will briefly discuss the notion of a coupled supersymmetry, particularly focusing on a family of coupled supersymmetries on \mathbb{R} first introduced in [35] of which the quantum harmonic oscillator is a special case. In Section 3, we will develop generalizations of the Segal-Bargmann space on which holomorphic representations of the coupled SUSYs live and remark on the reproducing kernel Hilbert space nature thereof. In Section 4, we will establish generalizations of the Segal-Bargmann transform that act as Lie algebra and Hilbert space isomorphisms between the two representations and discuss the relationships between the generalized Segal-Bargmann spaces. In Section 5, we will discuss the relationship of the coupled SUSY Segal-Bargmann transforms to short-time transforms and coherent states, particularly the work developed in [8] on SU(1,1)SU(1,1) coherent states.

2. A Class of Coupled Supersymmetries

2.1. Coupled Supersymmetry

For the sake of self-containment, we will restate some of the basics of coupled supersymmetry, including basic results without proof.

Definition 2.1.

Let 1\mathfrak{H}_{1} and 2\mathfrak{H}_{2} be Hilbert spaces, a:12a:\mathfrak{H}_{1}\to\mathfrak{H}_{2}, and b:21b:\mathfrak{H}_{2}\to\mathfrak{H}_{1} be closed, densely-defined operators such that the domains are such that the proceeding identities are well-defined. We say that the ordered quadruplet {a,b,γ,δ}\{a,b,\gamma,\delta\} where γ,δ\gamma,\delta\in\mathbb{R} form a coupled supersymmetry (coupled SUSY) if they satisfy the following relations

(2.1) aa\displaystyle a^{*}a =bb+γ11\displaystyle=bb^{*}+\gamma 1_{\mathfrak{H}_{1}}
(2.2) aa\displaystyle aa^{*} =bb+δ12.\displaystyle=b^{*}b+\delta 1_{\mathfrak{H}_{2}}.

In what follows, we suppress the notation 11_{\mathfrak{H}} for brevity as it is implied. The operators aaa^{*}a, aaaa^{*}, bbb^{*}b, and bbbb^{*} are called Hamiltonians. A coupled SUSY is unbroken if ker(a)\ker(a) and ker(b)\ker(b) are both nontrivial, partially unbroken if exactly one of ker(a)\ker(a) and ker(b)\ker(b) is nontrivial, and broken if ker(a)\ker(a) and ker(b)\ker(b) are both trivial.

Remark 1.

Note that taking b=ab=a and γ=1=δ-\gamma=1=\delta in a coupled SUSY gives the Heisenberg-Weyl commutation relations so that the abstract quantum harmonic oscillator is indeed a special case of a coupled SUSY. However, coupled SUSYs extend beyond the normal harmonic oscillator as will be seen in Definition 2.2.

In many cases, the coupled SUSY operators (aa, bb, aa^{*}, bb^{*}) map a Hilbert space back to itself, but distinguishing the domain and codomain and allowing for the two to be the same makes the diagram chasing in what follows simpler; moreover, the nature of the operators aa and bb may necessitate defining them on separate densely-defined subspaces of L2L^{2} spaces, so distinguishing their domains serves a further purpose. In traditional SUSY terms, 1\mathfrak{H}_{1} represents the Hilbert space for the first (bosonic) sector, whereas 2\mathfrak{H}_{2} represents the Hilbert space for the second (fermionic) sector [12].

Coupled SUSYs have an 𝔰𝔲(1,1)\mathfrak{su}(1,1) Lie algebra structure generated by the raising and lowering operators aba^{*}b^{*} (or bab^{*}a^{*}) and baba (or abab) and the coupled SUSY Hamiltonians aaa^{*}a and bbbb^{*} (or aaaa^{*} and bbb^{*}b):

(2.3) [aa,ab]\displaystyle[a^{*}a,a^{*}b^{*}] =(δγ)ab\displaystyle=(\delta-\gamma)a^{*}b^{*}
(2.4) [aa,ba]\displaystyle[a^{*}a,ba] =(δγ)ba\displaystyle=-(\delta-\gamma)ba
(2.5) [ab,ba]\displaystyle[a^{*}b^{*},ba] =2(δγ)(aaγ2)\displaystyle=-2(\delta-\gamma)\bigg{(}a^{*}a-\frac{\gamma}{2}\bigg{)}

The Lie structure determines the eigenstructure of coupled SUSYs. In an unbroken coupled SUSY the eigenvalue structure is uniquely determined as ker(a)\ker(a) and ker(b)\ker(b) being nontrivial force zero points in the spectra of aaa^{*}a and bbb^{*}b. In (partially) broken coupled SUSYs, the the spectra will exhibit an overall shift. In an unbroken coupled SUSY, the eigenvalues of aaa^{*}a are given by m(δγ)m(\delta-\gamma) and m(δγ)+δm(\delta-\gamma)+\delta where m0m\in\mathbb{N}_{0}.

Remark 2.

This ladder structure resembles that of the usual quantum harmonic oscillator with the squares of the creation and annihilation operators which also enjoy an 𝔰𝔲(1,1)\mathfrak{su}(1,1) Lie algebra structure:

(2.6) [aa+12,(a)2]=2(a)2,[aa+12,a2]=2a2,[(a)2,a2]=4(aa+12).\bigg{[}a^{*}a+\frac{1}{2},\big{(}a^{*}\big{)}^{2}\bigg{]}=2\big{(}a^{*}\big{)}^{2},\qquad\bigg{[}a^{*}a+\frac{1}{2},a^{2}\bigg{]}=-2a^{2},\qquad\bigg{[}\big{(}a^{*}\big{)}^{2},a^{2}\bigg{]}=-4\bigg{(}a^{*}a+\frac{1}{2}\bigg{)}.

This does not, however, encapsulate all coupled SUSYs as seen in Definition 2.2. Another connection exists between coupled SUSYs and the harmonic oscillator by way of the “spinification” of a two-particle quantum harmonic oscillator system as discussed in [35].

2.2. Concrete Realizations of Coupled SUSYs

In this paper, we will be concerned with a family of coupled SUSYs indexed by nn\in\mathbb{N} where γ=1\gamma=-1 and δ=2n1\delta=2n-1, i.e.

(2.7) aa\displaystyle a^{*}a =bb1\displaystyle=bb^{*}-1
(2.8) aa\displaystyle aa^{*} =bb+2n1.\displaystyle=b^{*}b+2n-1.

We will suppress the explicit dependence on nn of the operators aa, aa^{*}, bb, and bb^{*} throughout for ease of notation.

Definition 2.2.

Define the operators aa and bb on L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) by

(2.9) a=12(1xn1ddx+xn),b=12(ddx1xn1+xn)a=\frac{1}{\sqrt{2}}\bigg{(}\frac{1}{x^{n-1}}\frac{\mathrm{d}}{\mathrm{d}x}+x^{n}\bigg{)},\qquad b=\frac{1}{\sqrt{2}}\bigg{(}\frac{\mathrm{d}}{\mathrm{d}x}\frac{1}{x^{n-1}}+x^{n}\bigg{)}

These operators were introduced in [35]. aa can be formally defined on those fL2(,dx)f\in L^{2}(\mathbb{R},\mathrm{d}x) such that xnfL2(,dx)x^{n}f\in L^{2}(\mathbb{R},\mathrm{d}x) and 1xn1fL2(,dx)\frac{1}{x^{n-1}}f^{\prime}\in L^{2}(\mathbb{R},\mathrm{d}x), similar for bb. Examples of such functions include the compactly supported smooth functions supported away from zero. A straightforward computation shows that on sufficiently nice functions, e.g. the compactly supported smooth functions supported away from zero, the adjoints aa^{*} and bb^{*} can be represented by

(2.10) a=12(ddx1xn1+xn),b=12(1xn1ddx+xn).a^{*}=\frac{1}{\sqrt{2}}\bigg{(}-\frac{\mathrm{d}}{\mathrm{d}x}\frac{1}{x^{n-1}}+x^{n}\bigg{)},\qquad b^{*}=\frac{1}{\sqrt{2}}\bigg{(}-\frac{1}{x^{n-1}}\frac{\mathrm{d}}{\mathrm{d}x}+x^{n}\bigg{)}.

Note the functional similarity between aa^{*} and bb and likewise aa and bb^{*}. This plays a crucial role in the analysis that follows.

Remark 3.

Taking n=1n=1 in the above reveals the quantum harmonic oscillator ladder operators so that the n=1n=1 case in what follows will reduce to the usual Segal-Bargmann transform and Segal-Bargmann space. We will remark on this special case throughout as we build up the coupled SUSY Segal-Bargmann spaces and transforms.

In a slight abuse of notation in light of the definition of a coupled SUSY, we will reserve 1\mathfrak{H}_{1} and 2\mathfrak{H}_{2} henceforth to be L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) on which the above representation lives. 1\mathfrak{H}_{1} represents the Hilbert space corresponding to the eigenfunctions of aaa^{*}a (or equivalently, bbbb^{*}), whereas 2\mathfrak{H}_{2} represents the Hilbert space corresponding to the eigenfunctions of aaaa^{*} (or equivalently, bbb^{*}b).

Definition 2.3.

Define the normalized functions ψ01\psi_{0}\in\mathfrak{H}_{1} and ψ~02\widetilde{\psi}_{0}\in\mathfrak{H}_{2} by

(2.11) ψ0(x)=n1214nΓ(12n)ex2n2n,ψ~0(x)=n14nΓ(112n)xn1ex2n2n\psi_{0}(x)=\frac{n^{\frac{1}{2}-\frac{1}{4n}}}{\sqrt{\Gamma\big{(}\frac{1}{2n}\big{)}}}e^{-\frac{x^{2n}}{2n}},\qquad\widetilde{\psi}_{0}(x)=\frac{n^{\frac{1}{4n}}}{\sqrt{\Gamma\big{(}1-\frac{1}{2n}\big{)}}}x^{n-1}e^{-\frac{x^{2n}}{2n}}

Further, define the higher (normalized) eigenfunctions ψl\psi_{l} of aaa^{*}a and ψ~l\widetilde{\psi}_{l} of aaaa^{*} by

(2.12) ψ2l\displaystyle\psi_{2l} =(ab)lψ0(ab)lψ0,\displaystyle=\frac{(a^{*}b^{*})^{l}\psi_{0}}{\|(a^{*}b^{*})^{l}\psi_{0}\|},
(2.13) ψ2l+1\displaystyle\psi_{2l+1} =(ab)laψ~0(ab)laψ~0,\displaystyle=\frac{(a^{*}b^{*})^{l}a^{*}\widetilde{\psi}_{0}}{\|(a^{*}b^{*})^{l}a^{*}\widetilde{\psi}_{0}\|},
(2.14) ψ~2l\displaystyle\widetilde{\psi}_{2l} =(ba)lψ~0(ba)lψ~0,\displaystyle=\frac{(b^{*}a^{*})^{l}\widetilde{\psi}_{0}}{\|(b^{*}a^{*})^{l}\widetilde{\psi}_{0}\|},
(2.15) ψ~2l+1\displaystyle\widetilde{\psi}_{2l+1} =(ba)lbψ0(ba)lbψ0.\displaystyle=\frac{(b^{*}a^{*})^{l}b^{*}\psi_{0}}{\|(b^{*}a^{*})^{l}b^{*}\psi_{0}\|}.

A simple computation shows that

(2.16) aψ0=0,bψ~0=0,a\psi_{0}=0,\qquad b\widetilde{\psi}_{0}=0,

so that aaψ0=0a^{*}a\psi_{0}=0 and bbψ~0=0b^{*}b\widetilde{\psi}_{0}=0 so that these are indeed eigenfunctions and similarly, ψl\psi_{l} and ψ~l\widetilde{\psi}_{l} are eigenfunctions of their corresponding Hamiltonians, seen by way of the 𝔰𝔲(1,1)\mathfrak{su}(1,1) Lie algebra structure.

Proposition 2.4.

The eigenfunctions of aaa^{*}a generated in this way have Rodrigues formulae resembling that for the Hermite-Gauss functions given by

(2.17) (ab)lex2n2n\displaystyle(a^{*}b^{*})^{l}e^{-\frac{x^{2n}}{2n}} =12lex2n2n(ddx1x2n2ddx)lex2nn\displaystyle=\frac{1}{2^{l}}e^{\frac{x^{2n}}{2n}}\bigg{(}\frac{\mathrm{d}}{\mathrm{d}x}\frac{1}{x^{2n-2}}\frac{\mathrm{d}}{\mathrm{d}x}\bigg{)}^{l}e^{-\frac{x^{2n}}{n}}
(2.18) (ab)l(2x2n1ex2n2n)\displaystyle(a^{*}b^{*})^{l}\bigg{(}2x^{2n-1}e^{-\frac{x^{2n}}{2n}}\bigg{)} =12lex2n2n(ddx1x2n2ddx)l(2x2n1ex2nn).\displaystyle=\frac{1}{2^{l}}e^{\frac{x^{2n}}{2n}}\bigg{(}\frac{\mathrm{d}}{\mathrm{d}x}\frac{1}{x^{2n-2}}\frac{\mathrm{d}}{\mathrm{d}x}\bigg{)}^{l}\bigg{(}2x^{2n-1}e^{-\frac{x^{2n}}{n}}\bigg{)}.

Similar formulae exist for the eigenfunctions of aaaa^{*}. The proof is a straightforward induction.

Remark 4.

When n=1n=1, ψ0(x)=π14ex22=ψ~0(x)\psi_{0}(x)=\pi^{-\frac{1}{4}}e^{-\frac{x^{2}}{2}}=\widetilde{\psi}_{0}(x), giving the usual quantum harmonic oscillator ground state, and successive application of the raising operators gives the quantum harmonic oscillator excited states, i.e. the Hermite-Gauss functions.

Proposition 2.5.

The eigenfunctions ψl\psi_{l} and ψ~l\widetilde{\psi}_{l} form orthonormal bases for 1\mathfrak{H}_{1} and 2\mathfrak{H}_{2}, respectively.

A proof of related facts can be found in [34] and is based on a proof by Akhiezer for the Fourier-Bessel (Hankel) transform [1]. Therein, functions directly related to the functions ψl\psi_{l} were shown to give an orthonormal basis for L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x). The proof can be adapted readily to ψl\psi_{l} and ψ~l\widetilde{\psi}_{l} to show completeness.

The coupled SUSY ladder structure is summarized in the diagrams below. Note that the notation differs slightly from [35] for ease of discussion.

1\mathfrak{H}_{1}2\mathfrak{H}_{2}ψ0\psi_{0}ψ1\psi_{1}ψ2\psi_{2}ψ3\psi_{3}ψ4\psi_{4}ψ5\psi_{5}ψ6\psi_{6}ψ~0\widetilde{\psi}_{0}ψ~1\widetilde{\psi}_{1}ψ~2\widetilde{\psi}_{2}ψ~3\widetilde{\psi}_{3}ψ~4\widetilde{\psi}_{4}ψ~5\widetilde{\psi}_{5}aaaa^{*}bb^{*}bb
Figure 2. The actions of aa, bb, aa^{*}, and bb^{*} in a coupled SUSY.
1\mathfrak{H}_{1}2\mathfrak{H}_{2}ψ0\psi_{0}ψ1\psi_{1}ψ2\psi_{2}ψ~0\widetilde{\psi}_{0}ψ~1\widetilde{\psi}_{1}aa^{*}bb^{*}aba^{*}b^{*}
Figure 3. The raising operator structure for the first sector in a coupled SUSY.
1\mathfrak{H}_{1}2\mathfrak{H}_{2}ψ0\psi_{0}ψ1\psi_{1}ψ2\psi_{2}ψ~0\widetilde{\psi}_{0}ψ~1\widetilde{\psi}_{1}aabbbaba
Figure 4. The lowering operator structure for the first sector in a coupled SUSY.

3. Constructing the Coupled SUSY Segal-Bargmann Spaces

The following relations hold for the complex variable zz\in\mathbb{C} and holomorphic derivative ddz\frac{\mathrm{d}}{\mathrm{d}z} as operators:

(3.1) (zn)(1zn1ddz)\displaystyle\big{(}z^{n}\big{)}\bigg{(}\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z}\bigg{)} =(ddz1zn1)(zn)1,\displaystyle=\bigg{(}\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}}\bigg{)}\big{(}z^{n}\big{)}-1,
(3.2) (1zn1ddz)(zn)\displaystyle\bigg{(}\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z}\bigg{)}\big{(}z^{n}\big{)} =(zn)(ddz1zn1)+2n1,\displaystyle=\big{(}z^{n}\big{)}\bigg{(}\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}}\bigg{)}+2n-1,

mimicking the coupled SUSY relations in (2.1) and (2.2).

In order for these operators and their corresponding products to be well-defined on entire functions, we must restrict the domains for 1zn1ddz\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z} and ddz1zn1\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}}. Define dom(1zn1ddz)\operatorname{dom}\big{(}\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z}\big{)} to be those entire functions spanned by {1,z2n1,z2n,z4n1,z4n,}\{1,z^{2n-1},z^{2n},z^{4n-1},z^{4n},\ldots\} and dom(ddz1zn1)\operatorname{dom}\big{(}\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}}\big{)} to be those entire functions spanned by {zn1,zn,z3n1,z3n,}\{z^{n-1},z^{n},z^{3n-1},z^{3n},\ldots\}. It is easy to see that these are the maximal subspaces for which the coupled SUSY domain and range conditions hold.

Definition 3.1.

To this end, we define the following not-yet-closed spaces of entire functions:

(3.3) 𝒪1()\displaystyle\mathcal{O}_{1}(\mathbb{C}) =span{1,z2n1,z2n,z4n1,z4n,}\displaystyle=\operatorname{span}\{1,z^{2n-1},z^{2n},z^{4n-1},z^{4n},\ldots\}
(3.4) 𝒪2()\displaystyle\mathcal{O}_{2}(\mathbb{C}) =span{zn1,zn,z3n1,z3n,}\displaystyle=\operatorname{span}\{z^{n-1},z^{n},z^{3n-1},z^{3n},\ldots\}
Remark 5.

Such choices for these spaces further distinguish this from the square operator representation given in (2.6) as these would lead to the usual Segal-Bargmann space—or perhaps the even and odd subspaces thereof, depending on interpretation.

Definition 3.2.

Let ρ1\rho_{1} and ρ2\rho_{2} be not-yet-defined weights on 𝒪1()\mathcal{O}_{1}(\mathbb{C}) and 𝒪2()\mathcal{O}_{2}(\mathbb{C}), respectively, defining inner products ,1\langle\cdot,\cdot\rangle_{1} and ,2\langle\cdot,\cdot\rangle_{2}, respectively, by

(3.5) f1,g11\displaystyle\langle f_{1},g_{1}\rangle_{1} =f1(z)g1(z)¯ρ1(z,z¯)dA(z)\displaystyle=\int_{\mathbb{C}}f_{1}(z)\overline{g_{1}(z)}\rho_{1}(z,\bar{z})\,\mathrm{d}A(z)
(3.6) f2,g22\displaystyle\langle f_{2},g_{2}\rangle_{2} =f2(z)g2(z)¯ρ2(z,z¯)dA(z).\displaystyle=\int_{\mathbb{C}}f_{2}(z)\overline{g_{2}(z)}\rho_{2}(z,\bar{z})\,\mathrm{d}A(z).

Let 1\|\cdot\|_{1} denote the norm resulting from ,1\langle\cdot,\cdot\rangle_{1} and Let 2\|\cdot\|_{2} denote the norm resulting from ,2\langle\cdot,\cdot\rangle_{2}.

Much like in [6], ρ1\rho_{1} and ρ2\rho_{2} will be defined so that the adjoints of 1zn1ddz\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z} and ddz1zn1\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}} are both znz^{n}, agreeing with the desired coupled SUSY structure.

If we wish to realize 1zn1ddz\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z} and ddz1zn1\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}} as a representation of the coupled SUSY given in (3.3) and (3.4), then the following relations must hold for f1,g1𝒪1()f_{1},g_{1}\in\mathcal{O}_{1}(\mathbb{C}) and f2,g2𝒪2()f_{2},g_{2}\in\mathcal{O}_{2}(\mathbb{C}):

(3.7) 1zn1ddzf1(z)g2(z)¯ρ2(z,z¯)dA(z)\displaystyle\int_{\mathbb{C}}\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z}f_{1}(z)\overline{g_{2}(z)}\rho_{2}(z,\bar{z})\,\mathrm{d}A(z) =f1(z)zng2(z)¯ρ1(z,z¯)dA(z)\displaystyle=\int_{\mathbb{C}}f_{1}(z)\overline{z^{n}g_{2}(z)}\rho_{1}(z,\bar{z})\,\mathrm{d}A(z)
(3.8) ddz1zn1f2(z)g1(z)¯ρ1(z,z¯)dA(z)\displaystyle\int_{\mathbb{C}}\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}}f_{2}(z)\overline{g_{1}(z)}\rho_{1}(z,\bar{z})\,\mathrm{d}A(z) =f2(z)zng1(z)¯ρ2(z,z¯)dA(z)\displaystyle=\int_{\mathbb{C}}f_{2}(z)\overline{z^{n}g_{1}(z)}\rho_{2}(z,\bar{z})\,\mathrm{d}A(z)

Integrating by parts in (3.7) and (3.8), making use of entirety of g1g_{1} and g2g_{2}, and assuming—for now—that the boundary terms go to zero gives

(3.9) f1(z)g2(z)¯z¯nρ1(z,z¯)dA(z)\displaystyle\int_{\mathbb{C}}f_{1}(z)\overline{g_{2}(z)}\overline{z}^{n}\rho_{1}(z,\bar{z})\,\mathrm{d}A(z) =f1(z)g2(z)¯(z1zn1ρ2(z,z¯))dA(z),\displaystyle=\int_{\mathbb{C}}f_{1}(z)\overline{g_{2}(z)}\bigg{(}-\frac{\partial}{\partial z}\frac{1}{z^{n-1}}\rho_{2}(z,\bar{z})\bigg{)}\,\mathrm{d}A(z),
(3.10) f2(z)g1(z)¯z¯nρ2(z,z¯)dA(z)\displaystyle\int_{\mathbb{C}}f_{2}(z)\overline{g_{1}(z)}\overline{z}^{n}\rho_{2}(z,\bar{z})\,\mathrm{d}A(z) =f2(z)g1(z)¯(1zn1zρ1(z,z¯))dA(z).\displaystyle=\int_{\mathbb{C}}f_{2}(z)\overline{g_{1}(z)}\bigg{(}-\frac{1}{z^{n-1}}\frac{\partial}{\partial z}\rho_{1}(z,\bar{z})\bigg{)}\,\mathrm{d}A(z).

It is easy to see that the following relationships for ρ1\rho_{1} and ρ2\rho_{2} guarantee that the above hold:

(3.11) z1zn1ρ2(z,z¯)\displaystyle-\dfrac{\partial}{\partial z}\dfrac{1}{z^{n-1}}\rho_{2}(z,\bar{z}) =z¯nρ1(z,z¯)\displaystyle=\bar{z}^{n}\rho_{1}(z,\bar{z})
(3.12) 1zn1zρ1(z,z¯)\displaystyle-\dfrac{1}{z^{n-1}}\dfrac{\partial}{\partial z}\rho_{1}(z,\bar{z}) =z¯nρ2(z,z¯).\displaystyle=\bar{z}^{n}\rho_{2}(z,\bar{z}).

Decoupling these equations leads to the following partial differential equations for ρ1\rho_{1} and ρ2\rho_{2}:

(3.13) z1z2n2zρ1(z,z¯)\displaystyle\frac{\partial}{\partial z}\frac{1}{z^{2n-2}}\frac{\partial}{\partial z}\rho_{1}(z,\bar{z}) =z¯2nρ1(z,z¯)\displaystyle=\bar{z}^{2n}\rho_{1}(z,\bar{z})
(3.14) 1zn12z21zn1ρ2(z,z¯)\displaystyle\frac{1}{z^{n-1}}\frac{\partial^{2}}{\partial z^{2}}\frac{1}{z^{n-1}}\rho_{2}(z,\bar{z}) =z¯2nρ2(z,z¯).\displaystyle=\bar{z}^{2n}\rho_{2}(z,\bar{z}).
Definition 3.3.

Define ρ1:2\rho_{1}:\mathbb{C}^{2}\to\mathbb{R} and ρ2:2\rho_{2}:\mathbb{C}^{2}\to\mathbb{R} to be

(3.15) ρ1(z,z¯)\displaystyle\rho_{1}(z,\overline{z}) =2(2n)12nπΓ(12n)(zz¯)n12K112n((zz¯)nn)\displaystyle=\frac{2}{(2n)^{\frac{1}{2n}}\pi\Gamma\big{(}\frac{1}{2n}\big{)}}(z\bar{z})^{n-\frac{1}{2}}K_{1-\frac{1}{2n}}\bigg{(}\frac{(z\bar{z})^{n}}{n}\bigg{)}
(3.16) ρ2(z,z¯)\displaystyle\rho_{2}(z,\overline{z}) =2(2n)12nπΓ(12n)(zz¯)n12K12n((zz¯)nn),\displaystyle=\frac{2}{(2n)^{\frac{1}{2n}}\pi\Gamma\big{(}\frac{1}{2n}\big{)}}(z\bar{z})^{n-\frac{1}{2}}K_{\frac{1}{2n}}\bigg{(}\frac{(z\bar{z})^{n}}{n}\bigg{)},

where KνK_{\nu} is the modified Bessel function of the second kind [27, Eq. 10.27.4] given by

(3.17) Kν(z)=π2sin(νπ)(Iν(z)Iν(z)),K_{\nu}(z)=\frac{\pi}{2\sin(\nu\pi)}(I_{-\nu}(z)-I_{\nu}(z)),

and IνI_{\nu} is the modified Bessel function of the first kind [27, Eq. 10.25.2] given by

(3.18) Iν(z)=(z2)νl=01Γ(l+ν+1)l!(z2)2l.I_{\nu}(z)=\bigg{(}\frac{z}{2}\bigg{)}^{\nu}\sum_{l=0}^{\infty}\frac{1}{\Gamma(l+\nu+1)l!}\bigg{(}\frac{z}{2}\bigg{)}^{2l}.

From these representations in terms of the modified Bessel function KνK_{\nu}, we can see that ρ1\rho_{1} and ρ2\rho_{2} solve the above partial differential equations. Note that both ρ1\rho_{1} and ρ2\rho_{2} share a normalization factor of 2(2n)12nπΓ(12n)\frac{2}{(2n)^{\frac{1}{2n}}\pi\Gamma\big{(}\frac{1}{2n}\big{)}}. This is due to the coupled nature of the two weight functions, i.e. they cannot be separately normalized to 11.

In order for the inner products ,1\langle\cdot,\cdot\rangle_{1} and ,2\langle\cdot,\cdot\rangle_{2} to be positive, ρ1\rho_{1} and ρ2\rho_{2} should be positive functions. The modified Bessel function KνK_{\nu} has the following integral representation [27, Eq. 10.32.9]:

(3.19) Kν(x)=0excosh(t)cosh(νt)dt,|ph(x)|<π2.K_{\nu}(x)=\int_{0}^{\infty}e^{-x\cosh(t)}\cosh(\nu t)\,\mathrm{d}t,\qquad\qquad|\operatorname{ph}(x)|<\frac{\pi}{2}.

For positive xx, the integrand is real and positive. Taking x=zz¯x=z\bar{z}, where zz\in\mathbb{C}, we see that Kν((zz¯)nn)K_{\nu}\big{(}\frac{(z\bar{z})^{n}}{n}\big{)} is positive and thus so are ρ1\rho_{1} and ρ2\rho_{2}.

Remark 6.

When n=1n=1, ρ1\rho_{1} and ρ2\rho_{2} are identical and reduce to a single weight ρ\rho which is given by

(3.20) ρ(z,z¯)=1π2π|z|K12(|z|2).\rho(z,\bar{z})=\frac{1}{\pi}\sqrt{\frac{2}{\pi}}|z|K_{\frac{1}{2}}\big{(}|z|^{2}\big{)}.

Noting that for x>0x>0, K12(x)=π2xexK_{\frac{1}{2}}(x)=\sqrt{\frac{\pi}{2x}}e^{-x}, [27, Eq. 10.39.2] we obtain

(3.21) ρ(z,z¯)=1πezz¯,\rho(z,\bar{z})=\frac{1}{\pi}e^{-z\bar{z}},

matching the usual Segal-Bargmann space weight function as expected.

The modified Bessel function KνK_{\nu} has a nice asymptotic expansion as xx\to\infty [27, Eq. 10.40.2]:

(3.22) Kν(x)π2xex(1+4ν218x+O(x2)).K_{\nu}(x)\sim\sqrt{\frac{\pi}{2x}}e^{-x}\bigg{(}1+\frac{4\nu^{2}-1}{8x}+O(x^{-2})\bigg{)}.

Thus, asymptotically, ρ1\rho_{1} and ρ2\rho_{2} have the following asymptotic expansions:

(3.23) ρ1(z,z¯)\displaystyle\rho_{1}(z,\overline{z}) 1(2n)12nΓ(12n)2π|z|n1e|z|2nn(1+α|z|2n+O(z4n))\displaystyle\sim\frac{1}{(2n)^{\frac{1}{2n}}\Gamma\big{(}\frac{1}{2n}\big{)}}\sqrt{\frac{2}{\pi}}|z|^{n-1}e^{-\frac{|z|^{2n}}{n}}\bigg{(}1+\alpha|z|^{-2n}+O(z^{-4n})\bigg{)}
(3.24) ρ2(z,z¯)\displaystyle\rho_{2}(z,\overline{z}) 1(2n)12nΓ(12n)2π|z|n1e|z|2nn(1+α|z|2n+O(z4n)).\displaystyle\sim\frac{1}{(2n)^{\frac{1}{2n}}\Gamma\big{(}\frac{1}{2n}\big{)}}\sqrt{\frac{2}{\pi}}|z|^{n-1}e^{-\frac{|z|^{2n}}{n}}\bigg{(}1+\alpha|z|^{-2n}+O(z^{-4n})\bigg{)}.

Therefore the weights ρ1\rho_{1} and ρ2\rho_{2} are exponentially decaying, much as in the case of the traditional Segal-Bargmann space, which gives that the monomials znz^{n} have finite norm. Indeed for n=1n=1, the asymptotic expansion is exact and is a Gaussian in |z||z|. This informs the assumption that the boundary terms in the integration by parts in (3.7) and (3.8) vanish.

Much as in the case of the Segal-Bargmann space, the functions z2nlz^{2nl} and z2nl+2n1z^{2nl+2n-1} in 𝒪1()\mathcal{O}_{1}(\mathbb{C}) and z2nl+n1z^{2nl+n-1} and z2nl+nz^{2nl+n} in 𝒪2()\mathcal{O}_{2}(\mathbb{C}) are not normalized.

Definition 3.4.

Define the normalized functions ele_{l} and e~l\widetilde{e}_{l} by

(3.25) el(z)\displaystyle e_{l}(z) ={Γ(12n)(2n)2kΓ(k+12n)k!z2nk,l=2kΓ(12n)(2n)2k+21nΓ(k+212n)k!z2nk+2n1,l=2k+1\displaystyle=\begin{cases}\sqrt{\frac{\Gamma\big{(}\frac{1}{2n}\big{)}}{(2n)^{2k}\Gamma\big{(}k+\frac{1}{2n}\big{)}k!}}z^{2nk},&l=2k\\ \sqrt{\frac{\Gamma\big{(}\frac{1}{2n}\big{)}}{(2n)^{2k+2-\frac{1}{n}}\Gamma\big{(}k+2-\frac{1}{2n}\big{)}k!}}z^{2nk+2n-1},&l=2k+1\end{cases}
(3.26) e~l(z)\displaystyle\widetilde{e}_{l}(z) ={Γ(12n)(2n)2k+11nΓ(k+112n)k!z2nk+n1,l=2kΓ(12n)(2n)2k+1Γ(k+1+12n)k!z2nk+n,l=2k+1\displaystyle=\begin{cases}\sqrt{\frac{\Gamma\big{(}\frac{1}{2n}\big{)}}{(2n)^{2k+1-\frac{1}{n}}\Gamma\big{(}k+1-\frac{1}{2n}\big{)}k!}}z^{2nk+n-1},&l=2k\\ \sqrt{\frac{\Gamma\big{(}\frac{1}{2n}\big{)}}{(2n)^{2k+1}\Gamma\big{(}k+1+\frac{1}{2n}\big{)}k!}}z^{2nk+n},&l=2k+1\end{cases}

Note that e01e_{0}\equiv 1, but e~01\widetilde{e}_{0}\not\equiv 1 in general, a sharp contrast from typical Segal-Bargmann spaces.

Definition 3.5.

Let 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} be the holomorphic function spaces corresponding to 𝒪1()\mathcal{O}_{1}(\mathbb{C}) and 𝒪2()\mathcal{O}_{2}(\mathbb{C}), respectively, i.e.

(3.27) 𝔉1\displaystyle\mathfrak{F}_{1} ={f1=lclel|l|cl|2<},\displaystyle=\bigg{\{}f_{1}=\sum_{l}c_{l}e_{l}\biggm{|}\sum_{l}|c_{l}|^{2}<\infty\bigg{\}},
(3.28) 𝔉2\displaystyle\mathfrak{F}_{2} ={f2=ldle~l|l|dl|2<}.\displaystyle=\bigg{\{}f_{2}=\sum_{l}d_{l}\widetilde{e}_{l}\biggm{|}\sum_{l}|d_{l}|^{2}<\infty\bigg{\}}.

𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} are Hilbert spaces as they are equivalent to an 2\ell^{2} space. Standard arguments [19] show that point evaluation is a bounded linear functional on 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} which allows norm convergence to pass to uniform convergence on compact sets. As uniform limits of holomorphic functions on compact sets are again holomorphic, we can conclude that 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} are indeed comprised of entire functions—not merely L2L^{2} limits of entire functions.

With the establishment of the holomorphic function spaces 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2}, we can define the holomorphic realization of the coupled SUSY operators.

Definition 3.6.

Define the operators 𝒶:𝔉1𝔉2\mathpzc{a}:\mathfrak{F}_{1}\to\mathfrak{F}_{2} and 𝒷:𝔉2𝔉1\mathpzc{b}:\mathfrak{F}_{2}\to\mathfrak{F}_{1} by

(3.29) 𝒶𝒻1(𝓏)\displaystyle\mathpzc{a}f_{1}(z) =1zn1ddzf1(z)\displaystyle=\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z}f_{1}(z)
(3.30) 𝒷𝒻2(𝓏)\displaystyle\mathpzc{b}f_{2}(z) =ddz1zn1f2(z).\displaystyle=\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}}f_{2}(z).

Since 𝒶:𝔉1𝔉2\mathpzc{a}:\mathfrak{F}_{1}\to\mathfrak{F}_{2}, 𝒶:𝔉2𝔉1\mathpzc{a}^{*}:\mathfrak{F}_{2}\to\mathfrak{F}_{1}, and similarly since 𝒷:𝔉2𝔉1\mathpzc{b}:\mathfrak{F}_{2}\to\mathfrak{F}_{1}, 𝒷:𝔉1𝔉2\mathpzc{b}^{*}:\mathfrak{F}_{1}\to\mathfrak{F}_{2}. From previous arguments by way of the constructions of ρ1\rho_{1} and ρ2\rho_{2}, we have that

(3.31) 𝒶𝒻2(𝓏)\displaystyle\mathpzc{a}^{*}f_{2}(z) =znf2(z)\displaystyle=z^{n}f_{2}(z)
(3.32) 𝒷𝒻1(𝓏)\displaystyle\mathpzc{b}^{*}f_{1}(z) =znf1(z).\displaystyle=z^{n}f_{1}(z).

With the adjoints now established, we see that the ordered quadruplet {𝒶,𝒷,1,2𝓃1}\{\mathpzc{a},\mathpzc{b},-1,2n-1\} forms a coupled SUSY on the spaces 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2}.

In a further parallel with the typical Segal-Bargmann space, the holomorphic function spaces 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} are reproducing kernel Hilbert spaces [28], a fact that will prove useful in the next section. We state the formal definition of a reproducing kernel Hilbert space (RKHS) given in [28] for self-containment.

Definition 3.7.

Let XX be a set and \mathfrak{H} be a vector subspace of the complex-valued functions defined on XX with Hilbert space structure. \mathfrak{H} is called a reproducing kernel Hilbert space if the evaluation functional is a continuous linear functional on \mathfrak{H}. The function fxf_{x} satisfying h,fx=h(x)\langle h,f_{x}\rangle=h(x) for all xXx\in X and hh\in\mathfrak{H} is called the reproducing kernel.

Theorem 3.8.

The holomorphic function spaces 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} are reproducing kernel Hilbert spaces with the respective reproducing kernels FwF_{w} and F~w\widetilde{F}_{w} given by

(3.33) Fw(z)\displaystyle F_{w}(z) =l=0el(z)el(w)¯\displaystyle=\sum_{l=0}^{\infty}e_{l}(z)\overline{e_{l}(w)}
(3.34) F~w(z)\displaystyle\widetilde{F}_{w}(z) =l=0e~l(z)e~l(w)¯\displaystyle=\sum_{l=0}^{\infty}\widetilde{e}_{l}(z)\overline{\widetilde{e}_{l}(w)}
Proof.

Elementary methods show that the series converge for all zz\in\mathbb{C} and thus guarantee the entirety of FwF_{w} and F~w\widetilde{F}_{w} (as functions of zz). From the definitions of ele_{l} and e~l\widetilde{e}_{l}, it is straightforward to see that FwF_{w} and F~w\widetilde{F}_{w} can be represented by the hypergeometric function F10{}_{0}F_{1} [27, Eq. 16.2.1]

(3.35) Fw(z)\displaystyle F_{w}(z) =F10(;12n;(zw¯)2n(2n)2)+Γ(12n)Γ(212n)(zw¯)2n1(2n)21nF10(;212n;(zw¯)2n(2n)2),\displaystyle={}_{0}F_{1}\bigg{(};\frac{1}{2n};\frac{(z\overline{w})^{2n}}{(2n)^{2}}\bigg{)}+\frac{\Gamma\big{(}\frac{1}{2n}\big{)}}{\Gamma\big{(}2-\frac{1}{2n}\big{)}}\frac{(z\overline{w})^{2n-1}}{(2n)^{2-\frac{1}{n}}}{}_{0}F_{1}\bigg{(};2-\frac{1}{2n};\frac{(z\overline{w})^{2n}}{(2n)^{2}}\bigg{)},
(3.36) F~w(z)\displaystyle\widetilde{F}_{w}(z) =Γ(12n)Γ(112n)(zw¯)n1(2n)11nF10(;112n;(zw¯)2n(2n)2)+(zw¯)nF10(;1+12n;(zw¯)2n(2n)2).\displaystyle=\frac{\Gamma\big{(}\frac{1}{2n}\big{)}}{\Gamma\big{(}1-\frac{1}{2n}\big{)}}\frac{(z\overline{w})^{n-1}}{(2n)^{1-\frac{1}{n}}}{}_{0}F_{1}\bigg{(};1-\frac{1}{2n};\frac{(z\overline{w})^{2n}}{(2n)^{2}}\bigg{)}+(z\overline{w})^{n}{}_{0}F_{1}\bigg{(};1+\frac{1}{2n};\frac{(z\overline{w})^{2n}}{(2n)^{2}}\bigg{)}.

We now show that Fw𝔉1F_{w}\in\mathfrak{F}_{1} and F~w𝔉2\widetilde{F}_{w}\in\mathfrak{F}_{2} for fixed ww\in\mathbb{C}. To this end we compute Fw,Fw1\langle F_{w},F_{w}\rangle_{1} and F~w,F~w2\langle\widetilde{F}_{w},\widetilde{F}_{w}\rangle_{2}. Noting that el,el1=1=e~l,e~l\langle e_{l},e_{l}\rangle_{1}=1=\langle\widetilde{e}_{l},\widetilde{e}_{l}\rangle for all ll,

(3.37) Fw,Fw1\displaystyle\langle F_{w},F_{w}\rangle_{1} =l=0Γ(12n)|w|4nl(2n)2lΓ(l+12n)l!+l=0Γ(12n)|w|4nl+4n2(2n)2l+21nΓ(l+212n)l!\displaystyle=\sum_{l=0}^{\infty}\frac{\Gamma\big{(}\frac{1}{2n}\big{)}|w|^{4nl}}{(2n)^{2l}\Gamma\big{(}l+\frac{1}{2n}\big{)}l!}+\sum_{l=0}^{\infty}\frac{\Gamma\big{(}\frac{1}{2n}\big{)}|w|^{4nl+4n-2}}{(2n)^{2l+2-\frac{1}{n}}\Gamma\big{(}l+2-\frac{1}{2n}\big{)}l!}
(3.38) F~w,F~w2\displaystyle\langle\widetilde{F}_{w},\widetilde{F}_{w}\rangle_{2} =l=0Γ(12n)|w|4nl+2n2(2n)2l+11nΓ(l+112n)l!+l=0Γ(12n)|w|4nl+2n(2n)2l+1Γ(l+1+12n)l!.\displaystyle=\sum_{l=0}^{\infty}\frac{\Gamma\big{(}\frac{1}{2n}\big{)}|w|^{4nl+2n-2}}{(2n)^{2l+1-\frac{1}{n}}\Gamma\big{(}l+1-\frac{1}{2n}\big{)}l!}+\sum_{l=0}^{\infty}\frac{\Gamma\big{(}\frac{1}{2n}\big{)}|w|^{4nl+2n}}{(2n)^{2l+1}\Gamma\big{(}l+1+\frac{1}{2n}\big{)}l!}.

A straightforward analysis shows that each series converges for all ww\in\mathbb{C} and so both inner products are finite. Thus Fw𝔉1F_{w}\in\mathfrak{F}_{1} and F~w𝔉2\widetilde{F}_{w}\in\mathfrak{F}_{2} for all ww\in\mathbb{C}.

That FwF_{w} and F~w\widetilde{F}_{w} are reproducing kernels follows directly from the fact that el,Fw1=el(w)\langle e_{l},F_{w}\rangle_{1}=e_{l}(w) and e~l,F~w2=e~l(w)\langle\widetilde{e}_{l},\widetilde{F}_{w}\rangle_{2}=\widetilde{e}_{l}(w), extended linearly to the whole space, noting that limits pass through by virtue of Fw1<\|F_{w}\|_{1}<\infty and F~w2<\|\widetilde{F}_{w}\|_{2}<\infty. ∎

Remark 7.

When n=1n=1, FwF_{w} and F~w\widetilde{F}_{w} simplify to ezw¯e^{z\bar{w}}, matching the reproducing kernel in the Segal-Bargmann space as expected.

In the next section, we will explore the relationship between the real line realization of the coupled SUSY and the holomorphic realization and develop a generalization of the Segal-Bargmann transform.

4. The Coupled SUSY Segal-Bargmann Transforms

As the ordered quadruplet {𝒶,𝒷,1,2𝓃1}\{\mathpzc{a},\mathpzc{b},-1,2n-1\} satisfies the same coupled SUSY relations as the ordered quadruplet {a,b,1,2n1}\{a,b,-1,2n-1\}, it is natural to find operators—the coupled SUSY Segal-Bargmann transforms—which make this isomorphism explicit. We would like the Segal-Bargmann transforms to map aa to 𝒶\mathpzc{a}, bb to 𝒷\mathpzc{b}, aa^{*} to 𝒶\mathpzc{a}^{*}, and bb^{*} to 𝒷\mathpzc{b}^{*}. Let the coupled SUSY Segal-Bargmann transforms 1:1𝔉1\mathcal{B}_{1}:\mathfrak{H}_{1}\to\mathfrak{F}_{1} and 2:2𝔉2\mathcal{B}_{2}:\mathfrak{H}_{2}\to\mathfrak{F}_{2} to be the bounded operators defined via the following commutative diagrams.

1\mathfrak{H}_{1}𝔉1\mathfrak{F}_{1}2\mathfrak{H}_{2}𝔉2\mathfrak{F}_{2}1\mathfrak{H}_{1}𝔉1\mathfrak{F}_{1}2\mathfrak{H}_{2}𝔉2\mathfrak{F}_{2}1\mathcal{B}_{1}aa𝒶\mathpzc{a}2\mathcal{B}_{2}1\mathcal{B}_{1}aa^{*}𝒶\mathpzc{a}^{*}2\mathcal{B}_{2}
1\mathfrak{H}_{1}𝔉1\mathfrak{F}_{1}2\mathfrak{H}_{2}𝔉2\mathfrak{F}_{2}1\mathfrak{H}_{1}𝔉1\mathfrak{F}_{1}2\mathfrak{H}_{2}𝔉2\mathfrak{F}_{2}1\mathcal{B}_{1}bb𝒷\mathpzc{b}2\mathcal{B}_{2}1\mathcal{B}_{1}bb^{*}𝒷\mathpzc{b}^{*}2\mathcal{B}_{2}
Figure 5. Commutative diagrams which define the coupled SUSY Segal-Bargmann transforms 1\mathcal{B}_{1} and 2\mathcal{B}_{2}

For f11f_{1}\in\mathfrak{H}_{1}, 1f1\mathcal{B}_{1}f_{1} is a holomorphic function and is therefore well-defined pointwise, similarly for f22f_{2}\in\mathfrak{H}_{2} and 2f2\mathcal{B}_{2}f_{2}. As the evaluation map z1f1(z)z\mapsto\mathcal{B}_{1}f_{1}(z) is bounded by way of Theorem 3.8, |1f1(z)|C|1f11|\mathcal{B}_{1}f_{1}(z)|\leq C\||\mathcal{B}_{1}f_{1}\|_{1}. Furthermore since 1\mathcal{B}_{1} is assumed to be bounded, |1f1(z)|Df1|\mathcal{B}_{1}f_{1}(z)|\leq D\|f_{1}\| and so by the Riesz representation theorem, there exists a function A1(z,)1A_{1}(z,\cdot)\in\mathfrak{H}_{1} for each fixed zz such that

(4.1) 1f1(z)=A1(z,x)f1(x)dx.\mathcal{B}_{1}f_{1}(z)=\int_{\mathbb{R}}A_{1}(z,x)f_{1}(x)\,\mathrm{d}x.

Similar is true for 2f2\mathcal{B}_{2}f_{2}. The commutative diagrams in conjunction with the above representation for 1\mathcal{B}_{1} and 2\mathcal{B}_{2} can be realized as

(4.2) 1zn1ddzA1(z,x)f1(x)dx\displaystyle\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z}\int_{\mathbb{R}}A_{1}(z,x)f_{1}(x)\,\mathrm{d}x =A2(z,x)12(1xn1ddx+xn)f1(x)dx\displaystyle=\int_{\mathbb{R}}A_{2}(z,x)\frac{1}{\sqrt{2}}\bigg{(}\frac{1}{x^{n-1}}\frac{\mathrm{d}}{\mathrm{d}x}+x^{n}\bigg{)}f_{1}(x)\,\mathrm{d}x
(4.3) znA1(z,x)f1(x)dx\displaystyle z^{n}\int_{\mathbb{R}}A_{1}(z,x)f_{1}(x)\,\mathrm{d}x =A2(z,x)12(1xn1ddx+xn)f1(x)dx\displaystyle=\int_{\mathbb{R}}A_{2}(z,x)\frac{1}{\sqrt{2}}\bigg{(}-\frac{1}{x^{n-1}}\frac{\mathrm{d}}{\mathrm{d}x}+x^{n}\bigg{)}f_{1}(x)\,\mathrm{d}x
(4.4) znA2(z,x)f2(x)dx\displaystyle z^{n}\int_{\mathbb{R}}A_{2}(z,x)f_{2}(x)\,\mathrm{d}x =A1(z,x)12(ddx1xn1+xn)f2(x)dx\displaystyle=\int_{\mathbb{R}}A_{1}(z,x)\frac{1}{\sqrt{2}}\bigg{(}-\frac{\mathrm{d}}{\mathrm{d}x}\frac{1}{x^{n-1}}+x^{n}\bigg{)}f_{2}(x)\,\mathrm{d}x
(4.5) ddz1zn1A2(z,x)f2(x)dx\displaystyle\frac{\mathrm{d}}{\mathrm{d}z}\frac{1}{z^{n-1}}\int_{\mathbb{R}}A_{2}(z,x)f_{2}(x)\,\mathrm{d}x =A1(z,x)12(ddx1xn1+xn)f2(x)dx.\displaystyle=\int_{\mathbb{R}}A_{1}(z,x)\frac{1}{\sqrt{2}}\bigg{(}\frac{\mathrm{d}}{\mathrm{d}x}\frac{1}{x^{n-1}}+x^{n}\bigg{)}f_{2}(x)\,\mathrm{d}x.

Since we wish for these identities to hold for all f1f_{1} and f2f_{2}, applying integrations by parts—assuming sufficiently nice (exponentially decaying) behavior for A1A_{1} and A2A_{2}—that the following relations hold:

(4.6) 1zn1zA1(z,x)\displaystyle\frac{1}{z^{n-1}}\frac{\partial}{\partial z}A_{1}(z,x) =12(x1xn1+xn)A2(z,x)\displaystyle=\frac{1}{\sqrt{2}}\bigg{(}-\frac{\partial}{\partial x}\frac{1}{x^{n-1}}+x^{n}\bigg{)}A_{2}(z,x)
(4.7) z1zn1A2(z,x)\displaystyle\frac{\partial}{\partial z}\frac{1}{z^{n-1}}A_{2}(z,x) =12(1xn1x+xn)A1(z,x)\displaystyle=\frac{1}{\sqrt{2}}\bigg{(}-\frac{1}{x^{n-1}}\frac{\partial}{\partial x}+x^{n}\bigg{)}A_{1}(z,x)
(4.8) znA1(z,x)\displaystyle z^{n}A_{1}(z,x) =12(x1xn1+xn)A2(z,x)\displaystyle=\frac{1}{\sqrt{2}}\bigg{(}\frac{\partial}{\partial x}\frac{1}{x^{n-1}}+x^{n}\bigg{)}A_{2}(z,x)
(4.9) znA2(z,x)\displaystyle z^{n}A_{2}(z,x) =12(1xn1x+xn)A1(z,x)\displaystyle=\frac{1}{\sqrt{2}}\bigg{(}\frac{1}{x^{n-1}}\frac{\partial}{\partial x}+x^{n}\bigg{)}A_{1}(z,x)

By taking linear combinations, we obtain

(4.10) 12(1zn1z+zn)A1(z,x)\displaystyle\frac{1}{\sqrt{2}}\bigg{(}\frac{1}{z^{n-1}}\frac{\partial}{\partial z}+z^{n}\bigg{)}A_{1}(z,x) =xnA2(z,x)\displaystyle=x^{n}A_{2}(z,x)
(4.11) 12(z1zn1+zn)A2(z,x)\displaystyle\frac{1}{\sqrt{2}}\bigg{(}\frac{\partial}{\partial z}\frac{1}{z^{n-1}}+z^{n}\bigg{)}A_{2}(z,x) =xnA1(z,x)\displaystyle=x^{n}A_{1}(z,x)

We must first note a pair of identities that will aid in the solution of the above partial differential equations.

Proposition 4.1.

The following basic “integrating factor”-like identities hold.

(4.12) (1tn1ddt+tn)et2n2nf(t)\displaystyle\bigg{(}\frac{1}{t^{n-1}}\frac{\mathrm{d}}{\mathrm{d}t}+t^{n}\bigg{)}e^{-\frac{t^{2n}}{2n}}f(t) =et2n2n1tn1ddtf(t),\displaystyle=e^{-\frac{t^{2n}}{2n}}\frac{1}{t^{n-1}}\frac{\mathrm{d}}{\mathrm{d}t}f(t),
(4.13) (ddt1tn1+tn)et2n2nf(t)\displaystyle\bigg{(}\frac{\mathrm{d}}{\mathrm{d}t}\frac{1}{t^{n-1}}+t^{n}\bigg{)}e^{-\frac{t^{2n}}{2n}}f(t) =et2n2nddt1tn1f(t).\displaystyle=e^{-\frac{t^{2n}}{2n}}\frac{\mathrm{d}}{\mathrm{d}t}\frac{1}{t^{n-1}}f(t).

Noting these identities and the symmetry in xx and zz in (4.8)–(4.11), define B1(z,x)=ez2n2nA1(z,x)ex2n2nB_{1}(z,x)=e^{\frac{z^{2n}}{2n}}A_{1}(z,x)e^{\frac{x^{2n}}{2n}} and B2(z,x)=ez2n2nA2(z,x)ex2n2nB_{2}(z,x)=e^{\frac{z^{2n}}{2n}}A_{2}(z,x)e^{\frac{x^{2n}}{2n}}. B1B_{1} and B2B_{2} then solve simpler coupled partial differential equations:

(4.14) 1xn1xB1(z,x)\displaystyle\frac{1}{x^{n-1}}\frac{\partial}{\partial x}B_{1}(z,x) =2znB2(z,x)\displaystyle=\sqrt{2}z^{n}B_{2}(z,x)
(4.15) 1zn1zB1(z,x)\displaystyle\frac{1}{z^{n-1}}\frac{\partial}{\partial z}B_{1}(z,x) =2xnB2(z,x)\displaystyle=\sqrt{2}x^{n}B_{2}(z,x)
(4.16) x1xn1B2(z,x)\displaystyle\frac{\partial}{\partial x}\frac{1}{x^{n-1}}B_{2}(z,x) =2znB1(z,x)\displaystyle=\sqrt{2}z^{n}B_{1}(z,x)
(4.17) z1zn1B2(z,x)\displaystyle\frac{\partial}{\partial z}\frac{1}{z^{n-1}}B_{2}(z,x) =2xnB1(z,x).\displaystyle=\sqrt{2}x^{n}B_{1}(z,x).

These equations are again symmetric in xx and zz and resemble the intertwining relations for ρ1\rho_{1} and ρ2\rho_{2} with the primary differences being the appearance of the multiplicative factor of 2\sqrt{2} and xx rather than z¯\bar{z}. Taking cues from the analysis for ρ1\rho_{1} and ρ2\rho_{2}, B1B_{1} and B2B_{2} have the following forms

(4.18) B1(z,x)\displaystyle B_{1}(z,x) =αl=02l(zx)2nl(2n)2lΓ(l+12n)l!+βl=02l+12(zx)2nl+2n1(2n)2l+1Γ(l+212n)l!,\displaystyle=\alpha\sum_{l=0}^{\infty}\frac{2^{l}(zx)^{2nl}}{(2n)^{2l}\Gamma\big{(}l+\frac{1}{2n}\big{)}l!}+\beta\sum_{l=0}^{\infty}\frac{2^{l+\frac{1}{2}}(zx)^{2nl+2n-1}}{(2n)^{2l+1}\Gamma(l+2-\frac{1}{2n}\big{)}l!},
(4.19) B2(z,x)\displaystyle B_{2}(z,x) =βl=02l(zx)2nl+n1(2n)2lΓ(l+112n)l!+αl=02l+12(zx)2nl+n(2n)2l+1Γ(l+1+12n)l!,\displaystyle=\beta\sum_{l=0}^{\infty}\frac{2^{l}(zx)^{2nl+n-1}}{(2n)^{2l}\Gamma\big{(}l+1-\frac{1}{2n}\big{)}l!}+\alpha\sum_{l=0}^{\infty}\frac{2^{l+\frac{1}{2}}(zx)^{2nl+n}}{(2n)^{2l+1}\Gamma(l+1+\frac{1}{2n}\big{)}l!},

where α\alpha and β\beta are to be determined. To uniquely identify α\alpha and β\beta (which will also give unitarity as we will see shortly), we require that

(4.20) e0(z)\displaystyle e_{0}(z) =A1(z,x)ψ0(x)dx\displaystyle=\int_{\mathbb{R}}A_{1}(z,x)\psi_{0}(x)\,\mathrm{d}x
(4.21) e~0(z)\displaystyle\widetilde{e}_{0}(z) =A2(z,x)ψ~0(x)dx\displaystyle=\int_{\mathbb{R}}A_{2}(z,x)\widetilde{\psi}_{0}(x)\,\mathrm{d}x

Under these assumptions, α\alpha and β\beta are given by

(4.22) α\displaystyle\alpha =n1214nΓ(12n)\displaystyle=n^{\frac{1}{2}-\frac{1}{4n}}\sqrt{\Gamma\bigg{(}\frac{1}{2n}\bigg{)}}
(4.23) β\displaystyle\beta =212+12nn12+34nΓ(12n).\displaystyle=2^{-\frac{1}{2}+\frac{1}{2n}}n^{-\frac{1}{2}+\frac{3}{4n}}\sqrt{\Gamma\bigg{(}\frac{1}{2n}\bigg{)}}.

Simply, to determine α\alpha, plug in z=0z=0 on both sides and evaluate the integral; to determine β\beta, divide both sides by zn1z^{n-1}, plug in z=0z=0, and evaluate the integral.

As before with FwF_{w} and F~w\widetilde{F}_{w}, representations in terms of the hypergeometric function F10{}_{0}F_{1} exist for B1B_{1} and B2B_{2} and thus A1A_{1} and A2A_{2}:

(4.24) B1(z,x)\displaystyle B_{1}(z,x) =αΓ(12n)F10(;12n;(zx)2n2n2)+β2nΓ(212n)(zx)2n1F10(;212n;(zx)2n2n2),\displaystyle=\frac{\alpha}{\Gamma\big{(}\frac{1}{2n}\big{)}}{}_{0}F_{1}\bigg{(};\frac{1}{2n};\frac{(zx)^{2n}}{2n^{2}}\bigg{)}+\frac{\beta}{\sqrt{2}n\Gamma\big{(}2-\frac{1}{2n}\big{)}}(zx)^{2n-1}{}_{0}F_{1}\bigg{(};2-\frac{1}{2n};\frac{(zx)^{2n}}{2n^{2}}\bigg{)},
(4.25) B2(z,x)\displaystyle B_{2}(z,x) =βΓ(112n)(zx)n1F10(;112n;(zx)2n2n2)+2αΓ(12n)(zx)nF10(;1+12n;(zx)2n2n2).\displaystyle=\frac{\beta}{\Gamma\big{(}1-\frac{1}{2n}\big{)}}(zx)^{n-1}{}_{0}F_{1}\bigg{(};1-\frac{1}{2n};\frac{(zx)^{2n}}{2n^{2}}\bigg{)}+\frac{\sqrt{2}\alpha}{\Gamma\big{(}\frac{1}{2n}\big{)}}(zx)^{n}{}_{0}F_{1}\bigg{(};1+\frac{1}{2n};\frac{(zx)^{2n}}{2n^{2}}\bigg{)}.
Remark 8.

When n=1n=1, α=π14=β\alpha=\pi^{\frac{1}{4}}=\beta and 1\mathcal{B}_{1} and 2\mathcal{B}_{2} simplify to a single integral operator given by

(4.26) f(z)=1π14ez22e2zxex22f(x)dx\mathcal{B}f(z)=\frac{1}{\pi^{\frac{1}{4}}}\int_{\mathbb{R}}e^{-\frac{z^{2}}{2}}e^{\sqrt{2}zx}e^{-\frac{x^{2}}{2}}f(x)\,\mathrm{d}x

which can be recognized as the usual Segal-Bargmann transform.

Remark 9.

Of particular note is the appearance of the generalized Gaussians ex2n2ne^{-\frac{x^{2n}}{2n}} and ez2n2ne^{-\frac{z^{2n}}{2n}} in the coupled SUSY Segal-Bargmann transform integral kernels. In the traditional Segal-Bargmann setting, the appearance of the Gaussians ez22e^{-\frac{z^{2}}{2}} and ex22e^{-\frac{x^{2}}{2}} and the exponential e2zxe^{\sqrt{2}zx} in the traditional Segal-Bargmann transform integral kernel seem like a happy accident; however in this general setting, it becomes clear that these are intrinsic and emerge from the “integrating factor”-like simplification of the ladder operators. Furthermore, the connection between the weight ρ\rho and the exponential is clear in this setting: both obey very similar differential equations.

Theorem 4.2.

The Segal-Bargmann transforms 1:1𝔉1\mathcal{B}_{1}:\mathfrak{H}_{1}\to\mathfrak{F}_{1} and 2:2𝔉2\mathcal{B}_{2}:\mathfrak{H}_{2}\to\mathfrak{F}_{2} are unitary, i.e. for f11f_{1}\in\mathfrak{H}_{1} and f22f_{2}\in\mathfrak{H}_{2},

(4.27) 1f1,1f11\displaystyle\langle\mathcal{B}_{1}f_{1},\mathcal{B}_{1}f_{1}\rangle_{1} =f1,f1,\displaystyle=\langle f_{1},f_{1}\rangle,
(4.28) 2f2,2f22\displaystyle\langle\mathcal{B}_{2}f_{2},\mathcal{B}_{2}f_{2}\rangle_{2} =f2,f2.\displaystyle=\langle f_{2},f_{2}\rangle.
Proof.

By construction of A1A_{1} and A2A_{2},

(4.29) 1ψ0\displaystyle\mathcal{B}_{1}\psi_{0} =e0,\displaystyle=e_{0},
(4.30) 2ψ~0\displaystyle\mathcal{B}_{2}\widetilde{\psi}_{0} =e~0.\displaystyle=\widetilde{e}_{0}.

Furthermore, a direct computation shows that 1ψ1=e1\mathcal{B}_{1}\psi_{1}=e_{1}. This can be shown by noting that aψ1=2n1ψ~0a\psi_{1}=\sqrt{2n-1}\widetilde{\psi}_{0} and using the coupled SUSY Segal-Bargmann transform relations to get that 𝒶1ψ1=2𝒶ψ1=2𝓃12ψ~0\mathpzc{a}\mathcal{B}_{1}\psi_{1}=\mathcal{B}_{2}a\psi_{1}=\sqrt{2n-1}\mathcal{B}_{2}\widetilde{\psi}_{0}. Solving the differential equation 1zn1ddz1ψ1=2n1e~0\frac{1}{z^{n-1}}\frac{\mathrm{d}}{\mathrm{d}z}\mathcal{B}_{1}\psi_{1}=\sqrt{2n-1}\widetilde{e}_{0} gives that 1ψ1=e1\mathcal{B}_{1}\psi_{1}=e_{1}. Showing 2ψ~1=e~1\mathcal{B}_{2}\widetilde{\psi}_{1}=\widetilde{e}_{1} proceeds similarly.

By successively applying raising operators and using these four base cases, inductively it can be shown more generally that

1ψl\displaystyle\mathcal{B}_{1}\psi_{l} =el,\displaystyle=e_{l},
2ψ~l\displaystyle\mathcal{B}_{2}\widetilde{\psi}_{l} =e~l.\displaystyle=\widetilde{e}_{l}.

This follows from the observation that abψlψl+2a^{*}b^{*}\psi_{l}\propto\psi_{l+2}. The proportionality constant is determined by the eigenvalue exactly, and both representations (real line and holomorphic representations) of the coupled SUSY in question have the same spectra as they are both unbroken and share the same γ\gamma and δ\delta.

Thus 1\mathcal{B}_{1} and 2\mathcal{B}_{2} are surjective and norm-preserving as all ψl\psi_{l}, ψ~l\widetilde{\psi}_{l}, ele_{l}, and e~l\widetilde{e}_{l} are normalized and form orthonormal bases for their respective spaces, and therefore 1\mathcal{B}_{1} and 2\mathcal{B}_{2} are unitary. ∎

Much like the typical Segal-Bargmann transform, we also have explicit expressions for the inverse coupled SUSY Segal-Bargmann transforms.

Corollary 4.3.

The coupled SUSY Segal-Bargmann transforms 1\mathcal{B}_{1} and 2\mathcal{B}_{2} have inverses 11:𝔉11\mathcal{B}_{1}^{-1}:\mathfrak{F}_{1}\to\mathfrak{H}_{1} and 21:𝔉22\mathcal{B}_{2}^{-1}:\mathfrak{F}_{2}\to\mathfrak{H}_{2}, respectively, given by

(4.31) 11f1(x)\displaystyle\mathcal{B}_{1}^{-1}f_{1}(x) =A1(z¯,x)f1(z)ρ1(z,z¯)dA(z)\displaystyle=\int_{\mathbb{C}}A_{1}(\bar{z},x)f_{1}(z)\rho_{1}(z,\bar{z})\,\mathrm{d}A(z)
(4.32) 21f2(x)\displaystyle\mathcal{B}_{2}^{-1}f_{2}(x) =A2(z¯,x)f2(z)ρ2(z,z¯)dA(z)\displaystyle=\int_{\mathbb{C}}A_{2}(\bar{z},x)f_{2}(z)\rho_{2}(z,\bar{z})\,\mathrm{d}A(z)
Proof.

The proof follows from a simple application of Fubini-Tonelli, noting the exponential decay of A1A_{1} and A2A_{2} as well as ρ1\rho_{1} and ρ2\rho_{2} or by acting directly on the basis elements ele_{l} and e~l\widetilde{e}_{l} and extending linearly. ∎

Remark 10.

A quick inspection shows that there is a canonical inclusion of the Segal-Bargmann spaces 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} into 𝔉\mathfrak{F} by sending zkz^{k} in 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2} to zkz^{k} in 𝔉\mathfrak{F}. A function in L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x) will in general have distinct holomorphic representations through different coupled SUSY Segal-Bargmann transforms, not in contradiction with the identity theorem as the representations are related by the Segal-Bargmann transforms and are not directly equal to each other.

5. The Connection to Short-Time Transforms and Coherent States

The usual Segal-Bargmann transform is closely related to the short-time Fourier transform and coherent states. The short-time Fourier transform is a time-frequency representation wherein the frequency content of segment of a function is analyzed [18]. This is achieved by taking the Fourier transform of the function with a moving window. Rigorously,

Definition 5.1.

Given (typically non-negative and normalized) gL2(,dx)g\in L^{2}(\mathbb{R},\mathrm{d}x), the short-time Fourier transform with window function gg is the map 𝒱g:L2(,dx)L2(2,dxdy)\mathcal{V}_{g}:L^{2}(\mathbb{R},\mathrm{d}x)\to L^{2}(\mathbb{R}^{2},\mathrm{d}x\,\mathrm{d}y) given by

(5.1) 𝒱gf(ω,t)=12πeiωτg(τt)f(τ)dτ.\mathcal{V}_{g}f(\omega,t)=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}e^{-i\omega\tau}g(\tau-t)f(\tau)\,\mathrm{d}\tau.

This is a well-defined integral operator as gtfL1(,dx)g_{t}f\in L^{1}(\mathbb{R},\mathrm{d}x) by Cauchy-Schwartz. That it maps to L2(2,dxdy)L^{2}(\mathbb{R}^{2},\mathrm{d}x\,\mathrm{d}y) follows from the unitarity of the Fourier transform on L2(,dx)L^{2}(\mathbb{R},\mathrm{d}x).

Coherent states have multiple different definitions that are not in general equivalent [9]. At least three competing definitions exist: a coherent state is an eigenfunction of the lowering operator, coherent states are generated from dilation operators, and coherent states are minimum uncertainty states (for a given pair of non-commuting operators). In some cases, these definitions can coincide, for instance in the oscillator algebra case. For our purposes, we will be concerned with eigenfunctions of a lowering operator.

The connection between the short-time Fourier transform and the Segal-Bargmann transform can be made by letting z=t2i2ωz=\frac{t}{\sqrt{2}}-i\sqrt{2}\omega and g(t)=1π14et22g(t)=\frac{1}{\pi^{\frac{1}{4}}}e^{-\frac{t^{2}}{2}} to get

(5.2) 𝒱gf(ω,t)=12πet24ω2e2iωtf(z).\displaystyle\mathcal{V}_{g}f(\omega,t)=\frac{1}{\sqrt{2\pi}}e^{-\frac{t^{2}}{4}-\omega^{2}}e^{2i\omega t}\mathcal{B}f(z).

The connection between coherent states and the Segal-Bargmann transform is immediate as the Segal-Bargmann kernel is an eigenfunction of the lowering operator.

In [34], the authors defined a short-time analogue for the Φn\Phi_{n} transforms built upon the concept of moving windows. The Φn\Phi_{n} transform of a sufficiently nice function as an integral transform is given by

(5.3) Φnf(y)=φn(xy)f(x)dx,\Phi_{n}f(y)=\int_{\mathbb{R}}\varphi_{n}(xy)f(x)\,\mathrm{d}x,

where φn\varphi_{n} is a solution to the differential equation

(5.4) ddx1x2n2ddxφn(xy)=y2nφn(xy)-\frac{\textrm{d}}{\textrm{d}x}\frac{1}{x^{2n-2}}\frac{\textrm{d}}{\textrm{d}x}\varphi_{n}(xy)=y^{2n}\varphi_{n}(xy)

with φn(0)=n(2n)12nΓ(12n)\varphi_{n}(0)=\frac{n}{(2n)^{\frac{1}{2n}}\Gamma\big{(}\frac{1}{2n}\big{)}} and φn(0)=in(2n)212nΓ(212n)\varphi_{n}^{\prime}(0)=-i\frac{n}{(2n)^{2-\frac{1}{2n}}\Gamma\big{(}2-\frac{1}{2n}\big{)}}. The short-time Φn\Phi_{n} transform of a function ff with window gg is then defined to be

(5.5) 𝒱g(n)(y,x)=φn(y(xx))g(xx)f(x)dx.\mathcal{V}_{g}^{(n)}(y,x)=\int_{\mathbb{R}}\varphi_{n}(y(x^{\prime}-x))g(x^{\prime}-x)f(x^{\prime})\,\mathrm{d}x^{\prime}.

The Φn\Phi_{n} transforms are closely related to the coupled SUSY involving aa and bb by the following identity:

(5.6) Φnab=abΦn.\Phi_{n}ab=-ab\Phi_{n}.

which can be confirmed from a simple integration by parts. However despite connection to the coupled SUSYs at hand, a simple inspection shows that the short-time Φn\Phi_{n} transforms are distinct from the coupled SUSY Segal-Bargmann transforms. This is a break from traditional Segal-Bargmann transform theory and is a direct result of the fact that the Φn\Phi_{n} transforms do not play nicely with translations in general—this is a (nearly) uniquely identifying property of the Fourier transform. Thus the coupled SUSY Segal-Bargmann transforms may be viewed as an altogether new time-frequency representation.

Despite the break from short-time Φn\Phi_{n} transform, the coupled SUSY Segal-Bargmann kernels can be viewed as coherent states and thus the coupled SUSY Segal-Bargmann transforms can be viewed as coherent state transforms. Rewriting (4.9) and (4.10) in a matricial form, we have that

(5.7) (0ba0)(A1A2)=zn(A1A2),\begin{pmatrix}0&b\\ a&0\end{pmatrix}\begin{pmatrix}A_{1}\\ A_{2}\end{pmatrix}=z^{n}\begin{pmatrix}A_{1}\\ A_{2}\end{pmatrix},

or by squaring,

(5.8) (ba00ab)(A1A2)=z2n(A1A2).\begin{pmatrix}ba&0\\ 0&ab\end{pmatrix}\begin{pmatrix}A_{1}\\ A_{2}\end{pmatrix}=z^{2n}\begin{pmatrix}A_{1}\\ A_{2}\end{pmatrix}.

Thus the combined state (A1A2)\begin{pmatrix}A_{1}\\ A_{2}\end{pmatrix} is an eigenfunction of a lowering operator (for a matricial Hamiltonian in this case) and thus a generalized 𝔰𝔲(1,1)\mathfrak{su}(1,1) coherent state.

This work is closely related to the coherent states and Segal-Bargmann transform developed in [8]. Therein, the Segal-Bargmann transform also takes the form of the hypergeometric function F10{}_{0}F_{1} and the weight functions and reproducing kernels are similar. The Segal-Bargmann space is also spanned by zlz^{l} for l=0,1,2,l=0,1,2,\ldots, whereas the coupled SUSY Segal-Bargmann spaces developed herein are stricter vector subspaces.

Furthermore, in [8], limits were taken to obtain the Heisenberg-Weyl algebra, imagining the 𝔰𝔲(1,1)\mathfrak{su}(1,1) Lie algebra as a qq-deformation of the Heisenberg-Weyl algebra, and thus the usual Segal-Bargmann transform is not directly generalized, whereas the quantum harmonic oscillator is baked directly into coupled SUSY and so direct generalizations exist in our case.

Moreover, the representations for the 𝔰𝔲(1,1)\mathfrak{su}(1,1) elements in [8] were of a relatively complicated form, particularly the lowering operator, but their analogues in 𝒶\mathpzc{a} and 𝒷\mathpzc{b} and their adjoints and products are of a very simple form in 𝔉1\mathfrak{F}_{1} and 𝔉2\mathfrak{F}_{2}. The work presented herein is also of a supersymmetric nature which is not present in their analysis.

The work presented herein elucidates some of the traditional Segal-Bargmann transform theory. In the traditional setting, due to properties of exponentials, there is some symmetry between the weight ρ\rho and the integral kernel AA which obfuscates some of the mathematics. In this general setting, we see that this symmetry no longer exists. However, when considering the asymptotic forms of the weights ρ1\rho_{1} and ρ2\rho_{2}, some of the aforementioned symmetry reveals itself again.

Acknowledgements

The author would like to thank Dr. John R. Klauder for the suggestion for this avenue of research and the late Dr. Donald J. Kouri for his mentorship in this project.

References

  • [1] N. I. Akhiezer. Lectures on Integral Transforms. American Mathematical Society, 1988.
  • [2] D. Alpay, F. Colombo, I. Sabadini, and G. Salomon. The Fock Space in the Slice Hyperholomorphic Setting. In S. Bernstein, U. Kähler, I. Sabadini, and F. Sommen, editors, Hypercomplex Analysis: New Perspectives and Applications, pages 43–59. Springer International Publishing, 2014.
  • [3] Daniel Alpay and Motke Porat. Generalized fock spaces and the stirling numbers. Journal of Mathematical Physics, 59(6):063509, 2018.
  • [4] S. Barbier and S. Claerebout. A Schrödinger model, Fock model and Intertwining Segal-Bargmann Transform for the Exceptional Lie Superalgebra D(2,1;α)D(2,1;\alpha). 4 2021.
  • [5] S. Barbier, S. Claerebout, and H. De Bie. A Fock Model and the Segal-Bargmann Transform for the Minimal Representation of the Orthosymplectic Lie Superalgebra 𝔬𝔰𝔭(m,2|2n)\mathfrak{osp}(m,2|2n). Symmetry, Integrability, and Geometry: Methods and Applications, 16, 2020.
  • [6] V. Bargmann. On a Hilbert Space of Analytic Functions and an Associated Integral Transform Part I. Communications on Pure and Applied Mathematics, 14(3):187–214, 1961.
  • [7] V. Bargmann. On a Hilbert Space of Analytie Functions and an Associated Integral Transform. Part II. A Family of Related Function Spaces Application to Distribution Theory. Communications on Pure and Applied Mathematics, 20(1):1–101, 1967.
  • [8] A. O. Barut and L. Girardello. New “Coherent” States associated with non-compact groups. Communications in Mathematical Physics, 21:41–55, 1971.
  • [9] C. Brif. SU(2) and SU(1,1) algebra eigenstates: A unified analytic approach to coherent and intelligent states. International Journal of Theoretical Physics, 36(7):1651–1682, 1997.
  • [10] Alice Z. Chan. The Segal-Bargmann transform on classical matrix Lie groups. Journal of Functional Analysis, 278(9):108430, 2020.
  • [11] L. Cnudde and H. De Bie. Slice Segal–Bargmann Transform. Journal of Physics A: Mathematical and Theoretical, 50(25):255207, jun 2017.
  • [12] F. Cooper, Khare A., and U. Sukhatme. Supersymmetry and Quantum Mechanics. Physics Reports, 251(5):267–385, 1995.
  • [13] J. David and C. Fernandez. Supersymmetric Quantum Mechanics. AIP Conference Proceedings, 1287(1):3–36, 2010.
  • [14] A. De Martino and M. Diki. On the Quaternionic Short-Time Fourier and Segal–Bargmann Transforms. Mediterranean Journal of Mathematics, 18(3), 2021.
  • [15] K. Diki and A. Ghanmi. A Quaternionic Analogue of the Segal–Bargmann Transform. Complex Analysis and Operator Theory, 11(2):457–473, 2017.
  • [16] B. K. Driver, B. C. Hall, and T. Kemp. The Complex-Time Segal–Bargmann Transform. Journal of Functional Analysis, 278(1):108303, 2020.
  • [17] S. Eaknipitsari and W. Lewkeeratiyutkul. Clifford Algebra-Valued Segal–Bargmann Transform and Taylor Isomorphism. Advances in Applied Clifford Algebras, 31(5), 2021.
  • [18] K. Gröchenig. Foundations of Time-Frequency Analysis. Birkhäuser, 2000.
  • [19] B. C. Hall. Holomorphic Methods in Mathematical Physics.
  • [20] B. C. Hall. The Segal-Bargmann “Coherent State” Transform for Compact Lie Groups. Journal of Functional Analysis, 122(1):103–151, 1994.
  • [21] B. C. Hall. Geometric Quantization and the Generalized Segal-Bargmann Transform for Lie Groups of Compact Type. Communications in Mathematical Physics, 226(2):233–268, 2002.
  • [22] J. Hilgert, T. Kobayashi, J. Möllers, and B. Ørsted. Fock Model and Segal–Bargmann Transform for Minimal Representations of Hermitian Lie Groups. Journal of Functional Analysis, 263(11):3492–3563, 2012.
  • [23] J. Hilgert and G. Zhang. Segal–Bargmann and Weyl Transforms on Compact Lie Groups. Monatshefte für Mathematik, 158:285–305, 2009.
  • [24] William D. Kirwin, José M. Mourão, and João P. Nunes. Complex Time Evolution in Geometric Quantization and Generalized Coherent State Transforms. Journal of Functional Analysis, 265(8):1460–1493, 2013.
  • [25] Hans van Leeuwen and Hans Maassen. A q deformation of the gauss distribution. Journal of Mathematical Physics, 36(9):4743–4756, 1995.
  • [26] José Mourão, João P. Nunes, and Tao Qian. Coherent State Transforms and the Weyl Equation in Clifford Analysis. Journal of Mathematical Physics, 58(1):013503, 2017.
  • [27] F. Oliver, D. Lozier, R. Boisvert, and C. Clark. The NIST Handbook of Mathematical Functions. Cambridge University Press, 2021.
  • [28] Vern I. Paulsen and Mrinal Raghupathi. An Introduction to the Theory of Reproducing Kernel Hilbert Spaces. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2016.
  • [29] Joel A. Rosenfeld, Benjamin Russo, and Warren E. Dixon. The mittag leffler reproducing kernel hilbert spaces of entire and analytic functions. Journal of Mathematical Analysis and Applications, 463(2):576–592, 2018.
  • [30] I. E. Segal. Mathematical Problems of Relativistic Physics. American Mathematical Society, 2 edition, 1963.
  • [31] M. B. Stenzel. The Segal–Bargmann Transform on a Symmetric Space of Compact Type. Journal of Functional Analysis, 165(1):44–58, 1999.
  • [32] R. F. Streater. The Representations of the Oscillator Group. Communications in Mathematical Physics, 4(3):217 – 236, 1967.
  • [33] H. P. Thienel. A Generalization of the Bargmann-Fock Representation to Supersymmetry by Holomorphic Differential Geometry. J. Phys. A, 29:6983, 1996.
  • [34] C. L. Williams, B. G. Bodmann, and D. J. Kouri. Fourier and Beyond: Invariance Properties of a Family of Fourier-like Integral Transforms. Journal of Fourier Analysis and Applications, page to appear.
  • [35] C. L. Williams, N. N. Pandya, B. G. Bodmann, and D. J. Kouri. Coupled Supersymmetry and Ladder Structures Beyond the Harmonic Oscillator. Molecular Physics, 116(19-20):2599–2612, 2018.