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

Phase retrieval for affine groups over prime fields

David Bartusel, Hartmut Führ, Vignon Oussa D. Bartusel, H. Führ
Lehrstuhl für Geometrie und Analysis
RWTH Aachen University
D-52056 Aachen
Germany
email: [email protected], [email protected]
V. Oussa
Department of Mathematics
Bridgewater State University
MA 02234
USA
email: [email protected]
Abstract

We study phase retrieval for group frames arising from permutation representations, focusing on the action of the affine group of a finite field. We investigate various versions of the phase retrieval problem, including conjugate phase retrieval, sign retrieval, and matrix recovery. Our main result establishes that the canonical irreducible representation of the affine group pp\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{\ast} (with pp prime), acting on the vectors in p\mathbb{C}^{p} with zero-sum, has the strongest retrieval property, allowing to reconstruct matrices from scalar products with a group orbit consisting of rank-one projections. We explicitly characterize the generating vectors that ensure this property, provide a linear matrix recovery algorithm and explicit examples of vectors that allow matrix recovery. We also comment on more general permutation representations.

Keywords: phase retrieval; conjugate phase retrieval; sign retrieval; matrix recovery; Pauli pairs; group frame; group Fourier transform

AMS Subject Classification: 42C15; 42A38; 94A12; 20B20

1 Introduction

The phase retrieval problem for frames was introduced in [3], with motivation coming from applications such as speech recognition and imaging. Since its introduction to the field of mathematical signal processing, phase retrieval has developed into an autonomous area of mathematical research with rich connections to diverse mathematical fields such as complex analysis, optimization, numerical analysis, algebraic geometry, harmonic analysis and representation theory, as witnessed by the review papers [16, 12] and their extensive lists of references.

The phase retrieval problem can be briefly described as follows: Given a system Ψ=(ψi)iId\Psi=(\psi_{i})_{i\in I}\subset\mathbb{C}^{d}, we aim to reconstruct fdf\in\mathbb{C}^{d}, up to a scalar factor α𝕋\alpha\in\mathbb{T}, from the family (|f,ψi|)iI\left(|\langle f,\psi_{i}\rangle|\right)_{i\in I}. Here 𝕋\mathbb{T} denotes the multiplicative group of complex numbers with modulus one. Clearly, a necessary condition for the system Ψ\Psi is that span(Ψ)=d{\rm span}(\Psi)=\mathbb{C}^{d}, i.e. the linear coefficient map f(f,ψi)iI)If\mapsto(\langle f,\psi_{i}\rangle)_{i\in I})\in\mathbb{C}^{I} is necessarily injective. This observation motivates the name “phase retrieval”: Solving the phase retrieval problem for a system Ψ\Psi spanning d\mathbb{C}^{d} amounts to reconstructing the phases of the expansion coefficients (f,ψi)iI(\langle f,\psi_{i}\rangle)_{i\in I} from their moduli (which is a nonlinear problem), and then inverting the coefficient map (which is a standard, linear procedure).

It is therefore clear that phase retrieval requires some redundancy of Ψ\Psi: A basis Ψ\Psi does not admit phase retrieval, since each expansion coefficient with respect to a basis (in particular its phase) can be chosen independently of the remaining ones. The necessary degree of redundancy has been quantified fairly sharply: Generic systems Ψ\Psi consisting of 4d4\geq 4d-4 vectors allow phase retrieval, and for certain dimensions dd, phase retrieval for systems with less than 4d44d-4 vectors is impossible [9].

A problem closely related to phase retrieval is matrix recovery [6, 2], asking whether any matrix Ad×dA\in\mathbb{C}^{d\times d} can be recovered from the family

d×dA(Aψi,ψi)iII\mathbb{C}^{d\times d}\ni A\mapsto(\langle A\psi_{i},\psi_{i}\rangle)_{i\in I}\in\mathbb{C}^{I}

of scalar products. This injectivity property was discussed in [2] by the name of maximal span property, picking up on earlier investigations of the problem under the name of quantum state tomography in physics.

Any system Ψ\Psi doing matrix recovery does phase retrieval. To see this, one only needs to realize that

|f,ψ|2=(ff)(ψ),ψ,|\langle f,\psi\rangle|^{2}=\langle(f\otimes f)(\psi),\psi\rangle~{},

where ffd×df\otimes f\in\mathbb{C}^{d\times d} is given by (ff)(m,n)=f(m)f(n)¯(f\otimes f)(m,n)=f(m)\overline{f(n)}. Hence if Ψ\Psi does matrix recovery, fff\otimes f can be recovered from (|f,ψi|2)iI(|\langle f,\psi_{i}\rangle|^{2})_{i\in I}, and ff can be recovered up to a phase factor from fff\otimes f. An interesting byproduct of this observation is that stability of linear inversion is well-understood and controllable, unlike the non-linear problem of phase recovery. However, there is also a price to pay in terms of redundancy: For dimension reasons, matrix recovery requires at least d2d^{2} measurements.

Phase retrieval and matrix recovery are often studied under additional assumptions on the analyzed signal or matrix, such as sparsity or low rank; see e.g. [10, 18]. Here one does not aim for recovery guarantees for all possible signals, thereby allowing a reduction in the number of measurements. The constructions in this context, as well as the recovery guarantees, are typically of a probabilistic nature.

Overview of the paper

In this paper we are interested in the deterministic construction of structured systems guaranteeing phase retrieval and matrix recovery, more precisely in the construction of group frames with phase retrieval. Group frames are characterized by the fact that they are generated as orbits Ψ=π(G)ψ\Psi=\pi(G)\psi of a (usually unitary) representation π\pi of a group GG acting on a suitably chosen vector ψ\psi; see [24] for group frames over finite groups, and [14] for an extensive discussion of group frames over locally compact groups.

The chief example of group frames allowing phase retrieval is provided by the case where G=G=\mathbb{H} is a Heisenberg group, and π\pi is a Schrödinger representation. This setting provided the earliest results concerning the use of group actions for quantum state tomography [22]. In the setting of Schrödinger representations, the phase retrieval properties of the system π(G)ψ\pi(G)\psi were seen to be closely related to properties of the ambiguity function Aψ:xψ,π(x)ψA_{\psi}:\mathbb{H}\ni x\mapsto\langle\psi,\pi(x)\psi\rangle (see [16]). It was then observed in [5] (or rather, rediscovered, in the light of [22]) that for finite Heisenberg groups, proper assumptions on AψA_{\psi} even allow to solve the matrix recovery problem. A further investigation of this property for more general group frames was then provided in [19, 8], with emphasis on irreducible projective representations of finite abelian groups. A quite far-reaching extension of these results to irreducible representations of general nilpotent groups was recently obtained in [15]. These developments notwithstanding, the problem of phase retrieval for group frames is far from being fully understood.

In this paper we focus on permutation representations, and in particular on actions of affine groups pp\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{*} of finite fields p\mathbb{Z}_{p} (with pp prime). Our main results, Theorem 6.5, Corollary 6.6 and Theorem 6.7, establish criteria for generating vectors guaranteeing matrix recovery, provide an explicit recovery formula, and simple concrete examples of generating vectors fulfilling these criteria. We also present a general recipe how to tackle the problem of matrix recovery for group frames (see Remark 6.8), and explain how the known instances (finite Heisenberg groups and affine groups) fit in this general framework (Remark 6.9). Remarks 6.11 and 6.12 exhibit some connections of our results to intriguing questions in finite Fourier analysis.

To put these results in perspective, we first point out that the retrieval problem for finite affine groups can be understood as a finite analog of the wavelet phase retrieval problem previously studied by Mallat and Waldspurger [21]. At this point in time it is safe to say that explicit constructions of systems for matrix recovery, with linear reconstruction formulae, are only known for very specific systems of vectors, such as equiangular tight frames, or unions of d+1d+1 mutually unbiased bases in dd-dimensional Hilbert spaces, see [25, 2]. It is considered highly doubtful that every dimension dd accommodates d+1d+1 mutually unbiased bases. Known explicit constructions of such systems exist only in very restricted dimensions, e.g. for dd equal to a prime power. By contrast to this, our representation theoretic construction of a system doing matrix recovery via the affine group of the field p\mathbb{Z}_{p} (pp prime) works in a space of dimension d=p1d=p-1, yielding a system of d(d+1)d(d+1) vectors. Hence, while our argument does not rely on the notion of mutually unbiased bases, it is intriguing to note that the cardinalities of the resulting group frames obey the same dimension-dependent relationship.

Throughout the paper, we also discuss variants of phase retrieval, such as sign and conjugate phase retrieval, and extensions to more general permutation representations. We present several results and examples of independent interest, mostly for the purpose of highlighting the challenges that the representation-theoretic understanding of phase retrieval properties faces.

2 Phase retrieval for group frames

Throughout the paper, we let 𝕂{,}\mathbb{K}\in\{\mathbb{R},\mathbb{C}\}. Given finite sets X,YX,Y, we let 𝕂X×Y\mathbb{K}^{X\times Y} denote the space of matrices A=(A(x,y))xX,yYA=(A(x,y))_{x\in X,y\in Y}, with entries in 𝕂\mathbb{K}. We identify matrices with the induced linear maps 𝕂Y𝕂X\mathbb{K}^{Y}\to\mathbb{K}^{X} if needed. We endow the matrix space 𝕂X×Y\mathbb{K}^{X\times Y} with the usual scalar product

A,B=trace(AB).\langle A,B\rangle={\rm trace}(AB^{*})~{}.

We will use the convention 𝕂d=𝕂{0,,d1}\mathbb{K}^{d}=\mathbb{K}^{\{0,\ldots,d-1\}} for dd\in\mathbb{N}. Special vectors in 𝕂d\mathbb{K}^{d} are given by 𝟏=(1,,1)T\mathbf{1}=(1,\ldots,1)^{T} and δk𝕂d\delta_{k}\in\mathbb{K}^{d}, δk(m)=1\delta_{k}(m)=1 for k=mk=m and 0 elsewhere.

We now formally introduce a variety of retrieval properties. Sign and phase retrieval for frames were introduced by Balan et.al. [3]. The intermediate notion of conjugate phase retrieval is due to Evans and Lai [11]. Matrix recovery and associated linear reconstruction methods were introduced to this discussion in [2]; see also [6].

Definition 2.1

Let \mathcal{H} denote a finite-dimensional real or complex Hilbert space. Given a system Ψ=(ψi)iI\Psi=(\psi_{i})_{i\in I}\subset\mathcal{H}, we define the associated coefficient transform as

VΨ:I,f(f,ψi)iI.V_{\Psi}:\mathcal{H}\to\mathbb{C}^{I}~{},~{}f\mapsto(\langle f,\psi_{i}\rangle)_{i\in I}~{}.

Recall that VΨV_{\Psi} is injective iff (ψi)iI(\psi_{i})_{i\in I} is total. Given a complex Hilbert space \mathcal{H}, we say that the system Ψ\Psi\subset\mathcal{H} does phase retrieval if the map

𝒜Ψ:/𝕋f|VΨf|I\mathcal{A}_{\Psi}:\mathcal{H}/\mathbb{T}\ni f\mapsto|V_{\Psi}f|\in\mathbb{R}^{I}

is one-to-one. Here /𝕋\mathcal{H}/\mathbb{T} denotes the orbit space under the canonical action of 𝕋\mathbb{T} on \mathcal{H}, and |VΨf||V_{\Psi}f| is obtained by pointwise application of the modulus function to VΨfV_{\Psi}f.

Given a real Hilbert space \mathcal{H}, and Ψ=(ψi)iI\Psi=(\psi_{i})_{i\in I}\subset\mathcal{H}, we define the sign retrieval property by requiring that

𝒜Ψ:/{±1}I,f|VΨf|\mathcal{A}_{\Psi}:\mathcal{H}/\{\pm 1\}\to\mathbb{R}^{I}~{},~{}f\mapsto|V_{\Psi}f|

is one-to-one.

We define the intermediate property of conjugate phase retrieval as follows: We assume that X\mathcal{H}\subset\mathbb{C}^{X} is a complex subspace, where XX is finite, and X\mathbb{C}^{X} is endowed with the usual scalar product. Furthermore, we assume ΨX\Psi\subset\mathbb{R}^{X}\cap\mathcal{H}. Then Ψ\Psi does conjugate phase retrieval if the equation 𝒜Ψ(f)=𝒜Ψ(g)\mathcal{A}_{\Psi}(f)=\mathcal{A}_{\Psi}(g) is equivalent to the existence of α𝕋\alpha\in\mathbb{T} such that either f=αgf=\alpha g or f=αg¯f=\alpha\overline{g} holds, where complex conjugation is understood to be applied entrywise.

Finally, we say that the system Ψ\Psi\subset\mathcal{H} does matrix recovery if the map

()AVΨΨAI,VΨΨA(i)=Aψi,ψi\mathcal{L}(\mathcal{H})\ni A\mapsto V_{\Psi\otimes\Psi}A\in\mathbb{C}^{I},~{}V_{\Psi\otimes\Psi}A(i)=\langle A\psi_{i},\psi_{i}\rangle

is one-to-one. Here ()\mathcal{L}(\mathcal{H}) denotes the space of linear mappings \mathcal{H}\to\mathcal{H}.

Remark 2.2

Consider the following statements, for a system Ψ=(ψi)iIX\Psi=(\psi_{i})_{i\in I}\subset\mathbb{C}^{X}:

  1. 1.)

    Ψ\Psi does matrix recovery.

  2. 2.)

    Ψ\Psi does phase retrieval.

  3. 3.)

    Ψ\Psi does conjugate phase retrieval.

  4. 4.)

    Ψ\Psi does sign retrieval.

  5. 5.)

    Ψ\Psi spans X\mathbb{C}^{X}.

Then one has the implications 1.)2.)5.)1.)\Rightarrow 2.)\Rightarrow 5.), as well as 3.)4.)5.)3.)\Rightarrow 4.)\Rightarrow 5.). Observe that conditions 3.), 4.) refer to real-valued systems, which never do phase retrieval, due to the equation f,ψ¯=f¯,ψ\overline{\langle f,\psi\rangle}=\langle\overline{f},\psi\rangle holding for real-valued ψ\psi.

Definition 2.3

Let π\pi denote a representation of a finite group GG on the finite-dimensional Hilbert space π\mathcal{H}_{\pi} over 𝕂\mathbb{K}. Given ψπ\psi\in\mathcal{H}_{\pi}, the associated group frame is given by Ψ=π(G)ψ=(π(x)ψ)xG\Psi=\pi(G)\psi=(\pi(x)\psi)_{x\in G}. We call ψ\psi the generating vector of the system Ψ\Psi.

Given ψ\psi, we write

Vψ:πG,Vψf(x)=f,π(x)ψ.V_{\psi}:\mathcal{H}_{\pi}\to\mathbb{C}^{G}~{},~{}V_{\psi}f(x)=\langle f,\pi(x)\psi\rangle~{}.

In case we want to emphasize the dependence on the representation π\pi, we will also use Vψπ=VψV_{\psi}^{\pi}=V_{\psi}.

We say that π\pi does phase/conjugate phase/sign retrieval if there exists ψ\psi\in\mathcal{H} such that Ψ=(π(x)ψ)xG\Psi=(\pi(x)\psi)_{x\in G} does phase/conjugate phase/sign retrieval.

In the case of conjugate phase retrieval, we additionally assume that π=X\mathcal{H}_{\pi}=\mathbb{C}^{X} for suitable XX, and that π(G)(X)X\pi(G)(\mathbb{R}^{X})\subset\mathbb{R}^{X} holds.

Similarly, π\pi does matrix recovery if there exists ψ\psi\in\mathcal{H} such that Ψ=(π(x)ψ)xG\Psi=(\pi(x)\psi)_{x\in G} does matrix recovery.

The following observation reformulates matrix recovery for group frames. Recall that a vector ψπ\psi\in\mathcal{H}_{\pi} is called cyclic for π\pi if VψπV_{\psi}^{\pi} is one-to-one. With this definition in mind, the proposition is seen to be a mere reformulation.

Proposition 2.4

Let π\pi denote a representation of a group GG acting on X\mathbb{C}^{X}.

Let by ϱ=(ππ¯)|ΔG\varrho=(\pi\otimes\overline{\pi})|_{\Delta_{G}}, i.e. the representation of GG acting on X×X\mathbb{C}^{X\times X} by

ϱ(x)A=π(x)Aπ(x).\varrho(x)A=\pi(x)\cdot A\cdot\pi(x)^{*}~{}.

Then ψ\psi\in\mathcal{H} does matrix recovery iff ψψ\psi\otimes\psi is a cyclic vector for ϱ\varrho.

We next prove a basic feature of the sets of vectors ψ\psi that guarantee one of the retrieval properties. In the following lemma, a set U𝕂dU\subset\mathbb{K}^{d} is called real Zariski open (by slight abuse of terminology) if there exist finitely many functions f1,,fkf_{1},\ldots,f_{k} on 𝕂d\mathbb{K}^{d} that are polynomials in the real and imaginary parts of the coordinates such that

U={x𝕂d: 1ik:fi(x)0}.U=\{x\in\mathbb{K}^{d}:\exists\text{ }1\leq i\leq k~{}:~{}f_{i}(x)\not=0\}~{}.

The terminology naturally extends to subset of 𝕂d×d\mathbb{K}^{d\times d}.

Lemma 2.5

Assume that π\pi does sign/conjugate phase/phase/matrix recovery. Then the set of vectors ψπ\psi\in\mathcal{H}_{\pi} that guarantee this property is real Zariski open.

Proof 1

We identify a system Ψ=(ψi)i=1m\Psi=(\psi_{i})_{i=1}^{m} of vectors in 𝕂d\mathbb{K}^{d} with the matrix Ψ𝕂d×m\Psi\in\mathbb{K}^{d\times m}. Then the set of all Ψ\Psi that perform phase or sign retrieval is real Zariski open in 𝕂d×m\mathbb{K}^{d\times m}, by [9]. The analogous observation has been made for conjugate phase retrieval in [11].

In order to apply these observations to group frames, it is enough to note that the map ψVψ\psi\mapsto V_{\psi} is conjugate-linear, and that the preimage of a real Zariski open set under such a map is real Zariski open.

In the case of matrix recovery, the map ψVψψϱ\psi\mapsto V_{\psi\otimes\psi}^{\varrho} is polynomial, and by definition, ψ\psi does matrix recovery iff the linear map VψψϱV^{\varrho}_{\psi\otimes\psi} has rank dπ2d_{\pi}^{2}. However, the set of maximal rank matrices is real Zariski open, being defined by the nonvanishing of at least one dπ2×dπ2d_{\pi}^{2}\times d_{\pi}^{2}-subdeterminant. Hence the set of vectors ψ\psi guaranteeing matrix recovery is the polynomial preimage of a real Zariski open set, hence real Zariski open.

Note that real Zariski open sets are open in the standard topology, and either empty or of full Lebesgue measure. This shows that once a retrieval property is established for π\pi, almost every vector ψ\psi guarantees it. However, verifying either that the real Zariski open set is nonempty, or that a given concrete vector is contained in it, remains a considerable challenge, and the results in the present papers bear witness to this fact.

Remark 2.6

So far, phase retrieval for group frames Ψ=π(G)ψ\Psi=\pi(G)\psi has been studied almost exclusively for Schrödinger representations of (finite or infinite) Heisenberg groups. An extension of these results to projective representations of abelian groups was performed in [19]. Phase retrieval for the continuous wavelet transform in dimension one was studied in [21, 1]. Here the underlying representation is the quasiregular representation of the affine group \mathbb{R}\rtimes\mathbb{R}^{*}, and it is fair to say that the phase retrieval problem even for this basic representation is far from being completely understood. Our paper provides an analysis for the finite analogs of \mathbb{R}\rtimes\mathbb{R}^{*}.

The implications from Remark 2.2 immediately yield the implications 1.)2.)5.)1.)\Rightarrow 2.)\Rightarrow 5.), and 3.)4.)5.)3.)\Rightarrow 4.)\Rightarrow 5.) between the following statements.

  1. 1.)

    π\pi does matrix recovery.

  2. 2.)

    π\pi does phase retrieval.

  3. 3.)

    π\pi does conjugate phase retrieval.

  4. 4.)

    π\pi does sign retrieval.

  5. 5.)

    π\pi is cyclic.

Unlike in the general case considered in Remark 2.2, the implication 2.)3.)2.)\Rightarrow 3.) does make sense as soon as π=X\mathcal{H}_{\pi}=\mathbb{C}^{X}, and π(G)(X)X\pi(G)(\mathbb{R}^{X})\subset\mathbb{R}^{X} holds. Note that here 2.)2.) asks for the existence of a particular complex-valued generating vector, whereas 3.)3.) requires a suitable real-valued generating vector. The relationship between conditions 2.)2.) and 3.)3.) is currently open.

At this point it is also worth noting that not all properties are invariant under unitary equivalence. While it is easily verified that matrix and phase retrieval of a representation are preserved under unitary equivalence, the property that π(G)(X)X\pi(G)(\mathbb{R}^{X})\subset\mathbb{R}^{X} depends on the realization of the representation. This follows from the fact that most intertwining operators do not commute with taking pointwise conjugates or real parts. A concrete example will be given by the representation π^0\widehat{\pi}_{0} studied more closely in Section 6, which is equivalent to the permutation representation π0\pi_{0}. The permutation representation satisfies π0(G)(X)X\pi_{0}(G)(\mathbb{R}^{X})\subset\mathbb{R}^{X}, but its unitary equivalent π^0\widehat{\pi}_{0} does not.

We finally point out that not all cyclic representations do phase retrieval. As an extreme case in point, we can take the regular representation λG\lambda_{G} of a finite group GG, which is cyclic. Any cyclic vector ψ\psi for λG\lambda_{G} necessarily fulfills VΨ(G)=GV_{\Psi}(\mathbb{C}^{G})=\mathbb{C}^{G} (for dimension reasons), and the map F|F|F\mapsto|F| is clearly not injective on G\mathbb{C}^{G}.

Remark 2.7

Recent research activity has extended the original setup to include projective representations of finite groups GG, see e.g. [19]. Recall that by definition, a (unitary) projective representation of a finite group GG is a map π:G𝒰(π)\pi:G\to\mathcal{U}(\mathcal{H}_{\pi}) such that

x,yG:π(xy)=α(x,y)π(x)π(y),α:G×G𝕋.\forall x,y\in G:\pi(xy)=\alpha(x,y)\pi(x)\pi(y)~{},\alpha:G\times G\to\mathbb{T}~{}.

This extension can be useful for the treatment of special cases, such as the Schrödinger representation of a finite Heisenberg group \mathbb{H}, which can be alternatively viewed as a projective representation of the quotient group /Z()\mathbb{H}/Z(\mathbb{H}), where Z()Z(\mathbb{H}) denotes the center of \mathbb{H}.

However, it is well-known that every projective representation π\pi of a finite group GG can be lifted to a standard unitary representation π\pi^{\prime} of a finite central extension GG^{\prime} of GG, and the system π(G)ψ\pi^{\prime}(G^{\prime})\psi differs from π(G)ψ\pi(G)\psi only by the inclusion of a fixed number of scalar multiples of ww, for every vector wπ(G)ψw\in\pi(G)\psi. None of the retrieval properties discussed here is affected by this change. Hence the extension to projective representations ultimately does not enlarge the realm of available constructions in any truly substantial manner.

3 Setup: Affine groups and permutation representations

Let p>2p>2 denote a prime, and let p=/p\mathbb{Z}_{p}=\mathbb{Z}/p\mathbb{Z}, the associated finite field of order pp. We typically denote p={0,1,,p1}\mathbb{Z}_{p}=\{0,1,\ldots,p-1\}, and understand arithmetic operations on this set to be defined modulo pp. p={0}\mathbb{Z}_{p}^{*}=\mathbb{Z}\setminus\{0\} then denotes the multiplicative group of p\mathbb{Z}_{p}. The affine group of p\mathbb{Z}_{p} is given by the semidirect product G=ppG=\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{\ast}, using the canonical multiplicative action of p\mathbb{Z}_{p}^{\ast} on p\mathbb{Z}_{p}. I.e. as a set, G=p×pG=\mathbb{Z}_{p}\times\mathbb{Z}_{p}^{\ast}, with group law (k,l)(k,l)=(k+lk,ll)(k,l)(k^{\prime},l^{\prime})=(k+lk^{\prime},ll^{\prime}). Unless otherwise specified, the symbol GG will always refer to this semidirect product.

Given nn\in\mathbb{N}, we let S(n)S(n) denote the symmetric group of {0,,n1}\{0,\ldots,n-1\}. Note that since G=ppG=\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{\ast} acts faithfully on p={0,1,,p1}\mathbb{Z}_{p}=\{0,1,\ldots,p-1\} via (k,l).m=k+lm(k,l).m=k+lm, we may regard GG as a subgroup of S(p)S(p). In particular, one has S(3)=33S(3)=\mathbb{Z}_{3}\rtimes\mathbb{Z}_{3}^{\ast}.

The quasiregular representation Π\Pi of S(n)S(n) acts on 𝕂n\mathbb{K}^{n} by

(Π(h)f)(m)=f(h1m).(\Pi(h)f)(m)=f(h^{-1}m)~{}. (3.1)

In the case of n=pn=p a prime number, the restriction of Π\Pi to the subgroup G=ppG=\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{*} will be denoted by π=Π|G\pi=\Pi|_{G}, and it is given explicitly by

(π(k,l)f)(m)=f(l1(mk)).(\pi(k,l)f)(m)=f(l^{-1}(m-k))~{}. (3.2)

Note that our notation does not explicitly distinguish the real and complex cases. In fact, we have Π(S(n))(n)n\Pi(S(n))(\mathbb{R}^{n})\subset\mathbb{R}^{n}, a fact which allows to discuss conjugate phase and sign retrieval for Π\Pi and its restrictions.

Remark 3.1

The semi-direct product pp\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{\ast} and its quasi-regular representation were studied in [13] as a tool for the construction of wavelets on finite fields. In this sense, this paper can also be seen as dealing with a finite field analog of wavelet phase retrieval, see e.g. [21, 1].

We let 𝟏=(1,,1)T𝕂n\mathbf{1}=(1,\ldots,1)^{T}\in\mathbb{K}^{n}. We make the following observation:

Lemma 3.2

The subspaces 1=𝕂𝟏\mathcal{H}_{1}=\mathbb{K}\cdot\mathbf{1} and 0=1\mathcal{H}_{0}=\mathcal{H}_{1}^{\bot} are invariant under Π\Pi, and therefore also under π\pi. The associated restrictions π0,π1\pi_{0},\pi_{1} of π\pi acting on 0,1\mathcal{H}_{0},\mathcal{H}_{1} respectively are irreducible.

In this paper we wish to address the various retrieval properties for π0\pi_{0} and/or π\pi. Ideally, we would like to have explicit and sharp criteria for generating vectors guaranteeing any of the properties, as well as concrete examples. We first note a simple observation regarding π\pi.

Corollary 3.3

The representation π\pi does not do matrix recovery.

Proof 2

Matrix recovery is equivalent to stating that span{π(x)ψπ(x)ψ:xG}=p×p{\rm span}\{\pi(x)\psi\otimes\pi(x)\psi:x\in G\}=\mathbb{C}^{p\times p}. For dimension reasons, this requires p2|G|=p(p1)p^{2}\leq|G|=p(p-1), which is obviously false.

Note that this argument does not apply to the codimension one subspace 0\mathcal{H}_{0}, and we will indeed show that π0\pi_{0} does matrix recovery.

4 Sign retrieval for π0\pi_{0}

In this section we consider the sign retrieval properties of the permutation representation π0\pi_{0} acting on the real vector space 0={xp:x,𝟏=0}\mathcal{H}_{0}=\{x\in\mathbb{R}^{p}~{}:~{}\langle x,\mathbf{1}\rangle=0\}.

Definition 4.1

  1. 1.

    Let \mathcal{H} denote a finite-dimensional Hilbert space, and (ψi)iI(\psi_{i})_{i\in I}\subset\mathcal{H}. Then (ψi)iI(\psi_{i})_{i\in I} has the complement property iff for all SIS\subset I, either span(ψi)iS={\rm span}(\psi_{i})_{i\in S}=\mathcal{H}, or span(ψi)iSI={\rm span}(\psi_{i})_{i\in S\setminus I}=\mathcal{H}.

  2. 2.

    The family has full spark if dim()=d{\rm dim}(\mathcal{H})=d\in\mathbb{N}, and for every subset SIS\subset I of cardinality dd, the system (ψi)iS(\psi_{i})_{i\in S} spans \mathcal{H}.

The following result are contained in [4]; see Theorems 3 and 7.

Theorem 4.2

Let Ψ=(ψi)iI\Psi=(\psi_{i})_{i\in I}\subset\mathcal{H} denote a system of vectors contained in the finite-dimensional Hilbert space \mathcal{H}.

  1. (a)

    If Ψ\Psi does phase or sign retrieval, then Ψ\Psi has the complement property.

  2. (b)

    If \mathcal{H} is a real Hilbert space, then Ψ\Psi does sign retrieval iff it has the complement property.

The following theorem is implied by Theorem 10 in [20].

Theorem 4.3

The quasiregular representation π0\pi_{0} on 0\mathcal{H}_{0} has full spark, which means that there exists a real-valued vector ψ0\psi\in\mathcal{H}_{0} such that (π(x)ψ)xG(\pi(x)\psi)_{x\in G} has full spark. Any such vector does sign retrieval.

Proof 3

The first statement is Theorem 10 in [20]; a closer inspection of its proof shows that a vector whose orbit is a full spark frame can indeed be chosen to be real-valued. Finally, since GG has p(p1)>2p1p(p-1)>2p-1 elements, every full spark frame indexed by GG necessarily has the complement property. Hence, Theorem 4.2 yields sign retrieval.

5 Conjugate phase retrieval for doubly transitive permutation groups

From now on we let 0=𝟏p\mathcal{H}_{0}=\langle\mathbf{1}\rangle\subset\mathbb{C}^{p}. Throughout this section, we may assume that ΓS(n)\Gamma\subset S(n) is a subgroup of the symmetric group. Additionally, we also assume that n3n\geq 3 to avoid trivialities. Recall next, that Γ\Gamma is called kk-fold transitive if for every pair of kk-tuples u,v{0,,n1}ku,v\in\{0,\ldots,n-1\}^{k}, where both uu and vv have pairwise distinct entries, there exists hΓh\in\Gamma such that

i=1,,k:vi=uh(i).\forall i=1,\ldots,k~{}:~{}v_{i}=u_{h(i)}~{}.

22-fold transitive groups are also called doubly transitive. An important class of doubly transitive groups is given by G=ppG=\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{\ast}, for p3p\geq 3.

We intend to prove that 22-fold transitive groups do conjugate phase retrieval. The argument will be based on a particular choice of generating vector, and the following, well-known property of the complex plane.

Remark 5.1

Let (yk)k=0,,n1(y_{k})_{k=0,\ldots,n-1}, (zk)k=0,,n1n(z_{k})_{k=0,\ldots,n-1}\in\mathbb{C}^{n} satisfy

l,k=0,,n1:|ylyk|=|zlzk|.\forall l,k=0,\ldots,n-1~{}:~{}|y_{l}-y_{k}|=|z_{l}-z_{k}|~{}. (5.1)

Then there exists an isometry φ:\varphi:\mathbb{C}\to\mathbb{C} with φ(yl)=zl\varphi(y_{l})=z_{l}, for all l=0,,n1l=0,\ldots,n-1.

Furthermore, the group of isometries of the Euclidean space 2\mathbb{C}\cong\mathbb{R}^{2} coincides with the Euclidean motion group, i.e. it is the composition of a translation and an orthogonal matrix. In addition, any orthogonal mapping on 2\mathbb{R}^{2} is given by (complex) multiplication with an element of 𝕋\mathbb{T}, possibly followed by conjugation.

In short, we have deduced from (5.1) that either

l=0,,n1:zl=αyl+w\forall l=0,\ldots,n-1~{}:~{}z_{l}=\alpha y_{l}+w

or

l=0,,n1:zl=αyl¯+w,\forall l=0,\ldots,n-1~{}:~{}z_{l}=\alpha\overline{y_{l}}+w~{},

holds, with α𝕋\alpha\in\mathbb{T} and ww\in\mathbb{C}.

We then have the following observation, which implies conjugate phase retrieval for π0\pi_{0}.

Theorem 5.2

Let Γ<S(n)\Gamma<S(n) denote a doubly transitive subgroup. Pick k0,l0{0,,n1}k_{0},l_{0}\in\{0,\ldots,n-1\} with k0l0k_{0}\not=l_{0}, and let ψ=δk0δl0\psi=\delta_{k_{0}}-\delta_{l_{0}}. Then Π(Γ)ψ\Pi(\Gamma)\psi does conjugate phase retrieval on the invariant subspace 0\mathcal{H}_{0}.

Proof 4

Let (k,l)(k,l) denote any distinct pair of elements, and let hΓh\in\Gamma such that h(k0)=k,h(l0)=lh(k_{0})=k,h(l_{0})=l. Then Vψf(h)=f(k)f(l)V_{\psi}f(h)=f(k)-f(l) and this shows for all f,g0f,g\in\mathcal{H}_{0}:

|Vψf|=|Vψg|(k,l){0,,n1}2:|f(k)f(l)|=|g(k)g(l)|.|V_{\psi}f|=|V_{\psi}g|\Longleftrightarrow\forall(k,l)\in\{0,\ldots,n-1\}^{2}:|f(k)-f(l)|=|g(k)-g(l)|~{}.

An application of Remark 5.1 yields either g(l)=αf(l)+wg(l)=\alpha f(l)+w or g(l)=αg(l)¯+wg(l)=\alpha\overline{g(l)}+w holds for all ll, with α𝕋\alpha\in\mathbb{T} and ww\in\mathbb{C}.

To complete the proof, it remains to show w=0w=0, but this follows (for the case without conjugation) from f,g0f,g\in\mathcal{H}_{0} via

0=l=0n1g(l)=l=0n1αf(l)+w=nw+αl=0n1f(l)=nw.0=\sum_{l=0}^{n-1}g(l)=\sum_{l=0}^{n-1}\alpha f(l)+w=nw+\alpha\sum_{l=0}^{n-1}f(l)=nw~{}.

The second case, involving complex conjugation, is treated entirely analogously. This establishes conjugate phase retrieval.

Remark 5.3

We now turn to the question of conjugate phase retrieval for Π\Pi on n\mathbb{C}^{n}, using the vector

ψ1=δk0δl0ψ0+n1/2𝟏.\psi_{1}=\underbrace{\delta_{k_{0}}-\delta_{l_{0}}}_{\psi_{0}}+n^{-1/2}\mathbf{1}~{}.

Note that the projection of ψ1\psi_{1} onto the subspace 0\mathcal{H}_{0} does conjugate phase retrieval for the restriction of π\pi to that subspace, and the same can be said (for trivial resasons) of the orthogonal complement of 0\mathcal{H}_{0}, the one-dimensional space 𝟏\mathbb{C}\cdot\mathbf{1}. Yet these properties are generally not enough to guarantee conjugate phase retrieval for the full space n\mathbb{C}^{n}, as the following example for n=3n=3 shows:

Let ξ=e2πi/3\xi=e^{2\pi i/3}, and define y=(y0,y1,y2)3y=(y_{0},y_{1},y_{2})\in\mathbb{C}^{3} by letting

yl=ξl.y_{l}=\xi^{l}~{}.

We then have kyk=0\sum_{k}y_{k}=0, and one easily verifies, for all xS(3)x\in S(3),

|y,Π(x)ψ1|=|1ξ|=||1ξ|3𝟏,Π(x)ψ1|.|\langle y,\Pi(x)\psi_{1}\rangle|=|1-\xi|=|\langle\frac{|1-\xi|}{\sqrt{3}}\mathbf{1},\Pi(x)\psi_{1}\rangle|~{}.

On the other hand, the vectors w=|1ξ|3𝟏w=\frac{|1-\xi|}{\sqrt{3}}\mathbf{1} and y=(y0,y1,y2)y=(y_{0},y_{1},y_{2}) are clearly distinct, even orthogonal.

Note that this example does not yet establish that Π\Pi does not do conjugate phase retrieval on 3\mathbb{C}^{3}, as it only applies to a particular choice of generating vector ψ1\psi_{1}. But it does illustrate that the problem of conjugate phase retrieval on all of 3\mathbb{C}^{3} cannot be addressed by being able to solve it on each invariant subspace 0\mathcal{H}_{0},1\mathcal{H}_{1} separately, somewhat contrary to common (linear) representation theoretic wisdom. This phenomenon emphasizes the nonlinear nature of the conjugate phase retrieval problem.

We do not know whether analogs of the counterexample constructed here are available in dimensions n>3n>3.

As a further application of the basic proof principle exploited in this section, we have the following observation:

Proposition 5.4

Let n3n\geq 3, and let Γ<S(n)\Gamma<S(n) denote a doubly transitive subgroup. Pick k0,l0{0,,n1}k_{0},l_{0}\in\{0,\ldots,n-1\} with k0l0k_{0}\not=l_{0}, and let ψ1=δk0δl0+n1/2𝟏\psi_{1}=\delta_{k_{0}}-\delta_{l_{0}}+n^{-1/2}\mathbf{1}. Then Π(Γ)ψ1\Pi(\Gamma)\psi_{1} does sign retrieval on n\mathbb{R}^{n}.

Proof 5

Assume f=f0+an1/2𝟏f=f_{0}+a\cdot n^{-1/2}\mathbf{1}, g=g0+bn1/2𝟏g=g_{0}+b\cdot n^{-1/2}\mathbf{1} with f0,g0n0f_{0},g_{0}\in\mathbb{R}^{n}\cap\mathcal{H}_{0} and a,ba,b\in\mathbb{R}, such that

xΓ:|f,Π(x)ψ1|=|g,Π(x)ψ1|.\forall x\in\Gamma~{}:~{}|\langle f,\Pi(x)\psi_{1}\rangle|=|\langle g,\Pi(x)\psi_{1}|~{}.

Since Γ\Gamma is doubly transitive, this is equivalent to

lk:|f(k)f(l)|2+a2=|g(k)g(l)|2+b2,a(f(k)f(l))=b(g(k)g(l)).\forall l\not=k~{}:~{}|f(k)-f(l)|^{2}+a^{2}=|g(k)-g(l)|^{2}+b^{2}~{},~{}a(f(k)-f(l))=b(g(k)-g(l))~{}. (5.2)

Assuming that ff is constant then yields either b=0b=0, or that gg is constant. The latter implies |a|=|b||a|=|b|, and g=±fg=\pm f, as desired (since f0=g0=0f_{0}=g_{0}=0). If b=0b=0, we get

|g(k)g(l)|=|a|,lk.|g(k)-g(l)|=|a|~{},~{}l\not=k~{}.

However, there do not exist n3n\geq 3 equidistant points on the real line, hence this case can be excluded.

Hence it remains to address the case that ff is not constant. Assuming a=0a=0 then requires that b(g(k)g(l))=0b(g(k)-g(l))=0. If b=0b=0, we obtain

lk:|f(k)f(l)|=|g(k)g(l)|\forall l\not=k~{}:~{}|f(k)-f(l)|=|g(k)-g(l)|

and thus g=±fg=\pm f by Theorem 5.2. If b0b\not=0, it follows that gg is constant, and thus

lk:|f(k)f(l)|2=b2a2\forall l\not=k~{}:~{}|f(k)-f(l)|^{2}=b^{2}-a^{2}

and since ff is not constant, b2a2>0b^{2}-a^{2}>0 follows. As above, the fact that there are no n3n\geq 3 equidistant points in \mathbb{R} excludes this possibility.

The final case to consider is therefore that ff is not constant and a0a\not=0. This implies that

lk:g(k)g(l)=ba(f(k)f(l)).\forall l\not=k~{}:~{}g(k)-g(l)=\frac{b}{a}(f(k)-f(l))~{}.

Plugging this into the first equation of (5.2) and simplifying yields

lk:|f(k)f(l)|2=b2a21b2/a2=a2,\forall l\not=k~{}:~{}|f(k)-f(l)|^{2}=\frac{b^{2}-a^{2}}{1-b^{2}/a^{2}}=-a^{2}~{},

which contradicts our assumptions on aa and ff.

6 Matrix recovery for π0\pi_{0}

In this section we establish the matrix recovery property for π0\pi_{0}. Recall that we are interested in finding specific cyclic vectors for the representation ϱ0\varrho_{0} obtained by conjugating with π0\pi_{0}. The basic strategy of the following is to first decompose ϱ0\varrho_{0} into suitable subrepresentations, with the help of an explicitly derived intertwining operator; see Proposition 6.4. This decomposition will provide fairly explicit access to criteria for cyclic vectors. The remaining task is to apply the intertwining operator from Proposition 6.4 to matrices of the type A=φφA=\varphi\otimes\varphi, and determine vectors φ\varphi that fulfill the criteria for cyclic vectors. As we will see, the quadratic dependence of AA on φ\varphi makes this final step a further nontrivial obstacle.

The starting point in this approach is to switch to a unitarily equivalent representation π^0\widehat{\pi}_{0}, obtained by conjugating π0\pi_{0} with the Fourier transform on p\mathbb{C}^{p}. In the following, we will denote elements of p={0,,p1}\mathbb{C}^{p}=\mathbb{C}^{\{0,\ldots,p-1\}} as f=(f(0),,f(p1))f=(f(0),\ldots,f(p-1)). With this notation, the Fourier transform is given by

:pp,(f)(m)=p1/2n=0p1f(n)e2πinm/p.\mathcal{F}:\mathbb{C}^{p}\to\mathbb{C}^{p}~{},~{}\mathcal{F}(f)(m)=p^{-1/2}\sum_{n=0}^{p-1}f(n)e^{-2\pi inm/p}~{}. (6.1)

We will also use the notation f^=(f)\widehat{f}=\mathcal{F}(f). With the normalization used here, the Plancherel theorem becomes f^2=f2\|\widehat{f}\|^{2}=\|f\|^{2}, whereas the convolution theorem for the convolution product

(fg)(m)=n=0p1f(n)g(mn).(f\ast g)(m)=\sum_{n=0}^{p-1}f(n)g(m-n)~{}.

reads as

(fg)(k)=p1/2f^(k)g^(k).(f\ast g)^{\wedge}(k)=p^{1/2}\widehat{f}(k)\widehat{g}(k)~{}.

The chief advantage of the Fourier transform is that it diagonalizes the translation action of the affine group. The effect of conjugation with \mathcal{F} is spelled out in the following lemma.

Lemma 6.1

Define π^:G𝒰(p)\widehat{\pi}:G\to\mathcal{U}(\mathbb{C}^{p}), π^(k,l)=π(k,l).\widehat{\pi}(k,l)=\mathcal{F}\pi(k,l)\mathcal{F}^{*}.

  1. (a)

    π^(k,l)f(m)=e2πikm/pf(lm)\widehat{\pi}(k,l)f(m)=e^{-2\pi ikm/p}f(lm), with (1)=δ0\mathcal{F}(\mathcal{H}_{1})=\mathbb{C}\cdot\delta_{0}, (0)={0}×{1,,p1}{1,,p1}\mathcal{F}(\mathcal{H}_{0})=\{0\}\times\mathbb{C}^{\{1,\ldots,p-1\}}\cong\mathbb{C}^{\{1,\ldots,p-1\}}.

  2. (b)

    Let ϱ1\varrho_{1} denote the conjugation action of π^\widehat{\pi} on {1,,p1}×{1,,p1}\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}. Then one has, for all A{1,,p1}×{1,,p1}A\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}

    (ϱ1(k,l)A)(m,n)=e2πik(mn)/pA(lm,ln).\left(\varrho_{1}(k,l)A\right)(m,n)=e^{-2\pi ik(m-n)/p}A(lm,ln)~{}.
Proof 6

We include the computations proving the lemma for completeness. For part (a), we have for all fpf\in\mathbb{C}^{p} that

(π(k,l)f)(m)\displaystyle\mathcal{F}(\pi(k,l)f)(m) =\displaystyle= n=0p1f(l1(nk))e2πinm/p\displaystyle\sum_{n=0}^{p-1}f(l^{-1}(n-k))e^{-2\pi inm/p}
=\displaystyle= n=0p1f(n)e2πi(ln+k)m/p\displaystyle\sum_{n=0}^{p-1}f(n)e^{-2\pi i(ln+k)m/p}
=\displaystyle= e2πikm/p(f)(lm).\displaystyle e^{-2\pi ikm/p}\mathcal{F}(f)(lm)~{}.

The statements concerning 0,1\mathcal{H}_{0},\mathcal{H}_{1} are obvious.

For the proof of (b), we fix A{1,,p1}×{1,,p1}A\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}} as well as fpf\in\mathbb{C}^{p}, and get by (a)

(Aπ^(k,l)f)(r)\displaystyle(A\cdot\widehat{\pi}(k,l)^{*}\cdot f)(r) =\displaystyle= ((Aπ(l1k,l1)f)(r)\displaystyle((A\cdot\pi(-l^{-1}k,l^{-1})\cdot f)(r)
=\displaystyle= n=1p1A(r,n)f(l1n)e2πi(l1k)n/p\displaystyle\sum_{n=1}^{p-1}A(r,n)f(l^{-1}n)e^{-2\pi i(-l^{-1}k)n/p}
=\displaystyle= n=1p1A(r,ln)f(n)e2πikn/p.\displaystyle\sum_{n=1}^{p-1}A(r,ln)f(n)e^{2\pi ikn/p}~{}.

It follows that

(π(k,l)Aπ(k,l)u)(m)\displaystyle(\pi(k,l)\cdot A\cdot\pi(k,l)^{*}\cdot u)(m) =\displaystyle= e2πikm/pn=0p1A(lm,ln)e2πikn/pu(n)\displaystyle e^{-2\pi ikm/p}\sum_{n=0}^{p-1}A(lm,ln)e^{2\pi ikn/p}u(n)
=\displaystyle= n=0p1e2πik(mn)A(lm,ln)u(n),\displaystyle\sum_{n=0}^{p-1}e^{-2\pi ik(m-n)}A(lm,ln)u(n)~{},

which establishes (b).

Remark 6.2

Our explicit matrix recovery algorithm in Corollary 6.6 makes use of the group Fourier transform of GG. In the following, we will briefly sketch the fundamentals of this transform, for a general finite group GG, and a concrete choice of unitary dual G^=((σ,σ))σG^\widehat{G}=\left((\sigma,\mathcal{H}_{\sigma})\right)_{\sigma\in\widehat{G}} of GG. By definition, the unitary dual is a system of representatives of the unitary equivalence classes of irreducible representations of GG. Such a system exists for every finite group GG, and it is necessarily finite. For ease of notation, we assume σ=dσ\mathcal{H}_{\sigma}=\mathbb{C}^{d_{\sigma}}, i.e., σ(x)dσ×dσ\sigma(x)\in\mathbb{C}^{d_{\sigma}\times d_{\sigma}}.

Given FGF\in G^{\mathbb{C}}, one can now define its group Fourier transform as the matrix-valued tuple

G(F)=(σ(F))σG^σG^dσ×dσ\mathcal{F}_{G}(F)=(\sigma(F))_{\sigma\in\widehat{G}}\in\prod_{\sigma\in\widehat{G}}\mathbb{C}^{d_{\sigma}\times d_{\sigma}}

given by

G(F)(σ)=σ(F)=gGF(g)σ(g).\mathcal{F}_{G}(F)(\sigma)=\sigma(F)=\sum_{g\in G}F(g)\sigma(g)~{}. (6.2)

Observe that the Fourier transform (6.1) for the cyclic case is a special case of this construction, since the characters of the additive group p\mathbb{Z}_{p} are precisely given by

χk(l)=e2πilk/p.\chi_{k}(l)=e^{-2\pi ilk/p}~{}.

Note however that \mathcal{F} in equation (6.1) is normalized by the factor p1/2p^{-1/2}, in order to obtain a unitary map, whereas we refrained from introducing a normalization to the group Fourier transform. In order to avoid confusion in the following, we will mostly stick to the notation σ(F)\sigma(F) instead of G(F)(σ)\mathcal{F}_{G}(F)(\sigma), when we refer to the group Fourier transform.

The Fourier transform fulfills the Plancherel formula

F22=|G|1σG^dσσ(F)2\|F\|_{2}^{2}=|G|^{-1}\sum_{\sigma\in\widehat{G}}d_{\sigma}\|\sigma(F)\|^{2}

where the norm on the right hand side is the Frobenius norm. Its inverse is given explicitly by

G1(Aσ)(g)=|G|1σG^dσtrace(Aσσ(g)),\mathcal{F}_{G}^{-1}(A_{\sigma})(g)=|G|^{-1}\sum_{\sigma\in\widehat{G}}d_{\sigma}{\rm trace}(A_{\sigma}\sigma(g)^{*})~{}, (6.3)

for all (Aσ)σG^σG^dσ×dσ(A_{\sigma})_{\sigma\in\widehat{G}}\in\prod_{\sigma\in\widehat{G}}\mathbb{C}^{d_{\sigma}\times d_{\sigma}}. For all these general facts, we refer the reader to [23, Chapter 15].

Returning to the special case of the affine group G=ppG=\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{\ast}, we have the following description of its unitary dual. We refer to [23, Chapter 17] for a proof.

Lemma 6.3

The unitary dual of G=ppG=\mathbb{Z}_{p}\rtimes\mathbb{Z}_{p}^{\ast} is given by the following set of representatives: The representation π^0\widehat{\pi}_{0}, as well as the set {χ~:χp^}\{\tilde{\chi}:\chi\in\widehat{\mathbb{Z}_{p}^{*}}\} of characters of the multiplicative group, lifted to the full group GG by letting

χ~(k,l):=χ(l),χp^.\tilde{\chi}(k,l):=\chi(l)~{},\chi\in\widehat{\mathbb{Z}_{p}^{*}}~{}.

We thus obtain the following group Fourier transform for the affine group GG: Given FGF\in\mathbb{C}^{G}, its group Fourier transform is the tuple

(F)=(χ~(F))χp^(π^0(F)).\mathcal{F}(F)=\left(\tilde{\chi}(F)\right)_{\chi\in\widehat{\mathbb{Z}_{p}^{\ast}}}\cup(\widehat{\pi}_{0}(F)).

consisting of the scalars

χ~(F)=p(kpF(k,l))χ(l),\tilde{\chi}(F)=\sum_{\ell\in\mathbb{Z}_{p}^{\ast}}\left(\sum_{k\in\mathbb{Z}_{p}}F(k,l)\right)\chi(l)~{},

and the matrix

π^0(F)=k,lF(k,l)π^0(k,l){1,,p1}×{1,,p1}\widehat{\pi}_{0}(F)=\sum_{k,l}F(k,l)\widehat{\pi}_{0}(k,l)\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}

The next result provides a decomposition of ϱ1\varrho_{1} into suitable subrepresentations, which is an important step towards establishing matrix recovery. We remind the reader that all algebraic operations involving numbers l,k,m,n{0,,p1}l,k,m,n\in\{0,\ldots,p-1\} are understood modulo pp.

Proposition 6.4

Define the unitary operator

S:{1,,p1}×{1,,p1}{1,,p1}×{1,,p1}S:\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}\to\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}

by

(SA)(m,n)={A(m,m) for n=1A(m(1n)1,mn(1n)1) for n{2,,p1}(SA)(m,n)=\left\{\begin{array}[]{ll}A(-m,-m)&\mbox{ for }n=1\\ A(m(1-n)^{-1},mn(1-n)^{-1})&\mbox{ for }n\in\{2,\ldots,p-1\}\end{array}\right.

Define ϱ2:G𝒰({1,,p1}×{1,,p1})\varrho_{2}:G\to\mathcal{U}(\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}) by ϱ2(k,l)=Sϱ1(k,l)S\varrho_{2}(k,l)=S\circ\varrho_{1}(k,l)\circ S^{*}. Then

(ϱ2(k,l)A)(m,n)={A(lm,n) for n=1e2πikm/pA(lm,n) for n{2,,p1}(\varrho_{2}(k,l)A)(m,n)=\left\{\begin{array}[]{ll}A(lm,n)&\mbox{ for }n=1\\ e^{-2\pi ikm/p}A(lm,n)&\mbox{ for }n\in\{2,\ldots,p-1\}\end{array}\right. (6.4)

Define 𝒦1{1,,p1}×{1,,p1}\mathcal{K}_{1}\subset\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}} as the subspace of matrices AA with A(m,n)=0A(m,n)=0 for all n1n\not=1, and let 𝒦2=𝒦1{1,,p1}×{2,,p1}\mathcal{K}_{2}=\mathcal{K}_{1}^{\bot}\cong\mathbb{C}^{\{1,\ldots,p-1\}\times\{2,\ldots,p-1\}}. These spaces are invariant under ϱ2\varrho_{2}, and we have

ϱ2|𝒦1λ~p,ϱ2|𝒦2π^0𝟏{2,,p1}.\varrho_{2}|_{\mathcal{K}_{1}}\simeq\tilde{\lambda}_{\mathbb{Z}_{p}^{\ast}}~{},~{}\varrho_{2}|_{\mathcal{K}_{2}}\simeq\widehat{\pi}_{0}\otimes\mathbf{1}_{\mathbb{C}^{\{2,\ldots,p-1\}}}~{}.

where λ~p\tilde{\lambda}_{\mathbb{Z}_{p}^{\ast}} is the regular representation of p\mathbb{Z}_{p}^{\ast}, lifted to GG.

Proof 7

It is clearly enough to prove (6.4), and we do this by explicit calculation. First, one quickly verifies that the map

(m,n){(m,m) for n=1(m(1n)1,mn(1n)1) for n{2,,p1}(m,n)\mapsto\left\{\begin{array}[]{ll}(-m,-m)&\mbox{ for }n=1\\ (m(1-n)^{-1},mn(1-n)^{-1})&\mbox{ for }n\in\{2,\ldots,p-1\}\end{array}\right.

is a bijection on {1,,p1}×{1,,p1}\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}, with inverse given by

(m,n){(m,1) for n=m(mn,m1n) for mn(m,n)\mapsto\left\{\begin{array}[]{ll}(-m,1)&\mbox{ for }n=m\\ (m-n,m^{-1}n)&\mbox{ for }m\not=n\end{array}\right.

Hence SS is indeed unitary, with

(SA)(m,n)={A(m,1) for n=mA(mn,m1n) for mn(S^{*}A)(m,n)=\left\{\begin{array}[]{ll}A(-m,1)&\mbox{ for }n=m\\ A(m-n,m^{-1}n)&\mbox{ for }m\not=n\end{array}\right.

Fixing A{1,,p1}×{1,,p1}A\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}, we let B1=ϱ1(k,l)S(A)B_{1}=\varrho_{1}(k,l)S^{*}(A) and B2=SB1=ϱ2(k,l)AB_{2}=SB_{1}=\varrho_{2}(k,l)A. We then get

B1(m,n)=e2πik(mn)/p{A(lm,1) for n=mA((mn),m1n) for mnB_{1}(m,n)=e^{-2\pi ik(m-n)/p}\left\{\begin{array}[]{ll}A(lm,1)&\mbox{ for }n=m\\ A(\ell(m-n),m^{-1}n)&\mbox{ for }m\not=n\end{array}\right.

Accordingly, for m=1,,p1m=1,\ldots,p-1,

B2(m,1)=B1(m,m)=A(lm,1).B_{2}(m,1)=B_{1}(m,m)=A(lm,1)~{}.

For n{2,,p1}n\in\{2,\ldots,p-1\} we fix (m0,n0)=(m(1n)1,mn(1n)1)(m_{0},n_{0})=(m(1-n)^{-1},mn(1-n)^{-1}), and get

m0n0=m(1n)1mn(1n)1=m(1n)(1n)1=m,m01n0=nm_{0}-n_{0}=m(1-n)^{-1}-mn(1-n)^{-1}=m(1-n)(1-n)^{-1}=m~{},~{}m_{0}^{-1}n_{0}=n

which allows to compute

B2(m,n)=B1(m0,n0)=e2πik(m0n0)/pA(l(m0n0),m01n0)=e2πikm/pA(lm,n),B_{2}(m,n)=B_{1}(m_{0},n_{0})=e^{-2\pi ik(m_{0}-n_{0})/p}A(l(m_{0}-n_{0}),m_{0}^{-1}n_{0})=e^{-2\pi ikm/p}A(lm,n)~{},

and (6.4) is established.

We next state criteria for matrix recovery.

Theorem 6.5

Let φ{1,,p1}\varphi\in\mathbb{C}^{\{1,\ldots,p-1\}}. Then π^0(G)φ\widehat{\pi}_{0}(G)\varphi does matrix recovery iff φ\varphi fulfills the following conditions:

  1. (i)

    All entries of cφp^c_{\varphi}\in\widehat{\mathbb{Z}_{p}^{*}}^{\mathbb{C}}, defined by

    cφ(χ)=m=1p1|φ(l)|2χ(l)c_{\varphi}(\chi)=\sum_{m=1}^{p-1}|\varphi(-l)|^{2}\chi(l)

    are nonzero.

  2. (ii)

    The matrix Bφ{1,,p1}×{1,,p2}B_{\varphi}\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-2\}} defined by

    Bφ(m,n)=φ(mn)φ(m(n+1))¯B_{\varphi}(m,n)=\varphi(mn)\overline{\varphi(m(n+1))}~{}

    has a left inverse Bφ{1,,p2}×{1,,p1}B_{\varphi}^{\dagger}\in\mathbb{C}^{\{1,\ldots,p-2\}\times\{1,\ldots,p-1\}}.

The first and most involved step of the proof (given below) consists in explicitly inverting the linear operator Vφφϱ1:{1,,p1}×{1,,p1}GV_{\varphi\otimes\varphi}^{\varrho_{1}}:\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}\to\mathbb{C}^{G}, assuming that φ\varphi fulfills conditions (i) and (ii). We give a detailed description of this procedure in the following corollary to the proof:

Corollary 6.6

Assume that φ{1,,p1}\varphi\in\mathbb{C}^{\{1,\ldots,p-1\}} fulfills conditions (i) and (ii) from the previous theorem. An explicit algorithm for the inversion of the linear map Vφφϱ1:C{1,,p1}×{1,,p1}GV_{\varphi\otimes\varphi}^{\varrho_{1}}:C^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}\to\mathbb{C}^{G} is described as follows: Let Ω0{1,,p1}×{1,,p1}\Omega_{0}\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}} denote the matrix describing the linear map

(Ω0f)(m)=f(m),(\Omega_{0}f)(m)=f(-m)~{},

and let Ω1{1,,p2}×{2,,p1}\Omega_{1}\in\mathbb{C}^{\{1,\ldots,p-2\}\times\{2,\ldots,p-1\}} be the matrix describing the linear map

(Ω1f)(n)=f(1+n1).(\Omega_{1}f)(n)=f(1+n^{-1})~{}.

Let A{1,,p1}×{1,,p1}A\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}} and let FGF\in\mathbb{C}^{G} be defined as

F(k,l)=(Vφφϱ1A)(k,l)=Aπ^0(k,l)φ,π^0(k,l)φ.F(k,l)=\left(V_{\varphi\otimes\varphi}^{\varrho_{1}}A\right)(k,l)=\langle A\widehat{\pi}_{0}(k,l)\varphi,\widehat{\pi}_{0}(k,l)\varphi\rangle~{}.

Then AA can be recovered from FF by the following steps:

  1. 1.

    Define a1{1,,p1}a_{1}\in\mathbb{C}^{\{1,\ldots,p-1\}} by

    a1(k)=1p(p1)χp^cφ(χ)1χ~(F)χ(k).a_{1}(k)=\frac{1}{p(p-1)}\sum_{\chi\in\widehat{\mathbb{Z}_{p}^{\ast}}}c_{\varphi}(\chi)^{-1}\tilde{\chi}(F)\chi(k)~{}.
  2. 2.

    Define A2=pπ0^(F)Ω0T(Bφ)Ω1{1,,p1}×{2,,p1}A_{2}^{\prime}=p\cdot\widehat{\pi_{0}}(F)\cdot\Omega_{0}^{T}\cdot(B_{\varphi}^{\dagger})^{\ast}\cdot\Omega_{1}\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{2,\ldots,p-1\}}.

  3. 3.

    With the unitary operator S:{1,,p1}×{1,,p1}{1,,p1}×{1,,p1}S:\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}}\to\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}} from Proposition 6.4, we have

    A=S((a1|A2))A=S^{*}\left((a_{1}|A_{2}^{\prime})\right)

In particular, f{1,,p1}f\in\mathbb{C}^{\{1,\ldots,p-1\}} can be recovered up to a phase factor by computing AA from F=|Vφπ0^f|2F=|V_{\varphi}^{\widehat{\pi_{0}}}f|^{2} using steps (1)-(3), and then determining ff as the eigenvector associated to the nonzero eigenvalue of AA, normalized by f22=trace(A)\|f\|_{2}^{2}={\rm trace}(A).

Proof 8

Throughout the proof, the symbol 𝟎\mathbf{0} will denote zero matrices of varying sizes. Since these sizes can be derived from the context, we will refrain from using more specific notation.

The intertwining property of SS yields

F(k,l)\displaystyle F(k,l) =\displaystyle= A,π(k,l)φπ(k,l)φ\displaystyle\langle A,\pi(k,l)\varphi\otimes\pi(k,l)\varphi\rangle
=\displaystyle= SA,ϱ2(k,l)S(φφ).\displaystyle\langle SA,\varrho_{2}(k,l)S(\varphi\otimes\varphi)\rangle~{}.

We write SA=A1+A2SA=A_{1}+A_{2} with Ai𝒦iA_{i}\in\mathcal{K}_{i}, and get

A1=(a1|𝟎),A2=(𝟎|A2),A_{1}=\left(a_{1}|\mathbf{0}\right)~{},~{}A_{2}=\left(\mathbf{0}|A_{2}^{\prime}\right)~{},

with a vector a1{1,,p1}a_{1}\in\mathbb{C}^{\{1,\ldots,p-1\}}, A2={1,,p1}×{2,,p1}A_{2}^{\prime}=\mathbb{C}^{\{1,\ldots,p-1\}\times\{2,\ldots,p-1\}}.

Similarly, we have

S(φφ)=(cφ|Cφ)=(cφ|𝟎)𝒦1+(𝟎|Cφ)𝒦2S(\varphi\otimes\varphi)=(c_{\varphi}|C_{\varphi}^{\prime})=\underbrace{(c_{\varphi}|\mathbf{0})}_{\in\mathcal{K}_{1}}+\underbrace{(\mathbf{0}|C_{\varphi}^{\prime})}_{\in\mathcal{K}_{2}}

where (cφ|Cφ)=Cφ(c_{\varphi}|C^{\prime}_{\varphi})=C_{\varphi} is obtained by plugging

(φφ)(m,n)=φ(m)φ(n)¯(\varphi\otimes\varphi)(m,n)=\varphi(m)\overline{\varphi(n)}

into the definition of SS, which leads to

Cφ(m,n)={|φ(m)|2 for n=1φ(m(1n)1)φ(mn(1n)1)¯ for n{2,,p1}C_{\varphi}(m,n)=\left\{\begin{array}[]{ll}|\varphi(-m)|^{2}&\mbox{ for }n=1\\ \varphi(m(1-n)^{-1})\overline{\varphi(mn(1-n)^{-1})}&\mbox{ for }n\in\{2,\ldots,p-1\}\end{array}\right. (6.5)

In particular,

cφ(n)=|φ(n)|2.c_{\varphi}(n)=|\varphi(-n)|^{2}~{}.

The overall proof strategy for the inversion formula is to recover both a1a_{1} and A2A_{2}^{\prime} from FF via group Fourier transform, which allows to reconstruct SASA. It then only remains to apply SS^{*} to the result.

By definition of the various matrices involved, we have

F(k,l)=A1,ϱ2(k,l)(cφ|𝟎)=F1(k,l)+A2,ϱ2(k,l)(𝟎|Cφ)=F2(k,l).F(k,l)=\underbrace{\langle A_{1},\varrho_{2}(k,l)(c_{\varphi}|\mathbf{0})\rangle}_{=F_{1}(k,l)}+\underbrace{\langle A_{2},\varrho_{2}(k,l)(\mathbf{0}|C_{\varphi}^{\prime})\rangle}_{=F_{2}(k,l)}~{}.

Now equation (6.4) yields

F1(k,l)=m=1p1a(m)|φ(lm)|2,F_{1}(k,l)=\sum_{m=1}^{p-1}a(m)|\varphi(-lm)|^{2}~{}, (6.6)

whereas

F2(k,l)=A2,π^0(k,l)(𝟎,Cφ)=trace(A2(𝟎|Cφ)π^0(k,l)).F_{2}(k,l)=\langle A_{2},\widehat{\pi}_{0}(k,l)(\mathbf{0},C_{\varphi}^{\prime})\rangle={\rm trace}\left(A_{2}\cdot(\mathbf{0}|C_{\varphi}^{\prime})^{*}\cdot\widehat{\pi}_{0}(k,l)^{*}\right)~{}. (6.7)

This reveals F1F_{1} as a convolution product over the factor group p\mathbb{Z}_{p}^{\ast}, whereas F2F_{2} is seen to be a special instance of the Fourier inversion formula (6.3). In particular, we have

F1span{χ~:χp^},F2span{χ~:χp^}F_{1}\in{\rm span}\{\tilde{\chi}:\chi\in\widehat{\mathbb{Z}_{p}^{*}}\}~{},~{}F_{2}~{}\bot~{}{\rm span}\{\tilde{\chi}:\chi\in\widehat{\mathbb{Z}_{p}^{*}}\} (6.8)

We now can use (6.8) to get for all χp^\chi\in\widehat{\mathbb{Z}_{p}^{\ast}} that

χ~(F)\displaystyle\tilde{\chi}(F) =\displaystyle= χ~(F1)\displaystyle\tilde{\chi}(F_{1})
=\displaystyle= pl=1p1m=1p1a(m)|φ(lm)|2χ(l)\displaystyle p\sum_{l=1}^{p-1}\sum_{m=1}^{p-1}a(m)|\varphi(-lm)|^{2}\chi(l)
=\displaystyle= pl=1p1m1p1a(m)χ¯(m)|φ(lm)|2χ(lm)\displaystyle p\sum_{l=1}^{p-1}\sum_{m-1}^{p-1}a(m)\overline{\chi}(m)|\varphi(-lm)|^{2}\chi(lm)
=\displaystyle= pl=1p1m1p1a(m)χ¯(m)|φ(l)|2χ(l)\displaystyle p\sum_{l=1}^{p-1}\sum_{m-1}^{p-1}a(m)\overline{\chi}(m)|\varphi(-l)|^{2}\chi(l)
=\displaystyle= pχ¯(a)cφ(χ).\displaystyle p\overline{\chi}(a)c_{\varphi}(\chi)~{}.

By assumption, cφ(χ)0c_{\varphi}(\chi)\not=0, which allows to solve this equation for χ¯(a)\overline{\chi}(a), and plug the result into the Fourier inversion on p\mathbb{Z}_{p}^{\ast}, yielding

a(k)\displaystyle a(k) =\displaystyle= 1p1χp^χ¯(a)χ(k)\displaystyle\frac{1}{p-1}\sum_{\chi\in\widehat{\mathbb{Z}_{p}^{\ast}}}\overline{\chi}(a)\chi(k)
=\displaystyle= 1p(p1)χp^cφ(χ)1χ~(F)χ(k),\displaystyle\frac{1}{p(p-1)}\sum_{\chi\in\widehat{\mathbb{Z}_{p}^{\ast}}}c_{\varphi}(\chi)^{-1}\tilde{\chi}(F)\chi(k)~{},

which is the formula from the first step.

For the second step, we recall that (6.7) allows to view F2F_{2} as an inverse Fourier transform, which together with (6.8) gives rise to

p1A2(Cφ)\displaystyle p^{-1}A^{\prime}_{2}\cdot(C_{\varphi}^{\prime})^{*} =\displaystyle= |G|dπ^0((𝟎|A2)(0|Cφ))\displaystyle\frac{|G|}{d_{\widehat{\pi}_{0}}}\left((\mathbf{0}|A^{\prime}_{2})\cdot(0|C_{\varphi}^{\prime})^{*}\right) (6.9)
=\displaystyle= π^0(F2)\displaystyle\widehat{\pi}_{0}(F_{2})
=\displaystyle= π0^(F)\displaystyle\widehat{\pi_{0}}(F)

We next show

Bφ=Ω0CφΩ1T.B_{\varphi}=\Omega_{0}\cdot C_{\varphi}^{\prime}\cdot\Omega_{1}^{T}~{}. (6.10)

For this purpose we introduce the change of variables ω:{1,,p2}{2,,p1}\omega:\{1,\ldots,p-2\}\to\{2,\ldots,p-1\}, ω(n)=1+n1\omega(n)=1+n^{-1}. ω\omega is a well-defined bijection, with inverse map given by ω1(n)=(n1)1\omega^{-1}(n)=(n-1)^{-1}. Thus the map Ω1:ffω\Omega_{1}:f\mapsto f\circ\omega is a well-defined unitary map with inverse Ω1T\Omega_{1}^{T}. We apply this change of variables to the row vectors of CφC_{\varphi}^{\prime}, and the formulas

(ω(n)1)1=n,ω(n)(ω(n)1)1=n+1(\omega(n)-1)^{-1}=n~{},~{}\omega(n)(\omega(n)-1)^{-1}=n+1

show that

(CφΩ1T)(m,n)\displaystyle\left(C_{\varphi}^{\prime}\cdot\Omega_{1}^{T}\right)(m,n) =\displaystyle= φ(m(1ω(n))1)φ(mω(n)(1ω(n))1)¯\displaystyle\varphi(m(1-\omega(n))^{-1})\overline{\varphi(m\omega(n)(1-\omega(n))^{-1})}
=\displaystyle= φ(mn)φ(m(n+1))¯.\displaystyle\varphi(-mn)\overline{\varphi(-m(n+1))}~{}.

But then left multiplication by Ω0\Omega_{0}, effecting a sign flip in mm, yields

(Ω0CφΩ1T)(m,n)=Bφ(m,n),\left(\Omega_{0}\cdot C_{\varphi}^{\prime}\cdot\Omega_{1}^{T}\right)(m,n)=B_{\varphi}(m,n)~{},

as desired. Hence (6.10) is established, which implies Cφ=Ω0TBφΩ1C_{\varphi}^{\prime}=\Omega_{0}^{T}\cdot B_{\varphi}\cdot\Omega_{1}. In combination with (6.9) this yields

pπ^0(F)Ω0T(Bφ)Ω1\displaystyle p\cdot\widehat{\pi}_{0}(F)\cdot\Omega_{0}^{T}\cdot(B_{\varphi}^{\dagger})^{*}\cdot\Omega_{1} =\displaystyle= A2(Cφ)Ω0T(Bφ)Ω1\displaystyle A_{2}^{\prime}\cdot(C_{\varphi}^{\prime})^{*}\cdot\Omega_{0}^{T}\cdot(B_{\varphi}^{\dagger})^{*}\cdot\Omega_{1}
=\displaystyle= A2(Ω1TBφΩ0)Ω0TBφ)Ω1\displaystyle A_{2}^{\prime}\cdot\left(\Omega_{1}^{T}B_{\varphi}^{*}\cdot\Omega_{0}\right)\cdot\Omega_{0}^{T}\cdot B_{\varphi}^{\dagger})^{*}\cdot\Omega_{1}
=\displaystyle= A2Ω1T(BφBφ)Ω1\displaystyle A_{2}^{\prime}\cdot\Omega_{1}^{T}\cdot(B_{\varphi}^{\dagger}\cdot B_{\varphi})^{*}\cdot\Omega_{1}
=\displaystyle= A2,\displaystyle A_{2}^{\prime}~{},

by choice of BφB_{\varphi}^{\dagger}.

Hence the recovery of SASA from FF via steps (1) and (2) is shown, and application of SS^{*} yields the desired result.

This establishes the inversion algorithm, as well as sufficiency of the conditions (i), (ii). It remains to show necessity of the conditions. First assume that condition (i) is violated, i.e. there exists a character χp^\chi\in\widehat{\mathbb{Z}_{p}^{*}} such that

m=1p1|φ(l)|2χ(l)=0.\sum_{m=1}^{p-1}|\varphi(-l)|^{2}\chi(l)=0~{}.

Fix a1=χ{1,,p1}a_{1}=\chi\in\mathbb{C}^{\{1,\ldots,p-1\}}, and let A=S(a1|𝟎)A=S^{*}(a_{1}|\mathbf{0}), as well as F=Vφφϱ1AF=V_{\varphi\otimes\varphi}^{\varrho_{1}}A. Using the above notations and calculations, we then have

F(k,l)=F1(k,l)=m=1p1a(m)|φ(lm)|2=χ¯(l)m=1p1χ(lm)|φ(lm)|2=0F(k,l)=F_{1}(k,l)=\sum_{m=1}^{p-1}a(m)|\varphi(-lm)|^{2}=\overline{\chi}(l)\sum_{m=1}^{p-1}\chi(lm)|\varphi(-lm)|^{2}=0

for all (k,l)G(k,l)\in G. On the other hand, A0A\not=0, since SS^{*} is unitary.

Assuming condition (ii) to be violated, we obtain the existence of a nonzero vector v{1,,p2}v\in\mathbb{C}^{\{1,\ldots,p-2\}} such that Bφv=0B_{\varphi}v=0. We pick an arbitrary nonzero w{1,,p1}w\in\mathbb{C}^{\{1,\ldots,p-1\}} and define A2=(wv)Ω1A_{2}=(w\otimes v)\cdot\Omega_{1}, and A=S(𝟎|A2)A=S^{*}(\mathbf{0}|A_{2}). Then AA is nonzero, but the above computations yield for all (k,l)G(k,l)\in G:

F(k,l)\displaystyle F(k,l) =\displaystyle= F2(k,l)\displaystyle F_{2}(k,l)
=\displaystyle= trace(A2(Cφ)π(k,l))\displaystyle{\rm trace}(A_{2}\cdot(C_{\varphi}^{\prime})^{\ast}\pi(k,l)^{*})
=\displaystyle= trace((wv)Ω1Ω1TBφΩ0π(k,l))\displaystyle{\rm trace}((w\otimes v)\cdot\Omega_{1}\cdot\Omega_{1}^{T}\cdot B_{\varphi}^{*}\cdot\Omega_{0}\cdot\pi(k,l)^{*})
=\displaystyle= trace((wBφv)=0Ω0π(k,l))\displaystyle{\rm trace}(\underbrace{(w\otimes B_{\varphi}v)}_{=0}\cdot\Omega_{0}\cdot\pi(k,l)^{*})
=\displaystyle= 0\displaystyle 0

by choice of vv. This concludes the proof.

The remaining open question is whether the conditions for matrix recovery established in the last theorem can actually be fulfilled by a suitably chosen vector φ\varphi. The positive answer is contained in the next theorem. Note that Lemma 2.5 then implies that almost every vector does matrix recovery.

Theorem 6.7

Let φ{1,,p1}\varphi\in\mathbb{C}^{\{1,\ldots,p-1\}} be given by

φ(1)=1,φ(2)=2,\varphi(1)=1~{},~{}\varphi(2)=2~{},

in the case p=3p=3, and by

φ(m)=1δ1(m),\varphi(m)=1-\delta_{1}(m)~{},

for p>3p>3. Then π^0(G)φ\widehat{\pi}_{0}(G)\varphi does matrix retrieval.

Proof 9

In the case p=3p=3, the matrix BφB_{\varphi} has dimensions {1,2}×{1}\{1,2\}\times\{1\}, and it fulfills the rank condition (ii) iff φ(1)φ(2)0\varphi(1)\varphi(2)\not=0. Condition (i) is fulfilled as soon as |φ(1)||φ(2)||\varphi(1)|\not=|\varphi(2)|, and both conditions are fulfilled by φ(1)=1,φ(2)=2\varphi(1)=1,\varphi(2)=2.

Now assume p>3p>3 and let φ(m)=1δ1(m)\varphi(m)=1-\delta_{1}(m). Then property (i)(i) is easy to verify: For the constant characer χ0\chi_{0} of p^\widehat{\mathbb{Z}_{p}^{\ast}}, we find |φ|2,χ0=p2\langle|\varphi|^{2},\chi_{0}\rangle=p-2. Noting that |φ|2=χ0δ1|\varphi|^{2}=\chi_{0}-\delta_{1}, we find for all remaining characters χ\chi that

|φ|2,χ=δ1,χ=χ(1)0.\langle|\varphi|^{2},\chi\rangle=\langle-\delta_{1},\chi\rangle=-\chi(1)\not=0~{}.

Therefore it remains to verify condition (ii), which is clearly equivalent to the equation rank(Bφ)=p2{\rm rank}(B_{\varphi})=p-2. By definition of φ\varphi we get Bφ(m,n){0,1}B_{\varphi}(m,n)\in\{0,1\} for all m,nm,n, with

Bφ(m,n)=0\displaystyle B_{\varphi}(m,n)=0 \displaystyle\Longleftrightarrow (mn=1)(m(n+1)=1)\displaystyle\left(mn=1\right)\lor\left(m(n+1)=1\right)
\displaystyle\Longleftrightarrow m{n1,(n+1)1}.\displaystyle m\in\{n^{-1},(n+1)^{-1}\}~{}.

Letting B~φ\tilde{B}_{\varphi} denote the matrix obtained by

B~φ(m,n)=Bφ(m1,n),\tilde{B}_{\varphi}(m,n)=B_{\varphi}(m^{-1},n)~{},

we have rank(B~φ)=rank(Bφ){\rm rank}(\tilde{B}_{\varphi})={\rm rank}(B_{\varphi}), and

B~φ(m,n)=0m{n,n+1},\tilde{B}_{\varphi}(m,n)=0\Longleftrightarrow m\in\{n,n+1\}~{},

with value 11 for the remaining entries.

The next step in the proof consists in successive applications of Gaussian elimination steps to the rows of B~φ\tilde{B}_{\varphi}: We first subtract the second row from the first, then the third row from the second,\ldots, the (p1)(p-1)th row from the (p2)(p-2)th. This yields

B˙φ(m,n)={1,m2,n=m11,mp3,n=m+10 elsewhere \dot{B}_{\varphi}(m,n)=\left\{\begin{array}[]{cc}-1,&m\geq 2,n=m-1\\ 1,&m\leq p-3,n=m+1\\ 0&\mbox{ elsewhere }\end{array}\right.

Removal of the first row (and an index shift) results in the square matrix Mφ{1,,p2}×{1,,p2}M_{\varphi}\in\mathbb{C}^{\{1,\ldots,p-2\}\times\{1,\ldots,p-2\}}, given by

Mφ(m,n)={1,mp2,n=m1,mp4,n=m+21,m=p2,np30,otherwiseM_{\varphi}(m,n)=\left\{\begin{array}[]{cc}-1,&m\leq p-2,n=m\\ 1,&m\leq p-4,n=m+2\\ 1,&m=p-2,n\leq p-3\\ 0,&\mbox{otherwise}\end{array}\right. (6.11)

i.e.

Mφ=(101101101101101110).M_{\varphi}=\left(\begin{array}[]{rrrrrrrr}-1&0&1&&&&\\ &-1&0&1&&&&\\ &&\ddots&\ddots&\ddots&&&\\ &&&\ddots&\ddots&\ddots&&\\ &&&&-1&0&1&\\ &&&&&-1&0&1\\ &&&&&&-1&0\\ 1&1&\ldots&\ldots&\ldots&\ldots&1&0\end{array}\right)~{}. (6.12)

Since MφM_{\varphi} is a submatrix of B˙φ\dot{B}_{\varphi}, the steps taken so far ensure that the proof is finished once we have shown that MφM_{\varphi} is invertible. For this purpose, we make one final Gaussian elimination step to take care of the last row of MφM_{\varphi}, by letting

M˙φ(p2,n)=Mφ(p2,n)+j=1(p3)/2jMφ(2j1,n)+jMφ(2j,n),\dot{M}_{\varphi}(p-2,n)=M_{\varphi}(p-2,n)+\sum_{j=1}^{(p-3)/2}jM_{\varphi}(2j-1,n)+jM_{\varphi}(2j,n)~{}, (6.13)

for n=1,,p2n=1,\ldots,p-2, and M˙(k,m)=M(k,m)\dot{M}(k,m)=M(k,m) for k<p2k<p-2.

For n=p2n=p-2, we then get

M˙φ(p2,n)=0+(p3)/20.\dot{M}_{\varphi}(p-2,n)=0+(p-3)/2\not=0~{}. (6.14)

Now fix n<p2n<p-2. In case n2n\leq 2, there is only a single contribution on the right-hand side, leading to

M˙(p2,n)=M(p2,n)1=0.\dot{M}(p-2,n)=M(p-2,n)-1=0~{}.

For 3np23\leq n\leq p-2, we distinguish between even and odd cases. In the case where n=2kn=2k, the only nonzero contributions on the right-hand side of (6.13) are provided by j=kj=k and j=k1j=k-1, leading to

M˙φ(p2,n)=1+(k1)Mφ(2k2,2k)=1+kMφ(2k,2k)=1=0\dot{M}_{\varphi}(p-2,n)=1+(k-1)\underbrace{M_{\varphi}(2k-2,2k)}_{=1}+k\underbrace{M_{\varphi}(2k,2k)}_{=-1}=0

In the case n=2k+1n=2k+1, the nonzero contributions also come from j=kj=k and j=k1j=k-1, with corresponding entries 1-1 resp. 11, resulting again in M˙φ(p2,n)=0\dot{M}_{\varphi}(p-2,n)=0.

In summary, we find that M˙φ\dot{M}_{\varphi} is upper triangular, with nonvanishing diagonal thanks to (6.11) and (6.14), and hence M˙φ\dot{M}_{\varphi} is invertible. This concludes the proof.

The appearance of the group Fourier transform in the matrix recovery algorithm is no coincidence, but rather a byproduct of the approach taken in this paper. The following remark explains these connections in some more detail, and sketches how Corollary 6.6 can possibly be generalized to other settings.

Remark 6.8

Recall from Proposition 2.4 that matrix recovery for group frames requires finding certain cyclic vectors for the conjugation representation

ϱ(x)(A)=π(x)Aπ(x),\varrho(x)(A)=\pi(x)\cdot A\cdot\pi(x)^{*}~{},

acting on dπ×dπ\mathbb{C}^{d_{\pi}\times d_{\pi}}. The natural representation-theoretic approach to this problem relies on the decomposition of ϱ\varrho into irreducibles, of the type

ϱσG^mσσ,\varrho\simeq\bigoplus_{\sigma\in\widehat{G}}m_{\sigma}\cdot\sigma~{},

where mσ0m_{\sigma}\in\mathbb{N}_{0} denotes the multiplicity with which σ\sigma occurs in ϱ\varrho.

Slightly rewriting this decomposition allows to introduce the group Fourier transform in a rather natural fashion. As a starting point, we make the observation that any cyclic vector AA for ϱ\varrho yields an injective operator VAϱ:Cdπ×dπGV_{A}^{\varrho}:C^{d_{\pi}\times d_{\pi}}\to\mathbb{C}^{G} intertwining ϱ\varrho with the left regular representation λG\lambda_{G} acting on G\mathbb{C}^{G}. A necessary and sufficient condition for the existence of such an intertwining operator is that mσm_{\sigma} is less than or equal to the multiplicity of σ\sigma in λG\lambda_{G}, and the latter coincides with dσd_{\sigma}.

Hence, we have obtained mσdσm_{\sigma}\leq d_{\sigma} for all σG^\sigma\in\widehat{G}, as a necessary condition for matrix recovery. In this case, one can in effect realize mσσm_{\sigma}\cdot\sigma by left action of σ\sigma on a suitable subspace 𝒦σdσ×dσ\mathcal{K}_{\sigma}\subset\mathbb{C}^{d_{\sigma}\times d_{\sigma}}, acting by left multiplication. This results in an intertwining operator

S:dπ×dπσG^𝒦σS:\mathbb{C}^{d_{\pi}\times d_{\pi}}\to\prod_{\sigma\in\widehat{G}}\mathcal{K}_{\sigma}

such that

Sϱ(x)S(Aσ)σG^=(σ(x)Aσ)σG^.S\varrho(x)S^{*}(A_{\sigma})_{\sigma\in\widehat{G}}=(\sigma(x)\cdot A_{\sigma})_{\sigma\in\widehat{G}}~{}. (6.15)

In this realization, we get for arbitrary A,Bdπ×dπA,B\in\mathbb{C}^{d_{\pi}\times d_{\pi}} with A^=(A^σ)σG^=SA\widehat{A}=(\widehat{A}_{\sigma})_{\sigma\in\widehat{G}}=SA and B^=(B^σ)σG^=SB\widehat{B}=(\widehat{B}_{\sigma})_{\sigma\in\widehat{G}}=SB, that the matrix coefficient VAϱBV_{A}^{\varrho}B can be computed as

VAϱB(x)\displaystyle V_{A}\varrho B(x) =\displaystyle= B,ϱ(x)A\displaystyle\langle B,\varrho(x)A\rangle
=\displaystyle= σG^B^σ,σ(x)A^σ\displaystyle\sum_{\sigma\in\widehat{G}}\langle\widehat{B}_{\sigma},\sigma(x)\widehat{A}_{\sigma}\rangle
=\displaystyle= σG^trace(B^σA^σσ(x)),\displaystyle\sum_{\sigma\in\widehat{G}}{\rm trace}(\widehat{B}_{\sigma}\cdot\widehat{A}_{\sigma}^{*}\cdot\sigma(x)^{*})~{},

which shows that the coefficient transform is very closely related to the Fourier inversion formula (6.3). As a consequence, we can make the following observations, using similar reasoning as in the proof of Theorem 6.5:

  1. (a)

    Invertibility of VAϱV_{A}^{\varrho} is equivalent to injectivity of BσBσAσB_{\sigma}\mapsto B_{\sigma}\cdot A_{\sigma}^{*} on each 𝒦σ\mathcal{K}_{\sigma}.

  2. (b)

    Recovery of BB from F=VAϱBF=V^{\varrho}_{A}B can be carried out as follows: (i) Compute B^σA^σ=|G|dσ1σ(F)\widehat{B}_{\sigma}\cdot\widehat{A}_{\sigma}^{*}=|G|d_{\sigma}^{-1}\sigma(F); (ii) for each σ\sigma, recover B^σ𝒦σ\widehat{B}_{\sigma}\in\mathcal{K}_{\sigma} from the result of the previous step (as guaranteed by the condition in (a)); and (iii) invert SS.

  3. (c)

    The additional challenge posed by matrix retrieval from measurements of the type

    Aπ(x)ψ,π(x)ψ\langle A\pi(x)\psi,\pi(x)\psi\rangle

    shows itself in applying the criteria and inversion methods from (a) and (b) to specific matrices A=ψψA=\psi\otimes\psi, which depends quadratically on the entries in ψ\psi. Note that the explicit formulation of these criteria also hinges on the concrete choice of the intertwining operator SS.

We emphasize that these observations apply to any finite group, and they provide a unified representation-theoretic perspective on matrix recovery via group frames. Remark 6.9 below explains how the well-understood case of Heisenberg groups fits in this framework. Note that this point of view is quite similar to the treatment of admissible vectors (i.e., special continuous group frames arising from unitary actions of locally compact groups, and their relation to the Plancherel formula of such groups), as developed in [14].

The criteria and inversion method in Theorem 6.5 are closely related to the general program sketched in (a)-(c) above, but they do not constitute a faithful implementation. Proposition 6.4 decomposes the representation space into two orthogonal subspaces, 𝒦1\mathcal{K}_{1} and 𝒦2\mathcal{K}_{2}. On 𝒦1\mathcal{K}_{1}, ϱ2\varrho_{2} is essentially the regular representation of the quotient group p\mathbb{Z}_{p}^{\ast}, whereas the action on 𝒦2\mathcal{K}_{2} is a multiple of the irreducible representation π^0\widehat{\pi}_{0}.

Now the decomposition of the regular representation of p\mathbb{Z}_{p}^{\ast} is effected by its characters, which can also be considered as (lifted) characters of the whole group GG. Thus the Fourier coefficients cφc_{\varphi} occurring in condition (i) from Theorem 6.5 and in Step (1) of the recovery algorithm could also be expressed in terms of the group Fourier transform of GG, which would increase the similarity to steps (a) and (b).

In the case of condition (ii) and recovery step (2) from Theorem 6.5, the group Fourier transform is more explicitly employed, in a way that corresponds quite closely to steps (a) and (b) above. The main difference is that we opted to formulate the injectivity condition (ii), which corresponds to (a), in terms of the matrix BφB_{\varphi}. This matrix has the advantage that its dependence on the vector φ\varphi is rather transparent. This simplifies verification of condition (ii), as can be seen in the proof of Theorem 6.7. The price to pay is the occurrence of the auxiliary matrices Ω0\Omega_{0} and Ω1\Omega_{1} in the inversion step (2).

Remark 6.9

In this remark we study phase retrieval and matrix recovery in the context of finite Gabor systems. The following is largely based on observations made in [5]. The main purpose of this remark, and the slightly novel contribution to the discussion of matrix recovery for the Schrödinger representation, is the explicit reference to the program outlined in Remark 6.8.

Fix nn\in\mathbb{N}, n2n\geq 2, and define the associated finite Heisenberg group by

=n3\mathbb{H}=\mathbb{Z}_{n}^{3}

with group law

(k,l,m)(k,l,m)=(k+k,l+l,m+mlk).(k,l,m)\cdot(k^{\prime},l^{\prime},m^{\prime})=(k+k^{\prime},l+l^{\prime},m+m^{\prime}-l^{\prime}k)~{}.

The center of this group is given by Z()={0}×{0}×nZ(\mathbb{H})=\{0\}\times\{0\}\times\mathbb{Z}_{n}, and /Z()n×n\mathbb{H}/Z(\mathbb{H})\cong\mathbb{Z}_{n}\times\mathbb{Z}_{n}.

The Schrödinger representation of n\mathbb{H}_{n} acts on n\mathbb{C}^{n} via

(π(k,l,m)f)(y)=e2πim/ne2πily/nf(yk),k,l,m,y{0,,n1},(\pi(k,l,m)f)(y)=e^{2\pi im/n}e^{2\pi ily/n}f(y-k)~{},k,l,m,y\in\{0,\ldots,n-1\}~{},

with yky-k understood modulo nn. The Schrödinger representation is irreducible, and the restriction of π\pi to the center Z()Z(\mathbb{H}) is a character. Since we are interested in matrix retrieval, we will systematically omit the central variable in the following: All scalar actions cancel out in the conjugation action, which means that Z()Z(\mathbb{H}) is necessarily in the kernel of ϱ\varrho. In particular, we will write π(k,l):=π(k,l,0)\pi(k,l):=\pi(k,l,0).

In order to establish criteria for matrix recovery, we make the following two observations:

k,l,k,l:trace(π(k,l)π(k,l))=nδk,kδl,l,\forall k,l,k^{\prime},l^{\prime}~{}:~{}{\rm trace}(\pi(k,l)\pi(k^{\prime},l^{\prime})^{*})=n\delta_{k,k^{\prime}}\delta_{l,l^{\prime}}~{}, (6.16)

which means that (for dimension reasons) the matrices (n1/2π(k,l))(k,l)n2(n^{-1/2}\pi(k,l))_{(k,l)\in\mathbb{Z}_{n}^{2}} constitute an orthonormal basis of n×n\mathbb{C}^{n\times n}. The second observation is

k,l,k,l:π(k,l)π(k,l)π(k,l)=e2πi(kl+kl)π(k,l).\forall k,l,k^{\prime},l^{\prime}~{}:~{}\pi(k,l)\pi(k^{\prime},l^{\prime})\pi(k,l)^{*}=e^{-2\pi i(kl^{\prime}+k^{\prime}l)}\pi(k^{\prime},l^{\prime})~{}. (6.17)

We are interested in the conjugation action ϱ\varrho, and we therefore immediately rewrite (6.17) to obtain

k,l,k,l:ϱ(k,l)(π(k,l))=e2π(lk+lk)/nπ(k,l).\forall k,l,k^{\prime},l^{\prime}~{}:~{}\varrho(k,l)\left(\pi(k^{\prime},l^{\prime})\right)=e^{-2\pi(l^{\prime}k+lk^{\prime})/n}\pi(k^{\prime},l^{\prime})~{}. (6.18)

Taking (6.16), (6.18) together, we obtain the unitary equivalence

ϱ(k,l)n2χk,l,\varrho\simeq\bigoplus_{(k^{\prime},l^{\prime})\in\mathbb{Z}_{n}^{2}}\chi_{k,l}~{},

with pairwise distinct homomorphisms χk,l:𝕋\chi_{k^{\prime},l^{\prime}}:\mathbb{H}\to\mathbb{T} defined by

χk,l(k,l,m)=e2π(lk+lk)/n.\chi_{k^{\prime},l^{\prime}}(k,l,m)=e^{-2\pi(l^{\prime}k+lk^{\prime})/n}~{}.

Hence we have decomposed ϱ\varrho into irreducible and pairwise inequivalent representations of \mathbb{H}, with the intertwining operator S:n×nn×nS:\mathbb{C}^{n\times n}\to\mathbb{C}^{n\times n} explicitly given by

SA=n1/2(A,π(k,l))(k,l)n2.SA=n^{-1/2}(\langle A,\pi(k,l)\rangle)_{(k,l)\in\mathbb{Z}_{n}^{2}}~{}.

For rank-one operators of the form A=φφA=\varphi\otimes\varphi this specializes to

SA(k,l)=n1/2φφ,π(k,l)=n1/2φ,π(k,l)φ,SA(k,l)=n^{-1/2}\langle\varphi\otimes\varphi,\pi(k,l)\rangle=n^{-1/2}\langle\varphi,\pi(k,l)\varphi\rangle~{},

which is the so-called ambiguity function AφA_{\varphi} of φ\varphi.

In order to see how this function enters into the computation of F=VφφϱAF=V_{\varphi\otimes\varphi}^{\varrho}A, for an arbitrary matrix AA, we use the intertwining property of SS to compute

F(k,l)\displaystyle F(k,l) =\displaystyle= (k,l)n1A,π(k,l)φ,π(k,l)φ¯χk,l(k,l)¯.\displaystyle\sum_{(k^{\prime},l^{\prime})}n^{-1}\langle A,\pi(k^{\prime},l^{\prime})\rangle\overline{\langle\varphi,\pi(k^{\prime},l^{\prime})\varphi\rangle}\overline{\chi_{k^{\prime},l^{\prime}}(k,l)}~{}.

Since (χk,l)k,l(\chi_{k^{\prime},l^{\prime}})_{k^{\prime},l^{\prime}} is just the set of characters of the abelian group n2\mathbb{Z}_{n}^{2}, this formula for FF is recognized as a Fourier transform over n2\mathbb{Z}_{n}^{2}. Fourier inversion then yields

A,π(k,l)φ,π(k,l)φ¯=n1(k,l)F(k,l)χk,l(k,l).\langle A,\pi(k^{\prime},l^{\prime})\rangle\overline{\langle\varphi,\pi(k^{\prime},l^{\prime})\varphi\rangle}=n^{-1}\sum_{(k,l)}F(k,l)\chi_{k^{\prime},l^{\prime}}(k,l)~{}. (6.19)

This can be solved for the expansion coefficients SA=(n1/2A,π(k,l))k,lSA=(n^{-1/2}\langle A,\pi(k^{\prime},l^{\prime})\rangle)_{k^{\prime},l^{\prime}} if (and only if) AφA_{\varphi} is nowhere vanishing. The remaining step is then to invert SS.

These observations are summarized as follows:

  1. (a)

    φ\varphi does matrix recovery iff the ambiguity function of φ\varphi is nowhere vanishing.

  2. (b)

    Assume that φ\varphi fulfills the condition from part (a). Then an arbitrary matrix An×nA\in\mathbb{C}^{n\times n} can be recovered from F:n2F:\mathbb{Z}_{n}^{2}\to\mathbb{C},

    F(k,l)=A,ϱ(k,l)(φφ)F(k,l)=\langle A,\varrho(k,l)(\varphi\otimes\varphi)\rangle

    by the following steps:

    1. (i)

      Compute F^(k,l)=(k,l)F(k,l)χk,l(k,l)\widehat{F}(k^{\prime},l^{\prime})=\sum_{(k,l)}F(k,l)\chi_{k^{\prime},l^{\prime}}(k,l).

    2. (ii)

      AA is given by

      A=(k,l)n2F^(k,l)π(k,l)φ,φπ(k,l).A=\sum_{(k^{\prime},l^{\prime})}n^{-2}\frac{\widehat{F}(k,l)}{\langle\pi(k,l)\varphi,\varphi\rangle}\pi(k^{\prime},l^{\prime})~{}.

We do not claim originality for these observations. The fact that a nonvanishing ambiguity function guarantees phase retrieval for the short-time Fourier transform is well-known. Sufficiency of criterion (a) for matrix recovery can be found in [16, Theorem 2.2], whereas necessity is implicit in [16, Theorem 2.3]. We are not aware of a previous formulation of the recovery procedure from Part (b). However, it is implicit in the proof of [16, Theorem 3.16] (although only for A=ffA=f\otimes f).

The parallels between items (a) and (b) in the previous paragraph and the corresponding steps in the program outlined in Remark 6.8 should now become obvious. Possibly the biggest obstacle to the proper understanding of this correspondence comes from the fact that the representation ϱ\varrho is trivial on the center of \mathbb{H}, i.e. it can be viewed as a representation of the abelian quotient group /()\mathbb{H}/\mathbb{Z}(\mathbb{H}). As a consequence, the decomposition of ϱ\varrho only involves characters, and all irreducible subspaces are one-dimensional, which greatly simplifies both the criterion for matrix recovery, and the inversion formula. At the same time, these simplifications somewhat obscure the role of the group Fourier transform of the (nonabelian) Heisenberg group itself in steps (a) and (b).

In fact, it can be argued with some justification that this role can be safely ignored, since all computations can be carried out over the abelian factor group. However, the slightly more involved point of view developed here allows to appreciate the wider representation-theoretic context, and a more direct comparison to the case of affine groups.

It is worthwhile to explicitly compute the generating vector for the unitarily equivalent permutation representation π0\pi_{0}, using Fourier inversion.

Corollary 6.10

Let p>2p>2 be a prime, and define ψ{0,,p1}\psi\in\mathbb{C}^{\{0,\ldots,p-1\}} by

ψ(k)=e2πik/3+2e4πik/3\psi(k)=e^{2\pi ik/3}+2e^{4\pi ik/3}

in the case p=3p=3, and

ψ(k)=δ0(k)p1p1e2πik/p\psi(k)=\delta_{0}(k)-p^{-1}-p^{-1}e^{2\pi ik/p}~{}

for p5p\geq 5. Then π0(G)ψ\pi_{0}(G)\psi does matrix recovery.

Remark 6.11

Condition (i) from Theorem 6.5 is equivalent to requiring that all diagonal operators A{1,,p1}×{1,,p1}A\in\mathbb{C}^{\{1,\ldots,p-1\}\times\{1,\ldots,p-1\}} can be recovered from their scalar products with ϱ1(k,l)(φφ)\varrho_{1}(k,l)(\varphi\otimes\varphi).

This property provides an interesting interpretation of the phase retrieval problem for the permutation representation π0\pi_{0}, acting on 0=𝟏\mathcal{H}_{0}=\mathbf{1}^{\bot}, by the observation that the diagonal operators in Fourier domain are just the convolution operators in time domain.

Hence condition (i) imposed on φ=ψ^\varphi=\widehat{\psi} is equivalent to the fact that all convolution operators AA on 0\mathcal{H}_{0} can be recovered from their scalar products Aπ(k,l)ψ,π(k,l)ψ,(k,l)G\langle A\pi(k,l)\psi,\pi(k,l)\psi\rangle,(k,l)\in G.

Now assume to be given vectors f,g0f,g\in\mathcal{H}_{0} with |Vψπ0f|=|Vψπ0g||V^{\pi_{0}}_{\psi}f|=|V^{\pi_{0}}_{\psi}g|. Using that Vψπ0f(k,l)=fψlV^{\pi_{0}}_{\psi}f(k,l)=f\ast\psi_{l}, with ψl(m)=ψ(l1m)¯\psi_{l}(m)=\overline{\psi(-l^{-1}m)}, and the analogous formula for Vψπ0gV^{\pi_{0}}_{\psi}g, together with the Plancherel formula and convolution theorem for p\mathbb{Z}_{p}, allow to compute for all lpl\in\mathbb{Z}_{p}^{\ast} that

m=1p1|f^(m)|2|ψ^(lm)|2\displaystyle\sum_{m=1}^{p-1}|\widehat{f}(m)|^{2}|\widehat{\psi}(-lm)|^{2} =\displaystyle= p1/2k=0p1|Vψπ0f(k,l)|2\displaystyle p^{-1/2}\sum_{k=0}^{p-1}|V_{\psi}^{\pi_{0}}f(k,l)|^{2}
=\displaystyle= p1/2k=0p1|Vψπ0g(k,l)|2\displaystyle p^{-1/2}\sum_{k=0}^{p-1}|V_{\psi}^{\pi_{0}}g(k,l)|^{2}
=\displaystyle= m=1p1|g^(m)|2|ψ^(lm)|2.\displaystyle\sum_{m=1}^{p-1}|\widehat{g}(m)|^{2}|\widehat{\psi}(-lm)|^{2}~{}.

Both ends of this equation are convolution products over the multiplicative group. Now condition (i) from Theorem 6.5 implies that convolution with |ψ^|2|\widehat{\psi}|^{2} is injective; compare the argument following equation (6.8). This implies

mp:|f^(m)|=|g^(m)|.\forall m\in\mathbb{Z}_{p}~{}:~{}|\widehat{f}(m)|=|\widehat{g}(m)|~{}.

With this equality established, the convolution theorem yields

|(fψl)|=p1/2|f^||ψ^l|=|(gψl)|\left|\left(f\ast\psi_{l}\right)^{\wedge}\right|=p^{1/2}|\widehat{f}|\cdot|\widehat{\psi}_{l}|=\left|\left(g\ast\psi_{l}\right)^{\wedge}\right|

This means that if we compare the fixed lines Fl=Vψπ0f(,l),Gl=Vψπ0g(,l)F_{l}=V_{\psi}^{\pi_{0}}f(\cdot,l),G_{l}=V_{\psi}^{\pi_{0}}g(\cdot,l), the assumption |Vψπ0f|=|Vψπ0g||V_{\psi}^{\pi_{0}}f|=|V_{\psi}^{\pi_{0}}g| is equivalent to the condition

l=1,,p1:|Fl|=|Gl| and |F^l|=|G^l|.\forall l=1,\ldots,p-1:|F_{l}|=|G_{l}|\mbox{ and }\left|\widehat{F}_{l}\right|=\left|\widehat{G}_{l}\right|~{}.

I.e., we get a family of so-called Pauli pairs. The study of Pauli pairs and Pauli uniqueness, i.e. the question whether ff can be recovered from |f||f| and |f^||\widehat{f}| is a phase retrieval problem in its own right, with origins in quantum physics; see e.g. [17, 16]. Note that the whole family (Fl,Gl)l=1,,p1(F_{l},G_{l})_{l=1,\ldots,p-1} of Pauli pairs is generated from a single choice of f,gf,g and ψ\psi.

Remark 6.12

Corollary 6.10 allows to derive an intriguing observation regarding phase retrieval for finite Fourier transforms. Given l{0,1,,p1}l\in\{0,1,\ldots,p-1\}, let χl(m)=e2πilmp\chi_{l}(m)=e^{2\pi ilm}\in\mathbb{C}^{p}. Define for each l{1,,p1}l\in\{1,\ldots,p-1\} the projection operator PlP_{l} onto χl0\chi_{l}^{\bot}\subset\mathcal{H}_{0}. In other words, PlP_{l} is computed from ff by omitting the llth frequency from the Fourier series of ff. Then using

ψ(k)=δ0(k)p1p1e2πik/p,ψ^=p1/2(𝟏δ0δ1)\psi(k)=\delta_{0}(k)-p^{-1}-p^{-1}e^{2\pi ik/p}~{},~{}\widehat{\psi}=p^{-1/2}\left(\mathbf{1}-\delta_{0}-\delta_{1}\right)

and ψl(k)=ψ(l1k)¯\psi_{l}(k)=\overline{\psi(-l^{-1}k)} allows to write

(fψl)=p1/2f^ψ^l=f^(𝟏δ0δl1),\left(f\ast\psi_{l}\right)^{\wedge}=p^{1/2}\widehat{f}\cdot\widehat{\psi}_{l}=\widehat{f}\cdot\left(\mathbf{1}-\delta_{0}-\delta_{l^{-1}}\right)~{},

whence

Vψπ0f(,l)=fψl=Pl1(f).V_{\psi}^{\pi_{0}}f(\cdot,l)=f\ast\psi_{l}=P_{l^{-1}}(f)~{}.

Thus the fact that ψ\psi does phase retrieval is equivalent to the following fact concerning Fourier phase retrieval:

For any pair f,g0f,g\in\mathcal{H}_{0}: |Plf|=|Plg||P_{l}f|=|P_{l}g| holds for all lpl\in\mathbb{Z}_{p}^{\ast} iff f𝕋gf\in\mathbb{T}\cdot g.

We are not aware of a previous source making this observation, nor of a more elementary proof than via Theorem 6.5. It is also unclear to us whether this observation holds for the Fourier transform over general cyclic groups, or only over those of prime order.

7 Phase retrieval for 33-fold transitive permutation actions

Recall that as a slight reformulation of the results from the previous section, we have that S(3)S(3), acting on 0\mathcal{H}_{0}, does phase retrieval. We will now use this observation to prove that every 33-fold transitive subgroup ΓS(n)\Gamma\subset S(n) does phase retrieval, when acting on 0\mathcal{H}_{0}.

We start by noting an auxiliary result which requires additional notation. Given A{0,,n1}A\subset\{0,\ldots,n-1\}, we denote

A={g0:supp(g)A},\mathcal{H}_{A}=\{g\in\mathcal{H}_{0}:{\rm supp}(g)\subset A\}~{},

and let QA:0AQ_{A}:\mathcal{H}_{0}\to\mathcal{H}_{A} denote the orthogonal projection. The following is a phase propagation result, in a spirit similar to results in [7].

Lemma 7.1

Let n3n\geq 3, and assume that f,g0f,g\in\mathcal{H}_{0} are such that for all three-element subsets A{0,,n1}A\subset\{0,\ldots,n-1\} there exists αA𝕋\alpha_{A}\in\mathbb{T} with QAg=αAQAfQ_{A}g=\alpha_{A}Q_{A}f. Then there exists α𝕋\alpha\in\mathbb{T} with g=αfg=\alpha f.

Proof 10

We proceed by induction, first noting that the case n=3n=3 is obvious. For the induction step, assume the desired statement has been established for {0,,n1}\mathbb{C}^{\{0,\ldots,n-1\}}, and let f,g0{0,,n}f,g\in\mathcal{H}_{0}\subset\mathbb{C}^{\{0,\ldots,n\}} be given, with QAg=αAQAfQ_{A}g=\alpha_{A}Q_{A}f for all 3-element subsets A{0,,n}A\subset\{0,\ldots,n\}.

W.l.o.g. ff is nonzero, and since the union of all A\mathcal{H}_{A} corresponding to three-element subsets spans 0\mathcal{H}_{0}, this implies that gg is nonzero, too. Furthermore, f0f\in\mathcal{H}_{0} implies that the support of ff contains at least two elements l,ll,l^{\prime}.

In the case where supp(f)={l,l}{\rm supp}(f)=\{l,l^{\prime}\}, the fact that QA(g)=αAQA(f)Q_{A}(g)=\alpha_{A}Q_{A}(f) holds for all 3-element sets of the type {l,l,k}\{l,l^{\prime},k\} implies g(k)=0g(k)=0 for all k{l,l}k\not\in\{l,l^{\prime}\}, and thus g=αfg=\alpha f follows.

We can therefore assume for the remainder of the proof that supp(f){\rm supp}(f) contains at least three elements, and we let B={0,,n}{l}B=\{0,\ldots,n\}\setminus\{l\} and B={0,,n}{l}B^{\prime}=\{0,\ldots,n\}\setminus\{l^{\prime}\}.

Now the induction hypothesis, applied with B,B\mathcal{H}_{B},\mathcal{H}_{B^{\prime}} and all three-element subsets ABA\subset B and ABA^{\prime}\subset B^{\prime}, respectively, provide us with αB,αB𝕋\alpha_{B},\alpha_{B^{\prime}}\in\mathbb{T} such that

QBg=αBQBf,QBg=αBQBf.Q_{B}g=\alpha_{B}Q_{B}f~{},~{}Q_{B^{\prime}}g=\alpha_{B^{\prime}}Q_{B^{\prime}}f~{}.

The fact that {l,l}supp(f)\{l,l^{\prime}\}\subsetneq{\rm supp}(f) implies that QBQBf=QBBf0Q_{B}Q_{B^{\prime}}f=Q_{B^{\prime}\cap B}f\not=0. But then we get on the one hand

QBQBg=αBQBQBf=αBQBBfQ_{B^{\prime}}Q_{B}g=\alpha_{B}Q_{B^{\prime}}Q_{B}f=\alpha_{B}Q_{B^{\prime}\cap B}f

and on the other

QBQBg=QBQBg=αBQBBf,Q_{B^{\prime}}Q_{B}g=Q_{B}Q_{B^{\prime}}g=\alpha_{B^{\prime}}Q_{B^{\prime}\cap B}f~{},

leading to

αBQBBf=αBQBBf,\alpha_{B^{\prime}}Q_{B^{\prime}\cap B}f=\alpha_{B}Q_{B^{\prime}\cap B}f~{},

with nonzero QBBfQ_{B^{\prime}\cap B}f. This finally entails αB=αB=:α\alpha_{B^{\prime}}=\alpha_{B}=:\alpha. Now the fact that BB\mathcal{H}_{B}\cup\mathcal{H}_{B^{\prime}} spans all of 0\mathcal{H}_{0} implies

g=αf.g=\alpha f~{}.
Theorem 7.2

Let Γ<S(3)\Gamma<S(3) denote a 33-fold transitive permutation group. Pick ψ0{0,1,2}\psi_{0}\in\mathbb{C}^{\{0,1,2\}} doing phase retrieval for 0\mathcal{H}_{0} under the action of S(3)S(3), and define ψ{0,,n1}\psi\in\mathbb{C}^{\{0,\ldots,n-1\}} as trivial extension of ψ0\psi_{0}. Then Π(Γ)ψ\Pi(\Gamma)\psi does phase retrieval on 0\mathcal{H}_{0}.

Proof 11

Fix any three-element subset A{0,,n1}A\subset\{0,\ldots,n-1\}, and consider the set

{Π(h)ψ:hΓ,Π(h)ψA}.\{\Pi(h)\psi~{}:~{}h\in\Gamma,\Pi(h)\psi\in\mathcal{H}_{A}\}~{}.

By a suitable reindexing of AA by {0,1,2}\{0,1,2\}, this set corresponds to the system π(S(3))ψ0\pi(S(3))\psi_{0} which was chosen to do phase retrieval on the subspace {0,1,2}\mathbb{C}^{\{0,1,2\}} perpendicular to the constant signal. It follows for all f,g0f,g\in\mathcal{H}_{0} satisfying |Vψf|=|Vψg||V_{\psi}f|=|V_{\psi}g|, and for all three-element subsets A{0,1,,n1}A\subset\{0,1,\ldots,n-1\}, that QAg=αAQAfQ_{A}g=\alpha_{A}Q_{A}f holds, with αA𝕋\alpha_{A}\in\mathbb{T}. Hence the assumptions of the previous lemma are fulfilled, and f=αgf=\alpha g follows.

References

  • [1] R. Alaifari, F. Bartolucci, and M. Wellershoff. Phase retrieval of bandlimited functions for the wavelet transform, 2020.
  • [2] R. Balan, B. G. Bodmann, P. G. Casazza, and D. Edidin. Painless reconstruction from magnitudes of frame coefficients. J. Fourier Anal. Appl., 15(4):488–501, 2009.
  • [3] R. Balan, P. Casazza, and D. Edidin. On signal reconstruction without phase. Appl. Comput. Harmon. Anal., 20(3):345–356, 2006.
  • [4] A. S. Bandeira, J. Cahill, D. G. Mixon, and A. A. Nelson. Saving phase: injectivity and stability for phase retrieval. Appl. Comput. Harmon. Anal., 37(1):106–125, 2014.
  • [5] I. Bojarovska and A. Flinth. Phase retrieval from Gabor measurements. J. Fourier Anal. Appl., 22(3):542–567, 2016.
  • [6] E. J. Candès, Y. C. Eldar, T. Strohmer, and V. Voroninski. Phase retrieval via matrix completion. SIAM J. Imaging Sci., 6(1):199–225, 2013.
  • [7] C. Cheng, I. Daubechies, N. Dym, and J. Lu. Stable phase retrieval from locally stable and conditionally connected measurements, 2020.
  • [8] C. Cheng and D. Han. On twisted group frames. Linear Algebra Appl., 569:285–310, 2019.
  • [9] A. Conca, D. Edidin, M. Hering, and C. Vinzant. An algebraic characterization of injectivity in phase retrieval. Appl. Comput. Harmon. Anal., 38(2):346–356, 2015.
  • [10] Y. C. Eldar and S. Mendelson. Phase retrieval: stability and recovery guarantees. Appl. Comput. Harmon. Anal., 36(3):473–494, 2014.
  • [11] L. Evans and C.-K. Lai. Conjugate phase retrieval on M\mathbb{C}^{M} by real vectors. Linear Algebra Appl., 587:45–69, 2020.
  • [12] A. Fannjiang and T. Strohmer. The numerics of phase retrieval. Acta Numer., 29:125–228, 2020.
  • [13] K. Flornes, A. Grossmann, M. Holschneider, and B. Torrésani. Wavelets on discrete fields. Appl. Comput. Harmon. Anal., 1(2):137–146, 1994.
  • [14] H. Führ. Abstract harmonic analysis of continuous wavelet transforms, volume 1863 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2005.
  • [15] H. Führ and V. Oussa. Phase retrieval for nilpotent groups. 2022.
  • [16] P. Grohs, S. Koppensteiner, and M. Rathmair. Phase retrieval: uniqueness and stability. SIAM Rev., 62(2):301–350, 2020.
  • [17] P. Jaming. Phase retrieval techniques for radar ambiguity problems. J. Fourier Anal. Appl., 5(4):309–329, 1999.
  • [18] R. Kueng, H. Rauhut, and U. Terstiege. Low rank matrix recovery from rank one measurements. Appl. Comput. Harmon. Anal., 42(1):88–116, 2017.
  • [19] L. Li, T. Juste, J. Brennan, C. Cheng, and D. Han. Phase retrievable projective representation frames for finite abelian groups. J. Fourier Anal. Appl., 25(1):86–100, 2019.
  • [20] R. D. Malikiosis and V. Oussa. Full spark frames in the orbit of a representation. Appl. Comput. Harmon. Anal., 49(3):791–814, 2020.
  • [21] S. Mallat and I. Waldspurger. Phase retrieval for the Cauchy wavelet transform. J. Fourier Anal. Appl., 21(6):1251–1309, 2015.
  • [22] J. Schwinger. Unitary operator bases. Proc. Nat. Acad. Sci. U.S.A., 46:570–579, 1960.
  • [23] A. Terras. Fourier analysis on finite groups and applications, volume 43 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 1999.
  • [24] S. Waldron. Group frames. In Finite frames, Appl. Numer. Harmon. Anal., pages 171–191. Birkhäuser/Springer, New York, 2013.
  • [25] W. K. Wootters and B. D. Fields. Optimal state-determination by mutually unbiased measurements. Ann. Physics, 191(2):363–381, 1989.