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

Uniqueness of phase retrieval from three measurements

Philippe Jaming philippe.jaming@math.u-bordeaux.fr  and  Martin Rathmair martin.rathmair@math.u-bordeaux.fr Univ. Bordeaux, IMB, UMR 5251, F-33400 Talence, France. CNRS, IMB, UMR 5251, F-33400 Talence, France.
Abstract.

In this paper we consider the question of finding an as small as possible family of operators (Tj)jJ(T_{j})_{j\in J} on L2()L^{2}({\mathbb{R}}) that does phase retrieval: every φ\varphi is uniquely determined (up to a constant phase factor) by the phaseless data (|Tjφ|)jJ(|T_{j}\varphi|)_{j\in J}. This problem arises in various fields of applied sciences where usually the operators obey further restrictions.

Of particular interest here are so-called coded diffraction paterns where the operators are of the form Tjφ=[mjf]T_{j}\varphi={\mathcal{F}}[m_{j}f], {\mathcal{F}} the Fourier transform and mjL()m_{j}\in L^{\infty}({\mathbb{R}}) are “masks”. Here we explicitely construct three real-valued masks m1,m2,m3L()m_{1},m_{2},m_{3}\in L^{\infty}({\mathbb{R}}) so that the associated coded diffraction patterns do phase retrieval. This implies that the three self-adjoint operators Tjφ=[mj1φ]T_{j}\varphi={\mathcal{F}}[m_{j}{\mathcal{F}}^{-1}\varphi] also do phase retrieval. The proof uses complex analysis.

We then show that some natural analogues of these operators in the finite dimensional setting do not always lead to the same uniqueness result due to an undersampling effect.

Key words and phrases:
phase retrieval; wright’s conjecture, holomorphic functions
1991 Mathematics Subject Classification:
30D05, 42B10, 94A12

1. Introduction

Generally speaking, phase retrieval refers to the problem of recovering a signal from phaseless linear measurements. Typical instances of phase retrieval tasks include the question of recovering a function from the magnitude of its Fourier transform or a variant therof. Such problems arise in various areas of natural sciences ranging from signal processing to quantum mechanics. This family of problems has recently attracted a lot of attention in the mathematical community, and we refer e.g. to [3, 15, 16, 20, 25, 32] for an overview of some recent developments, as well as for references to concrete problems.

1.1. Problem setting

Within this article we will predominantly deal with signals ff of one real variable, i.e. fL2()f\in L^{2}(\mathbb{R}). The phase retrieval problems we shall consider will be associated to a given family of operators.

Definition 1.1.

Let 𝒯=(Tj)jJ\mathcal{T}=(T_{j})_{j\in J} be a family of linear operators on L2()L^{2}(\mathbb{R}), i.e. Tj:L2()ΩjT_{j}:L^{2}(\mathbb{R})\rightarrow\mathbb{C}^{\Omega_{j}} linear. We say that 𝒯\mathcal{T} does phase retrieval if

ϕ,ψL2():|Tjψ|=|Tjϕ|,jJc𝕋:s.t.ψ=cϕ\phi,\psi\in L^{2}(\mathbb{R}):~{}|T_{j}\psi|=|T_{j}\phi|,~{}j\in J\quad\Rightarrow\quad\exists c\in\mathbb{T}:~{}\text{s.t.}~{}\psi=c\phi

with 𝕋\mathbb{T} the set of complex numbers of modulus 11.

Clearly, due to the linearity of the operators, ϕ\phi and cϕc\phi produce the same phaseless measurements when |c|=1|c|=1. Thus, the notion of uniqueness introduced in Definition 1.1 is the best one can hope for.
In practice, an arbitrary linear operator TT will in general not represent an attainable measurement. Moreover, measurements may be costly resulting in natural restrictions on the number of operators employed. To put it casually, the objective in this article is to identify families 𝒯\mathcal{T} which do phase retrieval subject to the two side constraints

  1. i)

    the operators represent physically meaningful objects and

  2. ii)

    use as few operators as possible.

Constraint i) is admittedly very vaguely phrased, and depends strongly on the concrete application one has in mind. In the subsequent paragraph, Section 1.2 we provide some possible physical context and concretize the question further.
To summarize, formally the task we are confronted with is the following.

Given a set 𝒜\mathcal{A} of admissible linear operators on L2()L^{2}(\mathbb{R}), find a family of operators 𝒯𝒜\mathcal{T}\subseteq\mathcal{A} which does phase retrieval and which is as small as possible.

1.2. Motivation

Next we briefly discuss two important applications in physics where the problem of lost phase information appears. Both of these instances naturally fit into the problem formulation outlined above.

1.2.1. Diffractive Imaging

Perhaps the most prominent example of a phase retrieval problem arises in diffraction imaging, where one seeks to determine an unknown object represented by fL2()f\in L^{2}(\mathbb{R}) given its so-called diffraction pattern, which is represented by |f^||\hat{f}|, the modulus of its Fourier transform. We refer to [12] for the derivation of this model from physical considerations. Here and in the remainder we normalize the Fourier transform according to

f^(ξ)=f(ξ):=f(x)e2πixξ𝑑x,ξ,\hat{f}(\xi)={\mathcal{F}}f(\xi):=\int_{\mathbb{R}}f(x)e^{-2\pi ix\xi}\,dx,\quad\xi\in\mathbb{R},

for fL1()f\in L^{1}(\mathbb{R}), and extend the definition to L2()L^{2}(\mathbb{R}) in the usual way.
The mapping f|f^|f\mapsto|\hat{f}| is far from injective: Clearly, for an arbitrary measurable phase function φ:\varphi:\mathbb{R}\rightarrow\mathbb{R} we get that fφ:=1[eiφf^]f_{\varphi}:={\mathcal{F}}^{-1}[e^{i\varphi}\hat{f}] has the same Fourier modulus as ff.

There are two rather obvious strategies to overcome these issues of non-uniqueness:

Restrict the signal space: A popular constraint is to assume the signal under consideration to be compactly supported. The problem then amounts to determining a band-limited function from its modulus. However, this additional assumption is known not to render the problem unique. To start with, it is still possible to modulate and conjugate the function (this are called trivial solutions) and more ambiguous solutions can be constructed by employing what is known as the “zero-flipping” operation. For details we refer to the articles of Akutowicz[1], Walther[35] and Hofstetter[17]. In particular, given a compactly supported fL2()f\in L^{2}(\mathbb{R}) there is in general a huge (uncountably infinite!) set of non-equivalent ambiguous solutions, all of which have compact support. Corresponding results hold true in the context of wide-banded signals, see [23].

Collect several diffraction patterns: The idea of this approach is to accumulate more information by acquiring several diffraction patterns making use of so-called masks. In our setup, a mask would then be a function γL()\gamma\in L^{\infty}(\mathbb{R}), which interacts multiplicatively with the unknown signal ff before computing its diffraction pattern |[γf]||{\mathcal{F}}[\gamma\cdot f]|. Measurements acquired in this manner are also known as coded-diffraction patterns (see e.g. [6, 16]). Therefore, in this particular context of diffraction imaging it appears natural to define the set of admissible operators by

(1) 𝒜DI={mγ:γL()},\mathcal{A}_{DI}=\left\{{\mathcal{F}}\circ m_{\gamma}:~{}\gamma\in L^{\infty}(\mathbb{R})\right\},

where mγ(f):=γfm_{\gamma}(f):=\gamma\cdot f denotes the multiplication operator, and seek for a (small) family of operators (Tj)jJ𝒜DI(T_{j})_{j\in J}\subseteq\mathcal{A}_{DI} with corresponding masks (γj)jJ(\gamma_{j})_{j\in J} and which does phase retrieval.

1.2.2. Quantum Mechanics

A second motivation for looking at this kind of problem comes from quantum mechanics, and in particular from a question stemming back from the work of W. Pauli. The aim here is to formulate everything in terms of a mathematical language. However, we dedicate the Appendix to bridging the gap between the physics literature and our formulations.

In a footnote to the Handbuch der Physik article on the general principle of wave mechanics [30], W. Pauli asked whether a wave function ψL2()\psi\in L^{2}(\mathbb{R}) is uniquely determined (up to a constant phase factor) by the pair (|ψ|,|ψ^|)(|\psi|,|\widehat{\psi}|), which is sometimes called the Pauli data. In our terms, the question Pauli posed amounts to

Does (Id,)(\operatorname{Id},{\mathcal{F}}) do phase retrieval?

The first counter-example seems to be due to Bargmann who considered the following simple example based on complex Gaussians:

Example 1.2.

Let ψ±(x)=e(1±i)πx2\psi_{\pm}(x)=e^{-(1\pm i)\pi x^{2}} so that ψ±^(ξ)=21/4eiπ/8e(1i)πξ2/2\widehat{\psi_{\pm}}(\xi)=2^{-1/4}e^{\mp i\pi/8}e^{-(1\mp i)\pi\xi^{2}/2}. It follows that ψ+=ψ¯\psi_{+}=\overline{\psi_{-}} and ψ+^=ψ^¯\widehat{\psi_{+}}=\overline{\widehat{\psi_{-}}} so that |ψ+|=|ψ||\psi_{+}|=|\psi_{-}| and |ψ+^|=|ψ^||\widehat{\psi_{+}}|=|\widehat{\psi_{-}}|.

However, we may introduce slightly less restrictive notions of equivalence, such as the following two:

– two states φ,ψ\varphi,\psi are equivalent up to a constant phase factor and conjugation, if there exists λ\lambda\in{\mathbb{C}} with |λ|=1|\lambda|=1 such that either φ=λψ\varphi=\lambda\psi or φ=λψ¯\varphi=\lambda\overline{\psi}

– two states φ,ψ\varphi,\psi are equivalent up to a constant phase factor and conjugate reflection, if there exists λ\lambda\in{\mathbb{C}} with |λ|=1|\lambda|=1 such that either φ=λψ\varphi=\lambda\psi or φ=λψ\varphi=\lambda\psi^{*} where ψ(x)=ψ(x)¯\psi^{*}(x)=\overline{\psi(-x)}.

In Bargmann’s example, the states ψ±\psi_{\pm} are still equivalent up to a constant phase factor and conjugation (or conjugate reflection). One may then ask for a class of states 𝒞L2(){\mathcal{C}}\subset L^{2}({\mathbb{R}}) such that φ,ψ𝒞\varphi,\psi\in{\mathcal{C}} with same Pauli data are necessarily equivalent (eventually up to a conjugation and reflection as well). This problem has attracted some attention over the years and more evolved counter-examples have been found, (see [9, 10, 19, 21, 24, 34] to name a few) some which are still equivalent up to a constant phase factor and conjugation or conjugate reflection, and some which are not equivalent in this less restrictive form.

From a quantum mechanical perspective it is natural to consider unitary operators as admissible. We once more refer to the Appendix where we elaborate on why this is a natural choice. To be more precise, we define

(2) 𝒜QM={T:L2()L2(,μ)unitary,withμa Borel measure}.\mathcal{A}_{QM}=\left\{T:L^{2}(\mathbb{R})\rightarrow L^{2}(\mathbb{R},\mu)~{}\text{unitary},~{}\text{with}~{}\mu~{}\text{a Borel measure}\right\}.

Obviously, Id\operatorname{Id} and {\mathcal{F}} belong to this class of operators. A natural extension to Pauli’s question is the following conjecture attributed to R. Wright (based on a degree of freedom argument) which is mentioned in [34]:

Conjecture 1.3 (R. Wright).

There exists a unitary operator T𝒜QMT\in\mathcal{A}_{QM} such that (Id,,T)(\operatorname{Id},{\mathcal{F}},T) does phase retrieval.

To the best of our knowledge, both Wright’s conjecture as well as the following relaxed version remain open up to this point in time.

Conjecture 1.4.

There exist T1,T2,T3𝒜QMT_{1},T_{2},T_{3}\in\mathcal{A}_{QM} such that (T1,T2,T3)(T_{1},T_{2},T_{3}) does phase retrieval.

If the restriction on the number of operators in Wright’s conjecture is dropped we can give a positive answer: For instance, we may consider the fractional Fourier transform defined as follows: for απ\alpha\in{\mathbb{R}}\setminus\pi{\mathbb{Z}}, let cα=expi2(απ2)|sinα|c_{\alpha}=\displaystyle\frac{\exp\frac{i}{2}\left(\alpha-\frac{\pi}{2}\right)}{\sqrt{|\sin\alpha|}} be a square root of 1icotα1-i\cot\alpha. For uL1()u\in L^{1}({\mathbb{R}}) and απ\alpha\notin\pi{\mathbb{Z}}, define

(3) αu(ξ)\displaystyle{\mathcal{F}}_{\alpha}u(\xi) =\displaystyle= cαeiπ|ξ|2cotαu(t)eiπ|t|2cotαe2iπtξ/sinαdt\displaystyle c_{\alpha}e^{-i\pi|\xi|^{2}\cot\alpha}\int_{{\mathbb{R}}}u(t)e^{-i\pi|t|^{2}\cot\alpha}e^{-2i\pi t\xi/\sin\alpha}\mbox{d}t
=\displaystyle= cαeiπ|ξ|2cotα[u(t)eiπ|t|2cotα](ξ/sinα).\displaystyle c_{\alpha}e^{-i\pi|\xi|^{2}\cot\alpha}{\mathcal{F}}[u(t)e^{-i\pi|t|^{2}\cot\alpha}](\xi/\sin\alpha).

The last expression shows that α{\mathcal{F}}_{\alpha} extends to a bounded operator on L2()L^{2}({\mathbb{R}}). Further π/2={\mathcal{F}}_{\pi/2}={\mathcal{F}} the usual Fourier transform and we define 2kπ=I{\mathcal{F}}_{2k\pi}=I the identity operator and (2k+1)πf(x)=f(x){\mathcal{F}}_{(2k+1)\pi}f(x)=f(-x) then α+β=αβ{\mathcal{F}}_{\alpha+\beta}={\mathcal{F}}_{\alpha}{\mathcal{F}}_{\beta}. Finally, α{\mathcal{F}}_{\alpha} is a unitary operator with α=α{\mathcal{F}}_{\alpha}^{*}={\mathcal{F}}_{-\alpha}. One of the authors showed that (α)α[0,2π)({\mathcal{F}}_{\alpha})_{\alpha\in[0,2\pi)} does phase retrieval [22, Proposition 4.2].
An equivalent formulation is that if for every time t0t\geq 0, the free Schrödinger evolution of φ,ψ\varphi,\psi have same modulus |eitΔφ|=|eitΔψ||e^{it\Delta}\varphi|=|e^{it\Delta}\psi| then φ,ψ\varphi,\psi are equivalent up to a constant phase factor (this was conjectured in [34]).

Proceeding from Conjecture 1.4 one may replace the constraint on the operators to be unitary by assuming them to be self-adjoint, and ask

Is there a triple of self-adjoint operators (T1,T2,T3)(T_{1},T_{2},T_{3}) on L2()L^{2}(\mathbb{R}) which does phase retrieval?

Again, the point is that we want a triple of self-adjoint operators. For instance, Vogt [34] stated (without proof) that the set of all rank one orthogonal projections does phase retrieval. An even smaller set of rank one projections is sufficient. One may for instance take an orthonormal basis (ei)i(e_{i})_{i\in{\mathbb{N}}} of L2()L^{2}(\mathbb{R}) and then consider the rank one projections on the spaces Span(ek)\mathrm{Span}(e_{k}), kk\in{\mathbb{N}}, Span(ek+e)\mathrm{Span}(e_{k}+e_{\ell}), and Span(ek+ie)\mathrm{Span}(e_{k}+ie_{\ell}) kk\not=\ell\in{\mathbb{N}}. It is then easy to show that the family of associated orthogonal projections does phase retrieval.

On the other hand, shifting the focus towards the minimality of the employed operator family, without requiring self-adjointness of the operators, the follwing was shown by one of the authors. Take γ=eπ2\gamma=e^{-\pi\cdot^{2}} the Gaussian and απ\alpha\in\mathbb{R}\setminus\pi\mathbb{Q}, then the pair T1ψ=γψT_{1}\psi=\gamma\ast\psi and T2ψ=γαψT_{2}\psi=\gamma\ast{\mathcal{F}}_{\alpha}\psi does phase retrieval [22, Proposition 4.1]. Note that while T1T_{1} is self-adjoint, T2T_{2} is not.

Remark 1.5.

Wright’s conjecture has also attracted considerable amount of interest in the finite dimensional setting. Translating Conjecture 1.4 to d\mathbb{C}^{d} amounts to asking whether there exist three orthonormal bases (ek(1))k=1d,(ek(2))k=1d(e_{k}^{(1)})_{k=1}^{d},(e_{k}^{(2)})_{k=1}^{d} and (ek(3))k=1d(e_{k}^{(3)})_{k=1}^{d}, such that each and every vector ψd\psi\in\mathbb{C}^{d} is uniquely determined (up to multiplication by a unimodular constant) by the measurements

|ek(j),ψ|,k=1,,d,j=1,2,3.|\langle e_{k}^{(j)},\psi\rangle|,\quad k=1,\ldots,d,~{}j=1,2,3.

This finite dimensional version has been disproved by Morov and Perelomov [29] in the early 90s. Further, one may relax the constraints and ask for a set of vectors (ek)k=1N(e_{k})_{k=1}^{N} such that |ek,ψ||{\left\langle{e_{k},\psi}\right\rangle}|, k=1,,Nk=1,\ldots,N uniquely determines ψ\psi up to a unimodular constant. Heinosaari, Mazzarella and Wolf [18] proved that the minimal number of vectors is 3d+αd\geq 3d+\alpha_{d} with αd+\alpha_{d}\to+\infty when d+d\to+\infty.
On the other hand, Mondragon and Voroninski [26]111This paper has not appeared yet. However, a construction somewhat similar to our argument for rank-one projections gives an explicit family of 55 unitaries that lead to uniqueness up to a constant phase factor [13]. proved that for four “generic” orthonormal bases are enough to determine all ψ\psi up to a constant phase factor.

1.3. Contribution of this paper

The purpose of this paper is to show that there are three simple and explicit masks such that the resulting coded diffraction patterns uniquely determine all univariate signals. More precisely we will show the following.

Theorem 1.6.

Let γ1=γ\gamma_{1}=\gamma be the standard Gaussian, γ(t)=eπt2\gamma(t)=e^{-\pi t^{2}}, and let γ2,γ3\gamma_{2},\gamma_{3} be defined by

γ2(t):=2πtγ1(t),γ3(t):=(12πt)γ1(t).\gamma_{2}(t):=2\pi t\gamma_{1}(t),\qquad\gamma_{3}(t):=(1-2\pi t)\gamma_{1}(t).
  1. (i)

    Let φ,ψL2()\varphi,\psi\in L^{2}({\mathbb{R}}) be such that |[γ1φ]|=|[γ1ψ]||{\mathcal{F}}[\gamma_{1}\varphi]|=|{\mathcal{F}}[\gamma_{1}\psi]| and |[γ2φ]|=|[γ2ψ]||{\mathcal{F}}[\gamma_{2}\varphi]|=|{\mathcal{F}}[\gamma_{2}\psi]| then φ\varphi and ψ\psi are equivalent up to a constant phase factor and conjugation-reflection: ψ=cφ\psi=c\varphi or ψ=cφ\psi=c\varphi^{*} with |c|=1|c|=1;

  2. (ii)

    if we further assume that |[γ3φ]|=|[γ3ψ]||{\mathcal{F}}[\gamma_{3}\varphi]|=|{\mathcal{F}}[\gamma_{3}\psi]| then φ\varphi and ψ\psi are equivalent up to a constant phase factor; in other words, (mγk)k=13({\mathcal{F}}\circ m_{\gamma_{k}})_{k=1}^{3} does phase retrieval.

Remark 1.7.

The actual result can be extended in multiple ways. For instance, the function γ1\gamma_{1} can be replaced by ea|x|e^{-a|x|}. We will also provide a second set of 3 operators that does phase retrieval. Finally, we will also show that the result can be extended to L2(d)L^{2}({\mathbb{R}}^{d}) where we need 2d+12d+1 operators.

Remark 1.8.

Note that the three masks γ1,γ2,γ3\gamma_{1},\gamma_{2},\gamma_{3} are real-valued, and consequently that Ak=mγkA_{k}={\mathcal{F}}\circ m_{\gamma_{k}}\circ{\mathcal{F}}^{\ast}, k=1,2,3k=1,2,3 define self-adjoint operators on L2()L^{2}(\mathbb{R}). It follows directly from Theorem 1.6 that (A1,A2,A3)(A_{1},A_{2},A_{3}) does phase retrieval, hence we simultaneously solve the question posed earlier in Section 1.2.

This result could be deduced from a result by McDonald [27]: the main result of that paper can be summarized as the identity and the derivation operator do phase retrieval (up to reflections) when restricted to band-limited or even to narrow-banded functions. We will however give a more direct proof and deduce our result from a bit more general facts. There are two possible strategies of proof. We could first establish (i) and then deduce (ii) from it. It turns out that this can be done in a more direct way using a simple lemma about analytic functions (Lemma 2.1). The proof of (i) is a bit more evolved and uses a lemma from the second author. Deducing (ii) from it follows essentially the same lines as the ones used to directly establishing (ii).

In a second part of this paper, we will move to the discrete setting. The operators we consider have natural discrete analogues. More precisely, we will identify ψd\psi\in{\mathbb{C}}^{d} with an analytic trigonometric polynomial PψP_{\psi}. The measurements we consider are then samples of |Pψ||P_{\psi}| and of |Pψ||P_{\psi}^{\prime}| the modulus of the derivative of PψP_{\psi}. We will show that this requires 4d24d-2 samples to lead to uniqueness (up to a constant phase factor) and provide an example of non uniqueness with less samples. This is of course coherent with the fact mentionned above that 3d3d phaseless measurements are not sufficient. However, it allows to show the role of the sampling rate and explains why 3d3d phaseless measurements may not have been the right analogue of Wright’s conjecture in d{\mathbb{C}}^{d}.

The remainder of this paper is organized as follows: the next section is devoted to the continuous setting, followed by a section devoted to the discrete case. We conclude with an appendix to clarify the role of unitaries in Wright’s conjecture, mainly aimed to mathematicians without background on quantum mechanics.

2. Continuous Level

2.1. Three Measurements

We begin with an auxiliary result which provides us with a uniqueness statement.

Lemma 2.1.

Let II\subseteq\mathbb{R} be an open interval and let 𝒜(I)\mathcal{A}(I) denote the space of complex-valued analytic functions on II. Then F𝒜(I)F\in\mathcal{A}(I) is uniquely determined (up to multiplication by a unimodular constant) by |F|2|F|^{2} and FF¯F^{\prime}\bar{F}.

Proof.

Suppose that F,G𝒜(I)F,G\in\mathcal{A}(I) are such that |F|2=|G|2|F|^{2}=|G|^{2} and FF¯=GG¯F^{\prime}\bar{F}=G^{\prime}\bar{G}. We may assume w.l.o.g. that |F|2|F|^{2} does not vanish identically. Therefore there exists a nonempty interval III^{\prime}\subseteq I such that |F|2=|G|2|F|^{2}=|G|^{2} has no zeros in II^{\prime}. Moreover, according to the assumption we have that

(log(G/F))=(logG)(logF)=GG¯GG¯FF¯FF¯=0,(\log(G/F))^{\prime}=(\log G)^{\prime}-(\log F)^{\prime}=\frac{G^{\prime}\overline{G}}{G\overline{G}}-\frac{F^{\prime}\overline{F}}{F\overline{F}}=0,

i.e. log(G/F)\log(G/F) is constant on II^{\prime}. This implies that G=λFG=\lambda F on II^{\prime} for some λ\lambda\in\mathbb{C}. Since |G|2=|F|2|G|^{2}=|F|^{2} we get that λ\lambda must be unimodular. Finally, by analyticity the identity G=λFG=\lambda F extends to all of II. ∎

We are now in position to prove the second part of the theorem:

Proposition 2.2.

Let γ1=γ\gamma_{1}=\gamma be the standard Gaussian and let γ2,γ3\gamma_{2},\gamma_{3} be defined by

γ2(t):=2πtγ(t),γ3(t):=(12πt)γ(t).\gamma_{2}(t):=2\pi t\gamma(t),\qquad\gamma_{3}(t):=(1-2\pi t)\gamma(t).

Let φ,ψL2()\varphi,\psi\in L^{2}({\mathbb{R}}) be two wave functions such that |[γjφ]|=|[γjψ]||{\mathcal{F}}[\gamma_{j}\varphi]|=|{\mathcal{F}}[\gamma_{j}\psi]| for j=1,2j=1,2 and 33. Then φ\varphi and ψ\psi are equivalent up to a unimodular constant only.

Proof.

First we define a pair of analytic function on the real line by F:=[γ1φ]F:={\mathcal{F}}[\gamma_{1}\varphi] and G:=[γ1ψ]G:={\mathcal{F}}[\gamma_{1}\psi].

It is enough to show that |[γjφ]|=|[γjψ]||{\mathcal{F}}[\gamma_{j}\varphi]|=|{\mathcal{F}}[\gamma_{j}\psi]| for j=1,2,3j=1,2,3 implies that F=λGF=\lambda G with |λ|=1|\lambda|=1 since then γ1φ=λγ1ψ\gamma_{1}\varphi=\lambda\gamma_{1}\psi. Then, as γ1\gamma_{1} does not vanish, we get that φ\varphi and ψ\psi are equivalent up to a constant phase factor.

Now note that F=i[γ2φ]F^{\prime}=i{\mathcal{F}}[\gamma_{2}\varphi], and therefore that we have the identities

|F|=|[γ1φ]|,|F|=|[γ2φ]|,|F+iF|=|[γ3φ]|.|F|=|{\mathcal{F}}[\gamma_{1}\varphi]|,\qquad|F^{\prime}|=|{\mathcal{F}}[\gamma_{2}\varphi]|,\qquad|F+iF^{\prime}|=|{\mathcal{F}}[\gamma_{3}\varphi]|.

Thus, it remains to show that a function FF analytic on the real line, is uniquely determined given |F|,|F||F|,|F^{\prime}| and |F+iF||F+iF^{\prime}|. To see this, first compute

|F+iF|2=|F|2+|F|2+2Re{iFF¯}=|F|2+|F|22Im{FF¯}.|F+iF^{\prime}|^{2}=|F|^{2}+|F^{\prime}|^{2}+2\operatorname{Re}\left\{iF^{\prime}\bar{F}\right\}=|F|^{2}+|F^{\prime}|^{2}-2\operatorname{Im}\left\{F^{\prime}\bar{F}\right\}.

Together with

Re{FF¯}=12(FF¯+F¯F)=12(|F|2)\operatorname{Re}\left\{F^{\prime}\bar{F}\right\}=\frac{1}{2}\left(F^{\prime}\bar{F}+\overline{F^{\prime}}F\right)=\frac{1}{2}\left(|F|^{2}\right)^{\prime}

we get that

(4) FF¯=12(|F|2)+i2(|F|2+|F|2|F+iF|2).F^{\prime}\bar{F}=\frac{1}{2}\left(|F|^{2}\right)^{\prime}+\frac{i}{2}\left(|F|^{2}+|F^{\prime}|^{2}-|F+iF^{\prime}|^{2}\right).

It follows that |F|=|G||F|=|G| and that FF¯=GG¯F^{\prime}\overline{F}=G^{\prime}\overline{G}. Applying Lemma 2.1 yields the desired statement. ∎

Remark 2.3.

In the next section, we are going to prove that |[γjφ]|=|[γjψ]||{\mathcal{F}}[\gamma_{j}\varphi]|=|{\mathcal{F}}[\gamma_{j}\psi]| for j=1,2j=1,2 implies that φ=cψ\varphi=c\psi or φ=cψ\varphi=c\psi^{*}. In this last case G=cF¯G=c\overline{F}. But then |γ3φ|=|γ3ψ||\gamma_{3}\varphi|=|\gamma_{3}\psi| reads |F+iF|2=|F¯+iF¯|2|F+iF^{\prime}|^{2}=|\bar{F}+i\bar{F}^{\prime}|^{2} which implies that Im{FF¯}=0\operatorname{Im}\left\{F^{\prime}\bar{F}\right\}=0. But then (4) simplifies to FF¯=12(|F|2)=GG¯F^{\prime}\bar{F}=\frac{1}{2}\left(|F|^{2}\right)^{\prime}=G^{\prime}\bar{G}. Again lemma 2.1 yields the desired statement.

The direct proof given here is substentially simpler.

Remark 2.4.

The Gaussian γ1\gamma_{1} only plays a mild role here:

– it implies that F=[fγ1]F={\mathcal{F}}[f\gamma_{1}] is holomorphic in a neighborhood of the real line so that we may replace γ1\gamma_{1} by any function that is O(ea|x|)O(e^{-a|x|}) for some a>0a>0,

γ1\gamma_{1} does not vanish on a set of positive measure so that ff is uniquely determined by fγ1f\gamma_{1}.

This shows that we could replace γ1\gamma_{1} by e.g ea|x|αe^{-a|x|^{\alpha}}, a>0a>0, α1\alpha\geq 1.

2.2. Two Measurements

We are now going to prove Theorem 1.6 (i). The origin of our choice for the three operators comes from the work of Mc Donald [27] who characterized entire functions of finite order F,GF,G such that |F|=|G||F|=|G| on the real line and |F|=|G||F^{\prime}|=|G^{\prime}|. Once one notices that F=[γ1ψ]F={\mathcal{F}}[\gamma_{1}\psi] is entire of order 2, Mc Donald’s result applies directly. We here propose another strategy of proof that does not use the growth properties of FF. In fact, our arguments do not even require that the functions under consideration are entire only that they are holomorphic in a neighborhood of the real line.

Lemma 2.5.

Let DD\subseteq\mathbb{C} be a nonempty, open disk centered on the real line. Let u,vu,v be two smooth real valued functions such that h(z):=u(x,y)+iv(x,y)h(z):=u(x,y)+iv(x,y) is holomorphic in z=x+iyDz=x+iy\in D. Assume that uu satisfies

u(x,0)=0andyu(x,0)=0for allxD.u(x,0)=0\quad\text{and}\quad\partial_{y}u(x,0)=0\quad\text{for\ all}\ x\in D\cap\mathbb{R}.

Then h=iah=ia for some aa\in\mathbb{R}.

Proof.

Without loss of generality we assume that DD is centered at the origin. We expand h(z)=kakzkh(z)=\sum_{k\in\mathbb{N}}a_{k}z^{k} as a power series. The identity

0=u(x,0)=Re{k0akxk}=k0Re{ak}xk0=u(x,0)=\operatorname{Re}\left\{\sum_{k\geq 0}a_{k}x^{k}\right\}=\sum_{k\geq 0}\operatorname{Re}\{a_{k}\}x^{k}

implies that each of the coefficients (ak)k0(a_{k})_{k\geq 0} is purely imaginary. Using that yh(x+iy)=k1akik(x+iy)k1\frac{\partial}{\partial y}h(x+iy)=\sum_{k\geq 1}a_{k}ik(x+iy)^{k-1} yields together with the second assumption that

0=uy(x,0)=Re{hy(x,0)}=Re{ik1akkxk1}=k1Im{ak}kxk1,0=u_{y}(x,0)=\operatorname{Re}\left\{h_{y}(x,0)\right\}=\operatorname{Re}\left\{i\sum_{k\geq 1}a_{k}kx^{k-1}\right\}=-\sum_{k\geq 1}\operatorname{Im}\{a_{k}\}kx^{k-1},

which implies that Im{ak}=0\operatorname{Im}\{a_{k}\}=0 for k1k\geq 1.
Therefore we have indeed that ak=0a_{k}=0 for k1k\geq 1 and Re{a0}=0\operatorname{Re}\{a_{0}\}=0, which yields the desired statement. ∎

Moreover, we require the following connection between the complex derivative of an analytic function and the gradient of its modulus.

Lemma 2.6.

Let DD\subseteq\mathbb{C} be a domain in the complex plane and h𝒪(D)h\in\mathcal{O}(D). Then it holds for all zDz\in D with h(z)0h(z)\neq 0 that

||h|(z)|=|h(z)|.|\nabla|h|(z)|=|h^{\prime}(z)|.
Proof.

This can be shown rather elementary using Cauchy-Riemann equations. See [14, Lemma 3.4] for a proof. ∎

Lemma 2.7.

Let F,GF,G be two analytic functions and assume that for every xx\in{\mathbb{R}}, |F(x)|=|G(x)||F(x)|=|G(x)| and |F(x)|=|G(x)||F^{\prime}(x)|=|G^{\prime}(x)|. Then there exists cc\in{\mathbb{C}} with |c|=1|c|=1 such that either G=cFG=cF or G=cF¯G=c\bar{F}.

This result can be found in [27] for entire functions of finite order and in [23] for so-called wide-banded functions.

Proof.

We resort to a nonempty open disk DD centered on the real line such that neither FF nor GG has any zeros in DD. Note that such a disk always exists unless FF (or GG) vanishes identically, in which case the statement is trivial.

By Lemma 2.6 we have for all xDx\in D\cap\mathbb{R} that

(y|G|)2(x+i0)\displaystyle(\partial_{y}|G|)^{2}(x+i0) =|G(x+i0)|2(x|G|)2(x+i0)\displaystyle=|G^{\prime}(x+i0)|^{2}-(\partial_{x}|G|)^{2}(x+i0)
=|F(x+i0)|2(x|F|)2(x+i0)=(y|F|)2(x+i0),\displaystyle=|F^{\prime}(x+i0)|^{2}-(\partial_{x}|F|)^{2}(x+i0)=(\partial_{y}|F|)^{2}(x+i0),

which implies that either

a) y|G|=y|F|\partial_{y}|G|=\partial_{y}|F| on DD\cap\mathbb{R}  or  b) y|G|=y|F|\partial_{y}|G|=-\partial_{y}|F| on DD\cap\mathbb{R}.

In case a) we consider h:=log(G/F)𝒪(D)h:=\log(G/F)\in\mathcal{O}(D) and observe that due to |F(x)|=|G(x)||F(x)|=|G(x)| for real xx,

Re{h}(x+i0)=log|G/F|(x+i0)=0,xD.\operatorname{Re}\{h\}(x+i0)=\log|G/F|(x+i0)=0,\quad x\in D\cap\mathbb{R}.

Moreover, we have that

yRe{h}(x+i0)\displaystyle\partial_{y}\operatorname{Re}\{h\}(x+i0) =y(log|G|log|F|)(x+i0)\displaystyle=\partial_{y}(\log|G|-\log|F|)(x+i0)
=(y|G||G|y|F||F|)(x+i0)=0,xD.\displaystyle=\left(\frac{\partial_{y}|G|}{|G|}-\frac{\partial_{y}|F|}{|F|}\right)(x+i0)=0,\quad x\in D\cap\mathbb{R}.

Applying Lemma 2.5 yields that h=ich=ic with cc\in\mathbb{R}, which implies that

G/F=exph=eic,G/F=\exp h=e^{ic},

and therefore G=eicFG=e^{ic}F as desired (by analyticity the identity holds on the full plane).
In case b) one considers h:=logG/F¯h:=\log G/\bar{F} and proceed similarly as in case a). ∎

We can now show (i) of Theorem 1.6:

Proposition 2.8.

Let γ\gamma denote the standard Gaussian and let

(5) γ1(t):=γ(t)andγ2(t):=2πtγ(t).\gamma_{1}(t):=\gamma(t)\quad\text{and}\quad\gamma_{2}(t):=2\pi t\gamma(t).

Assume that φ,ψL2()\varphi,\psi\in L^{2}({\mathbb{R}}) are two wave functions such that |[γkφ]|=|[γkψ]||{\mathcal{F}}[\gamma_{k}\varphi]|=|{\mathcal{F}}[\gamma_{k}\psi]| for k=1,2k=1,2. Then φ\varphi and ψ\psi are equivalent up to a constant phase factor and conjugation.

Proof.

Once more, we set F:=[γ1φ]F:={\mathcal{F}}[\gamma_{1}\varphi] and G:=[γ1ψ]G:={\mathcal{F}}[\gamma_{1}\psi] and note that these functions are analytic, even more so they extend to entire functions on the plane. According to the assumption we have that

(6) |F(x+i0)|=|G(x+i0)|for allx,|F(x+i0)|=|G(x+i0)|\quad\text{for all}~{}x\in\mathbb{R},

as well as

|F(x+i0)|=|[γ2f](x)|=|[γ2g](x)|=|G(x+i0)|,x.|F^{\prime}(x+i0)|=|{\mathcal{F}}[\gamma_{2}f](x)|=|{\mathcal{F}}[\gamma_{2}g](x)|=|G^{\prime}(x+i0)|,\quad x\in\mathbb{R}.

Lemma 2.7 then shows that G=cFG=cF or G=cF¯G=c\bar{F} which is equivalent to ψ=cφ{\mathcal{F}}^{*}\psi=c{\mathcal{F}}^{*}\varphi or ψ(ξ)=c[φ](ξ)¯{\mathcal{F}}^{*}\psi(\xi)=c\overline{{\mathcal{F}}^{*}[\varphi](-\xi)} since γ1=γ1\gamma_{1}=\gamma_{1}^{*} does not vanish. In turn, this is then equivalent to ψ=cφ\psi=c\varphi or ψ=cφ¯\psi=c\overline{\varphi}. ∎

Remark 2.9.

We have only used that FF is holomorphic in a neighborhood of the real line. As for Remark 2.4, the same proof thus applies if γ\gamma is replaced by ea|x|αe^{-a|x|^{\alpha}}, a>0a>0, α1\alpha\geq 1. Note the for α=1\alpha=1, FF is only holomorphic in a strip.

2.3. A second family of three operators

Proposition 2.10.

Let γ\gamma denote the standard Gaussian and let a,b>0a,b>0 be such that ab\dfrac{a}{b}\notin{\mathbb{Q}}, and let

(7) γ1(t):=γ(t),γ2(t):=sin(aπt)γ(t)andγ3(t):=sin(bπt)γ(t)\gamma_{1}(t):=\gamma(t),\quad\gamma_{2}(t):=\sin(a\pi t)\gamma(t)\quad\text{and}\quad\gamma_{3}(t):=\sin(b\pi t)\gamma(t)

Assume that φ,ψL2()\varphi,\psi\in L^{2}({\mathbb{R}}) are two wave functions such that |[γkφ]|=|[γkψ]||{\mathcal{F}}[\gamma_{k}\varphi]|=|{\mathcal{F}}[\gamma_{k}\psi]| for k=1,2k=1,2 and 33. Then φ\varphi and ψ\psi are equivalent up to a constant phase factor.

Proof.

We again introduce F=[γ1φ]F={\mathcal{F}}[\gamma_{1}\varphi], G=[γ1ψ]G={\mathcal{F}}[\gamma_{1}\psi] and notice that F,GF,G are entire functions of order 2 and that |F(x)|=|G(x)||F(x)|=|G(x)|.

Further, using the standard fact that the Fourier transform of a modulation is the translation of the Fourier transform and that sinα=eiαeiα2i\sin\alpha=\dfrac{e^{i\alpha}-e^{-i\alpha}}{2i}, it is straightforward to see that

|[γ2φ]|=|[γ2ψ]||{\mathcal{F}}[\gamma_{2}\varphi]|=|{\mathcal{F}}[\gamma_{2}\psi]| if and only if |F(x)F(xa)|=|G(x)G(xa)||F(x)-F(x-a)|=|G(x)-G(x-a)|,

|[γ3φ]|=|[γ3ψ]||{\mathcal{F}}[\gamma_{3}\varphi]|=|{\mathcal{F}}[\gamma_{3}\psi]| if and only if |F(x)F(xb)|=|G(x)G(xb)||F(x)-F(x-b)|=|G(x)-G(x-b)|.

Applying twice the main result of [27] we get that there exist two periodic functions Wa,WbW_{a},W_{b} with repective period aa and bb and such that both are meromorphic and continuous over {\mathbb{R}} with |Wa(x)|=|Wb(x)|=1|W_{a}(x)|=|W_{b}(x)|=1 for real xx and satisfy G=WaF=WbFG=W_{a}F=W_{b}F. In particular, Wa=WbW_{a}=W_{b} on {\mathbb{R}} so that WaW_{a} is both aa and bb-periodic. But then for every k,k,\ell\in{\mathbb{Z}} we have that Wa(ak+b)=Wa(0)W_{a}(ak+b\ell)=W_{a}(0). As a/ba/b\notin{\mathbb{Q}}, {ak+b,k,}\{ak+b\ell,k,\ell\in{\mathbb{Z}}\} is dense in {\mathbb{R}} and by continuity of WaW_{a} we get that WaW_{a} is a constant of modulus one. Finally, as G=WaFG=W_{a}F we get ψ=Waφ\psi=W_{a}\varphi as claimed. ∎

Remark 2.11.

The same proof applies if γ\gamma is replaced by eα|x|pe^{-\alpha|x|^{p}}, p1p\geq 1. Note the for p=1p=1, FF would only be holomorphic in a strip and McDonald’s result no longer applies. In this case, one needs the extension of McDonald’s result in [23].

Remark 2.12.

The condition a/ba/b\notin{\mathbb{Q}} is essential. Indeed, let a0a\not=0 and b=pqab=\dfrac{p}{q}a with p,qp,q\in{\mathbb{Q}}, q0q\not=0 and β=qa\beta=\dfrac{q}{a}. We have chosen β\beta so that e2iπβa=e2iπβb=1e^{2i\pi\beta a}=e^{2i\pi\beta b}=1.

Let φ0\varphi\not=0 be smooth and compactly supported. Define F=[γφ]F={\mathcal{F}}[\gamma\varphi] and define ψ\psi by

ψ(t)=γ(t+β)γ(t)φ(t+β)=e2πβtπβ2φ(t+β).\psi(t)=\frac{\gamma(t+\beta)}{\gamma(t)}\varphi(t+\beta)=e^{-2\pi\beta t-\pi\beta^{2}}\varphi(t+\beta).

A direct computation shows that

G(x):=[γψ](x)=[γ(+β)φ(+β)](x)=e2iπβx[γφ](x)=e2iπβxF(x).G(x):={\mathcal{F}}[\gamma\psi](x)={\mathcal{F}}[\gamma(\cdot+\beta)\varphi(\cdot+\beta)](x)=e^{2i\pi\beta x}{\mathcal{F}}[\gamma\varphi](x)=e^{2i\pi\beta x}F(x).

But then |G(x)|=|F(x)||G(x)|=|F(x)|,

|G(x)G(x+a)|=|e2iπβxF(x)e2iπβxe2iπβaF(x+a)|=|F(x)F(x+a)||G(x)-G(x+a)|=|e^{2i\pi\beta x}F(x)-e^{2i\pi\beta x}e^{2i\pi\beta a}F(x+a)|=|F(x)-F(x+a)|

since e2iπβa=1e^{2i\pi\beta a}=1 and, replacing aa by bb in this computation, |G(x)G(x+b)|=|F(x)F(x+b)||G(x)-G(x+b)|=|F(x)-F(x+b)|. The proof of Proposition 2.10 then shows that |[γkφ]|=|[γkψ]||{\mathcal{F}}[\gamma_{k}\varphi]|=|{\mathcal{F}}[\gamma_{k}\psi]| for k=1,2k=1,2 and 33. Of course, ψ\psi is not a constant multiple of φ\varphi.

2.4. An extension to higher dimensions

We will now give an extension to several variables. Let us start with a simple lemma about several variable holomorphic functions. We will make use of the following notation: for j{1,,d}j\in\{1,\ldots,d\} and x=(x1,,xd)dx=(x_{1},\ldots,x_{d})\in{\mathbb{R}}^{d}, write x(j)=(x1,,xj1,xj+1,,xd)d1x^{(j)}=(x_{1},\ldots,x_{j-1},x_{j+1},\ldots,x_{d})\in{\mathbb{R}}^{d-1}.

Lemma 2.13.

Let F,GF,G be two non-zero holomorphic functions on d{\mathbb{C}}^{d} and assume that there are functions φ1,,φd:d1𝕋\varphi_{1},\ldots,\varphi_{d}\,:{\mathbb{R}}^{d-1}\to{\mathbb{T}} such that, for every j{1,,d}j\in\{1,\ldots,d\} and every xdx\in{\mathbb{R}}^{d}, F(x)=φj(x(j))G(x)F(x)=\varphi_{j}(x^{(j)})G(x). Then there is a c𝕋c\in{\mathbb{T}} such that F=cGF=cG.

Proof.

First, as FF is continuous and non-zero, there exists a ball B(x0,r)B(x_{0},r) of d{\mathbb{R}}^{d} such that FF does not vanish on B(x0,r)B(x_{0},r). Without loss of generality, we may assume that x0=0x_{0}=0. Then as |F(x)|=|φj(x(j))||G(x)|=|G(x)||F(x)|=|\varphi_{j}(x^{(j)})||G(x)|=|G(x)|, GG does also not vanish on B(0,r)B(0,r) and therefore φj(x(j))=F(x)G(x)\varphi_{j}(x^{(j)})=\dfrac{F(x)}{G(x)} for all jj does not depend on jj. But this implies that F(x)G(x)\dfrac{F(x)}{G(x)} does not depend on any of the variables x1,,xjx_{1},\ldots,x_{j} on B(0,r)B(0,r) and is thus a constant cc i.e. F=cGF=cG on B(0,r)B(0,r). From the holomorphy of FF and GG we conclude that F=cGF=cG on d{\mathbb{C}}^{d}. ∎

Corollary 2.14.

Let γ\gamma be the Gaussian on d{\mathbb{R}}^{d}, γ(t)=eπ|t|2\gamma(t)=e^{-\pi|t|^{2}}. Let f,gL2(d)f,g\in L^{2}({\mathbb{R}}^{d}) be non-zero and such that |[γf]|=|[γg]||{\mathcal{F}}[\gamma f]|=|{\mathcal{F}}[\gamma g]|. Assume further that one of the two following conditions are satisfied:

– for all j{1,,d}j\in\{1,\ldots,d\}, |[2πtjγf]|=|[2πtjγg]||{\mathcal{F}}[2\pi t_{j}\gamma f]|=|{\mathcal{F}}[2\pi t_{j}\gamma g]| and |[(12πtj)γf]|=|[(12πtj)γg]||{\mathcal{F}}[(1-2\pi t_{j})\gamma f]|=|{\mathcal{F}}[(1-2\pi t_{j})\gamma g]| on d{\mathbb{R}}^{d};

or

– for all j{1,,d}j\in\{1,\ldots,d\}, |[sinπajtjγf]|=|[sinπajγg]||{\mathcal{F}}[\sin\pi a_{j}t_{j}\gamma f]|=|{\mathcal{F}}[\sin\pi a_{j}\gamma g]| and |[sinπbjtjγf]|=|[sinπbjγg]||{\mathcal{F}}[\sin\pi b_{j}t_{j}\gamma f]|=|{\mathcal{F}}[\sin\pi b_{j}\gamma g]| on d{\mathbb{R}}^{d} wth aj,bj>0a_{j},b_{j}>0, ajbj\dfrac{a_{j}}{b_{j}}\notin{\mathbb{Q}};

then there is a c𝕋c\in{\mathbb{T}} with g=cfg=cf.

Proof.

In both cases, consider F=[γf]F={\mathcal{F}}[\gamma f] and G=[γg]G={\mathcal{F}}[\gamma g] so that F,GF,G extend to holomorphic functions over d{\mathbb{C}}^{d}.

Let us consider the first set of hypothesis. Fix ξ¯=(ξ2,,ξd)d1\underline{\xi}=(\xi_{2},\ldots,\xi_{d})\in{\mathbb{R}}^{d-1} and denote by

fξ¯(x)=d1eπ|x¯|2f(x,x¯)e2iπx¯,ξ¯dx¯f_{\underline{\xi}}(x)=\int_{{\mathbb{R}}^{d-1}}e^{-\pi|\underline{x}|^{2}}f(x,\underline{x})e^{-2i\pi{\left\langle{\underline{x},\underline{\xi}}\right\rangle}}\,\mbox{d}\underline{x}

and use a similar notation for gg.

Let γ1\gamma_{1} be the Gaussian on {\mathbb{R}} and 1{\mathcal{F}}_{1} be the 11-variable Fourier transform, then

|1[γ1fξ¯]|=|[γf](ξ1,ξ¯)|=|[γg](ξ1,ξ¯)|=|1[γ1gξ¯]||{\mathcal{F}}_{1}[\gamma_{1}f_{\underline{\xi}}]|=|{\mathcal{F}}[\gamma f](\xi_{1},\underline{\xi})|=|{\mathcal{F}}[\gamma g](\xi_{1},\underline{\xi})|=|{\mathcal{F}}_{1}[\gamma_{1}g_{\underline{\xi}}]|

and similarily |1[2πtγ1fξ¯]|=|1[2πtγ1gξ¯]||{\mathcal{F}}_{1}[2\pi t\gamma_{1}f_{\underline{\xi}}]|=|{\mathcal{F}}_{1}[2\pi t\gamma_{1}g_{\underline{\xi}}]| and |1[(12πt)γ1fξ¯]|=|1[(12πt)γ1gξ¯]||{\mathcal{F}}_{1}[(1-2\pi t)\gamma_{1}f_{\underline{\xi}}]|=|{\mathcal{F}}_{1}[(1-2\pi t)\gamma_{1}g_{\underline{\xi}}]|. Proposition 2.2 then implies that there exists c(ξ¯)𝕋c(\underline{\xi})\in{\mathbb{T}} such that fξ¯(x)=c(ξ¯)gξ¯(x)f_{\underline{\xi}}(x)=c(\underline{\xi})g_{\underline{\xi}}(x). Multiplying by γ1\gamma_{1} and taking Fourier transform, we then get F(ξ1,ξ¯)=c(ξ¯)G(ξ1,ξ¯)F(\xi_{1},\underline{\xi})=c(\underline{\xi})G(\xi_{1},\underline{\xi}) for every ξ1\xi_{1}\in{\mathbb{R}} and every ξ¯d1\underline{\xi}\in{\mathbb{R}}^{d-1}.

Doing the same for each variable, we see that the conditions of Lemma 2.13 are fullfilled. There is then c𝕋c\in{\mathbb{T}} such that F=cGF=cG which implies that f=cgf=cg.

Replacing Proposition 2.2 by 2.10, we get that the same is valid for the second set of conditions. ∎

Note that one can obtain the same result by imposing the first set of condition for some coordinates and the second set for the others.

On the other hand, taking functions of the tensor form

f(x1,,xd)=f1(x1)fd(xd)f(x_{1},\ldots,x_{d})=f_{1}(x_{1})\cdots f_{d}(x_{d})

it is easy to see that the full set of conditions is needed.

3. Discretizations

3.1. Continuous derivative

We now turn to a discrete setting. We consider ψ=(ψ0,,ψN1)N\psi=(\psi_{0},\ldots,\psi_{N-1})\in{\mathbb{C}}^{N} which can be identified with a the polynomial

Pψ(x)=j=0N1ψje2iπjx.P_{\psi}(x)=\sum_{j=0}^{N-1}\psi_{j}e^{2i\pi jx}.

It is crucial to notice that PψP_{\psi} is a so-called analytic trigonometric polynomial i.e. it has no negative frequencies. In particular P¯ψ\bar{P}_{\psi} is not an analytic trigonometric polynomial and can therefore not be of the form PφP_{\varphi}. We will use this fact below.

Remark 3.1.

Note that that if MNM\geq N,

Pφ(kM)=j=0N1ψje2iπjk/MP_{\varphi}\left(\frac{k}{M}\right)=\sum_{j=0}^{N-1}\psi_{j}e^{2i\pi jk/M}

is the MM-dimensional discrete Fourier transform M[ψ(M)]{\mathcal{F}}_{M}[\psi^{(M)}] where ψ(M)\psi^{(M)} is the 0-padded sequence ψ(M)=(ψ0,,ψN1,0,,0)M\psi^{(M)}=(\psi_{0},\ldots,\psi_{N-1},0,\ldots,0)\in{\mathbb{C}}^{M}.

Candés et al proved that {|N[ψ](k)|,k=0,N1}\{|{\mathcal{F}}_{N}[\psi](k)|,k=0,\ldots N-1\} together with the two difference sequences {|N[ψ](k)N[ψ](k1)|,k=0,N1}\{|{\mathcal{F}}_{N}[\psi](k)-{\mathcal{F}}_{N}[\psi](k-1)|,k=0,\ldots N-1\} and {|N[ψ](k)iN[ψ](k1)|,k=0,N1}\{|{\mathcal{F}}_{N}[\psi](k)-i{\mathcal{F}}_{N}[\psi](k-1)|,k=0,\ldots N-1\} determine almost every ψN\psi\in{\mathbb{C}}^{N}.

One can see N[ψ](k)N[ψ](k1){\mathcal{F}}_{N}[\psi](k)-{\mathcal{F}}_{N}[\psi](k-1) as the discrete derivative of the sequence N[ψ](k){\mathcal{F}}_{N}[\psi](k) and this result can thus be seen as a discrete analogue of Theorem 1.6 (ii).

Instead of a discrete derivative, let us first inverstigate what is happening if we consider the continuous derivative, that is Pψ(x)=2iπj=0N1jψje2iπjx=PjψP_{\psi}^{\prime}(x)=\displaystyle 2i\pi\sum_{j=0}^{N-1}j\psi_{j}e^{2i\pi jx}=P_{j\psi} with jψ=(0,ψ1,,(N1)ψN1)j\psi=(0,\psi_{1},\ldots,(N-1)\psi_{N-1}). We are here asking whether for some MNM\geq N

(8) {|Pφ(kM)|=|Pψ(kM)||Pφ(kM)|=|Pψ(kM)|for k=0,,M1\left\{\begin{matrix}{\left|{P_{\varphi}\left(\frac{k}{M}\right)}\right|}&=&{\left|{P_{\psi}\left(\frac{k}{M}\right)}\right|}\\[9.0pt] {\left|{P_{\varphi}^{\prime}\left(\frac{k}{M}\right)}\right|}&=&{\left|{P_{\psi}^{\prime}\left(\frac{k}{M}\right)}\right|}\end{matrix}\right.\quad\text{for }k=0,\ldots,M-1

implies Pφ=λPψP_{\varphi}=\lambda P_{\psi} where |λ|=1|\lambda|=1 so that φ=λψ\varphi=\lambda\psi. In other words, we are asking whether

{|M[ψ(M)](k)|,|M[jψ(M)](k)|,k=0,M1}\{|{\mathcal{F}}_{M}[\psi^{(M)}](k)|,|{\mathcal{F}}_{M}[j\psi^{(M)}](k)|,k=0\ldots,M-1\}

determines ψ\psi up to a constant phase factor.

Now notice that |Pψ(x)|2=Pψ(x)Pψ(x)¯=j,k=0N1ψjψk¯e2iπ(jk)x|P_{\psi}(x)|^{2}=P_{\psi}(x)\overline{P_{\psi}(x)}=\sum_{j,k=0}^{N-1}\psi_{j}\overline{\psi_{k}}e^{2i\pi(j-k)x} is a trigonometric polynomial of degree NN. We may write it in the form

|Pψ(x)|2=e2iπ(N1)x=02N2ce2iπx|P_{\psi}(x)|^{2}=e^{-2i\pi(N-1)x}\sum_{\ell=0}^{2N-2}c_{\ell}e^{2i\pi\ell x}

which shows that, up to the factor e2iπ(N1)xe^{-2i\pi(N-1)x}, |Pψ(x)|2|P_{\psi}(x)|^{2} is a polynomial of degree 2N22N-2 evaluated on the unit circle. Therefore it is determined by 2N12N-1 distinct values. The same applies to |P||P^{\prime}|. For instance

|Pψ(k2N1)|,|Pψ(k2N1)|,k=0,,2N2{\left|{P_{\psi}\left(\frac{k}{2N-1}\right)}\right|},{\left|{P_{\psi}^{\prime}\left(\frac{k}{2N-1}\right)}\right|},\quad k=0,\ldots,2N-2

uniquely determine |Pψ|,|Pψ||P_{\psi}|,|P_{\psi}^{\prime}|. We can then apply Lemma 2.7 wich then shows that for M=2N1M=2N-1, (8) implies that there is a unimodular complex number λ\lambda such that Pψ(x)=λPφ(x)P_{\psi}(x)=\lambda P_{\varphi}(x) or Pψ(x)=λPφ¯(x)P_{\psi}(x)=\lambda\overline{P_{\varphi}}(x). As said above, PφP_{\varphi} and PψP_{\psi} are analytic trigonometric polynomials so that the later case can not occur. In conclusion

Proposition 3.2.

Let ψ,φN\psi,\varphi\in{\mathbb{C}}^{N} and assume that the corresponding trigonometric polynomials satisfy

{|Pφ(k2N1)|=|Pψ(k2N1)||Pφ(k2N1)|=|Pψ(k2N1)|for k=0,,2N2\left\{\begin{matrix}{\left|{P_{\varphi}\left(\frac{k}{2N-1}\right)}\right|}&=&{\left|{P_{\psi}\left(\frac{k}{2N-1}\right)}\right|}\\[9.0pt] {\left|{P_{\varphi}^{\prime}\left(\frac{k}{2N-1}\right)}\right|}&=&{\left|{P_{\psi}^{\prime}\left(\frac{k}{2N-1}\right)}\right|}\end{matrix}\right.\quad\text{for }k=0,\ldots,2N-2

then there exists λ\lambda\in{\mathbb{C}} with |λ|=1|\lambda|=1 such that ψ=λφ\psi=\lambda\varphi.

We are going to prove that this result is sharp in the sense that (8) for M=2(N1)M=2(N-1) is not sufficient for φ,ψ\varphi,\psi to be identical up to a constant phase factor. We start with N=3N=3.

Lemma 3.3.

Let p(z)=zp(z)=z and q(z)=z22+3i2q(z)=\frac{z^{2}}{2}+\frac{\sqrt{3}i}{2}. Moreover, let Λ={1,i,1,i}\Lambda=\{1,i,-1,-i\}.
Then it holds that |p(λ)|=|q(λ)||p(\lambda)|=|q(\lambda)| and |p(λ)|=|q(λ)|=1|p^{\prime}(\lambda)|=|q^{\prime}(\lambda)|=1 for all λΛ\lambda\in\Lambda.

Proof.

Obviously for |z|=1|z|=1 it holds that |p(z)|=|p(z)|=|q(z)|=1|p(z)|=|p^{\prime}(z)|=|q^{\prime}(z)|=1. Thus, it remains to check that q(λ)q(\lambda) is of unit modulus for λΛ\lambda\in\Lambda. Indeed we have that

q(±1)\displaystyle q(\pm 1) =1+3i2=eπi/3,\displaystyle=\frac{1+\sqrt{3}i}{2}=e^{\pi i/3},
q(±i)\displaystyle q(\pm i) =1+3i2=e2πi/3.\displaystyle=\frac{-1+\sqrt{3}i}{2}=e^{2\pi i/3}.

Proposition 3.4.

Let N=2m+1N=2m+1 be an odd integer 3\geq 3. There exist φ,ψN\varphi,\psi\in\mathbb{C}^{N} which are not equivalent up to a constant phase factor while

{|Pφ(k2N2)|=|Pψ(k2N2)||Pφ(k2N2)|=|Pψ(k2N2)|for k=0,,2N3.\left\{\begin{matrix}{\left|{P_{\varphi}\left(\frac{k}{2N-2}\right)}\right|}&=&{\left|{P_{\psi}\left(\frac{k}{2N-2}\right)}\right|}\\[9.0pt] {\left|{P_{\varphi}^{\prime}\left(\frac{k}{2N-2}\right)}\right|}&=&{\left|{P_{\psi}^{\prime}\left(\frac{k}{2N-2}\right)}\right|}\end{matrix}\right.\quad\text{for }k=0,\ldots,2N-3.
Proof.

We use the polynomials from Lemma 3.3 and define φ\varphi to be the sequence of the coefficients of the polynomial p~(z):=p(zm)\tilde{p}(z):=p(z^{m}) and analogously, ψ\psi to consist of the coefficents of q~(z):=q(zm)\tilde{q}(z):=q(z^{m}). In other words Pφ(x)=p(e2imπx)P_{\varphi}(x)=p(e^{2im\pi x}) and Pψ(x)=q(e2imπx)P_{\psi}(x)=q(e^{2im\pi x}).

But then Pφ(k2N2)=Pφ(k4m)=p(eikπ/2)P_{\varphi}\left(\frac{k}{2N-2}\right)=P_{\varphi}\left(\frac{k}{4m}\right)=p(e^{ik\pi/2}) and analogously, Pψ(k2N2)=q(eikπ/2)P_{\psi}\left(\frac{k}{2N-2}\right)=q(e^{ik\pi/2}). In particular |Pφ(k2N2)|=|Pψ(k2N2)|\displaystyle{\left|{P_{\varphi}\left(\frac{k}{2N-2}\right)}\right|}={\left|{P_{\psi}\left(\frac{k}{2N-2}\right)}\right|} for k=0,,2N2k=0,\ldots,2N-2.

On the other hand

Pφ(x)=2iπme2imπxp(e2imπx)andPψ(x)=2iπme2imπxq(e2imπx);P_{\varphi}^{\prime}(x)=2i\pi me^{2im\pi x}p^{\prime}(e^{2im\pi x})\quad\text{and}\quad P_{\psi}^{\prime}(x)=2i\pi me^{2im\pi x}q^{\prime}(e^{2im\pi x});

thus, we get that |Pφ(x)|=|Pψ(x)|=2πm|P_{\varphi}^{\prime}(x)|=|P_{\psi}^{\prime}(x)|=2\pi m so that φ,ψ\varphi,\psi satisfy the condition of the theorem.

Finally, since the number of non-zero coefficients φ\varphi and ψ\psi are different, it is obvious that φ\varphi and ψ\psi are not equivalent. ∎

In view of the result by Candés al [7] it seems natural to ask

Question 3.5.

For which MM is it true that for almost every φN\varphi\in{\mathbb{C}}^{N}, every ψN\psi\in{\mathbb{C}}^{N} such that

(9) {|Pφ(kM)|=|Pψ(kM)||Pφ(kM)|=|Pψ(kM)|for k=0,,M1\left\{\begin{matrix}{\left|{P_{\varphi}\left(\frac{k}{M}\right)}\right|}&=&{\left|{P_{\psi}\left(\frac{k}{M}\right)}\right|}\\[9.0pt] {\left|{P_{\varphi}^{\prime}\left(\frac{k}{M}\right)}\right|}&=&{\left|{P_{\psi}^{\prime}\left(\frac{k}{M}\right)}\right|}\end{matrix}\right.\quad\text{for }k=0,\ldots,M-1

is equivalent to φ\varphi up to a constant phase factor? In other words, for which MM is the set of vectors which possess nontrivial ambiguous solutions a set of measure zero?

3.2. Discrete derivative

In this section we consider again samples of PφP_{\varphi} : uk=Pφ(kM)u_{k}=P_{\varphi}\left(\dfrac{k}{M}\right) (seen as an MM-periodic sequence) and we ask whether |uk||u_{k}| and its discrete derivative |ukuk1||u_{k}-u_{k-1}| determine φ\varphi up to a constant phase factor.

This will follow from the following proposition:

Proposition 3.6.

Let P,QP,Q be two polynomials of degree N\leq N and 0<b<2π/N0<b<2\pi/N. Assume that for every xx\in{\mathbb{R}},

{|P(eix)|=|Q(eix)||P(ei(x+b))P(eix)|=|Q(ei(x+b))Q(eix)|\left\{\begin{matrix}|P(e^{ix})|=|Q(e^{ix})|\\ |P(e^{i(x+b)})-P(e^{ix})|=|Q(e^{i(x+b)})-Q(e^{ix})|\end{matrix}\right.

then there is a unimodular constant such that P=cQP=cQ.

Proof.

The proof is divided into two steps. The first one is folklore and the second part is an elaboration on a result by McDonalds [27].

Write P(x)=αxkj=1K(xxj)P(x)=\displaystyle\alpha x^{k}\prod_{j=1}^{K}(x-x_{j}) and Q(x)=βxlj=1L(xyk)Q(x)=\displaystyle\beta x^{l}\prod_{j=1}^{L}(x-y_{k}) with xk,yk0x_{k},y_{k}\not=0 and, without loss of generality KLK\geq L. Then

|P(e2iπx)|2=|α|2P(e2iπx)P(e2iπx)¯=|α|2j=1K(e2iπxxj)(e2iπxxj¯)=|α|2e2iπKxj=1K(e2iπxxj)(1xj¯e2iπx)|P(e^{2i\pi x})|^{2}=|\alpha|^{2}P(e^{2i\pi x})\overline{P(e^{2i\pi x})}=|\alpha|^{2}\prod_{j=1}^{K}(e^{2i\pi x}-x_{j})(e^{-2i\pi x}-\overline{x_{j}})\\ =|\alpha|^{2}e^{-2i\pi Kx}\prod_{j=1}^{K}(e^{2i\pi x}-x_{j})(1-\overline{x_{j}}e^{2i\pi x})

while

|Q(e2iπx)|2=|β|2e2iπLxj=1L(e2iπxyj)(1yj¯e2iπx).|Q(e^{2i\pi x})|^{2}=|\beta|^{2}e^{-2i\pi Lx}\prod_{j=1}^{L}(e^{2i\pi x}-y_{j})(1-\overline{y_{j}}e^{2i\pi x}).

It follows that |P|=|Q||P|=|Q| on the unit circle implies that if ζ=e2iπx\zeta=e^{2i\pi x}

|α|2j=1K(ζxj)(1xj¯ζ)=|β|2ζKLj=1L(e2iπxζ)(1yj¯ζ).|\alpha|^{2}\prod_{j=1}^{K}(\zeta-x_{j})(1-\overline{x_{j}}\zeta)=|\beta|^{2}\zeta^{K-L}\prod_{j=1}^{L}(e^{2i\pi x}-\zeta)(1-\overline{y_{j}}\zeta).

This is an identity between two polynomials. As it is valid on the unit circle, it is valid over {\mathbb{C}}. As a consequence, as the left hand side does not vanish at zero, so does the right hand side and K=LK=L. Further, the two polynomials have same zeros. The zeros of the left hand side (counted with multiplicity) are {xj,1/x¯j,j=1,,K}\{x_{j},1/\bar{x}_{j},j=1,\ldots,K\} and those of the right hand side are {yj,1/y¯j,j=1,,K}\{y_{j},1/\bar{y}_{j},j=1,\ldots,K\} thus for every jj, yj=xjy_{j}=x_{j} or yj=1/x¯jy_{j}=1/\bar{x}_{j}, the reflection of xjx_{j} with respect to the unit circle. In particular, note that if |xj|=1|x_{j}|=1 then it is a common zero of PP and QQ and 1/x¯j=xj1/\bar{x}_{j}=x_{j}.

It follows that, up to reordering the zeroes, we may first list the zeros that are not reflected and then those that are reflected:

Q(x)=βzlj=1J(xxj)j=J+1K(x1/xj¯).Q(x)=\beta z^{l}\prod_{j=1}^{J}(x-x_{j})\prod_{j=J+1}^{K}(x-1/\overline{x_{j}}).

In order to remove some ambiguities, note that one may have a pair of zeros of the form {x,1/x¯}\{x,1/\bar{x}\}, i.e. there are j,kj,k such that xj=xx_{j}=x and xk=1/x¯x_{k}=1/\bar{x}. Up to reordering the zeroes, we may assume that those j,kj,k’s are J\leq J. We can thus write P(z)=αzkP1(z)P2(z)P(z)=\alpha z^{k}P_{1}(z)P_{2}(z) and Q(z)=βzlP1(z)P2(z)Q(z)=\beta z^{l}P_{1}(z)P_{2}^{*}(z) with P2(z)=j=J+1K(xxj)P_{2}(z)=\prod_{j=J+1}^{K}(x-x_{j}) and P2(z)=j=J+1K(x1/xj¯)P_{2}^{*}(z)=\prod_{j=J+1}^{K}(x-1/\overline{x_{j}}). Moreover, assume that if j,kJ+1j,k\geq J+1 then xj1/xk¯x_{j}\not=1/\overline{x_{k}} since the corresponding terms can be put into P1P_{1}.

Our aim is to show that this factor P2P_{2} is not present here. From now one we argue towards a contradiction by assuming that there is at least one reflected zero, so that P2P_{2} has at least one zero xJx_{J}. Further, up to re-ordering the zeroes, we may assume that |xj||x_{j}| is non-decreasing for jJj\geq J.

From McDonald [27] we know that Q(eix)=W(x)P(eix)Q(e^{ix})=W(x)P(e^{ix}) with WW meromorphic, periodic of period bb with |W(x)|=1|W(x)|=1 for xx real, continuous on the real line. The previous argument shows that

W(x)=βαei(lk)xj=J+1K(eix1/xj¯)j=J+1K(eixxj)W(x)=\frac{\beta}{\alpha}e^{i(l-k)x}\frac{\prod_{j=J+1}^{K}(e^{ix}-1/\overline{x_{j}})}{\prod_{j=J+1}^{K}(e^{ix}-x_{j})}

so that

W(x+b)=βαei(lk)(x+b)j=J+1K(eixeib/xj¯)j=J+1K(eixxjeib).W(x+b)=\frac{\beta}{\alpha}e^{i(l-k)(x+b)}\frac{\prod_{j=J+1}^{K}(e^{ix}-e^{-ib}/\overline{x_{j}})}{\prod_{j=J+1}^{K}(e^{ix}-x_{j}e^{-ib})}.

But then W(x)=W(x+b)W(x)=W(x+b) implies that

j=J+1K(eix1/xj¯)j=J+1K(eixxjeib)=ei(lk)bj=J+1K(eixeib/xj¯)j=J+1K(eixxj)\prod_{j=J+1}^{K}(e^{ix}-1/\overline{x_{j}})\prod_{j=J+1}^{K}(e^{ix}-x_{j}e^{-ib})=e^{i(l-k)b}\prod_{j=J+1}^{K}(e^{ix}-e^{-ib}/\overline{x_{j}})\prod_{j=J+1}^{K}(e^{ix}-x_{j})

for every xx\in{\mathbb{R}} so that we have the identity between polynomials

j=J+1K(X1/xj¯)j=J+1K(Xxjeib)=ei(lk)bj=J+1K(Xeib/xj¯)j=J+1K(Xxj)\prod_{j=J+1}^{K}(X-1/\overline{x_{j}})\prod_{j=J+1}^{K}(X-x_{j}e^{-ib})=e^{i(l-k)b}\prod_{j=J+1}^{K}(X-e^{-ib}/\overline{x_{j}})\prod_{j=J+1}^{K}(X-x_{j})

Therefore the sets of zeros {xjeib,j=J+1,,K}{1/xj¯,j=J+1,,K}\{x_{j}e^{-ib},j=J+1,\ldots,K\}\cup\{1/\overline{x_{j}},j=J+1,\ldots,K\} and {xj,j=J+1,,K}{eib/xj¯,j=J+1,,K}\{x_{j},j=J+1,\ldots,K\}\cup\{e^{-ib}/\overline{x_{j}},j=J+1,\ldots,K\} are equal (counting multiplicity).

Let LKNL\leq K\leq N be such that |xJ+j|=|xJ+1||x_{J+j}|=|x_{J+1}| for j=1,,Lj=1,\ldots,L. If we had {xj+Jeib,j=1,,L}={xj+J,j=1,,L}\{x_{j+J}e^{-ib},j=1,\ldots,L\}=\{x_{j+J},j=1,\ldots,L\} with multiplicity then this set would be invariant under multiplication by eibe^{-ib}. In particular, it contains {xJeikb,k}\{x_{J}e^{-ikb},k\in{\mathbb{Z}}\} but we have chosen b<2π/N2π/Lb<2\pi/N\leq 2\pi/L so P2P_{2} would have more than LL zeros, a contradiction. Thus there is a j,kj,k such that xjeib=eib/xk¯x_{j}e^{-ib}=e^{-ib}/\overline{x_{k}}. that is xj=1/xk¯x_{j}=1/\overline{x_{k}}, again a contradiction.

We are then left with W(x)=βαei(kl)xW(x)=\dfrac{\beta}{\alpha}e^{i(k-l)x} which is bb-periodic. As b<2π/Nb<2\pi/N it follows that k=lk=l. Thus WW is a constant of modulus 11 and Q=WPQ=WP as claimed. ∎

Note that the argument also works if bπb\in{\mathbb{R}}\setminus{\mathbb{Q}}\pi.

Corollary 3.7.

Let φ,ψN\varphi,\psi\in{\mathbb{C}}^{N} and assume that for k=0,,2N2k=0,\ldots,2N-2,

{|Pφ(k2N1)|=|Pψ(k2N1)||Pφ(k+12N1)Pφ(k2N1)|=|Pψ(k+12N1)Pψ(k2N1)|\left\{\begin{matrix}{\left|{P_{\varphi}\left(\dfrac{k}{2N-1}\right)}\right|}={\left|{P_{\psi}\left(\dfrac{k}{2N-1}\right)}\right|}\\[9.0pt] {\left|{P_{\varphi}\left(\dfrac{k+1}{2N-1}\right)-P_{\varphi}\left(\dfrac{k}{2N-1}\right)}\right|}={\left|{P_{\psi}\left(\dfrac{k+1}{2N-1}\right)-P_{\psi}\left(\dfrac{k}{2N-1}\right)}\right|}\end{matrix}\right.

then φ,ψ\varphi,\psi are equivalent up to a phase factor.

Proof.

As in the previous section, |Pφ(k2N1)|,k=0,,2N2\displaystyle{\left|{P_{\varphi}\left(\dfrac{k}{2N-1}\right)}\right|},k=0,\ldots,2N-2 fully defines |Pφ||P_{\varphi}| while

|Pφ(k+12N1)Pφ(k2N1)|,k=0,,2N2\displaystyle{\left|{P_{\varphi}\left(\dfrac{k+1}{2N-1}\right)-P_{\varphi}\left(\dfrac{k}{2N-1}\right)}\right|},k=0,\ldots,2N-2

fully determines |Pφ(x+(2N1)1)Pφ(x)||P_{\varphi}(x+(2N-1)^{-1})-P_{\varphi}(x)|. Applying Proposition (3.6) implies that there is λ\lambda\in{\mathbb{C}} with |λ|=1|\lambda|=1 such that Pψ=λPφP_{\psi}=\lambda P_{\varphi} which gives the result. ∎

We will now show that the result is false if we sample at a rate 1/(2N2)1/(2N-2) instead of 1/(2N1)1/(2N-1).

Lemma 3.8.

Let p(z)=zp(z)=z and q(z)=z2+i2q(z)=\frac{z^{2}+i}{\sqrt{2}}. Moreover, let Λ={1,i,1,i}\Lambda=\{1,i,-1,-i\}. Then it holds that |p(λ)|=|q(λ)||p(\lambda)|=|q(\lambda)| and |p(λ)p(iλ)|=|q(λ)q(iλ)||p(\lambda)-p(i\lambda)|=|q(\lambda)-q(i\lambda)| for all λΛ\lambda\in\Lambda.

Proof.

This is easily checked by direct computation. ∎

Proposition 3.9.

Let N=2m+1N=2m+1 be an odd integer 3\geq 3. There exist signals φ,ψN\varphi,\psi\in\mathbb{C}^{N} which are not equivalent up to a phase factor such that for k=0,,2N3k=0,\ldots,2N-3,

(10) {|Pφ(k2N2)|=|Pψ(k2N2)||Pφ(k2N2)Pφ(k12N2)|=|Pψ(k2N2)Pψ(k12N2)|.\left\{\begin{matrix}{\left|{P_{\varphi}\left(\frac{k}{2N-2}\right)}\right|}={\left|{P_{\psi}\left(\frac{k}{2N-2}\right)}\right|}\\[9.0pt] {\left|{P_{\varphi}\left(\frac{k}{2N-2}\right)-P_{\varphi}\left(\frac{k-1}{2N-2}\right)}\right|}={\left|{P_{\psi}\left(\frac{k}{2N-2}\right)-P_{\psi}\left(\frac{k-1}{2N-2}\right)}\right|}\end{matrix}\right..
Proof.

We use a similar construction as in the proof of Proposition 3.4 and define φ,ψ\varphi,\psi to be the sequence of coefficients of p~(z):=p(zm)\tilde{p}(z):=p(z^{m}) and q~(z):=q(zm)\tilde{q}(z):=q(z^{m}), respecitvely, where p,qp,q are the polynomials from Lemma 3.8. Again, since the number of non-zero coefficients does not agree we find that φ\varphi and ψ\psi are not equivalent.

To see that they satisfy (10) note that from Lemma 3.8 we deduce that

|Pφ(k2N2)|\displaystyle{\left|{P_{\varphi}\left(\frac{k}{2N-2}\right)}\right|} =\displaystyle= |p(e2iπmk4m)|=|p(eikπ/2)|\displaystyle{\left|{p(e^{2i\pi m\frac{k}{4m}})}\right|}={\left|{p(e^{ik\pi/2})}\right|}
=\displaystyle= |q(eikπ/2)|=|Pψ(k2N2)|\displaystyle{\left|{q(e^{ik\pi/2})}\right|}={\left|{P_{\psi}\left(\frac{k}{2N-2}\right)}\right|}

while

|Pφ(k+12N2)Pφ(k2N2)|=|p(ieikπ/2)p(eikπ/2)|=|q(ieikπ/2)q(eikπ/2)|=|Pψ(k+12N2)Pψ(k2N2)|{\left|{P_{\varphi}\left(\frac{k+1}{2N-2}\right)-P_{\varphi}\left(\frac{k}{2N-2}\right)}\right|}={\left|{p(ie^{ik\pi/2})-p(e^{ik\pi/2})}\right|}\\ ={\left|{q(ie^{ik\pi/2})-q(e^{ik\pi/2})}\right|}={\left|{P_{\psi}\left(\frac{k+1}{2N-2}\right)-P_{\psi}\left(\frac{k}{2N-2}\right)}\right|}

which implies the claim. ∎

Appendix

The purpose of this section is to provide some background on Wright’s conjecture, as formulated in Conjecture 1.3.
We begin with introducing the main objects and notions appearing in quantum mechanics that we need here. The space of all possible states of a quantum mechanical system is represented by L2()L^{2}(\mathbb{R}). A state ψL2()\psi\in L^{2}(\mathbb{R}) is also called a wave function. Two wave functions ψ\psi and φ\varphi are considered equivalent if they agree up to multiplication by a unimodular constant.
Quantities of a system that can be measured are called observables and represented by densely-defined self-adjoint operators on L2()L^{2}(\mathbb{R}). The expected value of the state ψD(A)\psi\in D(A) in the observable AA is defined as

Eψ(A)=ψ,Aψ.E_{\psi}(A)={\left\langle{\psi,A\psi}\right\rangle}.

The two following examples are essential in this paper. Let u,vL()u,v\in L^{\infty}({\mathbb{R}}) be real valued. To uu and vv associate the following two observables222Note that we normalized the Fourier transform {\mathcal{F}} so that it is unitary. Its adjoint is thus the inverse Fourier transform φ(x)=1φ(x)=φ(x){\mathcal{F}}^{*}\varphi(x)={\mathcal{F}}^{-1}\varphi(x)={\mathcal{F}}\varphi(-x).

Muφ=uφandvφ=[v[φ]].M_{u}\varphi=u\varphi\qquad\mbox{and}\quad{\mathcal{M}}_{v}\varphi={\mathcal{F}}^{*}\bigl{[}v{\mathcal{F}}[\varphi]\bigr{]}.

Then

Eψ(Mu)=u(x)|ψ(x)|2dxE_{\psi}(M_{u})=\int_{\mathbb{R}}u(x)|\psi(x)|^{2}\,\mbox{d}x

and

Eψ(v)=v(ξ)|ψ^(ξ)|2dξ.E_{\psi}({\mathcal{M}}_{v})=\int_{\mathbb{R}}v(\xi)|\widehat{\psi}(\xi)|^{2}\,\mbox{d}\xi.

Here, we keep the convention of notation in mathematics where the position variable is denoted by xx and the momentum variable is denoted by ξ\xi instead of pp.

Let {\mathcal{B}} be the set of Borel subsets of {\mathbb{R}}. It is then obvious that |ψ(x)||\psi(x)| is uniquely determined by

Q={Eψ(M𝟏B)}B\mathcal{E}_{Q}=\big{\{}E_{\psi}(M_{\mathbf{1}_{B}})\big{\}}_{B\in{\mathcal{B}}}

which is called the distribution of the state ψ\psi with respect to position since

Q={𝟏B(Q)ψ}B\mathcal{E}_{Q}=\big{\{}\|\mathbf{1}_{B}(Q)\psi\|\big{\}}_{B\in{\mathcal{B}}}

where 𝟏B(Q)\mathbf{1}_{B}(Q) are the spectral projections associated to the position operator.

On the other hand |ψ^(ξ)||\widehat{\psi}(\xi)| is uniquely determined by distribution of the state ψ\psi with respect to momentum:

P:={Eψ(v):v=𝟏B,B a Borel set}={𝟏B(P)ψ,B a Borel set}:=𝒮P\mathcal{E}_{P}:=\{E_{\psi}({\mathcal{M}}_{v})\,:\ v=\mathbf{1}_{B},\ B\mbox{ a Borel set}\}=\{\|\mathbf{1}_{B}(P)\psi\|,\ B\mbox{ a Borel set}\}:=\mathcal{S}_{P}

where 𝟏B(P)\mathbf{1}_{B}(P) are the spectral projections associated to the momentum operator.

In a footnote to the Handbuch der Physik article on the general principle of wave mechanics [30], W. Pauli asked whether a wave function ψ\psi is uniquely determined (up to a constant phase factor) by one of the equivalent quantities

  • the Pauli data (|ψ|,|ψ^|)(|\psi|,|\widehat{\psi}|);

  • {𝟏B(Q)ψ}B\big{\{}\|\mathbf{1}_{B}(Q)\psi\|\big{\}}_{B\in{\mathcal{B}}}, {𝟏B(P)ψ}B\big{\{}\|\mathbf{1}_{B}(P)\psi\|\big{\}}_{B\in{\mathcal{B}}};

  • {Eψ(M𝟏B)}B\big{\{}E_{\psi}(M_{\mathbf{1}_{B}})\big{\}}_{B\in{\mathcal{B}}}, {Eψ(𝟏B)}B\big{\{}E_{\psi}({\mathcal{M}}_{\mathbf{1}_{B}})\big{\}}_{B\in{\mathcal{B}}}.

The question can also be found e.g. in the book by H. Reichenbach [31] and in Busch & Lahti [5].

As mentionned in the introduction, it is known that in general the Pauli data does not uniquely determine the state ψ\psi (up to a constant phase factor).

It is then natural to ask whether there exists a set of observables (Aj)jJ(A_{j})_{j\in J} (preferably including position and momentum or at least having a physical meaning) such that the associated sets built from spectral projections

j:={𝟏B(Aj)ψ}B={Eψ(𝟏B(Aj))}B\mathcal{E}_{j}:=\big{\{}\|\mathbf{1}_{B}(A_{j})\psi\|\big{\}}_{B\in{\mathcal{B}}}=\big{\{}E_{\psi}\bigl{(}\mathbf{1}_{B}(A_{j})\bigr{)}\big{\}}_{B\in{\mathcal{B}}}

uniquely determine every state ψ\psi.

Using the spectral theorem, to a self-adjoint operator AjA_{j} we can associate a unitary operator UjU_{j} and a multiplication operator MjM_{j} on a space L2(μj)L^{2}(\mu_{j}) such that Aj=UjMjUjA_{j}=U_{j}^{*}M_{j}U_{j}. Then the data j\mathcal{E}_{j}, jJj\in J uniquely determine |Ujψ||U_{j}\psi|, jJj\in J. This then directly leads to Wright’s Conjecture 1.3 and to its relaxation 1.4: find a set of measures μj\mu_{j} and unitary operators Uj:L2(d)L2(μj)U_{j}\,:L^{2}({\mathbb{R}}^{d})\to L^{2}(\mu_{j}) such that |Ujψ|=|Ujφ||U_{j}\psi|=|U_{j}\varphi|, jJj\in J, implies that ψ\psi and φ\varphi are equivalent up to a constant phase factor.

A relaxed version is to find a set {Tj}jJ\{T_{j}\}_{j\in J} of bounded self-adjoint (or even only bounded) operators on L2()L^{2}({\mathbb{R}}) such that |Tjψ|=|Tjφ||T_{j}\psi|=|T_{j}\varphi|, jJj\in J, implies that ψ\psi and φ\varphi are equivalent up to a constant phase factor. The data |Tjψ||T_{j}\psi| can also be interpreted as an expectation of the state ψ\psi with respect to a family of observables. To be more precise, to a bounded operator TT, we may associate the self-adjoint operator Au=TMuTA_{u}=T^{*}M_{u}T whith uL()u\in L^{\infty}({\mathbb{R}}) real valued. Then

ψ,Auψ=u(x)|Tψ(x)|2dx{\left\langle{\psi,A_{u}\psi}\right\rangle}=\int_{\mathbb{R}}u(x)|T\psi(x)|^{2}\,\mbox{d}x

so that |Tψ||T\psi| is uniquely determined by

T:={Eψ(TM𝟏BT)}B.{\mathcal{E}}_{T}:=\big{\{}E_{\psi}(T^{*}M_{\mathbf{1}_{B}}T)\big{\}}_{B\in{\mathcal{B}}}.

However, it does not seem possible to reformulate this family of measurements in terms of spectral projections associated to a single self-adjoint operator.

4. Data availability

No data has been generated or analysed during this study.

5. Funding and/or Conflicts of interests/Competing interests

The second author was supported by an Erwin-Schrödinger Fellowship (J-4523) of the Austrian Science Fund FWF.

The authors have no relevant financial or non-financial interests to disclose.

References

  • [1] E. Akutowicz, On the determination of the phase of a Fourier integral. I. Trans. Amer. Math. Soc. 83 (1956), 179–192.
  • [2] R. Balan, P. Casazza & D. Edidin, On signal reconstruction without phase. Appl. Comp. Harmon. Anal. 20 (2006), 345–356.
  • [3] T. Bendory, R. Beinert & Y. C. Eldar, Fourier phase retrieval: uniqueness and algorithms. In Compressed sensing and its applications, Appl. Numer. Harmon. Anal., pages 55-–91. Birkhäuser/Springer, Cham, 2017.
  • [4] B. G. Bodmann & N. Hammen, Stable phase retrieval with low-redundancy frames. Adv. Comput. Math. 41 (2015), 317–331.
  • [5] P. Busch & P. J. Lahti, The determination of the past and the future of a physical system in quantum mechanics. Found. Phys. 19 (1989), 633–78.
  • [6] E. Candes, X. Li & M. Soltanolkotabi, Phase Retrieval from Coded Diffraction Patterns. Appl. Comput. Harmon. Anal., 39 (2015), 277–299.
  • [7] E. J. Candès, Y. Eldar, T. Strohmer & V. Voroninski, Phase Retrieval via Matrix Completion. SIAM Review 57 (2015), 225–251.
  • [8] A. Conca, D. Edidin, M. Hering & C. Vinzant, An algebraic characterization of injectivity in phase retrieval. Appl. Comp. Harmon. Anal. 38 (2015), 346–356.
  • [9] J. V. Corbett, The Pauli problem, state reconstruction and quantum-real numbers. Rep. Math. Phys., 57 (2006), 53–68.
  • [10] J. V. Corbett & C. A. Hurst, Are wave functions uniquely determined by their position and momentum distributions? J. Austral. Math. Soc B, 20 (1978), 182–201.
  • [11] M. Fickus, D. G. Mixon, A. A. Nelson & Y. Wang, Phase retrieval from very few measurements. Linear Algebra Appl. 449 (2014), 475–499.
  • [12] J. Goodman, Introduction to Fourier Optics. McGraw-Hill physical and quantum electronics series. W. H. Freeman, 2005.
  • [13] D. Goyeneche, G. Cat̃as, S. Etcheverry, E. S. Gómez, G. B. Xavier, G. Lima & A. Delgado, Five Measurement Bases Determine Pure Quantum States on Any Dimension. Phys. Rev. Lett. 115 (2015) 090401.
  • [14] P. Grohs & M. Rathmair, Stable Gabor Phase Retrieval and Spectral Clustering. Comm. Pure Appl. Math., 72 (2019), 981–1043.
  • [15] P. Grohs, S. Koppensteiner & M. Rathmair, Phase Retrieval: Uniqueness and Stability. SIAM Review, 62 (2020), 301–350.
  • [16] D. Gross, F. Krahmer & R. Kueng, Improved recovery guarantees for phase retrieval from coded diffraction patterns. Appl. Comput. Harmon. Anal., 42 (2017), 37–64.
  • [17] E. Hofstetter, Construction of time-limited functions with specified autocorrelation functions, IEEE Trans. Inform. Theory, 10 (1964), pp. 119-126.
  • [18] T. Heinosaari, L. Mazzarella & M. M. Wolf, Quantum Tomography under Prior Information. Commun. Math. Phys. 318 (2013), 355–374.
  • [19] R. S. Ismagilov, On the Pauli problem. Funksional Anal i Prilozhen 30 (1996), 82–84.
  • [20] K. Jaganathan, Y. C. Eldar & B. Hassibi, Phase retrieval: an overview of recent developments. In Optical compressive imaging, Ser. Opt. Optoelectron., pages 263–296. CRC Press, Boca Raton, FL, 2017
  • [21] Ph. Jaming, Phase retrieval techniques for radar ambiguity functions. J. Fourier Anal. Appl., 5 (1999), 313–333.
  • [22] Ph. Jaming, Uniqueness results in an extension of Pauli’s phase retrieval problem. Applied and Comp. Harm. Anal. 37 (2014) 413–441.
  • [23] Ph. Jaming, K. Kellay & R. Perez III, Phase Retrieval for Wide Band Signals. J. Fourier Anal. Appl. 26 (2020) No 54.
  • [24] A.J.E.M. Janssen, The Zak transform and some counterexamples in time-frequency analysis. IEEE Trans. Inform. Theory 38 (1992) 168–171.
  • [25] M. V. Klibanov, P. E. Sacks & A. V. Tikhonravov, The phase retrieval problem. Inverse Problems, 11 (1995),.
  • [26] D. Mondragon & V. Voroninski Determination of all pure quantum states from a minimal number of observables. arXiv:1306.1214 [math-ph].
  • [27] J. Mc Donald, Phase retrieval and magnitude retrieval of entire functions, J. Fourier Anal. Appl., 10 (2004), 259–267.
  • [28] D. G. Mixon, Phase Transitions in Phase Retrieval. In: Balan R., Begué M., Benedetto J., Czaja W., Okoudjou K. (eds) Excursions in Harmonic Analysis, Volume 4.(2015) Applied and Numerical Harmonic Analysis. Birkhäuser.
  • [29] B. Z. Morov & A. M. Perelomov, On a problem posed by Pauli. Theor. Math. Phys. 101 (1994), 1200–1204.
  • [30] W. Pauli, Die allgemeinen Prinzipien der Wellenmechanik. In: H. Geiger, K. Scheel (Eds.), Handbuch der Physik, 24, Springer-Verlag, Berlin, 1933, pp. 83-272; English translation: General Principles of Quantum Mechanics, Springer-Verlag, Berlin, 1980.
  • [31] H. Reichenbach, Philosophic Foundations of Quantum Mechanics. Univesity of California Press, Berkeley, 1944.
  • [32] Y. Shechtman, Y. C. Eldar, O. Cohen, H. Chapman, J. Miao & M. Segev, Phase Retrieval with Application to Optical Imaging: A contemporary overview. 32 (2015), 87–109.
  • [33] C. Vinzant, A small frame and a certificate of its injectivity. Sampling Theory and Applications (SampTA) Conference Proceedings. (2015), pp. 197 - 200.
  • [34] A. Vogt, Position and momentum distributions do not determine the quantum mechanical state. In A.R. Marlow, editor, Mathematical Foundations of Quantum Theory. Academic Press, New York, 1978.
  • [35] A. Walther, The question of phase retrieval in optics. Optica Acta 10 (1963), 41–49.