Uniqueness of phase retrieval from three measurements
Abstract.
In this paper we consider the question of finding an as small as possible family of operators on that does phase retrieval: every is uniquely determined (up to a constant phase factor) by the phaseless data . 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 , the Fourier transform and are “masks”. Here we explicitely construct three real-valued masks so that the associated coded diffraction patterns do phase retrieval. This implies that the three self-adjoint operators 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 functions1991 Mathematics Subject Classification:
30D05, 42B10, 94A121. 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 of one real variable, i.e. . The phase retrieval problems we shall consider will be associated to a given family of operators.
Definition 1.1.
Let be a family of linear operators on , i.e. linear. We say that does phase retrieval if
with the set of complex numbers of modulus .
Clearly, due to the linearity of the operators, and produce the same phaseless measurements when .
Thus, the notion of uniqueness introduced in Definition 1.1 is the best one can hope for.
In practice, an arbitrary linear operator 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 which do phase retrieval subject to the two side constraints
-
i)
the operators represent physically meaningful objects and
-
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 of admissible linear operators on , find a family of operators 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 given its so-called diffraction pattern, which is represented by , 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
for , and extend the definition to in the usual way.
The mapping is far from injective: Clearly, for an arbitrary measurable phase function we get that has the same Fourier modulus as .
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 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 , which interacts multiplicatively with the unknown signal before computing its diffraction pattern . 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) |
where denotes the multiplication operator, and seek for a (small) family of operators with corresponding masks 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 is uniquely determined (up to a constant phase factor) by the pair , which is sometimes called the Pauli data. In our terms, the question Pauli posed amounts to
Does 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 so that . It follows that and so that and .
However, we may introduce slightly less restrictive notions of equivalence, such as the following two:
– two states are equivalent up to a constant phase factor and conjugation, if there exists with such that either or
– two states are equivalent up to a constant phase factor and conjugate reflection, if there exists with such that either or where .
In Bargmann’s example, the states are still equivalent up to a constant phase factor
and conjugation (or conjugate reflection). One may then ask for a class of states such that
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) |
Obviously, and 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 such that 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 such that 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 , let be a square root of . For and , define
(3) | |||||
The last expression shows that extends to a bounded operator on .
Further the usual Fourier transform and we define
the identity operator and then .
Finally, is a unitary operator with .
One of the authors showed that does phase retrieval [22, Proposition 4.2].
An equivalent formulation is
that if for every time , the free Schrödinger evolution of
have same modulus then 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 on 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 of and then consider the rank one projections on the spaces , , , and . 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 the Gaussian and , then the pair and does phase retrieval [22, Proposition 4.1]. Note that while is self-adjoint, 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 amounts to asking whether there exist three orthonormal bases and , such that each and every vector is uniquely determined (up to multiplication by a unimodular constant) by the measurements
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 such that ,
uniquely determines up to a unimodular constant.
Heinosaari, Mazzarella and Wolf [18] proved that the minimal number of vectors is
with when .
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
unitaries that lead to uniqueness up to a constant phase factor [13].
proved that for four “generic” orthonormal bases are enough to
determine all 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 be the standard Gaussian, , and let be defined by
-
(i)
Let be such that and then and are equivalent up to a constant phase factor and conjugation-reflection: or with ;
-
(ii)
if we further assume that then and are equivalent up to a constant phase factor; in other words, does phase retrieval.
Remark 1.7.
The actual result can be extended in multiple ways. For instance, the function can be replaced by . 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 where we need operators.
Remark 1.8.
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 with an analytic trigonometric polynomial . The measurements we consider are then samples of and of the modulus of the derivative of . We will show that this requires 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 phaseless measurements are not sufficient. However, it allows to show the role of the sampling rate and explains why phaseless measurements may not have been the right analogue of Wright’s conjecture in .
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 be an open interval and let denote the space of complex-valued analytic functions on . Then is uniquely determined (up to multiplication by a unimodular constant) by and .
Proof.
Suppose that are such that and . We may assume w.l.o.g. that does not vanish identically. Therefore there exists a nonempty interval such that has no zeros in . Moreover, according to the assumption we have that
i.e. is constant on . This implies that on for some . Since we get that must be unimodular. Finally, by analyticity the identity extends to all of . ∎
We are now in position to prove the second part of the theorem:
Proposition 2.2.
Let be the standard Gaussian and let be defined by
Let be two wave functions such that for and . Then and are equivalent up to a unimodular constant only.
Proof.
First we define a pair of analytic function on the real line by and .
It is enough to show that for implies that with since then . Then, as does not vanish, we get that and are equivalent up to a constant phase factor.
Now note that , and therefore that we have the identities
Thus, it remains to show that a function analytic on the real line, is uniquely determined given and . To see this, first compute
Together with
we get that
(4) |
It follows that and that . Applying Lemma 2.1 yields the desired statement. ∎
Remark 2.3.
In the next section, we are going to prove that for implies that or . In this last case . But then reads which implies that . But then (4) simplifies to . Again lemma 2.1 yields the desired statement.
The direct proof given here is substentially simpler.
Remark 2.4.
The Gaussian only plays a mild role here:
– it implies that is holomorphic in a neighborhood of the real line so that we may replace by any function that is for some ,
– does not vanish on a set of positive measure so that is uniquely determined by .
This shows that we could replace by e.g , , .
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 such that on the real line and . Once one notices that 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 . 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 be a nonempty, open disk centered on the real line. Let be two smooth real valued functions such that is holomorphic in . Assume that satisfies
Then for some .
Proof.
Without loss of generality we assume that is centered at the origin. We expand as a power series. The identity
implies that each of the coefficients is purely imaginary. Using that yields together with the second assumption that
which implies that for .
Therefore we have indeed that for and , 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 be a domain in the complex plane and . Then it holds for all with that
Proof.
This can be shown rather elementary using Cauchy-Riemann equations. See [14, Lemma 3.4] for a proof. ∎
Lemma 2.7.
Let be two analytic functions and assume that for every , and . Then there exists with such that either or .
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 centered on the real line such that neither nor has any zeros in . Note that such a disk always exists unless (or ) vanishes identically, in which case the statement is trivial.
By Lemma 2.6 we have for all that
which implies that either
a) on or b) on .
In case a) we consider and observe that due to for real ,
Moreover, we have that
Applying Lemma 2.5 yields that with , which implies that
and therefore as desired (by analyticity the identity holds on the full plane).
In case b) one considers and proceed similarly as in case a).
∎
Proposition 2.8.
Let denote the standard Gaussian and let
(5) |
Assume that are two wave functions such that for . Then and are equivalent up to a constant phase factor and conjugation.
Proof.
Once more, we set and 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) |
as well as
Lemma 2.7 then shows that or which is equivalent to or since does not vanish. In turn, this is then equivalent to or . ∎
Remark 2.9.
We have only used that is holomorphic in a neighborhood of the real line. As for Remark 2.4, the same proof thus applies if is replaced by , , . Note the for , is only holomorphic in a strip.
2.3. A second family of three operators
Proposition 2.10.
Let denote the standard Gaussian and let be such that , and let
(7) |
Assume that are two wave functions such that for and . Then and are equivalent up to a constant phase factor.
Proof.
We again introduce , and notice that are entire functions of order 2 and that .
Further, using the standard fact that the Fourier transform of a modulation is the translation of the Fourier transform and that , it is straightforward to see that
– if and only if ,
– if and only if .
Applying twice the main result of [27] we get that there exist two periodic functions with repective period and and such that both are meromorphic and continuous over with for real and satisfy . In particular, on so that is both and -periodic. But then for every we have that . As , is dense in and by continuity of we get that is a constant of modulus one. Finally, as we get as claimed. ∎
Remark 2.11.
The same proof applies if is replaced by , . Note the for , 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 is essential. Indeed, let and with , and . We have chosen so that .
Let be smooth and compactly supported. Define and define by
A direct computation shows that
But then ,
since and, replacing by in this computation, . The proof of Proposition 2.10 then shows that for and . Of course, is not a constant multiple of .
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 and , write .
Lemma 2.13.
Let be two non-zero holomorphic functions on and assume that there are functions such that, for every and every , . Then there is a such that .
Proof.
First, as is continuous and non-zero, there exists a ball of such that does not vanish on . Without loss of generality, we may assume that . Then as , does also not vanish on and therefore for all does not depend on . But this implies that does not depend on any of the variables on and is thus a constant i.e. on . From the holomorphy of and we conclude that on . ∎
Corollary 2.14.
Let be the Gaussian on , . Let be non-zero and such that . Assume further that one of the two following conditions are satisfied:
– for all , and on ;
or
– for all , and on wth , ;
then there is a with .
Proof.
In both cases, consider and so that extend to holomorphic functions over .
Let us consider the first set of hypothesis. Fix and denote by
and use a similar notation for .
Let be the Gaussian on and be the -variable Fourier transform, then
and similarily and . Proposition 2.2 then implies that there exists such that . Multiplying by and taking Fourier transform, we then get for every and every .
Doing the same for each variable, we see that the conditions of Lemma 2.13 are fullfilled. There is then such that which implies that .
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
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 which can be identified with a the polynomial
It is crucial to notice that is a so-called analytic trigonometric polynomial i.e. it has no negative frequencies. In particular is not an analytic trigonometric polynomial and can therefore not be of the form . We will use this fact below.
Remark 3.1.
Note that that if ,
is the -dimensional discrete Fourier transform where is the -padded sequence .
Candés et al proved that together with the two difference sequences and determine almost every .
Instead of a discrete derivative, let us first inverstigate what is happening if we consider the continuous derivative, that is with . We are here asking whether for some
(8) |
implies where so that . In other words, we are asking whether
determines up to a constant phase factor.
Now notice that is a trigonometric polynomial of degree . We may write it in the form
which shows that, up to the factor , is a polynomial of degree evaluated on the unit circle. Therefore it is determined by distinct values. The same applies to . For instance
uniquely determine . We can then apply Lemma 2.7 wich then shows that for , (8) implies that there is a unimodular complex number such that or . As said above, and are analytic trigonometric polynomials so that the later case can not occur. In conclusion
Proposition 3.2.
Let and assume that the corresponding trigonometric polynomials satisfy
then there exists with such that .
We are going to prove that this result is sharp in the sense that (8) for is not sufficient for to be identical up to a constant phase factor. We start with .
Lemma 3.3.
Let and .
Moreover, let .
Then it holds that and for all .
Proof.
Obviously for it holds that . Thus, it remains to check that is of unit modulus for . Indeed we have that
∎
Proposition 3.4.
Let be an odd integer . There exist which are not equivalent up to a constant phase factor while
Proof.
We use the polynomials from Lemma 3.3 and define to be the sequence of the coefficients of the polynomial and analogously, to consist of the coefficents of . In other words and .
But then and analogously, . In particular for .
On the other hand
thus, we get that so that satisfy the condition of the theorem.
Finally, since the number of non-zero coefficients and are different, it is obvious that and are not equivalent. ∎
In view of the result by Candés al [7] it seems natural to ask
Question 3.5.
For which is it true that for almost every , every such that
(9) |
is equivalent to up to a constant phase factor? In other words, for which 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 : (seen as an -periodic sequence) and we ask whether and its discrete derivative determine up to a constant phase factor.
This will follow from the following proposition:
Proposition 3.6.
Let be two polynomials of degree and . Assume that for every ,
then there is a unimodular constant such that .
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 and with and, without loss of generality . Then
while
It follows that on the unit circle implies that if
This is an identity between two polynomials. As it is valid on the unit circle, it is valid over . As a consequence, as the left hand side does not vanish at zero, so does the right hand side and . Further, the two polynomials have same zeros. The zeros of the left hand side (counted with multiplicity) are and those of the right hand side are thus for every , or , the reflection of with respect to the unit circle. In particular, note that if then it is a common zero of and and .
It follows that, up to reordering the zeroes, we may first list the zeros that are not reflected and then those that are reflected:
In order to remove some ambiguities, note that one may have a pair of zeros of the form , i.e. there are such that and . Up to reordering the zeroes, we may assume that those ’s are . We can thus write and with and . Moreover, assume that if then since the corresponding terms can be put into .
Our aim is to show that this factor is not present here. From now one we argue towards a contradiction by assuming that there is at least one reflected zero, so that has at least one zero . Further, up to re-ordering the zeroes, we may assume that is non-decreasing for .
From McDonald [27] we know that with meromorphic, periodic of period with for real, continuous on the real line. The previous argument shows that
so that
But then implies that
for every so that we have the identity between polynomials
Therefore the sets of zeros and are equal (counting multiplicity).
Let be such that for . If we had with multiplicity then this set would be invariant under multiplication by . In particular, it contains but we have chosen so would have more than zeros, a contradiction. Thus there is a such that . that is , again a contradiction.
We are then left with which is -periodic. As it follows that . Thus is a constant of modulus and as claimed. ∎
Note that the argument also works if .
Corollary 3.7.
Let and assume that for ,
then are equivalent up to a phase factor.
Proof.
As in the previous section, fully defines while
fully determines . Applying Proposition (3.6) implies that there is with such that which gives the result. ∎
We will now show that the result is false if we sample at a rate instead of .
Lemma 3.8.
Let and . Moreover, let . Then it holds that and for all .
Proof.
This is easily checked by direct computation. ∎
Proposition 3.9.
Let be an odd integer . There exist signals which are not equivalent up to a phase factor such that for ,
(10) |
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 . A state is also called a wave function.
Two wave functions and 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 .
The expected value of the state in the observable is
defined as
The two following examples are essential in this paper. Let be real valued. To and associate the following two observables222Note that we normalized the Fourier transform so that it is unitary. Its adjoint is thus the inverse Fourier transform .
Then
and
Here, we keep the convention of notation in mathematics where the position variable is denoted by and the momentum variable is denoted by instead of .
Let be the set of Borel subsets of . It is then obvious that is uniquely determined by
which is called the distribution of the state with respect to position since
where are the spectral projections associated to the position operator.
On the other hand is uniquely determined by distribution of the state with respect to momentum:
where 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 is uniquely determined (up to a constant phase factor) by one of the equivalent quantities
-
•
the Pauli data ;
-
•
, ;
-
•
, .
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 (up to a constant phase factor).
It is then natural to ask whether there exists a set of observables (preferably including position and momentum or at least having a physical meaning) such that the associated sets built from spectral projections
uniquely determine every state .
Using the spectral theorem, to a self-adjoint operator we can associate a unitary operator and a multiplication operator on a space such that . Then the data , uniquely determine , . This then directly leads to Wright’s Conjecture 1.3 and to its relaxation 1.4: find a set of measures and unitary operators such that , , implies that and are equivalent up to a constant phase factor.
A relaxed version is to find a set of bounded self-adjoint (or even only bounded) operators on such that , , implies that and are equivalent up to a constant phase factor. The data can also be interpreted as an expectation of the state with respect to a family of observables. To be more precise, to a bounded operator , we may associate the self-adjoint operator whith real valued. Then
so that is uniquely determined by
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.