geometry and symmetry in short-and-sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfgeometry and...

11
Geometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 12 Yuqian Zhang 3 Yenson Lau 12 John Wright 124 Abstract We study the Short-and-Sparse (SaS) deconvo- lution problem of recovering a short signal a 0 and a sparse signal x 0 from their convolution. We propose a method based on nonconvex opti- mization, which under certain conditions recovers the target short and sparse signals, up to signed shift symmetry which is intrinsic to this model. This symmetry plays a central role in shaping the optimization landscape for deconvolution. We give a regional analysis, which characterizes this landscape geometrically, on a union of subspaces. Our geometric characterization holds when the length-p 0 short signal a 0 has shift coherence μ, and x 0 follows a random sparsity model with rate θ h c1 p0 , c2 p0 μ+ p0 i · 1 log 2 p0 . Based on this geometry, we give a provable method that successfully solves SaS deconvolution with high probability. 1. Introduction Datasets in a wide range of areas, including neuroscience (Lewicki, 1998), microscopy (Cheung et al., 2017) and as- tronomy (Saha, 2007), can be modeled as superpositions of translations of a basic motif. Data of this nature can be modeled mathematically as a convolution y = a 0 * x 0 , between a short signal a 0 (the motif) and a longer sparse signal x 0 , whose nonzero entries indicate where in the sam- ple the motif is present. A very similar structure arises in image deblurring (Chan & Wong, 1998), where y is a blurry image, a 0 the blur kernel, and x 0 the (edge map) of the target sharp image. 1 Data Science Institute, Columbia University, New York City, NY, USA 2 Department of Electrical Engineering, Columbia University, New York City, NY, USA 3 Department of Com- puter Science, Cornell University, Ithaca, NY, USA 4 Department of Applied Math and Applied Physics, Columbia University, New York City, NY, USA. Correspondence to: Han-Wen Kuo <[email protected]>. Proceedings of the 36 th International Conference on Machine Learning, Long Beach, California, PMLR 97, 2019. Copyright 2019 by the author(s). Motivated by these and related problems in imaging and scientific data analysis, we study the Short-and-Sparse (SaS) Deconvolution problem of recovering a short signal a 0 R p0 and a sparse signal x 0 R n (n p 0 ) from their length-n cyclic convolution 1 y = a 0 * x 0 R n . This SaS model exhibits a basic scaled shift symmetry: for any nonzero scalar α and cyclic shift s [·], αs [a 0 ] * 1 α s -[x 0 ] = y. (1.1) Because of this symmetry, we only expect to recover a 0 and x 0 up to a signed shift (see Figure 1). Our problem of interest can be stated more formally as: Problem 1.1 (Short-and-Sparse Deconvolution). Given the cyclic convolution y = a 0 * x 0 R n of a 0 R p0 short (p 0 n), and x 0 R n sparse, recover a 0 and x 0 , up to a scaled shift. Despite a long history and many applications, until recently very little algorithmic theory was available for SaS decon- volution. Much of this difficulty can be attributed to the scale-shift symmetry: natural convex relaxations fail, and nonconvex formulations exhibit a complicated optimization landscape, with many equivalent global minimizers (scaled shifts of the ground truth) and additional local minimizers (scaled shift truncations of the ground truth), and a vari- ety of critical points (Zhang et al., 2017; 2018). Currently available theory guarantees approximate recovery of a trun- cation 2 of a shift s [a 0 ], rather than guaranteeing recovery of a 0 as a whole, and requires certain (complicated) condi- tions on the convolution matrix associated with a 0 (Zhang et al., 2018). In this paper, describe an algorithm which, under sim- pler conditions, exactly recovers a scaled shift of the pair (a 0 , x 0 ). Our algorithm is based on a formulation first intro- duced in (Zhang et al., 2017), which casts the deconvolution problem as (nonconvex) optimization over the sphere. We characterize the geometry of this objective function, and show that near a certain union of subspaces, every local minimizer is very close to a signed shift of a 0 . Based on 1 Our result applies to direct convolution by zero padding both a0 and x0. 2 I.e., the portion of the shifted signal s [a0] that falls in the window {0,...,p0 - 1}.

Upload: others

Post on 22-Mar-2020

22 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

Geometry and Symmetry in Short-and-Sparse Deconvolution

Han-Wen Kuo 1 2 Yuqian Zhang 3 Yenson Lau 1 2 John Wright 1 2 4

Abstract

We study the Short-and-Sparse (SaS) deconvo-lution problem of recovering a short signal a0

and a sparse signal x0 from their convolution.We propose a method based on nonconvex opti-mization, which under certain conditions recoversthe target short and sparse signals, up to signedshift symmetry which is intrinsic to this model.This symmetry plays a central role in shaping theoptimization landscape for deconvolution. Wegive a regional analysis, which characterizes thislandscape geometrically, on a union of subspaces.Our geometric characterization holds when thelength-p0 short signal a0 has shift coherence µ,and x0 follows a random sparsity model withrate θ ∈

[c1p0, c2p0√µ+√p0

]· 1

log2 p0. Based on

this geometry, we give a provable method thatsuccessfully solves SaS deconvolution with highprobability.

1. IntroductionDatasets in a wide range of areas, including neuroscience(Lewicki, 1998), microscopy (Cheung et al., 2017) and as-tronomy (Saha, 2007), can be modeled as superpositionsof translations of a basic motif. Data of this nature canbe modeled mathematically as a convolution y = a0 ∗ x0,between a short signal a0 (the motif) and a longer sparsesignal x0, whose nonzero entries indicate where in the sam-ple the motif is present. A very similar structure arises inimage deblurring (Chan & Wong, 1998), where y is a blurryimage, a0 the blur kernel, and x0 the (edge map) of thetarget sharp image.

1Data Science Institute, Columbia University, New YorkCity, NY, USA 2Department of Electrical Engineering, ColumbiaUniversity, New York City, NY, USA 3Department of Com-puter Science, Cornell University, Ithaca, NY, USA 4Departmentof Applied Math and Applied Physics, Columbia University,New York City, NY, USA. Correspondence to: Han-Wen Kuo<[email protected]>.

Proceedings of the 36 th International Conference on MachineLearning, Long Beach, California, PMLR 97, 2019. Copyright2019 by the author(s).

Motivated by these and related problems in imaging andscientific data analysis, we study the Short-and-Sparse (SaS)Deconvolution problem of recovering a short signal a0 ∈Rp0 and a sparse signal x0 ∈ Rn (n � p0) from theirlength-n cyclic convolution1 y = a0 ∗ x0 ∈ Rn. ThisSaS model exhibits a basic scaled shift symmetry: for anynonzero scalar α and cyclic shift s`[·],(

α s`[a0])∗(

1α s−`[x0]

)= y. (1.1)

Because of this symmetry, we only expect to recover a0

and x0 up to a signed shift (see Figure 1). Our problem ofinterest can be stated more formally as:

Problem 1.1 (Short-and-Sparse Deconvolution). Given thecyclic convolution y = a0 ∗ x0 ∈ Rn of a0 ∈ Rp0 short(p0 � n), and x0 ∈ Rn sparse, recover a0 and x0, up to ascaled shift.

Despite a long history and many applications, until recentlyvery little algorithmic theory was available for SaS decon-volution. Much of this difficulty can be attributed to thescale-shift symmetry: natural convex relaxations fail, andnonconvex formulations exhibit a complicated optimizationlandscape, with many equivalent global minimizers (scaledshifts of the ground truth) and additional local minimizers(scaled shift truncations of the ground truth), and a vari-ety of critical points (Zhang et al., 2017; 2018). Currentlyavailable theory guarantees approximate recovery of a trun-cation2 of a shift s`[a0], rather than guaranteeing recoveryof a0 as a whole, and requires certain (complicated) condi-tions on the convolution matrix associated with a0 (Zhanget al., 2018).

In this paper, describe an algorithm which, under sim-pler conditions, exactly recovers a scaled shift of the pair(a0,x0). Our algorithm is based on a formulation first intro-duced in (Zhang et al., 2017), which casts the deconvolutionproblem as (nonconvex) optimization over the sphere. Wecharacterize the geometry of this objective function, andshow that near a certain union of subspaces, every localminimizer is very close to a signed shift of a0. Based on

1Our result applies to direct convolution by zero padding botha0 and x0.

2I.e., the portion of the shifted signal s`[a0] that falls in thewindow {0, . . . , p0 − 1}.

Page 2: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

-1 0 1

-1 0

0 1

0 2

-2 0

*

*=

Figure 1. Shift symmetry in Short-and-Sparse deconvolution. Anobservation y (left) is a convolution of a short signal a0 and asparse signal x0 (top right) can be equivalently expressed as aconvolution of s`[a0] and s−`[x0], where s`[·] denotes a shift `samples. The ground truth signals a0 and x0 can only be identifiedup to a scaled shift.

this geometric analysis, we give provable methods for SaSdeconvolution that exactly recover a scaled shift of (a0,x0)whenever a0 is shift-incoherent and x0 is a sufficientlysparse random vector. Our geometric analysis highlights therole of symmetry in shaping the objective landscape for SaSdeconvolution.

The remainder of this paper is organized as follows. Sec-tion 2 introduces our optimization approach and modelingassumptions. Section 3 introduces our main results — bothgeometric and algorithmic — and compares them to theliterature. In Section 4 we present a experimental resultwhich corroborates our theoretical claim. Finally, Section 5discusses two main limitations of our analysis and describesdirections for future work.

2. Formulation and Assumptions2.1. Nonconvex SaS over the Sphere

Our starting point is the (natural) formulation

mina,x

12 ‖a ∗ x− y‖

22

Data Fidelity+ λ ‖x‖1

Sparsitys.t. ‖a‖2 = 1. (2.1)

We term this optimization problem the Bilinear Lasso, for itsresemblance to Lasso estimator in statistics. Indeed, letting

ϕlasso(a) ≡ minx

{12 ‖a ∗ x− y‖

22 + λ ‖x‖1

}(2.2)

denote the optimal Lasso cost, we see that (2.1) simplyoptimizes ϕlasso with respect to a:

mina ϕlasso(a) s.t. ‖a‖2 = 1. (2.3)

In (2.1)-(2.3), we constrain a to have unit `2 norm. Thisconstraint breaks the scale ambiguity between a and x.Moreover, the choice of constraint manifold has surpris-ingly strong implications for computation: if a is insteadconstrained to the simplex, the problem admits trivialglobal minimizers. In contrast, local minima of the sphere-constrained formulation often correspond to shifts (or shifttruncations (Zhang et al., 2017)) of the ground truth a0.

The problem (2.3) is defined in terms of the optimal Lassocost. This function is challenging to analyze, especially faraway from a0. (Zhang et al., 2017) analyzes the local min-ima of a simplification of (2.3), obtained by approximating3

the data fidelity term as12 ‖a ∗ x− y‖

22 = 1

2 ‖a ∗ x‖22 − 〈a ∗ x,y〉+ 1

2 ‖y‖22 ,

≈ 12 ‖x‖

22 − 〈a ∗ x,y〉+ 1

2 ‖y‖22 . (2.4)

This yields a simpler objective function

ϕ`1(a) = minx

{12 ‖x‖

22 − 〈a ∗ x,y〉+ 1

2 ‖y‖22 + λ ‖x‖1

}.

(2.5)We make one further simplification to this problem, replac-ing the nondifferentiable penalty ‖·‖1 with a smooth ap-proximation ρ(x).4 Our analysis allows for a variety ofsmooth sparsity surrogates ρ(x); for concreteness, we stateour main results for the particular penalty5

ρ(x) =∑i

(x2i + δ2

)1/2. (2.6)

For δ > 0, this is a smooth function of x; as δ ↘ 0 itapproaches ‖x‖1. Replacing ‖·‖1 with ρ(·), we obtain theobjective function which will be our main object of study,

ϕρ(a) = minx

{12 ‖x‖

22 − 〈a ∗ x,y〉+ 1

2 ‖y‖22 + λρ(x)

}.

(2.7)As in (Zhang et al., 2017), we optimize ϕρ(a) over thesphere Sp−1:

mina

ϕρ(a) s.t. a ∈ Sp−1 (2.8)

Here, we set p = 3p0 − 2. As we will see, optimizing overthis slightly higher dimensional sphere enables us to recovera (full) shift of a0, rather than a truncated shift. Our ap-proach will leverage the following fact: if we view a ∈ Sp−1as indexed by coordinates W = {−p0 + 1, . . . , 2p0 − 1} ,then for any shifts ` ∈ {−p0 + 1, . . . , p0 − 1}, the supportof `-shifted short signal s`[a0] is entirely contained in inter-val W . We will give a provable method which recovers ascaled version of one of these canonical shifts.

2.2. Analysis Setting and Assumptions

For convenience, we assume that a0 has unit `2 norm, i.e.,a0 ∈ Sp0−16. Our analysis makes two main assumptions,on the short motif a0 and the sparse map x0, respectively:

3For a generic a, we have 〈si[a], sj [a]〉 ≈ 0 and hence‖a ∗ x‖22 ≈ ‖e0 ∗ x‖

22 = ‖x‖22.

4Objective ϕ`1 is not twice differentiable everywhere, hencecannot be minimized with conventional second order methods.

5This surrogate is often named as the pseudo-Huber function.6This is purely a technical convenience. Our theory guarantees

recovery of a signed shift (±s`[a0],±s−`[x0]) of the truth. If‖a0‖2 6= 1, identical reasoning implies that our method recoversa scaled shift

(αs`[a0], α−1s−`[x0]

)with α = ± 1

‖a0‖2.

2

Page 3: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

Spiky Generic Tapered Generic Lowpass

µ ≈ 0 µ ≈ p−1/20

µ ≈ β

θ ≈ p−1/20

(√p0 events every p0)

θ ≈ p−3/40

( 4√p0 events every p0)

θ ≈ p−10

(1 event every p0)

a0

x0

Figure 2. Sparsity-coherence tradeoff: We show three families ofmotifs a0 with varying coherence µ (top), with their maximumallowable sparsity θ and number of copies θp0 within each length-p0 window respectively (bot). When the target motif has smallershift-coherence µ, our result allows larger θ, and vise versa. Thissparsity-coherence tradeoff is made precise in our main resultTheorem 3.1, which, loosely speaking, asserts that when θ /1/(p0

√µ+√p0), our method succeeds.

The first is that distinct shifts a0 have small inner product.We define the shift coherence of µ(a0) to be the largestinner product between distinct shifts:

µ(a0) = max` 6=0|〈a0, s`[a0]〉| . (2.9)

µ(a0) is bounded between 0 and 1. Our theory allows any µsmaller than some numerical constant. Figure 2 shows threeexamples of families of a0 that satisfy this assumption:

• Spiky. When a0 is close to the Dirac delta δ0, the shiftcoherence µ(a0) ≈ 0.7 Here, the observed signal yconsists of a superposition of sharp pulses. This isarguably the easiest instance of SaS deconvolution.

• Generic. If a0 is chosen uniformly at random fromthe sphere Sp0−1, its coherence is bounded as µ(a0) /√

1/p0 with high probability.

• Tapered Generic Lowpass. Here, a0 is generated bytaki

• ng a random conjugate symmetric superposition of thefirst L length-p0 Discrete Fourier Transform (DFT)basis signals, windowing (e.g., with Hamming win-dow) and normalizing to unit `2 norm. When L =p0√

1− β, with high probability µ(a0) / β. In thismodel, µ does not have to diminish as p0 grows – it

7The use of “≈” here suppresses constant and log factors.

can be a fixed constant.8

Intuitively speaking, problems with smaller µ are easier tosolve, a claim which will be made precise in our results.

We assume that x0 is a sparse random vector, underBernoulli-Gaussian distribution, with rate θ. Concretelyspeaking, we assume x0i = ωigi, where ωi ∼ Ber(θ),gi ∼ N (0, 1) with all random variables are jointly indepen-dent. We write this as

x0 ∼i.i.d. BG(θ). (2.10)

Here, θ is the probability that a given entry x0i is nonzero.Problems with smaller θ are easier to solve. In the extremecase, when θ � 1/p0, the observation y contains manyisolated copies of the motif a0, and a0 can be determined bydirect inspection. Our analysis will focus on the nontrivialscenario, when θ ' 1/p0.

Our technical results will articulate sparsity-coherence trade-offs, in which smaller coherence µ enables larger θ, andvice-versa. More specifically, in our main theorem, thesparsity-coherence relationship is captured in the form

θ / 1/(p0√µ+√p0). (2.11)

When a0 is very shift-incoherent (µ ≈ 0), our method suc-ceeds when each length-p0 window contains about

√p0

copies of a0. When µ is larger (as in the generic low-pass model), our method succeeds as long as relatively fewcopies of a0 overlap in the observed signal. In Figure 2,we illustrate these tradeoffs for the three models describedabove.

3. Main Results: Geometry and Algorithms3.1. Geometry of the Objective ϕρ

The goal in SaS deconvolution is to recover a0 (and x0) upto a signed shift — i.e., we wish to recover some ±s`[a0].The shifts ±s`[a0] play a key role in shaping the landscapeof ϕρ. In particular, we will argue that over a certain subsetof the sphere, every local minimum of ϕρ is close to some±s`[a0].

To gain intuition into the properties of ϕρ, we first visualizethis function in the vicinity of a single shift s`[a0] of theground truth a0. In Figure 3-(a), we plot the function valueof ϕρ over B`2,r(s`[a0]) ∩ Sp−1, where B`2,r(a) is a ballof radius r around a. We make two observations:

8The upper right panel of Figure 2 is generated using randomDFT components with frequencies smaller then one-third Nyquist.Such a kernel is incoherent, with high probability. Many commonlyoccurring low-pass kernels have µ(a0) larger – very close to one.One of the most important limitations of our results is that they donot provide guarantees in this highly coherent situation.

3

Page 4: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

B`2,r(s`[a0]) ∩ Sp−1

s`[a0]

ϕρ(a)

s`1 [a0]

s`2 [a0]

S{`1,`2}

S{`1,`2} ∩ Sp−1

ϕρ(a)

S{`1,`2,`3} ∩ Sp−1

ϕρ(a)

s`1 [a0]

s`2 [a0]s`3 [a0]

S{`1,`2,`3}

(i) (ii) (iii)

Figure 3. Geometry of ϕρ near span of shift(s) of a0. (i) A portion of the sphere Sp−1 near s`[a0] (bot) colored according to height of ϕρ(top); ϕρ is strongly convex in this region, and it has a minimizer very close to s`[a0]. (ii) Each pair of shifts s`1 [a0], s`2 [a0] defines alinear subspace S{`1,`2} of Rp (left), in which every local minimum of ϕρ near S{`1,`2} is close to either s`1 [a0] or s`2 [a0] (right); thereis a negative curvature in the middle of s`1 [a0], s`2 [a0], and ϕρ is convex in direction away from S{`1,`2}. (iii) The subspace S{`1,`2,`3}is three-dimensional; its intersection with the sphere Sp−1 is isomorphic to a two-dimensional sphere (left). On this set, ϕρ has localminimizers near each of the s`i [a0], and are the only minimizers near S{`1,`2,`3} (right).

• The objective function ϕρ is strongly convex on thisneighborhood of s`[a0].

• There is a local minimizer very close to s`[a0].

We next visualize the objective function ϕρ near the linearspan of two different shifts s`1 [a0] and s`2 [a0]. More pre-cisely, we plot ϕρ near the intersection (Figure 3-(b)) of thesphere Sp−1 and the linear subspace

S{`1,`2} ={α1s`1 [a0] +α2s`2 [a0]

∣∣α ∈ R2}.

We make three observations:

• Again, there is a local minimizer near each shift s`[a0].

• These are the only local minimizers in the vicinity ofS{`1,`2}. In particular, the objective function ϕ exhibitsnegative curvature along S{`1,`2} at any superpositionα1s`1 [a0] +α2s`2 [a0] whose weights α1 and α2 arebalanced, i.e., |α1| ≈ |α2|.

• Furthermore, the function ϕρ exhibits positive curva-ture in directions away from the subspace S`1,`2 .

Finally, we visualize ϕρ over the intersection (Figure 3-(c)) of the sphere Sp−1 with the linear span of three shiftss`1 [a0], s`2 [a0], s`3 [a0] of the true kernel a0:

S{`1,`2,`3} ={α1s`1 [a0] +α2s`2 [a0] +α3s`3 [a0]

∣∣α ∈ R3}.

Again, there is a local minimizer near each signed shift.At roughly balanced superpositions of shifts, the objectivefunction exhibits negative curvature. As a result, again, theonly local minimizers are close to signed shifts.

Our main geometric result will show that these propertiesobtain on every subspace spanned by a few shifts of a0.Indeed, for each subset

τ ⊆ {−p0 + 1, . . . , p0 − 1} , (3.1)

define a linear subspace

Sτ ={∑

`∈τ α`s`[a0]∣∣α ∈ R2p0−1

}. (3.2)

The subspace Sτ is the linear span of the shifts s`[a0] in-dexed by ` in the set τ . Our geometric theory will show thatwith high probability the function ϕρ has no spurious localminimizers near any Sτ for which τ is not too large – say,|τ | ≤ 4θp0. Combining all of these subspaces into a singlegeometric object, define the union of subspaces

Σ4θp0 =⋃|τ |≤4θp0 Sτ . (3.3)

Figure 4 gives a schematic portrait of this set. We claim:

• In the neighborhood of Σ4θp0 , all local minimizers arenear signed shifts.

• The value of ϕρ grows in directions away from Σ4θp0 .

Our main result formalizes the above observations, undertwo key assumptions: first, that the sparsity rate θ is suffi-ciently small (relative to the shift coherence µ of p0), and,second, the signal length n is sufficiently large:

Theorem 3.1 (Main Geometric Theorem). Let y = a0 ∗x0

with a0 ∈ Sp0−1 µ-shift coherent and x0 ∼i.i.d. BG(θ) ∈Rn with sparsity rate

θ ∈[c1p0,

c2p0√µ+√p0

]· 1

log2 p0. (3.4)

Choose ρ(x) =√x2 + δ2 and set λ = 0.1/

√p0θ in ϕρ.

Then there exists δ > 0 and numerical constant c suchthat if n ≥ poly(p0), with high probability, every localminimizer a of ϕρ over Σ4θp0 satisfies ‖a− σs`[a0]‖2 ≤cmax

{µ, p−10

}for some signed shift σs`[a0] of the true

kernel. Above, c1, c2 > 0 are positive numerical constants.4

Page 5: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

S`1,`2 S`1,`3

S`2,`3

Σ4θp0

ϕρ(a)

Figure 4. Geometry of ϕρ over the union of subspaces Σ4θp0 . Weshow schematic representation of the union of subspaces Σ4θp0

(left). For each set τ of at most 4θp0 shifts, we have a subspaceSτ , by which ϕρ has good geometry near (right).

Proof. See Appendix B.

The upper bound on θ in (3.4) yields the tradeoff betweencoherence and sparsity described in Figure 2. Simply put,when a0 is better conditioned (as a kernel), its coherence µis smaller and x0 can be denser.

At a technical level, our proof of Theorem 3.1 shows that (i)ϕρ(a) is strongly convex in the vicinity of each signed shift,and that at every other point a near Σ4θp0 , there is either(ii) a nonzero gradient or (iii) a direction of strict negativecurvature; furthermore (iv) the function ϕρ grows awayfrom Σ4θp0 . Points (ii)-(iii) imply that near Σ4θp0 there areno “flat” saddles: every saddle point has a direction of strictnegative curvature. We will leverage these properties topropose an efficient algorithm for finding a local minimizernear Σ4θp0 . Moreover, this minimizer is close enough toa shift (here, ‖a− s`[a0]‖2 / µ) for us to exactly recovers`[a0]: we will give a refinement algorithm that produces(±s`[a0],±s−`[x0]).

3.2. Provable Algorithm for SaS Deconvolution

The objective function ϕρ has good geometric properties on(and near!) the union of subspaces Σ4θp0 . In this section,we show how to use give an efficient method that exactlyrecovers a0 and x0, up to shift symmetry. Although ourgeometric analysis only controls ϕρ near Σ4θp0 , we willgive a descent method which, with appropriate initializationa(0), produces iterates a(1), . . . ,a(k), . . . that remain closeto Σ4θp0 for all k. In short, it is easy to start near Σ4θp0

and easy to stay near Σ4θp0 . After finding a local minimizera, we refine it to produce a signed shift of (a0,x0) usingalternating minimization.

Our algorithm starts with a initialization scheme which gen-erates a(0) near the union of subspaces Σ4θp0 , which con-sists of linear combinations of just a few shifts of a0. How

Data y Kernel a0 Sparse x0

Windowed Data a(−1) αisi[a0] + αjsj [a0]Initialization a(0)

= ∗

Figure 5. Data-driven initialization. Using a piece of the observeddata y to generate an initial point that is close to a superpositionof shifts s`[a0] of the ground truth. Data y = a0 ∗ x0 is asuperposition of shifts of the true kernel a0 (top). A length-p0windowed y contains pieces of just a few shifts as a(−1), one stepof the generalized power method approximately fills in its missingpieces, yielding a(0) as a near superposition of shifts of a0 (bot).

can we find a point near this union? Notice that the data yalso consists of a linear combination of just a few shifts ofa0 Indeed:

y = a0 ∗ x0 =∑`∈supp(x0)

x0`s`[a0]. (3.5)

A length-p0 segment of data y0,...,p0−1 = [y0, . . . ,yp0−1]T

captures portions of roughly 2θp0 � 4θp0 shifts s`[a0].

Many of these copies of a0 are truncated by the restriction to{0, . . . , p0 − 1}. A relatively simple remedy is as follows:first, we zero-pad y0,...,p0−1 to length p = 3p0 − 2, giving[

0p0−1;y0; · · · ;yp0−1;0p0−1]. (3.6)

Zero padding provides enough space to accommodate anyshift s`[a0] with ` ∈ τ . We then perform one step of thegeneralized power method9, writing

a(0) = −PSp−1∇ϕ`1(PSp−1

[0p0−1;y0; · · · ;yp0−1;0p0−1

]),

(3.7)where PSp−1 projects onto the sphere. The reasoning behindthis construction may seem obscure, but can be clarifiedafter interpreting the gradient ∇ϕρ in terms of its actionon the shifts s`[a0] (see appendix). For now, we note thatthis operation has the effect of (approximately) filling in themissing pieces of the truncated shifts s`[a0] – see Figure 5for an example. We will prove that with high probabilitya(0) is indeed close to Σ4θp0 .

The next key observation is that the function ϕρ grows aswe move away from the subspace Sτ , as shown in Figure 3.

9The power method for minimizing a quadratic form ξ(a) =12a∗Ma over the sphere consists of the iteration a 7→−PSp−1Ma. Notice that in this mapping, −Ma = −∇ξ(a).The generalized power method, for minimizing a function ϕ overthe sphere consists of repeatedly projecting −∇ϕ onto the sphere,giving the iteration a 7→ −PSp−1∇ϕ(a). (3.7) can be interpretedas one step of the generalized power method for the objectivefunction ϕρ.

5

Page 6: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

Because of this, a small-stepping descent method will notmove far away from Σ4θp0 . For concreteness, we will ana-lyze a variant of the curvilinear search method (Goldfarb,1980; Goldfarb et al., 2017) , which moves in a linear combi-nation of the negative gradient direction −g and a negativecurvature direction −v. At the k-th iteration, the algorithmupdates a(k+1) as

a(k+1) ← PSp−1

[a(k) − tg(k) − t2v(k)

](3.8)

with appropriately chosen step size t. The inclusion of anegative curvature direction allows the method to avoidstagnation near saddle points. Indeed, we will prove thatstarting from initialization a(0), this method produces asequence a(1),a(2), . . . which efficiently converges to alocal minimizer a that is near some signed shift ±s`[a0] ofthe ground truth.

The second step of our algorithm rounds the local minimizera ≈ σs`[a0] to produce an exact solution a = σs`[a0].As a byproduct, it also exactly recovers the correspondingsigned shift of the true sparse signal, x = σs−`[x0].

Our rounding algorithm is an alternating minimizationscheme, which alternates between minimizing the Lassocost over a with x fixed, and minimizing the Lasso costover x with a fixed. We make two modifications to thisbasic idea, both of which are important for obtaining exactrecovery. First, unlike the standard Lasso cost, which penal-izes all of the entries of x, we maintain a running estimateI(k) of the support of x0, and only penalize those entriesthat are not in I(k):

12 ‖a ∗ x− y‖

22 + λ

∑i 6∈I(k) |xi| . (3.9)

This can be viewed as an extreme form of reweighting (Can-des et al., 2008). Second, our algorithm gradually decreasespenalty variable λ to 0, so that eventually

a ∗ x ≈ y. (3.10)

This can be viewed as a homotopy or continuation method(Osborne et al., 2000; Efron et al., 2004). For concreteness,at k-th iteration the algorithm reads:Update x:

x(k+1) ← argminx12‖a

(k) ∗ x− y‖22 + λ(k)∑i6∈I(k) |xi| ,

Update a:

a(k+1) ← PSp−1

[argmina

12‖a ∗ x

(k+1) − y‖22],

Update λ and I:

λ(k+1) ← 12λ

(k), I(k+1) ← supp(x(k+1)

). (3.11)

We prove that the iterates produced by this sequence ofoperations converge to the ground truth at a linear rate, aslong as the initializer a is sufficiently nearby.

Our overall algorithm is summarized as Algorithm 1. Fig-ure 6 illustrates the main steps of this algorithm. Our main

Initial a(0) a(100) Converged a Est. a & True a0

Figure 6. Local minimization and refinement. Data-driven initial-ization a(0) consists of a near-superposition of two shifts (left),and minimizing ϕρ produces a near shift of a0 as a (mid). Finallythe rounded solution a using the Lasso is almost identical to a shiftof a0 (right).

algorithmic result states that under closely related hypothe-ses as above, Algorithm 1 produces a signed shift of theground truth (a0,x0):

Algorithm 1 Short and Sparse Deconvolution

input Observation y, motif length p0, sparsity θ, shift-coherence µ, and curvature threshold −ηv .Minimization:Initialize a(0) ← −PSp−1∇ϕρ

(PSp−1

[0p0−1;y0; · · · ;

yp0−1;0p0−1])

, λ = 0.1/√p0θ

10and δ > 0 in ϕρ.for k = 1 to K1 do

a(k+1) ← PSp−1 [a(k) − tg(k) − t2v(k)]Here, g(k) is the Riemannian gradient; v(k) is theeigenvector of smallest Riemannian Hessian eigen-value if less then−ηv with

⟨v(k), g(k)

⟩≥ 0, otherwise

let v(k) = 0; and t ∈ (0, 0.1/nθ] satisfiesϕρ(a

(k+1)) < ϕρ(a(k))− 1

2 t‖g(k)‖22 − 1

4 t4ηv‖v(k)‖22

end forObtain a near local minimizer a← a(K1).Refinement:Initialize a(0) ← a, λ(0) ← 10(pθ + log n)(µ + 1/p)and I(0) ← Sλ(0) [supp(y

∧∗ a]).for k = 1 to K2 dox(k+1)←argminx

12‖a

(k)∗x−y‖22+λ(k)∑i 6∈I(k) |xi|

a(k+1)← PSp−1

[argmina

12‖a ∗ x

(k+1) − y‖22]

λ(k+1)← λ(k)/2, I(k+1) ←supp(x(k+1))end for

output (a, x)← (a(K2),x(K2))

Theorem 3.2 (Main Algorithmic Theorem). Suppose y =a0∗x0 where a0 ∈ Sp0−1 is µ-truncated shift coherent suchthat maxi 6=j

∣∣⟨ι∗p0si[a0], ι∗p0sj [a0]⟩∣∣ ≤ µ and x0 ∼i.i.d.

BG(θ) ∈ Rn with θ, µ satisfying

θ ∈

[c1p0,

c2(p0√µ+√p0)

log2 p0

], µ ≤ c3

log2 n(3.12)

for some constant c1, c2, c3 > 0. If the signal lengths n, p0satisfy n > poly(p0) and p0 > polylog(n), then thereexist δ, ηv > 0 such that with high probability, Algorithm 1

6

Page 7: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

produces (a, x) that are equal to the ground truth up tosigned shift symmetry:∥∥(a, x)− σ(s`[a0], s−`[x0]

)∥∥2≤ ε (3.13)

for some σ ∈ {−1, 1} and ` ∈ {−p0 + 1, . . . , p0 − 1} ifK1 > poly(n, p0) and K2 > polylog(n, p0, ε

−1).

Proof. See Appendix C.

3.3. Relationship to the Literature

Blind deconvolution is a classical problem in signal pro-cessing (Stockham et al., 1975; Cannon, 1976), and hasbeen studied under a variety of hypotheses. In this section,we first discuss the relationship between our results andthe existing literature on the short-and-sparse version ofthis problem, and then briefly discuss other deconvolutionvariants in the theoretical literature.

The short-and-sparse model arises in a number of applica-tions. One class of applications involves finding basic motifs(repeated patterns) in datasets. This motif discovery problemarises in extracellular spike sorting (Lewicki, 1998; Ekanad-ham et al., 2011) and calcium imaging (Pnevmatikakis et al.,2016), where the observed signal exhibits repetitive shortneuron excitation patterns occurring sparsely across timeand/or space. Similarly, electron microscopy images (Che-ung et al., 2017) arising in study of nanomaterials oftenexhibit repeated motifs.

Another significant application of SaS deconvolution is im-age deblurring. Typically, the blur kernel is small relativeto the image size (short) (Ayers & Dainty, 1988; You &Kaveh, 1996; Carasso, 2001; Levin et al., 2007; 2011). Innatural image deblurring, the target image is often assumedto have relatively few sharp edges (Fergus et al., 2006; Joshiet al., 2008; Levin et al., 2011), and hence have sparsederivatives. In scientific image deblurring, e.g., in astron-omy (Lane, 1992; Harmeling et al., 2009; Briers et al., 2013)and geophysics (Kaaresen & Taxt, 1998), the target imageis often sparse, either in the spatial or wavelet domains,again leading to variants of the SaS model. The literatureon blind image deconvolution is large; see, e.g., (Kundur &Hatzinakos, 1996; Campisi & Egiazarian, 2016) for surveys.

Variants of the SaS deconvolution problem arise in manyother areas of engineering as well. Examples include blindequalization in communications (Sato, 1975; Shalvi & Wein-stein, 1990; Johnson et al., 1998), dereverberation in soundengineering (Miyoshi & Kaneda, 1988; Naylor & Gaubitch,2010) and image super-resolution (Baker & Kanade, 2002;Shtengel et al., 2009; Yang et al., 2010).

These applications have motivated a great deal of algorith-

10In practice, we suggest setting λ = cλ/√p0θ with cλ ∈

[0.5, 0.8].

mic work on variants of the SaS problem (Lane & Bates,1987; Bones et al., 1995; Bell & Sejnowski, 1995; Kundur &Hatzinakos, 1996; Markham & Conchello, 1999; Campisi &Egiazarian, 2016; Walk et al., 2017). In contrast, relativelylittle theory is available to explain when and why algorithmssucceed. Our algorithm minimizes ϕρ as an approximationto the Lasso cost over the sphere. Our formulation andresults have strong precedent in the literature. Lasso-like ob-jective functions have been widely used in image deblurring(You & Kaveh, 1996; Chan & Wong, 1998; Fergus et al.,2006; Levin et al., 2007; Shan et al., 2008; Xu & Jia, 2010;Dong et al., 2011; Krishnan et al., 2011; Levin et al., 2011;Wipf & Zhang, 2014; Perrone & Favaro, 2014; Zhang et al.,2017). A number of insights have been obtained into thegeometry of sparse deconvolution – in particular, into theeffect of various constraints on a on the presence or absenceof spurious local minimizers. In image deblurring, a simplexconstraint (a ≥ 0 and ‖a‖1 = 1) arises naturally from thephysical structure of the problem (You & Kaveh, 1996; Chan& Wong, 1998). Perhaps surprisingly, simplex-constraineddeconvolution admits trivial global minimizers, at whichthe recovered kernel a is a spike, rather than the target blurkernel (Levin et al., 2011; Benichoux et al., 2013).

(Wipf & Zhang, 2014) imposes the `2 regularization on aand observes that this alternative constraint gives more re-liable algorithm. (Zhang et al., 2017) studies the geometryof the simplified objective ϕ`1 over the sphere, and provesthat in the dilute limit in which x0 has one nonzero entry,all strict local minima of ϕ`1 are close to signed shifts trun-cations of a0. By adopting a different objective function(based on `4 maximization) over the sphere, (Zhang et al.,2018) proves that on a certain region of the sphere everylocal minimum is near a truncated signed shift of a0, i.e.,the restriction of s`[a0] to the window {0, . . . , p0 − 1}. Theanalysis of (Zhang et al., 2018) allows the sparse sequencex0 to be denser (θ ∼ p

−2/30 for a generic kernel a0, as op-

posed to θ . p−3/40 in our result). Both (Zhang et al., 2017)

and (Zhang et al., 2018) guarantee approximate recoveryof a portion of s`[a0], under complicated conditions on thekernel a0. Our core optimization problem is very similar to(Zhang et al., 2017). However, we obtains exact recoveryof both a0 and relatively dense x0, under the much simplerassumption of shift incoherence.

Other aspects of the SaS problem have been studied theo-retically. One basic question is under what circumstancesthe problem is identifiable, up to the scaled shift ambiguity.(Choudhary & Mitra, 2015) shows that the problem ill-posedfor worst case (a0,x0) – in particular, for certain supportpatterns in which x0 does not have any isolated nonzeroentries. This demonstrates that some modeling assumptionson the support of the sparse term are needed. Nevertheless,this worst case structure is unlikely to occur, either underthe Bernoulli model, or in practical deconvolution problems.

7

Page 8: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

Motivated by a variety of applications, many low-dimensional deconvolution models have been studied inthe theoretical literature. In communication applications,the signals a0 and x0 either live in known low-dimensionalsubspaces, or are sparse in some known dictionary (Ahmedet al., 2014; Li et al., 2016; Chi, 2016; Ling & Strohmer,2015; Li et al., 2017; Ling & Strohmer, 2017; Kech &Krahmer, 2017). These theoretical works assume thatthe subspace / dictionary are chosen at random. Thislow-dimensional deconvolution model does not exhibit thesigned shift ambiguity; nonconvex formulations for thismodel exhibit a different structure from that studied here.In fact, the variant in which both signals belong to knownsubspaces can be solved by convex relaxation (Ahmed et al.,2014). The SaS model does not appear to be amenable toconvexification, and exhibits a more complicated nonconvexgeometry, due to the shift ambiguity. The main motivationfor tackling this model lies in the aforementioned applica-tions in imaging and data analysis.

(Wang & Chi, 2016; Li & Bresler, 2018) study the re-lated multi-instance sparse blind deconvolution problem(MISBD), where there are K observations yi = a0 ∗ xiconsisting of multiple convolutions i = 1, . . . ,K of a ker-nel a0 and different sparse vectors xi. Both works developprovable algorithms. There are several key differences withour work. First, both the proposed algorithms and theiranalysis require a0 to be invertible. Second, SaS model andMISBD are not equivalent despite the apparent similaritybetween them. It might seem possible to reduce SaS toMISBD by dividing the single observation y into K pieces;this apparent reduction fails due to boundary effects.

4. ExperimentsWe demonstrate that the tradeoffs between the motif lengthp0 and sparsity rate θ produce a transition region for success-ful SaS deconvolution under generic choices of a0 and x0.For fixed values of θ ∈ [10−3, 10−2] and p0 ∈ [103, 104],we draw 50 instances of synthetic data by choosing a0 ∼Unif(Sp0−1) and x0 ∈ Rn with x0 ∼i.i.d. BG(θ) wheren = 5 × 105. Note that choosing a0 this way impliesµ(a0) ≈ 1√

p0.

For each instance, we recover a0 and x0 from y = a0 ∗ x0

by minimizing problem (2.5). For ease of computation, wemodify Algorithm 1 by replacing curvilinear search withaccelerated Riemannian gradient descent method (See ap-pendix M). In Figure 7, we say the local minimizer amin issufficiently close to a solution of SaS deconvolution prob-lem, if

success(amin, ;a0) := {max` |〈s`[a0],amin〉| > 0.95 } .(4.1)

10-3 10-2.8 10-2.6 10-2.4 10-2.2 10-2

104

103.8

103.6

103.4

103.2

103

θ (log scale)

p0

(log

scal

e)

Figure 7. Success probability of SaS deconvolution under generica0, x0 with varying kernel length p0, and sparsity rate θ. Whensparsity rate decreases sufficiently with respect to kernel length,successful recovery becomes very likely (brighter), and vice versa(darker). A transition line is shown with slope log p0

log θ≈ −2,

implying our method works with high probability when θ / 1√p0

in generic case.

5. DiscussionThe main drawback of our proposed method is that it doesnot succeed when the target motif a0 has shift coherencevery close to 1. For instance, a common scenario in imageblind deconvolution involves deblurring an image with asmooth, low-pass point spread function (e.g., Gaussian blur).Both our analysis and numerical experiments show that inthis situation minimizing ϕρ does not find the generatingsignal pairs (a0,x0) consistently—the minimizer of ϕρ isoften spurious and is not close to any particular shift of a0.We do not suggest minimizing ϕρ in this situation. On theother hand, minimizing the bilinear lasso objective ϕlasso

over the sphere often succeeds even if the true signal pair(a0,x0) is coherent and dense.

In light of the above observations, we view the analysisof the bilinear lasso as the most important direction forfuture theoretical work on SaS deconvolution. The dropquadratic formulation studied here has commonalities withthe bilinear lasso: both exhibit local minima at signed shifts,and both exhibit negative curvature in symmetry breakingdirections. A major difference (and hence, major challenge)is that gradient methods for bilinear lasso do not retract toa union of subspaces – they retract to a more complicated,nonlinear set.

Finally, there are several directions in which our analysiscould be improved. Our lower bounds on the length n ofthe random vector x0 required for success are clearly sub-optimal. We also suspect our sparsity-coherence tradeoff be-tween µ, θ (roughly, θ / 1/(

õp0)) is suboptimal, even for

8

Page 9: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

the ϕρ objective. Articulating optimal sparsity-coherencetradeoffs for is another interesting direction for future work.

AcknowledgementsThe authors gratefully acknowledge support from NSF1343282, NSF CCF 1527809, and NSF IIS 1546411.

ReferencesAhmed, A., Recht, B., and Romberg, J. Blind deconvolu-

tion using convex programming. IEEE Transactions onInformation Theory, 60(3):1711–1732, 2014.

Ayers, G. and Dainty, J. C. Iterative blind deconvolutionmethod and its applications. Optics letters, 13(7):547–549, 1988.

Baker, S. and Kanade, T. Limits on super-resolution and howto break them. IEEE Transactions on Pattern Analysisand Machine Intelligence, 24(9):1167–1183, 2002.

Bell, A. J. and Sejnowski, T. J. An information-maximization approach to blind separation and blinddeconvolution. Neural computation, 7(6):1129–1159,1995.

Benichoux, A., Vincent, E., and Gribonval, R. A fundamen-tal pitfall in blind deconvolution with sparse and shift-invariant priors. In ICASSP-38th International Confer-ence on Acoustics, Speech, and Signal Processing-2013,2013.

Bones, P., Parker, C., Satherley, B., and Watson, R. De-convolution and phase retrieval with use of zero sheets.JOSA A, 12(9):1842–1857, 1995.

Briers, D., Duncan, D. D., Hirst, E. R., Kirkpatrick, S. J.,Larsson, M., Steenbergen, W., Stromberg, T., and Thomp-son, O. B. Laser speckle contrast imaging: theoreticaland practical limitations. Journal of biomedical optics,18(6):066018, 2013.

Campisi, P. and Egiazarian, K. Blind image deconvolution:theory and applications. CRC press, 2016.

Candes, E. J., Wakin, M. B., and Boyd, S. P. Enhancingsparsity by reweighted âDS 1 minimization. Journalof Fourier analysis and applications, 14(5-6):877–905,2008.

Cannon, M. Blind deconvolution of spatially invariant im-age blurs with phase. IEEE Transactions on Acoustics,Speech, and Signal Processing, 24(1):58–63, 1976.

Carasso, A. S. Direct blind deconvolution. SIAM Journalon Applied Mathematics, 61(6):1980–2007, 2001.

Chan, T. F. and Wong, C.-K. Total variation blind decon-volution. IEEE transactions on Image Processing, 7(3):370–375, 1998.

Cheung, S., Lau, Y., Chen, Z., Sun, J., Zhang, Y., Wright,J., and Pasupathy, A. Beyond the fourier transform: Anonconvex optimization approach to microscopy analysis.Submitted, 2017.

Chi, Y. Guaranteed blind sparse spikes deconvolution vialifting and convex optimization. IEEE Journal of SelectedTopics in Signal Processing, 10(4):782–794, June 2016.

Choudhary, S. and Mitra, U. Fundamental limits of blind de-convolution part ii: Sparsity-ambiguity trade-offs. arXivpreprint arXiv:1503.03184, 2015.

Dong, W., Zhang, L., Shi, G., and Wu, X. Image deblurringand super-resolution by adaptive sparse domain selectionand adaptive regularization. IEEE Transactions on ImageProcessing, 20(7):1838–1857, 2011.

Efron, B., Hastie, T., Johnstone, I., Tibshirani, R., et al.Least angle regression. The Annals of statistics, 32(2):407–499, 2004.

Ekanadham, C., Tranchina, D., and Simoncelli, E. P. Ablind sparse deconvolution method for neural spike iden-tification. In Advances in Neural Information ProcessingSystems 24, pp. 1440–1448. 2011.

Fergus, R., Singh, B., Hertzmann, A., Roweis, S. T., andFreeman, W. T. Removing camera shake from a singlephotograph. In ACM transactions on graphics (TOG),volume 25, pp. 787–794. ACM, 2006.

Goldfarb, D. Curvilinear path steplength algorithms forminimization which use directions of negative curvature.Mathematical programming, 18(1):31–40, 1980.

Goldfarb, D., Mu, C., Wright, J., and Zhou, C. Using neg-ative curvature in solving nonlinear programs. Compu-tational Optimization and Applications, 68(3):479–502,2017.

Harmeling, S., Hirsch, M., Sra, S., and Scholkopf, B. On-line blind deconvolution for astronomical imaging. In2009 IEEE International Conference on ComputationalPhotography (ICCP 2009), pp. 1–7. IEEE, 2009.

Johnson, R., Schniter, P., Endres, T. J., Behm, J. D., Brown,D. R., and Casas, R. A. Blind equalization using theconstant modulus criterion: A review. Proceedings of theIEEE, 86(10):1927–1950, 1998.

Joshi, N., Szeliski, R., and Kriegman, D. J. Psf estimationusing sharp edge prediction. In Computer Vision andPattern Recognition, 2008. CVPR 2008. IEEE Conferenceon, pp. 1–8. IEEE, 2008.

9

Page 10: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

Kaaresen, K. F. and Taxt, T. Multichannel blind deconvo-lution of seismic signals. Geophysics, 63(6):2093–2107,1998.

Kech, M. and Krahmer, F. Optimal injectivity conditionsfor bilinear inverse problems with applications to iden-tifiability of deconvolution problems. SIAM Journal onApplied Algebra and Geometry, 1(1):20–37, 2017.

Krishnan, D., Tay, T., and Fergus, R. Blind deconvolutionusing a normalized sparsity measure. In Computer Visionand Pattern Recognition (CVPR), 2011 IEEE Conferenceon, pp. 233–240. IEEE, 2011.

Kundur, D. and Hatzinakos, D. Blind image deconvolution.IEEE signal processing magazine, 13(3):43–64, 1996.

Lane, R. and Bates, R. Automatic multidimensional decon-volution. JOSA A, 4(1):180–188, 1987.

Lane, R. G. Blind deconvolution of speckle images. JOSAA, 9(9):1508–1514, 1992.

Levin, A., Fergus, R., Durand, F., and Freeman, W. T. De-convolution using natural image priors. MassachusettsInstitute of Technology, Computer Science and ArtificialIntelligence Laboratory, 3, 2007.

Levin, A., Weiss, Y., Durand, F., and Freeman, W. T. Un-derstanding blind deconvolution algorithms. IEEE trans-actions on pattern analysis and machine intelligence, 33(12):2354–2367, 2011.

Lewicki, M. S. A review of methods for spike sorting:the detection and classification of neural action potentials.Network: Computation in Neural Systems, 9(4):R53–R78,1998.

Li, Y. and Bresler, Y. Global geometry of multichannelsparse blind deconvolution on the sphere. arXiv preprintarXiv:1404.4104, 2018.

Li, Y., Lee, K., and Bresler, Y. Identifiability in blind de-convolution with subspace or sparsity constraints. IEEETransactions on Information Theory, 62(7):4266–4275,2016.

Li, Y., Lee, K., and Bresler, Y. Identifiability and stability inblind deconvolution under minimal assumptions. IEEETransaction of Information Theory, 2017.

Ling, S. and Strohmer, T. Self-calibration and biconvexcompressive sensing. Inverse Problems, 31(11):115002,2015.

Ling, S. and Strohmer, T. Blind deconvolution meets blinddemixing: Algorithms and performance bounds. IEEETransactions on Information Theory, 63(7):4497–4520,2017.

Markham, J. and Conchello, J.-A. Parametric blind decon-volution: a robust method for the simultaneous estimationof image and blur. JOSA A, 16(10):2377–2391, 1999.

Miyoshi, M. and Kaneda, Y. Inverse filtering of room acous-tics. IEEE Transactions on acoustics, speech, and signalprocessing, 36(2):145–152, 1988.

Naylor, P. A. and Gaubitch, N. D. Speech dereverberation.Springer Science & Business Media, 2010.

Osborne, M. R., Presnell, B., and Turlach, B. A. A newapproach to variable selection in least squares problems.IMA journal of numerical analysis, 20(3):389–403, 2000.

Perrone, D. and Favaro, P. Total variation blind deconvo-lution: The devil is in the details. In Proceedings ofthe IEEE Conference on Computer Vision and PatternRecognition, pp. 2909–2916, 2014.

Pnevmatikakis, E. A., Soudry, D., Gao, Y., Machado, T. A.,Merel, J., Pfau, D., Reardon, T., Mu, Y., Lacefield, C.,Yang, W., et al. Simultaneous denoising, deconvolution,and demixing of calcium imaging data. Neuron, 89(2):285–299, 2016.

Saha, S. K. Diffraction-limited imaging with large andmoderate telescopes. World Scientific, 2007.

Sato, Y. A method of self-recovering equalization for multi-level amplitude-modulation systems. IEEE Transactionson communications, 23(6):679–682, 1975.

Shalvi, O. and Weinstein, E. New criteria for blind de-convolution of nonminimum phase systems (channels).IEEE Transactions on information theory, 36(2):312–321,1990.

Shan, Q., Jia, J., and Agarwala, A. High-quality motiondeblurring from a single image. In Acm transactions ongraphics (tog), volume 27, pp. 73. ACM, 2008.

Shtengel, G., Galbraith, J. A., Galbraith, C. G., Lippincott-Schwartz, J., Gillette, J. M., Manley, S., Sougrat, R.,Waterman, C. M., Kanchanawong, P., Davidson, M. W.,et al. Interferometric fluorescent super-resolution mi-croscopy resolves 3d cellular ultrastructure. Proceedingsof the National Academy of Sciences, 106(9):3125–3130,2009.

Stockham, T. G., Cannon, T. M., and Ingebretsen, R. B.Blind deconvolution through digital signal processing.Proceedings of the IEEE, 63(4):678–692, 1975.

Walk, P., Jung, P., Pfander, G. E., and Hassibi, B. Blinddeconvolution with additional autocorrelations via convexprograms. arXiv preprint arXiv:1701.04890, 2017.

10

Page 11: Geometry and Symmetry in Short-and-Sparse …proceedings.mlr.press/v97/kuo19a/kuo19a.pdfGeometry and Symmetry in Short-and-Sparse Deconvolution Han-Wen Kuo 1 2Yuqian Zhang3 Yenson

Wang, L. and Chi, Y. Blind deconvolution from multiplesparse inputs. IEEE Signal Processing Letters, 23(10):1384–1388, 2016.

Wipf, D. and Zhang, H. Revisiting bayesian blind deconvo-lution. The Journal of Machine Learning Research, 15(1):3595–3634, 2014.

Xu, L. and Jia, J. Two-phase kernel estimation for robustmotion deblurring. In European conference on computervision, pp. 157–170. Springer, 2010.

Yang, J., Wright, J., Huang, T. S., and Ma, Y. Image super-resolution via sparse representation. IEEE transactionson image processing, 19(11):2861–2873, 2010.

You, Y.-L. and Kaveh, M. Anisotropic blind image restora-tion. In Image Processing, 1996. Proceedings., Inter-national Conference on, volume 2, pp. 461–464. IEEE,1996.

Zhang, Y., Lau, Y., Kuo, H.-w., Cheung, S., Pasupathy,A., and Wright, J. On the global geometry of sphere-constrained sparse blind deconvolution. In Proceedingsof the IEEE Conference on Computer Vision and PatternRecognition, pp. 4894–4902, 2017.

Zhang, Y., Kuo, H.-W., and Wright, J. Structured localoptima in sparse blind deconvolution. arXiv preprintarXiv:1806.00338, 2018.

11