Research Article -Goodness for Low-Rank Matrix Recovery Lingchen Kong, Levent Tunçel,

advertisement
Hindawi Publishing Corporation
Abstract and Applied Analysis
Volume 2013, Article ID 101974, 9 pages
http://dx.doi.org/10.1155/2013/101974
Research Article
𝑠-Goodness for Low-Rank Matrix Recovery
Lingchen Kong,1 Levent Tunçel,2 and Naihua Xiu1
1
2
Department of Applied Mathematics, Beijing Jiaotong University, Beijing 100044, China
Department of Combinatorics and Optimization, Faculty of Mathematics, University of Waterloo,
Waterloo, ON, Canada N2L 3G1
Correspondence should be addressed to Lingchen Kong; konglchen@126.com
Received 21 January 2013; Accepted 17 March 2013
Academic Editor: Jein-Shan Chen
Copyright © 2013 Lingchen Kong et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Low-rank matrix recovery (LMR) is a rank minimization problem subject to linear equality constraints, and it arises in many fields
such as signal and image processing, statistics, computer vision, and system identification and control. This class of optimization
problems is generally NP hard. A popular approach replaces the rank function with the nuclear norm of the matrix variable. In this
paper, we extend and characterize the concept of 𝑠-goodness for a sensing matrix in sparse signal recovery (proposed by Juditsky
and Nemirovski (Math Program, 2011)) to linear transformations in LMR. Using the two characteristic 𝑠-goodness constants, 𝛾𝑠 and
𝛾̂𝑠 , of a linear transformation, we derive necessary and sufficient conditions for a linear transformation to be 𝑠-good. Moreover, we
establish the equivalence of 𝑠-goodness and the null space properties. Therefore, 𝑠-goodness is a necessary and sufficient condition
for exact 𝑠-rank matrix recovery via the nuclear norm minimization.
1. Introduction
Low-rank matrix recovery (LMR for short) is a rank minimization problem (RMP) with linear constraints or the
affine matrix rank minimization problem which is defined as
follows:
rank (𝑋) ,
minimize
subject to
A𝑋 = 𝑏,
(1)
where 𝑋 ∈ Rπ‘š×𝑛 is the matrix variable, A : Rπ‘š×𝑛 → R𝑝
is a linear transformation, and 𝑏 ∈ R𝑝 . Although specific
instances can often be solved by specialized algorithms, the
LMR is NP hard. A popular approach for solving LMR
in the systems and control community is to minimize the
trace of a positive semidefinite matrix variable instead of its
rank (see, e.g., [1, 2]). A generalization of this approach to
nonsymmetric matrices introduced by Fazel et al. [3] is the
famous convex relaxation of LMR (1), which is called nuclear
norm minimization (NNM):
min
‖𝑋‖∗
s.t.
A𝑋 = 𝑏,
(2)
where ‖𝑋‖∗ is the nuclear norm of 𝑋, that is, the sum of
its singular values. When π‘š = 𝑛 and the matrix 𝑋 :=
Diag(π‘₯), π‘₯ ∈ R𝑛 , is diagonal, the LMR (1) reduces to
sparse signal recovery (SSR), which is the so-called cardinality
minimization problem (CMP):
min
β€–π‘₯β€–0
s.t.
Φπ‘₯ = 𝑏,
(3)
where β€–π‘₯β€–0 denotes the number of nonzero entries in the
vector π‘₯ and Φ ∈ Rπ‘š×𝑛 is a given sensing matrix. A
well-known heuristic for SSR is the β„“1 -norm minimization
relaxation (basis pursuit problem):
min
β€–π‘₯β€–1
s.t.
Φπ‘₯ = 𝑏,
(4)
where β€–π‘₯β€–1 is the β„“1 -norm of π‘₯, that is, the sum of absolute
values of its entries.
LMR problems have many applications and they appeared
in the literature of a diverse set of fields including signal
and image processing, statistics, computer vision, and system
identification and control. For more details, see the recent
2
Abstract and Applied Analysis
paper [4]. LMR and NNM have been the focus of some recent
research in the optimization community, see; for example, [4–
15]. Although there are many papers dealing with algorithms
for NNM such as interior-point methods, fixed point and
Bregman iterative methods, and proximal point methods,
there are fewer papers dealing with the conditions that
guarantee the success of the low-rank matrix recovery via
NNM. For instance, following the program laid out in the
work of CandeΜ€s and Tao in compressed sensing (CS, see, e.g.,
[16–18]), Recht et al. [4] provided a certain restricted isometry
property (RIP) condition on the linear transformation which
guarantees that the minimum nuclear norm solution is
the minimum rank solution. Recht et al. [14, 19] gave the
null space property (NSP) which characterizes a particular
property of the null space of the linear transformation, which
is also discussed by Oymak et al. [20, 21]. Note that NSP states
a necessary and sufficient condition for exactly recovering the
low-rank matrix via nuclear norm minimization. Recently,
Chandrasekaran et al. [22] proposed that a fixed 𝑠-rank
matrix 𝑋0 can be recovered if and only if the null space of
A does not intersect the tangent cone of the nuclear norm
ball at 𝑋0 .
In the setting of CS, there are other characterizations of
the sensing matrix, under which β„“1 -norm minimization can
be guaranteed to yield an optimal solution to SSR, in addition
to RIP and null-space properties, see; for example, [23–
26]. In particular, Juditsky and Nemirovski [24] established
necessary and sufficient conditions for a Sensing matrix to
be “𝑠-good” to allow for exact β„“1 -recovery of sparse signals
with 𝑠 nonzero entries when no measurement noise is present.
They also demonstrated that these characteristics, although
difficult to evaluate, lead to verifiable sufficient conditions
for exact SSR and to efficiently computable upper bounds on
those 𝑠 for which a given sensing matrix is 𝑠-good. Furthermore, they established instructive links between 𝑠-goodness
and RIP in the CS context. One may wonder whether we can
generalize the 𝑠-goodness concept to LMR and still maintain
many of the nice properties as done in [24]. Here, we deal
with this issue. Our approach is based on the singular value
decomposition (SVD) of a matrix and the partition technique
generalized from CS. In the next section, following Juditsky
and Nemirovski’s terminology, we propose definitions of 𝑠goodness and 𝐺-numbers, 𝛾𝑠 and 𝛾̂𝑠 , of a linear transformation in LMR and then we provide some basic properties of
𝐺-numbers. In Section 3, we characterize 𝑠-goodness of a
linear transformation in LMR via 𝐺-numbers. We consider
the connections between the 𝑠-goodness, NSP, and RIP in
Section 4. We eventually obtain that 𝛿2𝑠 < 0.472 ⇒ A
satisfying NSP ⇔ 𝛾̂𝑠 (A) < 1/2 ⇔ 𝛾𝑠 (A) < 1 ⇔ A is 𝑠-good.
Let π‘Š ∈ Rπ‘š×𝑛 , π‘Ÿ := min{π‘š, 𝑛}, and let π‘Š =
π‘ˆ Diag(𝜎(π‘Š))𝑉𝑇 be an SVD of π‘Š, where π‘ˆ ∈ Rπ‘š×π‘Ÿ , 𝑉 ∈
R𝑛×π‘Ÿ , and Diag(𝜎(π‘Š)) is the diagonal matrix of 𝜎(π‘Š) =
(𝜎1 (π‘Š), . . . , πœŽπ‘Ÿ (π‘Š))𝑇 which is the vector of the singular values
of π‘Š. Also let Ξ(π‘Š) denote the set of pairs of matrices (π‘ˆ, 𝑉)
in the SVD of π‘Š; that is,
Ξ (π‘Š) := { (π‘ˆ, 𝑉) : π‘ˆ ∈ Rπ‘š×π‘Ÿ , 𝑉 ∈ R𝑛×π‘Ÿ ,
π‘Š = π‘ˆ Diag (𝜎 (π‘Š)) 𝑉𝑇 } .
(5)
For 𝑠 ∈ {0, 1, 2, . . . , π‘Ÿ}, we say π‘Š ∈ Rπ‘š×𝑛 is an 𝑠-rank matrix
to mean that the rank of π‘Š is no more than 𝑠. For an 𝑠𝑇
rank matrix π‘Š, it is convenient to take π‘Š = π‘ˆπ‘š×𝑠 π‘Šπ‘  𝑉𝑛×𝑠
π‘š×𝑠
𝑛×𝑠
as its SVD where π‘ˆπ‘š×𝑠 ∈ R , 𝑉𝑛×𝑠 ∈ R are orthogonal
matrices and π‘Šπ‘  = Diag((𝜎1 (π‘Š), . . . , πœŽπ‘  (π‘Š))𝑇 ). For a vector
𝑦 ∈ R𝑝 , let β€– ⋅ ‖𝑑 be the dual norm of β€– ⋅ β€– specified by
‖𝑦‖𝑑 := maxV {⟨V, π‘¦βŸ© : β€–Vβ€– ≤ 1}. In particular, β€– ⋅ β€–∞ is the
dual norm of β€– ⋅ β€–1 for a vector. Let ‖𝑋‖ denote the spectral
or the operator norm of a matrix 𝑋 ∈ Rπ‘š×𝑛 , that is, the
largest singular value of 𝑋. In fact, ‖𝑋‖ is the dual norm of
‖𝑋‖∗ . Let ‖𝑋‖𝐹 := √βŸ¨π‘‹, π‘‹βŸ© = √Tr(𝑋𝑇 𝑋) be the Frobenius
norm of 𝑋, which is equal to the β„“2 -norm of the vector of
its singular values. We denote by 𝑋𝑇 the transpose of 𝑋. For
a linear transformation A : Rπ‘š×𝑛 → R𝑝 , we denote by
A∗ : R𝑝 → Rπ‘š×𝑛 the adjoint of A.
2. Definitions and Basic Properties
2.1. Definitions. We first go over some concepts related to 𝑠goodness of the linear transformation in LMR (RMP). These
are extensions of those given for SSR (CMP) in [24].
Definition 1. Let A : Rπ‘š×𝑛 → R𝑝 be a linear transformation
and 𝑠 ∈ {0, 1, 2, . . . , π‘Ÿ}. One says that A is 𝑠-good, if for every
𝑠-rank matrix π‘Š ∈ Rπ‘š×𝑛 , π‘Š is the unique optimal solution
to the optimization problem
min {‖𝑋‖∗ : A𝑋 = Aπ‘Š} .
𝑋∈Rπ‘š×𝑛
(6)
We denote by 𝑠∗ (A) the largest integer 𝑠 for which A
is 𝑠-good. Clearly, 𝑠∗ (A) ∈ {0, 1, . . . , π‘Ÿ}. To characterize 𝑠goodness we introduce two useful 𝑠-goodness constants: 𝛾𝑠 and
𝛾̂𝑠 . We call 𝛾𝑠 and 𝛾̂𝑠 𝐺-numbers.
Definition 2. Let A : Rπ‘š×𝑛 → R𝑝 be a linear transformation,
𝛽 ∈ [0, +∞] and 𝑠 ∈ {0, 1, 2, . . . , π‘Ÿ}. Then we have the
following.
(i) 𝐺-number 𝛾𝑠 (A, 𝛽) is the infimum of 𝛾 ≥ 0 such that
for every matrix 𝑋 ∈ Rπ‘š×𝑛 with singular value decomposi𝑇
tion 𝑋 = π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
(i.e., 𝑠 nonzero singular values, all equal
to 1), there exists a vector 𝑦 ∈ R𝑝 such that
σ΅„©σ΅„© σ΅„©σ΅„©
(7)
A∗ 𝑦 = π‘ˆ Diag (𝜎 (A∗ 𝑦)) 𝑉𝑇 ,
󡄩󡄩𝑦󡄩󡄩𝑑 ≤ 𝛽,
where π‘ˆ = [π‘ˆπ‘š×𝑠 π‘ˆπ‘š×(π‘Ÿ−𝑠) ], 𝑉 = [𝑉𝑛×𝑠 𝑉𝑛×(π‘Ÿ−𝑠) ] are orthogonal matrices, and
if πœŽπ‘– (𝑋) = 1,
{= 1,
πœŽπ‘– (A∗ 𝑦) {
∈ [0, 𝛾] , if πœŽπ‘– (𝑋) = 0,
{
𝑖 ∈ {1, 2, . . . , π‘Ÿ} .
(8)
If there does not exist such 𝑦 for some 𝑋 as above, we set
𝛾𝑠 (A, 𝛽) = +∞.
(ii) 𝐺-number 𝛾̂𝑠 (A, 𝛽) is the infimum of 𝛾 ≥ 0 such that
for every matrix 𝑋 ∈ Rπ‘š×𝑛 with 𝑠 nonzero singular values, all
equal to 1, there exists a vector 𝑦 ∈ R𝑝 such that A∗ 𝑦 and 𝑋
share the same orthogonal row and column spaces:
σ΅„©σ΅„© σ΅„©σ΅„©
σ΅„©σ΅„© ∗
σ΅„©σ΅„©
(9)
󡄩󡄩𝑦󡄩󡄩𝑑 ≤ 𝛽,
σ΅„©σ΅„©A 𝑦 − 𝑋󡄩󡄩 ≤ 𝛾.
Abstract and Applied Analysis
3
If there does not exist such 𝑦 for some 𝑋 as above, we set
𝛾𝑠 (A, 𝛽) = +∞ and to be compatible with the special case
given by [24] we write 𝛾𝑠 (A), 𝛾̂𝑠 (A) instead of 𝛾𝑠 (A, +∞),
𝛾̂𝑠 (A, +∞), respectively.
From the above definition, we easily see that the set of
values that 𝛾 takes is closed. Thus, when 𝛾𝑠 (A, 𝛽) < +∞, for
every matrix 𝑋 ∈ Rπ‘š×𝑛 with 𝑠 nonzero singular values, all
equal to 1, there exists a vector 𝑦 ∈ R𝑝 such that
σ΅„©σ΅„© σ΅„©σ΅„©
󡄩󡄩𝑦󡄩󡄩𝑑 ≤ 𝛽,
if πœŽπ‘– (𝑋) = 1,
= 1,
πœŽπ‘– (A 𝑦) {
∈ [0, 𝛾𝑠 (A, 𝛽)] ,
∗
if πœŽπ‘– (𝑋) = 0,
Similarly, for every matrix 𝑋 ∈ R
with 𝑠 nonzero singular
values, all equal to 1, there exists a vector 𝑦̂ ∈ R𝑝 such
that A∗ 𝑦̂ and 𝑋 share the same orthogonal row and column
spaces:
σ΅„©σ΅„© ∗ Μ‚
σ΅„©
σ΅„©σ΅„© Μ‚σ΅„©σ΅„©
(11)
σ΅„©σ΅„©A 𝑦 − 𝑋󡄩󡄩󡄩 ≤ 𝛾̂𝑠 (A, 𝛽) .
󡄩󡄩𝑦󡄩󡄩𝑑 ≤ 𝛽,
∗
Observing that the set {A 𝑦 : ‖𝑦‖𝑑 ≤ 𝛽} is convex, we obtain
that if 𝛾𝑠 (A, 𝛽) < +∞ then for every matrix 𝑋 with at most
𝑠 nonzero singular values and ‖𝑋‖ ≤ 1 there exist vectors 𝑦
satisfying (10) and there exist vectors 𝑦̂ satisfying (11).
2.2. Basic Properties of 𝐺-Numbers. In order to characterize
the 𝑠-goodness of a linear transformation A, we study the
basic properties of 𝐺-numbers. We begin with the result that
𝐺-numbers 𝛾𝑠 (A, 𝛽) and 𝛾̂𝑠 (A, 𝛽) are convex nonincreasing
functions of 𝛽.
Proposition 3. For every linear transformation A and every
𝑠 ∈ {0, 1, . . . , π‘Ÿ}, 𝐺-numbers 𝛾𝑠 (A, 𝛽) and 𝛾̂𝑠 (A, 𝛽) are convex
nonincreasing functions of 𝛽 ∈ [0, +∞].
Proof. We only need to demonstrate that the quantity
𝛾𝑠 (A, 𝛽) is a convex nonincreasing function of 𝛽 ∈ [0, +∞]. It
is evident from the definition that 𝛾𝑠 (A, 𝛽) is nonincreasing
for given A, 𝑠. It remains to show that 𝛾𝑠 (A, 𝛽) is a convex
function of 𝛽. In other words, for every pair 𝛽1 , 𝛽2 ∈ [0, +∞],
we need to verify that
𝛾𝑠 (A, 𝛼𝛽1 + (1 − 𝛼) 𝛽2 )
∀𝛼 ∈ [0, 1] .
(12)
The above inequality follows immediately if one of 𝛽1 , 𝛽2 is
+∞. Thus, we may assume 𝛽1 , 𝛽2 ∈ [0, +∞). In fact, from the
argument around (10) and the definition of 𝛾𝑠 (A, ⋅), we know
that for every matrix 𝑋 = π‘ˆ Diag(𝜎(𝑋))𝑉𝑇 with 𝑠 nonzero
singular values, all equal to 1, there exist vectors 𝑦1 , 𝑦2 ∈ R𝑝
such that for π‘˜ ∈ {1, 2}
σ΅„©σ΅„© σ΅„©σ΅„©
σ΅„©σ΅„©π‘¦π‘˜ 󡄩󡄩𝑑 ≤ π›½π‘˜ ,
𝑖 ∈ {1, 2, . . . , π‘Ÿ} .
rank (π‘Œπ‘˜ ) ≤ π‘Ÿ − 𝑠,
π‘‹π‘Œπ‘˜π‘‡ = 0,
σ΅„©σ΅„© σ΅„©σ΅„©
σ΅„©σ΅„©π‘Œπ‘˜ σ΅„©σ΅„© ≤ 𝛾𝑠 (A, π›½π‘˜ ) .
(13)
(14)
This implies that for every 𝛼 ∈ [0, 1]
(10)
𝑖 ∈ {1, 2, . . . , π‘Ÿ} .
= 1,
if πœŽπ‘– (𝑋) = 1,
πœŽπ‘– (A∗ π‘¦π‘˜ ) {
∈ [0, 𝛾𝑠 (A, π›½π‘˜ )] , if πœŽπ‘– (𝑋) = 0,
𝑋𝑇 π‘Œπ‘˜ = 0,
𝑋𝑇 [π›Όπ‘Œ1 + (1 − 𝛼) π‘Œ2 ] = 0,
π‘š×𝑛
≤ 𝛼𝛾𝑠 (A, 𝛽1 ) + (1 − 𝛼) 𝛾𝑠 (A, 𝛽2 ) ,
It is immediate from (13) that ‖𝛼𝑦1 + (1 − 𝛼)𝑦2 ‖𝑑 ≤ 𝛼𝛽1 + (1 −
𝛼)𝛽2 . Moreover, from the above information on the singular
values of A∗ 𝑦1 , A∗ 𝑦2 , we may set A∗ π‘¦π‘˜ = 𝑋 + π‘Œπ‘˜ , π‘˜ ∈ {1, 2}
such that
𝑇
𝑋[π›Όπ‘Œ1 + (1 − 𝛼)π‘Œ2 ] = 0,
(15)
and hence rank[π›Όπ‘Œ1 +(1−𝛼)π‘Œ2 ] ≤ π‘Ÿ−𝑠, 𝑋, and [π›Όπ‘Œ1 +(1−𝛼)π‘Œ2 ]
have orthogonal row and column spaces. Thus, noting that
A∗ [𝛼𝑦1 + (1 − 𝛼)𝑦2 ] = 𝑋 + π›Όπ‘Œ1 + (1 − 𝛼)π‘Œ2 , we obtain that
‖𝛼𝑦1 + (1 − 𝛼)𝑦2 ‖𝑑 ≤ 𝛼𝛽1 + (1 − 𝛼)𝛽2 and
πœŽπ‘– (A∗ (𝛼𝑦1 + (1 − 𝛼) 𝑦2 ))
if πœŽπ‘– (𝑋) = 1,
{1,
={
𝜎 (π›Όπ‘Œ1 + (1 − 𝛼) π‘Œ2 ) , if πœŽπ‘– (𝑋) = 0,
{ 𝑖
(16)
for every 𝛼 ∈ [0, 1]. Combining this with the fact
σ΅„© σ΅„©
σ΅„© σ΅„©
σ΅„©
σ΅„©σ΅„©
σ΅„©σ΅„©π›Όπ‘Œ1 + (1 − 𝛼) π‘Œ2 σ΅„©σ΅„©σ΅„© ≤ 𝛼 σ΅„©σ΅„©σ΅„©π‘Œ1 σ΅„©σ΅„©σ΅„© + (1 − 𝛼) σ΅„©σ΅„©σ΅„©π‘Œ2 σ΅„©σ΅„©σ΅„©
≤ 𝛼𝛾𝑠 (A, 𝛽1 ) + (1 − 𝛼) 𝛾𝑠 (A, 𝛽2 ) ,
(17)
we obtain the desired conclusion.
The following observation that 𝐺-numbers 𝛾𝑠 (A, 𝛽),
𝛾̂𝑠 (A, 𝛽) are nondecreasing in 𝑠 is immediate.
Proposition 4. For every 𝑠󸀠 ≤ 𝑠, one has 𝛾𝑠󸀠 (A, 𝛽) ≤ 𝛾𝑠 (A, 𝛽),
𝛾̂𝑠󸀠 (A, 𝛽) ≤ 𝛾̂𝑠 (A, 𝛽).
We further investigate the relationship between the 𝐺numbers 𝛾𝑠 (A, 𝛽) and 𝛾̂𝑠 (A, 𝛽).
Proposition 5. Let A : Rπ‘š×𝑛 → R𝑝 be a linear transformation, 𝛽 ∈ [0, +∞], and 𝑠 ∈ {0, 1, 2, . . . , π‘Ÿ}. Then one has
𝛾 := 𝛾𝑠 (A, 𝛽) < 1 󳨐⇒ 𝛾̂𝑠 (A,
=
1
𝛽)
1+𝛾
𝛾
1
< ,
1+𝛾 2
1
1
𝛽)
𝛾̂ := 𝛾̂𝑠 (A, 𝛽) < 󳨐⇒ 𝛾𝑠 (A,
2
1 − 𝛾̂
=
(18)
𝛾̂
< 1.
1 − 𝛾̂
Proof. Let 𝛾 := 𝛾𝑠 (A, 𝛽) < 1. Then, for every matrix 𝑍 ∈ Rπ‘š×𝑛
with 𝑠 nonzero singular values, all equal to 1, there exists
𝑦 ∈ R𝑝 , ‖𝑦‖𝑑 ≤ 𝛽, such that A∗ 𝑦 = 𝑍 + π‘Š, where β€–π‘Šβ€– ≤
𝛾 and π‘Š and 𝑍 have orthogonal row and column spaces.
4
Abstract and Applied Analysis
For a given pair 𝑍, 𝑦 as above, take 𝑦̃ := (1/(1 + 𝛾))𝑦. Then
Μƒ 𝑑 ≤ (1/(1 + 𝛾))𝛽 and
we have ‖𝑦‖
𝛾
𝛾
1
σ΅„©σ΅„© ∗ Μƒ
σ΅„©
,
}=
,
σ΅„©σ΅„©A 𝑦 − 𝑍󡄩󡄩󡄩 ≤ max {1 −
1+𝛾 1+𝛾
1+𝛾
(19)
where the first term under the maximum comes from the fact
that A∗ 𝑦 and 𝑍 agree on the subspace corresponding to the
nonzero singular values of 𝑍. Therefore, we obtain
𝛾̂𝑠 (A,
𝛾
1
1
𝛽) ≤
< .
1+𝛾
1+𝛾 2
(20)
Now, we assume that 𝛾̂ := 𝛾̂𝑠 (A, 𝛽) < 1/2. Fix orthogonal
matrices π‘ˆ ∈ Rπ‘š×π‘Ÿ , 𝑉 ∈ R𝑛×π‘Ÿ . For an 𝑠-element subset 𝐽 of
the index set {1, 2, . . . , π‘Ÿ}, we define a set 𝑆𝐽 with respect to
orthogonal matrices π‘ˆ, 𝑉 as
σ΅„© σ΅„©
𝑆𝐽 := {π‘₯ ∈ Rπ‘Ÿ : ∃𝑦 ∈ R𝑝 , 󡄩󡄩󡄩𝑦󡄩󡄩󡄩𝑑 ≤ 𝛽,
󡄨 󡄨
A∗ 𝑦 = π‘ˆ Diag (π‘₯) 𝑉𝑇 with 󡄨󡄨󡄨π‘₯𝑖 󡄨󡄨󡄨 ≤ 𝛾̂ for 𝑖 ∈ 𝐽} .
(21)
In the above, 𝐽 denotes the complement of 𝐽. It is immediately
seen that 𝑆𝐽 is a closed convex set in Rπ‘Ÿ . Moreover, we have
the following
Claim 1. 𝑆𝐽 contains the β€– ⋅ β€–∞ -ball of radius (1 − 𝛾̂) centered
at the origin in Rπ‘Ÿ .
Proof. Note that 𝑆𝐽 is closed and convex. Moreover, 𝑆𝐽 is the
direct sum of its projections onto the pair of subspaces
where π‘Š and π‘ˆ Diag(𝑧)𝑉𝑇 have the same orthogonal
row and column spaces, β€–A∗ 𝑦 − π‘ˆ Diag(𝑧)𝑉𝑇 β€– ≤ 𝛾̂ and
β€–πœŽ(A∗ 𝑦) − 𝑧‖∞ ≤ 𝛾̂. Together with the definitions of 𝑆𝐽 and
𝑄, this means that 𝑄 contains a vector V with |V𝑖 − sign(𝑒𝑖 )| ≤
𝛾̂, ∀𝑖 ∈ {1, 2, . . . , 𝑠}. Therefore,
𝑠
󡄨 󡄨
𝑒𝑇 V ≥ ∑ 󡄨󡄨󡄨𝑒𝑖 󡄨󡄨󡄨 (1 − 𝛾̂) = (1 − 𝛾̂) ‖𝑒‖1 = 1 − 𝛾̂.
By V ∈ 𝐡𝑠 and the definition of 𝑒, we obtain
1 − 𝛾̂ ≥ β€–Vβ€–∞ = ‖𝑒‖1 β€–Vβ€–∞ ≥ 𝑒𝑇 V > 𝑒𝑇V ≥ 1 − 𝛾̂,
Using the above claim, we conclude that for every 𝐽 ⊆
{1, 2, . . . , π‘Ÿ} with cardinality 𝑠, there exists an π‘₯ ∈ 𝑆𝐽 such that
π‘₯𝑖 = (1 − 𝛾̂), for all 𝑖 ∈ 𝐽. From the definition of 𝑆𝐽 , we obtain
that there exists 𝑦 ∈ R𝑝 with ‖𝑦‖𝑑 ≤ (1 − 𝛾̂)−1 𝛽 such that
A∗ 𝑦 = π‘ˆ Diag (𝜎 (A∗ 𝑦)) 𝑉𝑇 ,
(22)
Let 𝑄 denote the projection of 𝑆𝐽 onto 𝐿 𝐽 . Then, 𝑄 is closed
and convex (because of the direct sum property above and
the fact that 𝑆𝐽 is closed and convex). Note that 𝐿 𝐽 can be
naturally identified with R𝑠 , and our claim is the image 𝑄 ⊂
R𝑠 of 𝑄 under this identification that contains the β€– ⋅ β€–∞ ball 𝐡𝑠 of radius (1 − 𝛾̂) centered at the origin in R𝑠 . For a
contradiction, suppose 𝐡𝑠 is not contained in 𝑄. Then there
exists V ∈ 𝐡𝑠 \𝑄. Since 𝑄 is closed and convex, by a separating
hyperplane theorem, there exists a vector 𝑒 ∈ R𝑠 , ‖𝑒‖1 = 1
such that
for every VσΈ€  ∈ 𝑄.
(23)
Let 𝑧 ∈ Rπ‘Ÿ be defined by
{1,
𝑧𝑖 := {
0,
{
𝑖 ∈ 𝐽,
otherwise.
(24)
By definition of 𝛾̂ = 𝛾̂𝑠 (A, 𝛽), for 𝑠-rank matrix π‘ˆ Diag(𝑧)𝑉𝑇 ,
there exists 𝑦 ∈ R𝑝 such that ‖𝑦‖𝑑 ≤ 𝛽 and
A∗ 𝑦 = π‘ˆ Diag (𝑧) 𝑉𝑇 + π‘Š,
(28)
where πœŽπ‘– (A∗ 𝑦) = (1 − 𝛾̂)−1 π‘₯𝑖 = 1 if 𝑖 ∈ 𝐽, and πœŽπ‘– (A∗ 𝑦)𝑖 ≤
(1 − 𝛾̂)−1 𝛾̂ if 𝑖 ∈ 𝐽. Thus, we obtain that
𝛾̂𝑠 := 𝛾̂𝑠 (A, 𝛽) <
𝛾̂
1
1
󳨐⇒ 𝛾𝑠 (A,
𝛽) ≤
< 1.
2
1 − 𝛾̂
1 − 𝛾̂
(29)
𝐿⊥𝐽 = {π‘₯ ∈ Rπ‘Ÿ : π‘₯𝑖 = 0, 𝑖 ∈ 𝐽} .
𝑒𝑇 V > 𝑒 𝑇 V σΈ€ 
(27)
where the strict inequality follows from the facts that V ∈ 𝑄
and 𝑒 separates V from 𝑄. The above string of inequalities is a
contradiction, and hence the desired claim holds.
𝐿 𝐽 := {π‘₯ ∈ Rπ‘Ÿ : π‘₯𝑖 = 0, 𝑖 ∈ 𝐽}
and its orthogonal complement
(26)
𝑖=1
(25)
To conclude the proof, we need to prove that the inequalities we established
𝛾̂𝑠 (A,
𝛾
1
𝛽) ≤
,
1 + 𝛾̂
1+𝛾
𝛾𝑠 (A,
𝛾̂
1
𝛽) ≤
1 − 𝛾̂
1 + 𝛾̂
(30)
are both equations. This is straightforward by an argument
similar to the one in the proof of [24, Theorem 1]. We omit it
for the sake of brevity.
We end this section with a simple argument which
illustrates that for a given pair (A, 𝑠), 𝛾𝑠 (A, 𝛽) = 𝛾𝑠 (A) and
𝛾̂𝑠 (A, 𝛽) = 𝛾̂𝑠 (A), for all 𝛽 large enough.
Proposition 6. Let A : Rπ‘š×𝑛 → R𝑝 be a linear transformation and 𝛽 ∈ [0, +∞]. Assume that for some 𝜌 > 0, the image
of the unit β€– ⋅ β€–∗ -ball in Rπ‘š×𝑛 under the mapping 𝑋 󳨃→ A𝑋
contains the ball 𝐡 = {π‘₯ ∈ R𝑝 : β€–π‘₯β€–1 ≤ 𝜌}. Then for every
𝑠 ∈ {1, 2, . . . , π‘Ÿ}
𝛽≥
1
,
𝜌
𝛾𝑠 (A) < 1 󳨐⇒ 𝛾𝑠 (A, 𝛽) = 𝛾𝑠 (A) ,
𝛽≥
1
,
𝜌
𝛾̂𝑠 (A) <
1
󳨐⇒ 𝛾̂𝑠 (A, 𝛽) = 𝛾̂𝑠 (A) .
2
(31)
Abstract and Applied Analysis
5
Proof. Fix 𝑠 ∈ {1, 2, . . . , π‘Ÿ}. We only need to show the first
implication. Let 𝛾 := 𝛾𝑠 (A) < 1. Then for every matrix π‘Š ∈
𝑇
Rπ‘š×𝑛 with its SVD π‘Š = π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
, there exists a vector 𝑦 ∈
𝑝
R such that
σ΅„©σ΅„© σ΅„©σ΅„©
󡄩󡄩𝑦󡄩󡄩𝑑 ≤ 𝛽,
A∗ 𝑦 = π‘ˆ Diag (𝜎 (A∗ 𝑦)) 𝑉𝑇 ,
if πœŽπ‘– (π‘Š) = 1,
if πœŽπ‘– (π‘Š) = 0,
(33)
𝑖 ∈ {1, 2, . . . , π‘Ÿ} .
Clearly, β€–A∗ 𝑦‖ ≤ 1. That is,
σ΅„©
σ΅„©
1 ≥ σ΅„©σ΅„©σ΅„©A∗ 𝑦󡄩󡄩󡄩 = max
{βŸ¨π‘‹, A∗ π‘¦βŸ© : ‖𝑋‖∗ ≤ 1}
𝑋∈Rπ‘š×𝑛
{βŸ¨π‘’, π‘¦βŸ© : 𝑒 = A𝑋, ‖𝑋‖∗ ≤ 1} .
= max
π‘š×𝑛
{= 1,
πœŽπ‘– (A∗ 𝑦) {
∈ 1] ,
{ [0,
(32)
where π‘ˆ = [π‘ˆπ‘š×𝑠 π‘ˆπ‘š×(π‘Ÿ−𝑠) ], 𝑉 = [𝑉𝑛×𝑠 𝑉𝑛×(π‘Ÿ−𝑠) ] are orthogonal matrices, and
{= 1,
πœŽπ‘– (A∗ 𝑦) {
∈ [0, 𝛾] ,
{
it follows that there exist matrices π‘ˆπ‘š×(π‘Ÿ−𝑠) , 𝑉𝑛×(π‘Ÿ−𝑠) such that
A∗ 𝑦 = π‘ˆ Diag(πœŽπ‘– (A∗ 𝑦))𝑉𝑇 where π‘ˆ = [π‘ˆπ‘š×𝑠 π‘ˆπ‘š×(π‘Ÿ−𝑠) ],
𝑉 = [𝑉𝑛×𝑠 𝑉𝑛×(π‘Ÿ−𝑠) ] are orthogonal matrices and
where 𝐽 := {𝑖 : πœŽπ‘– (π‘Š) =ΜΈ 0} and 𝐽 := {1, 2, . . . , π‘Ÿ} \ 𝐽. Therefore,
the optimal objective value of the optimization problem
if 𝑖 ∈ 𝐽, }
{
{= 1,
∗
∗
min
𝑦
∈
πœ•β€–π‘Šβ€–
,
𝜎
(A
𝑦)
𝛾
:
A
∗
𝑖
}
{
𝑦,𝛾 {
∈ [0, 𝛾] , if 𝑖 ∈ 𝐽,
{
}
{
(38)
(34)
=
Π := {conv {𝑀 ∈ Rπ‘š×𝑛 : the SVD of 𝑀 is
𝑇
0
0
𝑀 = [π‘ˆπ‘š×𝑠 π‘ˆπ‘š×(π‘Ÿ−𝑠) ] ( 0𝑠 𝜎 (𝑀)) [𝑉𝑛×𝑠 𝑉𝑛×(π‘Ÿ−𝑠) ] } .
(39)
From the inclusion assumption, we obtain that
max {βŸ¨π‘’, π‘¦βŸ© : 𝑒 = A𝑋, ‖𝑋‖∗ ≤ 1}
σ΅„© σ΅„©
σ΅„© σ΅„©
≥ max𝑝 {βŸ¨π‘’, π‘¦βŸ© : ‖𝑒‖1 ≤ 𝜌} = πœŒσ΅„©σ΅„©σ΅„©π‘¦σ΅„©σ΅„©σ΅„©∞ = πœŒσ΅„©σ΅„©σ΅„©π‘¦σ΅„©σ΅„©σ΅„©π‘‘ .
𝑒∈R
(37)
if 𝑖 ∈ 𝐽,
is at most one. For the given π‘Š with its SVD π‘Š
𝑇
, let
π‘ˆπ‘š×𝑠 π‘Šπ‘  𝑉𝑛×𝑠
𝑋∈R
𝑋∈Rπ‘š×𝑛
if 𝑖 ∈ 𝐽,
(35)
Combining the above two strings of relations, we derive the
desired conclusion.
It is easy to see that Π is a subspace and its normal
cone (in the sense of variational analysis, see, e.g., [28] for
details) is specified by Π⊥ . Thus, the above problem (38)
is equivalent to the following convex optimization problem
with set constraint:
𝑇
− 𝑀 = 0, 𝑀 ∈ Π} .
min {‖𝑀‖ : A∗ 𝑦 − π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
3. 𝑠-Goodness and 𝐺-Numbers
𝑦,𝑀
(40)
We first give the following characterization result of 𝑠-goodness of a linear transformation A via the 𝐺-number 𝛾𝑠 (A),
which explains the importance of 𝛾𝑠 (A) in LMR.
We will show that the optimal value is less than 1. For a
contradiction, suppose that the optimal value is one. Then, by
[28, Theorem 10.1 and Exercise 10.52], there exists a Lagrange
multiplier 𝐷 ∈ Rπ‘š×𝑛 such that the function
Theorem 7. Let A : Rπ‘š×𝑛 → R𝑝 be a linear transformation,
and 𝑠 be an integer 𝑠 ∈ {0, 1, 2, . . . , π‘Ÿ}. Then A is 𝑠-good if and
only if 𝛾𝑠 (A) < 1.
𝑇
− π‘€βŸ© + 𝛿Р(𝑀)
𝐿 (𝑦, 𝑀) = ‖𝑀‖ + ⟨𝐷, A∗ 𝑦 − π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
(41)
Proof. Suppose A is 𝑠-good. Let π‘Š ∈ Rπ‘š×𝑛 be a matrix of
rank 𝑠 ∈ {1, 2, . . . , π‘Ÿ}. Without loss of generality, let π‘Š =
𝑇
be its SVD where π‘ˆπ‘š×𝑠 ∈ Rπ‘š×𝑠 , 𝑉𝑛×𝑠 ∈ R𝑛×𝑠 are
π‘ˆπ‘š×𝑠 π‘Šπ‘  𝑉𝑛×𝑠
orthogonal matrices and π‘Šπ‘  = Diag((𝜎1 (π‘Š), . . . , πœŽπ‘  (π‘Š))𝑇 ).
By the definition of 𝑠-goodness of A, π‘Š is the unique
solution to the optimization problem (6). Using the firstorder optimality conditions, we obtain that there exists 𝑦 ∈
R𝑝 such that the function 𝑓𝑦 (π‘₯) = ‖𝑋‖∗ − 𝑦𝑇 [A𝑋 − Aπ‘Š]
attains its minimum value over 𝑋 ∈ Rπ‘š×𝑛 at 𝑋 = π‘Š. So
0 ∈ πœ•π‘“π‘¦ (π‘Š) or A∗ 𝑦 ∈ πœ•β€–π‘Šβ€–∗ . Using the fact (see, e.g., [27])
𝑇
πœ•β€–π‘Šβ€–∗ = {π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
+ 𝑀 : π‘Š and 𝑀 have orthogonal
row and column spaces, and ‖𝑀‖ ≤ 1} ,
(36)
has unconstrained minimum in (𝑦, 𝑀) equal to 1, where 𝛿Р(⋅)
is the indicator function of Π. Let (𝑦∗ , 𝑀∗ ) be an optimal
solution. Then, by the optimality condition 0 ∈ πœ•πΏ, we obtain
that
0 ∈ πœ•π‘¦ 𝐿 (𝑦∗ , 𝑀∗ ) ,
0 ∈ πœ•π‘€πΏ (𝑦∗ , 𝑀∗ ) .
(42)
Direct calculation yields that
A𝐷 = 0,
σ΅„© σ΅„©
0 ∈ −𝐷 + πœ• 󡄩󡄩󡄩𝑀∗ σ΅„©σ΅„©σ΅„© + Π⊥ .
(43)
Then there exist 𝐷𝐽 ∈ Π⊥ and 𝐷𝐽 ∈ πœ•β€–π‘€∗ β€– such that
𝐷 = 𝐷𝐽 + 𝐷𝐽 . Notice that [29, Corollary 6.4] implies that
for 𝐷𝐽 ∈ πœ•β€–π‘€∗ β€–, 𝐷𝐽 ∈ Π and ‖𝐷𝐽 β€–∗ ≤ 1. Therefore,
𝑇
𝑇
⟨𝐷, π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
⟩ = ⟨𝐷𝐽 , π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
⟩ and ⟨𝐷, 𝑀∗ ⟩ = ⟨𝐷𝐽 , 𝑀∗ ⟩.
Moreover, ⟨𝐷𝐽 , 𝑀∗ ⟩ ≤ ‖𝑀∗ β€– by the definition of the dual
6
Abstract and Applied Analysis
norm of β€– ⋅ β€–. This together with the facts A𝐷 = 0, 𝐷𝐽 ∈ Π⊥
and 𝐷𝐽 ∈ πœ•β€–π‘€∗ β€– ⊆ Π yields
σ΅„© σ΅„©
𝐿 (𝑦∗ , 𝑀∗ ) = 󡄩󡄩󡄩𝑀∗ σ΅„©σ΅„©σ΅„© − ⟨𝐷𝐽 , 𝑀∗ ⟩ + ⟨𝐷, A∗ 𝑦∗ ⟩
𝑇
− ⟨𝐷𝐽 , π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
⟩ + 𝛿Р(𝑀∗ )
(44)
𝑇
≥ −⟨𝐷𝐽 , π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
⟩ + 𝛿Р(𝑀∗ ) .
Thus, the minimum value of 𝐿(𝑦, 𝑀) is attained, 𝐿(𝑦∗ , 𝑀∗ ) =
𝑇
−⟨𝐷𝐽 , π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
⟩, when 𝑀∗ ∈ Π, ⟨𝐷𝐽 , 𝑀∗ ⟩ = ‖𝑀∗ β€–. We
obtain that ‖𝐷𝐽 β€–∗ = 1. By assumption, 1 = 𝐿(𝑦∗ , 𝑀∗ ) =
𝑇
𝑇
−⟨𝐷𝐽 , π‘ˆπ‘š×𝑠 𝑉𝑛×𝑠
⟩. That is, ∑𝑠𝑖=1 (π‘ˆπ‘š×𝑠
𝐷𝑉𝑛×𝑠 )𝑖𝑖 = −1. Without
loss of generality, let SVD of the optimal 𝑀∗ be 𝑀∗ =
Μƒ (0𝑠 0 ∗ ) 𝑉
Μƒπ‘š×(π‘Ÿ−𝑠) ] and 𝑉
̃𝑇 , where π‘ˆ
Μƒ := [π‘ˆπ‘š×𝑠 π‘ˆ
Μƒ :=
π‘ˆ
0 𝜎(𝑀 )
̃𝑛×(π‘Ÿ−𝑠) ]. From the above arguments, we obtain that
[𝑉𝑛×𝑠 𝑉
(i) A𝐷 = 0,
𝑇
̃𝑇 𝐷𝑉)
Μƒ 𝑖𝑖 = −1,
(ii) ∑𝑠𝑖=1 (π‘ˆπ‘š×𝑠
𝐷𝑉𝑛×𝑠 )𝑖𝑖 = ∑𝑖∈𝐽 (π‘ˆ
̃𝑇
Μƒ 𝑖𝑖 = 1.
(iii) ∑𝑖∈𝐽 (π‘ˆ 𝐷𝑉)
Clearly, for every 𝑑 ∈ R, the matrices 𝑋𝑑 := π‘Š+𝑑𝐷 are feasible
in (6). Note that
𝑇
π‘Š = π‘ˆπ‘š×𝑠 π‘Šπ‘  𝑉𝑛×𝑠
= [π‘ˆπ‘š×𝑠
𝑇
Μƒπ‘š×(π‘Ÿ−𝑠) ] (π‘Šπ‘  0) [𝑉𝑛×𝑠 𝑉
̃𝑛×(π‘Ÿ−𝑠) ] .
π‘ˆ
0 0
(45)
̃𝑇 π‘Šπ‘‰β€–
Μƒ ∗ = Tr(π‘ˆ
̃𝑇 π‘Šπ‘‰).
Μƒ From the above
Then, β€–π‘Šβ€–∗ = β€–π‘ˆ
equations, we obtain that ‖𝑋𝑑 β€–∗ = β€–π‘Šβ€–∗ for all small enough
𝑑 > 0 (since πœŽπ‘– (π‘Š) > 0, 𝑖 ∈ {1, 2, . . . , 𝑠}). Noting that π‘Š is
the unique optimal solution to (6), we have 𝑋𝑑 = π‘Š, which
Μƒ 𝑖𝑖 = 0 for 𝑖 ∈ 𝐽. This is a contradiction,
̃𝑇 𝐷𝑉)
means that (π‘ˆ
and hence the desired conclusion holds.
We next prove that A is 𝑠-good if 𝛾𝑠 (A) < 1. That is,
we let π‘Š be an 𝑠-rank matrix and we show that π‘Š is the
unique optimal solution to (6). Without loss of generality,
𝑇
let π‘Š be a matrix of rank 𝑠󸀠 =ΜΈ 0 and π‘ˆπ‘š×𝑠󸀠 π‘Šπ‘ σΈ€  𝑉𝑛×𝑠
σΈ€  its SVD,
π‘š×𝑠󸀠
𝑛×𝑠󸀠
where π‘ˆπ‘š×𝑠󸀠 ∈ R
, 𝑉𝑛×𝑠󸀠 ∈ R
are orthogonal matrices
and π‘Šπ‘ σΈ€  = Diag((𝜎1 (π‘Š), . . . , πœŽπ‘ σΈ€  (π‘Š))𝑇 ). It follows from
Proposition 4 that 𝛾𝑠󸀠 (A) ≤ 𝛾𝑠 (A) < 1. By the definition of 𝛾𝑠 (A), there exists 𝑦 ∈ R𝑝 such that A∗ 𝑦 =
π‘ˆ Diag(𝜎(A∗ 𝑦))𝑉𝑇 , where π‘ˆ = [π‘ˆπ‘š×𝑠󸀠 π‘ˆπ‘š×(π‘Ÿ−𝑠󸀠 ) ], 𝑉 =
[𝑉𝑛×𝑠󸀠 𝑉𝑛×(π‘Ÿ−𝑠󸀠 ) ], and
if πœŽπ‘– (π‘Š) =ΜΈ 0,
{= 1,
πœŽπ‘– (A∗ 𝑦) {
∈ 1) , if πœŽπ‘– (π‘Š) = 0.
{ [0,
(46)
Now, we have the optimization problem of minimizing the
function
𝑓 (𝑋) = ‖𝑋‖∗ − 𝑦𝑇 [A𝑋 − Aπ‘Š]
∗
= ‖𝑋‖∗ − ⟨A 𝑦, π‘‹βŸ© + β€–π‘Šβ€–∗
(47)
over all 𝑋 ∈ Rπ‘š×𝑛 such that A𝑋 = Aπ‘Š. Note that
⟨A∗ 𝑦, π‘‹βŸ© ≤ ‖𝑋‖∗ by β€–A∗ 𝑦‖ ≤ 1 and the definition of dual
norm. So 𝑓(𝑋) ≥ ‖𝑋‖∗ − ‖𝑋‖∗ + β€–π‘Šβ€–∗ = β€–π‘Šβ€–∗ and this
function attains its unconstrained minimum in 𝑋 at 𝑋 = π‘Š.
Hence 𝑋 = π‘Š is an optimal solution to (6). It remains to show
that this optimal solution is unique. Let 𝑍 be another optimal
solution to the problem. Then 𝑓(𝑍)−𝑓(π‘Š) = ‖𝑍‖∗ −𝑦𝑇 A𝑍 =
‖𝑍‖∗ − ⟨A∗ 𝑦, π‘βŸ© = 0. This together with the fact β€–A∗ 𝑦‖ ≤ 1
implies that there exist SVDs for A∗ 𝑦 and 𝑍 such that
Μƒ Diag (𝜎 (A∗ 𝑦)) 𝑉
̃𝑇 ,
A∗ 𝑦 = π‘ˆ
(48)
Μƒ Diag (𝜎 (𝑍)) 𝑉
̃𝑇 ,
𝑍=π‘ˆ
Μƒ ∈ R𝑛×π‘Ÿ are orthogonal matrices,
Μƒ ∈ Rπ‘š×π‘Ÿ and 𝑉
where π‘ˆ
∗
and πœŽπ‘– (𝑍) = 0 if πœŽπ‘– (A 𝑦) =ΜΈ 1. Thus, for πœŽπ‘– (A∗ 𝑦) = 0, for all
𝑖 ∈ {𝑠󸀠 + 1, . . . , π‘Ÿ}, we must have πœŽπ‘– (𝑍) = πœŽπ‘– (π‘Š) = 0. By the
𝑇
̃𝑇
Μƒ
two forms of SVDs of A∗ 𝑦 as above, π‘ˆπ‘š×𝑠󸀠 𝑉𝑛×𝑠
σΈ€  = π‘ˆπ‘š×𝑠󸀠 𝑉𝑛×𝑠󸀠
𝑇
Μƒπ‘š×𝑠󸀠 , 𝑉
Μƒ σΈ€  are the corresponding submatrices of π‘ˆ,
Μƒ 𝑉,
Μƒ
where π‘ˆ
𝑛×𝑠
respectively. Without loss of generality, let
π‘ˆ = [𝑒1 , 𝑒2 , . . . , π‘’π‘Ÿ ] ,
𝑉 = [V1 , V2 , . . . , Vπ‘Ÿ ] ,
Μƒ = [Μƒ
π‘ˆ
𝑒1 , 𝑒̃2 , . . . , π‘’Μƒπ‘Ÿ ] ,
Μƒ = [ΜƒV1 , ΜƒV2 , . . . , ΜƒVπ‘Ÿ ] ,
𝑉
(49)
where 𝑒𝑗 = 𝑒̃𝑗 and V𝑗 = ΜƒV𝑗 for the corresponding index 𝑗 ∈
{𝑖 : πœŽπ‘– (A∗ 𝑦) = 0, 𝑖 ∈ {𝑠󸀠 + 1, . . . , π‘Ÿ}}. Then we have
𝑍=
𝑠󸀠
𝑠󸀠
∑πœŽπ‘– (𝑍) 𝑒̃𝑖 ΜƒV𝑖𝑇 ,
𝑖=1
π‘Š = ∑πœŽπ‘– (π‘Š) 𝑒𝑖 V𝑖𝑇 .
(50)
𝑖=1
𝑇
̃𝑇
Μƒ
From π‘ˆπ‘š×𝑠󸀠 𝑉𝑛×𝑠
σΈ€  = π‘ˆπ‘š×𝑠󸀠 𝑉𝑛×𝑠󸀠 , we obtain that
π‘Ÿ
π‘Ÿ
𝑖=𝑠󸀠 +1
𝑖=𝑠󸀠 +1
∑ πœŽπ‘– (A∗ 𝑦) 𝑒̃𝑖 ΜƒV𝑖𝑇 = ∑ πœŽπ‘– (A∗ 𝑦) 𝑒𝑖 V𝑖𝑇 .
(51)
Therefore, we deduce
π‘Ÿ
πœŽπ‘– (A∗ 𝑦) 𝑒̃𝑖 ΜƒV𝑖𝑇 +
∑
𝑖=𝑠󸀠 +1,πœŽπ‘– (A∗ 𝑦) =ΜΈ 0
π‘Ÿ
=
∑
π‘Ÿ
∑
𝑖=𝑠󸀠 +1,πœŽπ‘– (A∗ 𝑦)=0
πœŽπ‘– (A∗ 𝑦) 𝑒𝑖 V𝑖𝑇 +
𝑖=𝑠󸀠 +1,πœŽπ‘– (A∗ 𝑦) =ΜΈ 0
𝑒̃𝑖 ΜƒV𝑖𝑇
π‘Ÿ
∑
𝑖=𝑠󸀠 +1,πœŽπ‘– (A∗ 𝑦)=0
𝑒𝑖 V𝑖𝑇
=: Ω.
(52)
Clearly, the rank of Ω is no less than π‘Ÿ − 𝑠󸀠 ≥ π‘Ÿ − 𝑠. From the
Μƒ 𝑉,
Μƒ we easily derive that
orthogonality property of π‘ˆ, 𝑉 and π‘ˆ,
Ω𝑇 𝑒̃𝑖 ΜƒV𝑖𝑇 = 0,
Ω𝑇 𝑒𝑖 V𝑖𝑇 = 0,
∀𝑖 ∈ {1, 2, . . . , 𝑠󸀠 } .
(53)
Thus, we obtain Ω𝑇 (𝑍 − π‘Š) = 0, which implies that the rank
of the matrix 𝑍 − π‘Š is no more than 𝑠. Since 𝛾𝑠 (A) < 1, there
exists 𝑦̃ such that
{= 1,
Μƒ {
πœŽπ‘– (A∗ 𝑦)
∈ 1)
{ [0,
if πœŽπ‘– (𝑍 − π‘Š) =ΜΈ 0,
if πœŽπ‘– (𝑍 − π‘Š) = 0.
(54)
Μƒ 𝑍 − π‘ŠβŸ© = ‖𝑍 − π‘Šβ€–∗ .
Therefore, 0 = 𝑦̃𝑇 A(𝑍 − π‘Š) = ⟨A∗ 𝑦,
Then 𝑍 = π‘Š.
Abstract and Applied Analysis
7
For the 𝐺-number 𝛾̂𝑠 (A), we directly obtain the following
equivalent theorem of 𝑠-goodness from Proposition 5 and
Theorem 7.
Theorem 8. Let A : Rπ‘š×𝑛 → R𝑝 be a linear transformation,
and 𝑠 ∈ {1, 2, . . . , π‘Ÿ}. Then A is 𝑠-good if and only if 𝛾̂𝑠 (A) <
1/2.
4. 𝑠-Goodness, NSP, and RIP
This section deals with the connections between 𝑠-goodness,
the null space property (NSP), and the restricted isometry
property (RIP). We start with establishing the equivalence of
NSP and 𝐺-number 𝛾̂𝑠 (A) < 1/2. Here, we say A satisfies
NSP if for every nonzero matrix 𝑋 ∈ Null(A) with the SVD
𝑋 = π‘ˆ Diag(𝜎(𝑋))𝑉𝑇 , then we have
𝑠
π‘Ÿ
𝑖=1
𝑖=𝑠+1
∑πœŽπ‘– (𝑋) < ∑ πœŽπ‘– (𝑋) .
(55)
For further details, see, for example, [14, 19–21] and references
therein.
Proposition 9. For the linear transformation A, 𝛾̂𝑠 (A) < 1/2
if and only if A satisfies NSP.
Proof. We first give an equivalent representation of the 𝐺number 𝛾̂𝑠 (A, 𝛽). We define a compact convex set first:
𝑃𝑠 := {𝑍 ∈ Rπ‘š×𝑛 : ‖𝑍‖∗ ≤ 𝑠, ‖𝑍‖ ≤ 1} .
(56)
Let 𝐡𝛽 := {𝑦 ∈ R𝑝 : ‖𝑦‖𝑑 ≤ 𝛽} and 𝐡 := {𝑋 ∈ Rπ‘š×𝑛 :
‖𝑋‖ ≤ 1}. By definition, 𝛾̂𝑠 (A, 𝛽) is the smallest 𝛾 such that
the closed convex set 𝐢𝛾,𝛽 := A∗ 𝐡𝛽 + 𝛾𝐡 contains all matrices
with 𝑠 nonzero singular values, all equal to 1. Equivalently,
𝐢𝛾,𝛽 contains the convex hull of these matrices, namely, 𝑃𝑠 .
Note that 𝛾 satisfies the inclusion 𝑃𝑠 ⊆ 𝐢𝛾,𝛽 if and only if for
every 𝑋 ∈ Rπ‘š×𝑛
max βŸ¨π‘, π‘‹βŸ© ≤ max βŸ¨π‘Œ, π‘‹βŸ©
𝑍∈𝑃𝑠
π‘Œ∈𝐢𝛾,𝛽
=
max
𝑦∈R𝑝 ,π‘Š∈Rπ‘š×𝑛
{βŸ¨π‘‹, A∗ π‘¦βŸ© + π›ΎβŸ¨π‘‹, π‘ŠβŸ©
σ΅„© σ΅„©
: 󡄩󡄩󡄩𝑦󡄩󡄩󡄩𝑑 ≤ 𝛽, β€–π‘Šβ€– ≤ 1}
For 𝑋 ∈ Rπ‘š×𝑛 with A𝑋 = 0, let 𝑋 = π‘ˆ Diag(𝜎(𝑋))𝑉𝑇
be its SVD. Then, we obtain the sum of the 𝑠 largest singular
values of 𝑋 as
‖𝑋‖𝑠,∗ = maxβŸ¨π‘, π‘‹βŸ©.
𝑍∈𝑃𝑠
(60)
From (59), we immediately obtain that 𝛾̂𝑠 (A) is the best upper
bound on ‖𝑋‖𝑠,∗ of matrices 𝑋 ∈ Null(A) such that ‖𝑋‖∗ ≤
1. Therefore, 𝛾̂𝑠 (A) < 1/2 implies that the maximum value
of β€– ⋅ ‖𝑠,∗ -norms of matrices 𝑋 ∈ Null(A) with ‖𝑋‖∗ = 1
is less than 1/2. That is, ∑𝑠𝑖=1 πœŽπ‘– (𝑋) < 1/2 ∑π‘Ÿπ‘–=1 πœŽπ‘– (𝑋). Thus,
∑𝑠𝑖=1 πœŽπ‘– (𝑋) < ∑π‘Ÿπ‘–=𝑠+1 πœŽπ‘– (𝑋) and hence A satisfies NSP. Now,
it is easy to see that A satisfies NSP if and only if 𝛾̂𝑠 (A) <
1/2.
Next, we consider the connection between restricted
isometry constants and 𝐺-number of the linear transformation in LMR. It is well known that, for a nonsingular matrix
(transformation) 𝑇 ∈ R𝑝×𝑝 , the RIP constants of A and
𝑇A can be very different, as shown by Zhang [30] for the
vector case. However, the 𝑠-goodness properties of A and
𝑇A are always the same for a nonsingular transformation
𝑇 ∈ R𝑝×𝑝 (i.e., 𝑠-goodness properties enjoy scale invariance
in this sense). Recall that the 𝑠-restricted isometry constant
𝛿𝑠 of a linear transformation A is defined as the smallest
constant such that the following holds for all 𝑠-rank matrices
𝑋 ∈ Rπ‘š×𝑛 :
(1 − 𝛿𝑠 ) ‖𝑋‖2𝐹 ≤ β€–A𝑋‖22 ≤ (1 + 𝛿𝑠 ) ‖𝑋‖2𝐹 .
(61)
In this case, we say A possesses the RI (𝛿𝑠 )-property (RIP) as
in the CS context. For details, see [4, 31–34] and the references
therein.
Proposition 10. Let A : Rπ‘š×𝑛 → R𝑝 be a linear
transformation and 𝑠 ∈ {0, 1, 2, . . . , π‘Ÿ}. For any nonsingular
transformation 𝑇 ∈ R𝑝×𝑝 , 𝛾̂𝑠 (A) = 𝛾̂𝑠 (𝑇A).
Proof. It follows from the nonsingularity of 𝑇 that {𝑋 : A𝑋 =
0} = {𝑋 : 𝑇A𝑋 = 0}. Then, by the equivalent representation
of the 𝐺-number 𝛾̂𝑠 (A, 𝛽) in (59),
𝛾̂𝑠 (A) = max {βŸ¨π‘, π‘‹βŸ© : 𝑍 ∈ 𝑃𝑠 , ‖𝑋‖∗ ≤ 1, A𝑋 = 0}
(57)
𝑍,𝑋
= max {βŸ¨π‘, π‘‹βŸ© : 𝑍 ∈ 𝑃𝑠 , ‖𝑋‖∗ ≤ 1, 𝑇A𝑋 = 0}
𝑍,𝑋
= 𝛾̂𝑠 (𝑇A) .
= 𝛽 β€–A𝑋‖ + 𝛾‖𝑋‖∗ .
(62)
For the above, we adopt the convention that whenever 𝛽 =
+∞, 𝛽‖A𝑋‖ is defined to be +∞ or 0 depending on whether
β€–A𝑋‖ > 0 or β€–A𝑋‖ = 0. Thus, 𝑃𝑠 ⊆ 𝐢𝛾,𝛽 if and only if
max𝑍∈𝑃𝑠 {βŸ¨π‘, π‘‹βŸ© − 𝛽‖A𝑋‖} ≤ 𝛾‖𝑋‖∗ . Using the homogeneity
of this last relation with respect to 𝑋, the above is equivalent
to
max {βŸ¨π‘, π‘‹βŸ© − 𝛽 β€–A𝑋‖ : 𝑍 ∈ 𝑃𝑠 , ‖𝑋‖∗ ≤ 1} ≤ 𝛾.
(58)
For the RIP constant 𝛿2𝑠 , Oymak et al. [21] gave the
current best bound on the restricted isometry constant
𝛿2𝑠 < 0.472, where they proposed a general technique for
translating results from SSR to LMR. Together with the above
arguments, we immediately obtain the following theorem.
Therefore, we obtain 𝛾̂𝑠 (A, 𝛽) = max𝑍,𝑋 {βŸ¨π‘, π‘‹βŸ© − 𝛽‖A𝑋‖ :
𝑍 ∈ 𝑃𝑠 , ‖𝑋‖∗ ≤ 1}. Furthermore,
Theorem 11. 𝛿2𝑠 < 0.472 ⇒ A satisfying 𝑁𝑆𝑃 ⇔ 𝛾̂𝑠 (A) <
1/2 ⇔ 𝛾𝑠 (A) < 1 ⇔ A is 𝑠-good.
𝑍,𝑋
𝛾̂𝑠 (A) = max {βŸ¨π‘, π‘‹βŸ© : 𝑍 ∈ 𝑃𝑠 , ‖𝑋‖∗ ≤ 1, A𝑋 = 0} .
𝑍,𝑋
(59)
Proof. It follows from [21, Theorem 1], Proposition 9, and
Theorems 7 and 8.
8
The above theorem says that 𝑠-goodness is a necessary
and sufficient condition for recovering the low-rank solution
exactly via nuclear norm minimization.
5. Conclusion
In this paper, we have shown that 𝑠-goodness of the linear
transformation in LMR is a necessary and sufficient conditions for exact 𝑠-rank matrix recovery via the nuclear norm
minimization, which is equivalent to the null space property.
Our analysis is based on the two characteristic 𝑠-goodness
constants, 𝛾𝑠 and 𝛾̂𝑠 , and the variational property of matrix
norm in convex optimization. This shows that 𝑠-goodness is
an elegant concept for low-rank matrix recovery, although
𝛾𝑠 and 𝛾̂𝑠 may not be easy to compute. Development of
efficiently computable bounds on these quantities is left to
future work. Even though we develop and use techniques
based on optimization, convex analysis, and geometry, we do
not provide explicit analogues to the results of Donoho [35]
where necessary and sufficient conditions for vector recovery
special case were derived based on the geometric notions
of face preservation and neighborliness. The corresponding
generalization to low-rank recovery is not known, currently
the closest one being [22]. Moreover, it is also important
to consider the semidefinite relaxation (SDR) for the rank
minimization with the positive semidefinite constraint since
the SDR convexifies nonconvex or discrete optimization
problems by removing the rank-one constraint. Another
future research topic is to extend the main results and the
techniques in this paper to the SDR.
Acknowledgments
The authors thank two anonymous referees for their very
useful comments. The work was supported in part by the
National Basic Research Program of China (2010CB732501),
the National Natural Science Foundation of China (11171018),
a Discovery Grant from NSERC, a research grant from the
University of Waterloo, ONR research Grant N00014-1210049, and the Fundamental Research Funds for the Central
Universities (2011JBM306, 2013JBM095).
References
[1] C. Beck and R. D’Andrea, “Computational study and comparisons of LFT reducibility methods,” in Proceedings of the
American Control Conference, Philadelphia, Pa, USA, June 1998.
[2] M. Mesbahi and G. P. Papavassilopoulos, “On the rank minimization problem over a positive semidefinite linear matrix
inequality,” IEEE Transactions on Automatic Control, vol. 42, no.
2, pp. 239–243, 1997.
[3] M. Fazel, H. Hindi, and S. P. Boyd, “A rank minimization heuristic with application to minimum order system approximation,”
in Proceedings of American Control Conference, pp. 4734–4739,
June 2001.
[4] B. Recht, M. Fazel, and P. A. Parrilo, “Guaranteed minimumrank solutions of linear matrix equations via nuclear norm
minimization,” SIAM Review, vol. 52, no. 3, pp. 471–501, 2010.
Abstract and Applied Analysis
[5] B. P. W. Ames and S. A. Vavasis, “Nuclear norm minimization
for the planted clique and biclique problems,” Mathematical
Programming, vol. 129, no. 1, pp. 69–89, 2011.
[6] J.-F. Cai, E. J. CandeΜ€s, and Z. Shen, “A singular value thresholding algorithm for matrix completion,” SIAM Journal on
Optimization, vol. 20, no. 4, pp. 1956–1982, 2010.
[7] E. J. CandeΜ€s and B. Recht, “Exact matrix completion via convex
optimization,” Foundations of Computational Mathematics, vol.
9, no. 6, pp. 717–772, 2009.
[8] C. Ding, D. Sun, and K. C. Toh, “An introduction to a class of
matrix cone programming,” Tech. Rep., 2010.
[9] Z. Lin, M. Chen, L. Wu, and Y. Ma, “The augmented Lagrange
multiplier method for exact recovery of a corrupted low-rank
matrices,” http://arxiv.org/abs/1009.5055.
[10] Y. Liu, D. Sun, and K. C. Toh, “An implementable proximal
point algorithmic framework for nuclear norm minimization,”
Mathematical Programming, vol. 133, no. 1-2, pp. 399–436, 2012.
[11] Z. Liu and L. Vandenberghe, “Interior-point method for nuclear
norm approximation with application to system identification,”
SIAM Journal on Matrix Analysis and Applications, vol. 31, no.
3, pp. 1235–1256, 2009.
[12] S. Ma, D. Goldfarb, and L. Chen, “Fixed point and Bregman
iterative methods for matrix rank minimization,” Mathematical
Programming, vol. 128, no. 1-2, pp. 321–353, 2011.
[13] Y. Peng, A. Ganesh, J. Wright, W. Xu, and Y. Ma, “RASL: robust
alignment by sparse and low-rank decomposition for linearly
correlated images,” IEEE Transactions on Pattern Analysis and
Machine Intelligence, vol. 34, no. 1, pp. 2233–2246, 2010.
[14] B. Recht, W. Xu, and B. Hassibi, “Necessary and sufficient
conditions for success of the nuclear norm heuristic for rank
minimization,” in Proceedings of the 47th IEEE Conference
on Decision and Control (CDC ’08), pp. 3065–3070, Cancun,
Mexico, December 2008.
[15] M. Tao and X. Yuan, “Recovering low-rank and sparse components of matrices from incomplete and noisy observations,”
SIAM Journal on Optimization, vol. 21, no. 1, pp. 57–81, 2011.
[16] E. J. CandeΜ€s, J. Romberg, and T. Tao, “Robust uncertainty
principles: exact signal reconstruction from highly incomplete
frequency information,” IEEE Transactions on Information Theory, vol. 52, no. 2, pp. 489–509, 2006.
[17] E. J. Candes and T. Tao, “Decoding by linear programming,”
IEEE Transactions on Information Theory, vol. 51, no. 12, pp.
4203–4215, 2005.
[18] D. L. Donoho, “Compressed sensing,” IEEE Transactions on
Information Theory, vol. 52, no. 4, pp. 1289–1306, 2006.
[19] B. Recht, W. Xu, and B. Hassibi, “Null space conditions and
thresholds for rank minimization,” Mathematical Programming,
vol. 127, no. 1, pp. 175–202, 2011.
[20] S. Oymak and B. Hassibi, “New null space results and recovery
thresholds for matrix rank minimization,” http://arxiv.org/abs/
1011.6326.
[21] S. Oymak, K. Mohan, M. Fazel, and B. Hassibi, “A simplified
approach to recovery conditions for low-rank matrices,” in
Proceedings of International Symposium on Information Theory
(ISIT ’11), pp. 2318–2322, August 2011.
[22] V. Chandrasekaran, B. Recht, P. A. Parrilo, and A. S. Willsky,
“The convex geometry of linear inverse problems,” in Proceedings of the 48th Annual Allerton Conference on Communication,
Control, and Computing, 2011.
[23] A. d’Aspremont and L. El Ghaoui, “Testing the nullspace property using semidefinite programming,” Mathematical Programming, vol. 127, no. 1, pp. 123–144, 2011.
Abstract and Applied Analysis
[24] A. Juditsky and A. Nemirovski, “On verifiable sufficient conditions for sparse signal recovery via β„“1 minimization,” Mathematical Programming, vol. 127, no. 1, pp. 57–88, 2011.
[25] A. Juditsky, F. Karzan, and A. Nemirovski, “Verifiable conditions of β„“1 -recovery for sparse signals with sign restrictions,”
Mathematical Programming, vol. 127, no. 1, pp. 89–122, 2011.
[26] A. Juditsky, F. Karzan, and A. S. Nemirovski, “Accuracy guarantees for β„“1 -recovery,” IEEE Transactions on Information Theory,
vol. 57, no. 12, pp. 7818–7839, 2011.
[27] G. A. Watson, “Characterization of the subdifferential of some
matrix norms,” Linear Algebra and Its Applications, vol. 170, pp.
33–45, 1992.
[28] R. T. Rockafellar and R. J.-B. Wets, Variational Analysis,
Springer, New York, NY, USA, 2nd edition, 2004.
[29] A. S. Lewis and H. S. Sendov, “Nonsmooth analysis of singular
values. II. Applications,” Set-Valued Analysis, vol. 13, no. 3, pp.
243–264, 2005.
[30] Y. Zhang, “Theory of Compressive Sensing via 𝐿 1 -Minimization: A Non-RIP Analysis and Extensions,” 2008.
[31] E. J. CandeΜ€s and Y. Plan, “Tight oracle inequalities for lowrank matrix recovery from a minimal number of noisy random
measurements,” IEEE Transactions on Information Theory, vol.
57, no. 4, pp. 2342–2359, 2011.
[32] K. Lee and Y. Bresler, “Guaranteed minimum rank approximation from linear observations by nuclear norm minimization
with an ellipsoidal constraint,” http://arxiv.org/abs/0903.4742.
[33] R. Meka, P. Jain, and I. S. Dhillon, “Guaranteed rank minimization via singular value projection,” http://arxiv.org/abs/
0909.5457.
[34] K. Mohan and M. Fazel, “New restricted isometry results for
noisy low-rank matrix recovery,” in Proceedings of IEEE International Symposium on Information Theory (ISIT ’10), Austin,
Tex, USA, June 2010.
[35] D. L. Donoho, “Neighborly polytopes and sparse solution of
underdetermined linear equations,” Tech. Rep., 2004.
9
Advances in
Operations Research
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Advances in
Decision Sciences
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Mathematical Problems
in Engineering
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Algebra
Hindawi Publishing Corporation
http://www.hindawi.com
Probability and Statistics
Volume 2014
The Scientific
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International Journal of
Differential Equations
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Volume 2014
Submit your manuscripts at
http://www.hindawi.com
International Journal of
Advances in
Combinatorics
Hindawi Publishing Corporation
http://www.hindawi.com
Mathematical Physics
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Complex Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International
Journal of
Mathematics and
Mathematical
Sciences
Journal of
Hindawi Publishing Corporation
http://www.hindawi.com
Stochastic Analysis
Abstract and
Applied Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
International Journal of
Mathematics
Volume 2014
Volume 2014
Discrete Dynamics in
Nature and Society
Volume 2014
Volume 2014
Journal of
Journal of
Discrete Mathematics
Journal of
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Applied Mathematics
Journal of
Function Spaces
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Optimization
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Download