• 沒有找到結果。

Approximation Bounds for Quadratic Optimization with Homogeneous Quadratic Constraints∗

N/A
N/A
Protected

Academic year: 2022

Share "Approximation Bounds for Quadratic Optimization with Homogeneous Quadratic Constraints∗"

Copied!
35
0
0

加載中.... (立即查看全文)

全文

(1)

Approximation Bounds for Quadratic Optimization with Homogeneous Quadratic Constraints

Zhi-Quan Luo, Nicholas D. Sidiropoulos, Paul Tseng§, and Shuzhong Zhang

June 16, 2006

Abstract

We consider the NP-hard problem of finding a minimum norm vector in n-dimensional real or complex Euclidean space, subject to m concave homogeneous quadratic con- straints. We show that a semidefinite programming (SDP) relaxation for this noncon- vex quadratically constrained quadratic program (QP) provides an O(m2) approxima- tion in the real case, and an O(m) approximation in the complex case. Moreover, we show that these bounds are tight up to a constant factor. When the Hessian of each constraint function is of rank 1 (namely, outer products of some given so-called steer- ing vectors) and the phase spread of the entries of these steering vectors are bounded away from π/2, we establish a certain “constant factor” approximation (depending on the phase spread but independent of m and n) for both the SDP relaxation and a convex QP restriction of the original NP-hard problem. Finally, we consider a re- lated problem of finding a maximum norm vector subject to m convex homogeneous quadratic constraints. We show that a SDP relaxation for this nonconvex QP provides an O(1/ ln(m)) approximation, which is analogous to a result of Nemirovski, Roos and Terlaky [14] for the real case.

The first author is supported in part by the National Science Foundation, Grant No. DMS-0312416, and by the Natural Sciences and Engineering Research Council of Canada, Grant No. OPG0090391. The second author is supported in part by the U.S. ARO under ERO, Contract No. N62558-03-C-0012, and the EU under U-BROAD STREP, Grant No. 506790. The third author is supported by the National Science Foundation, Grant No. DMS-0511283. The fourth author is supported by Hong Kong RGC Earmarked Grant CUHK418505.

Department of Electrical and Computer Engineering, University of Minnesota, 200 Union Street SE, Minneapolis, MN 55455, U.S.A. (luozq@ece.umn.edu)

Department of Electronic and Computer Engineering, Technical University of Crete, 73100 Chania - Crete, Greece. (nikos@telecom.tuc.gr)

§Department of Mathematics, University of Washington, Seattle, Washington 98195, U.S.A.

(tseng@math.washington.edu)

Department of Systems Engineering and Engineering Management, The Chinese University of Hong

(2)

1 Introduction

Consider the quadratic optimization problem with concave homogeneous quadratic con- straints:

υqp := min kzk2 s.t. X

`∈Ii

|hH` z|2 ≥ 1, i = 1, ..., m, z ∈ IFn,

(1)

where IF is either IR or IC, k · k denotes the Euclidean norm in IFn, m ≥ 1, each h` is a given vector in IFn, and I1, ..., Im are nonempty, mutually disjoint index sets satisfying I1 ∪ · · · ∪ Im = {1, ..., M }. Throughout, the superscript “H” will denote the complex Hermitian transpose, i.e., for z = x + iy, where x, y ∈ IRn and i2 = −1, zH = xT − iyT. Geometrically, the above problem (1) corresponds to finding a least norm vector in a region defined by the intersection of the exteriors of m co-centered ellipsoids. If the vectors h1, ..., hM are linearly independent, then M equals the sum of the rank of the matrices defining these m ellipsoids. Notice that the problem (1) is easily solved for the case of n = 1, so we assume n ≥ 2.

We assume that P`∈Iikh`k 6= 0 for all i, which is clearly a necessary condition for (1) to be feasible. This is also a sufficient condition (sinceSmi=1{z | P`∈Ii|hH` z|2 = 0} is a finite union of proper subspaces of IFn, so its complement is nonempty and any point in its complement can be scaled to be feasible for (1)). Thus, the above problem (1) always has an optimal solution (not necessarily unique) since its objective function is coercive, continuous, and its feasible set is nonempty, closed. Notice, however, that the feasible set of (1) is typically nonconvex and disconnected, with an exponential number of connected components exhibiting little symmetry. This is in contrast to the quadratic problems with convex feasible set but nonconvex objective function considered in [13, 14, 22]. Furthermore, unlike the class of quadratic problems studied in [1, 7, 8, 15, 16, 21, 23, 24, 25, 26], the constraint functions in (1) do not depend on z21, ..., zn2 only.

Our interest in the nonconvex QP (1) is motivated by the transmit beamforming problem for multicasting applications [20] and by the wireless sensor network localization problem [6]. In the transmit beamforming problem, a transmitter utilizes an array of n transmitting antennas to broadcast information within its service area to m radio receivers, with receiver i ∈ {1, ..., m} equipped with |Ii| receiving antennas. Let h`, ` ∈ Ii, denote the n × 1 complex steering vector modelling propagation loss and phase shift from the transmitting antennas to the `th receiving antenna of receiver i. Assuming that each receiver performs spatially matched filtering / maximum ratio combining, which is the

(3)

optimal combining strategy under standard mild assumptions, then the constraint X

`∈Ii

|hH` z|2 ≥ 1

models the requirement that the total received signal power at receiver i must be above a given threshold (normalized to 1). This constraint is also equivalent to a signal-to-noise ratio (SNR) condition commonly used in data communication. Thus, to minimize the total transmit power subject to individual SNR requirements (one at each receiver), we are led to the QP (1). In the special case where each radio receiver is equipped with a single receiving antenna, the problem reduces to [20]:

min kzk2

s.t. |hH` z|2 ≥ 1, ` = 1, ..., m, z ∈ IFn,

(2)

This problem is a special case of (1) whereby each ellipsoid lies in IFnand the corresponding matrix has rank 1.

In this paper, we first show that the nonconvex QP (2) is NP-hard in either the real or the complex case, which further implies the NP-hardness of the general problem (1).

Then, we consider a semidefinite programming (SDP) relaxation of (1) and a convex QP restriction of (2) and study their worst-case performance. In particular, let υsdp, υcqp and υqp denote the optimal values of the SDP relaxation, the convex QP restriction, and the original QP (1), respectively. We establish a performance ratio of υqpsdp = O(m2) for the SDP relaxation in the real case, and we give an example showing that this bound is tight up to a constant factor. Similarly, we establish a performance ratio of υqpsdp= O(m) in the complex case, and we give an example showing the tightness of this bound. We further show that, in the case when the phase spread of the entries of h1, ..., hM is bounded away from π/2, the performance ratios υqpsdp and υcqpqp for the SDP relaxation and the convex QP restriction, respectively, are independent of m and n.

In recent years, there have been extensive studies of the performance of SDP relaxations for nonconvex QP. However, to our knowledge, this is the first performance analysis of SDP relaxation for QP with concave quadratic constraints. Our proof techniques also extend to a maximization version of the QP (1) with convex homogeneous quadratic constraints.

In particular, we give a simple proof of a result analogous to one of Nemirovski, Roos and Terlaky [14] (also see [13, Theorem 4.7]) for the real case, namely, the SDP relaxation for this nonconvex QP has a performance ratio of O(1/ ln(m)).

(4)

2 NP-hardness

In this section, we show that the nonconvex QP (1) is NP-hard in general. First, we notice that, by a linear transformation if necessary, the following problem

minimize zHQz

subject to |z`| ≥ 1, ` = 1, ..., n, z ∈ IFn,

(3)

is a special case of (1), where Q ∈ IFn×n is a Hermitian positive definite matrix (i.e., Q Â 0), and z` denotes the `th component of z. Hence, it suffices to establish the NP- hardness of (3). To this end, we consider a reduction from the NP-complete partition problem: Given positive integers a1, a2, ..., aN, decide whether there exists a subset I of {1, ..., N } satisfying

X

`∈I

a`= 1 2

XN

`=1

a`. (4)

Our reductions differ for the real and complex cases. As will be seen, the NP-hardness proof in the complex case1 is more intricate than in the real case.

2.1 The Real Case

We consider the real case of IF = IR. Let n := N and a := (a1, . . . , aN)T, Q := aaT + In  0, where Indenotes the n × n identity matrix.

We show that a subset I satisfying (4) exists if and only if the optimization problem (3) has a minimum value of n. Since

zTQz = |aTz|2+ Xn

`=1

|z`|2 ≥ n whenever |z`| ≥ 1 ∀ `, z ∈ IRn,

we see that (3) has a minimum value of n if and only if there exists a z ∈ IRn satisfying aTz = 0, |z`| = 1 ∀ `.

The above condition is equivalent to the existence of a subset I satisfying (4), with the correspondence I = {` | z` = 1}. This completes the proof.

1This NP-hardness proof was first presented in an appendix of [20] and is included here for completeness;

also see [26, Proposition 3.5] for a related proof.

(5)

2.2 The Complex Case

We consider the complex case of IF = IC. Let n := 2N + 1 and a := (a1, . . . , aN)T,

A :=

à IN IN −eN aT 0TN 12aTeN

! , Q := ATA + In  0,

where eN denotes the N -dimensional vector of ones, 0N denotes the N -dimensional vector of zeros, and In and IN are identity matrices of sizes n × n and N × N , respectively.

We show that a subset I satisfying (4) exists if and only if the optimization problem (3) has a minimum value of n. Since

zHQz = kAzk2+ Xn

`=1

|z`|2 ≥ n whenever |z`| ≥ 1 ∀ `, z ∈ ICn,

we see that (3) has a minimum value of n if and only if there exists a z ∈ ICn satisfying Az = 0, |z`| = 1 ∀ `.

Expanding Az = 0 gives the following set of linear equations:

0 = z`+ zN +`− zn, ` = 1, ..., N, (5) 0 =

XN

`=1

a`z`1 2

ÃN X

`=1

a`

!

zn. (6)

For ` = 1, ..., 2N , since |z`| = |zn| = 1 so that z`/zn = e` for some θ` ∈ [0, 2π), we can rewrite (5) as

cos θ`+ cos θN +` = 1,

sin θ`+ sin θN +` = 0, ` = 1, ..., N.

These equations imply that θ` ∈ {−π/3, π/3} for all ` 6= n. In fact, these equations further imply that cos θ`= cos θN +`= 1/2 for ` = 1, ..., N , so that

Re ÃN

X

`=1

a`z` zn 1

2 ÃN

X

`=1

a`

!!

= 0.

Therefore, (6) is satisfied if and only if Im

ÃN X

`=1

a`z` zn 1

2 ÃN

X

`=1

a`

!!

= Im ÃN

X

`=1

a`z` zn

!

= 0,

which is further equivalent to the existence of a subset I satisfying (4), with the corre- spondence I = {` | θ`= π/3}. This completes the proof.

(6)

3 Performance analysis of SDP relaxation

In this section, we study the performance of an SDP relaxation of (2). Let Hi := X

`∈Ii

h`hH` , i = 1, ..., m.

The well-known SDP relaxation of (1) [11, 19] is υsdp := min Tr(Z)

s.t. Tr(HiZ) ≥ 1, i = 1, ..., m, Z º 0, Z ∈ IFn×n is Hermitian.

(7)

An optimal solution of the SDP relaxation (7) can be computed efficiently using, say, interior-point methods; see [18] and references therein.

Clearly υsdp ≤ υqp. We are interested in upper bounds for the relaxation performance of the form

υqp≤ Cυsdp,

where C ≥ 1. Since we assume Hi 6= 0 for all i, it is easily checked that (7) has an optimal solution, which we denote by Z.

3.1 General steering vectors: the real case

We consider the real case of IF = IR. Upon obtaining an optimal solution Z of (7), we construct a feasible solution of (1) using the following randomization procedure:

1. Generate a random vector ξ ∈ IRn from the real-valued normal distri- bution N (0, Z).

2. Let z(ξ) = ξ/ min

1≤i≤m

q ξTHiξ.

We will use z(ξ) to analyze the performance of the SDP relaxation. Similar procedures have been used for related problems [1, 3, 4, 5, 14]. First, we need to develop two lemmas.

The first lemma estimates the left-tail of the distribution of a convex quadratic form of a Gaussian random vector.

(7)

Lemma 1 Let H ∈ IRn×n, Z ∈ IRn×n be two symmetric positive semidefinite matrices (i.e., H º 0, Z º 0). Suppose ξ ∈ IRn is a random vector generated from the real-valued normal distribution N (0, Z). Then, for any γ > 0,

Prob³ξTHξ < γE(ξTHξ)´≤ max

½

γ,2(¯r − 1)γ π − 2

¾

, (8)

where ¯r := min{rank (H), rank (Z)}.

Proof. Since the covariance matrix Z º 0 has rank r := rank (Z), we can write Z = U UT, for some U ∈ IRn×r satisfying UTZU = Ir. Let ¯ξ := QTUTξ ∈ IRr, where Q ∈ IRr×r is an orthogonal matrix corresponding to the eigen-decomposition of the matrix

UTHU = QΛQT,

for some diagonal matrix Λ = Diag{λ1, λ2, ..., λr}, with λ1 ≥ λ2 ≥ ... ≥ λr ≥ 0. Since UTHU has rank at most ¯r, we have λi = 0 for all i > ¯r. It is readily checked that ¯ξ has the normal distribution N (0, Ir). Moreover, ξ is statistically identical to U Q¯ξ, so that ξTHξ is statistically identical to

ξ¯TQTUTHU Q¯ξ = ¯ξTΛ¯ξ =

¯

Xr i=1

λi|¯ξi|2. Then, we have

Prob³ξTHξ < γE(ξTHξ)´ = Prob à ¯r

X

i=1

λi|¯ξi|2 < γE Ã ¯r

X

i=1

λi|¯ξi|2

!!

= Prob à ¯r

X

i=1

λi|¯ξi|2 < γ

¯

Xr i=1

λi

! .

If λ1 = 0, then this probability is zero, which proves (8). Thus, we will assume that λ1> 0. Let ¯λi := λi/(λ1+ · · · + λr¯), for i = 1, ..., ¯r. Clearly, we have

¯λ1+ · · · + ¯λ¯r= 1, λ¯1 ≥ ¯λ2 ≥ . . . ≥ ¯λ¯r≥ 0.

We consider two cases. First, suppose ¯λ1≥ α, where 0 < α < 1. Then, we can bound the above probability as follows:

Prob³ξTHξ < γE(ξTHξ)´ = Prob à r¯

X

i=1

λ¯i|¯ξi|2 < γ

!

≤ Prob³λ¯1|¯ξ1|2 < γ´

≤ Prob³|¯ξ1|2< γ/α´ (9)

r

πα,

(8)

where the last step is due to the fact that ¯ξ1 is a real-valued zero mean Gaussian random variable with unit variance.

In the second case, we have ¯λ1 < α, so that

λ¯2+ · · · + ¯λr¯= 1 − ¯λ1> 1 − α.

This further implies (¯r − 1)¯λ2 ≥ ¯λ2+ · · · + ¯λr¯> 1 − α. Hence λ¯1≥ ¯λ2 > 1 − α

¯ r − 1.

Using this bound, we obtain the following probability estimate:

Prob³ξTHξ < γE(ξTHξ)´ = Prob à r¯

X

i=1

¯λi|¯ξi|2< γ

!

≤ Prob³¯λ1|¯ξ1|2< γ, ¯λ2|¯ξ2|2< γ´

= Prob³¯λ1|¯ξ1|2< γ´· Prob³λ¯2|¯ξ2|2 < γ´ (10)

s π¯λ1 ·

s π¯λ2

2(¯r − 1)γ π(1 − α).

Combining the estimates for the above two cases and setting α = 2/π, we immediately obtain the desired bound (8).

Lemma 2 Let IF = IR. Let Zº 0 be a feasible solution of (7) and let z(ξ) be generated by the randomization procedure described earlier. Then, with probability 1, z(ξ) is well defined and feasible for (1). Moreover, for every γ > 0 and µ > 0,

Prob µ

1≤i≤mmin ξTHiξ ≥ γ, kξk2 ≤ µTr(Z)

≥ 1 − m · max

½

γ,2(r − 1)γ π − 2

¾

1

µ, (11) where r := rank (Z).

Proof. Since Z º 0 is feasible for (7), it follows that Tr(HiZ) ≥ 1 for all i = 1, ..., m.

Since E(ξTHiξ) = Tr(HiZ) ≥ 1 and the density of ξTHiξ is absolutely continuous, the probability of ξTHiξ = 0 is zero, implying that z(ξ) is well defined with probability 1.

The feasibility of z(ξ) is easily verified.

(9)

To prove (11), we first note that E(ξξT) = Z. Thus, for any γ > 0 and µ > 0, Prob

µ

1≤i≤mmin ξTHiξ ≥ γ, kξk2≤ µTr(Z)

= Prob³ξTHiξ ≥ γ ∀ i = 1, ..., m and kξk2 ≤ µTr(Z)´

≥ Prob³ξTHiξ ≥ γTr(HiZ) ∀ i = 1, ..., m and kξk2 ≤ µTr(Z)´

= Prob³ξTHiξ ≥ γE(ξTHiξ) ∀ i = 1, ..., m and kξk2 ≤ µE(kξk2)´

= 1 − Prob³ξTHiξ < γE(ξTHiξ) for some i or kξk2> µE(kξk2)´

≥ 1 − Xm i=1

Prob³ξTHiξ < γE(ξTHiξ)´− Prob³kξk2> µE(kξk2)´

> 1 − m · max

½

γ,2(r − 1)γ π − 2

¾

1 µ,

where the last step uses Lemma 1 as well as Markov’s inequality:

Prob³kξk2 > µE(kξk2)´ 1 µ. This completes the proof.

We now use Lemma 2 to bound the performance of the SDP relaxation.

Theorem 1 Let IF = IR. For the QP (1) and its SDP relaxation (7), we have υqp= υsdp if m ≤ 2, and otherwise

υqp 27m2 π υsdp.

Proof. By applying a suitable rank reduction procedure if necessary, we can assume that the rank r of the optimal SDP solution Z satisfies r(r + 1)/2 ≤ m; see e.g. [17]. Thus r <√

2m. If m ≤ 2, then r = 1, implying that Z = z(z)T for some z ∈ IRn and it is readily seen that z is an optimal solution of (1), so that υqp= υsdp. Otherwise, we apply the randomization procedure to Z. We also choose

µ = 3, γ = π 4m2

µ 1 − 1

µ

2

= π

9m2. Then, it is easily verified using r <√

2m that

√γ ≥ 2(r − 1)γ

π − 2 ∀ m = 1, 2, ...

(10)

Plugging these choices of γ and µ into (11), we see that there is a positive probability (independent of problem size) of at least

1 − m√ γ − 1

µ = 1 −

√π 3 1

3 = 0.0758...

that ξ generated by the randomization procedure satisfies

1≤i≤mmin ξTHiξ ≥ π

9m2 and kξk2≤ 3 Tr(Z).

Let ξ be any vector satisfying these two conditions.2 Then, z(ξ) is feasible for (1), so that

υqp≤ kz(ξ)k2 = kξk2

miniξTHiξ 3 Tr(Z)

(π/9m2) = 27m2 π υsdp, where the last equality uses Tr(Z) = υsdp.

In the above proof, other choices of µ can also be used, but the resulting bound seems not as sharp. Theorem 1 suggests that the worst-case performance of the SDP relaxation deteriorates quadratically with the number of quadratic constraints. Below we give an example demonstrating that this bound is in fact tight up to a constant factor.

Example 1: For any m ≥ 2 and n ≥ 2, consider a special instance of (2), corresponding to (1) with |Ii| = 1 (i.e., each Hi has rank 1), whereby

h`= µ

cos µ

m

, sin

µ m

, 0, . . . , 0

T

, ` = 1, ...., m.

Let z = (z1, . . . , zn)T ∈ IRn be an optimal solution of (2) corresponding to the above choice of steering vectors h`. We can write

(z1, z2) = ρ(cos θ, sin θ), for some θ ∈ [0, 2π).

Since {`π/m, ` = 1, ..., m} is uniformly spaced on [0, π), there must exist an integer ` such that

either

¯¯

¯¯θ −`π m −π

2

¯¯

¯¯ π 2m or

¯¯

¯¯θ −`π m + π

2

¯¯

¯¯ π 2m.

For simplicity, we assume the first case. (The second case can be treated similarly.) Since the last (n − 2) entries of h` are zero, it is readily checked that

|hT`z| = ρ

¯¯

¯¯cos µ

θ −`π m

¶¯¯

¯¯= ρ

¯¯

¯¯sin µ

θ −`π m −π

2

¶¯¯

¯¯≤ ρ

¯¯

¯¯sin µ π

2m

¶¯¯

¯¯ ρπ 2m.

2The probability that no such ξ is generated after N independent trials is at most (1−0.0758..)N, which for N = 100 equals 0.000375.. Thus, such ξ requires relatively few trials to generate.

(11)

Since z satisfies the constraint |hT`z| ≥ 1, it follows that kzk ≥ ρ ≥ 2m|hT`z|

π 2m

π , implying

υqp= kzk2 4m2 π2 . On the other hand, the positive semidefinite matrix

Z= Diag{1, 1, 0, . . . , 0}

is feasible for the SDP relaxation (7), and it has an objective value of Tr(Z) = 2. Thus, for this instance, we have

υqp 2m2 π2 υsdp.

The preceding example and Theorem 1 show that the SDP relaxation (7) can be weak if the number of quadratic constraints is large, especially when the steering vectors h` are in a certain sense “uniformly distributed” in space.

3.2 General steering vectors: the complex case

We consider the complex case of IF = IC. We will show that the performance ratio of the SDP relaxation (7) improves to O(m) in the complex case (as opposed to O(m2) in the real case). Similar to the real case, upon obtaining an optimal solution Z of (7), we construct a feasible solution of (1) using the following randomization procedure:

1. Generate a random vector ξ ∈ ICn from the complex-valued normal distribution Nc(0, Z) [2, 26].

2. Let z(ξ) = ξ/ min

1≤i≤m

q ξHHiξ.

Most of the ensuing performance analysis is similar to that of the real case. In partic- ular, we will also need the following two lemmas analogous to Lemmas 1 and 2.

Lemma 3 Let H ∈ ICn×n, Z ∈ ICn×n be two Hermitian positive semidefinite matrices (i.e., H º 0, Z º 0). Suppose ξ ∈ ICn is a random vector generated from the complex-valued normal distribution Nc(0, Z). Then, for any γ > 0,

Prob³ξHHξ < γE(ξHHξ)´≤ max

½4

3γ, 16(¯r − 1)2γ2

¾

, (12)

(12)

where ¯r := min{rank (H), rank (Z)}.

Proof. We follow the same notations and proof as for Lemma 1, except for two blanket changes:

matrix transpose → Hermitian transpose, orthogonal matrix → unitary matrix.

Also, ¯ξ has the complex-valued normal distribution Nc(0, Ir). With these changes, we consider the same two cases: ¯λ1 ≥ α and ¯λ1 < α, where 0 < α < 1. In the first case, we have similar to (9) that

Prob³ξHHξ < γE(ξHHξ)´≤ Prob³|¯ξ1|2 < γ/α´. (13) Recall that the density function of a complex-valued circular normal random variable u ∼ Nc(0, σ2), where σ is the standard deviation, is

1

πσ2e|u|2σ2 ∀ u ∈ IC.

In polar coordinates, the density function can be written as f (ρ, θ) = ρ

πσ2eρ2σ2 ∀ ρ ∈ [0, +∞), θ ∈ [0, 2π).

In fact, a complex-valued normal distribution can be viewed as a joint distribution of its modulus and its argument, with the following particular properties: (1) the modulus and argument are independently distributed; (2) the argument is uniformly distributed over [0, 2π); (3) the modulus follows a Weibull distribution with density

f (ρ) =

σ2eρ2σ2, if ρ ≥ 0;

0, if ρ < 0, and distribution function

Prob {|u| ≤ t} = 1 − eσ2t2. (14) Since ¯ξ1 ∼ Nc(0, 1), substituting this into (13) yields

Prob³ξHHξ < γE(ξHHξ)´≤ Prob³|¯ξ1|2< γ/α´≤ 1 − e−γ/α≤ γ/α, where the last inequality uses the convexity of the exponential function.

(13)

In the second case of ¯λ1< α, we have similar to (10) that

Prob³ξHHξ < γE(ξHHξ)´ ≤ Prob³¯λ1|¯ξ1|2< γ´· Prob³λ¯2|¯ξ2|2 < γ´

= (1 − e−γ/¯λ1)(1 − e−γ/¯λ2)

γ2 λ¯1¯λ2

r − 1)2γ2 (1 − α)2 ,

where last step uses the fact that ¯λ1 ≥ ¯λ2 ≥ (1 − α)/(¯r − 1). Combining the estimates for the above two cases and setting α = 3/4, we immediately obtain the desired bound (12).

Lemma 4 Let IF = IC. Let Z º 0 be a feasible solution of (7) and let z(ξ) be generated by the randomization procedure described earlier. Then, with probability 1, z(ξ) is well defined and feasible for (1). Moreover, for every γ > 0 and µ > 0,

Prob µ

1≤i≤mmin ξHHiξ ≥ γ, kξk2 ≤ µTr(Z)

≥ 1 − m · max

½4

3γ, 16(r − 1)2γ2

¾

1 µ, where r := rank (Z).

Proof. The proof is mostly the same as that for the real case (see Lemma 2). In particular, for any γ > 0 and µ > 0, we still have

Prob µ

1≤i≤mmin ξHHiξ ≥ γ, kξk2 ≤ µTr(Z)

≥ 1 − Xm i=1

Prob³ξHHiξ < γE(ξHHiξ)´− Prob³kξk2> µE(kξk2)´.

Therefore, we can invoke Lemma 3 to obtain Prob

µ

1≤i≤mmin ξHHiξ ≥ γ, kξk2≤ µTr(Z)

≥ 1 − m · max

½4

3γ, 16(r − 1)2γ2

¾

− Prob³kξk2 > µE(kξk2)´

≥ 1 − m · max

½4

3γ, 16(r − 1)2γ2

¾

1 µ, which completes the proof.

(14)

Theorem 2 Let IF = IC. For the QP (1) and its SDP relaxation (7), we have vsdp = vqp if m ≤ 3 and otherwise

vqp≤ 8m · vsdp.

Proof. By applying a suitable rank reduction procedure if necessary, we can assume that the rank r of the optimal SDP solution Z satisfies r = 1 if m ≤ 3 and r ≤√

m if m ≥ 4;

see [9, Section 5]. Thus, if m ≤ 3, then Z = z(z)H for some z ∈ ICn and it is readily seen that z is an optimal solution of (1), so that vsdp = vqp. Otherwise, we apply the randomization procedure to Z. By choosing µ = 2 and γ = 4m1 , it is easily verified using r ≤√

m that

4

3γ ≥ 16(r − 1)2γ2 ∀ m = 1, 2, ...

Therefore, it follows from Lemma 4 that Prob

½

1≤i≤mmin ξHHiξ ≥ γ, kξk2 ≤ µTr(Z)

¾

≥ 1 − m4 3γ − 1

µ = 1 6.

Then, similar to the proof of Theorem 1, we obtain that with probability of at least 1/6, z(ξ) is a feasible solution of (1) and vqp≤ kz(ξ)k2≤ 8m · vsdp.3

The proof of Theorem 2 shows that, by repeating the randomization procedure, the probability of generating a feasible solution with a performance ratio no more than 8m ap- proaches 1 exponentially fast (independent of problem size). Alternatively, a de-randomization technique from theoretical computer science can perhaps convert the above randomization procedure into a polynomial-time deterministic algorithm [12]; also see [14].

Theorem 2 shows that the worst-case performance of SDP relaxation deteriorates lin- early with the number of quadratic constraints. This contrasts with the quadratic rate of deterioration in the real case (see Theorem 1). Thus, the SDP relaxation can yield better performance in the complex case. This is in the same spirit as the recent results in [26]

which showed that the quality of SDP relaxation improves by a constant factor for certain quadratic maximization problems when the space is changed from IRn to ICn. Below we give an example demonstrating that this approximation bound is tight up to a constant factor.

Example 2: For any m ≥ 2 and n ≥ 2, let K = d√

me (so K ≥ 2). Consider a special instance of (2), corresponding to (1) with |Ii| = 1 (i.e., each Hi has rank 1), whereby

h` = µ

cos

K, sinjπ

Kei2kπK , 0, . . . , 0

T

with ` = jK − K + k, j, k = 1, ..., K.

3The probability that no such ξ is generated after N independent trials is at most (5/6)N, which for N = 30 equals 0.00421.. Thus, such ξ requires relatively few trials to generate.

(15)

Hence there are K2 complex rank-1 constraints. Let z= (z1, . . . , zn)T ∈ ICnbe an optimal solution of (2) corresponding to the above choice of d√

me2 steering vectors h`. By a phase rotation if necessary, we can without loss of generality assume that z1 is real and write

(z1, z2) = ρ(cos θ, sin θe), for some θ, ψ ∈ [0, 2π).

Since {2kπ/K, k = 1, ..., K} and {jπ/K, j = 1, ..., K} are uniformly spaced in [0, 2π) and [0, π) respectively, there must exist integers j and k such that

¯¯

¯¯ψ −2kπ K

¯¯

¯¯ π

K and either

¯¯

¯¯θ − K −π

2

¯¯

¯¯ π 2K or

¯¯

¯¯θ −jπ K +π

2

¯¯

¯¯ π 2K. Without loss of generality, we assume

¯¯

¯¯θ −jπ K −π

2

¯¯

¯¯ π 2K.

Since the last (n − 2) entries of each h` are zero, it is readily seen that, for ` = jK − K + k,

¯¯

¯Re(hH` z)¯¯¯ = ρ

¯¯

¯¯cos θ cosjπ

K + sin θ sinjπ K cos

µ

ψ −2kπ K

¶¯¯

¯¯

= ρ

¯¯

¯¯cos µ

θ − K

+ sin θ sinjπ K

µ cos

µ

ψ −2kπ K

− 1

¶¯¯

¯¯

= ρ

¯¯

¯¯sin µ

θ − K −π

2

− 2 sin θ sinjπ K sin2

µKψ − 2kπ 2K

¶¯¯

¯¯

≤ ρ

¯¯

¯¯sin π 2K

¯¯

¯¯+ 2ρ sin2 π 2K

ρπ

2K + ρπ2 2K2. In addition, we have

¯¯

¯Im(hH` z)¯¯¯ = ρ

¯¯

¯¯sin θ sinjπ K sin

µ

ψ −2kπ K

¶¯¯

¯¯

≤ ρ

¯¯

¯¯sin µ

ψ −2kπ K

¶¯¯

¯¯

≤ ρ

¯¯

¯¯ψ −2kπ K

¯¯

¯¯ ρπ K. Combining the above two bounds, we obtain

¯¯

¯hH` z¯¯¯¯¯¯Re(hH` z)¯¯¯+¯¯¯Im(hH` z)¯¯¯ 3ρπ

2K + ρπ2 2K2. Since z satisfies the constraint |hH` z| ≥ 1, it follows that

kzk ≥ ρ ≥ 2K2|hH` z|

π(3K + π) 2K2 π(3K + π),

(16)

implying

υqp= kzk2 4K4

π2(3K + π)2 = 4d√ me4 π2(3d√

me + π)2. On the other hand, the positive semidefinite matrix

Z= Diag{1, 1, 0, . . . , 0}

is feasible for the SDP relaxation (7), and it has an objective value of Tr(Z) = 2. Thus, for this instance, we have

υqp 2d√ me4 π2(3d√

me + π)2 υsdp 2m

π2(3 + π/2)2 υsdp.

The preceding example and Theorem 2 show that the SDP relaxation (7) can be weak if the number of quadratic constraints is large, especially when the steering vectors h` are in a certain sense “uniformly distributed” in space. In the next subsection, we will tighten the approximation bound in Theorem 2 by considering special cases where the steering vectors are “not too spread out in space”.

3.3 Specially configured steering vectors: the complex case

We consider the complex case of IF = IC. Let Z be any optimal solution of (7). Since Z is feasible for (7), Z6= 0. Then

Z= Xr k=1

wkwkH, (15)

for some nonzero wk ∈ ICn, where r := rank (Z) ≥ 1. By decomposing wk = uk+ vk, with uk ∈ span{h1, ..., hM} and vk ∈ span{h1, ..., hM}, it is easily checked that ˜Z :=

Pr

k=1ukuHk is feasible for (7) and hI, Zi =

Xr k=1

kuk+ vkk2 = Xr k=1

(kukk2+ kvkk2) = hI, ˜Zi + Xr k=1

kvkk2.

This implies vk= 0 for all k, so that

wk∈ span{h1, ..., hM}. (16)

Below we show that the SDP relaxation (7) provides a constant factor approximation to the QP (1) when the phase spread of the entries of h` is bounded away from π/2.

(17)

Theorem 3 Suppose that

h` = Xp

i=1

βi`gi ∀ ` = 1, ..., M, (17)

for some p ≥ 1, βi`∈ IC and gi ∈ ICn such that kgik = 1 and gHi gj = 0 for all i 6= j. Then the following results hold.

(a) If Re(βi`Hβj`) > 0 whenever βi`Hβj`6= 0, then υqp ≤ Cυsdp, where

C := max

i,j,` | βi`Hβj`6=0

Ã

1 +|Im(βi`Hβj`)|2

|Re(βi`Hβj`)|2

!1/2

. (18)

(b) If βi`= |βi`|ei`, where

φi`∈ [ ¯φ`− φ, ¯φ`+ φ] ∀ i, `, for some 0 ≤ φ < π

4 and some ¯φ` ∈ IR, (19) then Re(βi`Hβj`) > 0 whenever βi`Hβj` 6= 0, and C given by (18) satisfies

C ≤ 1

cos(2φ). (20)

Proof. (a) By (16), we have

wk= Xp

i=1

αkigi, for some αki∈ IC. This together with (15) yields

hI, Zi = Xr k=1

kwkk2 = Xr k=1

°°

°°

° Xp

i=1

αkigi

°°

°°

°

2

= Xr k=1

Xp i=1

ki|2 = Xp i=1

λ2i,

where the third equality uses the orthonormal properties of g1, ..., gp, and the last equality uses λi :=¡Prk=1ki|2¢1/2 = k(αki)rk=1k.

Let

z :=

Xp

i=1

λigi.

參考文獻

相關文件

Numerical results are reported for some convex second-order cone programs (SOCPs) by solving the unconstrained minimization reformulation of the KKT optimality conditions,

Numerical experiments are done for a class of quasi-convex optimization problems where the function f (x) is a composition of a quadratic convex function from IR n to IR and

Particularly, combining the numerical results of the two papers, we may obtain such a conclusion that the merit function method based on ϕ p has a better a global convergence and

Then, it is easy to see that there are 9 problems for which the iterative numbers of the algorithm using ψ α,θ,p in the case of θ = 1 and p = 3 are less than the one of the

By exploiting the Cartesian P -properties for a nonlinear transformation, we show that the class of regularized merit functions provides a global error bound for the solution of

volume suppressed mass: (TeV) 2 /M P ∼ 10 −4 eV → mm range can be experimentally tested for any number of extra dimensions - Light U(1) gauge bosons: no derivative couplings. =&gt;

Define instead the imaginary.. potential, magnetic field, lattice…) Dirac-BdG Hamiltonian:. with small, and matrix

incapable to extract any quantities from QCD, nor to tackle the most interesting physics, namely, the spontaneously chiral symmetry breaking and the color confinement.. 