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

Implicit Redundancy and Degeneracy
in Conic Program

Haesol Im
Abstract

This paper examines the feasible region of a standard conic program represented as the intersection of a closed convex cone and a set of linear equalities. It is recently shown that when Slater constraint qualification (strict feasibility) fails for the classes of linear and semidefinite programs, two key properties emerge within the feasible region; (a) every point in the feasible region is degenerate; (b) the constraint system inherits implicit redundancies. In this paper we show that degeneracy and implicit redundancies are inherent and universal traits of all conic programs in the absence of strict feasibility.

Keywords: Degeneracy, Facial reduction, Implicit redundancy

1 Introduction

We consider the feasible region of a standard conic program represented as the intersection of a closed convex cone 𝒦{\mathcal{K}} and an affine subspace {\mathcal{L}}. Throughout the paper we consider 𝒦{\mathcal{K}} and {\mathcal{L}} in a finite dimensional Euclidean space 𝔼\mathbb{E}. We represent the affine subspace by

={x𝔼:𝒜x=bm},{\mathcal{L}}=\{x\in\mathbb{E}:{\mathcal{A}}x=b\in\mathbb{R}^{m}\},

where 𝒜:𝔼m{\mathcal{A}}:\mathbb{E}\to\mathbb{R}^{m} is a linear map. We define the feasible region of a conic program

:=𝒦={x𝒦:𝒜x=b}.{\mathcal{F}}:={\mathcal{K}}\cap{\mathcal{L}}=\{x\in{\mathcal{K}}:{\mathcal{A}}x=b\}\neq\emptyset.

We assume that 𝒜{\mathcal{A}} is surjective. Since {\mathcal{F}} is not empty, 𝒜{\mathcal{A}} being surjective means that 𝒜x=b{\mathcal{A}}x=b itself does not contain any redundant equalities. If 𝒜{\mathcal{A}} is not given surjective, numerical methods [24, 1, 8] can be used for sorting out redundant equalities from the system 𝒜x=b{\mathcal{A}}x=b.

{\mathcal{F}} is said to hold strict feasibility (Slater constraint qualification) if {\mathcal{F}} contains a point the relative interior of 𝒦{\mathcal{K}}. This paper particularly focuses on {\mathcal{F}} that fails strict feasibility, i.e., {\mathcal{F}} that lies on the relative boundary of 𝒦{\mathcal{K}}. It is well-known that conic programs that fail this regularity condition does not guarantee strong duality, i.e., the primal and dual optimal values may not agree or the dual optimal value is not attained by a dual feasible point; see [31, Section 2]. For this reason, many optimization algorithms are developed under this regularity condition. The class of linear programs (𝒦=+n{\mathcal{K}}=\mathbb{R}_{+}^{n}) avoids this pathology and consequences of absence of strict feasibility for linear programs are rarely discussed in the literature. A recent work [18] highlights the importance of strict feasibility for linear programs and reveals the connection between degeneracy of extreme points and strict feasibility.

The cone of nn-by-nn positive semidefinite matrices, 𝕊+n\mathbb{S}_{+}^{n}, is a well-studied object. The computationally efficient facial structure of 𝕊+n\mathbb{S}_{+}^{n} is clearly identified and there are successful numerical solvers that take advantage of this knowledge [2, 30, 27]. Let 𝕊+n:={x𝕊+n:𝒜x=b}{\mathcal{F}}_{\mathbb{S}_{+}^{n}}:=\{x\in\mathbb{S}_{+}^{n}:{\mathcal{A}}x=b\}. When 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} fails strict feasibility, two interesting properties are shown recently:

  1. 1.

    Every point in 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} is degenerate [16].

  2. 2.

    When the domain of 𝒜{\mathcal{A}} is restricted to be the minimal face of 𝕊+n\mathbb{S}_{+}^{n} containing 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}}, 𝒜{\mathcal{A}} is not surjective. In other words, 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} contains implicitly redundant linear constraints [15].

We explain the meanings of degeneracy and implicit redundancy in Sections 2.3 and 3 respectively. In this paper we show that these two properties are universal properties of all conic programs that fail strict feasibility, not just restricted to linear and semidefinite programs.

1.1 Notation

Given a linear map 𝒯{\mathcal{T}}, we let 𝒯{\mathcal{T}}^{*} be the adjoint of 𝒯{\mathcal{T}}. We use ,𝔼\langle\cdot,\cdot\rangle_{\mathbb{E}} to denote the inner product between two vectors in the Euclidean space 𝔼\mathbb{E}; we omit the subscript 𝔼\mathbb{E} when the meaning is clear. We use (𝒯){\mathcal{R}}({\mathcal{T}}) to denote the range of the map 𝒯{\mathcal{T}}. Given a matrix XX, range(X)\operatorname{range}(X) denotes the range of XX, formed by the linear combinations of columns of XX. null(X)\operatorname{null}(X) is used to denote the null-space of XX.

Given a set 𝒳𝔼{\mathcal{X}}\subset\mathbb{E}, int(𝒳)\operatorname{{int}}({\mathcal{X}}) (relint(𝒳)\operatorname{{relint}}({\mathcal{X}}), resp.) means the interior (relative interior, resp.) of 𝒳{\mathcal{X}}. The dimension and the span of 𝒳{\mathcal{X}} are denoted dim(𝒳)\dim({\mathcal{X}}) and span(𝒳)\operatorname{{span}}({\mathcal{X}}), respectively. We use 𝒳{\mathcal{X}}^{\perp} to indicate the orthogonal complement of 𝒳{\mathcal{X}}. Given a vector xx, we often use xx^{\perp} to mean {x}\{x\}^{\perp}. n\mathbb{R}^{n} and 𝕊n\mathbb{S}^{n} denote the usual vector spaces of vector of nn coordinates and nn-by-nn symmetric matrices, respectively. Given a matrix AA of size mm-by-nn and an index set {\mathcal{I}}, we use A(:,)A(:,{\mathcal{I}}) for the submatrix of AA formed by the columns of AA that correspond to {\mathcal{I}}.

+n\mathbb{R}_{+}^{n} is used for the nonnegative orthant {xn:xi0,i}\{x\in\mathbb{R}^{n}:x_{i}\geq 0,\forall i\} and 𝕊+n\mathbb{S}_{+}^{n} is used for the set of nn-by-nn positive semidefinite matrices, {X𝕊n:y,Xyn0,yn}\{X\in\mathbb{S}^{n}:\langle y,Xy\rangle_{\mathbb{R}^{n}}\geq 0,\forall y\in\mathbb{R}^{n}\}. When a general conic program is discussed, we omit the subscript and use the symbol {\mathcal{F}}. We use the notation 𝒦{\mathcal{F}}_{\mathcal{K}} to distinguish a particular setting for the cone 𝒦{\mathcal{K}} from a general conic program. For example, when the feasible region of the semidefinite program is discussed, we use 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}}.

1.2 Related Work

The concept of implicit redundancies is introduced recently [18, 15]. Using this notion, [18] shows that every extreme point (basic feasible solution) of +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} is degenerate in the absence of strict feasibility. [15] shows that 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} also contains implicit redundant constraints in the absence of strict feasibility. Using the notion of implicit redundancies, [15] tightens the Barvinok-Pataki bound [3, 21], and further improves the bound presented in [17]. On the degeneracy side, a recent work [16] links the definition of degeneracy introduced by Pataki [34, Definition 3.3.1] to realize that every point in 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} is degenerate. This is done by exploiting the facial structure of 𝕊+n\mathbb{S}_{+}^{n}. Since a linear program is a special case of a semidefinite program, it immediately implies that every point in +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} is degenerate, not just limited to the extreme points.

2 Background

This section reviews basic notions and known results in the literature related to the main results. We present basic definitions including cones and faces, the goal of facial reduction process, and the notion of degeneracy.

2.1 Faces of Cones

A closed convex set 𝒦𝔼{\mathcal{K}}\subseteq\mathbb{E} is said to be a cone if x𝒦x\in{\mathcal{K}} implies αx𝒦\alpha x\in{\mathcal{K}}, α0\forall\alpha\geq 0. The dual cone of 𝒦{\mathcal{K}} is 𝒦:={z𝔼:z,x0,x𝒦}{\mathcal{K}}^{*}:=\{z\in\mathbb{E}:\langle z,x\rangle\geq 0,\forall x\in{\mathcal{K}}\}. A convex set ff is called a face of 𝒳𝔼{\mathcal{X}}\subseteq\mathbb{E} (denoted f𝒳f\unlhd{\mathcal{X}}) if

 for x,y𝒳 with {λx+(1λ)y:λ(0,1)}fx,yf.\text{ for }x,y\in{\mathcal{X}}\text{ with }\{\lambda x+(1-\lambda)y:\lambda\in(0,1)\}\subset f\implies x,y\in f.

A face f𝒦f\unlhd{\mathcal{K}} is said to be exposed if f=𝒦{z}f={\mathcal{K}}\cap\{z\}^{\perp} for some z𝒦z\in{\mathcal{K}}^{*}. 𝒦{\mathcal{K}} is said to be facially exposed if every face of 𝒦{\mathcal{K}} is exposed. Given a convex set 𝒳𝒦{\mathcal{X}}\subset{\mathcal{K}}, the minimal face of 𝒦{\mathcal{K}} containing 𝒳{\mathcal{X}}, denoted face(𝒳,𝒦)\operatorname{face}({\mathcal{X}},{\mathcal{K}}), is the intersection of faces of 𝒦{\mathcal{K}} that contains ff, i.e.,

face(𝒳,𝒦):={f𝒦:𝒳f}.\operatorname{face}({\mathcal{X}},{\mathcal{K}}):=\cap\{f\unlhd{\mathcal{K}}:{\mathcal{X}}\subseteq f\}.

It is known that a point x^relint(𝒳)\hat{x}\in\operatorname{{relint}}({\mathcal{X}}) provides the characterization face(𝒳,𝒦)=face(x^,𝒦)\operatorname{face}({\mathcal{X}},{\mathcal{K}})=\operatorname{face}(\hat{x},{\mathcal{K}}); see [9, Proposition 2.2.5]. Given a face f𝒦f\unlhd{\mathcal{K}}, the conjugate face of ff, denoted fΔf^{\Delta}, is

fΔ:={z𝒦:z,x=0,xf}.f^{\Delta}:=\{z\in{\mathcal{K}}^{*}:\langle z,x\rangle=0,\ \forall x\in f\}.

Understanding the facial structure of a cone is important for designing a computationally efficient algorithm. Examples of cones that demonstrate great computational powers include the nonnegative orthant +n\mathbb{R}_{+}^{n}, the positive semidefinite cone 𝕊+n\mathbb{S}_{+}^{n}, and the second-order (Lorentz) cone. Moreover, these cones lead to many successful interior point methods for solving the linear, semidefinite, second-order cone programs and the programs with Cartesian products of these cones [2, 30, 27].

We list facial characterizations of two cones. Every face f+nf\unlhd\mathbb{R}_{+}^{n} has the form

f=+n{xn:xi=0,i}={Vv:v0}=V+|c|,f=\mathbb{R}_{+}^{n}\cap\{x\in\mathbb{R}^{n}:x_{i}=0,i\in{\mathcal{I}}\}=\{Vv:v\geq 0\}=V\mathbb{R}_{+}^{|{\mathcal{I}}^{c}|}, (2.1)

where V=In(:,c)V=I_{n}(:,{\mathcal{I}}^{c}) for some index set {1,,n}{\mathcal{I}}\subseteq\{1,\ldots,n\} and InI_{n} is the nn-by-nn identity matrix. Here, c={1,,n}{\mathcal{I}}^{c}=\{1,\ldots,n\}\setminus{\mathcal{I}}, the complement of {\mathcal{I}}. 𝕊+n\mathbb{S}_{+}^{n} is another example of a well-understood cone. Every face f𝕊+nf\unlhd\mathbb{S}_{+}^{n} has the form

f=𝕊+n(QQT)=V𝕊+rank(X^)VT,f=\mathbb{S}_{+}^{n}\cap(QQ^{T})^{\perp}=V\mathbb{S}_{+}^{\operatorname{{rank}}(\hat{X})}V^{T}, (2.2)

where range(Q)=null(X^)\operatorname{range}(Q)=\operatorname{null}(\hat{X}) for some X^relint(f)\hat{X}\in\operatorname{{relint}}(f) and Vn×rank(X^)V\in\mathbb{R}^{n\times\operatorname{{rank}}(\hat{X})} satisfies range(V)=range(X^)\operatorname{range}(V)=\operatorname{range}(\hat{X}). Both +n\mathbb{R}_{+}^{n} and 𝕊+n\mathbb{S}_{+}^{n} are facially exposed. Other examples of facially exposed cones are the second-order cone, the doubly nonnegative cone, the hyperbolocity cone [23, Theorem 23] and the cone of Euclidean distance matrices [29]. On the other hand, not every cone is facially exposed. For example, the exponential cone [19], the completely positive cone [35], and the copositive cone [10] are known to be not facially exposed.

2.2 Facial Reduction for Conic Programs

Facial reduction (FR) is first introduced by Borwein and Wolkowicz in the 19801980’s [5, 6]. In-depth analysis for conic programs with linear objective functions, especially for semidefinite programs, has been conducted for the last two decades. Successful applications of FR are the semidefinite relaxations of NP-hard binary combinatorial optimization problems (e.g., see [36, 7]); and an analytical tool for measuring hardness of solving a problem numerically (e.g., see [28, 26]).

Facial reduction seeks to form an equivalent problem that satisfies strict feasibility. FR strives to find the minimal face of 𝒦{\mathcal{K}} that contains {\mathcal{F}} in order to represent the set

={x𝒦:𝒜x=b}={xface(,𝒦):𝒜x=b}.{\mathcal{F}}=\{x\in{\mathcal{K}}:{\mathcal{A}}x=b\}=\{x\in\operatorname{face}({\mathcal{F}},{\mathcal{K}}):{\mathcal{A}}x=b\}.

Finding face(,𝒦)\operatorname{face}({\mathcal{F}},{\mathcal{K}}) is often a challenging task and hence we often rely on an auxiliary lemma. Lemma 2.1 below lies in the heart of the FR algorithm.

Lemma 2.1.

[11, Theorem 3.1.3][Theorem of the Alternative] For a feasible system {\mathcal{F}}, exactly one of the following holds.

  1. 1.

    Strict feasibility holds for {\mathcal{F}};

  2. 2.

    There exists ymy\in\mathbb{R}^{m} such that

    𝒜y𝒦{0}, and b,y=0.{\mathcal{A}}^{*}y\in{\mathcal{K}}^{*}\setminus\{0\},\text{ and }\langle b,y\rangle=0. (2.3)

We note that a solution yy to (2.3) yields

0=b,ym=𝒜x,ym=x,𝒜y𝔼,x.0=\langle b,y\rangle_{\mathbb{R}^{m}}=\langle{\mathcal{A}}x,y\rangle_{\mathbb{R}^{m}}=\langle x,{\mathcal{A}}^{*}y\rangle_{\mathbb{E}},\ \forall x\in{\mathcal{F}}.

In other words, every feasible point xx\in{\mathcal{F}} lies in the orthogonal complement of 𝒜y{\mathcal{A}}^{*}y, i.e.,

𝒦(𝒜y).{\mathcal{F}}\subseteq{\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp}.

Consequently, we can rewrite the set {\mathcal{F}} by

={x𝒦(𝒜y):𝒜x=b}.{\mathcal{F}}=\{x\in{\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp}:{\mathcal{A}}x=b\}. (2.4)

The ambiant space now reduces from 𝔼=span(𝒦)\mathbb{E}=\operatorname{{span}}({\mathcal{K}}) to span(𝒦(𝒜y))\operatorname{{span}}({\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp}); the dimension reduction is an attractive by-product of the FR process. This completes one iteration of the FR algorithm. Then the FR algorithm replaces the cone as in (2.4) (i.e., 𝒦𝒦(𝒜y){\mathcal{K}}\leftarrow{\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp}) and solves the system 2.3 repeatedly until a strictly feasible point with respect to the reduced cone is guaranteed. Algorithm 2.1 summarizes the FR process.

  Given ={x𝒦:𝒜x=b}{\mathcal{F}}=\{x\in{\mathcal{K}}:{\mathcal{A}}x=b\}, 0={\mathcal{F}}^{0}={\mathcal{F}}, 𝒦0=𝒦{\mathcal{K}}^{0}={\mathcal{K}}, k=0k=0
  while strict feasibility fails for k{\mathcal{F}}^{k} do
     kk+1k\leftarrow k+1
     Find a vector yky^{k} satisfying 𝒜yk𝒦k1{0}{\mathcal{A}}^{*}y^{k}\in{\mathcal{K}}^{k-1}\setminus\{0\} and b,yk=0\langle b,y^{k}\rangle=0
     𝒦k:=𝒦k1(𝒜yk){\mathcal{K}}^{k}:={\mathcal{K}}^{k-1}\cap({\mathcal{A}}^{*}y^{k})^{\perp}
     k:={x𝒦k:𝒜x=b}{\mathcal{F}}^{k}:=\{x\in{\mathcal{K}}^{k}:{\mathcal{A}}x=b\}
  end while
Algorithm 2.1 FR Process

Upon the termination of the FR process Algorithm 2.1, we obtain

𝒦(𝒜y1)(𝒜yk).{\mathcal{F}}\subseteq{\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp}\cap\cdots\cap({\mathcal{A}}^{*}y^{k})^{\perp}.

Thus FR process allows an alternative representation of {\mathcal{F}} that satisfies strict feasibility:

={x𝒦(𝒜y1)(𝒜yk):𝒜x=b}.{\mathcal{F}}=\left\{x\in{\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp}\cap\cdots\cap({\mathcal{A}}^{*}y^{k})^{\perp}:{\mathcal{A}}x=b\right\}.

It is important to note that the FR process does not alter the points in the set {\mathcal{F}}, i.e.,

=1=2=k,{\mathcal{F}}={\mathcal{F}}^{1}={\mathcal{F}}^{2}=\cdots{\mathcal{F}}^{k}, (2.5)

where each i{\mathcal{F}}^{i} is defined in Algorithm 2.1. We emphasize that the representation of i{\mathcal{F}}^{i} is different from j{\mathcal{F}}^{j}, when iji\neq j.

The shortest length of FR process is called the singularity degree of {\mathcal{F}}, sd()\operatorname{sd}({\mathcal{F}}), first proposed by Sturm [28] for the class of semidefinite programs. This implies that Algorithm 2.1 requires at least sd()\operatorname{sd}({\mathcal{F}}) number of iterations. Indicated by (2.5), we have =sd(){\mathcal{F}}={\mathcal{F}}^{\operatorname{sd}({\mathcal{F}})} and we use the superscript to distinguish the different representations of the set {\mathcal{F}}. Hence strict feasibility fails for {\mathcal{F}}, sd()sd(sd())\operatorname{sd}({\mathcal{F}})\neq\operatorname{sd}({\mathcal{F}}^{\operatorname{sd}({\mathcal{F}})}) and sd(sd())=0\operatorname{sd}({\mathcal{F}}^{\operatorname{sd}({\mathcal{F}})})=0. While it is possible to construct a general conic program that has an arbitrarily large singularity degree, linear programs have sd(+n)1\operatorname{sd}({\mathcal{F}}_{\mathbb{R}_{+}^{n}})\leq 1 regardless of the problem dimension. sd()\operatorname{sd}({\mathcal{F}}) is often used as a measure of the hardness of solving a problem numerically by engaging error bounds; see [26, 28].

Example 2.2 below is an illustration of the FR process. We leave more details on the FR algorithm to [9, 11] for general conic programs, [25, 18] for 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} and +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} and related examples.

Example 2.2 (FR on the exponential cone).

Let

Kexp:={(x,y,z):y>0,zyex/y}{(x,y,z):x0,y=0,z0}K_{\text{exp}}:=\{(x,y,z):y>0,z\geq ye^{x/y}\}\cup\{(x,y,z):x\leq 0,y=0,z\geq 0\}

be the exponential cone. Visual illustrations of KexpK_{\text{exp}} can be found in [19, 2]. The dual cone of KexpK_{\text{exp}} is

Kexp={(x,y,z):x<0,ezxey/x}{(x,y,z):x=0,y0,z0}.K_{\text{exp}}^{*}=\{(x,y,z):x<0,ez\geq-xe^{y/x}\}\cup\{(x,y,z):x=0,y\geq 0,z\geq 0\}.

With A=[010110]A=\begin{bmatrix}0&1&0\\ 1&-1&0\end{bmatrix} and b=(00)b=\begin{pmatrix}0\\ 0\end{pmatrix}, consider the following feasible set

Kexp:={(x,y,z)Kexp:A(x,y,z)T=b}={(0,0,z):z0}.{\mathcal{F}}_{K_{\text{exp}}}:=\left\{(x,y,z)\in K_{\text{exp}}:A(x,y,z)^{T}=b\right\}=\{(0,0,z):z\geq 0\}.

It is known that {(0,0,z):z0}\{(0,0,z):z\geq 0\} is a non-exposed extreme ray of KexpK_{\text{exp}}; see [19].

Let λ1=(1,0)T\lambda^{1}=(1,0)^{T}. Then we have Aλ1=(0,1,0)TKexpA^{*}\lambda^{1}=(0,1,0)^{T}\in K_{\text{exp}}^{*} and Lemma 2.1 yields that

KexpKexp(Aλ1)={(x,y,z):x0,y=0,z0}.{\mathcal{F}}_{K_{\text{exp}}}\subseteq K_{\text{exp}}\cap(A^{*}\lambda^{1})^{\perp}=\{(x,y,z):x\leq 0,y=0,z\geq 0\}.

Now λ2=(1,1)T\lambda^{2}=(-1,-1)^{T} yields Aλ2=(1,0,0)TA^{*}\lambda^{2}=(-1,0,0)^{T}. Consequently we obtain

KexpKexp(Aλ1)(Aλ2)={(0,0,z):z0}.{\mathcal{F}}_{K_{\text{exp}}}\subseteq K_{\text{exp}}\cap(A^{*}\lambda^{1})^{\perp}\cap(A^{*}\lambda^{2})^{\perp}=\{(0,0,z):z\geq 0\}.

Note that λ1=(λ11,λ21)T\lambda^{1}=(\lambda^{1}_{1},\lambda^{1}_{2})^{T} with a nonzero second element cannot satisfy (2.3). If λ210\lambda^{1}_{2}\neq 0, Aλ1=(λ21,λ11λ21,0)TA^{*}\lambda^{1}=(\lambda^{1}_{2},\lambda^{1}_{1}-\lambda^{1}_{2},0)^{T} must belong to {(x,y,z):x<0,ezxey/x}\{(x,y,z):x<0,ez\geq-xe^{y/x}\}. However, this is impossible since the third element of Aλ1A^{*}\lambda^{1} is equal to 0 and λ21<0\lambda^{1}_{2}<0. Thus λ1=(1,0)T\lambda^{1}=(1,0)^{T} is the only vector that satisfies (2.3) up to a scalar multiple. We conclude that the singularity degree of this instance is 22.

It is well-known that FR applied to +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} and 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} generates +k{\mathcal{F}}_{\mathbb{R}_{+}^{k}} and 𝕊+k{\mathcal{F}}_{{\mathbb{S}^{k}_{+}}\,}, where k<nk<n. In plain language, FR for linear and semidefinite programs produces another semidefinite and linear programs in smaller dimensional spaces. This can be seen from the facial characterizations (2.1) and (2.2). It is because +n\mathbb{R}_{+}^{n} and 𝕊+n\mathbb{S}_{+}^{n} are self-reproducing cones. The FR process illustrated in Example 2.2 shows that the self-reproducibility does not apply a general conic program. Note that KexpK_{\text{exp}} is not polyhedral but the first FR iteration in Example 2.2 produces a polyhedral cone Kexp(Aλ1)K_{\text{exp}}\cap(A^{*}\lambda^{1})^{\perp}.

The feasible system {\mathcal{F}} is known to hold strict feasibility generically [22] and FR may seem unnecessary. However, many practical real-life examples seem to fail strict feasibility, e.g., see [18, 14]. As addressed in [18], a significant number of instances from the well-known NETLIB collection fails strict feasibility. We note that the significance of FR on +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} has only been a recent topic of discussion.

It is important to note that FR enables the restriction of the domain within which we can operate with 𝒜{\mathcal{A}}. Without FR, we cannot restrict the domain of 𝒜{\mathcal{A}} without any further analysis. We utilize this observation to discuss the main results in Section 3 below.

2.3 Degeneracy

In this section we introduce a known notion of degeneracy and discuss related results.

Definition 2.3 (Pataki).

[34, Definition 3.3.1] Let x¯\bar{x}\in{\mathcal{F}}. We say that x¯\bar{x} is nondegenerate if

linface(x¯,𝒦)Δ(𝒜)={0}.\operatorname{lin}\operatorname{face}(\bar{x},{\mathcal{K}})^{\Delta}\cap{\mathcal{R}}({\mathcal{A}}^{*})=\{0\}.

Characterizations of Definition 2.3 are available for some cones. Given x¯+n\bar{x}\in{\mathcal{F}}_{\mathbb{R}_{+}^{n}}, let supp(x¯)\operatorname{{supp}}(\bar{x}) be the support of x¯\bar{x}, {i{1,,n}:x¯i0}\{i\in\{1,\ldots,n\}:\bar{x}_{i}\neq 0\}. Let AA be a matrix representation of 𝒜:nm{\mathcal{A}}:\mathbb{R}^{n}\to\mathbb{R}^{m}. Then x¯\bar{x} is called degenerate if A(:,supp(x¯))A(:,\operatorname{{supp}}(\bar{x})) has linearly dependent rows; see [34, Example 3.3.1]. We note that the usual degeneracy of a basic feasible solution for the standard linear program agrees with the degeneracy discussed in [34, Example 3.3.1]. A basic feasible solution xx^{\prime} of the standard linear program is called degenerate if there is a basic variable equal to 0; see [4, Section 2]. This implies that A(:,supp(x))A(:,\operatorname{{supp}}(x^{\prime})) (a submatrix of the so-called basis matrix) is a full-column rank matrix with more rows than columns. The characterization immediately indicates that xx^{\prime} is degenerate.

A nice characterization of Definition 2.3 customized for 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} is available as well. For X¯𝕊+n\bar{X}\in{\mathcal{F}}_{\mathbb{S}_{+}^{n}}, let X¯=[V1V2][R000][V1TV2T]\bar{X}=\begin{bmatrix}V_{1}&V_{2}\end{bmatrix}\begin{bmatrix}R&0\\ 0&0\end{bmatrix}\begin{bmatrix}V_{1}^{T}\\ V_{2}^{T}\end{bmatrix} be a spectral decomposition of X¯\bar{X}, where RR is a positive definite matrix of order rank(X¯)\operatorname{{rank}}(\bar{X}). Let Ai,X𝕊n=bi\langle A_{i},X\rangle_{\mathbb{S}^{n}}=b_{i} be the ii-th equality of the system 𝒜X=bm{\mathcal{A}}X=b\in\mathbb{R}^{m}. Then [34, Corollary 3.3.2] states

X¯ is degenerate  {[V1TAiV1V1TAiV2]}i=1m contains linearly dependent matrices.\text{$\bar{X}$ is degenerate $\iff$ }\left\{\begin{bmatrix}V_{1}^{T}A_{i}V_{1}&V_{1}^{T}A_{i}V_{2}\end{bmatrix}\right\}_{i=1}^{m}\text{ contains linearly dependent matrices.} (2.6)

In [16, Section 2], the degeneracy of each point in 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} is realized using the characterization (2.6). We briefly explain the steps for recognizing the degeneracy. Since 𝕊+n\mathbb{S}_{+}^{n} is facially exposed, we always obtain face(,𝕊+n)=V𝕊+rVT\operatorname{face}({\mathcal{F}},\mathbb{S}_{+}^{n})=V\mathbb{S}_{+}^{r}V^{T} for some full column rank matrix Vn×rV\in\mathbb{R}^{n\times r}. Then a solution yy to (2.3) is used to detect the linear dependence of the matrices in (2.6) after FR. We leave more details to [16, Section 2].

The characterizations of degeneracies in the case of +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} and +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} owe to the computationally amenable facial structures (2.1) and (2.2). This again emphasizes the importance of understanding facial geometry of cones.

3 Main Result

We now present the main results of this paper.

Theorem 3.1 (Degeneracy in the absence of strict feasibility).

Suppose that strict feasibility fails for {\mathcal{F}}. Then every point in {\mathcal{F}} is degenerate.

Proof.

Suppose that strict feasibility fails for {\mathcal{F}}. Then by Lemma 2.1, there exists ymy\in\mathbb{R}^{m} that solves (2.3). Note that 𝒜y(𝒜){\mathcal{A}}^{*}y\in{\mathcal{R}}({\mathcal{A}}^{*}) and 𝒜y0{\mathcal{A}}^{*}y\neq 0. Now let x¯\bar{x}\in{\mathcal{F}}. Then

𝒜y,x¯=y,𝒜x¯=y,b=0 and 𝒜y𝒦.\langle{\mathcal{A}}^{*}y,\bar{x}\rangle=\langle y,{\mathcal{A}}\bar{x}\rangle=\langle y,b\rangle=0\text{ and }{\mathcal{A}}^{*}y\in{\mathcal{K}}^{*}.

Hence 𝒜yface(x¯,𝒦)Δ{\mathcal{A}}^{*}y\in\operatorname{face}(\bar{x},{\mathcal{K}})^{\Delta}. Therefore we conclude that 𝒜ylinface(x¯,𝒦)Δ{\mathcal{A}}^{*}y\in\operatorname{lin}\operatorname{face}(\bar{x},{\mathcal{K}})^{\Delta}. Consequently, by Definition 2.3, x¯\bar{x} is degenerate. ∎

Now the results in [18, 16] can be viewed as a consequence of Theorem 3.1. An immediate application of Theorem 3.1 yields that {\mathcal{F}} that has a nondegenerate point holds strict feasibility. Moreover, {\mathcal{F}} that holds strict feasibility always has a nondegenerate point as we see in Proposition 3.2 below.

Proposition 3.2.

Every strictly feasible point is nondegenerate.

Proof.

Let x¯\bar{x} be a strictly feasible point. Then face(x¯,𝒦)=𝒦\operatorname{face}(\bar{x},{\mathcal{K}})={\mathcal{K}}. If int𝒦\operatorname{{int}}{\mathcal{K}}\neq\emptyset, then Definition 2.3 immediately follows. Suppose to the contrary that

0ϕlin𝒦Δ(𝒜).0\neq\phi\in\operatorname{lin}{\mathcal{K}}^{\Delta}\cap{\mathcal{R}}({\mathcal{A}}^{*}).

Then ϕ=𝒜z\phi={\mathcal{A}}^{*}z for some zz. Note that 𝒜zlin𝒦Δ{\mathcal{A}}^{*}z\in\operatorname{lin}{\mathcal{K}}^{\Delta} and hence 𝒜z,x=0\langle{\mathcal{A}}^{*}z,x\rangle=0, for all x𝒦x\in{\mathcal{K}}. This yields 𝒜z𝒦{\mathcal{A}}^{*}z\in{\mathcal{K}}^{*} and 0=𝒜z,x¯=z,𝒜x¯=z,b0=\langle{\mathcal{A}}^{*}z,\bar{x}\rangle=\langle z,{\mathcal{A}}\bar{x}\rangle=\langle z,b\rangle. Therefore zz satisfies (2.3) and this contradicts the strict feasibility assumption. ∎

Theorem 3.1 and Proposition 3.2 lead to the following conclusion.

Corollary 3.3.

Given any feasible set {\mathcal{F}}, strict feasibility fails if, and only if, every point is degenerate. Moreover, strict feasibility holds if, and only if, {\mathcal{F}} contains a nondegenerate point.

If we restrict our attention to the extreme points of {\mathcal{F}} only, the converse of Theorem 3.1 is not true. For example, +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} that represents the set of doubly stochastic matrices (the feasible region of the linear assignment problem) has a Slater point, but every extreme point is degenerate; see [18].

We now proceed to the second main result. We first introduce a useful lemma. We include the proof of Lemma 3.4 for completeness.

Lemma 3.4.

[33, Lemma 3.6] Suppose that strict feasibility fails for {\mathcal{F}}. Let yy be a solution to (2.3). Then

𝒜(𝒦(𝒜y))=𝒜(𝒦)y.{\mathcal{A}}({\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp})={\mathcal{A}}({\mathcal{K}})\cap y^{\perp}. (3.1)
Proof.

Let w𝒜(𝒦(𝒜y))w\in{\mathcal{A}}({\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp}). Then w=𝒜xw={\mathcal{A}}x for some x𝒦(𝒜y)x\in{\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp}. Clearly w𝒜(𝒦)w\in{\mathcal{A}}({\mathcal{K}}). Note that w,y=𝒜x,y=b,y=0\langle w,y\rangle=\langle{\mathcal{A}}x,y\rangle=\langle b,y\rangle=0. Hence 𝒜(𝒦(𝒜y))𝒜(𝒦){y}{\mathcal{A}}({\mathcal{K}}\cap({\mathcal{A}}^{*}y)^{\perp})\subseteq{\mathcal{A}}({\mathcal{K}})\cap\{y\}^{\perp} holds. For the opposite containment, let w𝒜(𝒦){y}w\in{\mathcal{A}}({\mathcal{K}})\cap\{y\}^{\perp}. Then w=𝒜xw={\mathcal{A}}x for some x𝒦x\in{\mathcal{K}}. Furthermore, 0=w,y=𝒜x,y=x,𝒜y0=\langle w,y\rangle=\langle{\mathcal{A}}x,y\rangle=\langle x,{\mathcal{A}}^{*}y\rangle. ∎

Recall that 𝒜{\mathcal{A}} is assumed to be surjective throughout this paper. We emphasize that 𝒜X=b{\mathcal{A}}X=b alone does not have redundant equalities. Theorem 3.5 below shows that 𝒜{\mathcal{A}} is no longer surjective when the domain of 𝒜{\mathcal{A}} is restricted by the orthogonal complement of 𝒜y{\mathcal{A}}^{*}y. That is, the constraint system {\mathcal{F}} contains implicit redundant constraints.

Theorem 3.5 (Implicit loss of surjectivity in the absence of strict feasibility).

Suppose strict feasibility fails for {\mathcal{F}}. Let {yi}i=1k\{y^{i}\}_{i=1}^{k} be the set of vectors obtained by the FR process. Let 𝒦¯=𝒦i=1k(𝒜yi)\bar{{\mathcal{K}}}={\mathcal{K}}\cap_{i=1}^{k}({\mathcal{A}}^{*}y^{i})^{\perp} and let 𝔼¯=span(𝒦¯)\bar{\mathbb{E}}=\operatorname{{span}}(\bar{{\mathcal{K}}}). Then the map 𝒜:𝔼¯m{\mathcal{A}}:\bar{\mathbb{E}}\to\mathbb{R}^{m} is not surjective. In other words, the equality constraint system in {x𝒦¯:𝒜x=b}\{x\in\bar{{\mathcal{K}}}:{\mathcal{A}}x=b\} contains redundant constraints.

Proof.

Suppose strict feasibility fails for {\mathcal{F}}. Let 𝒦¯\bar{{\mathcal{K}}} be the closed convex cone obtained by the FR process, i.e., 𝒦¯:=𝒦i=1k(𝒜yi)\bar{{\mathcal{K}}}:={\mathcal{K}}\cap_{i=1}^{k}({\mathcal{A}}^{*}y^{i})^{\perp}, where {yi}i=1k\{y^{i}\}_{i=1}^{k} is obtained by solving (2.3). Then by Lemma 3.4, y1my^{1}\in\mathbb{R}^{m} satisfies (3.1). Thus the containment below follows:

𝒜(𝒦¯)𝒜(𝒦(𝒜y1))=𝒜(𝒦)(y1).{\mathcal{A}}(\bar{{\mathcal{K}}})\subseteq{\mathcal{A}}({\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp})={\mathcal{A}}({\mathcal{K}})\cap(y^{1})^{\perp}.

Note that y1y^{1} is nonzero since 𝒜y10{\mathcal{A}}^{*}y^{1}\neq 0. Hence the range of 𝒜{\mathcal{A}} with dom(𝒜)=span(𝒦¯)\operatorname{{dom}}({\mathcal{A}})=\operatorname{{span}}(\bar{{\mathcal{K}}}) cannot span m\mathbb{R}^{m}. ∎

Theorem 3.5 leads to the notion of implicit problem singularity, ips\operatorname{ips}, discussed [18, 15, 16]. We state the definition of ips\operatorname{ips} for a general conic program.

Definition 3.6.

Let 𝒦(𝒜y1)(𝒜yk){\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp}\cap\cdots\cap({\mathcal{A}}^{*}y^{k})^{\perp} be the cone obtained by the FR process. The implicit problem singularity of {\mathcal{F}}, ips()\operatorname{ips}({\mathcal{F}}), is the number of redundant equalities in the system

{x𝒦(𝒜y1)(𝒜yk):𝒜x=b}.\left\{x\in{\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp}\cap\cdots\cap({\mathcal{A}}^{*}y^{k})^{\perp}:{\mathcal{A}}x=b\right\}.

We discuss notions related to Definition 3.6 and their usages in Section 4 below.

Remark 3.7.

The implicit redundancies provide an interesting implication to the minimal representation of the affine space discussed in [32]. Let 𝒩(𝒜){\mathcal{N}}({\mathcal{A}}) be the null-space of 𝒜{\mathcal{A}}. Then, given x^={x𝔼:𝒜x=b}\hat{x}\in{\mathcal{L}}=\{x\in\mathbb{E}:{\mathcal{A}}x=b\}, we obtain =x^+𝒩(𝒜){\mathcal{L}}=\hat{x}+{\mathcal{N}}({\mathcal{A}}). Since 𝒦=face(,𝒦){\mathcal{K}}\cap{\mathcal{L}}=\operatorname{face}({\mathcal{F}},{\mathcal{K}})\cap{\mathcal{L}} and (face(,𝒦)face(,𝒦))𝒦=face(,𝒦)(\operatorname{face}({\mathcal{F}},{\mathcal{K}})-\operatorname{face}({\mathcal{F}},{\mathcal{K}}))\cap{\mathcal{K}}=\operatorname{face}({\mathcal{F}},{\mathcal{K}}), [32] recognizes

=(x^+𝒩(𝒜))𝒦=(x^+𝒩(𝒜))(face(,𝒦)face(,𝒦))𝒦.{\mathcal{F}}=(\hat{x}+{\mathcal{N}}({\mathcal{A}}))\cap{\mathcal{K}}=(\hat{x}+{\mathcal{N}}({\mathcal{A}}))\cap(\operatorname{face}({\mathcal{F}},{\mathcal{K}})-\operatorname{face}({\mathcal{F}},{\mathcal{K}}))\cap{\mathcal{K}}.

Then [32, equation (2.18)] establishes DM=(x^+𝒩(𝒜))(face(,𝒦)face(,𝒦)){\mathcal{L}}_{DM}=(\hat{x}+{\mathcal{N}}({\mathcal{A}}))\cap(\operatorname{face}({\mathcal{F}},{\mathcal{K}})-\operatorname{face}({\mathcal{F}},{\mathcal{K}})) as ‘the minimal subspace representation’ of {\mathcal{L}}. Theorem 3.5 and Definition 3.6 certify that DM{\mathcal{L}}_{DM} indeed has a smaller dimension than {\mathcal{L}} when strict feasibility fails for {\mathcal{F}}.

4 Discussions

In this section we connect the main results to some topics in the literature, including the notions of different singularities and their usages, ill-conditioning that arises in the interior point methods, and differentiability of the optimal value function.

Other Singularity Notions

Theorem 3.5 indicates that the image under 𝒜{\mathcal{A}} with its domain restricted by the FR process cannot span m\mathbb{R}^{m}. Yet, it does not specify how different the dimension of the image is from m\mathbb{R}^{m} in the absence of strict feasibility. We aim to provide a bound for the lost dimension.

As a related definition to the singularity degree, a notion of max-singularity degree is proposed in [18]. The max-singularity degree of {\mathcal{F}}, maxsd()\operatorname{maxsd}({\mathcal{F}}), is the longest nontrivial FR iterations for {\mathcal{F}}. We discuss some properties related to these singularity notions. Proposition 4.1 below is proved for the class of semidefinite programs using the facial structure of 𝕊+n\mathbb{S}_{+}^{n} explicitly; see [25, Lemma 3.5.2]. We present a generalization.

Proposition 4.1.

Suppose that {\mathcal{F}} fails strict feasibility. Let {yi}i=1k\{y^{i}\}_{i=1}^{k} generated by solving (2.3). Then the vectors in {yi}i=1k\{y^{i}\}_{i=1}^{k} are linearly independent.

Proof.

Suppose strict feasibility fails for {\mathcal{F}}. Let {yi}i=1k\{y^{i}\}_{i=1}^{k} be a sequence of vectors generated by FR iterations from Algorithm 2.1. We first show that the members in {𝒜yi}i=1k\{{\mathcal{A}}^{*}y^{i}\}_{i=1}^{k} are linearly independent. Let

𝒦k1:=𝒦(𝒜y1)(𝒜yk1).{\mathcal{K}}^{k-1}:={\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp}\cap\cdots\cap({\mathcal{A}}^{*}y^{k-1})^{\perp}.

Note that 𝒦k1{0}{\mathcal{K}}^{k-1}\neq\{0\}, since 𝒦k1={0}{\mathcal{K}}^{k-1}=\{0\} implies that ={0}{\mathcal{F}}=\{0\} and the FR process terminates before the kk-th iteration. Since

𝒜yk(𝒦k1){0} and b,y=0,{\mathcal{A}}^{*}y^{k}\in({\mathcal{K}}^{k-1})^{*}\setminus\{0\}\text{ and }\langle b,y\rangle=0,

we have

𝒜yk,x0,x𝒦k1.\langle{\mathcal{A}}^{*}y^{k},x\rangle\geq 0,\ \forall x\in{\mathcal{K}}^{k-1}.

We consider two cases below:

  1. 1.

    case 1: 𝒜yk,x¯=0\langle{\mathcal{A}}^{*}y^{k},\bar{x}\rangle=0, for some x¯𝒦k1{0}\bar{x}\in{\mathcal{K}}^{k-1}\setminus\{0\}.

    Note that x¯(𝒜y)\bar{x}\in({\mathcal{A}}^{*}y^{\ell})^{\perp}, {1,,k1}\forall\ell\in\{1,\ldots,k-1\}. Thus 𝒜yk{\mathcal{A}}^{*}y^{k} must be orthogonal to all members in {𝒜y}=1k1\{{\mathcal{A}}^{*}y^{\ell}\}_{\ell=1}^{k-1}. Therefore 𝒜ykspan{𝒜y}=1k1{\mathcal{A}}^{*}y^{k}\notin\operatorname{{span}}\{{\mathcal{A}}^{*}y^{\ell}\}_{\ell=1}^{k-1}.

  2. 2.

    case 2: 𝒜yk,x¯>0\langle{\mathcal{A}}^{*}y^{k},\bar{x}\rangle>0, for some x¯𝒦k1\bar{x}\in{\mathcal{K}}^{k-1}.

    Note that

     for each {1,,k1},𝒜y,x=0,x𝒦k1.\text{ for each }\ell\in\{1,\ldots,k-1\},\ \langle{\mathcal{A}}^{*}y^{\ell},x\rangle=0,\ \forall x\in{\mathcal{K}}^{k-1}.

    If 𝒜ykspan{𝒜y}=1k1{\mathcal{A}}^{*}y^{k}\in\operatorname{{span}}\{{\mathcal{A}}^{*}y^{\ell}\}_{\ell=1}^{k-1} holds, then 𝒜yk,x¯=0\langle{\mathcal{A}}^{*}y^{k},\bar{x}\rangle=0. This is a contradiction.

Since the members in {𝒜yi}i=1k\{{\mathcal{A}}^{*}y^{i}\}_{i=1}^{k} are linearly independent, it immediately follows that the vectors in {yi}i=1k\{y^{i}\}_{i=1}^{k} are linearly independent. ∎

Note that the property stated in Proposition 4.1 holds for any FR process. Thus Corollary 4.2 below follows.

Corollary 4.2.

Suppose that strict feasibility fails for {\mathcal{F}} and let 𝒦¯\bar{{\mathcal{K}}} be the cone obtained by the FR process. Then 𝒜(span𝒦¯){\mathcal{A}}(\operatorname{{span}}\bar{{\mathcal{K}}}) generates at most mmaxsd()m-\operatorname{maxsd}({\mathcal{F}}) dimensional space. Moreover, {x𝒦¯:𝒜x=b}\{x\in\bar{{\mathcal{K}}}:{\mathcal{A}}x=b\} contains at least maxsd()\operatorname{maxsd}({\mathcal{F}}) number of implicitly redundant constraints.

Proof.

Let {yi}i=1maxsd()\{y^{i}\}_{i=1}^{\operatorname{maxsd}({\mathcal{F}})} be the sequence of vectors obtained by the FR iterations by solving (2.3). Then Lemma 3.4 implies that

𝒜(𝒦¯)=𝒜(𝒦(𝒜y1)(𝒜ymaxsd()))=𝒜(𝒦){y1}{ymaxsd()}.{\mathcal{A}}(\bar{{\mathcal{K}}})={\mathcal{A}}\left({\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp}\cap\cdots({\mathcal{A}}^{*}y^{\operatorname{maxsd}({\mathcal{F}})})^{\perp}\right)={\mathcal{A}}({\mathcal{K}})\cap\{y^{1}\}^{\perp}\cap\cdots\{y^{\operatorname{maxsd}({\mathcal{F}})}\}^{\perp}.

Finally, Proposition 4.1 implies that 𝒜(span𝒦¯){\mathcal{A}}(\operatorname{{span}}\bar{{\mathcal{K}}}) spans at most mmaxsd()m-\operatorname{maxsd}({\mathcal{F}}) dimensional space. ∎

Corollary 4.2 gives rise to the following relationship:

ips()maxsd()sd().\operatorname{ips}({\mathcal{F}})\geq\operatorname{maxsd}({\mathcal{F}})\geq\operatorname{sd}({\mathcal{F}}).

This motivates the following question: how different can the ips,maxsd\operatorname{ips},\operatorname{maxsd}, and sd\operatorname{sd} be? Various instances are presented for +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} and 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} in the literature; sd(𝕊+n)=maxsd(𝕊+n)<ips(𝕊+n)\operatorname{sd}({\mathcal{F}}_{\mathbb{S}_{+}^{n}})=\operatorname{maxsd}({\mathcal{F}}_{\mathbb{S}_{+}^{n}})<\operatorname{ips}({\mathcal{F}}_{\mathbb{S}_{+}^{n}}) [15, Example 3.2.13]; sd(𝕊+n)=maxsd(𝕊+n)=ips(𝕊+n)\operatorname{sd}({\mathcal{F}}_{\mathbb{S}_{+}^{n}})=\operatorname{maxsd}({\mathcal{F}}_{\mathbb{S}_{+}^{n}})=\operatorname{ips}({\mathcal{F}}_{\mathbb{S}_{+}^{n}}) [25, Example 4.2.6]; and sd(+n)=1<ips(+n)=275\operatorname{sd}({\mathcal{F}}_{\mathbb{R}_{+}^{n}})=1<\operatorname{ips}({\mathcal{F}}_{\mathbb{R}_{+}^{n}})=275 [18, Section 4.2.2]. Example 2.2 above shows that sd(Kexp)=maxsd(Kexp)=ips(Kexp)=2\operatorname{sd}({\mathcal{F}}_{K_{\text{exp}}})=\operatorname{maxsd}({\mathcal{F}}_{K_{\text{exp}}})=\operatorname{ips}({\mathcal{F}}_{K_{\text{exp}}})=2.

The implicit redundancies can be used to tighten bounds that engage the number of linear constraints. An interesting usage of ips\operatorname{ips} is shown for +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}} and 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}} in the literature. For the case of +n{\mathcal{F}}_{\mathbb{R}_{+}^{n}}, ips\operatorname{ips} is used for measuring the degree of degeneracy of a basic feasible solution [15, Corollary 4.1.12]. More specifically, every basic feasible solution has at most mips(+n)m-\operatorname{ips}({\mathcal{F}}_{\mathbb{R}_{+}^{n}}) number of nonzero basic variables. This also translates to

every basic feasible solution has at least ips(+n)\operatorname{ips}({\mathcal{F}}_{\mathbb{R}_{+}^{n}}) degenerate basic variables. (4.1)

In the case of 𝕊+n{\mathcal{F}}_{\mathbb{S}_{+}^{n}}, ips\operatorname{ips} is used to tighten the Barvinok-Pataki bound [15, Theorem 3.2.7]:

every extreme point X𝕊+n satisfies rank(X)(rank(X)+1)/2mips(𝕊+n).\text{every extreme point }X\in{\mathcal{F}}_{\mathbb{S}_{+}^{n}}\text{ satisfies }\operatorname{{rank}}(X)(\operatorname{{rank}}(X)+1)/2\leq m-\operatorname{ips}({\mathcal{F}}_{\mathbb{S}_{+}^{n}}). (4.2)

We observe that a large ips(𝕊+n)\operatorname{ips}({\mathcal{F}}_{\mathbb{S}_{+}^{n}}) provides an especially useful bound since the term on the left-hand-side of the inequality (4.2) is quadratic in rank(X)\operatorname{{rank}}(X).

We now aim to generalize (4.1) and (4.2) to an arbitrary conic program. We first present a lemma. Note that Lemma 4.3 can be used to produce the well-known Barvinok-Pataki bound [3, 21].

Lemma 4.3.

(Pataki) [34, Theorem 3.3.1] Let x¯\bar{x}\in{\mathcal{F}}. Then

dim[face(x¯,)]=dim[face(x¯,𝒦)]m+dim[face(x¯,𝒦)(𝒜)].\dim[\operatorname{face}(\bar{x},{\mathcal{F}})]=\dim[\operatorname{face}(\bar{x},{\mathcal{K}})]-m+\dim[\operatorname{face}(\bar{x},{\mathcal{K}})^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})].
Theorem 4.4.

Let x¯\bar{x} be an extreme point of {\mathcal{F}}. Then

dimface(x¯,𝒦)mips().\dim\operatorname{face}(\bar{x},{\mathcal{K}})\leq m-\operatorname{ips}({\mathcal{F}}).
Proof.

Let {yi}i=1k\{y^{i}\}_{i=1}^{k} be a set of vectors obtained by the FR iterations. Let 𝒦¯:=𝒦(𝒜y1)(𝒜yk)\bar{{\mathcal{K}}}:={\mathcal{K}}\cap({\mathcal{A}}^{*}y^{1})^{\perp}\cap\cdots\cap({\mathcal{A}}^{*}y^{k})^{\perp}. Let 𝒩(𝒜){\mathcal{N}}({\mathcal{A}}) be the null-space of 𝒜{\mathcal{A}}. Note that

m=dim(𝒜(𝒦¯𝒦¯))dim(𝒜(𝒦¯))+dim(𝒜(𝒦¯)).m=\dim\left({\mathcal{A}}\left(\bar{{\mathcal{K}}}\cup\bar{{\mathcal{K}}}^{\perp}\right)\right)\leq\dim({\mathcal{A}}\left(\bar{{\mathcal{K}}})\right)+\dim\left({\mathcal{A}}(\bar{{\mathcal{K}}}^{\perp})\right). (4.3)

Since

𝒜(𝒦¯)=𝒜((𝒦¯(𝒜))𝒩(𝒜))𝒜(𝒦¯(𝒜))𝒜(𝒩(𝒜))=𝒜(𝒦¯(𝒜)),\begin{array}[]{rcl}{\mathcal{A}}(\bar{{\mathcal{K}}}^{\perp})&=&{\mathcal{A}}\left(\left(\bar{{\mathcal{K}}}^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})\right)\cup{\mathcal{N}}({\mathcal{A}})\right)\\ &\subseteq&{\mathcal{A}}\left(\bar{{\mathcal{K}}}^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})\right)\cup{\mathcal{A}}\left({\mathcal{N}}({\mathcal{A}})\right)\\ &=&{\mathcal{A}}\left(\bar{{\mathcal{K}}}^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})\right),\end{array}

we have

dim[𝒜(𝒦¯)]dim[𝒜(𝒦¯(𝒜))]dim[𝒦¯(𝒜)]dim[face(,𝒦)(𝒜)].\begin{array}[]{rcl}\dim[{\mathcal{A}}(\bar{{\mathcal{K}}}^{\perp})]&\leq&\dim[{\mathcal{A}}(\bar{{\mathcal{K}}}^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*}))]\\ &\leq&\dim[\bar{{\mathcal{K}}}^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})]\\ &\leq&\dim[\operatorname{face}({\mathcal{F}},{\mathcal{K}})^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})].\end{array} (4.4)

Since x¯\bar{x} is an extreme point of {\mathcal{F}}, dim[face(x¯,)]=0\dim[\operatorname{face}(\bar{x},{\mathcal{F}})]=0. Then Lemma 4.3 implies

dim[face(x¯,𝒦)]=mdim[face(x¯,𝒦)(𝒜)].\dim[\operatorname{face}(\bar{x},{\mathcal{K}})]=m-\dim[\operatorname{face}(\bar{x},{\mathcal{K}})^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})]. (4.5)

Note that

face(x¯,𝒦)face(,𝒦)dim[face(x¯,𝒦)(𝒜)]dim[face(,𝒦)(𝒜)].\operatorname{face}(\bar{x},{\mathcal{K}})^{\perp}\supseteq\operatorname{face}({\mathcal{F}},{\mathcal{K}})^{\perp}\implies\dim\left[\operatorname{face}(\bar{x},{\mathcal{K}})^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})\right]\geq\dim\left[\operatorname{face}({\mathcal{F}},{\mathcal{K}})^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})\right].

Then we obtain

dim[face(x¯,𝒦)]mdim[face(,𝒦)(𝒜)]by (4.5)mdim[𝒜(𝒦¯)]by (4.4)dim[𝒜(𝒦¯)]by (4.3)mips().\begin{array}[]{rcll}\dim[\operatorname{face}(\bar{x},{\mathcal{K}})]&\leq&m-\dim\left[\operatorname{face}({\mathcal{F}},{\mathcal{K}})^{\perp}\cap{\mathcal{R}}({\mathcal{A}}^{*})\right]&\text{by {(\ref{eq:dimfacexK_almostubd})}}\\ &\leq&m-\dim[{\mathcal{A}}(\bar{{\mathcal{K}}}^{\perp})]&\text{by {(\ref{eq:proof_aid1})}}\\ &\leq&\dim[{\mathcal{A}}(\bar{{\mathcal{K}}})]&\text{by {(\ref{Adimbreakup})}}\\ &\leq&m-\operatorname{ips}({\mathcal{F}}).\end{array}

Theorem 4.4 can be used for cones for which facial dimensions are known. For example, dimface(x¯,+n)=|supp(x¯)|\dim\operatorname{face}(\bar{x},\mathbb{R}_{+}^{n})=|\operatorname{{supp}}(\bar{x})|. Thus (4.1) can be viewed as a consequence of Theorem 4.4. It is well known that dimface(X¯,𝕊+n)=rank(X¯)(rank(X¯)+1)/2\dim\operatorname{face}(\bar{X},\mathbb{S}_{+}^{n})=\operatorname{{rank}}(\bar{X})(\operatorname{{rank}}(\bar{X})+1)/2 and (4.2) also follows from Theorem 4.4. Let COPn:={X𝕊n:y,Xyn,y+n}\text{COP}^{n}:=\{X\in\mathbb{S}^{n}:\langle y,Xy\rangle_{\mathbb{R}^{n}},\forall y\in\mathbb{R}_{+}^{n}\} be the copositive cone. The dimensions of the so-called maximal faces of the copositive cone are identified. A maximal face ff of COPn\text{COP}^{n} is characterized by f={XCOPn:vTXv=0}f=\{X\in\text{COP}^{n}:v^{T}Xv=0\} for some v+n{0}v\in\mathbb{R}_{+}^{n}\setminus\{0\}; see [10, Theorem 5.5]. Moreover, dimf=12n(n+1)|supp(v)|\dim f=\frac{1}{2}n(n+1)-|\operatorname{{supp}}(v)|. Then by Theorem 4.4, we conclude that an extreme point X¯\bar{X} of COPn{\mathcal{F}}_{\text{COP}^{n}}, where face(X¯,COPn)\operatorname{face}(\bar{X},\text{COP}^{n}) is maximal, satisfies

12n(n+1)|supp(v)|mips().\frac{1}{2}n(n+1)-|\operatorname{{supp}}(v)|\leq m-\operatorname{ips}({\mathcal{F}}).

Interior Point Methods

Interior point methods are the most popular class of algorithms for solving a linear conic program

minx{c,x:𝒜x=b,x𝒦}.\min_{x}\{\langle c,x\rangle:{\mathcal{A}}x=b,x\in{\mathcal{K}}\ \}.

A general framework of the interior point method applicable to the standard conic program is discussed in [20]. We aim to identify a repercussion that arises when the primal path-following interior point method is applied to an instance that fails strict feasibility. We follow the steps presented in [20]. First we construct the so-called logarithmically homogeneous self-concordant barrier (LHSCB) F:int(𝒦)F:\operatorname{{int}}({\mathcal{K}})\to\mathbb{R} for the cone 𝒦{\mathcal{K}}. For t>0t>0, consider

(PBt)minx{tc,x+F(x):Ax=b,x𝒦}.(PB_{t})\quad\min_{x}\{\ t\langle c,x\rangle+F(x):Ax=b,x\in{\mathcal{K}}\ \}.

The barrier function FF maintains the variable xx in int(𝒦)\operatorname{{int}}({\mathcal{K}}). The barrier functions F(x)=jlnxjF(x)=-\sum_{j}\ln x_{j}, and F(X)=lndetXF(X)=-\ln\det X are presented as examples for linear and semidefinite programs, respectively. The first-order optimality conditions for (PBt)(PB_{t}) are

𝒜y+s=c𝒜x=bF(x)+ts=0,xint(𝒦),sint(𝒦)\begin{array}[]{l}{\mathcal{A}}^{*}y+s=c\\ {\mathcal{A}}x=b\\ \nabla F(x)+ts=0,x\in\operatorname{{int}}({\mathcal{K}}),s\in\operatorname{{int}}({\mathcal{K}}^{*})\end{array} (4.6)

The Naewton step for solving the optimality conditions (4.6) can be computed by solving

𝒜Δy+Δs=0𝒜Δx=01t2F(x)Δx+Δs=s1tF(x)\begin{array}[]{l}{\mathcal{A}}^{*}\Delta y+\Delta s=0\\ {\mathcal{A}}\Delta x=0\\ \frac{1}{t}\nabla^{2}F(x)\Delta x+\Delta s=-s-\frac{1}{t}\nabla F(x)\end{array} (4.7)

Then by substituting Δs,Δx\Delta s,\Delta x into the middle block in (4.7), we obtain [20, equation 3.9]:

(𝒜[F2(x)]1𝒜)Δy¯=𝒜[2F(x)]1(s+t+1F(x)).({\mathcal{A}}[\nabla F^{2}(x)]^{-1}{\mathcal{A}}^{*})\overline{\Delta y}={\mathcal{A}}[\nabla^{2}F(x)]^{-1}(s+t_{+}^{-1}\nabla F(x)). (4.8)

Note that for infeasible-interior point method, some appropriate residuals are added to the right-hand-side of (4.8) but the data on the left-hand-side remains the same.

We now observe the limiting behaviour of the left-hand-side matrix A[F2(x)]1ATA[\nabla F^{2}(x)]^{-1}A^{T} in (4.8) with the barrier functions introduced in [20]. For the case of the linear program, we see that [2F(x)]1=Diag(x)2[\nabla^{2}F(x)]^{-1}=\operatorname{{Diag}}(x)^{2}, where Diag(x)\operatorname{{Diag}}(x) is the diagonal matrix with xx placed on the diagonal elements. Let xkx_{k} and the kk-th iterate. Let xx^{*} be an optimal solution to (PBt)(PB_{t}) and let Ax=ADiag(x)A_{x*}=A\operatorname{{Diag}}(x^{*}). We note that ADiag(x)2AT=AxAxTA\operatorname{{Diag}}(x^{*})^{2}A^{T}=A_{x*}A_{x*}^{T} and range(Ax)m\operatorname{range}(A_{x*})\subsetneq\mathbb{R}^{m}. Thus, ADiag(x)2ATA\operatorname{{Diag}}(x^{*})^{2}A^{T} is a singular matrix. As xkxx_{k}\to x^{*}, ill-conditioning arises. For the case of the semidefinite program, the (i,j)(i,j)-th entry of the data matrix in (4.8) is Ai,XAjX\langle A_{i},XA_{j}X\rangle. Let XX^{*} be an optimal solution and let MXM_{X^{*}} be the matrix where (MX)i,j=Ai,XAjX(M_{X^{*}})_{i,j}=\langle A_{i},X^{*}A_{j}X^{*}\rangle. Let yy be a vector that solves (2.3). Then we observe

MXy=[jyjtrace(A1XAjX)jyjtrace(AmXAjX)]=[trace(A1XjyjAjX)trace(AmXjyjAjX)]=0,M_{X^{*}}y=\begin{bmatrix}\sum_{j}y_{j}\operatorname{{trace}}(A_{1}X^{*}A_{j}X^{*})\\ \vdots\\ \sum_{j}y_{j}\operatorname{{trace}}(A_{m}X^{*}A_{j}X^{*})\end{bmatrix}=\begin{bmatrix}\operatorname{{trace}}\left(A_{1}X^{*}\sum_{j}y_{j}A_{j}X^{*}\right)\\ \vdots\\ \operatorname{{trace}}\left(A_{m}X^{*}\sum_{j}y_{j}A_{j}X^{*}\right)\end{bmatrix}=0,

since (jyjAj)X=0(\sum_{j}y_{j}A_{j})X^{*}=0. Hence MXM_{X^{*}} is a singular matrix. Similarly, ill-conditioning arises in (4.8) as iterate XkX_{k} approaches to XX^{*}.

Beside the fact that the dual optimal set is unbounded in the absence of strict feasibility, we see an ill-conditioned linear system that arises when an interior point method is used. The importance of strict feasibility for different classes numerical solvers is addressed in [18, 16]. [18] links the degeneracy to the stalling phenomena of the simplex method and the ill-conditioning that arises in the primal-dual path-following interior point methods for linear programs (even through the self-dual embedding). [16] presents a semi-smooth Newton method for solving the nearest point problem and why strict feasibility is needed for having a good performance of the algorithm.

Differentiability of the Optimal Value Function

The differentiability of the optimal value function is an interesting topic in the field of perturbation analysis. For the class of linear programs, many results are applicable under the nondegeneracy setting [13, 12]. Let f:𝔼f:\mathbb{E}\to\mathbb{R} be a convex function. We show that the optimal value function p:m{}p^{*}:\mathbb{R}^{m}\to\mathbb{R}\cup\{\infty\} defined by

p(Δb):=minx{f(x):𝒜x=b+Δb,x𝒦}p^{*}(\Delta b):=\min_{x}\{f(x):{\mathcal{A}}x=b+\Delta b,x\in{\mathcal{K}}\} (4.9)

is not differentiable at 0 in the absence of strict feasibility. Proposition 4.5 is established under 𝒦=+n{\mathcal{K}}=\mathbb{R}_{+}^{n} and 𝒦=𝕊+n{\mathcal{K}}=\mathbb{S}_{+}^{n} in [15, 18]. We list the result in this paper as a generalization to [15, 18].

Proposition 4.5.

For any {\mathcal{F}}, perturbing bb by a solution y¯\bar{y} of (2.3) leads to an infeasible problem.

Proof.

Let y¯\bar{y} be a solution to (2.3). We consider b=ϵy¯b=-\epsilon\bar{y} for some ϵ>0\epsilon>0. Then replacing the right-hand-side vector bb of {\mathcal{F}} by bϵy¯b-\epsilon\bar{y} implies

𝒜y𝒦, and bϵy¯,y¯=0ϵy2<0.{\mathcal{A}}^{*}y\in{\mathcal{K}}^{*},\text{ and }\langle b-\epsilon\bar{y},\bar{y}\rangle=0-\epsilon\|y\|^{2}<0.

By Farkas’ lemma, the perturbed system is infeasible. ∎

Proposition 4.5 immediately implies Corollary 4.6 below.

Corollary 4.6.

Suppose that strict feasibility fails for {\mathcal{F}}. Then the optimal value function p(Δb)p^{*}(\Delta b) defined in (4.9) is not differentiable at 0.

Proof.

Setting Δb=ϵy¯\Delta b=-\epsilon\bar{y} for an arbitrary small ϵ>0\epsilon>0 yields p(Δb)=p^{*}(\Delta b)=\infty. ∎

5 Conclusions

We showed two universal properties of the standard conic program that fails strict feasibility. First we showed that every point is degenerate in the absence of strict feasibility. We also noted that the application of FR can alter the algebraic interpretation of a point. For instance, in the absence of strict feasibility, a strictly feasible point in the facially reduced set is nondegenerate, whereas the same point is degenerate before FR. Understanding the points for which FR influences the degeneracy status presents an interesting avenue for exploration. Second, we showed that the constraint system that fails strict feasibility inherits implicitly redundant constraints. We saw different usages of the singularity notions conveyed by the FR process. While sd()\operatorname{sd}({\mathcal{F}}) quantifies the computational challenge associated with solving a given problem, maxsd()\operatorname{maxsd}({\mathcal{F}}) provides a measure for the number of implicitly redundant linear constraints. We further elucidated on how the implicit redundancies impact the bound that engages the number of constraints and the near optimal behaviour of the interior point methods.

Acknowledgement

The author would like to thank professor Henry Wolkowicz for providing helpful comments.

References

  • [1] E.D. Andersen. Finding all linearly dependent rows in large-scale linear programming. Optimization methods & software, 6(3):219–227, 1995.
  • [2] MOSEK ApS. The MOSEK optimization toolbox for MATLAB manual. Version 10.0., 2022.
  • [3] A. I. Barvinok. A remark on the rank of positive semidefinite matrices subject to affine constraints. Discrete and Computational Geometry, 25:23–31, 2001.
  • [4] D. Bertsimas and J. Tsitsiklis. Introduction to Linear Optimization. Athena Scientific, Belmont, MA, 1997.
  • [5] J.M. Borwein and H. Wolkowicz. Characterization of optimality for the abstract convex program with finite-dimensional range. J. Austral. Math. Soc. Ser. A, 30(4):390–411, 1980/81.
  • [6] J.M. Borwein and H. Wolkowicz. Regularizing the abstract convex program. J. Math. Anal. Appl., 83(2):495–530, 1981.
  • [7] F. Burkowski, Y-L. Cheung, and H. Wolkowicz. Efficient use of semidefinite programming for selection of rotamers in protein conformations. INFORMS J. Comput., 26(4):748–766, 2014.
  • [8] C. Cartis and N. Gould. Finding a point in the relative interior of a polyhedron. Council for the Central Laboratory of the Research Councils, pages 1–59, 2006.
  • [9] Y.-L. Cheung. Preprocessing and Reduction for Semidefinite Programming via Facial Reduction: Theory and Practice. PhD thesis, University of Waterloo, 2013.
  • [10] P. J.C. Dickinson. Geometry of the copositive and completely positive cones. Journal of Mathematical Analysis and Applications, 380(1):377 – 395, 2011.
  • [11] D. Drusvyatskiy and H. Wolkowicz. The many faces of degeneracy in conic optimization. Foundations and Trends® in Optimization, 3(2):77–170, 2017.
  • [12] A.V. Fiacco. Introduction to Sensitivity and Stability Analysis in Nonlinear Programming, volume 165 of Mathematics in Science and Engineering. Academic Press, 1983.
  • [13] R.M. Freund. Postoptimal analysis of a linear program under simultaneous changes in matrix coefficients. Number 24, pages 1–13, 1985. Mathematical programming, I.
  • [14] H. Hu, H. Im, J. Lin, N. Lütkenhaus, and H. Wolkowicz. Robust Interior Point Method for Quantum Key Distribution Rate Computation. Quantum, 6:792, September 2022.
  • [15] H. Im. Implicit Loss of Surjectivity and Facial Reduction: Theory and Applications. PhD thesis, University of Waterloo, 2023.
  • [16] H. Im, W. Jung, W.M. Moursi, D. Torregrosa-Belen, and H. Wolkowicz. Projection, stability and singularity degree for spectrahedra. Technical report, in progress. 30 pages.
  • [17] H. Im and H. Wolkowicz. A strengthened Barvinok-Pataki bound on SDP rank. Oper. Res. Lett., 49(6):837–841, 2021.
  • [18] H. Im and H. Wolkowicz. Revisiting degeneracy, strict feasibility, stability, in linear programming. European J. Oper. Res., 310(2):495–510, 2023.
  • [19] S.B. Lindstrom, B.F. Lourenço, and T.-K. Pong. Error bounds, facial residual functions and applications to the exponential cone. Mathematical programming, 200(1):229–278, 2023.
  • [20] A.S. Nemirovski and M.J. Todd. Interior-point methods for optimization. Acta Numer., 17:191–234, 2008.
  • [21] G. Pataki. On the rank of extreme matrices in semidefinite programs and the multiplicity of optimal eigenvalues. 23(2):339–358, 1998.
  • [22] G. Pataki and L. Tunçel. On the generic properties of convex optimization problems in conic form. Math. Programming, to appear.
  • [23] J. Renegar. Hyperbolic programs, and their derivative relaxations. Found. Comput. Math., 6(1):59–79, 2006.
  • [24] L. Schork and J. Gondzio. Rank revealing Gaussian elimination by the maximum volume concept. Linear Algebra Appl., 592:1–19, 2020.
  • [25] S. Sremac. Error bounds and singularity degree in semidefinite programming. PhD thesis, University of Waterloo, 2019.
  • [26] S. Sremac, H.J. Woerdeman, and H. Wolkowicz. Error bounds and singularity degree in semidefinite programming. SIAM J. Optim., 31(1):812–836, 2021.
  • [27] J.F Sturm. Using sedumi 1.02, a matlab toolbox for optimization over symmetric cones. Optimization methods and software, 11(1-4):625–653, 1999.
  • [28] J.F. Sturm. Error bounds for linear matrix inequalities. SIAM J. Optim., 10(4):1228–1248 (electronic), 2000.
  • [29] P. Tarazaga. Faces of the cone of Euclidean distance matrices: characterizations, structure and induced geometry. Linear Algebra Appl., 408:1–13, 2005.
  • [30] K.-C. Toh, M.J. Todd, and R.H. Tütüncü. On the implementation and usage of sdpt3 – a matlab software package for semidefinite-quadratic-linear programming, version 4.0. In Handbook on Semidefinite, Conic and Polynomial Optimization, International Series in Operations Research &\& Management Science, pages 715–754. Springer US, New York, NY, 2012.
  • [31] L. Tunçel. Polyhedral and semidefinite programming methods in combinatorial optimization. Fields Institute monographs ; 27. American Mathematical Society, Providence, RI, 2010.
  • [32] L. Tunçel and H. Wolkowicz. Strong duality and minimal representations for cone optimization. coap, 53(2):619–648, 2012.
  • [33] F. Wang and H. Wolkowicz. Singularity degree of non-facially exposed faces. Technical report, 2022 submitted. 19 pages.
  • [34] H. Wolkowicz, R. Saigal, and L. Vandenberghe, editors. Handbook of semidefinite programming. International Series in Operations Research & Management Science, 27. Kluwer Academic Publishers, Boston, MA, 2000. Theory, algorithms, and applications.
  • [35] Q. Zhang. Completely positive cones: Are they facially exposed? Linear algebra and its applications, 558:195–204, 2018.
  • [36] Q. Zhao, S.E. Karisch, F. Rendl, and H. Wolkowicz. Semidefinite programming relaxations for the quadratic assignment problem. volume 2, pages 71–109. 1998. Semidefinite programming and interior-point approaches for combinatorial optimization problems (Toronto, ON, 1996).