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

A note on faces of convex sets

Stephan Weis
(Date: 29 September 2024)
Abstract.

The faces of a convex set owe their relevance to an interplay between convexity and topology that is systematically studied in the work of Rockafellar. Infinite-dimensional convex sets are excluded from this theory as their relative interiors may be empty. Shirokov and the present author answered this issue by proving that every point in a convex set lies in the relative algebraic interior of the face it generates. This theorem is proved here in a simpler way, connecting ideas scattered throughout the literature. This article summarizes and develops methods for faces and their relative algebraic interiors and applies them to spaces of probability measures.

Key words and phrases:
Convex set, extreme set, face, face generated by a point, relative interior, relative algebraic interior, Radon-Nikodym derivative, convex core
2020 Mathematics Subject Classification:
52A05, 46E27

Dedicated to R. Tyrrell Rockafellar on His Ninetieth Birthday

1. Introduction

A face of a convex set KK in a real vector space VV is a convex subset of KK including every pair of points in KK that are the endpoints of some open segment intersected by this convex subset. One-point faces are in a one-to-one correspondence with extreme points and are useful in functional analysis [1, 5, 8, 15, 21]. Larger faces play a minor role beyond finite dimensions. A notable exception is Alfsen and Shultz’ work on operator algebras [2].

Faces of finite-dimensional convex sets owe their success to an interplay of convexity and topology that appears in Grünbaum’s work [13, Sec. 2.4] and is systematically studied by Rockafellar [17]. A central notion is the relative interior ri(K)\operatorname{ri}(K) of KK, the interior of KK in the induced topology on the affine hull aff(K)\operatorname{aff}(K) of KK. Lacking monotonicity, the operator ri\operatorname{ri} is not the interior operator of a topology. Imagine a side KK of a triangle LL in the Euclidean plane VV; although KLK\subset L holds, ri(K)\operatorname{ri}(K) and ri(L)\operatorname{ri}(L) are disjoint and nonempty. Rockafellar and Wets [18, p. 75] stress that the closure of ri(K)\operatorname{ri}(K) includes KK in a Euclidean space VV​. Notably, KK\neq\emptyset implies ri(K)\operatorname{ri}(K)\neq\emptyset.

The last assertion is false in a Hausdorff topological vector space, as the interior of KK can be empty and the affine hull of KK can still be equal to VV​.

  1. 1)

    A first example is the convex set K={xV:(x)>0}K=\{x\in V:\ell(x)>0\} defined by a discontinuous linear functional :V\ell:V\to\mathbb{R}, because ker()\ker(\ell) is dense in VV​.

  2. 2)

    The closed convex set K={fV:f0 a.e.}K=\{f\in V:\mbox{$f\geq 0$ a.e.}\} in the Banach space V=Lp([0,1])V=L^{p}([0,1]) of pp-integrable real functions, 1p<1\leq p<\infty, has no interior points. Every function fKf\in K is the limit of f1(1/n,1]1[0,1/n]f\cdot 1_{(1/n,1]}-1_{[0,1/n]} as nn\to\infty, where 1A1_{A} denotes the indicator function on a measurable set AA.

Avoiding empty interiors, Borwein and Goebel [7, Thm. 2.8] study modified interiors of convex sets in Banach spaces. This article, instead, focuses on those faces of a convex set that happen to have nonempty relative interiors.

Among all vector topologies on VV​, the greatest interior (with respect to inclusion) of a convex set is achieved by the finest locally convex topology 𝔗ω(V)\mathfrak{T}_{\omega}(V). We deduce this from the continuity of the map V\mathbb{R}\to V​, λx+λv\lambda\mapsto x+\lambda v, for every x,vVx,v\in V​, bearing in mind that a point xKx\in K lies in the interior of KK for 𝔗ω(V)\mathfrak{T}_{\omega}(V), if and only if for every line gg in VV containing xx, the intersection gKg\cap K includes an open segment containing xx [8, II.26]. The set of all such points xx is called the algebraic interior [5, 9] or “core” [9, 15] of KK.

Example 1) above ceases to exist in the topology 𝔗ω(V)\mathfrak{T}_{\omega}(V), which renders every linear functional continuous [8, II.26]. The empty interior of Example 2) persists in the topology 𝔗ω(V)\mathfrak{T}_{\omega}(V), as the interior of every closed convex subset of a Banach space is the algebraic interior of that set [9, p. 46]. Another example with empty algebraic interior is the set of univariate polynomials with real coefficients and positive leading coefficients, which Barvinok examines in several revealing exercises [5, III.1.6]. By contrast, if dim(V)<\dim(V)<\infty then 𝔗ω(V)\mathfrak{T}_{\omega}(V) is the Euclidean topology and, again, KK\neq\emptyset implies ri(K)\operatorname{ri}(K)\neq\emptyset.

The relative interior ri(K)\operatorname{ri}(K) for the topology 𝔗ω(V)\mathfrak{T}_{\omega}(V) can be described in terms of the Euclidean topology on a line. Shirokov and the present author [20] define the relative algebraic interior rai(K)\operatorname{rai}(K) of KK as the set of points xKx\in K such that for every line gg in aff(K)\operatorname{aff}(K) containing xx, the intersection gKg\cap K includes an open segment containing xx. This definition differs from that of the algebraic interior just in the affine space confining the lines. Holmes [15] calls rai(K)\operatorname{rai}(K) the “intrinsic core”. We prove ri(K)=rai(K)\operatorname{ri}(K)=\operatorname{rai}(K) for 𝔗ω(V)\mathfrak{T}_{\omega}(V) in Sec. 3. Topological vector spaces and relative interiors are ignored subsequently; instead, relative algebraic interiors are used consistently.

In their studies of constrained density operators [20], Shirokov and this author employ the face of KK generated by a point xKx\in K. This is the smallest face of KK containing xx, which we denote by FK(x)F_{K}(x). A basic property is that xrai(FK(x))x\in\operatorname{rai}(F_{K}(x)) holds for all xKx\in K.

This key result is deduced in Sec. 4 from two elementary assertions, whereas our prior proof unnecessarily employs the Kuratowski-Zorn lemma. The first elementary assertion (Prop. 4.1) is Alfsen’s formula [1, p. 121]

FK(x)={yKϵ>0:x+ϵ(xy)K}.F_{K}(x)=\left\{y\in K\mid\exists\epsilon>0\colon x+\epsilon(x-y)\in K\right\}\,\text{.} (1)

The second one (Prop. 4.4) is Borwein and Goebel’s observation [7] that a point xKx\in K lies in rai(K)\operatorname{rai}(K) if and only if for all yKy\in K there is ϵ>0\epsilon>0 such that x+ϵ(xy)Kx+\epsilon(x-y)\in K. A point xx satisfying the latter proposition is called an internal point [11] or a “relatively absorbing point” [7] of KK.

Consequences of xrai(FK(x))x\in\operatorname{rai}(F_{K}(x)), xKx\in K, are organized roughly as follows. A review in Sec. 5 and new findings in Sec. 6 extend selected results from Secs. 6 and 18 of Rockafellar’s monograph [17]. Secs. 7 and 8 refer to Secs. 2–4 of Dubins’ paper [11] on infinite-dimensional convexity.

The methods of this paper are suitable to study the space of probability measures on a measurable space. The face generated by a probability measure is described in Sec. 9. The face generated by a Borel probability measure μ\mu on d\mathbb{R}^{d} is related to Csiszár and Matúš’ notion [10] of the convex core of μ\mu in Sec. 10. Sec. 11 studies measures on the set of natural numbers.

2. Main definitions

Throughout this paper, VV denotes a real vector space and KK a convex subset of VV​, unless stated otherwise. If xyx\neq y are distinct points of VV​, then

(x,y)\displaystyle(x,y) ={(1λ)x+λy:λ(0,1)}\displaystyle=\left\{(1-\lambda)x+\lambda y\colon\lambda\in(0,1)\right\}
resp.[x,y]\displaystyle\text{resp.}\quad[x,y] ={(1λ)x+λy:λ[0,1]}\displaystyle=\left\{(1-\lambda)x+\lambda y\colon\lambda\in[0,1]\right\}

is called the open segment resp. closed segment with endpoints x,yx,y. Each of the symbols (x,x)=[x,x](x,x)=[x,x] denotes the singleton {x}\{x\} containing xVx\in V​.

An extreme set [3] of KK is a subset EE of KK including the closed segment [x,y][x,y] for all points xyx\neq y in KK for which the open segment (x,y)(x,y) intersects EE. A point xKx\in K is an extreme point of KK if {x}\{x\} is an extreme set of KK. A face of KK is a convex extreme set of KK. Clearly, any union or intersection of extreme sets of KK is an extreme set of KK. Since any intersection of convex sets is convex, any intersection of faces of KK is a face of KK. In particular, the intersection of all faces containing a point xKx\in K is a face of KK, which is called the face of KK generated by xx, and which is denoted by FK(x)F_{K}(x).

The algebraic interior of KK is the set of all points xx in KK such that for every line gg in VV containing xx, the intersection gKg\cap K includes an open segment containing xx [5]. The convex set KK is algebraically open if it is equal to its algebraic interior. The relative algebraic interior rai(K)\operatorname{rai}(K) of KK is the set of all points xx in KK such that for every line gg in aff(K)\operatorname{aff}(K) containing xx, the intersection gKg\cap K includes an open segment containing xx [20]. The convex set KK is relative algebraically open if K=rai(K)K=\operatorname{rai}(K). A point xKx\in K is an internal point [11] of KK if for all yxy\neq x in KK there is ϵ>0\epsilon>0 such that x+ϵ(xy)Kx+\epsilon(x-y)\in K.

3. The finest locally convex topology

One of the innovations of this paper is that Thm. 4.5 is a theorem of Zermelo-Fraenkel set theory. However, the use of the relative interior in a topological vector space creates a new dependence on the axiom of choice by Rem. 3.2. Coro. 3.3 allows us to avoid this problem simply by dismissing topological vector spaces altogether, and relying on relative algebraic interiors of convex sets instead. This comes at the modest price that some theorems of topological vector spaces require a proof in this paper. For example, the theorem that the interior of a convex set is convex [8, II.14] would make Lemma 5.2 superfluous. The assertion that the interior of a set is open would essentially supersede Thm. 6.6 (up to questions regarding affine hulls).

A fundamental system of neighborhoods of a point xx in a topological space is any set 𝔖\mathfrak{S} of neighborhoods of xx such that for each neighborhood UU of xx there is a neighborhood W𝔖W\in\mathfrak{S} such that WUW\subset U. A topological real vector space is locally convex if there exists a fundamental system of neighborhoods of 0 that are convex sets [8, II.23].

It is known that a convex subset CVC\subset V is open for the finest locally convex topology 𝔗ω(V)\mathfrak{T}_{\omega}(V) on VV if and only if CC is algebraically open; a subset UVU\subset V is open for 𝔗ω(V)\mathfrak{T}_{\omega}(V) if and only if UU is a union of convex open subsets of VV [8, II.26]. Every nonempty affine subspace AVA\subset V is isomorphic to the vector space of translations AA={xyx,yA}A-A=\{x-y\mid x,y\in A\} by means of the affine isomorphism α:xxx0\alpha:x\mapsto x-x_{0} defined by some point x0Ax_{0}\in A. A topology is defined on AA for which a subset UAU\subset A is open if and only if α(U)\alpha(U) is open for 𝔗ω(AA)\mathfrak{T}_{\omega}(A-A). This topology does not depend on x0x_{0} and is denoted by 𝔗ω(A)\mathfrak{T}_{\omega}(A). The topology induced on AA by 𝔗ω(V)\mathfrak{T}_{\omega}(V) is denoted by 𝔗ω(V)|A\mathfrak{T}_{\omega}(V)|A. By definition, a subset UAU\subset A is open for 𝔗ω(V)|A\mathfrak{T}_{\omega}(V)|A if and only if there exists an open set WW for 𝔗ω(V)\mathfrak{T}_{\omega}(V) such that U=AWU=A\cap W. Let S,SVS,S^{\prime}\subset V be linear subspaces. Then SS^{\prime} is a complementary subspace of SS in VV if SS={0}S\cap S^{\prime}=\{0\} and if for all xVx\in V there is sSs\in S and sSs^{\prime}\in S^{\prime} such that x=s+sx=s+s^{\prime}.

Proposition 3.1.

If AA is an affine subspace of VV​, then 𝔗ω(A)=𝔗ω(V)|A\mathfrak{T}_{\omega}(A)=\mathfrak{T}_{\omega}(V)|A.

Proof.

Using the isomorphism AAAA\cong A-A, we assume that 0 lies in AA. Clearly, the induced topology 𝔗ω(V)|A\mathfrak{T}_{\omega}(V)|A is locally convex, which shows that 𝔗ω(A)\mathfrak{T}_{\omega}(A) is finer than 𝔗ω(V)|A\mathfrak{T}_{\omega}(V)|A. Conversely, let UU be a convex open subset of AA for the topology 𝔗ω(A)\mathfrak{T}_{\omega}(A). Let BB be a complementary subspace of AA in VV​. It is easy to see that the convex set U+B={u+b:uU,bB}U+B=\{u+b:u\in U,b\in B\} is open for 𝔗ω(V)\mathfrak{T}_{\omega}(V) and that A(U+B)=UA\cap(U+B)=U holds. This shows that 𝔗ω(V)|A\mathfrak{T}_{\omega}(V)|A is finer than 𝔗ω(A)\mathfrak{T}_{\omega}(A) and completes the proof. ∎

Remark 3.2.

The complementary subspace used in Prop. 3.1 exist by the axiom of choice. Conversely, the existence of a complementary subspace for all subspaces of all vector spaces over the reals (or over any other field) implies a weak version of the axiom of choice [6, Lemma 2] that is equivalent to the full axiom of choice in Zermelo-Fraenkel set theory [16, Thm. 9.1].

Corollary 3.3.

We have ri(K)=rai(K)\operatorname{ri}(K)=\operatorname{rai}(K) for the topology 𝔗ω(V)\mathfrak{T}_{\omega}(V).

Proof.

This follows immediately from Prop. 3.1. ∎

4. Every point lies in the relative algebraic interior of the face it generates

This section provides proofs for Alfsen’s formula (1) describing the face generated by a point, and Borwein and Goebel’s observation [7] that a point is an internal point of KK if and only if it lies in rai(K)\operatorname{rai}(K). An immediate corollary is Shirokov and the present author’s theorem [20] that every point lies in the relative algebraic interior of the face it generates. Let xKx\in K and define

SK(x)\displaystyle S_{K}(x) ={yKϵ>0:x+ϵ(xy)K},\displaystyle=\left\{y\in K\mid\exists\epsilon>0\colon x+\epsilon(x-y)\in K\right\}\,\text{,}
CK(x)\displaystyle C_{K}(x) ={yVϵ>0:xϵ(xy)K},\displaystyle=\left\{y\in V\mid\exists\epsilon>0\colon x-\epsilon(x-y)\in K\right\}\,\text{,}
AK(x)\displaystyle A_{K}(x) ={yVϵ>0:x±ϵ(xy)K}.\displaystyle=\left\{y\in V\mid\exists\epsilon>0\colon x\pm\epsilon(x-y)\in K\right\}\,\text{.}

a)Refer to captionb)Refer to caption

Figure 1. Configurations (in the plane) for Prop. 4.1. The set SK(x)S_{K}(x) is a) an extreme set and b) convex.
Proposition 4.1 (Alfsen).

For all xKx\in K we have FK(x)=SK(x)F_{K}(x)=S_{K}(x).

Proof.

Let yy be a point in SK(x)S_{K}(x) and yxy\neq x. Then there is ϵ>0\epsilon>0 such that xx lies in the open segment with endpoints yy and x+ϵ(xy)x+\epsilon(x-y). As FK(x)F_{K}(x) is an extreme set containing xx, it follows yFK(x)y\in F_{K}(x), which proves SK(x)FK(x)S_{K}(x)\subset F_{K}(x).

We finish the proof by showing that SK(x)S_{K}(x) is a face of KK. Thus, we consider distinct points aba\neq b in KK and a point yy in the open segment (a,b)(a,b), see Fig. 1. Let η(0,1)\eta\in(0,1) such that y=(1η)a+ηby=(1-\eta)a+\eta b. The coefficients for the following constructions are obtained from Menelaus’ theorem [4, 12].

We show that SK(x)S_{K}(x) is an extreme set, see Fig. 1 a). Assuming ySK(x)y\in S_{K}(x), there is ϵ>0\epsilon>0 such that y:=x+ϵ(xy)y^{\prime}:=x+\epsilon(x-y) lies in KK. Let ϵa=ϵ(1η)/(1+ϵη)\epsilon_{a}=\epsilon(1-\eta)/(1+\epsilon\eta) and ξa=ϵη/(1+ϵη)\xi_{a}=\epsilon\eta/(1+\epsilon\eta). Then

a:=x+ϵa(xa)=(1ξa)y+ξaba^{\prime}:=x+\epsilon_{a}(x-a)=(1-\xi_{a})y^{\prime}+\xi_{a}b

lies in KK, and hence aSK(x)a\in S_{K}(x). Similarly, bSK(x)b\in S_{K}(x).

We show that SK(x)S_{K}(x) is convex, see Fig. 1 b). We assume a,bSK(x)a,b\in S_{K}(x). Let ϵa,ϵb>0\epsilon_{a},\epsilon_{b}>0 such that a:=x+ϵa(xa)a^{\prime}:=x+\epsilon_{a}(x-a) and b:=x+ϵb(xb)b^{\prime}:=x+\epsilon_{b}(x-b) lie in KK; and let

ϵ=ϵaϵb(1η)ϵb+ηϵaandξ=ηϵa(1η)ϵb+ηϵa.\textstyle\epsilon=\frac{\epsilon_{a}\epsilon_{b}}{(1-\eta)\epsilon_{b}+\eta\epsilon_{a}}\quad\text{and}\quad\xi=\frac{\eta\epsilon_{a}}{(1-\eta)\epsilon_{b}+\eta\epsilon_{a}}\,\text{.}

Then

y:=x+ϵ(xy)=(1ξ)a+ξby^{\prime}:=x+\epsilon(x-y)=(1-\xi)a^{\prime}+\xi b^{\prime}

lies in KK, and hence ySK(x)y\in S_{K}(x). ∎

The union in Coro. 4.2 extends over all closed segments in KK, whose respective open segments contain xx, and the singleton {x}=(x,x)=[x,x]\{x\}=(x,x)=[x,x].

Corollary 4.2.

Every point xKx\in K is an internal point of FK(x)F_{K}(x) and we have FK(x)=y,zK,x(y,z)[y,z]F_{K}(x)=\bigcup_{y,z\in K,x\in(y,z)}[y,z].

Proof.

The first assertion follows from the definition of SK(x)S_{K}(x) and from Prop. 4.1, which proves FK(x)=SK(x)F_{K}(x)=S_{K}(x). Regarding the second assertion, the inclusion “\supset” holds because FK(x)F_{K}(x) is an extreme set of KK containing xx. The inclusion “\subset” is implied by a proof of SK(x)y,zK,x(y,z)[y,z]S_{K}(x)\subset\bigcup_{y,z\in K,x\in(y,z)}[y,z]. Let ySK(x)y\in S_{K}(x). Then there is ϵ>0\epsilon>0 such that z:=x+ϵ(xy)Kz:=x+\epsilon(x-y)\in K. Hence, x=1ϵ+1(ϵy+z)(y,z)x=\frac{1}{\epsilon+1}(\epsilon y+z)\in(y,z) and y[y,z]y\in[y,z] complete the proof. ∎

Note that CK(x)=cone(Kx)+xC_{K}(x)=\operatorname{cone}(K-x)+x holds for all xKx\in K, where cone(C)\operatorname{cone}(C) denotes the set λ0{λx:xC}\bigcup_{\lambda\geq 0}\{\lambda x:x\in C\} for every convex set CVC\subset V containing the origin. Borwein and Goebel [7, p. 2544] suggest that a preliminary step to Prop. 4.4 should be a proof that xx is an internal point of KK if and only if aff(K)=CK(x)\operatorname{aff}(K)=C_{K}(x) holds. Instead, we use the following Lemma 4.3.

Lemma 4.3.

Let xx be an internal point of KK. Then aff(K)AK(x)\operatorname{aff}(K)\subset A_{K}(x) holds.

Proof.

Let yxy\neq x be a point in aff(K)\operatorname{aff}(K). It suffices to find ϵ>0\epsilon>0 such that the two points x±ϵ(xy)x\pm\epsilon(x-y) are both contained in KK.

Since x,yaff(K)x,y\in\operatorname{aff}(K) and since yxy\neq x, there exist yiKy_{i}\in K and αi\alpha_{i}\in\mathbb{R}, i=1,,ni=1,\dots,n, not all numbers αi\alpha_{i} being zero, such that

y=x+iαiyiandiαi=0.\textstyle y=x+\sum_{i}\alpha_{i}y_{i}\quad\text{and}\quad\sum_{i}\alpha_{i}=0\,\text{.}

By the assumption that xx is an internal point of KK, there is ϵi>0\epsilon_{i}>0 such that

yi:=x+ϵi(xyi)y_{i}^{\prime}:=x+\epsilon_{i}(x-y_{i})

is contained in KK, i=1,,ni=1,\dots,n. Let α1=i=1n|αi|\|\alpha\|_{1}=\sum_{i=1}^{n}|\alpha_{i}| and let ϵ\epsilon be the minimum of miniϵi/α1\min_{i}\epsilon_{i}/\|\alpha\|_{1} and 1/α11/\|\alpha\|_{1}. Then ϵ\epsilon is strictly positive. We have

x±ϵ(xy)\displaystyle\textstyle x\pm\epsilon(x-y) =xϵiαiyi=x±ϵiαi(xyi)\displaystyle\textstyle=x\mp\epsilon\sum_{i}\alpha_{i}y_{i}=x\pm\epsilon\sum_{i}\alpha_{i}(x-y_{i})
=i|αi|α1(x±sgn(αi)ϵα1(xyi))zi:=.\displaystyle\textstyle=\sum_{i}\frac{|\alpha_{i}|}{\|\alpha\|_{1}}\underbrace{\left(x\pm\operatorname{sgn}(\alpha_{i})\epsilon\|\alpha\|_{1}(x-y_{i})\right)}_{z_{i}:=}\,\text{.}

If ±sgn(αi)=1\pm\operatorname{sgn}(\alpha_{i})=-1, then zi[yi,x]Kz_{i}\in[y_{i},x]\subset K holds because of ϵα11\epsilon\|\alpha\|_{1}\leq 1. If sgn(αi)=0\operatorname{sgn}(\alpha_{i})=0, then zi=xKz_{i}=x\in K. If ±sgn(αi)=+1\pm\operatorname{sgn}(\alpha_{i})=+1, then zi[x,yi]Kz_{i}\in[x,y_{i}^{\prime}]\subset K because ϵα1ϵi\epsilon\|\alpha\|_{1}\leq\epsilon_{i}. This shows that the points x±ϵ(xy)x\pm\epsilon(x-y) are convex combinations of points in KK, and therefore are themselves points in KK. ∎

Prop. 4.4 is mentioned on p. 2544 of [7].

Proposition 4.4 (Borwein-Goebel).

For all xKx\in K, the following assertions are equivalent.

  1. 1)

    The point xx is an internal point of KK.

  2. 2)

    We have xrai(K)x\in\operatorname{rai}(K).

  3. 3)

    We have AK(x)=CK(x)A_{K}(x)=C_{K}(x).

  4. 4)

    We have CK(x)=aff(K)C_{K}(x)=\operatorname{aff}(K).

Proof.

If xx is an internal point of KK, then Lemma 4.3 shows aff(K)AK(x)\operatorname{aff}(K)\subset A_{K}(x). This implies 2) by the definition of the relative algebraic interior. It also implies 3) and 4), because the inclusions AK(x)CK(x)aff(K)A_{K}(x)\subset C_{K}(x)\subset\operatorname{aff}(K) are trivial. As 2) \Rightarrow 1) is clear, it suffices to prove n) \Rightarrow 1) for n=3,4n=3,4.

Assume 3) is true and let yxy\neq x be a point of KK. Since KCK(x)K\subset C_{K}(x), the point yy lies in AK(x)A_{K}(x). This provides ϵ>0\epsilon>0 such that x+ϵ(xy)x+\epsilon(x-y) lies in KK. Hence, xx is an internal point of KK.

Assume 4) is true and let yxy\neq x be a point of KK. Then z:=2xyz:=2x-y lies in aff(K)\operatorname{aff}(K) and hence in CK(x)C_{K}(x). This provides ϵ>0\epsilon>0 such that

x+ϵ(xy)=xϵ(xz)Kx+\epsilon(x-y)=x-\epsilon(x-z)\in K

and shows that xx is an internal point of KK. ∎

This section’s main result is a novel proof for Thm. 2.3 in [20], which states the following.

Theorem 4.5 (W​.-Shirokov).

For all xKx\in K we have xrai(FK(x))x\in\operatorname{rai}(F_{K}(x)).

Proof.

The point xx is an internal point of FK(x)F_{K}(x) by Coro. 4.2. It lies in the relative algebraic interior of FK(x)F_{K}(x) by Prop. 4.4. ∎

Corollary 4.6.

For all xKx\in K we have aff(FK(x))=AK(x)\operatorname{aff}\left(F_{K}(x)\right)=A_{K}(x).

Proof.

Thm. 4.5 and Prop. 4.4 prove aff(FK(x))=AFK(x)(x)\operatorname{aff}\left(F_{K}(x)\right)=A_{F_{K}(x)}(x). The inclusion AFK(x)(x)AK(x)A_{F_{K}(x)}(x)\subset A_{K}(x) is clear. Conversely, if yAK(x)y\in A_{K}(x), then there is ϵ>0\epsilon>0 such that x±ϵ(xy)Kx\pm\epsilon(x-y)\in K. Since FK(x)F_{K}(x) is an extreme set of KK containing xx, this implies x±ϵ(xy)FK(x)x\pm\epsilon(x-y)\in F_{K}(x), and hence yAFK(x)(x)y\in A_{F_{K}(x)}(x). ∎

5. Review on the face generated by a point

Here we review some of our prior work from [20, Sec. 2]. Lemma 5.1.1 matches [17, Thm. 18.1]. See Lemma 2.1 in [20] for a proof.

Lemma 5.1.

Let CKC\subset K be a convex subset of KK, let EKE\subset K be an extreme set of KK, let FKF\subset K be a face of KK, and let xKx\in K be a point in KK. Then

  1. 1)

    rai(C)ECE\operatorname{rai}(C)\cap E\neq\emptyset\implies C\subset E,

  2. 2)

    xFFK(x)Fx\in F\iff F_{K}(x)\subset F,

  3. 3)

    xrai(F)F=FK(x)x\in\operatorname{rai}(F)\implies F=F_{K}(x).

Lemma 5.2 is proved in Lemma 2.2 in [20].

Lemma 5.2.

The complement Krai(K)K\setminus\operatorname{rai}(K) of the relative algebraic interior rai(K)\operatorname{rai}(K) is an extreme set of KK and rai(K)\operatorname{rai}(K) is a convex set.

Whereas Lemma 5.1 and Lemma 5.2 are rather easy to prove, the remainder of this section relies on Thm. 4.5.

Corollary 5.3.

Let SKS\subset K. The following assertions are equivalent.

  1. 1)

    SS is an extreme set of KK.

  2. 2)

    SS includes the face FK(x)F_{K}(x) of KK generated by any point xx in SS.

  3. 3)

    SS is the union of the faces FK(x)F_{K}(x) of KK generated by the points xx in SS.

Proof.

See Coro. 2.5 in [20]; a proof is provided for easy reference. 1) \Rightarrow 2) follows from Lemma 5.1.1 as xrai(FK(x))x\in\operatorname{rai}(F_{K}(x)) holds for all xSx\in S by Thm. 4.5. 2) \Rightarrow 3) follows from xFK(x)x\in F_{K}(x) for all xSx\in S. 3) \Rightarrow 1) follows directly from the definition of an extreme set. ∎

Coro. 5.4, and Thm. 6.8 below, match [17, Thm. 18.2]. Let

𝔘1\displaystyle\mathfrak{U}_{1} ={rai(FK(x)):xK},\displaystyle=\left\{\operatorname{rai}(F_{K}(x))\colon x\in K\right\}\,\text{,}
𝔘2\displaystyle\mathfrak{U}_{2} ={rai(F):F is a face of K}{}.\displaystyle=\left\{\operatorname{rai}(F)\colon\text{$F$ is a face of $K$}\right\}\setminus\{\emptyset\}\,\text{.}

We recall that a partition of KK is a family of nonempty subsets of KK whose elements are mutually disjoint and whose union is KK.

Corollary 5.4.

We have 𝔘1=𝔘2\mathfrak{U}_{1}=\mathfrak{U}_{2}, the family 𝔘2\mathfrak{U}_{2} is a partition of KK, and

{F is a face of K:rai(F)}𝔘2,Frai(F)\left\{\text{$F$ is a face of $K$}\colon\operatorname{rai}(F)\neq\emptyset\right\}\longrightarrow\mathfrak{U}_{2},\quad F\longmapsto\operatorname{rai}(F)

is a bijection.

Proof.

See Coro. 2.6 in [20]; a proof is provided for easy reference. The union of the family 𝔘1\mathfrak{U}_{1} covers KK as xrai(FK(x))x\in\operatorname{rai}(F_{K}(x)) by Thm. 4.5. Since 𝔘1𝔘2\mathfrak{U}_{1}\subset\mathfrak{U}_{2} is clear, proving that the elements of 𝔘2\mathfrak{U}_{2} are mutually disjoint implies that 𝔘1=𝔘2\mathfrak{U}_{1}=\mathfrak{U}_{2} and that 𝔘2\mathfrak{U}_{2} is a partition of KK. Let F,GF,G be faces of KK and let xrai(F)rai(G)x\in\operatorname{rai}(F)\cap\operatorname{rai}(G). Then Lemma 5.1.3 shows F=FK(x)=GF=F_{K}(x)=G. This also shows that the map in question is injective. Its surjectivity is clear. ∎

Remark 5.5 (Partitions of extreme sets).
  1. 1)

    Every extreme set EE of KK is the union of the family {rai(FK(x)):xE}\{\operatorname{rai}(F_{K}(x)):x\in E\} by Coro. 5.3.2 and Thm. 4.5. This family is a partition of EE by Coro. 5.4.

  2. 2)

    There exist subfamilies of the partition 𝔘1\mathfrak{U}_{1} in Coro. 5.4, whose union is not an extreme set of KK. An example is the subfamily having as its only element the open unit interval (0,1)(0,1) when K=[0,1]K=[0,1].

  3. 3)

    If the present concept of an extreme set is replaced with that underlying Dubins’ work [11], then the union of any subfamily of 𝔘1\mathfrak{U}_{1} will be an extreme set, cf. Thm. 8.1.

Coro. 5.6 and Coro. 5.7 are similar to [20, Coro. 2.7]. Proofs are provided for easy reference.

Corollary 5.6.

Let FF be a face of KK and let xKx\in K. The following statements are equivalent.

  1. 1)

    We have xrai(F)x\in\operatorname{rai}(F).

  2. 2)

    We have F=FK(x)F=F_{K}(x).

  3. 3)

    We have rai(F)=rai(FK(x))\operatorname{rai}(F)=\operatorname{rai}(F_{K}(x)).

Proof.

1) \Rightarrow 2) is Lemma 5.1.3. 2) \Rightarrow 3) is clear. 3) \Rightarrow 1) is implied by Thm. 4.5. ∎

Corollary 5.7.

Let x,yKx,y\in K. The following statements are equivalent.

  1. 1)

    We have xrai(F(y))x\in\operatorname{rai}(F(y)).

  2. 2)

    We have FK(y)=FK(x)F_{K}(y)=F_{K}(x).

  3. 3)

    We have rai(FK(y))=rai(FK(x))\operatorname{rai}(F_{K}(y))=\operatorname{rai}(F_{K}(x)).

  4. 4)

    We have xFK(y)x\in F_{K}(y) and yFK(x)y\in F_{K}(x).

Proof.

1) \Leftrightarrow 2) \Leftrightarrow 3) is the special case of Coro. 5.6 when FF is replaced with FK(y)F_{K}(y). 2) \Rightarrow 4) is clear, and 4) \Rightarrow 2) follows from Lemma 5.1.2. ∎

Prop. 5.8 complements [11, Thm. 4.3] but is not equivalent to it, as different concepts of a “face” are in use (see Sec. 8 below).

Proposition 5.8.

Let K,LVK,L\subset V be convex sets and let xKLx\in K\cap L. Then

  1. 1)

    FKL(x)F_{K\cap L}(x)

    == FK(x)FL(x)F_{K}(x)\cap F_{L}(x),

  2. 2)

    rai(FKL(x))\operatorname{rai}\big{(}F_{K\cap L}(x)\big{)}

    == rai(FK(x))rai(FL(x))\operatorname{rai}\big{(}F_{K}(x)\big{)}\cap\operatorname{rai}\big{(}F_{L}(x)\big{)},

  3. 3)

    aff(FKL(x))\operatorname{aff}\big{(}F_{K\cap L}(x)\big{)}

    == aff(FK(x))aff(FL(x))\operatorname{aff}\big{(}F_{K}(x)\big{)}\cap\operatorname{aff}\big{(}F_{L}(x)\big{)}.

Proof.

See [20, Prop. 2.13]. ∎

Corollary 5.9.

Let K,LVK,L\subset V be convex sets and let FF be a nonempty face of KLK\cap L with rai(F)\operatorname{rai}(F)\neq\emptyset. Then FF is the intersection of a face of KK and a face of LL. A sufficient condition for rai(F)\operatorname{rai}(F)\neq\emptyset is that FF have finite dimension.

Proof.

Let xrai(F)x\in\operatorname{rai}(F). Lemma 5.1.3 shows F=FKL(x)F=F_{K\cap L}(x) and Prop. 5.8.1 proves the first claim. If dim(F)<\dim(F)<\infty, then 𝔗ω(aff(F))\mathfrak{T}_{\omega}(\operatorname{aff}(F)) is the Euclidean topology on aff(F)\operatorname{aff}(F) [8, II.26]. Hence, the relative interior of FF is nonempty by Thm. 6.2 in [17]. Coro. 3.3 shows that the relative algebraic interior is nonempty, too. ∎

Is the assumption of rai(F)\operatorname{rai}(F)\neq\emptyset in Coro. 5.9 necessary?

Open Problem 5.10.

Are there convex sets KK and LL and a face FF of KLK\cap L that can not be written as the intersection of a face of KK and a face of LL?

The analogue of Prop. 5.8 for infinitely many convex sets is wrong.

Example 5.11.

We consider the open segment (ϵ,1+ϵ)(-\epsilon,1+\epsilon) for every ϵ>0\epsilon>0. The intersection ϵ>0(ϵ,1+ϵ)\bigcap_{\epsilon>0}(-\epsilon,1+\epsilon) is the closed unit interval [0,1][0,1]. The only face of (ϵ,1+ϵ)(-\epsilon,1+\epsilon) containing the extreme point 0 of [0,1][0,1] is (ϵ,1+ϵ)(-\epsilon,1+\epsilon) itself. Therefore, {0}\{0\} is not an intersection of faces of {(ϵ,1+ϵ)}ϵ>0\{(-\epsilon,1+\epsilon)\}_{\epsilon>0}, as such an intersection includes [0,1][0,1]. More examples are obtained by replacing some or all of the open segments (ϵ,1+ϵ)(-\epsilon,1+\epsilon) with closed segments [ϵ,1+ϵ][-\epsilon,1+\epsilon].

6. Novel results on faces and relative algebraic interiors

This section is inspired by properties of relative interiors of convex sets in finite dimensions [17]. Various insights into relative algebraic interiors are facilitated by Prop. 6.1 and Coro. 6.2. The unions in these statements extend over all open segments in KK containing xx, and the singleton {x}=(x,x)\{x\}=(x,x). Prop. 6.1 matches part of [17, Thm. 6.1].

Proposition 6.1.

For all xKx\in K we have rai(FK(x))=y,zK,x(y,z)(y,z)\operatorname{rai}(F_{K}(x))=\bigcup_{y,z\in K,x\in(y,z)}(y,z).

Proof.

The point xx lies in the left-hand side of the equation by Thm. 4.5 and in the right-hand side by definition. Let aKa\in K and axa\neq x. Then

arai(FK(x))Coro.5.7.4aFK(x) and xFK(a)Coro.4.2y,zK:x(a,y) and a(x,z)axy,zK:x,a(y,z)\begin{array}[]{rcl}a\in\operatorname{rai}(F_{K}(x))&\stackrel{{\scriptstyle\rm{Coro.\leavevmode\nobreak\ \ref{cor:char-ri-Fx}.4}}}{{\iff}}&\mbox{$a\in F_{K}(x)$ and $x\in F_{K}(a)$}\\ &\stackrel{{\scriptstyle\rm{Coro.\leavevmode\nobreak\ \ref{cor:facex}}}}{{\iff}}&\exists y,z\in K:\mbox{$x\in(a,y)$ and $a\in(x,z)$}\\ &\stackrel{{\scriptstyle a\neq x}}{{\iff}}&\exists y,z\in K:x,a\in(y,z)\end{array}

proves the claim. ∎

Corollary 6.2.

Let xKx\in K. Then rai(K)=y,zK,x(y,z)(y,z)\operatorname{rai}(K)=\bigcup_{y,z\in K,x\in(y,z)}(y,z) holds if and only if xrai(K)x\in\operatorname{rai}(K).

Proof.

Coro. 5.6 shows that xrai(K)x\in\operatorname{rai}(K) is equivalent to rai(K)=rai(FK(x))\operatorname{rai}(K)=\operatorname{rai}(F_{K}(x)). Now, the claim follows from Prop. 6.1. ∎

Coro. 6.3 matches part of [17, Thm. 6.2].

Corollary 6.3.

If rai(K)\operatorname{rai}(K)\neq\emptyset, then aff(rai(K))=aff(K)\operatorname{aff}(\operatorname{rai}(K))=\operatorname{aff}(K) holds.

Proof.

Let xrai(K)x\in\operatorname{rai}(K) and yaff(K)y\in\operatorname{aff}(K). By the definition of the relative algebraic interior, there is ϵ>0\epsilon>0 such that x±ϵ(xy)Kx\pm\epsilon(x-y)\in K. Coro. 6.2 shows y±=x±ϵ2(xy)rai(K)y_{\pm}=x\pm\frac{\epsilon}{2}(x-y)\in\operatorname{rai}(K), which implies

y=(121ϵ)y++(12+1ϵ)yaff(rai(K)).\textstyle y=\left(\frac{1}{2}-\frac{1}{\epsilon}\right)y_{+}+\left(\frac{1}{2}+\frac{1}{\epsilon}\right)y_{-}\in\operatorname{aff}(\operatorname{rai}(K))\,\text{.}

The opposite inclusion is obvious. ∎

Coro. 6.4 matches part of [17, Thm. 6.5].

Corollary 6.4.

Let K,LVK,L\subset V be convex sets and rai(K)rai(L)\operatorname{rai}(K)\cap\operatorname{rai}(L)\neq\emptyset. Then

rai(KL)=rai(K)rai(L)andaff(KL)=aff(K)aff(L).\operatorname{rai}(K\cap L)=\operatorname{rai}(K)\cap\operatorname{rai}(L)\quad\text{and}\quad\operatorname{aff}(K\cap L)=\operatorname{aff}(K)\cap\operatorname{aff}(L)\,\text{.}

The intersection of two relative algebraically open convex sets is relative algebraically open.

Proof.

Let xrai(K)rai(L)x\in\operatorname{rai}(K)\cap\operatorname{rai}(L). Then K=FK(x)K=F_{K}(x) and L=FL(x)L=F_{L}(x) by Lemma 5.1.3 and the first claim follows from Prop. 5.8. If KK and LL are relative algebraically open, then

KL=rai(K)rai(L)=rai(KL)K\cap L=\operatorname{rai}(K)\cap\operatorname{rai}(L)=\operatorname{rai}(K\cap L)

shows that the intersection KLK\cap L is relative algebraically open. ∎

Thm. 6.5 matches part of [17, Thm. 6.6].

Theorem 6.5.

Let rai(K)\operatorname{rai}(K)\neq\emptyset and let α:VW\alpha:V\to W be an affine map to a real vector space WW​. Then rai(α(K))=α(rai(K))\operatorname{rai}(\alpha(K))=\alpha(\operatorname{rai}(K)).

Proof.

Let xrai(K)x\in\operatorname{rai}(K). First, we show α(x)rai(α(K))\alpha(x)\in\operatorname{rai}(\alpha(K)). Let yα(x)y\neq\alpha(x) be a point in α(K)\alpha(K) and choose any yα|K1(y)y^{\prime}\in\alpha|_{K}^{-1}(y). Since xx is an internal point of KK, there exists ϵ>0\epsilon>0 such that x+ϵ(xy)x+\epsilon(x-y^{\prime}) lies in KK. Applying α\alpha shows that α(x)+ϵ(α(x)y)\alpha(x)+\epsilon(\alpha(x)-y) lies in α(K)\alpha(K), hence α(x)\alpha(x) is an internal point of α(K)\alpha(K). The implication 1) \Rightarrow 2) of Prop. 4.4 implies α(x)rai(α(K))\alpha(x)\in\operatorname{rai}(\alpha(K)). Second, Coro. 6.2 proves

α(rai(K))\displaystyle\alpha(\operatorname{rai}(K)) =α(y,zK,x(y,z)(y,z))\displaystyle=\textstyle\alpha\left(\bigcup_{y,z\in K,x\in(y,z)}(y,z)\right)
=y,zK,x(y,z)(α(y),α(z))\displaystyle=\textstyle\bigcup_{y,z\in K,x\in(y,z)}(\alpha(y),\alpha(z))
=y,zα(K),α(x)(y,z)(y,z).\displaystyle=\textstyle\bigcup_{y^{\prime},z^{\prime}\in\alpha(K),\alpha(x)\in(y^{\prime},z^{\prime})}(y^{\prime},z^{\prime})\,\text{.}

As α(x)rai(α(K))\alpha(x)\in\operatorname{rai}(\alpha(K)), the last expression of this equation equals rai(α(K))\operatorname{rai}(\alpha(K)), again by Coro. 6.2. ∎

Thm. 6.6 matches Problem 2 in [5, III.1.6].

Theorem 6.6.

The set rai(K)\operatorname{rai}(K) is a relative algebraically open convex set.

Proof.

The convexity of rai(K)\operatorname{rai}(K) is provided by Lemma 5.2. That rai(K)\operatorname{rai}(K) is relative algebraically open can be proved by showing for all xrai(K)x\in\operatorname{rai}(K) that xrai(rai(K))x\in\operatorname{rai}(\operatorname{rai}(K)), or equivalently that xx is an internal point of rai(K)\operatorname{rai}(K), as per the implication 1) \Rightarrow 2) of Prop. 4.4. Let yxy\neq x lie in rai(K)\operatorname{rai}(K). Coro. 6.2 provides a,bKa,b\in K such that xx and yy are in the open segment (a,b)(a,b). Relabel aa and bb, if necessary, such that x=(1λ)a+λbx=(1-\lambda)a+\lambda b and y=(1μ)a+μby=(1-\mu)a+\mu b for scalars 0<λ<μ<10<\lambda<\mu<1. Then 12(a+x)\frac{1}{2}(a+x) lies in rai(K)\operatorname{rai}(K), again by Coro. 6.2, and

x+λ2(μλ)(xy)=12(a+x)\textstyle x+\frac{\lambda}{2(\mu-\lambda)}(x-y)=\frac{1}{2}(a+x)

shows that xx is an internal point of rai(K)\operatorname{rai}(K). ∎

Coro. 6.7 matches part of [11, Thm. 2.1], see also Coro. 8.2 below.

Corollary 6.7.

For every xKx\in K, the set rai(FK(x))\operatorname{rai}(F_{K}(x)) is the greatest relative algebraically open convex subset (with respect to inclusion) of KK that contains xx.

Proof.

Thm. 6.6 shows that rai(FK(x))\operatorname{rai}(F_{K}(x)) is a relative algebraically open convex set, which contains xx by Thm. 4.5. Let CKC\subset K be relative algebraically open and convex, and let xCx\in C. For each yCy\in C, Coro. 6.2 provides a,bCa,b\in C such that x,y(a,b)x,y\in(a,b). Hence, Prop. 6.1 shows yrai(FK(x))y\in\operatorname{rai}(F_{K}(x)). ∎

Thm. 6.8 generalizes [17, Thm. 18.2].

Theorem 6.8.

Let KK\neq\emptyset. The family of maximal relative algebraically open convex subsets (with respect to inclusion) of KK is

𝔘={rai(FK(x)):xK}={rai(F):F is a face of K}{}.\mathfrak{U}=\left\{\operatorname{rai}(F_{K}(x))\colon x\in K\right\}=\left\{\operatorname{rai}(F)\colon\text{$F$ is a face of $K$}\right\}\setminus\{\emptyset\}\,\text{.}

The family 𝔘\mathfrak{U} is a partition of KK. Each nonempty relative algebraically open convex subset of KK is included in a unique element of 𝔘\mathfrak{U}.

Proof.

Let CC be a nonempty relative algebraically open convex subset of KK and let xx be a point in CC. Then Crai(FK(x))C\subset\operatorname{rai}(F_{K}(x)) holds by Coro. 6.7. The set CC can not be included in any other element of 𝔘\mathfrak{U} because 𝔘\mathfrak{U} is a partition by Coro. 5.4, which also proves the equality between the two descriptions of 𝔘\mathfrak{U}.

Let CC be a relative algebraically open convex subset of KK. We show that the elements of 𝔘\mathfrak{U} are maximal. Assume that CC includes rai(FK(x))\operatorname{rai}(F_{K}(x)) for some xKx\in K. Then CC contains xx by Thm. 4.5 and Crai(FK(x))C\subset\operatorname{rai}(F_{K}(x)) follows from Coro. 6.7. We show that KK has no other maximal relative algebraically open convex subsets. Assume that CC is maximal. As 𝔘\mathfrak{U}\neq\emptyset, the set CC contains some point xKx\in K. This implies C=rai(FK(x))C=\operatorname{rai}(F_{K}(x)) by Coro. 6.7. ∎

7. Novel results on generators of faces

This section is motivated by theorems in [11], but differs from them due to the nonequivalent concepts of a “face” (see Sec. 8).

Lemma 7.1.

Let KiK_{i} be a convex subset of VV​, let xirai(Ki)x_{i}\in\operatorname{rai}(K_{i}), and let λi>0\lambda_{i}>0, i=1,,ni=1,\ldots,n, such that λ1++λn=1\lambda_{1}+\ldots+\lambda_{n}=1. Then the point λ1x1++λnxn\lambda_{1}x_{1}+\ldots+\lambda_{n}x_{n} lies in the relative algebraic interior of the convex hull of K1KnK_{1}\cup\cdots\cup K_{n}.

Proof.

Let n=2n=2, let λ:=λ2(0,1)\lambda:=\lambda_{2}\in(0,1), x:=(1λ)x1+λx2x:=(1-\lambda)x_{1}+\lambda x_{2}, and let CC denote the convex hull of K1K2K_{1}\cup K_{2}. Below, we construct for every yCy\in C a number η>0\eta>0 such that x+η(xy)x+\eta(x-y) lies in CC. This shows that xx is an internal point of CC, and the implication 1) \Rightarrow 2) of Prop. 4.4 proves xrai(C)x\in\operatorname{rai}(C).

As yCy\in C, there is μ[0,1]\mu\in[0,1] and there are y1K1y_{1}\in K_{1} and y2K2y_{2}\in K_{2} such that y=(1μ)y1+μy2y=(1-\mu)y_{1}+\mu y_{2}. Since xix_{i} is an internal point of KiK_{i}, there exists ϵ>0\epsilon>0 such that zi:=xi+ϵ(xiyi)z_{i}:=x_{i}+\epsilon(x_{i}-y_{i}) lies in KiK_{i}, i=1,2i=1,2. We distinguish two cases. If μλ\mu\leq\lambda, then ν:=1μ+ϵ(λμ)\nu:=1-\mu+\epsilon(\lambda-\mu) and η:=ϵ(1λ)/ν\eta:=\epsilon(1-\lambda)/\nu are positive and

x+η(xy)=1ν[ϵ(λμ)y2+(1λ)(1μ)z1+λ(1μ)z2]K.\textstyle x+\eta(x-y)=\frac{1}{\nu}\left[\epsilon(\lambda-\mu)y_{2}+(1-\lambda)(1-\mu)z_{1}+\lambda(1-\mu)z_{2}\right]\in K\,\text{.}

If μλ\mu\geq\lambda, then ν:=μ+ϵ(μλ)\nu:=\mu+\epsilon(\mu-\lambda) and η:=ϵλ/ν\eta:=\epsilon\lambda/\nu are positive and

x+η(xy)=1ν[ϵ(μλ)y1+(1λ)μz1+λμz2]K.\textstyle x+\eta(x-y)=\frac{1}{\nu}\left[\epsilon(\mu-\lambda)y_{1}+(1-\lambda)\mu z_{1}+\lambda\mu z_{2}\right]\in K\,\text{.}

Induction extends the claim from n=2n=2 to all nn\in\mathbb{N}. ∎

Thm. 7.2 matches [11, (3.1)].

Theorem 7.2.

Let CKC\subset K be convex. Then xCFK(x)\bigcup_{x\in C}F_{K}(x) is a face of KK.

Proof.

The set E:=xCFK(x)E:=\bigcup_{x\in C}F_{K}(x) is a union of extreme sets, and hence an extreme set itself. We show that EE is convex. Let aiEa_{i}\in E and let ciCc_{i}\in C such that aiFK(ci)a_{i}\in F_{K}(c_{i}), i=1,2i=1,2. By Coro. 4.2, there is a point biKb_{i}\in K such that cic_{i} lies in the relative algebraic interior of [ai,bi][a_{i},b_{i}], i=1,2i=1,2. Lemma 7.1 shows that the point c:=(c1+c2)/2c:=(c_{1}+c_{2})/2 lies in the relative algebraic interior of the convex hull DD of {a1,a2,b1,b2}\{a_{1},a_{2},b_{1},b_{2}\}. Since CC is convex, it contains cc. Hence cc lies in EE, and Lemma 5.1.1 proves DED\subset E. It follows that [a1,a2]DE[a_{1},a_{2}]\subset D\subset E. ∎

We define the face generated by a subset SKS\subset K as the smallest face of KK containing SS. We denote this face by FK(S)F_{K}(S). Coro. 7.3 matches [11, (3.3)].

Corollary 7.3.

Let SKS\subset K. Then FK(S)=xCFK(x)F_{K}(S)=\bigcup_{x\in C}F_{K}(x), where CC is the convex hull of SS.

Proof.

The union F:=xCFK(x)F:=\bigcup_{x\in C}F_{K}(x) is a face of KK by Thm. 7.2. Let GG be any face containing SS. As GG is convex it includes CC and it also includes the face FK(x)F_{K}(x) for every xCx\in C by Lemma 5.1.2. This proves FGF\subset G. ∎

Corollary 7.4.

Let xiKx_{i}\in K, i=1,,ni=1,\ldots,n. Let S=i=1nFK(xi)S=\bigcup_{i=1}^{n}F_{K}(x_{i}). Let λi>0\lambda_{i}>0, i=1,,ni=1,\ldots,n, such that λ1++λn=1\lambda_{1}+\ldots+\lambda_{n}=1, and put x=λ1x1++λnxnx=\lambda_{1}x_{1}+\ldots+\lambda_{n}x_{n}. Then

FK(S)=FK({x1,,xn})=FK(x).F_{K}(S)=F_{K}(\{x_{1},\ldots,x_{n}\})=F_{K}(x)\,\text{.}
Proof.

As the inclusions FK(S)FK({x1,,xn})FK(x)F_{K}(S)\supset F_{K}(\{x_{1},\ldots,x_{n}\})\supset F_{K}(x) follow from Coro. 7.3, it suffices to prove FK(S)FK(x)F_{K}(S)\subset F_{K}(x). Lemma 7.1 shows that xx lies in the relative algebraic interior of the convex hull CC of {x1,,xn}\{x_{1},\ldots,x_{n}\}. Hence CC is included in FK(x)F_{K}(x) by Lemma 5.1.1. Lemma 5.1.2 then shows that FK(xi)F_{K}(x_{i}) is included in FK(x)F_{K}(x) for all i=1,,ni=1,\ldots,n, which proves FK(S)FK(x)F_{K}(S)\subset F_{K}(x). ∎

Coro. 7.5 matches [11, (4.7)].

Corollary 7.5.

Let K,LVK,L\subset V be convex sets and let KK be relative algebraically open. Then every extreme set resp. face of KLK\cap L is the intersection of KK and an extreme set resp. face of LL.

Proof.

Let EE be an extreme set of KLK\cap L. Coro. 5.3.3 and Prop. 5.8.1 show

E=xEFKL(x)=xE(FK(x)FL(x)).\textstyle E=\bigcup_{x\in E}F_{K\cap L}(x)=\textstyle\bigcup_{x\in E}\left(F_{K}(x)\cap F_{L}(x)\right)\,\text{.}

As KK is relative algebraically open, Lemma 5.1.3 implies FK(x)=KF_{K}(x)=K for all xKx\in K, hence

E=xE(KFL(x))=KxEFL(x).\textstyle E=\textstyle\bigcup_{x\in E}\left(K\cap F_{L}(x)\right)=\textstyle K\cap\bigcup_{x\in E}F_{L}(x)\,\text{.}

The set xEFL(x)\bigcup_{x\in E}F_{L}(x) is a union of extreme sets of LL and hence an extreme set of LL itself. If EE is a face of KLK\cap L, then EE is convex and Thm. 7.2 completes the proof. ∎

8. Dubins’ terminology

A d-extreme set of KK is a subset EE of KK including the open segment (x,y)(x,y) for all points xyx\neq y in KK for which (x,y)(x,y) intersects EE. A point xKx\in K is a d-extreme point of KK if {x}\{x\} is a d-extreme set of KK. A d-face of KK is a convex d-extreme set of KK. Clearly, any union or intersection of d-extreme sets of KK is a d-extreme set of KK. Hence, the intersection of all d-faces containing a point xKx\in K is a d-face of KK, which we call the d-face of KK generated by xx. This is the smallest d-face of KK containing xx.

Theorem 8.1.

A subset of KK is a d-extreme set of KK if and only if it is equal to the union xSrai(FK(x))\bigcup_{x\in S}\operatorname{rai}(F_{K}(x)) for some subset SKS\subset K. The d-face generated by xKx\in K is rai(FK(x))\operatorname{rai}(F_{K}(x)). A point in KK is a d-extreme point of KK if and only if it is an extreme point of KK.

Proof.

Let EE be a d-extreme set of KK and let xEx\in E. As xrai(FK(x))Ex\in\operatorname{rai}(F_{K}(x))\subset E holds by Prop. 6.1, we have E=xErai(FK(x))E=\bigcup_{x\in E}\operatorname{rai}(F_{K}(x)). Conversely, let SKS\subset K be any subset and assume that a point yy lies in the open segment (a,b)(a,b) with endpoints aba\neq b in KK and in the set rai(FK(x))\operatorname{rai}(F_{K}(x)) for some xSx\in S. The open segment (a,b)(a,b) is included in rai(FK(y))\operatorname{rai}(F_{K}(y)) by Coro. 6.7 and hence in rai(FK(x))\operatorname{rai}(F_{K}(x)), as rai(FK(y))=rai(FK(x))\operatorname{rai}(F_{K}(y))=\operatorname{rai}(F_{K}(x)) holds by Coro. 5.7.

The set rai(FK(x))\operatorname{rai}(F_{K}(x)) contains xx by Thm. 4.5 and is convex by Lemma 5.2. Hence, it is the smallest d-face containing xx by the first part of this theorem.

That “d-extreme point” and “extreme point” are equivalent terms is implied by the fact that a singleton cannot contain a segment no matter whether it is an open segment or a closed segment. ∎

Coro. 8.2 provides an alternative proof of [11, Thm. 2.1].

Corollary 8.2.

For all xKx\in K, the d-face of KK generated by xx is equal to rai(FK(x))\operatorname{rai}(F_{K}(x)) and equal to the greatest relative algebraically open convex subset of KK containing xx.

Proof.

Thm. 8.1 shows that rai(FK(x))\operatorname{rai}(F_{K}(x)) is the d-face of KK generated by xx. That rai(FK(x))\operatorname{rai}(F_{K}(x)) is the greatest relative algebraically open convex subset of KK containing xx is proved in Coro. 6.7. ∎

By definition [11], an elementary d-face of KK is a nonempty relative algebraically open d-face of KK.

Corollary 8.3.

Let KK\neq\emptyset. A subset of KK is an elementary d-face of KK if and only if it equals rai(FK(x))\operatorname{rai}(F_{K}(x)) for some xKx\in K.

Proof.

Let FF be a d-face of KK. By Thm. 8.1, FF is a union of a subfamily of 𝔘:={rai(FK(x)):xK}\mathfrak{U}:=\left\{\operatorname{rai}(F_{K}(x))\colon x\in K\right\}. If FF is an elementary d-face of KK, then FF cannot be a union of more than one element of 𝔘\mathfrak{U}, because the elements of 𝔘\mathfrak{U} are the maximal relative algebraically open convex subsets of KK by Thm. 6.8. ∎

Corollary 8.4.

Let KK\neq\emptyset. The following statements are equivalent.

  1. 1)

    KK is relative algebraically open.

  2. 2)

    KK has exactly two d-faces (which are \emptyset and KK).

  3. 3)

    KK has exactly two faces (which are \emptyset and KK).

Proof.

Assuming 1), rai(K)=K\operatorname{rai}(K)=K is nonempty. Hence, the partition of KK

𝔘={rai(FK(x)):xK}\mathfrak{U}=\left\{\operatorname{rai}(F_{K}(x))\colon x\in K\right\}

contains rai(K)\operatorname{rai}(K) as one of its elements by Coro. 5.4. Hence, \emptyset and KK are the only d-faces of KK by Thm. 8.1.

The statement 2) implies that KK has at most two faces, as every face of KK is a d-face of KK. Since KK\neq\emptyset, the convex set KK has exactly two faces.

Assuming 3), the convex set KK has only one nonempty face. Then {rai(K)}\{\operatorname{rai}(K)\} is a partition of KK by Coro. 5.4, which implies K=rai(K)K=\operatorname{rai}(K). ∎

9. Examples 1: Spaces of probability measures

Let 𝒫=𝒫(Ω,𝒜)\mathcal{P}=\mathcal{P}(\Omega,\mathcal{A}) denote the convex set of probability measures on a measurable space (Ω,𝒜)(\Omega,\mathcal{A}). A probability measure λ𝒫\lambda\in\mathcal{P} is absolutely continuous with respect to μ𝒫\mu\in\mathcal{P}, symbolically λμ\lambda\ll\mu, if every μ\mu-null set is a λ\lambda-null set. The measures are equivalent, λμ\lambda\equiv\mu, if λμ\lambda\ll\mu and μλ\mu\ll\lambda. If λμ\lambda\ll\mu, then we denote by dλdμ:Ω[0,)\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}:\Omega\to[0,\infty) the Radon-Nikodym derivative of λ\lambda with respect to μ\mu, which is a measurable function satisfying λ(A)=Adλdμdμ\lambda(A)=\int_{A}\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\,\operatorname{\mathrm{d}}\mu for all A𝒜A\in\mathcal{A}, see for example Halmos [14, Sec. 31].

Lemma 9.1.

For every μ𝒫\mu\in\mathcal{P}, the set {λ𝒫:λμ}\{\lambda\in\mathcal{P}:\lambda\ll\mu\} is a face of 𝒫\mathcal{P}.

Proof.

Let D(μ):={λ𝒫:λμ}D(\mu):=\{\lambda\in\mathcal{P}:\lambda\ll\mu\}. The set D(μ)D(\mu) is an extreme set of 𝒫\mathcal{P}. Let λ1,λ2𝒫\lambda_{1},\lambda_{2}\in\mathcal{P}, s(0,1)s\in(0,1), and λ:=(1s)λ1+sλ2\lambda:=(1-s)\lambda_{1}+s\lambda_{2} be contained in D(μ)D(\mu). If AA is a μ\mu-null set, then AA is a λ\lambda-null set and hence a λi\lambda_{i}-null set for i=1,2i=1,2.

The set D(μ)D(\mu) is convex, as (1s)λ1+sλ2(1-s)\lambda_{1}+s\lambda_{2} has the Radon-Nikodym derivative (1s)dλ1dμ+sdλ2dμ(1-s)\frac{\operatorname{\mathrm{d}}\lambda_{1}}{\operatorname{\mathrm{d}}\mu}+s\frac{\operatorname{\mathrm{d}}\lambda_{2}}{\operatorname{\mathrm{d}}\mu} with respect to μ\mu for all λ1,λ2D(μ)\lambda_{1},\lambda_{2}\in D(\mu) and s[0,1]s\in[0,1]. ∎

If μ𝒫\mu\in\mathcal{P}, then we say a proposition π(ω)\pi(\omega), ωΩ\omega\in\Omega, is true μ\mu-almost surely, which we abbreviate as μ\mu-a.s., if μ({ωΩ:π(ω) is false})=0\mu(\{\omega\in\Omega\colon\text{$\pi(\omega)$ is false}\})=0.

Theorem 9.2.

Let λ,μ𝒫\lambda,\mu\in\mathcal{P}. The following assertions are equivalent.

  1. 1)

    The measure λ\lambda lies in the face F𝒫(μ)F_{\mathcal{P}}(\mu) of 𝒫\mathcal{P} generated by μ\mu.

  2. 2)

    There is c[1,)c\in[1,\infty) such that λ(A)cμ(A)\lambda(A)\leq c\,\mu(A) holds for all A𝒜A\in\mathcal{A}.

  3. 3)

    We have λμ\lambda\ll\mu and there is c[1,)c\in[1,\infty) such that dλdμc\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\leq c holds μ\mu-a.s..

Proof.

Prop. 4.1 shows that a probability measure λ𝒫\lambda\in\mathcal{P} lies in F𝒫(μ)F_{\mathcal{P}}(\mu) if and only if there is ϵ>0\epsilon>0 such that μ+ϵ(μλ)𝒫\mu+\epsilon(\mu-\lambda)\in\mathcal{P}. The latter condition is equivalent to the nonnegativity of the set function μ+ϵ(μλ)\mu+\epsilon(\mu-\lambda), and hence to part 2) of the theorem. It remains to prove the equivalence 2) \Leftrightarrow 3).

If λμ\lambda\ll\mu and if there is c1c\geq 1 such that dλdμ(ω)c\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}(\omega)\leq c holds μ\mu-a.s., then part 2) follows (with the same constant cc),

λ(A)=Adλdμdμcμ(A)for all A𝒜.\textstyle\lambda(A)=\int_{A}\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\,\operatorname{\mathrm{d}}\mu\leq c\,\mu(A)\quad\text{for all $A\in\mathcal{A}$.}

Conversely, if λμ\lambda\ll\mu is false, then there is A𝒜A\in\mathcal{A} such that λ(A)>μ(A)=0\lambda(A)>\mu(A)=0, making part 2) impossible. If λμ\lambda\ll\mu is true but dλdμ\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu} is not bounded μ\mu-a.s., then for every c>0c>0 there is A𝒜A\in\mathcal{A} such that μ(A)>0\mu(A)>0 and dλdμ(ω)>c\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}(\omega)>c holds for all ωA\omega\in A. Then

λ(A)=Adλdμdμ>cμ(A)\textstyle\lambda(A)=\int_{A}\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\,\operatorname{\mathrm{d}}\mu>c\,\mu(A)

proves that part 2) fails. ∎

Dubins in [11] asserts that a probability measure λ𝒫\lambda\in\mathcal{P} is contained in the smallest d-face of 𝒫\mathcal{P} generated by μ𝒫\mu\in\mathcal{P} if and only if λμ\lambda\ll\mu and there exists c>0c>0, such that for all A𝒜A\in\mathcal{A} we have μ(A)cλ(A)c2μ(A)\mu(A)\leq c\,\lambda(A)\leq c^{2}\mu(A). Coro. 8.2 translates this assertion into Coro. 9.3.

Corollary 9.3 (Dubins).

Let λ,μ𝒫\lambda,\mu\in\mathcal{P}. The following statements are equivalent.

  1. 1)

    The measure λ\lambda lies in rai(F𝒫(μ))\operatorname{rai}(F_{\mathcal{P}}(\mu)).

  2. 2)

    There is c1c\geq 1 such that μ(A)/cλ(A)cμ(A)\mu(A)/c\leq\lambda(A)\leq c\,\mu(A) for all A𝒜A\in\mathcal{A}.

  3. 3)

    We have λμ\lambda\equiv\mu and there are c1,c2[1,)c_{1},c_{2}\in[1,\infty) such that dλdμc1\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\leq c_{1} holds μ\mu-a.s. and dμdλc2\frac{\operatorname{\mathrm{d}}\mu}{\operatorname{\mathrm{d}}\lambda}\leq c_{2} holds λ\lambda-a.s..

  4. 4)

    We have λμ\lambda\ll\mu and there is c[1,)c\in[1,\infty) such that 1cdλdμc\frac{1}{c}\leq\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\leq c holds μ\mu-a.s..

Proof.

Coro. 5.7 shows that λrai(F𝒫(μ))\lambda\in\operatorname{rai}(F_{\mathcal{P}}(\mu)) is equivalent to λF𝒫(μ)\lambda\in F_{\mathcal{P}}(\mu) and μF𝒫(λ)\mu\in F_{\mathcal{P}}(\lambda), so the equivalences 1) \Leftrightarrow 2) \Leftrightarrow 3) follow from those of Thm. 9.2.

Note that λ\lambda-a.s. is the same as μ\mu-a.s. if λμ\lambda\equiv\mu. Hence, 3) implies that dμdλdλdμ=1\frac{\operatorname{\mathrm{d}}\mu}{\operatorname{\mathrm{d}}\lambda}\cdot\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}=1 and hence dλdμ=(dμdλ)11/c2\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}=(\frac{\operatorname{\mathrm{d}}\mu}{\operatorname{\mathrm{d}}\lambda})^{-1}\geq 1/c_{2} holds μ\mu-a.s., see for example [14, Thm. A, p. 133]. Conversely, if λμ\lambda\ll\mu and if there is c[1,)c\in[1,\infty) such that 1cdλdμ\frac{1}{c}\leq\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu} holds μ\mu-a.s., then

λ(A)=Adλdμdμ1cμ(A)for all A𝒜\textstyle\lambda(A)=\int_{A}\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\,\operatorname{\mathrm{d}}\mu\geq\frac{1}{c}\,\mu(A)\quad\text{for all $A\in\mathcal{A}$}

implies that μλ\mu\equiv\lambda and that dμdλ=(dλdμ)1c\frac{\operatorname{\mathrm{d}}\mu}{\operatorname{\mathrm{d}}\lambda}=(\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu})^{-1}\leq c holds λ\lambda-a.s.. ∎

10. Examples 2: Convex cores

In a second example, we consider the Borel σ\sigma-algebra (d)\mathcal{B}(d) of d\mathbb{R}^{d}. The convex core cc(μ)\operatorname{cc}(\mu) of μ𝒫=𝒫(d,(d))\mu\in\mathcal{P}=\mathcal{P}(\mathbb{R}^{d},\mathcal{B}(d)) is the intersection of all convex sets C(d)C\in\mathcal{B}(d) of full measure μ(C)=μ(d)\mu(C)=\mu(\mathbb{R}^{d}). The convex core was introduced in [10] to extend exponential families in a natural way, such that information projections become properly defined. The mean of μ\mu is the integral m(μ)=dxdμ(x)dm(\mu)=\int_{\mathbb{R}^{d}}x\operatorname{\mathrm{d}}\mu(x)\in\mathbb{R}^{d}, provided that each coordinate function is μ\mu-integrable; otherwise, μ\mu does not have a mean.

Theorem 10.1 (Csiszár and Matúš).

Let μ𝒫(d,(d))\mu\in\mathcal{P}(\mathbb{R}^{d},\mathcal{B}(d)) have a mean. Then the convex core of μ\mu equals cc(μ)=m({λ𝒫:λμ})\operatorname{cc}(\mu)=m\left(\left\{\lambda\in\mathcal{P}\colon\lambda\ll\mu\right\}\right). Moreover, to each acc(μ)a\in\operatorname{cc}(\mu) there exists λ𝒫\lambda\in\mathcal{P} with λμ\lambda\ll\mu and mean m(λ)=am(\lambda)=a such that dλdμ\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu} is bounded μ\mu-a.s..

Thm. 10.1 is proved in Thm. 3 of [10]. We derive from it a description of the relative algebraic interior of the convex core.

Corollary 10.2.

Let μ𝒫(d,(d))\mu\in\mathcal{P}(\mathbb{R}^{d},\mathcal{B}(d)) have a mean. Then cc(μ)=m(F𝒫(μ))\operatorname{cc}(\mu)=m(F_{\mathcal{P}}(\mu)). The relative algebraic interior of cc(μ)\operatorname{cc}(\mu) is rai(cc(μ))=m(rai(F𝒫(μ)))\operatorname{rai}(\operatorname{cc}(\mu))=m(\operatorname{rai}(F_{\mathcal{P}}(\mu))), which equals

rai(cc(μ))=m({λ𝒫λμ,c(1,):1cdλdμc μ-a.s.}).\textstyle\operatorname{rai}(\operatorname{cc}(\mu))=m\left(\left\{\lambda\in\mathcal{P}\mid\lambda\equiv\mu,\exists c\in(1,\infty)\colon\frac{1}{c}\leq\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu}\leq c\mbox{\leavevmode\nobreak\ $\mu$-a.s.}\right\}\right)\,\text{.}
Proof.

Thm. 9.2 and Thm. 10.1 show that cc(μ)=m(F𝒫(μ))\operatorname{cc}(\mu)=m(F_{\mathcal{P}}(\mu)). As μ\mu lies in the relative algebraic interior of F𝒫(μ)F_{\mathcal{P}}(\mu) by Thm. 4.5, we obtain

rai(cc(μ))=m(rai(F𝒫(μ)))\operatorname{rai}(\operatorname{cc}(\mu))=m(\operatorname{rai}(F_{\mathcal{P}}(\mu)))

from Thm. 6.5. Coro. 9.3 completes the proof. ∎

The characterization of rai(cc(μ))\operatorname{rai}(\operatorname{cc}(\mu)) in Coro. 10.2 is somewhat stronger than that in Lemma 5 of [10], which ignores the lower bound 0<1cdλdμ0<\frac{1}{c}\leq\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\mu} μ\mu-a.s.. Lemma 5 of [10] also shows rai(cc(μ))=m({λ𝒫:λμ})\operatorname{rai}(\operatorname{cc}(\mu))=m\left(\left\{\lambda\in\mathcal{P}\colon\lambda\equiv\mu\right\}\right), which cannot be deduced from Thm. 10.1 with the methods developed in this paper, without the assistance of other methods.

11. Examples 3: Discrete probability measures

In a third example, we consider the discrete σ\sigma-algebra 22^{\mathbb{N}} of all subsets of \mathbb{N}. Consider the Banach space 1={x:x1<}\ell^{1}=\{x:\mathbb{N}\to\mathbb{C}\mid\|x\|_{1}<\infty\} of absolutely summable sequences endowed with the 1\ell^{1}-norm x1=n=1|x(n)|\|x\|_{1}=\sum_{n=1}^{\infty}|x(n)|. We study 𝒫=𝒫(,2)\mathcal{P}=\mathcal{P}(\mathbb{N},2^{\mathbb{N}}) in terms of the set of probability mass functions

Δ={p:n:p(n)0 and p1=1}.\textstyle\Delta_{\mathbb{N}}=\left\{p:\mathbb{N}\to\mathbb{R}\mid\forall n\in\mathbb{N}\colon p(n)\geq 0\mbox{\leavevmode\nobreak\ and\leavevmode\nobreak\ }\|p\|_{1}=1\right\}\,\text{.}

The map 𝒫Δ\mathcal{P}\to\Delta_{\mathbb{N}} that maps a probability measure μ𝒫\mu\in\mathcal{P} to its Radon-Nikodym derivative dμdν\frac{\operatorname{\mathrm{d}}\mu}{\operatorname{\mathrm{d}}\nu} with respect to the counting measure ν\nu, is an affine isomorphism. We endow 𝒫\mathcal{P} with the distance in variation

μλ:=2supA|μ(A)λ(A)|,λ,μ𝒫.\textstyle\|\mu-\lambda\|:=2\sup_{A\subset\mathbb{N}}|\mu(A)-\lambda(A)|\,\text{,}\quad\lambda,\mu\in\mathcal{P}\,\text{.}

Then 𝒫Δ\mathcal{P}\to\Delta_{\mathbb{N}} is an isometry, as μλ=dμdνdλdν1\|\mu-\lambda\|=\|\frac{\operatorname{\mathrm{d}}\mu}{\operatorname{\mathrm{d}}\nu}-\frac{\operatorname{\mathrm{d}}\lambda}{\operatorname{\mathrm{d}}\nu}\|_{1} holds [19, Sec. 3.9]. The support of pΔp\in\Delta_{\mathbb{N}} is spt(p):={n:p(n)>0}\operatorname{spt}(p):=\{n\in\mathbb{N}\colon p(n)>0\}. Lemma 11.1 is proved in [20, Lemma 2.12], and follows also from Thm. 9.2 and Coro. 9.3.

Lemma 11.1.

For every pΔp\in\Delta_{\mathbb{N}} we have

FΔ(p)\displaystyle\textstyle F_{\Delta_{\mathbb{N}}}(p) ={qΔspt(p):supnspt(p)q(n)/p(n)<},\displaystyle=\textstyle\left\{q\in\Delta_{\operatorname{spt}(p)}\colon\sup_{n\in\operatorname{spt}(p)}q(n)/p(n)<\infty\right\}\,\text{,}
rai(FΔ(p))\displaystyle\operatorname{rai}(F_{\Delta_{\mathbb{N}}}(p)) ={qFΔ(p):infnspt(p)q(n)/p(n)>0}.\displaystyle=\textstyle\left\{q\in F_{\Delta_{\mathbb{N}}}(p)\colon\inf_{n\in\operatorname{spt}(p)}q(n)/p(n)>0\right\}\,\text{.}
Example 11.2 (Faces with empty relative algebraic interiors).

For all subsets II\subset\mathbb{N} we define

ΔI:={pΔ:spt(p)I}andΔI,fin:={pΔI:ν(spt(p))<}.\Delta_{I}:=\{p\in\Delta_{\mathbb{N}}\colon\operatorname{spt}(p)\subset I\}\quad\text{and}\quad\Delta_{I,\text{fin}}:=\{p\in\Delta_{I}\colon\nu(\operatorname{spt}(p))<\infty\}\,\text{.}

The sets ΔI\Delta_{I} and ΔI,fin\Delta_{I,\text{fin}} are faces of Δ\Delta_{\mathbb{N}}, see [20, Example 2.10]. If II is an infinite set, then rai(ΔI)=rai(ΔI,fin)=\operatorname{rai}(\Delta_{I})=\operatorname{rai}(\Delta_{I,\text{fin}})=\emptyset holds by Lemma 5.1.3, as each of the faces ΔI\Delta_{I} and ΔI,fin\Delta_{I,\text{fin}} is strictly larger than the face generated by any of its points pp. This is clear if J:=spt(p)J:=\operatorname{spt}(p) is finite. Otherwise, if JJ is infinite, let pH(n):=p(n)/(rn+rn+1)p_{H}(n):=p(n)/(\sqrt{r_{n}}+\sqrt{r_{n+1}}) for nn\in\mathbb{N}, where rn:=mnp(m)r_{n}:=\sum_{m\geq n}p(m). Then pHΔJΔIp_{H}\in\Delta_{J}\subset\Delta_{I} but pHFΔ(p)p_{H}\not\in F_{\Delta_{\mathbb{N}}}(p) holds by Lemma 11.1.

In [20, Sec. 2], we raise the question as to whether Δ\Delta_{\mathbb{N}} has other faces with empty relative algebraic interiors, aside from those described in Example 11.2. Example 11.3 shows the answer is yes.

Example 11.3 (Another face with empty relative algebraic interior).

For every s>1s>1, let ps:p_{s}:\mathbb{N}\to\mathbb{R}, nζ(s)1nsn\mapsto\zeta(s)^{-1}\cdot n^{-s}, where ζ(s)=nns\zeta(s)=\sum_{n\in\mathbb{N}}n^{-s} is the Euler-Riemann zeta function. For all s,t>1s,t>1, Lemma 11.1 shows that the probability mass function psp_{s} is included in FΔ(pt)F_{\Delta_{\mathbb{N}}}(p_{t}) if and only if sts\geq t. Hence, Lemma 5.1.2 proves

FΔ(ps)FΔ(pt)st,s,t>1.F_{\Delta_{\mathbb{N}}}(p_{s})\subset F_{\Delta_{\mathbb{N}}}(p_{t})\iff s\geq t\,\text{,}\qquad s,t>1\,\text{.}

By Lemma 11.4 below, for every t1t\geq 1, the union

Ft:=s>tFΔ(ps)\textstyle F_{t}:=\bigcup_{s>t}F_{\Delta_{\mathbb{N}}}(p_{s})

is a face of Δ\Delta_{\mathbb{N}} and rai(Ft)=\operatorname{rai}(F_{t})=\emptyset. Let t>1t>1. Then FtF_{t} is included in FΔ(pt)F_{\Delta_{\mathbb{N}}}(p_{t}), which is properly included in Δ\Delta_{\mathbb{N}} by Ex. 11.2. We also have FtΔI,finF_{t}\neq\Delta_{I,\text{fin}} and FtΔIF_{t}\neq\Delta_{I} for all II\subset\mathbb{N}, because FtF_{t} contains the point pt+1p_{t+1} of support \mathbb{N}.

Lemma 11.4.

Let {xα}αA\{x_{\alpha}\}_{\alpha\in A} be a set of points in a convex set KK indexed by a totally ordered set AA that has no greatest element, such that αβ\alpha\leq\beta if and only if FK(xα)FK(xβ)F_{K}(x_{\alpha})\subset F_{K}(x_{\beta}) holds for all α,βA\alpha,\beta\in A. Then F=αAFK(xα)F=\bigcup_{\alpha\in A}F_{K}(x_{\alpha}) is a face of KK and rai(F)=\operatorname{rai}(F)=\emptyset.

Proof.

The set FF is convex. If a,bFa,b\in F, then aFK(xα)a\in F_{K}(x_{\alpha}) and bFK(xβ)b\in F_{K}(x_{\beta}) for some α,βA\alpha,\beta\in A. Both points xαx_{\alpha} and xβx_{\beta} lie in FK(xmax(α,β))F_{K}(x_{\max(\alpha,\beta)}), hence the closed segment with endpoints a,ba,b lies in FK(xmax(α,β))FF_{K}(x_{\max(\alpha,\beta)})\subset F. The set FF is an extreme set. If the open segment with endpoints aba\neq b in KK intersects FF, then it intersects FK(xα)F_{K}(x_{\alpha}) for some αA\alpha\in A. It follows that a,bFK(xα)Fa,b\in F_{K}(x_{\alpha})\subset F.

Assume there is arai(F)a\in\operatorname{rai}(F). Since aFa\in F, there is αA\alpha\in A with aFK(xα)a\in F_{K}(x_{\alpha}). Lemma 5.1.1 implies FFK(xα)F\subset F_{K}(x_{\alpha}), which shows that α\alpha is the greatest element of AA. This is excluded from the assumptions. ∎

Whereas Δ\Delta_{\mathbb{N}} is closed in the 1\ell^{1}-norm [21], the face FΔ(p)F_{\Delta_{\mathbb{N}}}(p) is not closed for any pΔp\in\Delta_{\mathbb{N}} of infinite support. Indeed, Lemma 11.5 shows that the closure of FΔ(p)F_{\Delta_{\mathbb{N}}}(p) is Δspt(p)\Delta_{\operatorname{spt}(p)} whereas FΔ(p)F_{\Delta_{\mathbb{N}}}(p) is strictly included in Δspt(p)\Delta_{\operatorname{spt}(p)} by Ex. 11.2. Let the function en1e_{n}\in\ell^{1} be defined by en(m)=1e_{n}(m)=1 if n=mn=m and en(m)=0e_{n}(m)=0 otherwise, m,nm,n\in\mathbb{N}.

Lemma 11.5.

The closure of any face FF of Δ\Delta_{\mathbb{N}} in the 1\ell^{1}-norm is ΔI(F)\Delta_{I(F)}, where I(F):=pFspt(p)I(F):=\bigcup_{p\in F}\operatorname{spt}(p). For every pΔp\in\Delta_{\mathbb{N}} we have I(FΔ(p))=spt(p)I(F_{\Delta_{\mathbb{N}}}(p))=\operatorname{spt}(p).

Proof.

First, the face ΔI,fin\Delta_{I,\text{fin}} of functions with finite support is dense in ΔI\Delta_{I} for all II\subset\mathbb{N}. To see this, let I=I=\mathbb{N} (without loss of generality) and let pΔp\in\Delta_{\mathbb{N}}. Then (pk)kΔ,fin(p_{k})_{k\in\mathbb{N}}\subset\Delta_{\mathbb{N},\text{fin}}, defined by

pk(n)={p(n)if n<k,mkp(m)if n=k,0else,k,n,p_{k}(n)=\left\{\begin{array}[]{ll}p(n)&\text{if $n<k$,}\\ \sum_{m\geq k}p(m)&\text{if $n=k$,}\\ 0&\text{else,}\end{array}\right.\quad k,n\in\mathbb{N}\,\text{,}

converges to pp, as ppk1=2m>kp(m)\|p-p_{k}\|_{1}=2\sum_{m>k}p(m) for all kk\in\mathbb{N}.

Second, if FF is a face of Δ\Delta_{\mathbb{N}} then ΔI(F),finFΔI(F)\Delta_{I(F),\text{fin}}\subset F\subset\Delta_{I(F)} holds. The right inclusion is obvious. To see the left inclusion, let nI(F)n\in I(F), and let pFp\in F such that nspt(p)n\in\operatorname{spt}(p). By Lemma 11.1, we have enFΔ(p)e_{n}\in F_{\Delta_{\mathbb{N}}}(p). Then enFe_{n}\in F follows from Lemma 5.1.2. This implies ΔI(F),finF\Delta_{I(F),\text{fin}}\subset F as FF is convex.

The preceding two arguments prove the first assertion. The second assertion is a special case of the first one and follows from Lemma 11.1. ∎

Lemma 11.5 shows that the norm closed faces of Δ\Delta_{\mathbb{N}} are in a one-to-one correspondence with the subsets of \mathbb{N}. This assertion is a special case of a more general property of von Neumann algebras [2]. The space 1\ell^{1} is the predual of the von Neumann algebra

={x:supn|x(n)|<}.\textstyle\ell^{\infty}=\{x:\mathbb{N}\to\mathbb{C}\mid\sup_{n\in\mathbb{N}}|x(n)|<\infty\}\,\text{.}

The set Δ1\Delta_{\mathbb{N}}\subset\ell^{1} is the normal state space of \ell^{\infty}, and the subsets II of \mathbb{N} are in a one-to-one correspondence with the projections in \ell^{\infty}, that is to say, functions {0,1}\mathbb{N}\to\{0,1\}. In a general von Neumann algebra, there is an order preserving isomorphism between the norm closed faces of the normal state space and the projections in the algebra [2, Thm. 3.35].

Acknowledgements

I thank Didier Henrion, Martin Kružík, Milan Korda, Milan Studený, and Tobias Fritz for inspiring discussions. This work was co-funded by the European Union under the project ROBOPROX (reg. no. CZ.02.01.01/00/22_008/0004590).

References

Stephan Weis

Czech Technical University in Prague

Faculty of Electrical Engineering

Karlovo náměstí 13

12000, Prague 2

Czech Republic

e-mail [email protected]