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

Linear and Sublinear Diversities

David Bryant and Paul Tupper
Abstract

Diversities are an extension of the concept of a metric space, where a non-negative value is assigned to every finite set of points, rather than just pairs. A general theory of diversities has been developed which exhibits many deep analogies to metric space theory but also veers off in new directions. Just as many of the most important aspects of metric space theory involve metrics defined on k\mathbb{R}^{k}, many applications of diversity theory require a specialized theory for diversities defined on k\mathbb{R}^{k}, as we develop here. We focus on two fundamental classes of diversities defined on k\mathbb{R}^{k}: those that are Minkowski linear and those that are Minkowski sublinear. Many well-known functions in convex analysis belong to these classes, including diameter, circumradius and mean width. We derive surprising characterizations of these classes, and establish elegant connections between them. Motivated by classical results in metric geometry, and connections with combinatorial optimization, we then examine embeddability of finite diversities into k\mathbb{R}^{k}. We prove that a finite diversity can be embedded into a linear diversity exactly when it has negative type and that it can be embedded into a sublinear diversity exactly when it corresponds to a generalized circumradius.

1 Introduction

A diversity [6] is a pair (X,δ)(X,\delta) where XX is a set and δ\delta is a non-negative function defined on finite subsets of XX satisfying

(D1) δ(A)0\delta(A)\geq 0 and δ(A)=0\delta(A)=0 if and only if |A|1|A|\leq 1,
(D2) δ(AC)δ(AB)+δ(BC)\delta(A\cup C)\leq\delta(A\cup B)+\delta(B\cup C) whenever BB\neq\emptyset.

As such, diversities are set-based analogues of metric spaces, and in fact the restriction of a diversity to pairs is a metric space [6].

Properties (D1) and (D2) are equivalent to (D1) together with monotonicity

(D3) δ(A)δ(B)\delta(A)\leq\delta(B) whenever ABA\subseteq B

and subadditivity on intersecting sets

(D4) δ(AB)δ(A)+δ(B)\delta(A\cup B)\leq\delta(A)+\delta(B) when ABA\cap B\neq\emptyset.

We say (X,δ)(X,\delta) is a semidiversity if (D1) is relaxed to

(D1) δ(A)0\delta(A)\geq 0 and δ(A)=0\delta(A)=0 if |A|1|A|\leq 1.

That is, sets with two or more points may have zero diversity. This terminology is analogous to at least some of the definitions of semimetrics.

Many well-known set functions are diversities: the diameter of a set; the length of a connecting Steiner tree; the circumradius; the length of a minimal traveling salesperson tour; the mean width; the size of a smallest enclosing zonotope. Two set functions which fail to be diversities are genetic diversity

π(A)=(|A|2)1a,bAd(a,b),\pi(A)=\mbox{$\binom{|A|}{2}^{-1}$}\sum_{a,b\in A}d(a,b),

which is not monotonic (D3), and volume of convex hull, which fails (D2) and (D4).

There are broad classes of diversities just like there are broad classes of metrics. The 1\ell_{1} metrics have the form

d1(a,b)=i|aibi|,d_{1}(a,b)=\sum_{i}|a_{i}-b_{i}|,

while 1\ell_{1} diversities [7] have the form

δ1(A)=imaxa,bA|aibi|.\delta_{1}(A)=\sum_{i}\max_{a,b\in A}|a_{i}-b_{i}|.

Negative-type metrics satisfy

a,bxaxbd(a,b)0\sum_{a,b}x_{a}x_{b}\,d(a,b)\leq 0

for all zero-sum vectors xx while negative type diversities [27] satisfy

A,BxAxBδ(AB)0\sum_{A,B}x_{A}x_{B}\,\delta(A\cup B)\leq 0

for all zero-sum vectors xx with x=0x_{\emptyset}=0.

The theory of diversities sometimes runs in parallel with that of metric spaces and other times veers off in new directions. In the first paper on diversities [6] we explored how concepts of hyperconvexity, injectivity and the tight span extended, and to an extent enriched, the analogous metric concepts. In [7] we showed that the ‘geometry of graphs’ [18] linking metric embeddings to approximation algorithms on graphs has a parallel ‘geometry of hypergraphs’ linking diversity embeddings to approximation algorithms on hypergraphs. Jozefiak and Shephard [17] use this approach to obtain the best known approximation algorithms for several hypergraph optimization problems.

Diversities turn out to be an exemplary class of metric structures, exhibiting fascinating connections with model theory and Urysohn’s universal space [4, 5, 15, 14]. Other directions that have been pursued are a diversity analogue of ultrametric and normed spaces [13, 20, 12] and new diversity-based results in fixed point theory [9, 22].

Our focus here is on the intersection of diversity theory, geometry and convex analysis. Recall that the Minkowski sum of two subsets A,BkA,B\subseteq\mathbb{R}^{k} is given by A+B={a+b:aA,bB}.A+B=\{a+b:a\in A,b\in B\}. We investigate diversities defined on k\mathbb{R}^{k} which are (Minkowski) linear [24]

(D5) δ(λA)=λδ(A) and δ(A+B)=δ(A)+δ(B)\delta(\lambda A)=\lambda\delta(A)\mbox{ and }\delta(A+B)=\delta(A)+\delta(B)

and those which are (Minkowski) sublinear

(D6) δ(λA)=λδ(A) and δ(A+B)δ(A)+δ(B),\delta(\lambda A)=\lambda\delta(A)\mbox{ and }\delta(A+B)\leq\delta(A)+\delta(B),

for λ0\lambda\geq 0 and A,BA,B nonempty finite subsets of k\mathbb{R}^{k}. Many familiar diversities defined on k\mathbb{R}^{k} are Minkowski linear or sublinear (see below). We explore their properties and characterization.

As per usual, we make repeated use of support functions when dealing with convex bodies and functions defined on them. The support function of a nonempty bounded set AA is defined

hA:k:xsup{ax:aA}.h_{A}:\mathbb{R}^{k}\rightarrow\mathbb{R}:x\mapsto\sup\{a\cdot x:a\in A\}.

Here axa\cdot x denotes the usual dot product in k\mathbb{R}^{k},

ax=i=1kaixi.a\cdot x=\sum_{i=1}^{k}a_{i}x_{i}.

We note that a set has the same support function as both its closure and convex hull.

We make use of the following properties of the support function, see [24, Chapter 1] for further details.

  1. 1.

    hA+B=hA+hBh_{A+B}=h_{A}+h_{B} and hλA=λhAh_{\lambda A}=\lambda h_{A} for for non-empty, bounded A,BA,B and λ0.\lambda\geq 0.

  2. 2.

    If A,BA,B are nonempty convex compact sets then ABA\subseteq B if and only hA(x)hB(x)h_{A}(x)\leq h_{B}(x) for all xkx\in\mathbb{R}^{k}

  3. 3.

    A function h:kh:\mathbb{R}^{k}\rightarrow\mathbb{R} is the support function for some bounded nonempty set if and only if h(x+y)h(x)+h(y)h(x+y)\leq h(x)+h(y) and h(λx)=λh(x)h(\lambda x)=\lambda h(x) for all x,ykx,y\in\mathbb{R}^{k} and λ0\lambda\geq 0 (that is, hh is sublinear).

We often consider support functions restricted to 𝕊k1\mathbb{S}^{k-1}, the unit sphere in k\mathbb{R}^{k}, noting that a support function is determined everywhere by its values on 𝕊k1\mathbb{S}^{k-1}. We note that the support function restricted to 𝕊k1\mathbb{S}^{k-1} of a nonempty set is bounded if and only if the set is bounded.

Our main results for diversities and semidiversities (k,δ)(\mathbb{R}^{k},\delta) are:

  1. 1.

    (Theorem 5) Linear diversities and semidiversities are exactly those which can be written in the form

    δ(A)=𝕊k1hA(x)dν(x)\delta(A)=\int_{\mathbb{S}^{k-1}}h_{A}(x)\mathrm{d}\nu(x)

    for a Borel measure ν\nu on the sphere 𝕊k1\mathbb{S}^{k-1} satisfying

    𝕊k1xdν(x)=0.\int_{\mathbb{S}^{k-1}}x\,\mathrm{d}\nu(x)=0. (1)
  2. 2.

    (Theorem 7) The extremal linear semidiversities are those where the support of ν\nu is a finite, affinely independent set, which in turn correspond to a generalized circumradius (a Minkowski semidiversity) based on the simplex.

  3. 3.

    (Theorem 8) A diversity or semidiversity is sublinear if and only if it is the maximum of linear semidiversities (just like a function is convex if and only if it is the maximum of linear functions).

We then shift to studying the embeddings of finite diversities into linear or sublinear diversities. Questions regarding embeddings and approximate embeddings of metrics in normed spaces are central to metric geometry and its applications. Consider, for example, Menger’s characterizations of when a metric can be embedded in Euclidean space, or the vast literature applying metric embeddings to combinatorial optimizations (reviewed in [19, 8] and [16]).

For finite diversities (X,δ)(X,\delta) we show:

  1. 1.

    (Theorem 11) A finite diversity can be embedded in a linear diversity if and only if it has negative type, meaning that

    A,BxAxBδ(AB)0\sum_{A,B}x_{A}x_{B}\delta(A\cup B)\leq 0

    for all vectors xx with zero sum and x=0x_{\emptyset}=0.

  2. 2.

    (Theorem 12) A finite diversity can be embedded as a sublinear diversity if and only if it can be embedded in a Minkowski diversity (that is, a generalized circumradius) if and only if it is the maximum of a collection of negative type diversities.

2 Linear and sublinear diversities

In this section we establish basic properties and characterizations for linear and sublinear diversities.

2.1 Examples of Linear and Sublinear Diversities

We start with examples of diversities which are linear or sublinear. Note that for all diversities (X,δ)(X,\delta) we have δ()=0\delta(\emptyset)=0, even if that is not stated explicitly below.

  1. 1.

    Let \|\cdot\| be any norm on k\mathbb{R}^{k}. The diameter diversity is given by

    δ(A)=maxa,bAab\delta(A)=\max_{a,b\in A}\|a-b\|

    for finite AkA\subseteq\mathbb{R}^{k}. The diameter diversity is sublinear [24, pg 49].

  2. 2.

    The 1\ell_{1} diversity (k,δ1)(\mathbb{R}^{k},\delta_{1}) is

    δ1(A)=i=1kmaxa,bA(aibi).\delta_{1}(A)=\sum_{i=1}^{k}\max_{a,b\in A}(a_{i}-b_{i}).

    for finite AkA\subseteq\mathbb{R}^{k} [7]. For finite A,BA,B and λ0\lambda\geq 0 we have

    δ1(λA+B)\displaystyle\delta_{1}(\lambda A+B) =i=1kmax{((λa+b)i(λa+b)i):a,aA,b,bB}\displaystyle=\sum_{i=1}^{k}\max\left\{\left((\lambda a+b)_{i}-(\lambda a^{\prime}+b^{\prime})_{i}\right):a,a^{\prime}\in A,\,b,b^{\prime}\in B\right\}
    =i=1kmax{λ(aiai)+(bibi):a,aA,b,bB}\displaystyle=\sum_{i=1}^{k}\max\{\lambda(a_{i}-a_{i}^{\prime})+(b_{i}-b_{i}^{\prime}):a,a^{\prime}\in A,\,b,b^{\prime}\in B\}
    =λδ1(A)+δ1(B).\displaystyle=\lambda\delta_{1}(A)+\delta_{1}(B).

    so (k,δ1)(\mathbb{R}^{k},\delta_{1}) is a linear diversity.

  3. 3.

    The circumradius of finite AkA\subset\mathbb{R}^{k} with respect to the unit ball \mathcal{B} is

    δ(A)=min{λ0:Aλ+x for some xk}.\delta(A)=\min\{\lambda\geq 0:A\subseteq\lambda\mathcal{B}+x\mbox{ for some $x\in\mathbb{R}^{k}$}\}.

    More generally, the Minkowski diversity (k,δK)(\mathbb{R}^{k},\delta_{K}) with kernel KK is equal to the generalized circumradius

    δK(A)=inf{λ0:AλK+x for some xk},\delta_{K}(A)=\inf\{\lambda\geq 0:A\subseteq\lambda K+x\mbox{ for some }x\in\mathbb{R}^{k}\},

    for finite AkA\subseteq\mathbb{R}^{k}. Minkowski diversities are sublinear [3] but are not, in general, linear. For example, consider the circumradius diversity (2,δ)(\mathbb{R}^{2},\delta). If A={(0,0),(1,0)}A=\{(0,0),(1,0)\} and B={(0,0),(0,1)}B=\{(0,0),(0,1)\} then δ(A+B)<δ(A)+δ(B)\delta(A+B)<\delta(A)+\delta(B).

    We assume that KK is closed, convex and has non-empty interior. We have elsewhere required that the kernel KK be bounded, however in this paper we will not require this. Note that if KK is an unbounded, (k,δK)(\mathbb{R}^{k},\delta_{K}) is a semidiversity rather than a diversity.

  4. 4.

    The mean-width diversity (k,δw)(\mathbb{R}^{k},\delta_{w}) is

    δw(A)=2ωk𝕊k1hA(x)𝑑ν(x)\delta_{w}(A)=\frac{2}{\omega_{k}}\int_{\mathbb{S}^{k-1}}h_{A}(x)\,d\nu(x)

    where ν(x)\nu(x) is the (uniform) Haar measure on the sphere and ωk=𝕊k1𝑑ν(x)\omega_{k}=\int_{\mathbb{S}^{k-1}}\,d\nu(x). Equivalently, δw(A)\delta_{w}(A) is the mean-width of the convex hull of AA. Mean-width diversities are linear [24, pg 50].

    Let wA(x)=max{x(ab):a,bA}w_{A}(x)=\max\{x\cdot(a-b):a,b\in A\} denote the width of AA in direction xx, so wA(x)=hAA(x)=hA(x)+hA(x)w_{A}(x)=h_{A-A}(x)=h_{A}(x)+h_{A}(-x). Then

    δw(A)=1ωk𝕊k1wA(x)𝑑ν(x).\delta_{w}(A)=\frac{1}{\omega_{k}}\int_{\mathbb{S}^{k-1}}w_{A}(x)\,d\nu(x).

    For 1p<1\leq p<\infty we define

    δw(p)(A)=1ωk[𝕊k1|wA(x)|p𝑑ν(x)]1/p.\delta^{(p)}_{w}(A)=\frac{1}{\omega_{k}}\left[\int_{\mathbb{S}^{k-1}}|w_{A}(x)|^{p}\,d\nu(x)\right]^{1/p}.

    That this is a sublinear diversity follows from the Minkowski inequality. See [13] Proposition 2.4, or [2] Proposition 10 in the case that p=2p=2.

  5. 5.

    A zonotope ZZ is a Minkowski sum of line segments and the length (Z)\ell(Z) of the zonotope equals the sum of the length of the line segments. We define the zonotope diversity (k,δz)(\mathbb{R}^{k},\delta_{z}) where δz(A)\delta_{z}(A) is the minimum length of a zonotope containing AA. We show that zonotope diversities are sublinear in Proposition 2.

In a Euclidean space k\mathbb{R}^{k}, any non-negative linear combination of sublinear semidiversities is sublinear, and any non-negative linear combination of linear semidiversities is linear. Hence the set of sublinear semidiversities forms a convex cone, as does the set of linear semidiversities.

2.2 Properties of linear and sublinear diversities

We establish some basic properties of sublinear diversities (and hence of linear diversities). This includes the continuous extension of sublinear diversities from finite sets to bounded sets.

Proposition 1.

Let δ\delta be a function on finite subsets of k\mathbb{R}^{k} which satisfies (D1), monotonicity (D3) and sublinearity (D6).

  1. 1.

    δ\delta is translation invariant: δ(A+x)=δ(A)\delta(A+x)=\delta(A) for all finite AkA\subseteq\mathbb{R}^{k} and xx\in\mathbb{R}.

  2. 2.

    (k,δ)(\mathbb{R}^{k},\delta) is a diversity.

  3. 3.

    If conv(A)=conv(B)\mathrm{conv}(A)=\mathrm{conv}(B) then δ(A)=δ(B)\delta(A)=\delta(B).

  4. 4.

    The map N:kN:\mathbb{R}^{k}\rightarrow\mathbb{R} given by N(x)=δ({0,x})N(x)=\delta(\{0,x\}) is a norm on k\mathbb{R}^{k}.

  5. 5.

    For all finite AkA\subseteq\mathbb{R}^{k} with |A|>2|A|>2 we have

    δ(A) |A|1|A|(|A|2)aAδ(A{a}).\delta(A)\leq\mbox{ $\frac{|A|-1}{|A|(|A|-2)}$}\sum_{a\in A}\delta(A\setminus\{a\}).

If δ\delta satisfies (D1) rather than (D1) then 1-5 still hold except that (k,δ)(\mathbb{R}^{k},\delta) is a semidiversity and NN is a seminorm.

Proof.
  1. 1.

    By sublinearity (D6) and (D1), we have

    δ(A+x)\displaystyle\delta(A+x) δ(A)+δ({x})=δ(A), and\displaystyle\leq\delta(A)+\delta(\{x\})=\delta(A),\mbox{ and }
    δ(A)\displaystyle\delta(A) δ(A+x)+δ({x})=δ(A+x).\displaystyle\leq\delta(A+x)+\delta(\{-x\})=\delta(A+x).
  2. 2.

    As δ\delta is monotonic and δ()=0\delta(\emptyset)=0, δ\delta is non-negative, and by part 1. δ\delta is translation invariant. We show that (X,δ)(X,\delta) satisfies (D4). Suppose that xABx\in A\cap B. Then 0(Ax)(Bx)0\in(A-x)\cap(B-x) and so (Ax)(Bx)(Ax)+(Bx)(A-x)\cup(B-x)\subseteq(A-x)+(B-x) and

    δ(AB)\displaystyle\delta(A\cup B) =δ((Ax)(Bx))\displaystyle=\delta\Big{(}(A-x)\cup(B-x)\Big{)}
    δ((Ax)+(Bx))\displaystyle\leq\delta\Big{(}(A-x)+(B-x)\Big{)}
    δ(Ax)+δ(Bx)\displaystyle\leq\delta(A-x)+\delta(B-x)
    =δ(A)+δ(B).\displaystyle=\delta(A)+\delta(B).

    Hence (k,δ)(\mathbb{R}^{k},\delta) satisfies (D1), (D3) and (D4).

  3. 3.

    Proposition 2.2b in [3].

  4. 4.

    By (D5) we have N(x+y)=δ({0,x+y})δ({0,x})+δ({0,y})=N(x)+N(y)N(x+y)=\delta(\{0,x+y\})\leq\delta(\{0,x\})+\delta(\{0,y\})=N(x)+N(y). If λ>0\lambda>0 then N(λx)=δ({0,λx})=λδ(0,x)=|λ|N(x)N(\lambda x)=\delta(\{0,\lambda x\})=\lambda\delta(0,x)=|\lambda|N(x), while if λ<0\lambda<0 we have

    N(λx)=δ({λx,0})=δ({0,λx})=|λ|N(x).N(\lambda x)=\delta(\{\lambda x,0\})=\delta(\{0,-\lambda x\})=|\lambda|N(x).

    Also, N(0)=δ({0})=0N(0)=\delta(\{0\})=0 if and only if δ({x,0})=0\delta(\{x,0\})=0 if and only if x=0x=0.

  5. 5.

    This follows from sublinearity and the following observation; see the proof of [1, Theorem 4.1]. By sublinearity we may assume aAa=0\sum_{a\in A}a=0. So for each aAa\in A

    1|A|1a=1|A|1aaaconv(A{a}).-\frac{1}{|A|-1}a=\frac{1}{|A|-1}\sum_{a^{\prime}\neq a}a^{\prime}\in\mathrm{conv}(A\setminus\{a\}).

    We also have aconv(Aa)a\in\mathrm{conv}(A\setminus a^{\prime}) for aaa\neq a^{\prime}. So for all aAa\in A

    (|A|2)|A||A|1a=(|A|1)a1|A|1abAconv(A{b}).\frac{(|A|-2)|A|}{|A|-1}a=(|A|-1)a-\frac{1}{|A|-1}a\in\sum_{b\in A}\mathrm{conv}(A\setminus\{b\}).

    This gives

    A|A|1|A|(|A|2)aAconv(A{a})A\subseteq\frac{|A|-1}{|A|(|A|-2)}\sum_{a\in A}\mathrm{conv}(A\setminus\{a\})

    and applying sublinearity gives the result.

It is now straightforward to show that the zonotope diversity introduced above is in fact a sublinear diversity.

Proposition 2.

The zonotope diversity (k,δz)(\mathbb{R}^{k},\delta_{z}) is a sublinear diversity.

Proof.

Recall that δz(A)\delta_{z}(A) is the shortest length of a zonotope containing AA. The function δz(A)\delta_{z}(A) is clearly monotonic, vanishes when |A|1|A|\leq 1, and is strictly positive when |A|>1|A|>1. Given finite A,BA,B, let ZAZ_{A} and ZBZ_{B} the the minimum length zonotopes containing AA and BB respectively. Then ZA+ZBZ_{A}+Z_{B} is a zonotope containing A+BA+B with length (ZA)+(ZB)\ell(Z_{A})+\ell(Z_{B}). By Proposition 1 part 2, (k,δz)(\mathbb{R}^{k},\delta_{z}) is a sublinear diversity. ∎

The zonotope diversity is not linear: let A={(0,0),(1,0),(0,1)}A=\{(0,0),(1,0),(0,1)\} and B=AB=-A. Then δz(A)=δz(B)=2\delta_{z}(A)=\delta_{z}(B)=2 but δz(A+B)=2+2\delta_{z}(A+B)=2+\sqrt{2}.

In a semidiversity, (D1) is replaced by (D1), and sets with more than one element can have diversity zero. When the semidiversity is sublinear, the sets with zero diversity are highly structured. Define the null set of a semidiversity (k,δ)(\mathbb{R}^{k},\delta) to be the set

null(δ)={x:δ({0,x})=0}\mathrm{null}(\delta)=\Big{\{}x:\delta(\{0,x\})=0\Big{\}}

and null(δ)={xk:xy=0 for all ynull(δ)}\mathrm{null}(\delta)^{\perp}=\{x\in\mathbb{R}^{k}:x\cdot y=0\mbox{ for all }y\in\mathrm{null}(\delta)\}.

Proposition 3.

Let (k,δ)(\mathbb{R}^{k},\delta) be a sublinear semidiversity.

  1. 1.

    null(δ)\mathrm{null}(\delta) is a linear subspace of k\mathbb{R}^{k}

  2. 2.

    δ\delta restricted to null(δ)\mathrm{null}(\delta)^{\perp} is a diversity

  3. 3.

    If PP is the projection operator for null(δ)\mathrm{null}(\delta)^{\perp} then δ(A)=δ(PA)\delta(A)=\delta(PA) for all finite AkA\subseteq\mathbb{R}^{k}.

Proof.
  1. 1.

    For x,ynull(δ)x,y\in\mathrm{null}(\delta) and α>0\alpha>0 we have δ({0,x+y})δ({0,x})+δ({0,y})=0\delta(\{0,x+y\})\leq\delta(\{0,x\})+\delta(\{0,y\})=0 and δ({0,αx})=αδ({0,x})=0\delta(\{0,\alpha x\})=\alpha\delta(\{0,x\})=0 so x+ynull(δ)x+y\in\mathrm{null}(\delta) and αxnull(δ)\alpha x\in\mathrm{null}(\delta). By translation invariance, δ({0,x})=δ({x,0})=0\delta(\{0,-x\})=\delta(\{x,0\})=0 and xnull(δ)-x\in\mathrm{null}(\delta).

  2. 2.

    Suppose x,ynull(δ)x,y\in\mathrm{null}(\delta)^{\perp} and δ({x,y})=0\delta(\{x,y\})=0. We have that xynull(δ)x-y\in\mathrm{null}(\delta)^{\perp} by part 1. By translation invariance δ({0,xy})=δ({x,y})=0\delta(\{0,x-y\})=\delta(\{x,y\})=0 which implies xynull(δ)x-y\in\mathrm{null}(\delta). Hence xyx-y is both in a subspace and its orthogonal complement, and so x=yx=y.

  3. 3.

    For all finite AkA\subseteq\mathbb{R}^{k} we have APA+BA\subseteq PA+B and PAA+CPA\subseteq A+C for some B,Cnull(δ)B,C\subset\mathrm{null}(\delta). We have δ(B)=0\delta(B)=0 since

    0δ(B)bBδ({0,b})=0,0\leq\delta(B)\leq\sum_{b\in B}\delta(\{0,b\})=0,

    and, likewise, δ(C)=0\delta(C)=0. By sublinearity, δ(A)=δ(PA)\delta(A)=\delta(PA).

Let \|\cdot\| be a norm on k\mathbb{R}^{k} with associated metric d(x,y)=xyd(x,y)=\|x-y\| and unit ball ={x:x1}\mathcal{B}=\{x:\|x\|\leq 1\}. The Hausdorff distance between two nonempty closed bounded sets KK and LL can be defined by [24, p. 61] :

dH(K,L)=min{λ:KL+λ and LK+λ}.d_{H}(K,L)=\min\{\lambda:K\subseteq L+\lambda\mathcal{B}\mbox{ and }L\subseteq K+\lambda\mathcal{B}\}.

For bounded KkK\subseteq\mathbb{R}^{k} define

δ(K)=sup{δ(A):AK finite}.\delta^{*}(K)=\sup\{\delta(A):A\subseteq K\mbox{ finite}\}. (2)
Proposition 4.

Let (k,δ)(\mathbb{R}^{k},\delta) be a sublinear semidiversity.

  1. 1.

    For all bounded KkK\subseteq\mathbb{R}^{k}, δ(K)<\delta^{*}(K)<\infty.

  2. 2.

    For all finite AkA\subseteq\mathbb{R}^{k} we have

    δ(conv(A))=δ(A).\delta^{*}(\mathrm{conv}(A))=\delta(A).
  3. 3.

    For all bounded K,LkK,L\subset\mathbb{R}^{k} and λ0\lambda\geq 0

    δ(K+L)δ(K)+δ(L)\delta^{*}(K+L)\leq\delta^{*}(K)+\delta^{*}(L)

    and

    δ(λK)=λδ(K).\delta^{*}(\lambda K)=\lambda\delta^{*}(K).
  4. 4.

    If (k,δ)(\mathbb{R}^{k},\delta) is linear then the restriction of δ\delta^{*} to the set of nonempty compact convex subsets of k\mathbb{R}^{k} is a valuation. That is,

    δ(KL)+δ(KL)=δ(K)+δ(L)\delta^{*}(K\cap L)+\delta^{*}(K\cup L)=\delta^{*}(K)+\delta^{*}(L)

    for all nonempty compact convex bodies K,LK,L such that KLK\cap L and KLK\cup L are non-empty and convex.

  5. 5.

    The restriction of δ\delta^{*} to the set of nonempty compact convex subsets of k\mathbb{R}^{k} is Lipschitz continuous with respect to the Hausdorff metric, with Lipschitz constant δ()\delta^{*}(\mathcal{B}).

Proof.
  1. 1.

    By equivalency of norms on k\mathbb{R}^{k} we have that KK is bounded with respect to metric dd if and only if it is bounded with respect to the induced metric of δ\delta. Let VV be the set of vertices of some polytope (e.g. a cube) containing KK. For all finite AKA\subseteq K we have by monotonicity and part 2 that

    δ(A)δ(AV)=δ(V)\delta(A)\leq\delta(A\cup V)=\delta(V)

    so that δ(K)δ(V)<.\delta^{*}(K)\leq\delta(V)<\infty.

  2. 2.

    Let AA be a finite subset of k\mathbb{R}^{k} and let K=conv(A)K=\mathrm{conv}(A). For any AKA^{\prime}\subseteq K we have conv(AA)=conv(A)\mathrm{conv}(A\cup A^{\prime})=\mathrm{conv}(A) so δ(A)δ(AA)=δ(A)\delta(A^{\prime})\leq\delta(A\cup A^{\prime})=\delta(A) by Proposition 1 (ii). Hence

    δ(A)sup{δ(A): finite AK)}δ(A).\delta(A)\leq\sup\{\delta(A^{\prime}):\mbox{ finite }A^{\prime}\subseteq K)\}\leq\delta(A).
  3. 3.

    Fix ϵ>0\epsilon>0 and suppose that CC is a finite subset of K+LK+L such that δ(C)>δ(K+L)ϵ\delta(C)>\delta^{*}(K+L)-\epsilon. For each cCc\in C there is acKa_{c}\in K and bcLb_{c}\in L such that c=ac+bcc=a_{c}+b_{c}. Let A={ac:cC}KA=\{a_{c}:c\in C\}\subseteq K and B={bc:cC}LB=\{b_{c}:c\in C\}\subseteq L so that CA+BC\subseteq A+B. It follows that

    δ(K+L)ϵ<δ(C)δ(A+B)δ(A)+δ(B)δ(K)+δ(L).\delta^{*}(K+L)-\epsilon<\delta(C)\leq\delta(A+B)\leq\delta(A)+\delta(B)\leq\delta^{*}(K)+\delta^{*}(L).

    Taking ϵ\epsilon to zero gives the result.

    Let AA be a finite subset of KK such that δ(A)>δ(K)ϵ\delta(A)>\delta^{*}(K)-\epsilon. As λAλK\lambda A\subseteq\lambda K we have

    λ(δ(K)ϵ)<λδ(A)=δ(λA)δ(λK).\lambda(\delta^{*}(K)-\epsilon)<\lambda\delta(A)=\delta(\lambda A)\leq\delta^{*}(\lambda K).

    Hence δ(λK)λδ(K)\delta^{*}(\lambda K)\geq\lambda\delta^{*}(K) from which equality follows by symmetry.

  4. 4.

    By Lemma 3.1.1 of [24] we have that if K,L,KLK,L,K\cup L and KLK\cap L are nonempty compact convex subsets then

    (KL)+(KL)=K+L.(K\cup L)+(K\cap L)=K+L.

    By linearity,

    δ(KL)+δ(KL)=δ(K)+δ(L).\delta^{*}(K\cup L)+\delta^{*}(K\cap L)=\delta^{*}(K)+\delta^{*}(L).
  5. 5.

    Suppose that K,LK,L are bounded nonempty subsets satisfying dH(K,L)=λd_{H}(K,L)=\lambda. For any ϵ>0\epsilon>0 there is a finite AKA\subseteq K such that δ(A)δ(K)<δ(A)+ϵ\delta(A)\leq\delta^{*}(K)<\delta(A)+\epsilon. We also have AKL+λA\subseteq K\subseteq L+\lambda\mathcal{B} so there is finite BLB\subseteq L and CC\subseteq\mathcal{B} such that AB+λCA\subseteq B+\lambda C. Hence

    δ(K)<δ(A)+ϵδ(B)+λδ(C)+ϵδ(L)+λδ()+ϵ.\delta^{*}(K)<\delta(A)+\epsilon\leq\delta(B)+\lambda\delta(C)+\epsilon\leq\delta^{*}(L)+\lambda\delta^{*}(\mathcal{B})+\epsilon.

    By a symmetric argument,

    δ(L)<δ(K)+λδ()+ϵ.\delta^{*}(L)<\delta^{*}(K)+\lambda\delta^{*}(\mathcal{B})+\epsilon.

    Taking ϵ\epsilon to zero, we have

    |δ(K)δ(L)|δ()dH(K,L).|\delta^{*}(K)-\delta^{*}(L)|\leq\delta^{*}(\mathcal{B})d_{H}(K,L).

    The bound is tight, as can be seen by letting K=K=\mathcal{B} and L=2L=2\mathcal{B}. Then dH(K,L)=1d_{H}(K,L)=1 and |δ(K)δ(L)|=δ()|\delta^{*}(K)-\delta^{*}(L)|=\delta^{*}(\mathcal{B}).

Bryant et al. [3] also describe an extension of Minkowski diversities from finite sets to bounded sets. They define δ~(P)=δ(vert(P))\widetilde{\delta}(P)=\delta(\mathrm{vert}(P)) for any polytope with vertex set vert(P)\mathrm{vert}(P), and extend that to general bounded convex sets KK by defining δ~(K)=limnδ~(Pn)\widetilde{\delta}(K)=\lim_{n\rightarrow\infty}\widetilde{\delta}(P_{n}) for any sequence of polytopes P1,P2,P_{1},P_{2},\ldots converging to KK. Proposition 4 part 2. gives that δ(P)=δ~(P)\delta^{*}(P)=\widetilde{\delta}(P) for any polytope, while from Proposition 4 part 5 we have that δ(K)=δ~(K)\delta^{*}(K)=\widetilde{\delta}(K). Hence δ\delta^{*} coincides with δ~\widetilde{\delta} for Minkowski diversities.

2.3 Characterization of linear diversities

The following characterization of linear diversities is essentially contained in the proof of the main theorem in Firey [10]; see also [21].

Theorem 5.

Let δ\delta be a function defined on finite subsets of k\mathbb{R}^{k}. Then (X,δ)(X,\delta) is a linear semidiversity if and only if there is a positive finite Borel measure ν\nu on the unit sphere 𝕊k1={xk:x2=1}\mathbb{S}^{k-1}=\{x\in\mathbb{R}^{k}:\|x\|_{2}=1\} such that

𝕊k1xdν(x)=0\int_{\mathbb{S}^{k-1}}x\,\mathrm{d}\nu(x)=0 (3)

and

δ(A)=𝕊k1hA(x)dν(x)\delta(A)=\int_{\mathbb{S}^{k-1}}h_{A}(x)\,\mathrm{d}\nu(x) (4)

for all finite AkA\subseteq\mathbb{R}^{k}. Such a measure is unique.

Proof.

First we show that δ\delta given by (4) is a linear semidiversity. For aka\in\mathbb{R}^{k} and finite ABkA\subseteq B\subseteq\mathbb{R}^{k} we have

δ({a})\displaystyle\delta(\{a\}) =𝕊k1h{a}(x)dν(x)=𝕊k1axdν(x)=a𝕊k1xdν(x)=0\displaystyle=\int_{\mathbb{S}^{k-1}}h_{\{a\}}(x)\,\mathrm{d}\nu(x)=\int_{\mathbb{S}^{k-1}}a\cdot x\,\mathrm{d}\nu(x)=a\cdot\int_{\mathbb{S}^{k-1}}x\,\mathrm{d}\nu(x)=0

and, since hA(x)hB(x)h_{A}(x)\leq h_{B}(x) and hA+B(x)=hA(x)+hB(x)h_{A+B}(x)=h_{A}(x)+h_{B}(x) for all xx we have δ(A)δ(B)\delta(A)\leq\delta(B) and δ(A+B)=δ(A)+δ(B)\delta(A+B)=\delta(A)+\delta(B). By Proposition 1, (k,δ)(\mathbb{R}^{k},\delta) is a linear semidiversity.

For the converse, let (k,δ)(\mathbb{R}^{k},\delta) be a linear semidiversity and define δ\delta^{*} as in (2). By Proposition 4 the restriction of δ\delta^{*} to nonempty compact convex subsets is Minkowski linear, monotonic and vanishes on singletons. From the proof of the main theorem in [10], we have that, for all compact convex sets KK,

δ(K)=𝕊k1hK(x)dν(x)\delta^{*}(K)=\int_{\mathbb{S}^{k-1}}h_{K}(x)\,\mathrm{d}\nu(x)

for some positive finite Borel measure ν\nu satisfying (3), and ν\nu is the unique such measure. (See [23, Thm 2.14] for details on the use of the Riesz Theorem in this case.) Now for any nonempty finite AA, let K=conv(A)K=\mathrm{conv}(A). Since δ(A)=δ(K)\delta(A)=\delta^{*}(K) and hA=hKh_{A}=h_{K}, the result follows for all nonempty finite AA. ∎

Refer to caption
Figure 1: Support of measures corresponding to (a) mean width; (b) the L1L_{1} diversity; and (c) a Minkowski diversity with a simplex kernel.

In Figure 1 we depict the support for the measures corresponding to mean width (uniform on the unit circle), the L1L_{1} diversity (±ei\pm e_{i}), and the Minkowski diversity for a simplex kernel. The first two of these are easy enough to demonstrate. We prove the third example below, after we have a characterization of extremal linear diversities.

2.4 Extremal linear diversities

The set of linear semidiversities on k\mathbb{R}^{k} forms a cone. A non-zero semidiversity δ\delta is extremal (or lies on an extremal ray) if it cannot be expressed as the convex combination of two linear semidiversities which are not its scale copies. We make use of Theorem 5 to characterize the extremal linear diversities and semidiversities. First we prove a technical result simplifying evaluation of the Minkowski diversity for a simplex.

Lemma 6.

Let v0,,vjkv_{0},\ldots,v_{j}\in\mathbb{R}^{k} be affinely independent with =0jcv=0\sum_{\ell=0}^{j}c_{\ell}v_{\ell}=0, for some c0c_{\ell}\geq 0, c=1\sum_{\ell}c_{\ell}=1. Define the polyhedron K={y:vy1,for all }K=\{y:v_{\ell}\cdot y\leq 1,\,\text{for all }\ell\}. Let δK\delta_{K} be the Minkowski semidiversity given by KK. Then

δK(A)==0jchA(v).\delta_{K}(A)=\sum_{\ell=0}^{j}c_{\ell}h_{A}(v_{\ell}).

for all finite AkA\subseteq\mathbb{R}^{k}.

Proof.

Let A={ai}i=1,,|A|A=\{a_{i}\}_{i=1,\ldots,|A|}. We express δK(A)\delta_{K}(A) as the solution to a linear program. Recall that δK(A)\delta_{K}(A) is the minimum λ\lambda such that there is some xkx\in\mathbb{R}^{k} such that aixλKa_{i}-x\in\lambda K for all ii. We can rewrite this constraint as v(aix)λv_{\ell}\cdot(a_{i}-x)\leq\lambda for all i,i,\ell. If we take λ\lambda and xx to be our primal variables we get the following linear program:

minimize λ=(1,0)(λ,x)\displaystyle\lambda=(1,0)\cdot(\lambda,x)
subject to λ+vxvai,for all i and .\displaystyle\lambda+v_{\ell}\cdot x\geq v_{\ell}\cdot a_{i},\mbox{for all }i\mbox{ and }\ell.

The dual linear program with dual variables yiy_{i\ell} is

maximize i(vai)yij\displaystyle\sum_{i\ell}(v_{\ell}\cdot a_{i})y_{ij}
subject to yi0,for all i and ,\displaystyle y_{i\ell}\geq 0,\mbox{for all }i\mbox{ and }\ell,
iyi=1,\displaystyle\sum_{i\ell}y_{i\ell}=1,
ivyi=0.\displaystyle\sum_{i\ell}v_{\ell}y_{i\ell}=0.

Let y¯=iyi\bar{y}_{\ell}=\sum_{i}y_{i\ell}. Then our dual constraints are equivalent to

y¯=1,y¯v=0.\sum_{\ell}\bar{y}_{\ell}=1,\ \ \ \sum_{\ell}\bar{y}_{\ell}v_{\ell}=0.

Since the vv_{\ell} are affinely independent and cv=0\sum_{\ell}c_{\ell}v_{\ell}=0, c=1\sum_{\ell}c_{\ell}=1, there is a unique solution given by y¯=c\bar{y}_{\ell}=c_{\ell} for all \ell. Now it remains to determine for each \ell the value of yiy_{i\ell} for each ii. We need to maximize i(vai)yi\sum_{i\ell}(v_{\ell}\cdot a_{i})y_{i\ell} given yi0y_{i\ell}\geq 0 and iyi=c\sum_{i}y_{i\ell}=c_{\ell}. For each \ell, the solution is to let yi=cy_{i\ell}=c_{\ell} for the ii that maximizes (vai)(v_{\ell}\cdot a_{i}), and 0 otherwise. This gives for the solution to the dual problem

cmaxaAva=chA(v).\sum_{\ell}c_{\ell}\max_{a\in A}v_{\ell}\cdot a=\sum_{\ell}c_{\ell}h_{A}(v_{\ell}).

The following theorem identifies extremal linear semidiversities as Minkowski diversities δK\delta_{K} with KK equal to a simplex or a simplex plus a subspace.

Theorem 7.

The following are equivalent for a semidiversity (k,δ)(\mathbb{R}^{k},\delta):

  1. (i)

    (k,δ)(\mathbb{R}^{k},\delta) is extremal in the class of linear semidiversities.

  2. (ii)

    (k,δ)(\mathbb{R}^{k},\delta) satisfies

    δ(A)=𝕊k1hA(x)dν(x)\delta(A)=\int_{\mathbb{S}^{k-1}}h_{A}(x)\,\mathrm{d}\nu(x)

    for all finite AkA\subseteq\mathbb{R}^{k}, where ν\nu is a measure on 𝕊k1\mathbb{S}^{k-1} with 𝕊k1xdν(x)=0\int_{\mathbb{S}^{k-1}}x\,\mathrm{d}\nu(x)=0, such that the support of ν\nu is a finite, affinely independent set.

  3. (iii)

    (k,δ)(\mathbb{R}^{k},\delta) is a Minkowski semidiversity with kernel KK of the form

    K=conv(W)+H,K=\mathrm{conv}(W)+H^{\perp},

    where WW is an affinely independent set of points, HH is the affine closure of WW,and HH^{\perp} is the orthogonal space to HH.

Proof.

(i) \Rightarrow (ii). Suppose δ\delta is extremal and the support of ν\nu is not affinely independent. Let HH be the affine hull of the support of ν\nu, with dimH=j\dim H=j, and let S=H𝕊k1S=H\cap\mathbb{S}^{k-1}. Affine dependence implies ν\nu is not supported on only j+1j+1 points or fewer. Therefore we can partition SS into S1,,Sj+2S_{1},\ldots,S_{j+2}, each with ν(Si)>0\nu(S_{i})>0. Let mi=Sidν(x)=ν(Si)>0m_{i}=\int_{S_{i}}\,\mathrm{d}\nu(x)=\nu(S_{i})>0. Then

m=Sdν(x)=iSi𝑑ν(x)=imim=\int_{S}\,\mathrm{d}\nu(x)=\sum_{i}\int_{S_{i}}d\nu(x)=\sum_{i}m_{i}

and

0=Sxdν(x)=jSixdν(x)=imixi0=\int_{S}x\,\mathrm{d}\nu(x)=\sum_{j}\int_{S_{i}}x\,\mathrm{d}\nu(x)=\sum_{i}m_{i}x_{i}

where xi=(Sixdν(x))/mix_{i}=(\int_{S_{i}}x\,\mathrm{d}\nu(x))/m_{i}. Choose a subset of the xix_{i} with k+1k+1 points so that 0 is in the convex hull of them. Let’s say they are x1,,xk+1x_{1},\ldots,x_{k+1}. Find μi\mu_{i} for i=1,,k+1i=1,...,k+1 such that iμixi=0\sum_{i}\mu_{i}x_{i}=0, and μi<mi\mu_{i}<m_{i}. Let μk+2=0\mu_{k+2}=0. Now define ν\nu^{\prime} by ν(A)=(μj/mj)ν(A)\nu^{\prime}(A)=(\mu_{j}/m_{j})\nu(A) for ASi,j=1k+1A\subseteq S_{i},j=1...k+1, and zero otherwise. Then νν\nu^{\prime}\leq\nu, and ν\nu^{\prime} has smaller support, because mk+2>0m_{k+2}>0 but μk+2=0\mu_{k+2}=0. Also

xdν(x)=iSixμjmj𝑑ν(x)=iμimjmixi=iμixi=0.\int x\,\mathrm{d}\nu^{\prime}(x)=\sum_{i}\int_{S_{i}}x\frac{\mu_{j}}{m_{j}}d\nu(x)=\sum_{i}\frac{\mu_{i}}{m_{j}}m_{i}x_{i}=\sum_{i}\mu_{i}x_{i}=0.

We can now write ν=(νν)+ν\nu=(\nu-\nu^{\prime})+\nu^{\prime} where νν\nu-\nu^{\prime} and ν\nu are not scale copies, so δ\delta is not extremal in the cone of linear semidiversities.
(ii) \Rightarrow (i). Suppose that

δ(A)=𝕊k1hA(x)dν(x)\delta(A)=\int_{\mathbb{S}^{k-1}}h_{A}(x)\,\mathrm{d}\nu(x)

for all finite AkA\subseteq\mathbb{R}^{k} and some measure ν\nu on 𝕊k1\mathbb{S}^{k-1} with affinely independent support. Let δ1\delta_{1} and δ2\delta_{2} be linear semidiversities with corresponding measures ν1\nu_{1} and ν2\nu_{2}. If supp(ν1)supp(ν)\mathrm{supp}(\nu_{1})\subseteq\mathrm{supp}(\nu) then affine independence and the constraint that 𝕊k1xdν1(x)=0\int_{\mathbb{S}^{k-1}}x\,\mathrm{d}\nu_{1}(x)=0 implies that ν1\nu_{1} is a scaled version of ν\nu. Likewise for ν2\nu_{2}. Hence if ν=λν1+(1λ)ν2\nu=\lambda\nu_{1}+(1-\lambda)\nu_{2} for λ[0,1]\lambda\in[0,1] we have that supp(ν1)supp(ν2)supp(ν)\mathrm{supp}(\nu_{1})\cup\mathrm{supp}(\nu_{2})\subseteq\mathrm{supp}(\nu) and both ν1\nu_{1} and ν2\nu_{2} are scale versions of ν\nu. This shows that δ\delta is an extremal linear diversity.
(ii) \Rightarrow (iii). Let the support of ν\nu be the points u0,,uju_{0},\ldots,u_{j} with weights m>0m_{\ell}>0 such that mu=0\sum_{\ell}m_{\ell}u_{\ell}=0. Let m=mm=\sum_{\ell}m_{\ell}, c=m/mc_{\ell}=m_{\ell}/m, and v=muv_{\ell}=mu_{\ell}, so that δ(A)=chA(v)\delta(A)=\sum_{\ell}c_{\ell}h_{A}(v_{\ell}), cv=0\sum_{\ell}c_{\ell}v_{\ell}=0, and c=1\sum_{\ell}c_{\ell}=1. Let V={v0,,vj}V=\{v_{0},\ldots,v_{j}\}, which is affinely independent because the uu_{\ell} are. Let HH be the span of VV and HH^{\perp} its orthogonal complement. By Lemma 6, δ=δK\delta=\delta_{K}, the Minkowski semidiversity for the set K={y:yv1, for all }K=\{y:y\cdot v_{\ell}\leq 1,\mbox{ for all }\ell\}. The intersection of KK with HH is a simplex; let it have vertices W={w0,,wj}W=\{w_{0},\ldots,w_{j}\}. Then K=conv(V)+HK=\mathrm{conv}(V)+H^{\perp} as required.
(iii) \Rightarrow (ii). By translating KK if necessary, we may assume 0 is in the relative interior of conv(W)\mathrm{conv}(W). We can write K={y:vy1}K=\{y:v_{\ell}\cdot y\leq 1\} for some affinely independent V={v0,,vj}V=\{v_{0},\ldots,v_{j}\}. Because conv(W)\mathrm{conv}(W) is bounded, 0conv(V)0\in\mathrm{conv}(V). So there are c0c_{\ell}\geq 0 with cv=0\sum_{\ell}c_{\ell}v_{\ell}=0 and c=1\sum_{\ell}c_{\ell}=1. By Lemma 6 we have

δK(A)=chA(v)\delta_{K}(A)=\sum_{\ell}c_{\ell}h_{A}(v_{\ell})

for all finite AA. Let m=c|v|m_{\ell}=c_{\ell}|v_{\ell}| and u=v/|v|u_{\ell}=v_{\ell}/|v_{\ell}|. Then the uu_{\ell} are also affinely independent. Let ν\nu be the measure that assigns mass mm_{\ell} to each uu_{\ell}. ∎

Points in a finite dimensional convex cone can always be written as convex combinations of extremal points. The cone of linear semidiversities has infinite dimensional, so proving that linear semidiversities are in the convex hull of extremal diversities requires a little more work.

Theorem 8.

A semidiversity (k,δ)(\mathbb{R}^{k},\delta) is linear if and only if δ\delta is a convex combination of extremal linear semidiversity functions.

Proof.

Since a weighted average of linear semidiversities is a linear semidiversity, one way is immediate. For the other, suppose that δ\delta is a linear semidiversity. By Theorem 5, there is a Borel measure ν\nu on 𝕊k1\mathbb{S}^{k-1} such that xdν(x)=0\int x\,\mathrm{d}\nu(x)=0 and δ(A)=hA(x)dν(x)=0\delta(A)=\int h_{A}(x)\,\mathrm{d}\nu(x)=0 for all finite AA. Let m=dν(x)m=\int\,\mathrm{d}\nu(x).

Let EE be the set of all signed Borel measures on 𝕊k1\mathbb{S}^{k-1}, which is a Hausdorff locally convex set [26, p. 134]. The space CC of measures on 𝕊k1\mathbb{S}^{k-1} with xdν(x)=m\int x\,\mathrm{d}\nu(x)=m and xdν(x)=0\int x\,\mathrm{d}\nu(x)=0 is compact by the Banach-Alaoglu theorem [26, p. 114], and is convex.

We claim that the set of extremal points of CC is closed. Let νn\nu_{n}, n1n\geq 1 be a sequence of extremal measures that converges in the vague topology, so that fdνn\int f\,\mathrm{d}\nu_{n} converges to fdν\int f\,\mathrm{d}\nu for some νC\nu\in C for all continuous bounded ff. By repeatedly taking subsequences, we can obtain a subsequence νnk=i=1jμi,kδxi,k\nu_{n_{k}}=\sum_{i=1}^{j}\mu_{i,k}\delta_{x_{i,k}} (where δx\delta_{x} is a unit mass measure at xx) where xi,kxix_{i,k}\rightarrow x_{i} and μi,kμi\mu_{i,k}\rightarrow\mu_{i} for some xi𝕊k1x_{i}\in\mathbb{S}^{k-1} and μi0\mu_{i}\geq 0. Since fdνnkfdν\int f\,\mathrm{d}\nu_{n_{k}}\rightarrow\int f\,\mathrm{d}\nu as kk\rightarrow\infty, we must have ν=iμiδxi\nu=\sum_{i}\mu_{i}\delta_{x_{i}}, showing that the limit is also an extremal point in CC. Hence the set of extremal measures is closed.

We can apply a version of the Krein-Milman Theorem ([26, Corollary 17.7]) to obtain that δ\delta is a weighted average of members of the closure of the extreme points of CC. Since the set of extremal points of CC is closed, the result follows. ∎

2.5 Characterization of sublinear diversities

We now turn our attention to sublinear diversities. We will show that the relationship between sublinear and linear diversities parallels that between convex and linear functions. Just as every convex function is the supremum of linear functions, every sublinear diversity is the supremum of linear diversities (Theorem 9). In fact, in our case, the supremum is attained for each set, so the value of every sublinear diversity on a set is the maximum of the value of a family of linear diversities on the set. Our proof relies heavily on the ‘Sandwich Theorem’ (Theorem 1.2.5) of [11].

Theorem 9.

Let δ\delta be a function on finite subsets of k\mathbb{R}^{k}. If (k,δ)(\mathbb{R}^{k},\delta) is a sublinear diversity or semidiversity then there is a collection {(k,δγ)}γΓ\{(\mathbb{R}^{k},\delta_{\gamma})\}_{\gamma\in\Gamma} of linear semidiversities such that

δ(A)=max{δγ(A):γΓ}.\delta(A)=\max\{\delta_{\gamma}(A):\gamma\in\Gamma\}.

Conversely, for any collection {(k,δγ)}γΓ\{(\mathbb{R}^{k},\delta_{\gamma})\}_{\gamma\in\Gamma} of linear semidiversities and δ\delta defined by δ(A)=supγΓδγ(A)\delta(A)=\sup_{\gamma\in\Gamma}\delta_{\gamma}(A), (k,δ)(\mathbb{R}^{k},\delta) is a sublinear semidiversity.

Proof.

Suppose that {(k,δγ)}γΓ\{(\mathbb{R}^{k},\delta_{\gamma})\}_{\gamma\in\Gamma} are linear semidiversities and

δ(A)=sup{δγ(A):γΓ}\delta(A)=\sup\{\delta_{\gamma}(A):\gamma\in\Gamma\}

for all finite AkA\subseteq\mathbb{R}^{k}. Note that δ\delta vanishes on singletons and is monotonic since each δγ\delta_{\gamma} has these properties. Suppose that A,BA,B are finite subsets of k\mathbb{R}^{k} and λ0\lambda\geq 0. Then

δ(λA)=sup{δγ(λA):γΓ}=sup{λδγ(A):γΓ}=λδ(A)\delta(\lambda A)=\sup\{\delta_{\gamma}(\lambda A):\gamma\in\Gamma\}=\sup\{\lambda\delta_{\gamma}(A):\gamma\in\Gamma\}=\lambda\delta(A)

and

δ(A+B)=sup{δγ(A+B):γΓ}=sup{δγ(A)+δγ(B):γΓ}δ(A)+δ(B).\delta(A+B)=\sup\{\delta_{\gamma}(A+B):\gamma\in\Gamma\}=\sup\{\delta_{\gamma}(A)+\delta_{\gamma}(B):\gamma\in\Gamma\}\leq\delta(A)+\delta(B).

So δ\delta is sublinear. By Proposition 1, δ\delta is a sublinear semidiversity.

For the converse, suppose that (k,δ)(\mathbb{R}^{k},\delta) is sublinear. Define \mathcal{H} to be the set of all support functions hAh_{A} for nonempty finite AkA\subseteq\mathbb{R}^{k}. Define pp on the convex cone \mathcal{H} by p(hA)=δ(A)p(h_{A})=\delta(A) for all finite sets AA. The function pp is sublinear (and convex in the terminology of [25]), as for any finite A,BA,B,

p(hA+hB)=p(hA+B)=δ(A+B)δ(A)+δ(B)=p(hA)+p(hB),p(h_{A}+h_{B})=p(h_{A+B})=\delta(A+B)\leq\delta(A)+\delta(B)=p(h_{A})+p(h_{B}),

and p(λhA)=λp(hA)p(\lambda h_{A})=\lambda p(h_{A}) for λ0\lambda\geq 0.

Fix finite BkB\subseteq\mathbb{R}^{k}. Define qBq_{B} on \mathcal{H} by

qB(hA)=sup{λ:λB+xconv(A) for some xk}.q_{B}(h_{A})=\sup\{\lambda:\lambda B+x\subseteq\mathrm{conv}(A)\mbox{ for some $x\in\mathbb{R}^{k}$}\}.

That is, qB(hA)q_{B}(h_{A}) is the largest we can scale BB so that a translate is contained in conv(A)\mathrm{conv}(A). Note that qB(hB)=1q_{B}(h_{B})=1. This tells us that p(hB)=δ(B)=δ(B)qB(hB)p(h_{B})=\delta(B)=\delta(B)q_{B}(h_{B}).

We show that qBq_{B} is superlinear. For all α0\alpha\geq 0 we have qB(αhA)=qB(hαA)=αqB(hA)q_{B}(\alpha h_{A})=q_{B}(h_{\alpha A})=\alpha q_{B}(h_{A}). Now suppose that A1,A2A_{1},A_{2} are finite and non-empty subsets of k\mathbb{R}^{k}. Given ϵ>0\epsilon>0 there are λ1>qB(hA1)ϵ/2\lambda_{1}>q_{B}(h_{A_{1}})-\epsilon/2, λ2>qB(hA2)ϵ/2\lambda_{2}>q_{B}(h_{A_{2}})-\epsilon/2, x1,x2kx_{1},x_{2}\in\mathbb{R}^{k} such that

λ1B+x1\displaystyle\lambda_{1}B+x_{1} conv(A1)\displaystyle\subseteq\mathrm{conv}(A_{1})
λ2B+x2\displaystyle\lambda_{2}B+x_{2} conv(A2)\displaystyle\subseteq\mathrm{conv}(A_{2})
and hence
(λ1+λ2)B+(x1+x2)\displaystyle(\lambda_{1}+\lambda_{2})B+(x_{1}+x_{2}) conv(A1)+conv(A2)\displaystyle\subseteq\mathrm{conv}(A_{1})+\mathrm{conv}(A_{2})
=conv(A1+A2),\displaystyle=\mathrm{conv}(A_{1}+A_{2}),

so that qB(hA1+A2)(λ1+λ2)>qB(hA1)+qB(hA2)ϵq_{B}(h_{A_{1}+A_{2}})\geq(\lambda_{1}+\lambda_{2})>q_{B}(h_{A_{1}})+q_{B}(h_{A_{2}})-\epsilon. Taking ϵ0\epsilon\rightarrow 0 gives superlinearity.

We now have that pp is monotonic and sublinear and that qBq_{B} is superlinear.

Furthermore, for any finite AkA\subseteq\mathbb{R}^{k} and ϵ>0\epsilon>0 there is λ\lambda such that qB(hA)ϵ<λqB(hA)q_{B}(h_{A})-\epsilon<\lambda\leq q_{B}(h_{A}) and xkx\in\mathbb{R}^{k} such that λB+xconv(A)\lambda B+x\subseteq\mathrm{conv}(A), and so

p(hA)\displaystyle p(h_{A}) p(hλB)\displaystyle\geq p(h_{\lambda B})
=λp(hB)\displaystyle=\lambda p(h_{B})
>(qB(hA)ϵ)δ(B).\displaystyle>(q_{B}(h_{A})-\epsilon)\delta(B).

Taking ϵ0\epsilon\rightarrow 0 we conclude that qB(hA)δ(B)p(hA)q_{B}(h_{A})\delta(B)\leq p(h_{A}) for all hAh_{A}\in\mathcal{H}. Recall that qB(hB)δ(B)=p(hB)q_{B}(h_{B})\delta(B)=p(h_{B}).

For each finite BB we have now satisfied the conditions for Theorem 1.2.5 of [11]:

Let FF be a pre-ordered cone and let p:F¯p:F\rightarrow\overline{\mathbb{R}} be monotone and sublinear, q:F¯q:F\rightarrow\overline{\mathbb{R}} superlinear with qpq\leq p. Then there is a monotone linear μ:F¯\mu:F\rightarrow\overline{\mathbb{R}} with qμpq\leq\mu\leq p.

In our example FF is the cone \mathcal{H} of support functions of finite sets. Let q(h)=δ(B)qB(h)q(h)=\delta(B)q_{B}(h). Let μB:\mu_{B}\colon\mathcal{H}\rightarrow\mathbb{R} be the linear map given by the theorem. It is monotone, linear, and

δ(B)qB(h)μB(h)p(h).\delta(B)q_{B}(h)\leq\mu_{B}(h)\leq p(h).

Since by definition p(h{a})=δ({a})=0p(h_{\{a\}})=\delta(\{a\})=0 for all aka\in\mathbb{R}^{k}, μB(h{a})=0\mu_{B}(h_{\{a\}})=0 for all aka\in\mathbb{R}^{k}.

Now define δB\delta_{B} by δB(A)=μB(hA)\delta_{B}(A)=\mu_{B}(h_{A}) for all finite AA. Then δB\delta_{B} vanishes on singletons, it is monotone, linear, and hence also sublinear. By Proposition 1 (k,δB)(\mathbb{R}^{k},\delta_{B}) is a linear semidiversity.

Because δ(B)qB(hB)=p(hB)\delta(B)q_{B}(h_{B})=p(h_{B}), we have that

δB(B)=μB(hB)=δ(B)qB(B)=δ(B)\delta_{B}(B)=\mu_{B}(h_{B})=\delta(B)q_{B}(B)=\delta(B)

and for general finite AA we have

δB(A)=μB(hA)p(hA)=δ(A).\delta_{B}(A)=\mu_{B}(h_{A})\leq p(h_{A})=\delta(A).

Repeating this process for all finite BkB\subseteq\mathbb{R}^{k} we obtain a set of linear semidiversities {δB}finite Bk\{\delta_{B}\}_{\text{finite }B\subseteq\mathbb{R}^{k}} such that δBδ\delta_{B}\leq\delta and δB(B)=δ(B)\delta_{B}(B)=\delta(B) for all finite BkB\subseteq\mathbb{R}^{k}. So for all finite AkA\subseteq\mathbb{R}^{k},

δ(A)=sup{δB(A): finite Bk}=max{δB(A): finite Bk},\delta(A)=\sup\{\delta_{B}(A):\mbox{ finite }B\subseteq\mathbb{R}^{k}\}=\max\{\delta_{B}(A):\mbox{ finite }B\subseteq\mathbb{R}^{k}\},

since the supremum is actually attained when B=AB=A. ∎

3 Embedding into linear and sublinear diversities

We now turn our attention from linear and sublinear diversities to the questions of when finite diversities can be isometrically embedded within linear or sublinear diversities. Questions about embedding of metric spaces have, of course, been central to metric geometry and its applications, particularly after Linial et al. [18] demonstrated the link between approximate embeddings and combinatorial optimization algorithms on graphs. We showed in [7] that an analogous link holds between approximate embeddings of diversities and combinatorial optimization algorithms on hypergraphs. Here we only consider embeddings without distortion, that is, exact rather than approximate embeddings.

A map f:X1X2f:X_{1}\mapsto X_{2} between two diversities (X1,δ1)(X_{1},\delta_{1}) and (X2,δ2)(X_{2},\delta_{2}) is an isometric embedding if δ2(f(A))=δ1(A)\delta_{2}(f(A))=\delta_{1}(A) for all finite AX1A\subseteq X_{1}. We say that a finite diversity (X,δ)(X,\delta) is linear-embeddable if there is an isometric embedding from (X,δ)(X,\delta) to a linear diversity on k\mathbb{R}^{k} for some kk and sublinear-embeddable if there is an isometric embedding to some sublinear diversity on k\mathbb{R}^{k}, for some kk. At this stage we allow the dimension kk to be arbitrary.

Theorem 11 gives a characterization of linear-embeddability while Theorem 12 gives a characterization of sublinear-embeddability. Minkowski diversities and negative type diversities were reviewed earlier. We first establish a lemma on finite diversities that are embeddable in extremal linear diversities.

Lemma 10.

If (k,δ)(\mathbb{R}^{k},\delta) is an extremal linear semidiversity, and XkX\subseteq\mathbb{R}_{k} is finite then (X,δ)(X,\delta) is Minkowski embeddable with a simplex kernel.

Proof.

By Theorem 7 (k,δ)(\mathbb{R}^{k},\delta) is the Minkowski semidiversity with kernel K=conv(W)+HK=\mathrm{conv}(W)+H^{\perp}, where WW is a set of affinely independent vectors lying in a subspace HH. Let TT be an orthogonal matrix so that TH=span({e1,,ej})TH=\mathrm{span}(\{e_{1},\ldots,e_{j}\}) and TH=span({ej+1,,ej})TH^{\perp}=\mathrm{span}(\{e_{j+1},\ldots,e_{j}\}). Let THT_{H} be the first jj rows of TT, so that THH=jT_{H}H=\mathbb{R}^{j}, THH={0}T_{H}H^{\perp}=\{0\} and THK=conv(THW)T_{H}K=\mathrm{conv}(T_{H}W) is a full-dimensional simplex in j\mathbb{R}^{j}. Then for all λ\lambda, A+xλKA+x\subset\lambda K for some xkx\in\mathbb{R}^{k} if and only if THA+yλconv(THW)T_{H}A+y\subset\lambda\mathrm{conv}(T_{H}W) for some yjy\in\mathbb{R}^{j}. So δ(A)=δconv(THW)(THA)\delta(A)=\delta_{\mathrm{conv}(T_{H}W)}(T_{H}A) for all finite AA, as required. ∎

Theorem 11.

Let (X,δ)(X,\delta) be a finite diversity. The following are equivalent:

  1. (i)

    (X,δ)(X,\delta) is linear-embeddable.

  2. (ii)

    (X,δ)(X,\delta) has negative type.

  3. (iii)

    (X,δ)(X,\delta) can be embedded into a Minkowski diversity (k,δK)(\mathbb{R}^{k},\delta_{K}) with kernel equal to a simplex KkK\subseteq\mathbb{R}^{k}.

Proof.

(i)\Rightarrow(ii) Without loss of generality assume XkX\subseteq\mathbb{R}^{k} where (k,δ)(\mathbb{R}^{k},\delta) is a linear diversity. From Theorem 8 we have that any linear diversity can be expressed as a convex combination of extremal linear semidiversities. By Lemma 10 each of these extremal linear semidiversities can be expressed as a Minkowski diversity with a simplex, each of which has negative type by Theorem 17 in [3]. As the set of negative type diversities forms a convex cone, (X,δ)(X,\delta) also has negative type.
(ii)\Leftrightarrow(iii) This is Theorem 17 in [3].
(iii)\Rightarrow(i) Theorem 8 shows that any Minkowski diversity with a simplex kernel (being a trivial example of a weighted average of such diversities) is linear. Therefore, if (X,δ)(X,\delta) is embeddable in a Minkowski diversity with a simplex kernel, it also embeddable in a linear diversity. ∎

Theorem 12.

Let (X,δ)(X,\delta) be a finite diversity. The following are equivalent:

  1. (i)

    (X,δ)(X,\delta) is sublinear-embeddable.

  2. (ii)

    (X,δ)(X,\delta) can be embedded into a Minkowski diversity.

  3. (iii)

    (X,δ)(X,\delta) is the maximum of a collection of negative type diversities.

Proof.

(ii)\Rightarrow(i) If (X,δ)(X,\delta) is embeddable as a Minkowski diversity, then it is sublinear-embeddable, since by Theorem 2.4 of [3] all Minkowski diversities are sublinear.
(i)\Rightarrow(ii) Let (X,δ)(X,\delta) be a sublinear-embeddable. We may assume XX is a subset of k\mathbb{R}^{k} where (k,δ)(\mathbb{R}^{k},\delta) is a sublinear diversity. By Theorem 9 there is a family of linear semidiversities δγ\delta_{\gamma} for γΓ\gamma\in\Gamma such that δ(A)=maxδγ(A)\delta(A)=\max\delta_{\gamma}(A). Since XX is finite, it has a finite number of subsets, and so we may assume that Γ\Gamma is finite. By Proposition 4.1 (a) in [3] if two finite diversities are Minkowski embeddable, then so is their maximum, and hence the same is true of any finite number of finite Minkowski embeddable diversities. Therefore (X,δ)(X,\delta) is Minkowski embeddable.
(i)\Rightarrow(iii) We may assume XkX\subseteq\mathbb{R}^{k} where (k,δ)(\mathbb{R}^{k},\delta) is a sublinear diversity. By Theorem 9 there is a family of linear semidiversities δγ\delta_{\gamma} for γΓ\gamma\in\Gamma such that δ(A)=maxδγ(A)\delta(A)=\max\delta_{\gamma}(A). Since XX is finite, it has a finite number of subsets, and so we may assume that Γ\Gamma is finite. By Theorem 11, each of (X,δγ)(X,\delta_{\gamma}) is negative type, and therefore (X,δ)(X,\delta) is the maximum of a collection of negative type diversities.
(iii)\Rightarrow(ii) Suppose (X,δ)(X,\delta) is the maximum of a collection of negative type diversities. By Theorem 11 (X,δ)(X,\delta) can then be represented as the maximum of a collection of Minkowski diversities, and since XX is finite, we may assume the collection is finite. By Proposition 4.1 (a) in [3] the maximum of a finite collection of Minkowski embeddable diversities is Minkowski embeddable. ∎

References

  • [1] René Brandenberg and Stefan König. No dimension-independent core-sets for containment under homothetics. Discrete and Computational Geometry, 49:3–21, 2013.
  • [2] David Bryant, Petru Cioica-Licht, Lisa Orloff Clark, and Rachael Young. Inner products for convex bodies. Journal of Convex Analysis, 28(4):1249–1264, 2021.
  • [3] David Bryant, Katharina T Huber, Vincent Moulton, and Paul F Tupper. Diversities and the generalized circumradius. Discrete & Computational Geometry, pages 1–22, 2023.
  • [4] David Bryant, André Nies, and Paul F. Tupper. A universal separable diversity. Analysis and geometry in metric spaces, 5(1):138–151, 2017.
  • [5] David Bryant, André Nies, and Paul F. Tupper. Fraïssé limits for relational metric structures. Journal of Symbolic Logic, 86(3):913–934, 2021.
  • [6] David Bryant and Paul F Tupper. Hyperconvexity and tight-span theory for diversities. Advances in Mathematics, 231(6):3172–3198, 2012.
  • [7] David Bryant and Paul F. Tupper. Diversities and the geometry of hypergraphs. Discrete Mathematics and Theoretical Computer Science, 16(2):1–20, 2014.
  • [8] Michel Marie Deza and Monique Laurent. Geometry of Cuts and Metrics, volume 15 of Algorithms and Combinatorics. Springer-Verlag, Berlin, 1997.
  • [9] Rafa Espínola and Bożena Piatek. Diversities, hyperconvexity and fixed points. Nonlinear Analysis: Theory, Methods & Applications, 95:229–245, 2014.
  • [10] William J Firey. A functional characterization of certain mixed volumes. Israel Journal of Mathematics, 24:274–281, 1976.
  • [11] Benno Fuchssteiner and Wolfgang Lusky. Convex cones. North-Holland, Amsterdam, 1981.
  • [12] Pouya Haghmaram, Shohreh Golpaigani Fard, and Kourosh Nourouzi. Diversity-normed spaces and diversity embeddings. Studia Mathematica, 267:19–35, 2022.
  • [13] Pouya Haghmaram and Kourosh Nourouzi. Ultradiversification of diversities. Analysis and Geometry in Metric Spaces, 8:410–417, 2020.
  • [14] Andreas Hallbäck. Automorphism groups of universal diversities. Topology and its Applications, 285:107381, 2020.
  • [15] Andreas Hallbäck. Metric model theory, Polish groups & diversities. PhD thesis, Université de Paris, 2020.
  • [16] Piotr Indyk, Jiří Matoušek, and Anastasios Sidiropoulos. Low-distortion embeddings of finite metric spaces. In Handbook of discrete and computational geometry, pages 211–231. Chapman and Hall/CRC, 2017.
  • [17] Adam D Jozefiak and F Bruce Shepherd. Diversity embeddings and the hypergraph sparsest cut. arXiv preprint arXiv:2303.04199, 2023.
  • [18] Nathan Linial, Eran London, and Yuri Rabinovich. The geometry of graphs and some of its algorithmic applications. Combinatorica, 15(2):215–245, 1995.
  • [19] Jirı Matoušek. Lectures on discrete geometry. Springer Science & Business Media, 2013.
  • [20] Gholamreza H. Mehrabani and Kourosh Nourouzi. Ultradiversities and their spherical completeness. Journal of Applied Analysis, 26(2):231–240, 2020.
  • [21] M Meyer, G Mokobodzki, and M Rogalski. Convex bodies and concave functions. Proceedings of the American Mathematical Society, 123(2):477–484, 1995.
  • [22] B. Pia̧tek. On the gluing of hyperconvex metrics and diversities. Annales Universitatis Paedagogicae Cracoviensis. 149, Studia Mathematica, 13(1):65–76, 2014.
  • [23] Walter Rudin. Real and complex analysis. McGraw-Hill, 3rd edition, 1987.
  • [24] Rolf Schneider. Convex bodies: the Brunn–Minkowski theory. Encyclopedia of Mathematics and It Applications. Cambridge University Press, Cambridge, 2nd edition, 2014.
  • [25] Fedor Sergeevich Stonyakin. An analogue of the Hahn–Banach theorem for functionals on abstract convex cones. Eurasian Mathematical Journal, 7(3):89–99, 2016.
  • [26] Jürgen Voigt. A course on topological vector spaces. Springer, 2020.
  • [27] Pei Wu, David Bryant, and Paul F. Tupper. Negative-type diversities, a multi-dimensional analogue of negative-type metrics. The Journal of Geometric Analysis, 31:1703–1720, 2021.