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

Diversities and the Generalized Circumradius

David Bryant Department of Mathematics and Statistics, University of Otago, Dunedin, New Zealand. email: [email protected] phone: +64 34797889. (Corresponding author) Katharina T. Huber School of Computing Sciences, University of East Anglia, NR4 7TJ, Norwich, United Kingdom Vincent Moulton School of Computing Sciences, University of East Anglia, NR4 7TJ, Norwich, United Kingdom Paul F. Tupper Department of Mathematics, Simon Fraser University, Burnaby, British Columbia V5A 1S6, Canada
Abstract

The generalized circumradius of a set of points AdA\subseteq\mathbb{R}^{d} with respect to a convex body KK equals the minimum value of λ0\lambda\geq 0 such that a translate of λK\lambda K contains AA. Each choice of KK gives a different function on the set of bounded subsets of d\mathbb{R}^{d}; we characterize which functions can arise in this way. Our characterization draws on the theory of diversities, a recently introduced generalization of metrics from functions on pairs to functions on finite subsets. We additionally investigate functions which arise by restricting the generalized circumradius to a finite subset of d\mathbb{R}^{d}. We obtain elegant characterizations in the case that KK is a simplex or parallelotope.

The circumradius of a set of points in the plane is the radius of the smallest circle containing them. The concept is key to optimal containment and facility location problems, including a classic problem studied by Sylvester [21, 32], since the center of the smallest enclosing circle minimizes the maximum distance to any of the points.

The generalized circumradius replaces the plane with d\mathbb{R}^{d} and the circle or ball with a general convex body, that is a compact, convex set with non-empty interior. For a convex body KK in d\mathbb{R}^{d} and bounded AdA\subseteq\mathbb{R}^{d} we say that the generalized circumradius of AA with respect to KK is

R(A,K)=inf{λ0:AλK+x for some xd}.R(A,K)=\inf\{\lambda\geq 0:A\subseteq\lambda K+x\mbox{ for some $x\in\mathbb{R}^{d}$}\}.

In other words, R(A,K)R(A,K) equals the minimal amount that KK must be scaled so that a translate covers AA (see Fig. 1). The set KK is called the kernel. See [2, 3, 16, 22] for properties and inequalities related to the generalized circumradius and [4, 18, 17, 19] for computational results.

Refer to caption
Figure 1: An example of the generalized circumradius. In this example R({a,b,d,g,k},K)=2R(\{a,b,d,g,k\},K)=2 since 2K2K is the smallest scaled version of KK which can be translated to cover {a,b,d,g,k}\{a,b,d,g,k\}. Similarly, R({c,e,f},K)=0.6R(\{c,e,f\},K)=0.6, since 0.6K0.6K is the smallest scaled version of KK which can be translated to cover {c,e,f}\{c,e,f\}, and R({h,i},K)=1R(\{h,i\},K)=1.

Our motivation for studying the generalized circumradius comes from connections with diversity theory. A (mathematical) diversity is an extension of the idea of a metric space. Instead of assigning values just to pairs of objects, a diversity assigns values to all finite sets of objects. More formally, a diversity is a pair (X,δ)(X,\delta) where XX is a set and δ\delta a function from finite subsets of XX to 0\mathbb{R}_{\geq 0} such that, for A,B,CA,B,C finite,

  1. (D1)

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

  2. (D2)

    If BB\neq\emptyset then δ(AC)δ(AB)+δ(BC)\delta(A\cup C)\leq\delta(A\cup B)+\delta(B\cup C).

Diversities were introduced in [9]. A consequence of (D1) and (D2) is that diversities are monotonic, that is, if ABA\subseteq B then δ(A)δ(B)\delta(A)\leq\delta(B). Furthermore, δ\delta restricted to pairs satisfies the definition of a metric; we call this the metric induced by (X,δ)(X,\delta). We say that a diversity (X,δ)(X,\delta) is finite if XX is finite. Note that on occasion we use the term ‘diversity’ to refer to the function δ\delta rather than the pair (X,δ)(X,\delta).

Many well-known functions on sets are diversities. Examples include

  1. 1.

    The diameter of a set

  2. 2.

    The length of a shortest Steiner tree connecting a set

  3. 3.

    The mean width of a set

  4. 4.

    The length of the shortest traveling salesman tour through a set

  5. 5.

    The L1L_{1} diversity (in d\mathbb{R}^{d})

    δ(A)=δ1(A)=i=1dmax{|aiai|:a,aA}\delta(A)=\delta_{1}(A)=\sum_{i=1}^{d}\max\{|a_{i}-a^{\prime}_{i}|:a,a^{\prime}\in A\}
  6. 6.

    The circumradius of a set.

We show that the generalized circumradius is also a diversity:

Theorem 0.

Let KK be a convex body in d\mathbb{R}^{d}. If we define δ(A)=R(A,K)\delta(A)=R(A,K) for all finite AA then (d,δ)(\mathbb{R}^{d},\delta) is a diversity.

In reference to the concept of a Minkowski norm we say that a diversity (d,δ)(\mathbb{R}^{d},\delta) is a Minkowski diversity if there is a convex body KK such that δ(A)=R(A,K)\delta(A)=R(A,K) for all finite AdA\subset\mathbb{R}^{d}.

Theorem 5 connects the generalized circumradius to a growing and varied literature on diversity theory. The first diversity paper [9] described diversity analogs to the metric tight span and metric hyperconvexity, leading to new results in analysis and fixed point theory [23, 26, 27]. It was shown in [10] that results of [25] and others on embedding of finite metrics in L1L_{1} can be extended to diversities, with potential algorithmic gains. There is a direct analog of the Urysohn metric space [7] for diversities and work on diversity theory within model theory [8, 20], lattice theory [6, 33], image analysis [15], machine learning [24] and phylogenetics [5, 9, 30, 31].

In this paper we are mainly concerned with characterizations and embeddings for Minkowski diversities—what characterizes these diversities and which finite diversities can be embedded into a Minkowski diversity. Such embeddings (possibly with distortion) should in future provide valuable graphical representations of diversities in addition to algorithmic and computational tools.

Regarding characterization we prove the following result. A real-valued function ff on subsets of d\mathbb{R}^{d} is said to be sublinear if

f(A+B)\displaystyle f(A+B) f(A)+f(B)\displaystyle\leq f(A)+f(B) for all A,BA,B,
f(λA)\displaystyle f(\lambda A) =λf(A)\displaystyle=\lambda f(A) for all A and λ0,\displaystyle\mbox{for all $A$ and $\lambda\geq 0$},

where A+BA+B denotes the Minkowski sum.

Theorem 0.

Let (d,δ)(\mathbb{R}^{d},\delta) be a diversity. Then δ\delta is a Minkowski diversity if and only if δ\delta is sublinear and for all finite A,BA,B there are a,bda,b\in\mathbb{R}^{d} such that

δ((A+a)(B+b))max{δ(A),δ(B)}.\delta((A+a)\cup(B+b))\leq\max\{\delta(A),\delta(B)\}.

We also show that the last result extends beyond diversities to functions defined on bounded subsets of d\mathbb{R}^{d}.

Theorem 0.

Let ff be a function on bounded subsets of d\mathbb{R}^{d}. Then there is a convex body KK such that f(A)=R(A,K)f(A)=R(A,K) for all bounded AA if and only if ff is sublinear, monotonic, and ff restricted to finite subsets is a Minkowski diversity.

Having characterized which diversities on d\mathbb{R}^{d} are Minkowski diversities, we turn to the more difficult problem of characterizing which diversities can be isometrically embedded into Minkowski diversities. An isometric embedding of a diversity (X,δ1)(X,\delta_{1}) into a diversity (d,δ2)(\mathbb{R}^{d},\delta_{2}) is a map ϕ:Xd\phi:X\rightarrow\mathbb{R}^{d} such that δ1(A)=δ2(ϕ(A))\delta_{1}(A)=\delta_{2}(\phi(A)) for all finite AXA\subseteq X. If there is an isometric embedding from a diversity (X,δ1)(X,\delta_{1}) into a Minkowski diversity (d,δK)(\mathbb{R}^{d},\delta_{K}) for some dd and some convex body KdK\subseteq\mathbb{R}^{d} then we say that (X,δ1)(X,\delta_{1}) is Minkowski-embeddable.

We provide a complete characterization of Minkowski-embeddability for diversities which are finite and symmetric, meaning that the diversity of a set is determined by a function of its cardinality.

Theorem 0.

Let (X,δ)(X,\delta) be a finite symmetric diversity. Then (X,δ)(X,\delta) is Minkowski-embeddable if and only if

δ(A{a})δ(A)|A|2|A|1\frac{\delta(A\setminus\{a\})}{\delta(A)}\geq\frac{|A|-2}{|A|-1} (1)

for all AXA\subseteq X with |A|2|A|\geq 2, aAa\in A.

A consequence of the theorem is that, even though any diversity on three elements is Minkowski-embeddable, there exist diversities on four elements which are not.

We then investigate which finite diversities (X,δ1)(X,\delta_{1}) can be embedded into the diversity (d,δ2)(\mathbb{R}^{d},\delta_{2}) when δ2\delta_{2} is a Minkowski diversity with kernel KK restricted to a particular class. A diversity (X,δ)(X,\delta) is a diameter diversity if δ(A)=max{δ({a,a}):a,aA}\delta(A)=\max\{\delta(\{a,a^{\prime}\}):a,a^{\prime}\in A\} for all finite AXA\subseteq X, see [9]. The following characterization for when KK is a cube (or non-degenerate transform of a cube) follows from an observation of [2].

Theorem 0.

A finite diversity (X,δ)(X,\delta) can be embedded in a Minkowski diversity with kernel equal to some parallelotope if and only if (X,δ)(X,\delta) is a diameter diversity.

The case when KK is a simplex is more complex. We say that a finite diversity (X,δ)(X,\delta) is of negative type if

ABxAxBδ(AB)0\sum_{A\neq\emptyset}\sum_{B\neq\emptyset}x_{A}x_{B}\delta(A\cup B)\leq 0

for all zero-sum vectors xx indexed by non-empty subsets of XX. Diversities of negative type are analogous to metrics of negative type, and several of the properties of metrics of negative type extend to diversities of negative type, see [35]. For negative-type diversities we prove the following characterization.

Theorem 0.

A finite diversity (X,δ)(X,\delta) can be embedded in a Minkowski diversity with kernel equal to some simplex if and only if (X,δ)(X,\delta) has negative type.

Significantly, the set of diversities with negative type is quite large, with the same dimension as the set of diversities on XX. The result shows that the class of Minkowski-embeddable diversities is even larger, potentially opening up possibilities for quite general theoretical and algorithmic results.

1 The generalized circumradius

In this section we collect together a number of fundamental results about the generalized circumradius. We begin with several observations from [21] (Proposition 3.2). Let conv(A)\mathrm{conv}(A) denote the convex hull of AA.

Proposition 1.

Let A,A,BA,A^{\prime},B be bounded subsets of d\mathbb{R}^{d}, let K,KK,K^{\prime} be convex bodies in d\mathbb{R}^{d}. Then

  1. (a)

    If AAA\subseteq A^{\prime} and KKK^{\prime}\subseteq K then R(A,K)R(A,K)R(A,K)\leq R(A^{\prime},K^{\prime}).

  2. (b)

    If conv(A)=conv(A)\mathrm{conv}(A)=\mathrm{conv}(A^{\prime}) then R(A,K)=R(A,K)R(A,K)=R(A^{\prime},K).

  3. (c)

    R(A+B,K)R(A,K)+R(B,K)R(A+B,K)\leq R(A,K)+R(B,K) with equality if B=αKB=\alpha K for some α0\alpha\geq 0.

  4. (d)

    R(A+x,K+y)=R(A,K)R(A+x,K+y)=R(A,K) for all x,ydx,y\in\mathbb{R}^{d}.

  5. (e)

    R(αA,βK)=(α/β)R(A,K)R(\alpha A,\beta K)=(\alpha/\beta)R(A,K) for α,β>0\alpha,\beta>0.

An indirect consequence of Helly’s theorem (see e.g. [13], Section 6.2) is that for bounded AdA\subseteq\mathbb{R}^{d} and a convex body KK we can find a small subset AAA^{\prime}\subseteq A such that |A|d+1|A^{\prime}|\leq d+1 and R(A,K)=R(A,K)R(A^{\prime},K)=R(A,K). The following more general result forms one part of Theorem 1.2 in [2].

Proposition 2.

Suppose that AdA\subset\mathbb{R}^{d} is bounded and KK is a convex body. For all ϵ0\epsilon\geq 0 there exists AAA^{\prime}\subseteq A such that |A|d1+ϵ+1|A^{\prime}|\leq\left\lceil\frac{d}{1+\epsilon}\right\rceil+1 and

R(A,K)R(A,K)(1+ϵ)R(A,K).R(A^{\prime},K)\leq R(A,K)\leq(1+\epsilon)R(A^{\prime},K).

Note that for particular choices of KK there can be much smaller bounds on |A||A^{\prime}|. For example, when KK is the Euclidean ball in d\mathbb{R}^{d} we have for all bounded AdA\subset\mathbb{R}^{d} and ϵ>0\epsilon>0 that there is a subset AAA^{\prime}\subseteq A such that R(A,K)(1+ϵ)R(A,K)R(A,K)\leq(1+\epsilon)R(A^{\prime},K) and

|A|12ϵ+ϵ2+1.|A^{\prime}|\leq\left\lceil\frac{1}{2\epsilon+\epsilon^{2}}\right\rceil+1.

This bound is independent of the dimension dd.

We will make use of the following general property for Minkowski addition which is established during the proof of Theorem 4.1 in [2]. Let AdA\subseteq\mathbb{R}^{d} be any set with cardinality k+1k+1, k2k\geq 2, and zero sum. Then

Ak(k+1)(k1)aAconv(A{a}).A\subseteq\frac{k}{(k+1)(k-1)}\sum_{a\in A}\mathrm{conv}(A\setminus\{a\}).

Combining this observation with Proposition 1 (d), (b) and (c) we have

Proposition 3.

Suppose AdA\subset\mathbb{R}^{d}, |A|=k+1|A|=k+1 and KK is a convex body. Then

R(A,K)k(k+1)(k1)aAR(A{a},K).R(A,K)\leq\frac{k}{(k+1)(k-1)}\sum_{a\in A}R(A\setminus\{a\},K).

For Cp×qC\subseteq\mathbb{R}^{p}\times\mathbb{R}^{q} we define the two projection operators

π1(C)\displaystyle\pi_{1}(C) ={c1p:(c1,c2)C for some c2}\displaystyle=\{c_{1}\in\mathbb{R}^{p}:(c_{1},c_{2})\in C\mbox{ for some $c_{2}$}\}
π2(C)\displaystyle\pi_{2}(C) ={c2q:(c1,c2)C for some c1}.\displaystyle=\{c_{2}\in\mathbb{R}^{q}:(c_{1},c_{2})\in C\mbox{ for some $c_{1}$}\}.

The following result will be useful for questions regarding embeddings.

Proposition 4.

Let AA be a bounded subset of p×q\mathbb{R}^{p}\times\mathbb{R}^{q}. If BB and CC are convex bodies in p\mathbb{R}^{p} and q\mathbb{R}^{q} respectively, then

R(A,B×C)=max{R(π1(A),B),R(π2(A),C)}.R(A,B\times C)=\max\{R(\pi_{1}(A),B),R(\pi_{2}(A),C)\}.
Proof.

Suppose λ=R(A,B×C)\lambda=R(A,B\times C) and that

A+(a1,a2)λ(B×C).A+(a_{1},a_{2})\subseteq\lambda(B\times C).

Applying π1\pi_{1} and π2\pi_{2} to both sides gives

π1(A)+a1λB\pi_{1}(A)+a_{1}\subseteq\lambda B

and

π2(A)+a2λC.\pi_{2}(A)+a_{2}\subseteq\lambda C.

Hence max{R(π1(A),B),R(π2(A),C)}R(A,B×C)\max\{R(\pi_{1}(A),B),R(\pi_{2}(A),C)\}\leq R(A,B\times C).

Conversely, if λ=max{R(π1(A),B),R(π2(A),C)}\lambda=\max\{R(\pi_{1}(A),B),R(\pi_{2}(A),C)\}, then there are a1,a2a_{1},a_{2} such that

π1(A)+a1\displaystyle\pi_{1}(A)+a_{1} λB\displaystyle\subseteq\lambda B
π2(A)+a2\displaystyle\pi_{2}(A)+a_{2} λC.\displaystyle\subseteq\lambda C.

We then have

A+(a1,a2)\displaystyle A+(a_{1},a_{2}) π1(A)×π2(A)+(a1,a2)\displaystyle\subseteq\pi_{1}(A)\times\pi_{2}(A)+(a_{1},a_{2})
=(π1(A)+a1)×(π2(A)+a2)\displaystyle=(\pi_{1}(A)+a_{1})\times(\pi_{2}(A)+a_{2})
λ(B×C)\displaystyle\subseteq\lambda(B\times C)

so that R(A,B×C)max{R(π1(A),B),R(π2(A),C)}R(A,B\times C)\leq\max\{R(\pi_{1}(A),B),R(\pi_{2}(A),C)\}. ∎

2 Characterization of Minkowski diversities

We begin this section by proving the first main result connecting the theory of generalized circumradii with diversity theory.

Theorem 5.

Let KK be a convex body in d\mathbb{R}^{d}. If δ(A)=R(A,K)\delta(A)=R(A,K) for all finite AA, then (d,δ)(\mathbb{R}^{d},\delta) is a diversity.

Proof.

Clearly δ(A)0\delta(A)\geq 0 for all AA and δ(A)=R(A,K)=0\delta(A)=R(A,K)=0 if and only if |A|1|A|\leq 1. Hence (D1) holds. By Proposition 1 (a), R(A,K)R(A,K) is monotonic in AA.

Suppose AA and BB are finite subsets of d\mathbb{R}^{d} and xABx\in A\cap B. Then (Ax)=(Ax)+0(Ax)+(Bx)(A-x)=(A-x)+0\subseteq(A-x)+(B-x) and (Bx)=0+(Bx)(Ax)+(Bx)(B-x)=0+(B-x)\subseteq(A-x)+(B-x). Hence

(Ax)(Bx)(Ax)+(Bx),(A-x)\cup(B-x)\subseteq(A-x)+(B-x),

and so by Proposition 1 (d) and (c),

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

This fact, together with monotonicity, implies the diversity triangle inequality (D2).

In the rest of this section we focus on characterizing which diversities on d\mathbb{R}^{d} can be obtained in this way (Theorem 8).

Let 𝔅\mathfrak{B} denote the Euclidean unit ball in d\mathbb{R}^{d}. The Hausdorff distance on (bounded) subsets of d\mathbb{R}^{d} is given by

dH(A,B)=inf{λ:AB+λ𝔅 and BA+λ𝔅},d_{H}(A,B)=\inf\{\lambda:A\subseteq B+\lambda\mathfrak{B}\mbox{ and }B\subseteq A+\lambda\mathfrak{B}\},

see [29]. We note that dHd_{H} becomes a metric when restricted to compact sets.

A diversity (d,δ)(\mathbb{R}^{d},\delta) is sublinear if δ\delta is a sublinear function, δ(A+B)δ(A)+δ(B)\delta(A+B)\leq\delta(A)+\delta(B) and δ(λA)=λδ(A)\delta(\lambda A)=\lambda\delta(A) for all finite A,BA,B and non-negative λ\lambda. Note that finite sublinear diversities are closely related to set-norms [12, Def. 2.1]. The L1L_{1} diversity (d,δ1)(\mathbb{R}^{d},\delta_{1}), as introduced earlier, is an example of a sublinear diversity. Here,

δ1(A)=i=1dmax{|aiai|:a,aA},\delta_{1}(A)=\sum_{i=1}^{d}\max\{|a_{i}-a^{\prime}_{i}|:a,a^{\prime}\in A\},

and δ1\delta_{1} is sublinear since, given λ0\lambda\geq 0 and finite subsets A,BdA,B\subseteq\mathbb{R}^{d} we have

δ1(λA)\displaystyle\delta_{1}(\lambda A) =i=1dmax{|λaiλai|:a,aA}\displaystyle=\sum_{i=1}^{d}\max\{|\lambda a_{i}-\lambda a^{\prime}_{i}|:a,a^{\prime}\in A\}
=λδ1(A),\displaystyle=\lambda\delta_{1}(A),
and
δ1(A+B)\displaystyle\delta_{1}(A+B) =i=1dmax{|ai+biaibi|:a,aA,b,bB}\displaystyle=\sum_{i=1}^{d}\max\{|a_{i}+b_{i}-a^{\prime}_{i}-b^{\prime}_{i}|:a,a^{\prime}\in A,b,b^{\prime}\in B\}
i=1dmax{|aiai|:a,aA}+max{|bibi|:b,bB}\displaystyle\leq\sum_{i=1}^{d}\max\{|a_{i}-a^{\prime}_{i}|:a,a^{\prime}\in A\}+\max\{|b_{i}-b^{\prime}_{i}|:b,b^{\prime}\in B\}
=δ1(A)+δ1(B).\displaystyle=\delta_{1}(A)+\delta_{1}(B).

Sublinearity has several important consequences for diversities.

Proposition 6.

Let (d,δ)(\mathbb{R}^{d},\delta) be a sublinear diversity. Then

  1. (a)

    δ\delta is translation invariant: δ(A+x)=δ(A)\delta(A+x)=\delta(A) for all xx.

  2. (b)

    δ\delta is determined by the convex hull: if A,BA,B finite in d\mathbb{R}^{d} and conv(A)=conv(B)\mathrm{conv}(A)=\mathrm{conv}(B), then δ(A)=δ(B)\delta(A)=\delta(B).

  3. (c)

    δ\delta is Lipschitz continuous with respect to the Hausdorff metric.

Proof.

(a) By (D1) and sublinearity, 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).

(b) Let BBB^{\prime}\subseteq B be a maximal subset of BB such that δ(AB)=δ(A)\delta(A\cup B^{\prime})=\delta(A). Suppose that there is bBBb\in B\setminus B^{\prime}. As bconv(B)=conv(A)b\in\mathrm{conv}(B)=\mathrm{conv}(A) there are non-negative {λa}aA\{\lambda_{a}\}_{a\in A} with unit sum such that

b=aλaab=\sum_{a}\lambda_{a}a

and hence

baλaA.b\in\sum_{a}\lambda_{a}A.

As

ABaλa(AB)A\cup B^{\prime}\subseteq\sum_{a}\lambda_{a}(A\cup B^{\prime})

we have

δ(AB{b})δ(aλa(AB))aλaδ(AB)=δ(AB)=δ(A).\delta(A\cup B^{\prime}\cup\{b\})\leq\delta\Big{(}\sum_{a}\lambda_{a}(A\cup B^{\prime})\Big{)}\leq\sum_{a}\lambda_{a}\delta(A\cup B^{\prime})=\delta(A\cup B^{\prime})=\delta(A).

This contradicts the choice of BB^{\prime}. Hence B=BB^{\prime}=B and δ(AB)=δ(A)\delta(A\cup B)=\delta(A). Exchanging AA and BB in this argument gives δ(AB)=δ(B)\delta(A\cup B)=\delta(B). Hence δ(A)=δ(B)\delta(A)=\delta(B).
(c) Let VV denote any finite set with convex hull conv(V)\mathrm{conv}(V) containing the unit ball 𝔅\mathfrak{B}. One example is V={1,1}dV=\{-1,1\}^{d}. Let κ=δ(V)\kappa=\delta(V) and suppose that dH(A,B)=λd_{H}(A,B)=\lambda. From the definition of dHd_{H} we have AB+λ𝔅A\subseteq B+\lambda\mathfrak{B} and BA+λ𝔅B\subseteq A+\lambda\mathfrak{B} so there are finite sets A,B𝔅A^{\prime},B^{\prime}\subseteq\mathfrak{B} such that

AB+λA and BA+λB.A\subseteq B+\lambda A^{\prime}\mbox{ and }B\subseteq A+\lambda B^{\prime}.

As 𝔅conv(V)\mathfrak{B}\subseteq\mathrm{conv}(V), we have conv(AV)=conv(BV)=conv(V)\mathrm{conv}(A^{\prime}\cup V)\!=\!\mathrm{conv}(B^{\prime}\cup V)\!=\!\mathrm{conv}(V). By sublinearity, monotonicity of δ\delta and part (b) we therefore have

δ(A)δ(B)+λδ(V) and δ(B)δ(A)+λδ(V)\delta(A)\leq\delta(B)+\lambda\delta(V)\mbox{ and }\delta(B)\leq\delta(A)+\lambda\delta(V)

giving |δ(A)δ(B)|κdH(A,B)|\delta(A)-\delta(B)|\leq\kappa d_{H}(A,B). ∎

Our next theorem gives a complete characterization of Minkowski diversities. A key idea in the proof of the theorem is that we can extend a sublinear diversity (d,δ)(\mathbb{R}^{d},\delta) to a function on convex bodies in d\mathbb{R}^{d}. More specifically, given a sublinear diversity δ\delta, define the function δ~\tilde{\delta} on the set of convex bodies in d\mathbb{R}^{d} by setting

δ~(P)=δ(Vert(P))\tilde{\delta}(P)=\delta(\mathrm{Vert}(P))

for all polytopes PP with vertex set Vert(P)\mathrm{Vert}(P) and

δ~(K)=limnδ~(Pn)\tilde{\delta}(K)=\ \lim_{n\rightarrow\infty}\tilde{\delta}(P_{n})

for any convex body KK and sequence P1,P2,P_{1},P_{2},\ldots of polytopes converging under the Hausdorff metric to KK.

Lemma 7.

Given a sublinear diversity δ\delta, the function δ~\tilde{\delta} is well-defined and Lipschitz continuous.

Proof.

Since δ~\tilde{\delta} is Lipschitz on the set of polytopes, it is uniformly continuous on the same set, and so can be uniquely extended to a continuous function on the closure of that set [28, Prob. 13, Ch. 4], the convex bodies. An expression for δ~\tilde{\delta} can then be obtain for convex bodies using the limits of sequences of polytopes as above, and this gives that the extension has the same Lipschitz constant. ∎

Using this observation, we now prove the main theorem for this section.

Theorem 8.

Let (d,δ)(\mathbb{R}^{d},\delta) be a diversity. Then (d,δ)(\mathbb{R}^{d},\delta) is a Minkowski diversity if and only if δ\delta is sublinear and for all finite A,BA,B there are a,bda,b\in\mathbb{R}^{d} such that

δ((A+a)(B+b))max{δ(A),δ(B)}.\delta((A+a)\cup(B+b))\leq\max\{\delta(A),\delta(B)\}. (2)
Proof.

Suppose that (d,δ)(\mathbb{R}^{d},\delta) is a Minkowski diversity, so that there is a convex body KdK\subseteq\mathbb{R}^{d} such that δ(A)=R(A,K)\delta(A)=R(A,K) for all finite AdA\subseteq\mathbb{R}^{d}. Sublinearity is given by Proposition 1 parts (c) and (e). Given finite AA and BB, there are a,bda,b\in\mathbb{R}^{d} such that Aδ(A)KaA\subseteq\delta(A)K-a and Bδ(B)KbB\subseteq\delta(B)K-b. Hence

(A+a)(B+b)max{δ(A),δ(B)}K(A+a)\cup(B+b)\subseteq\max\{\delta(A),\delta(B)\}K

and

δ((A+a)(B+b))\displaystyle\delta((A+a)\cup(B+b)) =R((A+a)(B+b),K)\displaystyle=R((A+a)\cup(B+b),K)
R(max{δ(A),δ(B)}K,K)\displaystyle\leq R(\max\{\delta(A),\delta(B)\}K,K)
=max{δ(A),δ(B)}R(K,K)\displaystyle=\max\{\delta(A),\delta(B)\}R(K,K)
=max{δ(A),δ(B)}.\displaystyle=\max\{\delta(A),\delta(B)\}.

Now suppose that (d,δ)(\mathbb{R}^{d},\delta) is sublinear and satisfies (2) for all finite A,BA,B. Let

ρ=max{xδ({0,x}):x0}=max{1δ({0,x}):x=1}.\rho=\max\left\{\frac{\|x\|}{\delta(\{0,x\})}:x\neq 0\right\}=\max\left\{\frac{1}{\delta(\{0,x\})}:\|x\|=1\right\}.

Then 0<ρ<0<\rho<\infty and xyρδ({x,y})\|x-y\|\leq\rho\delta(\{x,y\}) for all x,yx,y. (Here and below we use \|\cdot\| to denote Euclidean norm).

We show that for any convex bodies L,KdL,K\subseteq\mathbb{R}^{d} and the function δ~\tilde{\delta} in Lemma 7 there is some xdx\in\mathbb{R}^{d} such that

δ~(conv(K(L+x)))max{δ~(K),δ~(L)}.\tilde{\delta}(\mathrm{conv}(K\cup(L+x)))\leq\max\{\tilde{\delta}(K),\tilde{\delta}(L)\}. (3)

Suppose that K1,K2,K_{1},K_{2},\ldots is a sequence of finite subsets of KK such that conv(Kn)K\mathrm{conv}(K_{n})\rightarrow K and L1,L2,L_{1},L_{2},\ldots is a sequence of finite subsets of LL such that conv(Ln)L\mathrm{conv}(L_{n})\rightarrow L. Applying (2) and translation invariance of δ\delta we have that for each nn\in\mathbb{N} there is an xndx_{n}\in\mathbb{R}^{d} such that

δ(Kn(Ln+xn))max{δ(Kn),δ(Ln)}.\delta(K_{n}\cup(L_{n}+x_{n}))\leq\max\{\delta(K_{n}),\delta(L_{n})\}.

Hence, for all nn,

δ~(conv(Kn(Ln+xn)))max{δ~(conv(Kn)),δ~(conv(Ln))}.\tilde{\delta}(\mathrm{conv}(K_{n}\cup(L_{n}+x_{n})))\leq\max\{\tilde{\delta}(\mathrm{conv}(K_{n})),\tilde{\delta}(\mathrm{conv}(L_{n}))\}. (4)

We show that the sequence xnx_{n} has a convergent subsequence. First note that since KK and LL are bounded, convergence of conv(Kn)\mathrm{conv}(K_{n}) and conv(Ln)\mathrm{conv}(L_{n}) to them in the Hausdorff metric implies that the union of all these sets is bounded, and hence the sequence max{δ(Kn),δ(Ln)}\max\{\delta(K_{n}),\delta(L_{n})\} is bounded. Choose knKnk_{n}\in K_{n} and nLn\ell_{n}\in L_{n} for each nn, which are then bounded over all nn. We then have

xn(knn)=(xn+n)knρδ(Kn(Ln+xn))max{δ(Kn),δ(Ln)}.\|x_{n}-(k_{n}-\ell_{n})\|=\|(x_{n}+\ell_{n})-k_{n}\|\leq\rho\delta(K_{n}\cup(L_{n}+x_{n}))\leq\max\{\delta(K_{n}),\delta(L_{n})\}.

So the set {x1,x2,x3,}\{x_{1},x_{2},x_{3},\ldots\} is bounded. Let xi1,xi2,x_{i_{1}},x_{i_{2}},\ldots be a convergence subsequence and let xdx\in\mathbb{R}^{d} be its limit.

Since conv(Ln)L\mathrm{conv}(L_{n})\rightarrow L and xnxx_{n}\rightarrow x, conv(Ln+xn)L+x\mathrm{conv}(L_{n}+x_{n})\rightarrow L+x. So, conv(Kn)conv(Ln+xn)K(L+x)\mathrm{conv}(K_{n})\cup\mathrm{conv}(L_{n}+x_{n})\rightarrow K\cup(L+x) and

conv(Kn(Ln+xn))=conv(conv(Kn)conv(Ln+xn))conv(K(L+x)).\mathrm{conv}(K_{n}\cup(L_{n}+x_{n}))=\mathrm{conv}(\mathrm{conv}(K_{n})\cup\mathrm{conv}(L_{n}+x_{n}))\rightarrow\mathrm{conv}(K\cup(L+x)).

Taking the limit as nn\rightarrow\infty of (4) and using the continuity of δ~\tilde{\delta} gives (3) which proves the claim.

Let \mathfrak{C} denote the set of convex bodies

={Ad:δ~(A)1}.\mathfrak{C}=\{A\subseteq\mathbb{R}^{d}:\tilde{\delta}(A)\leq 1\}.

The set \mathfrak{C} is closed under the Hausdorff metric and both volume and δ~\tilde{\delta} are continuous with respect to the Hausdorff metric [29, Sec 1.8]. It follows that there is some KK\in\mathfrak{C} such that the volume of KK is at least as large as any other element in \mathfrak{C}. The convex body KK is necessarily inclusion-maximum: if KK was a proper subset of some KK^{\prime}\in\mathfrak{C} then the volume of KK^{\prime} would be strictly greater than the volume of KK.

We claim that R(A,K)=1R(A,K)=1 for all finite AA such that δ(A)=1\delta(A)=1.

Let AA be finite with δ(A)=1\delta(A)=1. By (3) there is xdx\in\mathbb{R}^{d} such that

δ~(conv(K(conv(A)+x)))max{δ~(K),δ~(conv(A))}=1.\tilde{\delta}(\mathrm{conv}(K\cup(\mathrm{conv}(A)+x)))\leq\max\{\tilde{\delta}(K),\tilde{\delta}(\mathrm{conv}(A))\}=1.

We therefore have conv(K(conv(A)+x))\mathrm{conv}(K\cup(\mathrm{conv}(A)+x))\in\mathfrak{C}. As KK is set inclusion-maximum in \mathfrak{C} and Kconv(K(conv(A)+x))K\subseteq\mathrm{conv}(K\cup(\mathrm{conv}(A)+x)) we have

K=conv(K(conv(A)+x))K=\mathrm{conv}(K\cup(\mathrm{conv}(A)+x))

and so A+xKA+x\subseteq K and R(A,K)1R(A,K)\leq 1. On the other hand, if there is some λ0\lambda\geq 0 and zdz\in\mathbb{R}^{d} such that AλK+zA\subseteq\lambda K+z then

δ(A)=δ~(conv(A))δ~(λK+z)=λδ~(K)=λ,\delta(A)=\tilde{\delta}(\mathrm{conv}(A))\leq\tilde{\delta}(\lambda K+z)=\lambda\tilde{\delta}(K)=\lambda,

showing that R(A,K)1R(A,K)\geq 1. Hence R(A,K)=1R(A,K)=1, as claimed.

It follows that δ(A)=R(A,K)\delta(A)=R(A,K) when δ(A)=1\delta(A)=1.

More generally, suppose δ(A)=d\delta(A)=d. The case d=0d=0 is straightforward, as then |A|1|A|\leq 1. If d>0d>0 then, by sublinearity, δ(A)=dδ(1dA)=dR(1dA,K)=R(A,K)\delta(A)=d\delta(\frac{1}{d}A)=dR(\frac{1}{d}A,K)=R(A,K). ∎

We note that there are diversities on d\mathbb{R}^{d} which are sublinear, but do not satisfy property (2) in Theorem 8 (and are hence not Minkowski diversities). For example, consider the L1L_{1} diversity in the plane (2,δ1)(\mathbb{R}^{2},\delta_{1}). We saw above that L1L_{1} diversities are sublinear but if

A={[00],[10]} and B={[00],[01]}A=\left\{\left[\begin{matrix}0\\ 0\end{matrix}\right],\left[\begin{matrix}1\\ 0\end{matrix}\right]\right\}\mbox{ and }B=\left\{\left[\begin{matrix}0\\ 0\end{matrix}\right],\left[\begin{matrix}0\\ 1\end{matrix}\right]\right\}

then for any a,b2a,b\in\mathbb{R}^{2} we have

δ1((A+a)(B+b))2>1=max{δ1(A),δ1(B)}.\delta_{1}((A+a)\cup(B+b))\geq 2>1=\max\{\delta_{1}(A),\delta_{1}(B)\}.

Refer to caption
Figure 2: The sets A,B,BA,B,B^{\prime} in the proof of Proposition 9 with (translated) kernels KK and KK^{\prime}.

The set of diversities on d\mathbb{R}^{d}, and indeed the the set of sublinear diversities on d\mathbb{R}^{d}, are both convex. Condition (2) in Theorem 8 suggests that the set of Minkowski diversities on d\mathbb{R}^{d} is not convex, as we now confirm.

Proposition 9.

The set of Minkowski diversities on d\mathbb{R}^{d}, d2d\geq 2, is not convex.

Proof.

We first establish this for d=2d=2. Let

A\displaystyle A ={[10],[01],[11]},\displaystyle=\left\{\left[\begin{matrix}1\\ 0\end{matrix}\right],\left[\begin{matrix}0\\ 1\end{matrix}\right],\left[\begin{matrix}1\\ 1\end{matrix}\right]\right\},
B\displaystyle B ={[10],[20],[11]},\displaystyle=\left\{\left[\begin{matrix}1\\ 0\end{matrix}\right],\left[\begin{matrix}2\\ 0\end{matrix}\right],\left[\begin{matrix}1\\ 1\end{matrix}\right]\right\},
B\displaystyle B^{\prime} ={[01],[11],[02]}.\displaystyle=\left\{\left[\begin{matrix}0\\ 1\end{matrix}\right],\left[\begin{matrix}1\\ 1\end{matrix}\right],\left[\begin{matrix}0\\ 2\end{matrix}\right]\right\}.

Let K=conv(AB)K=\mathrm{conv}(A\cup B) and K=conv(AB)K^{\prime}=\mathrm{conv}(A\cup B^{\prime}). Let (X,δ)(X,\delta) be the diversity on 2\mathbb{R}^{2} given by δ(Y)=12R(Y,K)+12R(Y,K)\delta(Y)=\mbox{$\frac{1}{2}$}R(Y,K)+\mbox{$\frac{1}{2}$}R(Y,K^{\prime}). We will show that for all a,b2a,b\in\mathbb{R}^{2}

δ((A+a)(B+b))>max{δ(A),δ(B)},\delta((A+a)\cup(B+b))>\max\{\delta(A),\delta(B)\},

from which it follows by Theorem 8 that (X,δ)(X,\delta) is not a Minkowski diversity.

First note that since δ\delta is translation invariant, we can assume a=0a=0. Now, note that

R(A,K)=R(A,K)=R(B,K)=R(B,K)=1R(A,K)=R(A,K^{\prime})=R(B,K)=R(B,K^{\prime})=1

so that

max{δ(A),δ(B)}=1.\max\{\delta(A),\delta(B)\}=1.

If b=0b=0 then R(A(B+b),K)=1R(A\cup(B+b),K)=1, otherwise R(A(B+b),K)>1R(A\cup(B+b),K)>1. If b=[11]b=\left[\begin{matrix}-1\\ 1\end{matrix}\right] then R(A(B+b),K)=1R(A\cup(B+b),K^{\prime})=1, otherwise R(A(B+b),K)>1R(A\cup(B+b),K^{\prime})>1. Hence we have

δ((A+a)(B+b))=12R(A(B+b),K)+12R(A(B+b),K)>1\delta((A+a)\cup(B+b))=\mbox{$\frac{1}{2}$}R(A\cup(B+b),K)+\mbox{$\frac{1}{2}$}R(A\cup(B+b),K^{\prime})>1

even though max{δ(A),δ(B)}=1\max\{\delta(A),\delta(B)\}=1.

For d>2d>2, in the above argument we replace KK and KK^{\prime} with their product with the unit hypercube in d2\mathbb{R}^{d-2}, and append d2d-2 zeros to the elements in AA and BB and to xx. ∎

3 Characterizing the general circumradius

In this section, we characterize functions ff for which there is a convex body KK such that f(A)=R(A,K)f(A)=R(A,K) for all bounded AA, noting that in the previous section we only considered finite AA. The main idea behind the proof is to show that a sublinear, monotonic function is Hausdorff continuous, after which the result follows almost immediately from Theorem 8.

Lemma 10.

Let ff be a sublinear, monotonic function on bounded subsets of d\mathbb{R}^{d}. Then ff is Hausdorff continuous.

Proof.

Let 𝔅\mathfrak{B} denote the Euclidean unit ball in d\mathbb{R}^{d}, let b=f(𝔅)b=f(\mathfrak{B}) and let ϵ>0\epsilon>0. Then for any bounded A,BA,B such that dH(A,B)ϵbd_{H}(A,B)\leq\frac{\epsilon}{b} we have

f(A)f(B+ϵb𝔅)f(B)+ϵf(A)\leq f\left(B+\frac{\epsilon}{b}\mathfrak{B}\right)\leq f(B)+\epsilon

and

f(B)f(A+ϵb𝔅)f(A)+ϵ.f(B)\leq f\left(A+\frac{\epsilon}{b}\mathfrak{B}\right)\leq f(A)+\epsilon.

Theorem 11.

Let ff be a function on bounded subsets of d\mathbb{R}^{d}. Then there is a convex body KK such that f(A)=R(A,K)f(A)=R(A,K) for all bounded AA if and only if ff is sublinear, monotonic, and ff restricted to finite subsets is a Minkowski diversity.

Proof.

Necessity follows from the arguments used for Theorem 8.

Suppose that ff is sublinear and monotonic, and ff restricted to finite subsets is a Minkowski diversity. By Lemma 10, ff is Hausdorff continuous. Let AA be a bounded subset of d\mathbb{R}^{d}. Then f(A¯)=f(A)f(\overline{A})=f(A), where A¯\overline{A} denotes the topological closure of AA.

By Theorem 8 there is a convex body KK such that f(B)=R(B,K)f(B)=R(B,K) for any finite BB. Given any natural number n1n\geq 1 there is a finite cover of A¯\overline{A} by balls of radius 1n\frac{1}{n}. Let AnA_{n} denote the set of centers of those balls, so that AnA¯A_{n}\rightarrow\overline{A}. We then have

f(A)=f(A¯)=limnf(An)=limnR(An,K)=R(A,K).f(A)=f(\overline{A})=\lim_{n\rightarrow\infty}f(A_{n})=\lim_{n\rightarrow\infty}R(A_{n},K)=R(A,K).

4 Embedding finite diversities

In this section we consider an (isometric) embedding problem: when can a given finite diversity (X,δ)(X,\delta) be embedded in a Minkowski diversity? There is a long history in mathematics regarding embedding of finite metrics into standard spaces. Perhaps best known is the characterization due to Cayley and Menger of when a finite metric can be embedded into Euclidean space [1, 14]. The theory of metric embeddings forms the basis of many methods for multi-dimensional scaling, an approximate low-dimensional embedding designed specifically for data reduction and representation. Approximate embeddings have proven exceptionally useful for algorithm design and approximations (e.g. [25]), work that has a direct analog in the mathematics of diversities [10].

To discuss embeddings, it is convenient to consider a slight generalization of diversities. A semimetric is a bivariate, symmetric map dd on XX that vanishes on the diagonal and satisfies the triangle inequality, but where we allow d(x,y)=0d(x,y)=0 even when xyx\neq y, x,yXx,y\in X (so, in particular a metric is a semimetric). Similarly, a pair (X,δ)(X,\delta) is a semidiversity if it satisfies (D2) and the following slightly weaker version of (D1)

  1. (D1’)

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

We say that a (semi)diversity (X,δ)(X,\delta) is Minkowski-embeddable if for some dd there is a map ϕ:Xd\phi:X\rightarrow\mathbb{R}^{d} and a convex body KK in d\mathbb{R}^{d} such that

δ(A)=R(ϕ(A),K)\delta(A)=R(\phi(A),K)

for all finite AXA\subseteq X.

For the rest of this section we shall focus on the embedding problem for symmetric diversities, where a (semi)diversity (X,δ)(X,\delta) is symmetric if δ(A)=δ(B)\delta(A)=\delta(B) whenever |A|=|B||A|=|B|, A,BXA,B\subseteq X, that is, the value of δ\delta on a set depends only upon the cardinality of the set, see [11]. We shall characterize when a finite symmetric diversity is Minkowski-embeddable. As a corollary we also show that not every diversity is Minkowski-embeddable

We start with some utility results on embeddings. For convenience, for the rest of this section we shall assume that XX is a finite set. Note that if (X,δ1)(X,\delta_{1}) and (X,δ2)(X,\delta_{2}) are two semidiversities then (X,δ1δ2)(X,\delta_{1}\vee\delta_{2}) denotes the semidiversity with

(δ1δ2)(A)=max{δ1(A),δ2(A)}(\delta_{1}\vee\delta_{2})(A)=\max\{\delta_{1}(A),\delta_{2}(A)\}

for all AXA\subseteq X. To see that this is a semidiversity, note that for all A,B,CA,B,C with BB\neq\emptyset we have without loss of generality,

(δ1δ2)(AC)=δ1(AC)(\delta_{1}\vee\delta_{2})(A\cup C)=\delta_{1}(A\cup C)

so that

(δ1δ2)(AC)δ1(AB)+δ1(BC)(δ1δ2)(AB)+(δ1δ2)(BC).(\delta_{1}\vee\delta_{2})(A\cup C)\leq\delta_{1}(A\cup B)+\delta_{1}(B\cup C)\leq(\delta_{1}\vee\delta_{2})(A\cup B)+(\delta_{1}\vee\delta_{2})(B\cup C).
Proposition 12.
  1. (a)

    Let (X,δ1)(X,\delta_{1}) and (X,δ2)(X,\delta_{2}) be Minkowski-embeddable semidiversities, and λ>0\lambda>0. Then both (X,λδ1)(X,\lambda\delta_{1}) and (X,δ1δ2)(X,\delta_{1}\vee\delta_{2}) are Minkowski-embeddable.

  2. (b)

    Suppose that KK is a convex body in d\mathbb{R}^{d} and ϕ:dd\phi:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} is a non-degenerate affine map. Then for all AA we have R(ϕ(A),ϕ(K))=R(A,K)R(\phi(A),\phi(K))=R(A,K). Hence if there is an isometric embedding from (X,δ)(X,\delta) into (d,K)(\mathbb{R}^{d},K) then there is also an isometric embedding from (X,δ)(X,\delta) into (d,ϕ(K))(\mathbb{R}^{d},\phi(K)).

  3. (c)

    If (X,δ)(X,\delta) is Minkowski-embeddable and |A|=k+1|A|=k+1, k2k\geq 2, then

    δ(A)k(k+1)(k1)aAδ(A{a}).\delta(A)\leq\frac{k}{(k+1)(k-1)}\sum_{a\in A}\delta(A\setminus\{a\}). (5)
Proof.

(a) There are maps ϕ1:Xp\phi_{1}:X\rightarrow\mathbb{R}^{p} and ϕ2:Xq\phi_{2}:X\rightarrow\mathbb{R}^{q} and convex bodies K1pK_{1}\subset\mathbb{R}^{p} and K2qK_{2}\subset\mathbb{R}^{q} such that δ1(A)=R(ϕ1(A),K1)\delta_{1}(A)=R(\phi_{1}(A),K_{1}) and δ2(A)=R(ϕ2(A),K2)\delta_{2}(A)=R(\phi_{2}(A),K_{2}) for all AXA\subseteq X.

Then for all AXA\subseteq X we have

λδ1(A)=R(ϕ1(A),λ1K1)\lambda\delta_{1}(A)=R(\phi_{1}(A),\lambda^{-1}K_{1})

and by Proposition 4

(δ1δ2)(A)=R(ϕ1(A)×ϕ2(A),K1×K2).(\delta_{1}\vee\delta_{2})(A)=R(\phi_{1}(A)\times\phi_{2}(A),K_{1}\times K_{2}).

(b) If there are λ0\lambda\geq 0 and xx such that A+xλKA+x\subseteq\lambda K then ϕ(A)+ϕ(x)λϕ(K)\phi(A)+\phi(x)\subseteq\lambda\phi(K) so R(ϕ(A),ϕ(K))R(A,K)R(\phi(A),\phi(K))\leq R(A,K). Applying the inverse map gives equality.
(c) There is a map ϕ:Xd\phi:X\rightarrow\mathbb{R}^{d} and a convex body KK such that R(ϕ(A),K)=δ(A)R(\phi(A),K)=\delta(A). Applying Proposition 3 gives the result. ∎

We now consider Minkowski-embeddability for a few key examples of symmetric diversities.

Proposition 13.
  1. (a)

    The diversity (X,δ)(X,\delta) with δ(A)=1\delta(A)=1 for all AXA\subseteq X with |A|>1|A|>1 is Minkowski-embeddable.

  2. (b)

    The diversity (X,δ)(X,\delta) with δ(A)=|A|1\delta(A)=|A|-1 for all non-empty AXA\subseteq X is Minkowski-embeddable.

  3. (c)

    Any diversity (X,δ)(X,\delta) with X={a,b,c}X=\{a,b,c\} and δ({a,b})=δ({a,c})=δ({b,c})=1\delta(\{a,b\})=\delta(\{a,c\})=\delta(\{b,c\})=1 is Minkowski-embeddable.

Proof.

(a) Let n=|X|1n=|X|-1. Let KK be the simplex with vertex set VV given by the standard basis vectors in n\mathbb{R}^{n} together with 0. Then for non-singleton subset VVV^{\prime}\subseteq V we have R(V,K)=1R(V^{\prime},K)=1. Hence any bijection from XX to VV gives an isometric embedding.
(b) Let KK be the same simplex as in (a) and now let VV be the vertex set of K-K. The proof of Theorem 4.1 in [2] then gives R(V,K)=|V|1R(V^{\prime},K)=|V^{\prime}|-1 for all non-empty subsets of VVV^{\prime}\subseteq V. The result follows.
(c) Let x=δ({a,b,c})x=\delta(\{a,b,c\}). As (X,δ)(X,\delta) is a diversity, 1x21\leq x\leq 2. Let (X,δ1)(X,\delta_{1}) be the diversity on XX with δ1(A)=1\delta_{1}(A)=1 for all non-singleton AXA\subseteq X and let (X,δ2)(X,\delta_{2}) be the diversity on XX with δ2(A)=|A|1\delta_{2}(A)=|A|-1 for all non-empty AXA\subseteq X. Then (X,δ1)(X,\delta_{1}) and (X,δ2)(X,\delta_{2}) are Minkowski-embeddable, and by Proposition 12 (a) so is

(X,δ)=(X,δ1(x2δ2)).(X,\delta)=(X,\delta_{1}\vee(\mbox{$\frac{x}{2}$}\delta_{2})).

We now give an exact characterization for when a finite symmetric diversity is Minkowski-embeddable.

Theorem 14.

Let (X,δ)(X,\delta) be a finite symmetric diversity. Then (X,δ)(X,\delta) is Minkowski-embeddable if and only if

δ(A{a})δ(A)|A|2|A|1\frac{\delta(A\setminus\{a\})}{\delta(A)}\geq\frac{|A|-2}{|A|-1} (6)

for all AXA\subseteq X with |A|2|A|\geq 2, aAa\in A.

Proof.

Suppose that (X,δ)(X,\delta) is Minkowski-embeddable, that AXA\subseteq X, |A|2|A|\geq 2 and aAa\in A.

If |A|=2|A|=2 then (6) holds trivially. If |A|>2|A|>2 then by Proposition 12 (c)

δ(A)\displaystyle\delta(A) |A|1|A|(|A|2)xAδ(A{x})\displaystyle\leq\frac{|A|-1}{|A|(|A|-2)}\sum_{x\in A}\delta(A\setminus\{x\})
=|A|1|A|(|A|2)|A|δ(A{a}),\displaystyle=\frac{|A|-1}{|A|(|A|-2)}|A|\delta(A\setminus\{a\}),

where the last line follows since δ\delta is symmetric. Hence (6) holds.

Conversely, suppose that (6) holds for all AXA\subseteq X such that |A|2|A|\geq 2 and aAa\in A. As (X,δ)(X,\delta) is symmetric, δ\delta is monotonic and there is an increasing function f:0f:\mathbb{Z}\rightarrow\mathbb{R}_{\geq 0} such that δ(A)=f(|A|1)\delta(A)=f(|A|-1) for all non-empty AXA\subseteq X. Note that for each k=2,,|X|k=2,\ldots,|X|, choosing an AXA\subseteq X with |A|=k|A|=k and substituting AA into (6) gives

f(k2)f(k1)k2k1,\frac{f(k-2)}{f(k-1)}\geq\frac{k-2}{k-1}, (7)

a fact that we shall use later on in the proof.

Let X={x1,x2,,xn}X=\{x_{1},x_{2},\ldots,x_{n}\} and for each mnm\leq n define Xm={x1,x2,,xm}X_{m}=\{x_{1},x_{2},\ldots,x_{m}\}. We will use induction on mm to show that for each mnm\leq n the symmetric diversity (X,δ)(X,\delta) restricted to XmX_{m} is Minkowski-embeddable. This is clearly the case when m2m\leq 2.

Suppose that 2<mn2<m\leq n and that (X,δ)(X,\delta) restricted to Xm1X_{m-1} is Minkowski-embeddable. Then there is a map ϕ:Xm1d\phi:X_{m-1}\rightarrow\mathbb{R}^{d} for some dd and a convex body KdK\subseteq\mathbb{R}^{d} such that

δK(ϕ(A))=δ(A)\delta_{K}(\phi(A))=\delta(A)

for all AXm1A\subseteq X_{m-1}. We now define a collection δ(i)\delta^{(i)} of (semi)diversities on XmX_{m}, 1im1\leq i\leq m.

First, for each i=1,2,,m1i=1,2,\ldots,m-1 define the map ϕ(i):Xmd\phi^{(i)}:X_{m}\rightarrow\mathbb{R}^{d} by ϕ(i)(x)=ϕ(x)\phi^{(i)}(x)=\phi(x) if xXm1x\in X_{m-1} and ϕ(i)(xm)=ϕ(xi)\phi^{(i)}(x_{m})=\phi(x_{i}). We then define (Xm,δ(i))(X_{m},\delta^{(i)}) by

δ(i)(A)=δK(ϕ(i)(A)).\delta^{(i)}(A)=\delta_{K}(\phi^{(i)}(A)).

Since (d,δK)(\mathbb{R}^{d},\delta_{K}) is a diversity, (Xm,δ(i))(X_{m},\delta^{(i)}) satisfies (D1’) and (D2), and so (Xm,δ(i))(X_{m},\delta^{(i)}) is a Minkowski-embeddable semidiversity. Moreover, the definitions of ϕ\phi and ϕ(i)\phi^{(i)} give

δ(i)(A)={f(|A|2) if xi,xmA f(|A|1) otherwise.\delta^{(i)}(A)=\begin{cases}f(|A|-2)&\mbox{ if $x_{i},x_{m}\in A$ }\\ f(|A|-1)&\mbox{ otherwise.}\end{cases}

Second, let (Xm,δ(m))(X_{m},\delta^{(m)}) be the diversity defined by setting

δ(m)(A)=|A|1m1f(m1)\delta^{(m)}(A)=\frac{|A|-1}{m-1}f(m-1)

for all non-empty AXmA\subseteq X_{m}. This is Minkowski-embeddable by Proposition 13 (b) and Proposition 12 (a).

We now claim that

δ(A)=max{δ(i)(A):i=1,,m}\delta(A)=\max\{\delta^{(i)}(A):i=1,\ldots,m\} (8)

holds for all AXmA\subseteq X_{m}. Proving this claim will complete the proof of the theorem by induction since each (semi)diversity δ(i)\delta^{(i)} is Minkowski-embeddable, and hence by Proposition 12 (a) (Xm,δ)(X_{m},\delta) is Minkowski-embeddable.

When |A|1|A|\leq 1, (8) holds trivially, as all relevant quantities are zero. Suppose |A|2|A|\geq 2. Three cases may hold:

  • xiAx_{i}\not\in A for some i=1,,m1i=1,\ldots,m-1; then δ(i)(A)=f(|A|1)\delta^{(i)}(A)=f(|A|-1).

  • xmAx_{m}\not\in A; then δ(i)(A)=f(|A|1)\delta^{(i)}(A)=f(|A|-1) for all i=1,,m1i=1,\ldots,m-1.

  • A=XmA=X_{m}; then δ(i)(A)=f(|A|2)\delta^{(i)}(A)=f(|A|-2) for all i=1,,m1i=1,\ldots,m-1.

Hence

max{δ(i)(A):i=1,2,,m1}={f(|A|1) if AXm;f(|A|2) if A=Xm.\max\{\delta^{(i)}(A):i=1,2,\ldots,m-1\}=\begin{cases}f(|A|-1)&\mbox{ if $A\neq X_{m}$;}\\ f(|A|-2)&\mbox{ if $A=X_{m}$.}\end{cases}

We now consider δ(m)\delta^{(m)} and AXmA\subseteq X_{m}. If A=XmA=X_{m} then

δ(m)(A)=(|A|1)(m1)f(m1)=f(m1).\delta^{(m)}(A)=\frac{(|A|-1)}{(m-1)}f(m-1)=f(m-1).

Otherwise suppose 1<|A|<m1<|A|<m and so

δ(m)(A)\displaystyle\delta^{(m)}(A) =(|A|1)(m1)f(m1),\displaystyle=\frac{(|A|-1)}{(m-1)}f(m-1),
=(|A|1)(m1)(k=|A|+1mf(k1)f(k2))f(|A|1),\displaystyle=\frac{(|A|-1)}{(m-1)}\left(\prod_{k=|A|+1}^{m}\frac{f(k-1)}{f(k-2)}\right)f(|A|-1),
(|A|1)(m1)(k=|A|+1mk1k2)f(|A|1)\displaystyle\leq\frac{(|A|-1)}{(m-1)}\left(\prod_{k=|A|+1}^{m}\frac{k-1}{k-2}\right)f(|A|-1) by (7),
=f(|A|1).\displaystyle=f(|A|-1).

Hence if AXmA\neq X_{m},

max{δ(i)(A):i=1,2,,m}=max{f(|A|1),δ(m)(A)}=f(|A|1),\max\{\delta^{(i)}(A):i=1,2,\ldots,m\}=\max\{f(|A|-1),\delta^{(m)}(A)\}=f(|A|-1),

while if A=XmA=X_{m} then since ff is increasing

max{δ(i)(A):i=1,2,,m}=max{f(|A|2),δ(m)(A)}=f(|A|1).\max\{\delta^{(i)}(A):i=1,2,\ldots,m\}=\max\{f(|A|-2),\delta^{(m)}(A)\}=f(|A|-1).

We conclude that δ(A)=max{δ(i)(A):i=1,,m}\delta(A)=\max\{\delta^{(i)}(A):i=1,\ldots,m\} for all AXmA\subseteq X_{m} which completes the proof of (8) and also the theorem. ∎

By considering the diversity on X={a,b,c,d}X=\{a,b,c,d\} with

δ(A)={2,|A|=4,1,2|A|3,0,otherwise,\delta(A)=\begin{cases}2,&|A|=4,\\ 1,&2\leq|A|\leq 3,\\ 0,&\mbox{otherwise,}\end{cases}

and taking A={b,c,d}A=\{b,c,d\} in Theorem 14 we immediately obtain

Corollary 15.

There exists a diversity a set of four elements that is not Minkowski-embeddable.

We show later (Corollary 18) that every diversity on three elements is Minkowski-embeddable.

5 Parallelotopes and simplices

We have shown that not every diversity is Minkowski-embeddable, and so the question now becomes one of characterizing which diversities are. In this section we characterize when we can embed the diameter and negative-type diversities defined in the introduction in terms of Minkowski diversities having kernels equal to parallelotopes and simplices, respectively.

We first consider diameter diversities.

Theorem 16.

A finite diversity (X,δ)(X,\delta) can be embedded in a Minkowski diversity with kernel equal to some parallelotope if and only if (X,δ)(X,\delta) is a diameter diversity.

Proof.

First note that if there is some such embedding then (X,δ)(X,\delta) is a diameter diversity by Proposition 3.4 of [2].

Conversely, suppose that X={x1,x2,,xn}X=\{x_{1},x_{2},\ldots,x_{n}\} and (X,δ)(X,\delta) is a diameter diversity. Let ϕ:Xn\phi:X\rightarrow\mathbb{R}^{n} be the standard Fréchet embedding

ϕ:Xn:y(d(x1,y),d(x2,y),,d(xn,y))\phi:X\rightarrow\mathbb{R}^{n}:y\mapsto(d(x_{1},y),d(x_{2},y),\ldots,d(x_{n},y))

of the metric dd induced by δ\delta. Then d(x,y)=ϕ(x)ϕ(y)d(x,y)=\|\phi(x)-\phi(y)\|_{\infty} for all x,yXx,y\in X.

Let KK be the unit cube in n\mathbb{R}^{n}. For all AXA\subseteq X we have

δ(A)\displaystyle\delta(A) =maxa,bAd(a,b)\displaystyle=\max_{a,b\in A}d(a,b)
=maxa,bAϕ(a)ϕ(b)\displaystyle=\max_{a,b\in A}\|\phi(a)-\phi(b)\|_{\infty}
=maximaxa,bA|ϕ(a)iϕ(b)i|\displaystyle=\max_{i}\max_{a,b\in A}|\phi(a)_{i}-\phi(b)_{i}|
=R(ϕ(A),K).\displaystyle=R(\phi(A),K).

We now consider finite diversities of negative type, the diversity analog of metrics of negative type [14]. Note that the cone of all diversities on a set XX of cardinality nn has dimension 2n(n+1)2^{n}-(n+1), the number of subsets AXA\subset X with |A|2|A|\geq 2, see [35]. The set of diversities of negative type forms a cone of the same dimension, indicating that an appropriately chosen ‘random’ diversity could have non-negative type with non-zero probability.

Theorem 17.

A finite diversity (X,δ)(X,\delta) can be embedded in a Minkowski diversity with kernel equal to some simplex if and only if (X,δ)(X,\delta) has negative type.

Proof.

Let n=|X|n=|X|. From Theorem 7 of [35], (X,δ)(X,\delta) is negative-type if and only if it can be embedded in (d,δneg)(\mathbb{R}^{d},\delta_{neg}) for some dd and

δneg(A)=i=1dmax{ai:aA}min{i=1dai:aA}.\delta_{neg}(A)=\sum_{i=1}^{d}\max\{a_{i}:a\in A\}-\min\left\{\sum_{i=1}^{d}a_{i}:a\in A\right\}.

Define the polytope

K\displaystyle K =conv(0,e1,e2,,ed)\displaystyle=-\mathrm{conv}(0,e_{1},e_{2},\ldots,e_{d})
={x0:i=1dxi1}.\displaystyle=\{x\leq 0:\sum_{i=1}^{d}x_{i}\geq-1\}.

Suppose λ=R(A,K)\lambda=R(A,K), so there is zz such that AλK+zA\subseteq\lambda K+z. Then for all aAa\in A and all ii we have aizi0a_{i}-z_{i}\leq 0. Hence

zimax{ai:aA}.z_{i}\geq\max\{a_{i}:a\in A\}.

Let aAa^{*}\in A minimize i=1dai\sum_{i=1}^{d}a^{*}_{i}. Then aλK+za^{*}\in\lambda K+z implies i=1d(aizi)λ\sum_{i=1}^{d}(a^{*}_{i}-z_{i})\geq-\lambda and so

λ\displaystyle\lambda i=1dzii=1dai\displaystyle\geq\sum_{i=1}^{d}z_{i}-\sum_{i=1}^{d}a^{*}_{i}
i=1dmax{ai:aA}minaAi=1dai\displaystyle\geq\sum_{i=1}^{d}\max\{a_{i}:a\in A\}-\min_{a\in A}\sum_{i=1}^{d}a_{i}
=δneg(A).\displaystyle=\delta_{neg}(A).

Now suppose λ=δneg(A)\lambda=\delta_{neg}(A). Let zi=max{ai:aA}z_{i}=\max\{a_{i}:a\in A\} so that aizi0a_{i}-z_{i}\leq 0 for all aAa\in A and all i=1,,di=1,\ldots,d. Furthermore

λ\displaystyle-\lambda =i=1dzi+min{i=1dai:aA}\displaystyle=-\sum_{i=1}^{d}z_{i}+\min\left\{\sum_{i=1}^{d}a_{i}:a\in A\right\}
i=1dzi+i=1dai\displaystyle\leq-\sum_{i=1}^{d}z_{i}+\sum_{i=1}^{d}a_{i} for all aAa\in A

so that i=1d(aizi)λ\sum_{i=1}^{d}(a_{i}-z_{i})\geq-\lambda and azλKa-z\in\lambda K for all aAa\in A. Hence R(A,K)δneg(A)R(A,K)\leq\delta_{neg}(A).

We have shown that (X,δ)(X,\delta) is of negative type if and only if it can be embedded into a Minkowski diversity with kernel equal to the particular simplex KK. The theorem now follows from Proposition 12(b) and the fact that every simplex in d\mathbb{R}^{d} can be transformed into another by a non-degenerate affine map.

The last theorem immediately implies that two further classes are Minkowski-embeddable.

Corollary 18.

If δ\delta is diversity on three elements or a finite diversity that can be embedded in L1L_{1}, then δ\delta is Minkowski-embeddable.

Proof.

All three-element diversities and L1L_{1}-embeddable diversities have negative type [35]. ∎

6 Open problems

We have characterized diversities and functions defined by the generalized circumradius R(A,K)R(A,K), and established preliminary results on embedding finite diversities into these spaces. Our results suggest several avenues for further investigation.

First, is there a complete characterization of when a finite diversity is Minkowski-embeddable? Indeed it is not even obvious which finite diversities can be embedded into sublinear diversities in d\mathbb{R}^{d}.

A second related question is algorithmic in nature: Are there efficient algorithms for determining whether or not a finite diversity can be embedded in some dimension? Interestingly, we note that for the classical case of a circumradius, even though we do not know a characterization for Minkowski-embeddability, we are able to give an efficient algorithm for deciding embeddability (for bounded dimension):

Proposition 19.

Let (X,δ)(X,\delta) be a finite diversity such that δ(A)=max{δ(A):AA,|A|d+1}\delta(A)=\max\{\delta(A^{\prime}):A^{\prime}\subseteq A,|A^{\prime}|\leq d+1\} for all AXA\subseteq X. For fixed dd, there is an algorithm which runs in polynomial time in n=|X|n=|X| to determine if (X,δ)(X,\delta) is Minkowski-embeddable in d\mathbb{R}^{d} with kernel equal to the unit ball 𝔅\mathfrak{B}.

Proof.

We begin with a useful observation. Suppose that there is an (unknown) embedding ϕ:Xd\phi:X\rightarrow\mathbb{R}^{d} such that δ(A)=δ𝔅(ϕ(A))\delta(A)=\delta_{\mathfrak{B}}(\phi(A)) for all AXA\subseteq X. Note that since the metric induced by δ𝔅\delta_{\mathfrak{B}} is Euclidean so is the one induced by δ\delta. Let ψ:Xd\psi:X\rightarrow\mathbb{R}^{d} be any map which preserves the metrics induced by δ\delta and δ𝔅\delta_{\mathfrak{B}}. In addition, let ff be the (unknown) isometry from ψ(X)\psi(X) to ϕ(X)\phi(X) given by f(ψ(x))=ϕ(x)f(\psi(x))=\phi(x) for all xXx\in X. As d\mathbb{R}^{d} is a finite dimensional Hilbert space, ff can be extended to an isometry on the whole space d\mathbb{R}^{d} (see, e.g., [34], Theorem 11.4). Moreover, for any AXA\subseteq X we have

δ𝔅(ψ(A))\displaystyle\delta_{\mathfrak{B}}(\psi(A)) =inf{sup{ψ(a)x2:aA}:xd}\displaystyle=\inf\Big{\{}\sup\{\|\psi(a)-x\|_{2}:a\in A\}:x\in\mathbb{R}^{d}\Big{\}}
=inf{sup{f(ψ(a))f(x)2:aA}:xd}\displaystyle=\inf\Big{\{}\sup\{\|f(\psi(a))-f(x)\|_{2}:a\in A\}:x\in\mathbb{R}^{d}\Big{\}}
=inf{sup{ϕ(a)y2:aA}:yd}\displaystyle=\inf\Big{\{}\sup\{\|\phi(a)-y\|_{2}:a\in A\}:y\in\mathbb{R}^{d}\Big{\}}
=δ𝔅(ϕ(A))\displaystyle=\delta_{\mathfrak{B}}(\phi(A))
=δ(A).\displaystyle=\delta(A).

Hence, if (X,δ)(X,\delta) is Minkowski embeddable into (d,δ𝔅)(\mathbb{R}^{d},\delta_{\mathfrak{B}}), then the map ψ\psi gives one embedding.

We now present an algorithm for deciding whether or not (X,δ)(X,\delta) is embeddable in (d,δ𝔅)(\mathbb{R}^{d},\delta_{\mathfrak{B}}):

  1. 1.

    Decide whether or not the metric induced by δ\delta on XX is Euclidean. If not, then (X,δ)(X,\delta) cannot be embedded in (d,δ𝔅)(\mathbb{R}^{d},\delta_{\mathfrak{B}}). Else, compute a (metric) embedding ψ\psi of XX in d\mathbb{R}^{d} which preserves the induced metrics.

  2. 2.

    If δ𝔅(ψ(A))=δ(A)\delta_{\mathfrak{B}}(\psi(A))=\delta(A) for all AA with |A|d+1|A|\leq d+1 then (X,δ)(X,\delta) can be embedded in (d,δ𝔅)(\mathbb{R}^{d},\delta_{\mathfrak{B}}), otherwise (X,δ)(X,\delta) cannot be embedded in (d,δ𝔅)(\mathbb{R}^{d},\delta_{\mathfrak{B}}).

The correctness of this algorithm follows by the observation above. To see that it also runs in polynomial time in nn (for fixed dd), note that Step 1 can be computed in polynomial time in nn by the results in e.g. [14, Section 6.2] or [34, Theorem 2.1], and that for Step 2 the definition of δ\delta and Proposition 2 imply that to determine whether δ𝔅(ψ(A))=δ(A)\delta_{\mathfrak{B}}(\psi(A))=\delta(A) for all AXA\subseteq X we need only check subsets AA with |A|d+1|A|\leq d+1. ∎

Another question is how to extend the embedding results to include distortion. Let (X1,δ1)(X_{1},\delta_{1}) and (X2,δ2)(X_{2},\delta_{2}) be two diversities. We say that a map ϕ:X1X2\phi:X_{1}\rightarrow X_{2} has distortion cc if there are c1,c2>0c_{1},c_{2}>0 such that c=c1c2c=c_{1}c_{2} and

1c1δ1(A)δ2(ϕ(A))c2δ1(A)\frac{1}{c_{1}}\delta_{1}(A)\leq\delta_{2}(\phi(A))\leq c_{2}\delta_{1}(A)

for all finite AXA\subseteq X. Continuing the program of [25], it is shown in [10] that bounds on the distortion of embeddings from diversities into L1L_{1}-diversities provide approximation algorithms for hypergraph generalizations of sparsest cut. It was shown in [35] that there are finite diversities of metric type which cannot be embedded into L1L_{1} without at least Ω(log|X|)\Omega(\sqrt{\log|X|}) distortion. This bound therefore holds for Minkowski-embeddable diversities. In general, questions concerning distortion seem intricately connected with core sets of the generalized circumradius [2].

Apart from potential algorithmic gains it would be good to explore embeddings with distortion for diversities into Minkowski diversities with low dimension, simply for their use in visualization and modelling of diversity type data.

Acknowledgments

We thank Pei Wu and Malcolm Jones for discussions and ideas which helped initiate this paper. PT is supported by an NSERC (Canada) Discovery Grant. DB, KTH and VM thank the Royal Society for its support. We also thank the reviewers for their helpful comments.

References

  • [1] Blumenthal, L. M. Theory and Applications of Distance Geometry, 2nd ed., vol. 242. Chelsea Publishing Company, 1970.
  • [2] Brandenberg, R., and König, S. No dimension-independent core-sets for containment under homothetics. Discrete and Computational Geometry 49 (2013), 3–21.
  • [3] Brandenberg, R., and König, S. Sharpening geometric inequalities using computable symmetry measures. Mathematika 61 (2015), 559–580.
  • [4] Brandenberg, R., and Roth, L. Minimal containment under homothetics: a simple cutting plane approach. Computational Optimization and Applications 48, 2 (2011), 325–340.
  • [5] Bryant, D., Cioica-Licht, P., Clark, L. O., and Young, R. Inner products for convex bodies. Journal of Convex Analysis 28, 4 (2021), 1249–1264.
  • [6] Bryant, D., Felipe, R., Toledo-Acosta, M., and Tupper, P. F. Lattice diversities. arXiv 2010.11442 (under submission), 2020.
  • [7] Bryant, D., Nies, A., and Tupper, P. F. A universal separable diversity. Analysis and geometry in metric spaces 5, 1 (2017), 138–151.
  • [8] Bryant, D., Nies, A., and Tupper, P. F. Fraïssé limits for relational metric structures. Journal of Symbolic Logic 86, 3 (2021), 913–934.
  • [9] Bryant, D., and Tupper, P. F. Hyperconvexity and tight-span theory for diversities. Advances in Mathematics 231, 6 (2012), 3172–3198.
  • [10] Bryant, D., and Tupper, P. F. Diversities and the geometry of hypergraphs. Discrete Mathematics and Theoretical Computer Science 16, 2 (2014), 1–20.
  • [11] Bryant, D., and Tupper, P. F. Constant distortion embeddings of symmetric diversities. Analysis and Geometry in Metric Spaces 4, 1 (2016), 326–335.
  • [12] Croitoru, A. Set-norm continuity of set multifunctions. ROMAI Journal 6, 1 (2010), 47–56.
  • [13] Danzer, L., Grünbaum, B., and Klee, V. Helly’s theorem and its relatives. In Convexity, V. Klee, Ed., vol. 7 of Proceedings of Symposia in Pure Mathematics. AMS, Providence, RI, 1963, pp. 101–180.
  • [14] Deza, M. M., and Laurent, M. Geometry of Cuts and Metrics, vol. 15 of Algorithms and Combinatorics. Springer-Verlag, Berlin, 1997.
  • [15] Dokania, P. K. High-Order Inference, Ranking, and Regularization Path for Structured SVM. PhD thesis, Université Paris Saclay (COmUE), 2016.
  • [16] González Merino, B., Jahn, T., and Richter, C. Uniqueness of circumcenters in generalized Minkowski spaces. Journal of Approximation Theory 237 (2019), 153–159.
  • [17] Gritzmann, P., and Klee, V. Inner and outer jj-radii of convex bodies in finite-dimensional normed spaces. Discrete & Computational Geometry 7 (1992), 255–280.
  • [18] Gritzmann, P., and Klee, V. Computational complexity of inner and outer jj-radii of polytopes in finite-dimensional normed spaces. Mathematical Programming 59 (1993), 163–213.
  • [19] Gritzmann, P., and Klee, V. On the complexity of some basic problems in computational convexity: I. Containment problems. Discrete Mathematics 136, 1-3 (1994), 129–174.
  • [20] Hallbäck, A. Metric model theory, Polish groups & diversities. PhD thesis, Université de Paris, 2020.
  • [21] Jahn, T. Extremal radii, diameter and minimum width in generalized Minkowski spaces. Rocky Mountain Journal of Mathematics 47, 3 (2017), 825–848.
  • [22] Jahn, T. Successive radii and ball operators in generalized Minkowski spaces. Advances in Geometry 17, 3 (2017), 347–354.
  • [23] Kirk, W., and Shahzad, N. Diversities. In Fixed Point Theory in Distance Spaces, W. Kirk and N. Shahzad, Eds. Springer International Publishing, Cham, 2014, pp. 153–158.
  • [24] Komodakis, N., Pawan Kumar, M., and Paragios, N. (Hyper)-graphs inference through convex relaxations and move making algorithms: Contributions and applications in artificial vision. Foundations and Trends in Computer Graphics and Vision 10, 1 (2014), 1–102.
  • [25] Linial, N., London, E., and Rabinovich, Y. The geometry of graphs and some of its algorithmic applications. Combinatorica 15, 2 (1995), 215–245.
  • [26] Pia̧tek, B. On the gluing of hyperconvex metrics and diversities. Annales Universitatis Paedagogicae Cracoviensis. 149, Studia Mathematica 13, 1 (2014), 65–76.
  • [27] Poelstra, A. On the topological and uniform structure of diversities. Journal of Function Spaces 2013 (2013), Article ID 675057.
  • [28] Rudin, W., et al. Principles of mathematical analysis, vol. 3. McGraw-hill New York, 1976.
  • [29] Schneider, R. Convex bodies: the Brunn–Minkowski theory, 2nd ed. Encyclopedia of Mathematics and It Applications. Cambridge University Press, Cambridge, 2014.
  • [30] Steel, M. Tracing evolutionary links between species. American Mathematical Monthly 121, 9 (2014), 771–792.
  • [31] Steel, M. Continuous phylogenies and distance-based tree reconstruction. In Phylogeny. Society for Industrial and Applied Mathematics, Philadelphia, PA, 2016, ch. 6, pp. 111–145.
  • [32] Sylvester, J. J. A question in the geometry of situation. Quarterly Journal of Pure and Applied Mathematics 1 (1857), 79–80.
  • [33] Toledo Acosta, G. M. Generalizaciones sobre la noción de diversidades. PhD thesis, Centro de Investigación en Matemáticas A.C., 2016.
  • [34] Wells, J. H., and Williams, L. R. Embeddings and Extensions in Analysis, vol. 84. Springer Science & Business Media, New York, 1975.
  • [35] Wu, P., Bryant, D., and Tupper, P. F. Negative-type diversities, a multi-dimensional analogue of negative-type metrics. The Journal of Geometric Analysis 31 (2021), 1703–1720.