Linear and Sublinear Diversities
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 , many applications of diversity theory require a specialized theory for diversities defined on , as we develop here. We focus on two fundamental classes of diversities defined on : 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 . 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 where is a set and is a non-negative function defined on finite subsets of satisfying
(D1) and if and only if ,
(D2) whenever .
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) whenever
and subadditivity on intersecting sets
(D4) when .
We say is a semidiversity if (D1) is relaxed to
(D1′) and if .
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
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 metrics have the form
while diversities [7] have the form
Negative-type metrics satisfy
for all zero-sum vectors while negative type diversities [27] satisfy
for all zero-sum vectors with .
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 is given by We investigate diversities defined on which are (Minkowski) linear [24]
(D5)
and those which are (Minkowski) sublinear
(D6)
for and nonempty finite subsets of . Many familiar diversities defined on 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 is defined
Here denotes the usual dot product in ,
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.
and for for non-empty, bounded and
-
2.
If are nonempty convex compact sets then if and only for all
-
3.
A function is the support function for some bounded nonempty set if and only if and for all and (that is, is sublinear).
We often consider support functions restricted to , the unit sphere in , noting that a support function is determined everywhere by its values on . We note that the support function restricted to of a nonempty set is bounded if and only if the set is bounded.
Our main results for diversities and semidiversities are:
-
1.
(Theorem 5) Linear diversities and semidiversities are exactly those which can be written in the form
for a Borel measure on the sphere satisfying
(1) -
2.
(Theorem 7) The extremal linear semidiversities are those where the support of is a finite, affinely independent set, which in turn correspond to a generalized circumradius (a Minkowski semidiversity) based on the simplex.
-
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 we show:
-
1.
(Theorem 11) A finite diversity can be embedded in a linear diversity if and only if it has negative type, meaning that
for all vectors with zero sum and .
-
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 we have , even if that is not stated explicitly below.
-
1.
Let be any norm on . The diameter diversity is given by
for finite . The diameter diversity is sublinear [24, pg 49].
- 2.
-
3.
The circumradius of finite with respect to the unit ball is
More generally, the Minkowski diversity with kernel is equal to the generalized circumradius
for finite . Minkowski diversities are sublinear [3] but are not, in general, linear. For example, consider the circumradius diversity . If and then .
We assume that is closed, convex and has non-empty interior. We have elsewhere required that the kernel be bounded, however in this paper we will not require this. Note that if is an unbounded, is a semidiversity rather than a diversity.
-
4.
The mean-width diversity is
where is the (uniform) Haar measure on the sphere and . Equivalently, is the mean-width of the convex hull of . Mean-width diversities are linear [24, pg 50].
-
5.
A zonotope is a Minkowski sum of line segments and the length of the zonotope equals the sum of the length of the line segments. We define the zonotope diversity where is the minimum length of a zonotope containing . We show that zonotope diversities are sublinear in Proposition 2.
In a Euclidean space , 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 be a function on finite subsets of which satisfies (D1), monotonicity (D3) and sublinearity (D6).
-
1.
is translation invariant: for all finite and .
-
2.
is a diversity.
-
3.
If then .
-
4.
The map given by is a norm on .
-
5.
For all finite with we have
If satisfies (D1′) rather than (D1) then 1-5 still hold except that is a semidiversity and is a seminorm.
Proof.
-
1.
By sublinearity (D6) and (D1), we have
-
2.
As is monotonic and , is non-negative, and by part 1. is translation invariant. We show that satisfies (D4). Suppose that . Then and so and
Hence satisfies (D1), (D3) and (D4).
-
3.
Proposition 2.2b in [3].
-
4.
By (D5) we have . If then , while if we have
Also, if and only if if and only if .
-
5.
This follows from sublinearity and the following observation; see the proof of [1, Theorem 4.1]. By sublinearity we may assume . So for each
We also have for . So for all
This gives
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 is a sublinear diversity.
Proof.
Recall that is the shortest length of a zonotope containing . The function is clearly monotonic, vanishes when , and is strictly positive when . Given finite , let and the the minimum length zonotopes containing and respectively. Then is a zonotope containing with length . By Proposition 1 part 2, is a sublinear diversity. ∎
The zonotope diversity is not linear: let and . Then but .
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 to be the set
and .
Proposition 3.
Let be a sublinear semidiversity.
-
1.
is a linear subspace of
-
2.
restricted to is a diversity
-
3.
If is the projection operator for then for all finite .
Proof.
-
1.
For and we have and so and . By translation invariance, and .
-
2.
Suppose and . We have that by part 1. By translation invariance which implies . Hence is both in a subspace and its orthogonal complement, and so .
-
3.
For all finite we have and for some . We have since
and, likewise, . By sublinearity, .
∎
Let be a norm on with associated metric and unit ball . The Hausdorff distance between two nonempty closed bounded sets and can be defined by [24, p. 61] :
For bounded define
(2) |
Proposition 4.
Let be a sublinear semidiversity.
-
1.
For all bounded , .
-
2.
For all finite we have
-
3.
For all bounded and
and
-
4.
If is linear then the restriction of to the set of nonempty compact convex subsets of is a valuation. That is,
for all nonempty compact convex bodies such that and are non-empty and convex.
-
5.
The restriction of to the set of nonempty compact convex subsets of is Lipschitz continuous with respect to the Hausdorff metric, with Lipschitz constant .
Proof.
-
1.
By equivalency of norms on we have that is bounded with respect to metric if and only if it is bounded with respect to the induced metric of . Let be the set of vertices of some polytope (e.g. a cube) containing . For all finite we have by monotonicity and part 2 that
so that
-
2.
Let be a finite subset of and let . For any we have so by Proposition 1 (ii). Hence
-
3.
Fix and suppose that is a finite subset of such that . For each there is and such that . Let and so that . It follows that
Taking to zero gives the result.
Let be a finite subset of such that . As we have
Hence from which equality follows by symmetry.
- 4.
-
5.
Suppose that are bounded nonempty subsets satisfying . For any there is a finite such that . We also have so there is finite and such that . Hence
By a symmetric argument,
Taking to zero, we have
The bound is tight, as can be seen by letting and . Then and .
∎
Bryant et al. [3] also describe an extension of Minkowski diversities from finite sets to bounded sets. They define for any polytope with vertex set , and extend that to general bounded convex sets by defining for any sequence of polytopes converging to . Proposition 4 part 2. gives that for any polytope, while from Proposition 4 part 5 we have that . Hence coincides with 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 be a function defined on finite subsets of . Then is a linear semidiversity if and only if there is a positive finite Borel measure on the unit sphere such that
(3) |
and
(4) |
for all finite . Such a measure is unique.
Proof.
First we show that given by (4) is a linear semidiversity. For and finite we have
and, since and for all we have and . By Proposition 1, is a linear semidiversity.
For the converse, let be a linear semidiversity and define as in (2). By Proposition 4 the restriction of 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 ,
for some positive finite Borel measure satisfying (3), and 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 , let . Since and , the result follows for all nonempty finite . ∎

In Figure 1 we depict the support for the measures corresponding to mean width (uniform on the unit circle), the diversity (), 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 forms a cone. A non-zero semidiversity 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 be affinely independent with , for some , . Define the polyhedron . Let be the Minkowski semidiversity given by . Then
for all finite .
Proof.
Let . We express as the solution to a linear program. Recall that is the minimum such that there is some such that for all . We can rewrite this constraint as for all . If we take and to be our primal variables we get the following linear program:
minimize | |||
subject to |
The dual linear program with dual variables is
maximize | |||
subject to | |||
Let . Then our dual constraints are equivalent to
Since the are affinely independent and , , there is a unique solution given by for all . Now it remains to determine for each the value of for each . We need to maximize given and . For each , the solution is to let for the that maximizes , and otherwise. This gives for the solution to the dual problem
∎
The following theorem identifies extremal linear semidiversities as Minkowski diversities with equal to a simplex or a simplex plus a subspace.
Theorem 7.
The following are equivalent for a semidiversity :
-
(i)
is extremal in the class of linear semidiversities.
-
(ii)
satisfies
for all finite , where is a measure on with , such that the support of is a finite, affinely independent set.
-
(iii)
is a Minkowski semidiversity with kernel of the form
where is an affinely independent set of points, is the affine closure of ,and is the orthogonal space to .
Proof.
(i) (ii). Suppose is extremal and the support of is not affinely independent. Let be the affine hull of the support of , with , and let . Affine dependence implies is not supported on only points or fewer. Therefore we can partition into , each with . Let . Then
and
where . Choose a subset of the with points so that 0 is in the convex hull of them. Let’s say they are . Find for such that , and . Let . Now define by for , and zero otherwise. Then , and has smaller support, because but . Also
We can now write where and are not scale copies, so is not extremal in the cone of linear semidiversities.
(ii) (i). Suppose that
for all finite and some measure on with affinely independent support. Let and be linear semidiversities with corresponding measures and . If then affine independence and the constraint that implies that is a scaled version of . Likewise for . Hence if for we have that and both and are scale versions of . This shows that is an extremal linear diversity.
(ii) (iii). Let the support of be the points with weights such that .
Let , , and , so that , , and .
Let , which is affinely independent because
the are. Let be the span of and its orthogonal complement. By Lemma 6, , the Minkowski semidiversity for the set . The intersection of with is a simplex; let it have vertices . Then as required.
(iii) (ii). By translating if necessary, we may assume is in the relative interior of .
We can write for some affinely independent .
Because is bounded,
.
So there are with and . By Lemma 6 we have
for all finite . Let and . Then the are also affinely independent. Let be the measure that assigns mass to each . ∎
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 is linear if and only if 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 is a linear semidiversity. By Theorem 5, there is a Borel measure on such that and for all finite . Let .
Let be the set of all signed Borel measures on , which is a Hausdorff locally convex set [26, p. 134]. The space of measures on with and is compact by the Banach-Alaoglu theorem [26, p. 114], and is convex.
We claim that the set of extremal points of is closed. Let , be a sequence of extremal measures that converges in the vague topology, so that converges to for some for all continuous bounded . By repeatedly taking subsequences, we can obtain a subsequence (where is a unit mass measure at ) where and for some and . Since as , we must have , showing that the limit is also an extremal point in . 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 is a weighted average of members of the closure of the extreme points of . Since the set of extremal points of 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 be a function on finite subsets of . If is a sublinear diversity or semidiversity then there is a collection of linear semidiversities such that
Conversely, for any collection of linear semidiversities and defined by , is a sublinear semidiversity.
Proof.
Suppose that are linear semidiversities and
for all finite . Note that vanishes on singletons and is monotonic since each has these properties. Suppose that are finite subsets of and . Then
and
So is sublinear. By Proposition 1, is a sublinear semidiversity.
For the converse, suppose that is sublinear. Define to be the set of all support functions for nonempty finite . Define on the convex cone by for all finite sets . The function is sublinear (and convex in the terminology of [25]), as for any finite ,
and for .
Fix finite . Define on by
That is, is the largest we can scale so that a translate is contained in . Note that . This tells us that .
We show that is superlinear. For all we have . Now suppose that are finite and non-empty subsets of . Given there are , , such that
and hence | ||||
so that . Taking gives superlinearity.
We now have that is monotonic and sublinear and that is superlinear.
Furthermore, for any finite and there is such that and such that , and so
Taking we conclude that for all . Recall that .
For each finite we have now satisfied the conditions for Theorem 1.2.5 of [11]:
Let be a pre-ordered cone and let be monotone and sublinear, superlinear with . Then there is a monotone linear with .
In our example is the cone of support functions of finite sets. Let . Let be the linear map given by the theorem. It is monotone, linear, and
Since by definition for all , for all .
Now define by for all finite . Then vanishes on singletons, it is monotone, linear, and hence also sublinear. By Proposition 1 is a linear semidiversity.
Because , we have that
and for general finite we have
Repeating this process for all finite we obtain a set of linear semidiversities such that and for all finite . So for all finite ,
since the supremum is actually attained when . ∎
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 between two diversities and is an isometric embedding if for all finite . We say that a finite diversity is linear-embeddable if there is an isometric embedding from to a linear diversity on for some and sublinear-embeddable if there is an isometric embedding to some sublinear diversity on , for some . At this stage we allow the dimension 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 is an extremal linear semidiversity, and is finite then is Minkowski embeddable with a simplex kernel.
Proof.
By Theorem 7 is the Minkowski semidiversity with kernel , where is a set of affinely independent vectors lying in a subspace . Let be an orthogonal matrix so that and . Let be the first rows of , so that , and is a full-dimensional simplex in . Then for all , for some if and only if for some . So for all finite , as required. ∎
Theorem 11.
Let be a finite diversity. The following are equivalent:
-
(i)
is linear-embeddable.
-
(ii)
has negative type.
-
(iii)
can be embedded into a Minkowski diversity with kernel equal to a simplex .
Proof.
(i)(ii)
Without loss of generality assume where 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, also has negative type.
(ii)(iii)
This is Theorem 17 in [3].
(iii)(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 is embeddable in a Minkowski diversity with a simplex kernel, it also embeddable in a linear diversity.
∎
Theorem 12.
Let be a finite diversity. The following are equivalent:
-
(i)
is sublinear-embeddable.
-
(ii)
can be embedded into a Minkowski diversity.
-
(iii)
is the maximum of a collection of negative type diversities.
Proof.
(ii)(i) If is embeddable as a Minkowski diversity, then it is sublinear-embeddable, since by Theorem 2.4 of [3] all Minkowski diversities are sublinear.
(i)(ii)
Let be a sublinear-embeddable. We may assume is a subset of where is a sublinear diversity.
By Theorem 9 there is a family of linear semidiversities for such that .
Since is finite, it has a finite number of subsets, and so we may assume that 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 is Minkowski embeddable.
(i)(iii)
We may assume where is a sublinear diversity. By Theorem 9 there is a family of linear semidiversities for such that . Since is finite, it has a finite number of subsets, and so we may assume that is finite. By Theorem 11, each of is negative type, and therefore is the maximum of a collection of negative type diversities.
(iii)(ii) Suppose is the maximum of a collection of negative type diversities. By Theorem 11 can then be represented as the maximum of a collection of Minkowski diversities, and since 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.