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

Total positivity of copulas from a Markov kernel perspective

S. Fuchs111Department for Artificial Intelligence & Human Interfaces, University of Salzburg, Hellbrunnerstrasse 34, 5020 Salzburg, Austria. [email protected], M. Tschimpke222Department for Artificial Intelligence & Human Interfaces, University of Salzburg, Hellbrunnerstrasse 34, 5020 Salzburg, Austria. [email protected]
(February 21, 2025)
Abstract

The underlying dependence structure between two random variables can be described in manifold ways. This includes the examination of certain dependence properties such as lower tail decreasingness (LTD), stochastic increasingness (SI) or total positivity of order 22, the latter usually considered for a copula (TP2) or (if existent) its density (d-TP2). In the present paper we investigate total positivity of order 22 for a copula’s Markov kernel (MK-TP2 for short), a positive dependence property that is stronger than TP2 and SI, weaker than d-TP2 but, unlike d-TP2, is not restricted to absolutely continuous copulas, making it presumably the strongest dependence property defined for any copula (including those with a singular part such as Marshall-Olkin copulas). We examine the MK-TP2 property for different copula families, among them the class of Archimedean copulas and the class of extreme value copulas. In particular we show that, within the class of Archimedean copulas, the dependence properties SI and MK-TP2 are equivalent.

Keywords: Copula, Markov kernel, Dependence property

1 Introduction

Copulas capture the many facets of dependence relationships between (continuous) random variables, and various ways exist to quantify and describe them. For the quantification of dependence, numerous measures are available that examine dependence from a wide range of perspectives. This includes (but is not limited to) the popular measures of concordance Kendall’s tau and Spearman’s rho (see, e.g., [1, 2]), the tail dependence coefficients and functions (see, e.g., [3]), or measures of directed or mutual complete dependence as presented in [4, 5, 6].
Apart from the option of assigning a single value to a dependence structure, an alternative way consists in checking whether a copula fulfills certain (positive) dependence properties such as positive quadrant dependence (PQD), left tail decreasingness (LTD), right tail increasingness (RTI), stochastic increasingness (SI), or total positivity of order 22, the latter usually considered for a copula (TP2) or (if existent) its density (d-TP2 for short), where the different notions of positive dependence are linked as given in Figure 1 below (see, e.g., [3, 2]). LTD, RTI and SI are defined coordinatewise while PQD, TP2 and d-TP2 are symmetric in the sense that permuting the coordinates has no effect on the respective dependence property.

Dependence properties may have direct impact on the interplay of different measures of dependence or the equivalence of different modes of convergence: For instance, as mentioned in [7, 8], Spearman’s rho is larger than Kendall’s tau for all those copulas that are LTD and RTI. Moreover, as shown in [9] pointwise convergence and weak conditional convergence (see, e.g., [10, 11]), i.e., weak convergence of almost all conditional distribution functions, are equivalent for stochastically increasing (SI) copulas.

The property total positivity of order 22 has been extensively studied for copulas (TP2) and their densities (d-TP2); see, e.g., [12, 13, 3, 14]. The d-TP2 property is one of the strongest notions of positive dependence - stronger than TP2 and also stronger than stochastic increasingness (SI) (the latter two notions are not related to each other, in general). Both notions, TP2 and SI, can be examined for any copula, however, the d-TP2 property is defined only for those copulas that are absolutely continuous, which automatically excludes many known copula families such as Fréchet copulas or Marshall-Olkin copulas.
In this paper we build the analysis on conditional distributions and investigate total positivity of order 22 for a copula’s Markov kernel (MK-TP2 for short), a positive dependence property that is stronger than TP2 and SI, weaker than d-TP2 (see Figure 1 below) but, unlike d-TP2, is not restricted to absolutely continuous copulas, making it presumably the strongest dependence property defined for any copula. From a conditional distribution point of view the MK-TP2 property has been studied in [15, 16] showing its interrelation with the other above-mentioned dependence properties.

TP2d-TP2MK-TP2LTD & RTIPQDSI
Figure 1: Relations between the different notions of positive dependence.

In the present paper, we investigate under which conditions the MK-TP2 property is fulfilled for certain copulas classes, among them the class of Archimedean copulas and the class of extreme value copulas (including Marshall-Olkin copulas). To the best of the authors’ knowledge, no results are known in this regard so far. We show that, for Archimedean copulas, the properties MK-TP2 and SI are equivalent, and can be characterized by the Archimedean generator. In passing we slightly extend a well-known result from literature (see, e.g., [7, 17]) by proving the equivalence of SI and log-convexity of the generator’s derivative for arbitrary Archimedean copulas (i.e., the generator neither needs to be strict nor twice differentiable). The equivalence of MK-TP2 and SI is consistent with the very restrictive, exchangeable structure of Archimedean copulas and the well-known equivalence of TP2 and LTD (see, e.g., [2]).
Another class of copulas generated by convex functions are extreme value copulas (EVC). It is well-known that EVCs are TP2 and SI, however, fail to be MK-TP2, in general. We present sufficient and necessary conditions (in terms of the Pickands dependence function) for an EVC to be MK-TP2 and, for certain parametric subclasses (including Marshall-Olkin copulas, Gumbel copulas, and EVCs occuring in the symmetric mixed model by Tawn), we determine exactly those parameters for which their elements are MK-TP2.

The rest of this contribution is organized as follows: Section 2 gathers preliminaries and notation that will be used throughout the paper. In Section 3, we formally define the positive dependence properties and examine the MK-TP2 property for different well-known copulas and copula families. Section 4 is devoted to Archimedean copulas and comprises the equivalence of MK-TP2 and SI property in the Archimedean setting. In Section 5 we then present sufficient and necessary conditions for an extreme value copula to be MK-TP2. Some auxiliary results and technical proofs can be found in Section 6.

2 Notation and preliminaries

Throughout this paper we will write 𝕀:=[0,1]{\mathbb{I}}:=[0,1] and let 𝒞\mathcal{C} denote the family of all bivariate copulas; MM will denote the comonotonicity copula, Π\Pi the independence copula and WW will denote the countermonotonicity copula.
For every C𝒞C\in\mathcal{C} the corresponding probability measure will be denoted by μC\mu_{C}, i.e. μC([0,u]×[0,v])=C(u,v)\mu_{C}([0,u]\times[0,v])=C(u,v) for all (u,v)𝕀2(u,v)\in{\mathbb{I}}^{2}; for more background on copulas and copula measures we refer to [1, 2]. For every metric space (Ω,δ)(\Omega,\delta) the Borel σ\sigma-field on Ω\Omega will be denoted by (Ω)\mathcal{B}(\Omega).

In what follows Markov kernels will play a decisive role: A Markov kernel from \mathbb{R} to ()\mathcal{B}(\mathbb{R}) is a mapping K:×()𝕀K:\mathbb{R}\times\mathcal{B}(\mathbb{R})\rightarrow{\mathbb{I}} such that for every fixed F()F\in\mathcal{B}(\mathbb{R}) the mapping xK(x,F)x\mapsto K(x,F) is measurable and for every fixed xx\in\mathbb{R} the mapping FK(x,F)F\mapsto K(x,F) is a probability measure. Given a real-valued random variable YY and a real-valued random variable XX on a probability space (Ω,𝒜,P)(\Omega,\mathcal{A},P) we say that a Markov kernel KK is a regular conditional distribution of YY given XX if

K(X(ω),F)=E(𝟙FY|X)(ω)\displaystyle K(X(\omega),F)=E(\mathds{1}_{F}\circ Y\,|\,X)(\omega)

holds PP-almost surely for every F()F\in\mathcal{B}(\mathbb{R}). It is well-known that for each random vector (X,Y)(X,Y) a regular conditional distribution K(.,.)K(.,.) of YY given XX always exists and is unique for PXP^{X}-a.e. xx\in\mathbb{R}, where PXP^{X} denotes the push-forward of PP under XX. If (X,Y)(X,Y) has distribution function HH (in which case we will also write (X,Y)H(X,Y)\sim H and let μH\mu_{H} denote the corresponding probability measure on (2)\mathcal{B}(\mathbb{R}^{2}) we will let KHK_{H} denote (a version of) the regular conditional distribution of YY given XX and simply refer to it as Markov kernel of HH. If C𝒞C\in\mathcal{C} is a copula then we will consider the Markov kernel of CC (with respect to the first coordinate) automatically as mapping KC:𝕀×(𝕀)𝕀K_{C}:{\mathbb{I}}\times\mathcal{B}({\mathbb{I}})\rightarrow{\mathbb{I}}. Defining the uu-section of a set G(𝕀2)G\in\mathcal{B}({\mathbb{I}}^{2}) as Gu:={v𝕀:(u,v)G}G_{u}:=\{v\in{\mathbb{I}}\,:\,(u,v)\in G\} the so-called disintegration theorem yields

μC(G)=𝕀KC(u,Gu)dλ(u)\displaystyle\mu_{C}(G)=\int_{{\mathbb{I}}}K_{C}(u,G_{u})\;\mathrm{d}\lambda(u) (2.1)

so, in particular, we have

μC(𝕀×F)=𝕀KC(u,F)dλ(u)=λ(F)\displaystyle\mu_{C}({\mathbb{I}}\times F)=\int_{{\mathbb{I}}}K_{C}(u,F)\;\mathrm{d}\lambda(u)=\lambda(F)

where λ\lambda denotes the Lebesgue measure on (𝕀)\mathcal{B}({\mathbb{I}}). For more background on conditional expectation and general disintegration we refer to [18, 19]; for more information on Markov kernels in the context of copulas we refer to [1, 10, 11].

3 Dependence properties

In the present section we first summarise some well-known standard notions of positive dependence viewed from a copula perspective; their probabilistic interpretation is presented in Figure 3 below. We then formally define the MK-TP2 property and discuss this property for different (parametric) copula families.

A copula C𝒞C\in\mathcal{C} is said to be

  • positively quadrant dependent (PQD) if C(u,v)Π(u,v)C(u,v)\geq\Pi(u,v) holds for all (u,v)(0,1)2(u,v)\in(0,1)^{2}.

  • left tail decreasing (LTD) if, for any v(0,1)v\in(0,1), the mapping (0,1)(0,1)\to\mathbb{R} given by uC(u,v)uu\mapsto\frac{C(u,v)}{u} is non-increasing.

  • stochastically increasing (SI) if, for (a version of) the Markov kernel KCK_{C} and any v(0,1)v\in(0,1), the mapping uKC(u,[0,v])u\mapsto K_{C}(u,[0,v]) is non-increasing.

According to [2] these three dependence properties are related as follows

SILTDPQDSI\Longrightarrow LTD\Longrightarrow PQD

Another notion of positive dependence that differs from the above mentioned dependence properties is total positivity of order 2 - a property applicable to various copula-related objects: the copula itself, its Markov kernel and its density, leading to three different but related positive dependence properties. In general, a function f:Ωf:\Omega\to\mathbb{R} with Ω2\Omega\subseteq\mathbb{R}^{2} is said to be totally positive of order 2 (TP2) on Ω\Omega if the inequality

f(u1,v1)f(u2,v2)f(u1,v2)f(u2,v1)0\displaystyle f(u_{1},v_{1})f(u_{2},v_{2})-f(u_{1},v_{2})f(u_{2},v_{1})\geq 0 (3.1)

holds for all u1u2u_{1}\leq u_{2} and all v1v2v_{1}\leq v_{2} such that [u1,u2]×[v1,v2]Ω[u_{1},u_{2}]\times[v_{1},v_{2}]\subseteq\Omega. The TP2 property has been extensively discussed for copulas and their densities (see, e.g., [7, 13, 3, 2]) and, to a certain extent, also for their partial derivatives respectively their conditional distribution functions (see, e.g., [15, 16]). A copula C𝒞C\in\mathcal{C} is said to be

  • TP2 if the copula itself is TP2, i.e., the inequality

    C(u1,v1)C(u2,v2)C(u1,v2)C(u2,v1)0C(u_{1},v_{1})\,C(u_{2},v_{2})-C(u_{1},v_{2})\,C(u_{2},v_{1})\geq 0 (3.2)

    holds for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all 0<v1v2<10<v_{1}\leq v_{2}<1; see Lemma 6.1 for a characterization in terms of the copula’s Markov kernel.

  • MK-TP2 (short for Markov kernel TP2) if (a version of) its Markov kernel is TP2, i.e., the inequality

    KC(u1,[0,v1])KC(u2,[0,v2])KC(u1,[0,v2])KC(u2,[0,v1])0K_{C}(u_{1},[0,v_{1}])\,K_{C}(u_{2},[0,v_{2}])-K_{C}(u_{1},[0,v_{2}])\,K_{C}(u_{2},[0,v_{1}])\geq 0 (3.3)

    holds for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all 0<v1v2<10<v_{1}\leq v_{2}<1.

If the copula has a density another dependence property can be formulated: CC is said to be

  • d-TP2 (short for density TP2) if (a version of) its density cc is TP2, i.e., the inequality

    c(u1,v1)c(u2,v2)c(u1,v2)c(u2,v1)0c(u_{1},v_{1})\,c(u_{2},v_{2})-c(u_{1},v_{2})\,c(u_{2},v_{1})\geq 0 (3.4)

    holds for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all 0<v1v2<10<v_{1}\leq v_{2}<1.

The property d-TP2 is also referred to as positively likelihood ratio dependence (see, e.g., [14]). Altogether the different notions of positive dependence are linked as depicted in Figure 2 (compare [16, 2]).

TP2d-TP2MK-TP2LTDPQDSI
Figure 2: Relations between the different notions of positive dependence.

Notice that the MK-TP2 property, unlike d-TP2, can be examined for any copula, making it presumably the strongest dependence property defined for any copula.

Remark 3.1.

By definition, a copula CC that is SI satisfies KC(u1,[0,v])KC(u2,[0,v])K_{C}(u_{1},[0,v])\geq K_{C}(u_{2},[0,v]) for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all v(0,1)v\in(0,1). Thus, the inequality

KC(u2,[0,v1])\displaystyle K_{C}(u_{2},[0,v_{1}]) \displaystyle\leq min{KC(u1,[0,v1]),KC(u2,[0,v2])}\displaystyle\min\{K_{C}(u_{1},[0,v_{1}]),K_{C}(u_{2},[0,v_{2}])\}
\displaystyle\leq max{KC(u1,[0,v1]),KC(u2,[0,v2])}\displaystyle\max\{K_{C}(u_{1},[0,v_{1}]),K_{C}(u_{2},[0,v_{2}])\}
\displaystyle\leq KC(u1,[0,v2])\displaystyle K_{C}(u_{1},[0,v_{2}])

holds for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all 0<v1v2<10<v_{1}\leq v_{2}<1. Therefore, to prove the MK-TP2 property for such copulas Equation (3.3) only needs to be shown for those rectangles 0<u1u2<10<u_{1}\leq u_{2}<1 and 0<v1v2<10<v_{1}\leq v_{2}<1 for which KC(u2,[0,v1])>0K_{C}(u_{2},[0,v_{1}])>0.

We now provide a probabilistic interpretation of the above-mentioned positive dependence properties in terms of continuous random variables XX and YY with connecting copula CC:

Dependence property Probabilistic interpretation
PQD (Yy|Xx)(Yy)\mathbb{P}(Y\leq y|X\leq x)\geq\mathbb{P}(Y\leq y)
LTD x(Yy|Xx)x\mapsto\mathbb{P}(Y\leq y|X\leq x) is non-increasing for any yy
SI x(Yy|X=x)x\mapsto\mathbb{P}(Y\leq y|X=x) is non-increasing (a.s.) for any yy
TP2 x(Yy|Xx)(Yy|Xx)x\mapsto\dfrac{\mathbb{P}(Y\leq y|X\leq x)}{\mathbb{P}(Y\leq y^{\prime}|X\leq x)} is non-increasing for any yyy\leq y^{\prime}
MK-TP2 x(Yy|X=x)(Yy|X=x)x\mapsto\dfrac{\mathbb{P}(Y\leq y|X=x)}{\mathbb{P}(Y\leq y^{\prime}|X=x)} is non-increasing (a.s.) for any yyy\leq y^{\prime}
d-TP2 xfY|X(y|x)fY|X(y|x)x\mapsto\dfrac{f_{Y|X}(y|x)}{f_{Y|X}(y^{\prime}|x)} is non-increasing (a.s.) for any yyy\leq y^{\prime}
Figure 3: Probabilistic interpretation of the positive dependence properties

In the remainder of this section, we examine the MK-TP2 property for different well-known copulas and copula families.

Example 3.2.
  1. 1.

    The independence copula Π\Pi is d-TP2 and hence MK-TP2.

  2. 2.

    The comonotonicity copulas MM is MK-TP2; since MM is singular, it cannot fulfill the d-TP2 property.

  3. 3.

    The countermonotonicity copula WW fulfills W(0.5,0.5)<Π(0.5,0.5)W(0.5,0.5)<\Pi(0.5,0.5). Thus it fails to be PQD and hence fails to satisfy any of the above-mentioned positive dependence properties.

The first two copula families under consideration - the Farlie-Gumbel-Morgenstern family of copulas (FGM copula) and the Gaussian family of copulas according to [1] - are exceptional since all notions of positive dependence depicted in (2) are equivalent:

Example 3.3.

(FGM copula)
For θ[1,1]\theta\in[-1,1], the mapping Cθ:𝕀2𝕀C_{\theta}:{\mathbb{I}}^{2}\to{\mathbb{I}} given by

Cθ(u,v):=uv+θuv(1u)(1v)\displaystyle C_{\theta}(u,v):=uv+\theta uv(1-u)(1-v)

is a copula and called Farlie-Gumbel-Morgenstern (FGM) copula. For an FGM copula CθC_{\theta} the following statements are equivalent:

  1. (a)

    CθC_{\theta} is PQD / LTD / SI / TP2 / MK-TP2 / d-TP2.

  2. (b)

    θ0\theta\geq 0.

Example 3.4.

(Gaussian copula)
For ρ(1,0)(0,1)\rho\in(-1,0)\cup(0,1), the mapping Cρ:𝕀2𝕀C_{\rho}:{\mathbb{I}}^{2}\to{\mathbb{I}} given by

Cρ(u,v)=(,ϕ1(u)]×(,ϕ1(v)]12π1ρ2exp(s22ρst+t22(1ρ2))dλ2(s,t)C_{\rho}(u,v)=\int\limits_{(-\infty,\phi^{-1}(u)]\times(-\infty,\phi^{-1}(v)]}\frac{1}{2\pi\sqrt{1-\rho^{2}}}\exp{\left(-\frac{s^{2}-2\rho st+t^{2}}{2(1-\rho^{2})}\right)}\;\mathrm{d}\lambda^{2}(s,t)

is a copula and called Gaussian copula, where ϕ1\phi^{-1} denotes the inverse of the standard Gaussian distribution function. According to [3], for a Gaussian copula CρC_{\rho} the following statements are equivalent:

  1. (a)

    CρC_{\rho} is PQD / LTD / SI / TP2 / MK-TP2 / d-TP2.

  2. (b)

    ρ(0,1)\rho\in(0,1).

We conclude this section with a brief discussion of the Fréchet family of copulas and the Marshall-Olkin family of copulas according to [1] - two copula subclasses for which the above-mentioned dependence properties are not equivalent; we thereby provide at the same time an outlook on a result from Section 5.

Example 3.5.

(Fréchet copula)
For α,β𝕀\alpha,\beta\in{\mathbb{I}} with α+β1\alpha+\beta\leq 1, the mapping Cα,β:𝕀2𝕀C_{\alpha,\beta}:{\mathbb{I}}^{2}\to{\mathbb{I}} given by

Cα,β(u,v)=αM(u,v)+(1αβ)Π(u,v)+βW(u,v)C_{\alpha,\beta}(u,v)=\alpha M(u,v)+(1-\alpha-\beta)\Pi(u,v)+\beta W(u,v)

is a copula and called Fréchet copula. The following statements can be easily verified:

  1. 1.

    Cα,βC_{\alpha,\beta} is PQD / LTD / SI / TP2 if and only if β=0\beta=0.

  2. 2.

    Cα,βC_{\alpha,\beta} is MK-TP2 if and only if α{0,1}\alpha\in\{0,1\} and β=0\beta=0.

  3. 3.

    Cα,βC_{\alpha,\beta} is d-TP2 if and only if α=1\alpha=1 and β=0\beta=0.

Example 3.6.

(Marshall-Olkin copula)
For α,β𝕀\alpha,\beta\in{\mathbb{I}}, the mapping Mα,β:𝕀2𝕀M_{\alpha,\beta}:{\mathbb{I}}^{2}\to{\mathbb{I}} given by

Mα,β(u,v)={u1αvuαvβuv1βuα<vβM_{\alpha,\beta}(u,v)=\begin{cases}u^{1-\alpha}v&u^{\alpha}\geq v^{\beta}\\ uv^{1-\beta}&u^{\alpha}<v^{\beta}\end{cases}

is a copula and called Marshall-Olkin (MO) copula. Since MO copulas are extreme value copulas (EVC) they are SI and TP2. If min{α,β}=0\min\{\alpha,\beta\}=0, then Mα,β=ΠM_{\alpha,\beta}=\Pi implying that in this case Mα,βM_{\alpha,\beta} is d-TP2. If, on the other hand, min{α,β}0\min\{\alpha,\beta\}\neq 0, Mα,βM_{\alpha,\beta} fails to be absolutely continuous (see, e.g., [1]) and hence fails to be d-TP2. In this case, the strongest verifiable notion of positive dependence is MK-TP2: According to Example 5.13, for α,β(0,1]\alpha,\beta\in(0,1], the following statements are equivalent:

  1. (a)

    Mα,βM_{\alpha,\beta} is MK-TP2.

  2. (b)

    β=1\beta=1.

In the next two sections we consider the family of Archimedean copulas and the family of extreme value copulas - both of which are copula families whose elements are generated by convex functions - and investigate under which conditions the corresponding copulas are MK-TP2.

4 Archimedean copulas

In this section we investigate under which conditions on the Archimedean generator (or its pseudo-inverse) the corresponding Archimedean copula is MK-TP2. It is well-known that an Archimedean copula CC is TP2 if and only if it is LTD (see, e.g., [2]). In what follows, we add a second characterisation to Figure 2 by showing that an Archimedean copula CC is MK-TP2 if and only if it is SI.

Recall that a generator of a bivariate Archimedean copula is a convex, strictly decreasing function φ:𝕀[0,]\varphi:{\mathbb{I}}\to[0,\infty] with φ(1)=0\varphi(1)=0 (see, e.g., [1, 2]). According to [10] we may, w.l.o.g., assume that all generators are right-continuous at 0. Every generator φ\varphi induces a symmetric copula CC via

C(u,v)=ψ(φ(u)+φ(v))\displaystyle C(u,v)=\psi(\varphi(u)+\varphi(v))

for all (u,v)𝕀2(u,v)\in{\mathbb{I}}^{2} where ψ:[0,]𝕀\psi:[0,\infty]\to{\mathbb{I}} denotes the pseudo-inverse of φ\varphi defined by

ψ(x):={φ1(x)if x[0,φ(0))0if xφ(0).\psi(x):=\begin{cases}\varphi^{-1}(x)&\text{if }x\in[0,\varphi(0))\\ 0&\text{if }x\geq\varphi(0).\end{cases}

The pseudo-inverse ψ\psi is convex, non-increasing and strictly decreasing on [0,φ(0))[0,\varphi(0)). Furthermore, it fulfills ψ(0)=1\psi(0)=1. Since most of the subsequent results are formulated in terms of ψ\psi, we call ψ\psi co-generator.
If φ(0)=\varphi(0)=\infty the induced copula CC is called strict, otherwise it is referred to as non-strict. Since for every generator φ\varphi and every constant a(0,)a\in(0,\infty) the generator aφa\varphi generates the same copula, w.l.o.g., we may assume that generators are normalized, i.e. φ(12)=1\varphi\big{(}\tfrac{1}{2}\big{)}=1, which allows for a one-to-one correspondence between generators and Archimedean copulas.
For every generator φ\varphi we denote by D+φ(u)D^{+}\varphi(u) (Dφ(u)D^{-}\varphi(u)) the right-hand (left-hand) derivative of φ\varphi at u(0,1)u\in(0,1), and for every co-generator ψ\psi we denote by D+ψ(x)D^{+}\psi(x) (Dψ(x)D^{-}\psi(x)) the right-hand (left-hand) derivative of ψ\psi at x(0,)x\in(0,\infty). Convexity of φ\varphi (respectively ψ\psi) implies that right-hand and left-hand derivative coincide almost surely, i.e., φ\varphi (respectively ψ\psi) is differentiable outside a countable subset of (0,1)(0,1) (respectively (0,)(0,\infty)). Setting D+φ(0)=D^{+}\varphi(0)=-\infty (Dψ()=0D^{-}\psi(\infty)=0) in case of strict CC allows to view D+φD^{+}\varphi (DψD^{-}\psi) as non-decreasing and right-continuous (left-continuous) function on [0,1)[0,1) ((0,](0,\infty]).

To investigate under which conditions an Archimedean copula CC is MK-TP2 we need to establish (a version of) the Markov kernel of CC. To this end, we first define the 0-level function f0:(0,1)𝕀f^{0}:(0,1)\to{\mathbb{I}} by f0(u):=ψ(φ(0)φ(u))f^{0}(u):=\psi(\varphi(0)-\varphi(u)).

Theorem 4.1.

Let CC be an Archimedean copula with generator φ\varphi and co-generator ψ\psi. If CC is strict, then

KC(u,[0,v])={1if u{0,1}Dψ(φ(u)+φ(v))Dψ(φ(u))if u(0,1)K_{C}(u,[0,v])=\begin{cases}1&\text{if }u\in\{0,1\}\\ \frac{D^{-}\psi(\varphi(u)+\varphi(v))}{D^{-}\psi(\varphi(u))}&\text{if }u\in(0,1)\end{cases}

is (a version of) its Markov kernel, and, if CC is non-strict, then

KC(u,[0,v])={1if u{0,1}Dψ(φ(u)+φ(v))Dψ(φ(u))if u(0,1) and vf0(u)0if u(0,1) and v<f0(u)K_{C}(u,[0,v])=\begin{cases}1&\text{if }u\in\{0,1\}\\ \frac{D^{-}\psi(\varphi(u)+\varphi(v))}{D^{-}\psi(\varphi(u))}&\text{if }u\in(0,1)\text{ and }v\geq f^{0}(u)\\ 0&\text{if }u\in(0,1)\text{ and }v<f^{0}(u)\end{cases}

is (a version of) its Markov kernel.

Proof.

According to [20, 10], if CC is strict, then

KC(u,[0,v])={1if u{0,1}D+φ(u)(D+φ)(C(u,v))if u(0,1)K_{C}(u,[0,v])=\begin{cases}1&\text{if }u\in\{0,1\}\\ \frac{D^{+}\varphi(u)}{\left(D^{+}\varphi\right)(C(u,v))}&\text{if }u\in(0,1)\end{cases}

is (a version of) its Markov kernel, and, if CC is non-strict, then

KC(u,[0,v])={1if u{0,1}D+φ(u)(D+φ)(C(u,v))if u(0,1) and vf0(u)0if u(0,1) and v<f0(u)K_{C}(u,[0,v])=\begin{cases}1&\text{if }u\in\{0,1\}\\ \frac{D^{+}\varphi(u)}{\left(D^{+}\varphi\right)(C(u,v))}&\text{if }u\in(0,1)\text{ and }v\geq f^{0}(u)\\ 0&\text{if }u\in(0,1)\text{ and }v<f^{0}(u)\end{cases}

is (a version of) its Markov kernel. Since φ\varphi is decreasing, for t(0,1)t\in(0,1) if CC is strict (and for t[0,1)t\in[0,1) if CC is non-strict), h>0h>0 such that t+h(0,1)t+h\in(0,1), and l:=φ(t+h)φ(t)<0l:=\varphi(t+h)-\varphi(t)<0, we first have

φ(t+h)φ(t)h=1ψ(φ(t+h))ψ(φ(t))φ(t+h)φ(t)=1ψ(φ(t)+l)ψ(φ(t))l\frac{\varphi(t+h)-\varphi(t)}{h}=\frac{1}{\frac{\psi(\varphi(t+h))-\psi(\varphi(t))}{\varphi(t+h)-\varphi(t)}}=\frac{1}{\frac{\psi(\varphi(t)+l)-\psi(\varphi(t))}{l}}

and hence

D+φ(t)=limh0+φ(t+h)φ(t)h=liml01ψ(φ(t)+l)ψ(φ(t))l=1Dψ(φ(t))D^{+}\varphi(t)=\lim_{h\to 0^{+}}\frac{\varphi(t+h)-\varphi(t)}{h}=\lim_{l\to 0^{-}}\frac{1}{\frac{\psi(\varphi(t)+l)-\psi(\varphi(t))}{l}}=\frac{1}{D^{-}\psi(\varphi(t))}

Now, consider u(0,1)u\in(0,1). If CC is strict,

D+φ(u)(D+φ)(C(u,v))\displaystyle\frac{D^{+}\varphi(u)}{\left(D^{+}\varphi\right)(C(u,v))} ={Dψ(φ(u)+φ(v))Dψ(φ(u))if φ(u)+φ(v)<φ(0)0if φ(u)+φ(v)=φ(0)\displaystyle=\begin{cases}\frac{D^{-}\psi(\varphi(u)+\varphi(v))}{D^{-}\psi(\varphi(u))}&\textrm{if }\varphi(u)+\varphi(v)<\varphi(0)\\ 0&\textrm{if }\varphi(u)+\varphi(v)=\varphi(0)\end{cases}
=Dψ(φ(u)+φ(v))Dψ(φ(u))\displaystyle=\frac{D^{-}\psi(\varphi(u)+\varphi(v))}{D^{-}\psi(\varphi(u))}

If CC is non-strict, then vf0(u)v\geq f^{0}(u) if and only if φ(u)+φ(v)φ(0)\varphi(u)+\varphi(v)\leq\varphi(0) and in this case

D+φ(u)(D+φ)(C(u,v))\displaystyle\frac{D^{+}\varphi(u)}{\left(D^{+}\varphi\right)(C(u,v))} ={Dψ(φ(u)+φ(v))Dψ(φ(u))if φ(u)+φ(v)<φ(0)Dψ(φ(0))Dψ(φ(u))if φ(u)+φ(v)=φ(0)\displaystyle=\begin{cases}\frac{D^{-}\psi(\varphi(u)+\varphi(v))}{D^{-}\psi(\varphi(u))}&\textrm{if }\varphi(u)+\varphi(v)<\varphi(0)\\ \frac{D^{-}\psi(\varphi(0))}{D^{-}\psi(\varphi(u))}&\textrm{if }\varphi(u)+\varphi(v)=\varphi(0)\end{cases}
=Dψ(φ(u)+φ(v))Dψ(φ(u))\displaystyle=\frac{D^{-}\psi(\varphi(u)+\varphi(v))}{D^{-}\psi(\varphi(u))}

This proves the assertion. ∎

For ψ\psi twice differentiable, (a version of) the Markov kernel of an Archimedean copula is given in [7, 17].

For completeness, we repeat the following necessary and sufficient condition for an Archimedean copula to be TP2 which is well known, see, e.g., [2]; having Lemma 6.1 in mind, the equivalence of (b) and (c) is straightforward. A function f:[0,](0,)f:[0,\infty]\to(0,\infty) is called log-convex if logf|(0,)\log f|_{(0,\infty)} is convex.

Proposition 4.2.

Let CC be an Archimedean copula with generator φ\varphi and co-generator ψ\psi. Then the following statements are equivalent:

  1. (a)

    CC is LTD.

  2. (b)

    CC is TP2.

  3. (c)

    ψ\psi is log-convex.

In the sequel we examine under which conditions on the co-generator ψ\psi the corresponding Archimedean copula CC is MK-TP2. We prove the result by showing that SI implies log-convexity of Dψ-D^{-}\psi (Lemma 4.4) which in turn implies MK-TP2 (Lemma 4.5). In passing we slightly extend a well-known result from literature (see, e.g., [7, 17]) by proving the equivalence of SI and log-convexity of Dψ-D^{-}\psi for arbitrary Archimedean copulas (i.e., CC needs not to be strict and ψ\psi needs not to be twice differentiable).

Recall that every TP2 (and also every SI) copula satisfies C(u,v)>0C(u,v)>0 for all (u,v)(0,1)2(u,v)\in(0,1)^{2}. In the Archimedean setting, the latter is equivalent to v>f0(u)v>f^{0}(u) for all (u,v)(0,1)2(u,v)\in(0,1)^{2}. This implies that a non-strict Archimedean copula is neither TP2 nor SI and hence not MK-TP2. Therefore, from now on we can restrict ourselves to strict Archimedean copulas. The following lemma imposes further restrictions on an Archimedean co-generator for the corresponding copula CC to be MK-TP2:

Lemma 4.3.

Let CC be a strict Archimedean copula with generator φ\varphi and co-generator ψ\psi. If DψD^{-}\psi has a discontinuity point then CC is not SI and hence not MK-TP2.

Proof.

Recall that CC is SI if and only if uKC(u,[0,v])u\mapsto K_{C}(u,[0,v]) is non-increasing. Further recall that DψD^{-}\psi is non-decreasing, left-continuous on (0,](0,\infty] and, since CC is strict, Dψ(x)<0D^{-}\psi(x)<0 holds for all x(0,)x\in(0,\infty).
Since DψD^{-}\psi has a discontinuity point there exists x(0,)x^{\ast}\in(0,\infty) and δ(0,1)\delta\in(0,1) such that y:=Dψ(x)<Dψ(x+)=y+δ<0y:=D^{-}\psi(x^{\ast})<D^{-}\psi(x^{\ast}+)=y+\delta<0. Further, there exists z(0,x)z^{\ast}\in(0,x^{\ast}) such that zDψ(z)z\mapsto D^{-}\psi(z) is continuous on (z,x)(z^{\ast},x^{\ast}). Choosing ε(0,δy/(y+δ))\varepsilon\in\big{(}0,\delta y/(y+\delta)\big{)}, there exists some w(z,x)w^{\ast}\in(z^{\ast},x^{\ast}) such that yεDψ(z)yy-\varepsilon\leq D^{-}\psi(z)\leq y for all z[w,x]z\in[w^{\ast},x^{\ast}]. Now, consider u1=ψ(x)u_{1}=\psi(x^{\ast}), u2=ψ(w)u_{2}=\psi(w^{\ast}), v1=ψ(1/n)v_{1}=\psi(1/n) for some nn\in\mathbb{N} such that w+1/n<xw^{\ast}+1/n<x^{\ast}, and v2:=1v_{2}:=1. Then, u1<u2u_{1}<u_{2}, v1<v2v_{1}<v_{2} and monotonicity implies

x+1/n\displaystyle x^{\ast}+1/n =\displaystyle= φ(u1)+φ(v1)\displaystyle\varphi(u_{1})+\varphi(v_{1})
\displaystyle\geq max{φ(u1)+φ(v2),φ(u2)+φ(v1)}\displaystyle\max\{\varphi(u_{1})+\varphi(v_{2}),\varphi(u_{2})+\varphi(v_{1})\}
\displaystyle\geq min{φ(u1)+φ(v2),φ(u2)+φ(v1)}\displaystyle\min\{\varphi(u_{1})+\varphi(v_{2}),\varphi(u_{2})+\varphi(v_{1})\}
\displaystyle\geq φ(u2)+φ(v2)\displaystyle\varphi(u_{2})+\varphi(v_{2})
=\displaystyle= w\displaystyle w^{\ast}

Thus, using φ(v2)=0\varphi(v_{2})=0

KC(u2,[0,v1])KC(u1,[0,v1])\displaystyle\frac{K_{C}(u_{2},[0,v_{1}])}{K_{C}(u_{1},[0,v_{1}])} =Dψ(φ(u2)+φ(v1))Dψ(φ(u1)+φ(v2))Dψ(φ(u2)+φ(v2))Dψ(φ(u1)+φ(v1))\displaystyle=\frac{D^{-}\psi(\varphi(u_{2})+\varphi(v_{1}))\,D^{-}\psi(\varphi(u_{1})+\varphi(v_{2}))}{D^{-}\psi(\varphi(u_{2})+\varphi(v_{2}))\,D^{-}\psi(\varphi(u_{1})+\varphi(v_{1}))} (4.1)
=Dψ(w+1/n)Dψ(x)Dψ(w)Dψ(x+1/n)=Dψ(w+1/n)yDψ(w)Dψ(x+1/n)\displaystyle=\frac{D^{-}\psi(w^{\ast}+1/n)\,D^{-}\psi(x^{\ast})}{D^{-}\psi(w^{\ast})\,D^{-}\psi(x^{\ast}+1/n)}=\frac{D^{-}\psi(w^{\ast}+1/n)\,y}{D^{-}\psi(w^{\ast})\,D^{-}\psi(x^{\ast}+1/n)}
y2(yε)(y+δ)>y2(yδyy+δ)(y+δ)=1\displaystyle\geq\frac{y^{2}}{(y-\varepsilon)(y+\delta)}>\frac{y^{2}}{\big{(}y-\tfrac{\delta y}{y+\delta}\big{)}(y+\delta)}=1

This proves the assertion. ∎

Equation (4.1) in the proof of Lemma 4.3 shows very clearly that in the Archimedean case the SI property is actually a TP2 property.

The next lemma provides a necessary condition for an Archimedean copula to be SI:

Lemma 4.4.

Let CC be an Archimedean copula with generator φ\varphi and co-generator ψ\psi. If CC is SI, then CC is strict, DψD^{-}\psi is continuous and Dψ-D^{-}\psi is log-convex.

Proof.

Setting f:=log(Dψ)f:=\log(-D^{-}\psi), the copula CC is SI if and only if

0\displaystyle 0 \displaystyle\leq log(KC(u1,[0,v])KC(u2,[0,v]))\displaystyle\log\left(\frac{K_{C}(u_{1},[0,v])}{K_{C}(u_{2},[0,v])}\right)
=\displaystyle= log(Dψ(φ(u1)+φ(v))Dψ(φ(u2))Dψ(φ(u1))Dψ(φ(u2)+φ(v)))\displaystyle\log\left(\frac{D^{-}\psi(\varphi(u_{1})+\varphi(v))\,D^{-}\psi(\varphi(u_{2}))}{D^{-}\psi(\varphi(u_{1}))\,D^{-}\psi(\varphi(u_{2})+\varphi(v))}\right)
=\displaystyle= f(φ(u1)+φ(v))+f(φ(u2))f(φ(u1))f(φ(u2)+φ(v))()\displaystyle f(\varphi(u_{1})+\varphi(v))+f(\varphi(u_{2}))-f(\varphi(u_{1}))-f(\varphi(u_{2})+\varphi(v))\qquad\qquad(\ast)

for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all v(0,1)v\in(0,1) for which Dψ(φ(u1)+φ(v))<0D^{-}\psi(\varphi(u_{1})+\varphi(v))<0.
Assume that CC is SI. Then CC is strict (and hence Dψ(x)<0D^{-}\psi(x)<0 for all x(0,)x\in(0,\infty)) and, by Lemma 4.3, DψD^{-}\psi is continuous. We now prove that ff is convex. Since ff is continuous, it remains to show mid-point convexity. To this end, consider x,y(0,)x,y\in(0,\infty) with x<yx<y and set u1:=ψ((x+y)/2)u_{1}:=\psi((x+y)/2), u2:=ψ(x)u_{2}:=\psi(x), and v:=ψ((yx)/2)v:=\psi((y-x)/2). Then (\ast) implies

0\displaystyle 0 \displaystyle\leq f(x+y2+yx2)+f(x)f(x+y2)f(x+yx2)\displaystyle f\big{(}\tfrac{x+y}{2}+\tfrac{y-x}{2}\big{)}+f(x)-f\big{(}\tfrac{x+y}{2}\big{)}-f\big{(}x+\tfrac{y-x}{2}\big{)}
=\displaystyle= f(y)+f(x)2f(x+y2)\displaystyle f(y)+f(x)-2f\big{(}\tfrac{x+y}{2}\big{)}

and hence f((x+y)/2)(f(y)+f(x))/2f\big{(}(x+y)/2\big{)}\leq\big{(}f(y)+f(x)\big{)}/2. This proves the assertion. ∎

The next result provides a sufficient condition for an Archimedean copula to be MK-TP2:

Lemma 4.5.

Let CC be an Archimedean copula with generator φ\varphi and co-generator ψ\psi. If Dψ-D^{-}\psi is log-convex, then CC is MK-TP2.

Proof.

Defining the functions f:(,0)f:(-\infty,0)\to\mathbb{R} and G:(0,1)2(,0)G:(0,1)^{2}\to(-\infty,0) by f(x):=log(Dψ(x))f(x):=\log(-D^{-}\psi(-x)) and G(u,v):=(φ(u)+φ(v))G(u,v):=-(\varphi(u)+\varphi(v)), the copula CC is MK-TP2 if

0\displaystyle 0 \displaystyle\leq log(KC(u1,[0,v1])KC(u2,[0,v2])KC(u1,[0,v2])KC(u2,[0,v1]))\displaystyle\log\left(\frac{K_{C}(u_{1},[0,v_{1}])\,K_{C}(u_{2},[0,v_{2}])}{K_{C}(u_{1},[0,v_{2}])\,K_{C}(u_{2},[0,v_{1}])}\right)
=\displaystyle= log(Dψ(G(u1,v1))Dψ(G(u2,v2))Dψ(G(u1,v2))Dψ(G(u2,v1)))\displaystyle\log\left(\frac{D^{-}\psi(-G(u_{1},v_{1}))\,D^{-}\psi(-G(u_{2},v_{2}))}{D^{-}\psi(-G(u_{1},v_{2}))\,D^{-}\psi(-G(u_{2},v_{1}))}\right)
=\displaystyle= (fG)(u1,v1)+(fG)(u2,v2)(fG)(u1,v2)(fG)(u2,v1)\displaystyle(f\circ G)(u_{1},v_{1})+(f\circ G)(u_{2},v_{2})-(f\circ G)(u_{1},v_{2})-(f\circ G)(u_{2},v_{1})

for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all 0<v1v2<10<v_{1}\leq v_{2}<1 for which Dψ(G(u1,v1))<0D^{-}\psi(-G(u_{1},v_{1}))<0. Since GG is a 22-increasing and coordinatewise non-decreasing function and ff is non-decreasing and, by assumption, convex (recall that convexity of log(Dψ)\log(-D^{-}\psi) implies convexity of ff), it follows from [21, p.219] that the composition fGf\circ G is 22-increasing. This proves the assertion. ∎

Combining Figure 2, Lemma 4.5 and Lemma 4.4, we are now in the position to state the main result of this section: an Archimedean copula is MK-TP2 if and only if it is SI if and only if Dψ-D^{-}\psi is log-convex. Recall that we slightly extend a well-known result from literature (see, e.g., [7, 17]) by proving the equivalence of (b) and (c) for arbitrary Archimedean copulas (i.e., CC needs not to be strict and ψ\psi needs not to be twice differentiable).

Theorem 4.6.

Let CC be an Archimedean copula with generator φ\varphi and pseudo-inverse ψ\psi. Then the following statements are equivalent:

  1. (a)

    CC is MK-TP2.

  2. (b)

    CC is SI.

  3. (c)

    Dψ-D^{-}\psi is log-convex.

The previous result is remarkable since for Archimedean copulas the dependence properties SI and MK-TP2 are equivalent.

Remark 4.7.

According to [7], for an Archimedean copula with twice differentiable co-generator ψ\psi, the following statements are equivalent:

  1. (a)

    CC is d-TP2.

  2. (b)

    ψ′′\psi^{{}^{\prime\prime}} is log-convex.

In sum, for Archimedean copulas, Figure 2 reduces to

d-TP2(1)MK-TP2SI(2)TP2LTDPQD\textrm{d-TP2}\overset{(1)}{\Longrightarrow}\textrm{MK-TP2}\Longleftrightarrow\textrm{SI}\overset{(2)}{\Longrightarrow}\textrm{TP2}\Longleftrightarrow\textrm{LTD}\Longrightarrow\textrm{PQD}

Remark 2.13 in [17] provides an Archimedean copula that is MK-TP2 and SI, but not d-TP2, i.e., the reverse of (1) does not hold, in general. An Archimedean copula that is LTD and TP2 but not MK-TP2 is given in [22, Example 19]: The mapping ψ:[0,]𝕀\psi:[0,\infty]\to{\mathbb{I}} given by ψ(x):=(x+1+x2)1/10\psi(x):=\big{(}x+\sqrt{1+x^{2}}\big{)}^{-1/10} is an Archimedean co-generator which is log-convex. However, since ψ-\psi^{\prime} is not log-convex, the reverse of (2) does not hold, in general.

An Archimedean co-generator ψ\psi is said to be completely monotone if its restiction ψ|(0,)\psi|_{(0,\infty)} has derivatives of all orders and satisfies (1)nψ(n)(x)0(-1)^{n}\psi^{(n)}(x)\geq 0 for all x(0,)x\in(0,\infty) and all nn\in\mathbb{N}, and it follows from [23] that completely monotone functions are log-convex. If ψ\psi is completely monotone then its negative derivative is completely monotone as well implying that (1)nψ(n)(-1)^{n}\psi^{(n)} is log-convex for all nn\in\mathbb{N}. Therefore, every Archimedean copula with completely monotone co-generator is d-TP2 and hence MK-TP2.

Example 4.8.

(Gumbel copula)
For α[1,)\alpha\in[1,\infty), the mapping ψα:[0,]𝕀\psi_{\alpha}:[0,\infty]\to{\mathbb{I}} given by

ψα(x):=exp(x1/αlog(2))\psi_{\alpha}(x):=\exp\big{(}-x^{1/\alpha}\,\log(2)\big{)}

is a (normalized) Archimedean co-generator and the corresponding Archimedean copula is called Gumbel copula. ψ\psi is completely monotone (see, e.g., [2]) and hence d-TP2 as well as MK-TP2.

Recall that Gumbel copulas are Archimedean and extreme value copulas (EVC) at the same time. EVCs are subject of the next section.

Further examples of Archimedean copulas with completely monotone co-generator may be found in [2, Chapter 4].

5 Extreme Value copulas

It is well-known that every extreme value copula (EVC) is TP2 (see, e.g., [13]) and SI (see, e.g., [16]). However, EVCs may fail to be MK-TP2, in general; see Example 5.13 below or [16]. In what follows we answer the question under which conditions on the Pickands dependence function the corresponding EVC is MK-TP2.

According to [1], a copula CC is called extreme value copula if one of the following two equivalent conditions is fulfilled (see also [24, 25, 26]):

  1. 1.

    There exists a copula B𝒞B\in\mathcal{C} such that for all (u,v)𝕀2(u,v)\in{\mathbb{I}}^{2} we have

    C(u,v)=limnBn(u1n,v1n)C(u,v)=\lim_{n\to\infty}B^{n}(u^{\frac{1}{n}},v^{\frac{1}{n}})
  2. 2.

    There exists a convex function A:𝕀𝕀A:{\mathbb{I}}\to{\mathbb{I}} fulfilling max{1t,t}A(t)1\max\{1-t,t\}\leq A(t)\leq 1 for all t𝕀t\in{\mathbb{I}} such that

    C(u,v)=(uv)A(ln(u)ln(uv))C(u,v)=\left(uv\right)^{A\left(\frac{\ln(u)}{\ln(uv)}\right)}

    for all (u,v)(0,1)2(u,v)\in(0,1)^{2}. AA is called Pickands dependence function.

Let 𝒜\mathcal{A} denote the class of all Pickands dependence function. Following [10, 27], for A𝒜A\in\mathcal{A} we will let D+AD^{+}A denote the right-hand derivative of AA on [0,1)[0,1) and DAD^{-}A the left-hand derivative of AA on (0,1](0,1]. Convexity of AA implies that it is differentiable outside a countable subset of (0,1)(0,1), i.e. the identity D+A(t)=DA(t)D^{+}A(t)=D^{-}A(t) holds for λ\lambda-almost every t(0,1)t\in(0,1). Furthermore D+AD^{+}A is non-decreasing and right-continuous. For more information on Pickands dependence functions and the approach via right-hand derivatives we refer to [28, 29].

In order to simplify notation, for a Pickands dependence function AA we define the function FA:[0,1)F_{A}:[0,1)\to\mathbb{R} by letting

FA(t):=A(t)+(1t)D+A(t)F_{A}(t):=A(t)+(1-t)D^{+}A(t)

and set FA(1):=1F_{A}(1):=1. Then, convexity of AA implies that FAF_{A} is non-decreasing and non-negative (see, e.g., [27, Lemma 5]). Additionally, we consider the function h:(0,1)2h:(0,1)^{2}\to\mathbb{R} given by h(u,v):=log(u)/log(uv)h(u,v):=\log(u)/\log(uv) and recall that uh(u,v)u\mapsto h(u,v) is strictly decreasing for all v(0,1)v\in(0,1), and that vh(u,v)v\mapsto h(u,v) is strictly increasing for all u(0,1)u\in(0,1). Moreover, for t(0,1)t\in(0,1), we define the function ft:𝕀𝕀f_{t}:{\mathbb{I}}\to{\mathbb{I}} by ft(u):=u1ttf_{t}(u):=u^{\frac{1-t}{t}}, and note that ftf_{t} is increasing and convex whenever t(0,1/2]t\in(0,1/2] and concave whenever t[1/2,1)t\in[1/2,1). Simple calculation yields h(u,ft(u))=th(u,f_{t}(u))=t.

According to [27, Lemma 3] and using the above-introduced functions,

KA(u,[0,v])={1if u{0,1}C(u,v)uFA(h(u,v))if u,v(0,1)vif (u,v)(0,1)×{0,1}K_{A}(u,[0,v])=\begin{cases}1&\text{if }u\in\{0,1\}\\ \frac{C(u,v)}{u}F_{A}(h(u,v))&\text{if }u,v\in(0,1)\\ v&\text{if }(u,v)\in(0,1)\times\{0,1\}\end{cases} (5.1)

is a (version of the) Markov kernel of the extreme value copula CC with Pickands dependence function AA.


In what follows we give sufficient and necessary conditions on AA for the corresponding EVC to be MK-TP2. As we will show, the behaviour of D+AD^{+}A at point 0 is crucial. The first result is evident:

Lemma 5.1.

D+A(0)=0D^{+}A(0)=0 if and only if A1A\equiv 1. In this case C=ΠC=\Pi is d-TP2 and hence MK-TP2.

Next we consider the class of Pickands dependence functions fulfilling D+A(0)(1,0)D^{+}A(0)\in(-1,0) and show that the corresponding EVCs fail to be MK-TP2.

Lemma 5.2.

If D+A(0)(1,0)D^{+}A(0)\in(-1,0) then the corresponding EVC is not MK-TP2.

Proof.

To prove the result we proceed in two steps: first, we consider the case when D+AD^{+}A has at least one discontinuity point and in the second step we assume that D+AD^{+}A is continuous.

First, suppose there exists some t1(0,1)t_{1}\in(0,1) such that D+A(t1)DA(t1)D^{+}A(t_{1})\neq D^{-}A(t_{1}). By assumption, D+A(0)(1,0)D^{+}A(0)\in(-1,0) which implies FA(0)=1+D+A(0)(0,1]F_{A}(0)=1+D^{+}A(0)\in(0,1]. Since FAF_{A} is non-decreasing and continuous outside of a countable subset of (0,1)(0,1) there further exists some t2(0,t1)t_{2}\in(0,t_{1}) such that FA(t)>0F_{A}(t)>0 for all t[t2,t1)t\in[t_{2},t_{1}) and tFA(t)t\mapsto F_{A}(t) is continuous on [t2,t1)[t_{2},t_{1}). Lemma 6.3 hence implies that the corresponding EVC is not MK-TP2.

Suppose now that D+AD^{+}A and thus also FAF_{A} are continuous. Notice that, since hh is continuous, the mapping uKA(u,[0,v])u\mapsto K_{A}(u,[0,v]) is continuous and non-increasing for every v(0,1)v\in(0,1) where the latter property follows from the fact that every EVC is stochastically increasing.
Define βA(0,)\beta_{A}\in(0,\infty) and g:[0,βA]g:[0,\beta_{A}]\to\mathbb{R} according to Lemma 6.4. Then, by Lemma 6.4, g(α)>g(βA)g(\alpha)>g(\beta_{A}), FA(1/(1+βA))<FA(1/(1+α))F_{A}\big{(}1/(1+\beta_{A})\big{)}<F_{A}\big{(}1/(1+\alpha)\big{)} and the inequality

γα:=log(FA(11+βA)FA(11+α))g(α)g(βA)<0\gamma_{\alpha}:=\frac{\log\left(\frac{F_{A}\big{(}\tfrac{1}{1+\beta_{A}}\big{)}}{F_{A}\big{(}\tfrac{1}{1+\alpha}\big{)}}\right)}{g(\alpha)-g(\beta_{A})}<0

holds for all α(0,βA)\alpha\in(0,\beta_{A}). Now, fix α(0,βA)\alpha\in(0,\beta_{A}) and choose u:=exp(γα/2)(exp(γα),1)u:=\exp(\gamma_{\alpha}/2)\in(\exp(\gamma_{\alpha}),1) and set u1:=uu_{1}:=u, v1:=uβAv_{1}:=u^{\beta_{A}}, v2:=uαv_{2}:=u^{\alpha} and u2,n:=11/nu_{2,n}:=1-1/n for nn\in\mathbb{N} with u<11/nu<1-1/n (as illustrated in Figure 4).

1111uuu2,nu_{2,n}v2v_{2}v1v_{1}f1/(1+α)f_{1/(1+\alpha)}f1/(1+βA)f_{1/(1+\beta_{A})}
Figure 4: Blue lines are the contour lines of f1/(1+α)f_{1/(1+\alpha)} and f1/(1+βA)f_{1/(1+\beta_{A})} with α<1<βA\alpha<1<\beta_{A}.

Then

limnKA(u2,n,[0,v])=limnC(u2,n,v)u2,nFA(h(u2,n,v))=vFA(0)>0\lim_{n\to\infty}K_{A}(u_{2,n},[0,v])=\lim_{n\to\infty}\frac{C(u_{2,n},v)}{u_{2,n}}F_{A}(h(u_{2,n},v))=vF_{A}(0)>0

for all v(0,1)v\in(0,1) and hence

limnKA(u2,n,[0,v2])KA(u2,n,[0,v1])=uαFA(0)uβAFA(0)=uαβA=exp(γα2(αβA))>1\lim_{n\to\infty}\frac{K_{A}(u_{2,n},[0,v_{2}])}{K_{A}(u_{2,n},[0,v_{1}])}=\frac{u^{\alpha}\,F_{A}(0)}{u^{\beta_{A}}\,F_{A}(0)}=u^{\alpha-\beta_{A}}=\exp\big{(}\tfrac{\gamma_{\alpha}}{2}(\alpha-\beta_{A})\big{)}>1

Continuity then implies the existence of some nn^{\ast}\in\mathbb{N} with nnn^{\ast}\geq n such that

KA(u2,n,[0,v2])KA(u2,n,[0,v1])<(FA(11+α)FA(11+βA))1/2>1uαβA\frac{K_{A}(u_{2,n^{\ast}},[0,v_{2}])}{K_{A}(u_{2,n^{\ast}},[0,v_{1}])}<\underbrace{\left(\frac{F_{A}\left(\tfrac{1}{1+\alpha}\right)}{F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}\right)^{1/2}}_{>1}u^{\alpha-\beta_{A}}

Moreover, we have

KA(u1,[0,v1])=C(u,uβA)uFA(11+βA)=u(1+βA)A(11+βA)1FA(11+βA)>0K_{A}(u_{1},[0,v_{1}])=\frac{C(u,u^{\beta_{A}})}{u}F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)=u^{(1+\beta_{A})A\big{(}\tfrac{1}{1+\beta_{A}}\big{)}-1}F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)>0

and

KA(u1,[0,v2])=C(u,uα)uFA(11+α)=u(1+α)A(11+α)1FA(11+α)>0K_{A}(u_{1},[0,v_{2}])=\frac{C(u,u^{\alpha})}{u}F_{A}\left(\tfrac{1}{1+\alpha}\right)=u^{(1+\alpha)A\left(\frac{1}{1+\alpha}\right)-1}F_{A}\left(\tfrac{1}{1+\alpha}\right)>0

Altogether, we finally obtain

KA(u1,[0,v1])KA(u2,n,[0,v2])KA(u1,[0,v2])KA(u2,n,[0,v1])\displaystyle\frac{K_{A}(u_{1},[0,v_{1}])\,K_{A}(u_{2,n^{\ast}},[0,v_{2}])}{K_{A}(u_{1},[0,v_{2}])\,K_{A}(u_{2,n^{\ast}},[0,v_{1}])} =\displaystyle= KA(u2,n,[0,v2])KA(u2,n,[0,v1])u(1+βA)A(11+βA)1FA(11+βA)u(1+α)A(11+α)1FA(11+α)\displaystyle\frac{K_{A}(u_{2,n^{\ast}},[0,v_{2}])}{K_{A}(u_{2,n^{\ast}},[0,v_{1}])}\;\frac{u^{(1+\beta_{A})A\big{(}\tfrac{1}{1+\beta_{A}}\big{)}-1}F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}{u^{(1+\alpha)A\left(\frac{1}{1+\alpha}\right)-1}F_{A}\left(\tfrac{1}{1+\alpha}\right)}
<\displaystyle< (FA(11+α)FA(11+βA))1/2uαβAug(βA)+βAFA(11+βA)ug(α)+αFA(11+α)\displaystyle\left(\frac{F_{A}\left(\tfrac{1}{1+\alpha}\right)}{F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}\right)^{1/2}u^{\alpha-\beta_{A}}\;\frac{u^{g(\beta_{A})+\beta_{A}}F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}{u^{g(\alpha)+\alpha}F_{A}\left(\tfrac{1}{1+\alpha}\right)}
=\displaystyle= (FA(11+α)FA(11+βA))1/2ug(βA)g(α)FA(11+βA)FA(11+α)\displaystyle\left(\frac{F_{A}\left(\tfrac{1}{1+\alpha}\right)}{F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}\right)^{1/2}\;u^{g(\beta_{A})-g(\alpha)}\;\frac{F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}{F_{A}\left(\tfrac{1}{1+\alpha}\right)}
=\displaystyle= (FA(11+α)FA(11+βA))1/2(FA(11+α)FA(11+βA))1/2(FA(11+βA)FA(11+α))\displaystyle\left(\frac{F_{A}\left(\tfrac{1}{1+\alpha}\right)}{F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}\right)^{1/2}\;\left(\frac{F_{A}\left(\tfrac{1}{1+\alpha}\right)}{F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}\right)^{1/2}\;\left(\frac{F_{A}\left(\tfrac{1}{1+\beta_{A}}\right)}{F_{A}\left(\tfrac{1}{1+\alpha}\right)}\right)
=\displaystyle= 1\displaystyle 1

Thus, the corresponding EVC is not MK-TP2. This proves the result. ∎

Example 5.3.

In [16] the authors mentioned that the EVC CθC_{\theta} occuring in Tawn’s model [30] and generated by the Pickands dependence function Aθ(t):=θt2θt+1A_{\theta}(t):=\theta t^{2}-\theta t+1, θ𝕀\theta\in{\mathbb{I}}, is MK-TP2 for θ=1\theta=1 but fails to be MK-TP2 for θ=1/5\theta=1/5.
Due to Lemma 5.1 and Lemma 5.2 we are now in the position to finally name exactly those EVCs in that class that are MK-TP2: Since D+(0)=θD^{+}(0)=-\theta, CθC_{\theta} is MK-TP2 if and only if θ{0,1}\theta\in\{0,1\}.

Due to Lemma 5.1 and Lemma 5.2, it remains to discuss only those EVCs for which the corresponding Pickands dependence function AA satisfies D+A(0)=1D^{+}A(0)=-1 (or, equivalently, FA(0)=0F_{A}(0)=0). We first focus on how many discontinuity points can occur in this setting:

Lemma 5.4.

Suppose that D+A(0)=1D^{+}A(0)=-1 and that D+AD^{+}A has at least two discontinuity points. Then the corresponding EVC is not MK-TP2.

Proof.

Denote by t1t_{1} and t2t_{2} two discontinuity points of D+AD^{+}A such that tD+A(t)t\mapsto D^{+}A(t) is continuous on [t1,t2)[t_{1},t_{2}). Then t1t_{1} and t2t_{2} are discontinuity points of FAF_{A}, FA(t)>0F_{A}(t)>0 for all t[t1,t2)t\in[t_{1},t_{2}) and the mapping tFA(t)t\mapsto F_{A}(t) is continuous on [t1,t2)[t_{1},t_{2}). The assertion now follows from Lemma 6.3. ∎

The previous lemma states that the Pickands dependence function of an EVC that is MK-TP2 can have at most one discontinuity point. We can even make the statement more precise for this case.

Lemma 5.5.

Suppose that D+A(0)=1D^{+}A(0)=-1 and that D+AD^{+}A has exactly one discontinuity point at t(0,1)t^{\ast}\in(0,1). If the corresponding EVC is MK-TP2 then AA fulfills A(t)=1tA(t)=1-t for all t[0,t]t\in[0,t^{\ast}]. In this case, t1/2t^{\ast}\leq 1/2.

Proof.

Assume that the EVC is MK-TP2. By Lemma 6.3, we immediately obtain FA(t)=0F_{A}(t)=0 for all t[0,t)t\in[0,t^{\ast}) and hence A(t)=1tA(t)=1-t for all t[0,t]t\in[0,t^{\ast}]. Thus, t1/2t^{\ast}\leq 1/2. This proves the assertion. ∎

Although quite simple, Lemma 5.5 is crucial, as from now on we can restrict ourselves to only those D+AD^{+}A for which there exists some t[0,1/2]t^{\ast}\in[0,1/2] such that D+A(t)=1D^{+}A(t)=-1 for all t[0,t)t\in[0,t^{\ast}) and being continuous on [t,1][t^{\ast},1]. This is equivalent to considering only those FAF_{A} for which there exists some t[0,1/2]t^{\ast}\in[0,1/2] such that FA(t)=0F_{A}(t)=0 for all t[0,t)t\in[0,t^{\ast}) and being continuous on [t,1][t^{\ast},1].
According to inequality (3.3) and Remark 3.1, for proving MK-TP2, it thus remains to consider only those rectangles 0<u1u2<10<u_{1}\leq u_{2}<1 and 0<v1v2<10<v_{1}\leq v_{2}<1 for which h(u2,v1)th(u_{2},v_{1})\geq t^{\ast} (since otherwise KA(u2,[0,v1])=0K_{A}(u_{2},[0,v_{1}])=0); setting f:=logFAf:=\log\circ F_{A} the copula CC hence is MK-TP2 if and only if

1(C(u1,v1)C(u2,v2)C(u1,v2)C(u2,v1))=:I1(FA(h(u1,v1))FA(h(u2,v2))FA(h(u1,v2))FA(h(u2,v1)))=:I21\leq\underbrace{\left(\frac{C(u_{1},v_{1})\,C(u_{2},v_{2})}{C(u_{1},v_{2})\,C(u_{2},v_{1})}\right)}_{=:I_{1}}\cdot\underbrace{\left(\frac{F_{A}(h(u_{1},v_{1}))\,F_{A}(h(u_{2},v_{2}))}{F_{A}(h(u_{1},v_{2}))\,F_{A}(h(u_{2},v_{1}))}\right)}_{=:I_{2}} (5.2)

for all those rectangles 0<u1u2<10<u_{1}\leq u_{2}<1 and 0<v1v2<10<v_{1}\leq v_{2}<1 for which h(u2,v1)th(u_{2},v_{1})\geq t^{\ast}. Since every EVC is TP2, we obvioulsy have I11I_{1}\geq 1, and I21I_{2}\geq 1 if and only if FAhF_{A}\circ h is TP2 (or equivalently, logFAh\log\circ F_{A}\circ h is 22-increasing) on h1([t,1])h^{-1}\big{(}[t^{\ast},1]\big{)}.

At this point we summarize our findings obtained so far:

Theorem 5.6.

Let CC be an EVC with Pickands dependence function AA.

  1. 1.

    If D+A(0)=0D^{+}A(0)=0, then CC is MK-TP2.

  2. 2.

    If D+A(0)(1,0)D^{+}A(0)\in(-1,0), then CC is not MK-TP2.

  3. 3.

    If D+A(0)=1D^{+}A(0)=-1 and

    1. (i)

      D+AD^{+}A has at least two discontinuity points, then CC is not MK-TP2.

    2. (ii)

      D+AD^{+}A has exactly one discontinuity point t(0,1)t^{\ast}\in(0,1) such that D+A(t)>1D^{+}A(t)>-1 for some t(0,t)t\in(0,t^{\ast}), then CC is not MK-TP2.

    3. (iii)

      there exists some t[0,1/2]t^{\ast}\in[0,1/2] such that D+A(t)=1D^{+}A(t)=-1 for all t[0,t)t\in[0,t^{\ast}), D+AD^{+}A is continuous on [t,1][t^{\ast},1] and FAhF_{A}\circ h is TP2 on h1([t,1])h^{-1}\big{(}[t^{\ast},1]\big{)}, then CC is MK-TP2.

For the remainder of this section we focus on part 3.(iii) in Theorem 5.6 and examine conditions under which the function FAhF_{A}\circ h is TP2. We start with the following lemma that relates the TP2 property of FAhF_{A}\circ h to log-concavity of FAF_{A}.

Lemma 5.7.

Suppose that D+A(0)=1D^{+}A(0)=-1, there exists some t[0,1/2]t^{\ast}\in[0,1/2] such that D+A(t)=1D^{+}A(t)=-1 for all t[0,t)t\in[0,t^{\ast}) and D+AD^{+}A is continuous on [t,1][t^{\ast},1].

  1. 1.

    If FAhF_{A}\circ h is TP2 on h1([t,1/2])h^{-1}\big{(}\big{[}t^{\ast},1/2\big{]}\big{)}, then FAF_{A} is log-concave on [t,1/2]\big{[}t^{\ast},1/2\big{]}.

  2. 2.

    If FAF_{A} is log-concave on [1/2,1)\big{[}1/2,1\big{)}, then FAhF_{A}\circ h is TP2 on h1([1/2,1))h^{-1}\big{(}\big{[}1/2,1\big{)}\big{)}.

Proof.

We first assume that FAhF_{A}\circ h is TP2, or equivalently logFAh\log\circ F_{A}\circ h is 22-increasing, on h1([t,1/2])h^{-1}\big{(}\big{[}t^{\ast},1/2\big{]}\big{)}. Since FAF_{A} is assumed to be continuous on [t,1/2][t^{\ast},1/2] it is sufficient to show that logFA\log\circ F_{A} is midpoint concave. To this end, for every ts<t1/2t^{\ast}\leq s<t\leq 1/2 we will construct a rectangle [u1,v1]×[u2,v2]h1([t,1/2])[u_{1},v_{1}]\times[u_{2},v_{2}]\subseteq h^{-1}\big{(}\big{[}t^{\ast},1/2\big{]}\big{)} with h(u1,v2)=t,h(u2,v1)=sh(u_{1},v_{2})=t,h(u_{2},v_{1})=s and max{h(u1,v1),h(u2,v2)}r:=(s+t)/2\max\{h(u_{1},v_{1}),h(u_{2},v_{2})\}\leq r:=(s+t)/2 and then use the 22-increasingness of logFAh\log\circ F_{A}\circ h to show log(FA(s))+log(FA(t))2log(FA(r))\log(F_{A}(s))+\log(F_{A}(t))\leq 2\log(F_{A}(r)) (see Figure 5 for an illustration).

1111u1u_{1}u2u_{2}v1v_{1}v2v_{2}
Figure 5: Blue lines are the contour lines of fs,frf_{s},f_{r} and ftf_{t} where s<r<t<1/2s<r<t<1/2.

Fix u1(0,1)u_{1}\in(0,1) and set v2:=ft(u1)v_{2}:=f_{t}(u_{1}). Since p(1p)/pp\mapsto(1-p)/p is positive, strictly decreasing and strictly log-convex on (0,1/2](0,1/2] we have

1ss1tt>(1rr)2 and hence 1>1ttr1r>1rrs1s.\tfrac{1-s}{s}\,\tfrac{1-t}{t}>\left(\tfrac{1-r}{r}\right)^{2}\text{ and hence }1>\tfrac{1-t}{t}\tfrac{r}{1-r}>\tfrac{1-r}{r}\tfrac{s}{1-s}.

We can thus choose u2(u1,1)u_{2}\in(u_{1},1) such that u11ttr1r<u2<u11rrs1su_{1}^{\frac{1-t}{t}\frac{r}{1-r}}<u_{2}<u_{1}^{\frac{1-r}{r}\frac{s}{1-s}} and set v1:=fs(u2)v_{1}:=f_{s}(u_{2}). Then, h(u1,v2)=th(u_{1},v_{2})=t, h(u2,v1)=sh(u_{2},v_{1})=s and monotonicity of hh yields

h(u1,v1)h(u1,u11rr)=r\displaystyle h(u_{1},v_{1})\leq h\big{(}u_{1},u_{1}^{\frac{1-r}{r}}\big{)}=r

and

h(u2,v2)h(u11ttr1r,u11tt)=1ttr1r1ttr1r+1tt=r\displaystyle h(u_{2},v_{2})\leq h\left(u_{1}^{\frac{1-t}{t}\frac{r}{1-r}},u_{1}^{\frac{1-t}{t}}\right)=\frac{\frac{1-t}{t}\frac{r}{1-r}}{\frac{1-t}{t}\frac{r}{1-r}+\frac{1-t}{t}}=r

Applying 22-increasingness of logFAh\log\circ F_{A}\circ h and monotonicity of logFA\log\circ F_{A} we finally obtain

(logFA)(s)+(logFA)(t)\displaystyle(\log\circ F_{A})(s)+(\log\circ F_{A})(t) =\displaystyle= (logFAh)(u2,v1)+(logFAh)(u1,v2)\displaystyle(\log\circ F_{A}\circ h)(u_{2},v_{1})+(\log\circ F_{A}\circ h)(u_{1},v_{2})
\displaystyle\leq (logFAh)(u1,v1)+(logFAh)(u2,v2)\displaystyle(\log\circ F_{A}\circ h)(u_{1},v_{1})+(\log\circ F_{A}\circ h)(u_{2},v_{2})
\displaystyle\leq 2log(FA(r))\displaystyle 2\log(F_{A}(r))

This proves the first assertion. The second assertion follows from Lemma 6.2 and the fact that hh is 22-increasing on h1([1/2,1))h^{-1}\big{(}\big{[}1/2,1\big{)}\big{)}. ∎


For those Pickands dependence functions AA that are three times differentiable, log-concavity of FAF_{A} (i.e., FA(t)FA′′(t)FA(t)20F_{A}(t)F_{A}^{\prime\prime}(t)-F_{A}^{\prime}(t)^{2}\leq 0) is even more visible and Lemma 5.7 follows immediately:

Theorem 5.8.

Suppose that D+A(0)=1D^{+}A(0)=-1, there exists some t[0,1/2]t^{\ast}\in[0,1/2] such that D+A(t)=1D^{+}A(t)=-1 for all t[0,t)t\in[0,t^{\ast}) and AA is three times differentiable on (t,1)(t^{\ast},1). Then the following statements are equivalent:

  1. (a)

    FAhF_{A}\circ h is TP2 on h1([t,1))h^{-1}\big{(}[t^{\ast},1)\big{)}.

  2. (b)

    The inequality

    FA(t)FA′′(t)FA(t)2FA2(t)t(1t)+FA(t)FA(t)(12t)0\frac{F_{A}(t)F_{A}^{\prime\prime}(t)-F_{A}^{\prime}(t)^{2}}{F_{A}^{2}(t)}\;t(1-t)+\frac{F_{A}^{\prime}(t)}{F_{A}(t)}\;(1-2t)\leq 0 (5.3)

    holds for all t(t,1)t\in(t^{\ast},1).

Proof.

Notice that (a) is equivalent to logFAh\log\circ F_{A}\circ h being 2-increasing on h1([t,1))h^{-1}([t^{\ast},1)), i.e.,

0\displaystyle 0 2uvlog(FA(h(u,v)))()\displaystyle\leq\frac{\partial^{2}}{\partial u\partial v}\log(F_{A}(h(u,v)))\hskip 227.62204pt(\ast)
=\displaystyle= FA(h(u,v))FA′′(h(u,v))FA(h(u,v))2FA(h(u,v))21h(u,v)2h(u,v)\displaystyle\frac{F_{A}(h(u,v))F_{A}^{\prime\prime}(h(u,v))-F_{A}^{\prime}(h(u,v))^{2}}{F_{A}(h(u,v))^{2}}\partial_{1}h(u,v)\,\partial_{2}h(u,v)
+FA(h(u,v))FA(h(u,v))12h(u,v)\displaystyle+\frac{F_{A}^{\prime}(h(u,v))}{F_{A}(h(u,v))}\partial_{12}h(u,v)

for all (u,v)h1([t,1))(u,v)\in h^{-1}([t^{\ast},1)), where

  • 1h(u,v):=uh(u,v)=log(v)u(log(uv))2<0\partial_{1}h(u,v):=\frac{\partial}{\partial u}h(u,v)=\frac{\log(v)}{u(\log(uv))^{2}}<0

  • 2h(u,v):=vh(u,v)=log(u)v(log(uv))2>0\partial_{2}h(u,v):=\frac{\partial}{\partial v}h(u,v)=\frac{-\log(u)}{v(\log(uv))^{2}}>0

  • 12h(u,v):=2uvh(u,v)=log(u)log(v)uv(log(uv))3\partial_{12}h(u,v):=\frac{\partial^{2}}{\partial u\partial v}h(u,v)=\frac{\log(u)-\log(v)}{uv(\log(uv))^{3}}

Since

12h(u,v)\displaystyle\partial_{12}h(u,v) =\displaystyle= 1h(u,v)2h(u,v)12h(u,v)h(u,v)(1h(u,v))\displaystyle\partial_{1}h(u,v)\,\partial_{2}h(u,v)\,\frac{1-2h(u,v)}{h(u,v)(1-h(u,v))}

we have

0\displaystyle 0 \displaystyle\leq 2uvlog(FA(h(u,v)))\displaystyle\frac{\partial^{2}}{\partial u\partial v}\log(F_{A}(h(u,v)))
=\displaystyle= (FA(h(u,v))FA′′(h(u,v))FA(h(u,v))2FA(h(u,v))2+FA(h(u,v))FA(h(u,v))12h(u,v)h(u,v)(1h(u,v)))\displaystyle\biggl{(}\frac{F_{A}(h(u,v))F_{A}^{\prime\prime}(h(u,v))-F_{A}^{\prime}(h(u,v))^{2}}{F_{A}(h(u,v))^{2}}+\frac{F_{A}^{\prime}(h(u,v))}{F_{A}(h(u,v))}\frac{1-2h(u,v)}{h(u,v)(1-h(u,v))}\biggr{)}
1h(u,v)2h(u,v)\displaystyle\cdot\,\partial_{1}h(u,v)\,\partial_{2}h(u,v)

Therefore, ()(\ast) holds if and only if

0FA(t)FA′′(t)FA(t)2FA2(t)t(1t)+FA(t)FA(t)(12t)0\geq\frac{F_{A}(t)F_{A}^{\prime\prime}(t)-F_{A}^{\prime}(t)^{2}}{F_{A}^{2}(t)}\;t(1-t)+\frac{F_{A}^{\prime}(t)}{F_{A}(t)}\;(1-2t)

for all t(t,1)t\in(t^{\ast},1). This proves the assertion. ∎

Remark 5.9.

It is worth mentioning that Inequality (5.3) is equivalent to the non-increasingness of

tt(1t)FA(t)FA(t)t\mapsto t\,(1-t)\,\frac{F_{A}^{\prime}(t)}{F_{A}(t)} (5.4)

on (t,1)(t^{\ast},1) which turns out to be neither a sufficient nor a necessary condition for log-concavity of FAF_{A} (i.e., tFA(t)/FA(t)t\mapsto F_{A}^{\prime}(t)/F_{A}(t) is non-increasing) on (t,1)(t^{\ast},1).

Lemma 5.7 and Theorem 5.8 suggest that log-concavity of FAF_{A} could be a sufficient condition for MK-TP2. The next example contradicts this conjecture.

Example 5.10.

The mapping A:𝕀𝕀A:{\mathbb{I}}\to{\mathbb{I}} given by

A(t):=log(t3/2+(1t)3/2)2/3+1A(t):=\log\left(t^{3/2}+(1-t)^{3/2}\right)^{2/3}+1

is a Pickands dependence function fulfilling D+A(0)=1D^{+}A(0)=-1 and being (at least) three times differentiable on (0,1)(0,1) with

FA(t)\displaystyle F_{A}(t) =\displaystyle= 23log(t3/2+(1t)3/2)+t1/2t3/2+(1t)3/2\displaystyle\frac{2}{3}\log\left(t^{3/2}+(1-t)^{3/2}\right)+\frac{t^{1/2}}{t^{3/2}+(1-t)^{3/2}}
FA(t)\displaystyle F_{A}^{\prime}(t) =\displaystyle= 1t(t3/2+(1t)3/2)21+4t(1t)2(1t)t2t(1t)\displaystyle\frac{1-t}{(t^{3/2}+(1-t)^{3/2})^{2}}\,\frac{1+4t(1-t)-2\sqrt{(1-t)t}}{2\sqrt{t(1-t)}}

Figure 6 depicts a sample of the corresponding EVC.

Refer to caption
Figure 6: Sample of size n=10.000n=10.000 drawn from the EVC discussed in Example 5.10 with Pickands dependence function A(t):=log(t3/2+(1t)3/2)2/3+1A(t):=\log\left(t^{3/2}+(1-t)^{3/2}\right)^{2/3}+1.

Tedious calculation yields that FAF_{A} is log-concave. However, for t1=0.1t_{1}=0.1 and t2=0.2t_{2}=0.2

FA(t1)FA(t1)t1(1t1)=0.474<0.505=FA(t2)FA(t2)t2(1t2)\frac{F_{A}^{\prime}(t_{1})}{F_{A}(t_{1})}t_{1}(1-t_{1})=0.474<0.505=\frac{F_{A}^{\prime}(t_{2})}{F_{A}(t_{2})}t_{2}(1-t_{2})

which contradicts (5.4), hence FAhF_{A}\circ h fails to be TP2 on (0,1)2(0,1)^{2}. It even holds that the corresponding EVC is not MK-TP2 which follows considering the rectangle [0.9,0.95]×[0.5,0.6][0.9,0.95]\times[0.5,0.6].

In what follows, we examine for selected classes of EVCs for which parameters the MK-TP2 property is fulfilled and for which it is not.

Example 5.11.

(Gumbel copula)
For α[1,)\alpha\in[1,\infty), the mapping Aα:𝕀𝕀A_{\alpha}:{\mathbb{I}}\to{\mathbb{I}} given by

Aα(t):=(tα+(1t)α)1/αA_{\alpha}(t):=(t^{\alpha}+(1-t)^{\alpha})^{1/\alpha}

is a Pickands dependence function; the corresponding EVC CαC_{\alpha} is called Gumbel copula (see, e.g., [1]). It immedately follows from Example 4.8 that CαC_{\alpha} is MK-TP2 by taking into account its Archimedean structure.

We now consider the asymmetric mixed model by Tawn (see, e.g., [31, 30]) which generalizes Example 5.3:

Example 5.12.

(Asymmetric mixed model by Tawn)
For θ\theta and κ\kappa satisfying θ0\theta\geq 0, θ+3κ0\theta+3\kappa\geq 0, θ+κ1\theta+\kappa\leq 1 and θ+2κ1\theta+2\kappa\leq 1, the mapping A:𝕀𝕀A:{\mathbb{I}}\to{\mathbb{I}} given by

A(t):=1(θ+κ)t+θt2+κt3A(t):=1-(\theta+\kappa)t+\theta t^{2}+\kappa t^{3}

is a Pickands dependence function; the corresponding EVC Cθ,κC_{\theta,\kappa} occurs in Tawn’s model [30]. Note that the parameter conditions imply 0θ+κ10\leq\theta+\kappa\leq 1. The following statements are equivalent

  1. (a)

    Cθ,κC_{\theta,\kappa} is MK-TP2.

  2. (b)

    θ+κ{0,1}\theta+\kappa\in\{0,1\}.

Indeed, considering A(t)=(θ+κ)+2θt+3κt2A^{\prime}(t)=-(\theta+\kappa)+2\theta t+3\kappa t^{2} and hence A(0)=(θ+κ)A^{\prime}(0)=-(\theta+\kappa) cases θ+κ=0\theta+\kappa=0 and θ+κ(0,1)\theta+\kappa\in(0,1) immediately follow from Theorem 5.6. Now assume that θ+κ=1\theta+\kappa=1. Then A(t)=1t+θt2+(1θ)t3A(t)=1-t+\theta t^{2}+(1-\theta)t^{3} with θ[1,3/2]\theta\in[1,3/2]. Simple calculation yields

  • FA(t)=2θt(1t)2+(32t)t2F_{A}(t)=2\theta t(1-t)^{2}+(3-2t)t^{2}

  • FA(t)=2(1t)(θ+3(1θ)t)F_{A}^{\prime}(t)=2(1-t)(\theta+3(1-\theta)t)

  • FA′′(t)=2θ+6(1θ)(1t)6(1θ)tF_{A}^{\prime\prime}(t)=-2\theta+6(1-\theta)(1-t)-6(1-\theta)t

In order to show inequality (5.3) we calculate

(FA(t)FA′′(t)FA(t)2)t(1t)+FA(t)FA(t)(12t)\displaystyle\big{(}F_{A}(t)F_{A}^{\prime\prime}(t)-F_{A}^{\prime}(t)^{2}\big{)}\;t(1-t)+F_{A}(t)\,F_{A}^{\prime}(t)\;(1-2t)
=\displaystyle= 2(1t)t2(6θ2(1t)36(2t)t2+θ(317t+30t212t3))\displaystyle 2(1-t)t^{2}\,\big{(}-6\theta^{2}(1-t)^{3}-6(2-t)t^{2}+\theta(3-17t+30t^{2}-12t^{3})\big{)}

The latter term is negative if and only if

θ(317t+30t212t3)>06θ2(1t)3+6(2t)t2\theta\underbrace{(3-17t+30t^{2}-12t^{3})}_{>0}\geq 6\theta^{2}(1-t)^{3}+6(2-t)t^{2}

for all t(0,1)t\in(0,1). Bounding θ\theta on the left-hand side from below by 11 and on the right-hand side from above by 3/23/2 yields a sufficient condition, namely

(317t+30t212t3)272(1t)3+6(2t)t2(3-17t+30t^{2}-12t^{3})\geq\tfrac{27}{2}(1-t)^{3}+6(2-t)t^{2}

which is true for all t(0,1)t\in(0,1). Therefore, inequality (5.3) holds implying that Cθ,κC_{\theta,\kappa} is MK-TP2.

We now resume Example 3.6 and show that a Marshall-Olkin copula Mα,βM_{\alpha,\beta} with α,β(0,1]\alpha,\beta\in(0,1] is MK-TP2 if and only if β=1\beta=1:

Example 5.13.

(Marshall-Olkin copula)
For α,β(0,1]\alpha,\beta\in(0,1], the mapping A:𝕀𝕀A:{\mathbb{I}}\to{\mathbb{I}} given by

A(t)={1βtt<αα+β1α(1t)tαα+βA(t)=\begin{cases}1-\beta t&t<\frac{\alpha}{\alpha+\beta}\\ 1-\alpha(1-t)&t\geq\frac{\alpha}{\alpha+\beta}\end{cases}

is a Pickands dependence function; the corresponding EVC Mα,βM_{\alpha,\beta} is called Marshall-Olkin copula (see, e.g., [1]). The following statements are equivalent

  1. (a)

    Mα,βM_{\alpha,\beta} is MK-TP2.

  2. (b)

    β=1\beta=1.

Indeed, we first have D+A(0)=βD^{+}A(0)=-\beta. Now, if β(0,1)\beta\in(0,1) then D+A(0)(1,0)D^{+}A(0)\in(-1,0) and Theorem 5.6 implies that Mα,βM_{\alpha,\beta} is not MK-TP2. If, otherwise, β=1\beta=1 then

FA(t)={0t<αα+11tαα+1=𝟙[αα+1,1](t)F_{A}(t)=\begin{cases}0&t<\frac{\alpha}{\alpha+1}\\ 1&t\geq\frac{\alpha}{\alpha+1}\end{cases}=\mathds{1}_{[\frac{\alpha}{\alpha+1},1]}(t)

implying Inequality (5.3) to hold for all t(α/(α+1),1)t\in\big{(}\alpha/(\alpha+1),1\big{)}.

We conclude this section by drawing the reader’s attention to another quite simple condition under which a given EVC fails to be MK-TP2; the proof is deferred to the Appendix.

Lemma 5.14.

Suppose that D+A(0)=1D^{+}A(0)=-1 and that there exist t1,t2(0,1)t_{1},t_{2}\in(0,1) with t1<t2t_{1}<t_{2} and some c(0,1)c\in(0,1) such that FA(t)=cF_{A}(t)=c on [t1,t2][t_{1},t_{2}]. Then the corresponding EVC is not MK-TP2.

In other words, if the function FAF_{A} is constant on an interval, then the corresponding EVC is not MK-TP2. We illustrate this result with an example:

Example 5.15.

The mapping A:𝕀𝕀A:{\mathbb{I}}\to{\mathbb{I}} given by

A(t):={1tt[0,18)151612tt[18,14)1216+(x12)2t[14,1]A(t):=\begin{cases}1-t&t\in\big{[}0,\tfrac{1}{8}\big{)}\\ \tfrac{15}{16}-\tfrac{1}{2}t&t\in\big{[}\tfrac{1}{8},\tfrac{1}{4}\big{)}\\ \tfrac{12}{16}+\big{(}x-\tfrac{1}{2}\big{)}^{2}&t\in\big{[}\tfrac{1}{4},1\big{]}\end{cases}

is a Pickands dependence function fulfilling D+A(t)=1D^{+}A(t)=-1 for all t[0,1/8)t\in\big{[}0,1/8\big{)} and

FA(t)={0t[0,18)716t[18,14)t(2t)t[14,1]F_{A}(t)=\begin{cases}0&t\in\big{[}0,\tfrac{1}{8}\big{)}\\ \tfrac{7}{16}&t\in\big{[}\tfrac{1}{8},\tfrac{1}{4}\big{)}\\ t(2-t)&t\in\big{[}\tfrac{1}{4},1\big{]}\end{cases}

(see Figure 7 for an illustration). It is straightforward to verify that FAhF_{A}\circ h fails to be TP2 on [1/8,1]\big{[}1/8,1\big{]} and it follows from Lemma 5.14 that the corresponding EVC is not MK-TP2.

1111
(a) Plot of Pickands dependence function AA.
1111
(b) Plot of function FAF_{A}.
Figure 7: Corresponding functions AA and FAF_{A} for Example 5.15.

Acknowledgement

The first author gratefully acknowledges the support of the WISS 2025 project ’IDA-lab Salzburg’ (20204-WISS/225/197-2019 and 20102-F1901166-KZP). The second author gratefully acknowledges the financial support from AMAG Austria Metall AG within the project ProSa.

References

  • [1] F. Durante, C. Sempi, Principles of copula theory, Taylor & Francis Group LLC, Boca Raton FL, 2016.
  • [2] R. Nelsen, An Introduction to Copulas, Springer-Verlag, Berlin Heidelberg, 2006.
  • [3] H. Joe, Dependence Modeling with Copulas, CRC press, 2014.
  • [4] S. Chatterjee, A new coefficient of correlation., J. Amer. Statist. Ass. 116 (536) (2021) 2009–2022.
  • [5] H. Dette, K. F. Siburg, P. A. Stoimenov, A copula-based non-parametric measure of regression dependence., Scand. J. Statist. 40 (1) (2013) 21–41.
  • [6] R. Junker, F. Griessenberger, W. Trutschnig, Estimating scale-invariant directed dependence of bivariate distributions., Comput. Statist. Data Anal. 153 (2020) Article ID 107058, 22 pages.
  • [7] P. Capéraà, C. Genest, Spearman’s ρ\rho is larger than kendall’s τ\tau for positively dependent random variables, J. Nonparametr. Stat. 2 (2) (1993) 183–194.
  • [8] G. A. Fredricks, R. B. Nelsen, On the relationship between Spearman’s rho and Kendall’s tau for pairs of continuous random variables., J. Statist. Plann. Inference 137 (2007) 2143–2150.
  • [9] K. F. Siburg, C. Strothmann, Stochastic monotonicity and the markov product for copulas, J. Math. Anal. Appl. 503 (2) (2021) 125348.
  • [10] T. Kasper, S. Fuchs, W. Trutschnig, On weak conditional convergence of bivariate Archimedean and Extreme Value copulas, and consequences to nonparametric estimation., Bernoulli 27 (2021) 2217–2240.
  • [11] T. Mroz, S. Fuchs, W. Trutschnig, How simplifying and flexible is the simplifying assumption in pair-copula constructions – analytic answers in dimension three and a glimpse beyond., Electron. J. Statist. 15 (1) (2021) 1951–1992.
  • [12] W. Hürlimann, Hutchinson-lai’s conjecture for bivariate extreme value copulas., Statist. Probab. Lett. 61 (2003) 191–198.
  • [13] H. Joe, Multivariate Models and Multivariate Dependence Concepts, CRC Press, 1997.
  • [14] E. Lehmann, Some concepts of dependence, Ann. Math. Statist. 37 (1966) 1137–1153.
  • [15] P. Capéraà, C. Genest, Concepts de dépendance et ordres stochastiques pour des lois bidimensionnelles, Canad. J. Statist. 18 (4) (1990) 315–326.
  • [16] A. I. G. Guillem, Structure de dépendance des lois de valeurs extrêmes bivariées, Comptes Rendus de l’Académie des Sciences-Series I-Mathematics 330 (7) (2000) 593–596.
  • [17] A. Müller, M. Scarsini, Archimedean copulae and positive dependence, J. Multivariate Anal. 93 (2005) 434–445.
  • [18] O. Kallenberg, Foundations of Modern Probability, Springer-Verlag, New York, 2002.
  • [19] A. Klenke, Wahrscheinlichkeitstheorie, Springer Lehrbuch Masterclass Series, Berlin Heidelberg, 2008.
  • [20] J. Fernández Sánchez, W. Trutschnig, Singularity aspects of archimedean copulas, J. Math. Anal. Appl. 432 (1) (2015) 103–113.
  • [21] A. Marshall, I. Olkin, B. Arnold, Inequalities: Theory of Majorization and Its Applications, 2nd Edition, Springer, New York, 2011.
  • [22] J. Spreeuw, Archimedean copulas derived from morgenstern utility functions, Available at SSRN 2215701 (2013).
  • [23] C. Niculescu, L.-E. Persson, Convex Functions and Their Applications, Springer, 2006.
  • [24] L. de Haan, S. Resnick, Limit theory for multivariate sample extremes 40 (1977) 317–337.
  • [25] G. Gudendorf, J. Segers, Nonparametric estimation of an extreme-value copula in arbitrary dimensions, J. Multivariate Anal. 102 (1) (2011) 37 – 47.
  • [26] J. Pickands, Multivariate extreme value distributions, Proceedings 43rd Session International Statistical Institute 2 (1981) 859–878.
  • [27] W. Trutschnig, M. Schreyer, J. Fernández-Sánchez, Mass distributions of two-dimensional extreme-value copulas and related results, Extremes 19 (3) (2016) 405–427.
  • [28] P. Capéraà, A.-L. Fougéres, C. Genest, A nonparametric estimation procedure for bivariate extreme value copulas, Biometrika 84 (1997) 567–577.
  • [29] K. Ghoudi, A. Khoudraji, L.-P. Rivest, Propriétés statistiques des copules de valeurs extremes bidimensionnelles, Canad. J. Statist. 26 (1998) 187–197.
  • [30] J. Tawn, Bivariate extreme value theory: models and estimation, Biometrika 75 (1988) 397–415.
  • [31] C. Genest, J. Segers, Rank-based inference for bivariate extreme-value copulas., Ann. Stat. 37 (5B) (2009) 2990–3022.
  • [32] I. J. Schoenberg, On Pólya frequency functions, i. the totally positive functions and their laplace transforms., J. Analyse Math. 1 (1951) 331–374.
  • [33] W. Walter, Gewöhnliche Differentialgleichungen: Eine Einführung, Springer-Verlag, 2013.

6 Supplementary material

In this appendix we gather some helpful results concerning the TP2 property and extreme value copulas.

By definition, a copula CC that is TP2 (or SI) satisfies C(u,v)Π(u,v)>0C(u,v)\geq\Pi(u,v)>0 for all (u,v)(0,1)2(u,v)\in(0,1)^{2}. The next result characterizes the TP2 property in terms of a monotonicity property:

Lemma 6.1.

Consider some copula CC satisfying C(u,v)>0C(u,v)>0 on (0,1)2(0,1)^{2}. Then the following statements are equivalent:

  1. (a)

    CC is TP2.

  2. (b)

    logC\log\circ C is 22-increasing.

  3. (c)

    For any u(0,1)u\in(0,1), the mapping (0,1)(0,1)\to\mathbb{R} given by

    vKC(u,[0,v])C(u,v)v\mapsto\frac{K_{C}(u,[0,v])}{C(u,v)}

    is non-decreasing.

Proof.

The equivalence of (a) and (b) is straightforward. Now, fix v(0,1)v\in(0,1) and define Fv:𝕀F_{v}:{\mathbb{I}}\to\mathbb{R} by letting Fv(u):=C(u,v)F_{v}(u):=C(u,v). Then FvF_{v} is a continuous distribution function with λ\lambda-density satisfying

Fv(u)=(0,u]KC(s,[0,v])dλ(s)F_{v}(u)=\int_{(0,u]}K_{C}(s,[0,v])\;\mathrm{d}\lambda(s)

Thus, for every u(0,1)u\in(0,1)

(logC)(u,v)\displaystyle(\log\circ C)(u,v) =\displaystyle= (C(u,v),1]1sdλ(s)\displaystyle-\,\int_{(C(u,v),1]}\frac{1}{s}\;\mathrm{d}\lambda(s)
=\displaystyle= (0,1)𝟙(u,1](Fv(s))1(FvFv)(s)dλ(s)\displaystyle\;\;-\,\int_{(0,1)}\mathds{1}_{(u,1]}(F_{v}^{\leftarrow}(s))\frac{1}{(F_{v}\circ F_{v}^{\leftarrow})(s)}\;\mathrm{d}\lambda(s)
=\displaystyle= (u,1]1C(s,v)dλFv(s)=(u,1]KC(s,[0,v])C(s,v)dλ(s)\displaystyle-\,\int_{(u,1]}\frac{1}{C(s,v)}\;\mathrm{d}\lambda^{F_{v}^{\leftarrow}}(s)\;\;=\;\;-\,\int_{(u,1]}\frac{K_{C}(s,[0,v])}{C(s,v)}\;\mathrm{d}\lambda(s)

Therefore, logC\log\circ C is 22-increasing if and only if, for any u(0,1)u\in(0,1)

vKC(u,[0,v])C(u,v)v\mapsto\frac{K_{C}(u,[0,v])}{C(u,v)}

is non-decreasing. This proves the result. ∎

It is well-known that if ϕ\phi is a 22-increasing and monotone (both coordinates in the same direction) function and ff is convex and non-decreasing, then the composition fϕf\circ\phi is 22-increasing (see [21, p.219]). In the next lemma we modify the assumption on ff for the case ϕ\phi is monotone in opposite direction. Lemma 6.2 applies to Lemma 5.7.

Lemma 6.2.

Suppose that ϕ:Ω2Ω1\phi:\Omega_{2}\to\Omega_{1}, Ω22\Omega_{2}\subseteq\mathbb{R}^{2}, Ω1\Omega_{1}\subseteq\mathbb{R}, is 22-increasing, xϕ(x,y)x\mapsto\phi(x,y) is non-increasing, yϕ(x,y)y\mapsto\phi(x,y) is non-decreasing, and f:Ω1f:\Omega_{1}\to\mathbb{R} is concave and non-decreasing. Then fϕf\circ\phi is 22-increasing.

Proof.

For x1x2x_{1}\leq x_{2} and y1y2y_{1}\leq y_{2} such that [x1,x2]×[y1,y2]Ω2[x_{1},x_{2}]\times[y_{1},y_{2}]\subseteq\Omega_{2} monotonicity of ϕ\phi implies

ϕ(x2,y1)min{ϕ(x1,y1),ϕ(x2,y2)}max{ϕ(x1,y1),ϕ(x2,y2)}ϕ(x1,y2).\displaystyle\phi(x_{2},y_{1})\leq\min\{\phi(x_{1},y_{1}),\phi(x_{2},y_{2})\}\leq\max\{\phi(x_{1},y_{1}),\phi(x_{2},y_{2})\}\leq\phi(x_{1},y_{2}).

and 22-increasingness yields ϕ(x1,y2)+ϕ(x2,y1)ϕ(x1,y1)+ϕ(x2,y2)\phi(x_{1},y_{2})+\phi(x_{2},y_{1})\leq\phi(x_{1},y_{1})+\phi(x_{2},y_{2}).
First, assume ϕ(x1,y1)<ϕ(x2,y2)\phi(x_{1},y_{1})<\phi(x_{2},y_{2}), choose s2s_{2} and t1<t2t_{1}<t_{2} such that s2t1=ϕ(x1,y2)s_{2}-t_{1}=\phi(x_{1},y_{2}) and s2t2=ϕ(x2,y2)s_{2}-t_{2}=\phi(x_{2},y_{2}), and set s1:=ϕ(x2,y1)+t2<s2s_{1}:=\phi(x_{2},y_{1})+t_{2}<s_{2}. Then (see Figure 8 for an illustration)

s1t1\displaystyle s_{1}-t_{1} =ϕ(x2,y1)+t2t1+ϕ(x1,y2)s2+t1\displaystyle=\phi(x_{2},y_{1})+t_{2}-t_{1}+\phi(x_{1},y_{2})-s_{2}+t_{1}
=ϕ(x2,y1)+ϕ(x1,y2)+t2s2\displaystyle=\phi(x_{2},y_{1})+\phi(x_{1},y_{2})+t_{2}-s_{2}
ϕ(x1,y1)+ϕ(x2,y2)+t2s2\displaystyle\leq\phi(x_{1},y_{1})+\phi(x_{2},y_{2})+t_{2}-s_{2}
=ϕ(x1,y1)\displaystyle=\phi(x_{1},y_{1})
ϕ(x1,y2)\phi(x_{1},y_{2})ϕ(x2,y1)\phi(x_{2},y_{1})ϕ(x1,y1)\phi(x_{1},y_{1})ϕ(x2,y2)\phi(x_{2},y_{2})s2t1s_{2}-t_{1}s2t2s_{2}-t_{2}s1t1s_{1}-t_{1}s1t2s_{1}-t_{2}
Figure 8: Illustration how the specific points are chosen and why s1t1ϕ(x1,y1)s_{1}-t_{1}\leq\phi(x_{1},y_{1}) holds in the case ϕ(x1,y1)<ϕ(x2,y2)\phi(x_{1},y_{1})<\phi(x_{2},y_{2}).

By assumption expf\exp\circ f is (non-negative), log-concave and non-decreasing. Analogous to [32] (see also [21, Example A.10]) log-concavity implies that the function (s,t)(expf)(st)(s,t)\mapsto(\exp\circ f)(s-t) is TP2 which gives

(expf)(ϕ(x1,y2))(expf)(ϕ(x2,y1))\displaystyle(\exp\circ f)(\phi(x_{1},y_{2}))\cdot(\exp\circ f)(\phi(x_{2},y_{1})) =(expf)(s2t1)(expf)(s1t2)\displaystyle=(\exp\circ f)(s_{2}-t_{1})\cdot(\exp\circ f)(s_{1}-t_{2})
(expf)(s2t2)(expf)(s1t1)\displaystyle\leq(\exp\circ f)(s_{2}-t_{2})\cdot(\exp\circ f)(s_{1}-t_{1})
=(expf)(ϕ(x2,y2))(expf)(s1t1)\displaystyle=(\exp\circ f)(\phi(x_{2},y_{2}))\cdot(\exp\circ f)(s_{1}-t_{1})

and monotonicity of expf\exp\circ f yields

(expf)(ϕ(x1,y2))(expf)(ϕ(x2,y1))(expf)(ϕ(x2,y2))(expf)(ϕ(x1,y1))\displaystyle(\exp\circ f)(\phi(x_{1},y_{2}))\cdot(\exp\circ f)(\phi(x_{2},y_{1}))\leq(\exp\circ f)(\phi(x_{2},y_{2}))\cdot(\exp\circ f)(\phi(x_{1},y_{1}))

This is equivalent to fϕf\circ\phi being 22-increasing. The case ϕ(x1,y1)>ϕ(x2,y2)\phi(x_{1},y_{1})>\phi(x_{2},y_{2}) follows analogously. This proves the desired assertion. ∎

We now present some auxiliary results that are used in Section 5. Lemma 6.3 provides a structure for FAF_{A} that contradicts the MK-TP2 property.

Lemma 6.3.

Suppose that FAF_{A} has discontinuity point tr(0,1)t_{r}\in(0,1) and there exists some tl(0,tr)t_{l}\in(0,t_{r}) such that FA(t)>0F_{A}(t)>0 for all t[tl,tr)t\in[t_{l},t_{r}) and tFA(t)t\mapsto F_{A}(t) is continuous on [tl,tr)[t_{l},t_{r}). Then the corresponding EVC is not MK-TP2.

Proof.

In what follows we construct a rectangle [u1,u2]×[v1,v2][u_{1},u_{2}]\times[v_{1},v_{2}] for which inequality (3.3) is not fulfilled (see Figure 9 for an illustration).
Recall that FAF_{A} is non-decreasing. First of all, discontinuity of FAF_{A} in trt_{r} implies the existence of some δ(0,1)\delta\in(0,1) such that FA(tr)=δ+yF_{A}(t_{r})=\delta+y where y:=supt<trFA(t)>0y:=\sup_{t<t_{r}}F_{A}(t)>0, and choose ε(0,δy/(δ+y))\varepsilon\in\bigl{(}0,\delta y/(\delta+y)\bigr{)}. Since FAF_{A} is continuous on [tl,tr)[t_{l},t_{r}) there exists some t[tl,tr)t^{\ast}\in[t_{l},t_{r}) such that

0<yεFA(t)y for all t[t,tr)\displaystyle 0<y-\varepsilon\leq F_{A}(t)\leq y\qquad\text{ for all }t\in[t^{\ast},t_{r})

where the first inequality follows from the fact that y>δy/(δ+y)y>\delta y/(\delta+y).

1111u1u_{1}u2u_{2}uu^{\ast}v2v_{2}v1v_{1}vv^{\ast}ftlf_{t_{l}}ftrf_{t_{r}}ftf_{t^{\ast}}
Figure 9: Blue lines are the contour lines of ftr,ftf_{t_{r}},f_{t^{\ast}} and ftlf_{t_{l}} where tl<t<trt_{l}<t^{\ast}<t_{r}.

Now, let u1(0,1)u_{1}\in(0,1) and set v2:=ftr(u1)=u11tr1(0,1)v_{2}:=f_{t_{r}}(u_{1})=u_{1}^{\frac{1}{t_{r}}-1}\in(0,1). Then we immediately obtain h(u1,v2)=trh(u_{1},v_{2})=t_{r} and monotonicity of hh implies

h(u,v)<min{h(u,v2),h(u1,v)}max{h(u,v2),h(u1,v)}<h(u1,v2)=trh(u,v)<\min\{h(u,v_{2}),h(u_{1},v)\}\leq\max\{h(u,v_{2}),h(u_{1},v)\}<h(u_{1},v_{2})=t_{r}

for all u(u1,1)u\in(u_{1},1) and all v(0,v2)v\in(0,v_{2}). Since h is continuous and has convex or concave contour lines respectively, there exist u(u1,1)u^{\ast}\in(u_{1},1) as well as v(0,v2)v^{\ast}\in(0,v_{2}) such that h(u,v)th(u,v)\geq t^{\ast} for all (u,v)[u1,u]×[v,v2](u,v)\in[u_{1},u^{\ast}]\times[v^{\ast},v_{2}] and hence

0<yεFA(h(u,v))y for all (u,v)([u1,u]×[v,v2])\{(u1,v2)}.0<y-\varepsilon\leq F_{A}(h(u,v))\leq y\qquad\text{ for all }(u,v)\in([u_{1},u^{\ast}]\times[v^{\ast},v_{2}])\backslash\{(u_{1},v_{2})\}.

We therefore have

supu(u1,u)v(v,v2)FA(h(u1,v))FA(h(u,v2))FA(h(u1,v2))FA(h(u,v))\displaystyle\sup\limits_{\begin{subarray}{c}u\in(u_{1},u^{\ast})\\ v\in(v^{\ast},v_{2})\end{subarray}}\frac{F_{A}(h(u_{1},v))F_{A}(h(u,v_{2}))}{F_{A}(h(u_{1},v_{2}))F_{A}(h(u,v))} y2(δ+y)(yε)\displaystyle\leq\frac{y^{2}}{(\delta+y)(y-\varepsilon)}
<y2(δ+y)(yδyδ+y)=y2y(δ+y)δy=1\displaystyle<\frac{y^{2}}{(\delta+y)(y-\frac{\delta y}{\delta+y})}=\frac{y^{2}}{y(\delta+y)-\delta y}=1

Setting β:=y2/[(y+δ)(yε)]<1\beta:=y^{2}/[(y+\delta)(y-\varepsilon)]<1 we then obtain

supu(u1,u)v(v,v2)FA(h(u1,v))FA(h(u,v2))FA(h(u1,v2))FA(h(u,v))β\sup\limits_{\begin{subarray}{c}u\in(u_{1},u^{\ast})\\ v\in(v^{\ast},v_{2})\end{subarray}}\frac{F_{A}(h(u_{1},v))F_{A}(h(u,v_{2}))}{F_{A}(h(u_{1},v_{2}))F_{A}(h(u,v))}\leq\beta

and continuity of copulas implies the existence of some u2(u1,u)u_{2}\in(u_{1},u^{\ast}) and some v1(v,v2)v_{1}\in(v^{\ast},v_{2}) such that C(u2,v1)>βC(u2,v2)C(u_{2},v_{1})>\beta\cdot C(u_{2},v_{2}).
We are now in the position to show that inequality (3.3) for the chosen rectangle [u1,u2]×[v1,v2][u_{1},u_{2}]\times[v_{1},v_{2}] fails to hold. First note that KC(u2,[0,v1])=C(u2,v1)/u2FA(h(u2,v1))>0K_{C}(u_{2},[0,v_{1}])=C(u_{2},v_{1})/u_{2}\cdot F_{A}(h(u_{2},v_{1}))>0, hence

KC(u1,[0,v1])KC(u2,[0,v2])KC(u1,[0,v2])KC(u2,[0,v1])\displaystyle\frac{K_{C}(u_{1},[0,v_{1}])\,K_{C}(u_{2},[0,v_{2}])}{K_{C}(u_{1},[0,v_{2}])\,K_{C}(u_{2},[0,v_{1}])} =\displaystyle= C(u1,v1)FA(h(u1,v1))C(u2,v2)FA(h(u2,v2))C(u1,v2)FA(h(u1,v2))C(u2,v1)FA(h(u2,v1))\displaystyle\frac{C(u_{1},v_{1})F_{A}(h(u_{1},v_{1}))\,C(u_{2},v_{2})F_{A}(h(u_{2},v_{2}))}{C(u_{1},v_{2})F_{A}(h(u_{1},v_{2}))\,C(u_{2},v_{1})F_{A}(h(u_{2},v_{1}))}
\displaystyle\leq C(u1,v1)C(u1,v2)βC(u2,v2)C(u2,v1)\displaystyle\frac{C(u_{1},v_{1})}{C(u_{1},v_{2})}\,\beta\,\frac{C(u_{2},v_{2})}{C(u_{2},v_{1})}
<\displaystyle< C(u1,v1)C(u1,v2)\displaystyle\frac{C(u_{1},v_{1})}{C(u_{1},v_{2})}
\displaystyle\leq 1\displaystyle 1

This proves the assertion. ∎

Next, consider some Pickands dependence function AA with A1A\neq 1. Since AA is convex it attains its minimum on (0,1)(0,1), i.e. there exists some δ(0,)\delta\in(0,\infty) such that

δ=argminγ(0,)A(11+γ)\delta=\operatorname*{arg\,min}_{\gamma\in(0,\infty)}A\left(\frac{1}{1+\gamma}\right)

Since there may exist several minimizer we set

βA:=sup{δ(0,):δ=argminγ(0,)A(11+γ)}\beta_{A}:=\sup\Big{\{}\delta\in(0,\infty):\delta=\operatorname*{arg\,min}\limits_{\gamma\in(0,\infty)}A\left(\frac{1}{1+\gamma}\right)\Big{\}}

Lemma 6.4 simplifies the proof of Lemma 5.2 in the continuous case.

Lemma 6.4.

Suppose that D+AD^{+}A is continuous with A1A\neq 1 and define the mapping g:[0,βA]g:[0,\beta_{A}]\to\mathbb{R} by

g(α):=(1+α)A(11+α)αg(\alpha):=(1+\alpha)A\left(\frac{1}{1+\alpha}\right)-\alpha

Then

  1. (1)

    gg is strictly decreasing on (0,βA)(0,\beta_{A}).

  2. (2)

    g(0)=1g(0)=1 and g(α)0g(\alpha)\geq 0 for all α(0,βA)\alpha\in(0,\beta_{A}).

  3. (3)

    the inequality

    FA(11+βA)<FA(11+α)F_{A}\left(\frac{1}{1+\beta_{A}}\right)<F_{A}\left(\frac{1}{1+\alpha}\right)

    holds for all α(0,βA)\alpha\in(0,\beta_{A}).

  4. (4)

    the inequality

    log(FA(11+βA)FA(11+α))g(α)g(βA)<0\frac{\log\left(\frac{F_{A}\left(\frac{1}{1+\beta_{A}}\right)}{F_{A}\left(\frac{1}{1+\alpha}\right)}\right)}{g(\alpha)-g(\beta_{A})}<0

    holds for all α(0,βA)\alpha\in(0,\beta_{A}).

Proof.

First of all, the derivative of gg on (0,βA)(0,\beta_{A}) fulfills

g(α)=(A(11+α)1)<0D+A(11+α)1+α0<0g^{\prime}(\alpha)=\underbrace{\left(A\left(\frac{1}{1+\alpha}\right)-1\right)}_{<0}-\underbrace{\frac{D^{+}A\left(\frac{1}{1+\alpha}\right)}{1+\alpha}}_{\geq 0}<0

for all α(0,βA)\alpha\in(0,\beta_{A}) which implies that gg is strictly decreasing on (0,βA)(0,\beta_{A}). This proves (1), and (2) is immediate from the fact that A(t)1tA(t)\geq 1-t for all t𝕀t\in{\mathbb{I}}. Now, we prove (3) and (4). Continuity of D+AD^{+}A implies

FA(11+βA)\displaystyle F_{A}\left(\frac{1}{1+\beta_{A}}\right) =A(11+βA)+(111+βA)D+A(11+βA)\displaystyle=A\left(\frac{1}{1+\beta_{A}}\right)+\left(1-\frac{1}{1+\beta_{A}}\right)D^{+}A\left(\frac{1}{1+\beta_{A}}\right)
=A(11+βA)\displaystyle=A\left(\frac{1}{1+\beta_{A}}\right)
<A(11+α)+(111+α)D+A(11+α)\displaystyle<A\left(\frac{1}{1+\alpha}\right)+\left(1-\frac{1}{1+\alpha}\right)D^{+}A\left(\frac{1}{1+\alpha}\right)
=FA(11+α)\displaystyle=F_{A}\left(\frac{1}{1+\alpha}\right)

for all α(0,βA)\alpha\in(0,\beta_{A}). Since FA(1/(1+βA))=A(1/(1+βA))>0F_{A}\big{(}1/(1+\beta_{A})\big{)}=A\big{(}1/(1+\beta_{A})\big{)}>0 we therefore obtain log(FA(1/(1+βA)))<log(FA(1/(1+α)))\log\big{(}F_{A}\big{(}1/(1+\beta_{A})\big{)}\big{)}<\log\big{(}F_{A}\big{(}1/(1+\alpha)\big{)}\big{)} and hence

log(FA(11+βA))log(FA(11+α))g(α)g(βA)<0\frac{\log\left(F_{A}\left(\frac{1}{1+\beta_{A}}\right)\right)-\log\left(F_{A}\left(\frac{1}{1+\alpha}\right)\right)}{g(\alpha)-g(\beta_{A})}<0

This proves the assertion. ∎

Lemma 5.14 provides another structure for FAF_{A} that contradicts the MK-TP2 property. Its proof is given below and employs the results presented in Lemma 6.5 and Lemma 6.6.

Lemma 6.5.

For every aa\in\mathbb{R} the identity

u1a(h(u1,v1)h(u1,v2))v1a(h(u1,v1)h(u2,v1))u2a(h(u2,v2)h(u2,v1))v2a(h(u2,v2)h(u1,v2))=1\displaystyle u_{1}^{a(h(u_{1},v_{1})-h(u_{1},v_{2}))}v_{1}^{a(h(u_{1},v_{1})-h(u_{2},v_{1}))}u_{2}^{a(h(u_{2},v_{2})-h(u_{2},v_{1}))}v_{2}^{a(h(u_{2},v_{2})-h(u_{1},v_{2}))}=1

holds for all 0<u1u2<10<u_{1}\leq u_{2}<1 and all 0<v1v2<10<v_{1}\leq v_{2}<1.

Proof.

For simplicity we use the following abbreviation: hij:=h(ui,vj)h_{ij}:=h(u_{i},v_{j}). W.l.o.g. assume a=1a=1. Then straightforward calculation yields

log(u1h11h12v1h11h21u2h22h21v2h22h12)\displaystyle\log\big{(}u_{1}^{h_{11}-h_{12}}v_{1}^{h_{11}-h_{21}}u_{2}^{h_{22}-h_{21}}v_{2}^{h_{22}-h_{12}}\big{)}
=\displaystyle= (h11h12)log(u1)+(h11h21)log(v1)+(h22h21)log(u2)\displaystyle(h_{11}-h_{12})\log(u_{1})+(h_{11}-h_{21})\log(v_{1})+(h_{22}-h_{21})\log(u_{2})
+(h22h12)log(v2)\displaystyle+\,(h_{22}-h_{12})\log(v_{2})
=\displaystyle= h11(log(u1)+log(v1))h12(log(u1)+log(v2))\displaystyle h_{11}(\log(u_{1})+\log(v_{1}))-h_{12}(\log(u_{1})+\log(v_{2}))
h21(log(u2))+log(v1))+h22(log(u2))+log(v2))\displaystyle-\,h_{21}(\log(u_{2}))+\log(v_{1}))+h_{22}(\log(u_{2}))+\log(v_{2}))
=\displaystyle= log(u1)log(u1)log(u2)+log(u2)\displaystyle\log(u_{1})-\log(u_{1})-\log(u_{2})+\log(u_{2})
=\displaystyle= 0\displaystyle 0

which proves the result. ∎

Lemma 6.6.

Suppose that 0<t1<t2<s<10<t_{1}<t_{2}<s<1 such that

1sst21t2>1t2t2t11t1\displaystyle\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}>\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}}

Then the inequality

h(u2,v1)=t1<min{h(u1,v1),h(u2,v2)}max{h(u1,v1),h(u2,v2)}t2<s=h(u1,v2)h(u_{2},v_{1})=t_{1}<\min\{h(u_{1},v_{1}),h(u_{2},v_{2})\}\leq\max\{h(u_{1},v_{1}),h(u_{2},v_{2})\}\leq t_{2}<s=h(u_{1},v_{2})

holds for all u1,u2(0,1)u_{1},u_{2}\in(0,1) such that u11sst21t2<u2<u11t2t2t11t1u_{1}^{\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}}<u_{2}<u_{1}^{\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}}}, v1=ft1(u2)v_{1}=f_{t_{1}}(u_{2}) and v2=fs(u1)v_{2}=f_{s}(u_{1}).

Proof.

For simplicity we use the following abbreviation: hij:=h(ui,vj)h_{ij}:=h(u_{i},v_{j}). By assumption, u11sst21t2<u11t2t2t11t1u_{1}^{\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}}<u_{1}^{\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}}} which allows choosing u2u_{2} in between. Since 1sst21t2<1\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}<1 we first have u1<u2u_{1}<u_{2}. We further show that v1<v2v_{1}<v_{2} which can be seen as follows: since 1sst11t1<1t2t2t11t1\frac{1-s}{s}\frac{t_{1}}{1-t_{1}}<\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}} we obtain u2<u11t2t2t11t1<u11sst11t1u_{2}<u_{1}^{\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}}}<u_{1}^{\frac{1-s}{s}\frac{t_{1}}{1-t_{1}}} or, equivalently, v1=ft1(u2)=u21t1t1<u11ss=fs(u1)=v2v_{1}=f_{t_{1}}(u_{2})=u_{2}^{\frac{1-t_{1}}{t_{1}}}<u_{1}^{\frac{1-s}{s}}=f_{s}(u_{1})=v_{2}. Direct computation then yields h21=t1h_{21}=t_{1} and h12=sh_{12}=s, and monotonicity of hh implies

h21<h22=log(u2)log(u2v2)<log(u11sst21t2)log(u11sst21t2u11ss)=1sst21t21ss(t21t2+1)=t2<s\displaystyle h_{21}<h_{22}=\frac{\log(u_{2})}{\log(u_{2}v_{2})}<\frac{\log(u_{1}^{\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}})}{\log(u_{1}^{\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}}u_{1}^{\frac{1-s}{s}})}=\frac{\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}}{\frac{1-s}{s}(\frac{t_{2}}{1-t_{2}}+1)}=t_{2}<s

and

h21<h11=log(u1)log(u1v1)<log(u21t1t1t21t2)log(u21t1t1t21t2u21t1t1)=1t1t1t21t21t1t1(t21t2+1)=t2<s\displaystyle h_{21}<h_{11}=\frac{\log(u_{1})}{\log(u_{1}v_{1})}<\frac{\log(u_{2}^{\frac{1-t_{1}}{t_{1}}\frac{t_{2}}{1-t_{2}}})}{\log(u_{2}^{\frac{1-t_{1}}{t_{1}}\frac{t_{2}}{1-t_{2}}}u_{2}^{\frac{1-t_{1}}{t_{1}}})}=\frac{\frac{1-t_{1}}{t_{1}}\frac{t_{2}}{1-t_{2}}}{\frac{1-t_{1}}{t_{1}}(\frac{t_{2}}{1-t_{2}}+1)}=t_{2}<s

from which the assertion hence follows. ∎

We are now in the position to prove Lemma 5.14:

Proof.

[of Lemma 5.14] For simplicity we use the following abbreviation: hij:=h(ui,vj)h_{ij}:=h(u_{i},v_{j}). By assumption, we have FA(t)>0F_{A}(t)>0 for all t[t1,1]t\in[t_{1},1]. If FAF_{A} has a discontinuity point in [t1,1)[t_{1},1) then Lemma 6.3 implies that the corresponding EVC is not MK-TP2. Therefore, it remains to prove the result for FAF_{A} continuous on [t1,1)[t_{1},1).

The solution of the first-order differential equation FA(t)=A(t)+(1t)D+A(t)=cF_{A}(t)=A(t)+(1-t)D^{+}A(t)=c on [t1,t2][t_{1},t_{2}] is a linear function (see, e.g., [33]), i.e., there exist a,ba,b\in\mathbb{R} such that A(t)=at+bA(t)=at+b on [t1,t2][t_{1},t_{2}], hence a+b=at+b+(1t)a=FA(t)=c(0,1)a+b=at+b+(1-t)a=F_{A}(t)=c\in(0,1) for all t[t1,t2]t\in[t_{1},t_{2}]. W.l.o.g. set t2:=sup{t𝕀:FA(t)c}t_{2}:=\sup\{t\in{\mathbb{I}}\,:\,F_{A}(t)\leq c\}. Then FA(t)>c=a+bF_{A}(t)>c=a+b for all t(t2,1]t\in(t_{2},1].

In what follows we choose a rectangle u1u2u_{1}\leq u_{2} and v1v2v_{1}\leq v_{2} such that the corresponding values fulfill h21,h11,h22[t1,t2]h_{21},h_{11},h_{22}\in[t_{1},t_{2}], h12>t2h_{12}>t_{2} and hence FA(h21)=FA(h11)=FA(h22)=c<FA(h12)F_{A}(h_{21})=F_{A}(h_{11})=F_{A}(h_{22})=c<F_{A}(h_{12}) (see Figure 10 for an illustration), which we then use to show that CC is not MK-TP2.

1111t1t_{1}t2t_{2}h21h_{21}h21,h11,h22\in h_{21},h_{11},h_{22}
Figure 10: Example of FAF_{A} which is constant on [t1,t2][t_{1},t_{2}].

Since t(1t)/tt\mapsto(1-t)/t is strictly decreasing and continuous on (0,1)(0,1), we have (1t2)/t2t1/(1t1)<1(1-t_{2})/t_{2}\cdot t_{1}/(1-t_{1})<1 and there exists some s1(t2,1)s_{1}\in(t_{2},1) such that

1sst21t2>1t2t2t11t1\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}>\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}}

for all s(t2,s1]s\in(t_{2},s_{1}]. Moreover, since b[0,1]b\in[0,1] and c(0,1)c\in(0,1) we have a=cb(1,1)a=c-b\in(-1,1) such that the Pickands dependence function AA does not coincide with the lower bound tmax{1t,t}t\mapsto\max\{1-t,t\} on (t2,1](t_{2},1] (otherwise D+AD^{+}A would have a discontinuity point which then would contradict Lemma 6.3). Therefore we can extend the linear part of AA to a certain degree and are still above the lower bound (see Figure 11). Formally this means that there exists some s2(t2,1)s_{2}\in(t_{2},1) such that at+b>max{1t,t}1/2at+b>\max\{1-t,t\}\geq 1/2 for all t(t2,s2]t\in(t_{2},s_{2}].

1111t1t_{1}t2t_{2}s2s_{2}
Figure 11: Example of Pickands dependence function which is linear on [t1,t2][t_{1},t_{2}] with the described extension (red line).

Setting s:=(t2+min{s1,s2})/2s:=(t_{2}+\min\{s_{1},s_{2}\})/2, we then have

  • FA(s)>a+bF_{A}(s)>a+b

  • 1sst21t2>1t2t2t11t1\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}>\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}}

  • A(s)(as+b)112=12A(s)-(as+b)\leq 1-\frac{1}{2}=\frac{1}{2}

Now, choose u1(0,1)u_{1}\in(0,1) large enough such that

a+b<u112sFA(s)(u11s)A(s)(as+b)FA(s)\displaystyle a+b<u_{1}^{\frac{1}{2s}}F_{A}(s)\leq\big{(}u_{1}^{\frac{1}{s}}\big{)}^{A(s)-(as+b)}F_{A}(s) (6.1)

By choosing u2(0,1)u_{2}\in(0,1) such that u11sst21t2<u2<u11t2t2t11t1u_{1}^{\frac{1-s}{s}\frac{t_{2}}{1-t_{2}}}<u_{2}<u_{1}^{\frac{1-t_{2}}{t_{2}}\frac{t_{1}}{1-t_{1}}} and setting v1=ft1(u2)v_{1}=f_{t_{1}}(u_{2}) and v2=fs(u1)v_{2}=f_{s}(u_{1}), applying Lemma 6.6 yields

h21=t1<min(h11,h22)max(h11,h22)t2<s=h12h_{21}=t_{1}<\min(h_{11},h_{22})\leq\max(h_{11},h_{22})\leq t_{2}<s=h_{12}

For the Markov kernels, we then obtain

KA(u1,[0,v1])KA(u2,[0,v2])\displaystyle K_{A}(u_{1},[0,v_{1}])\,K_{A}(u_{2},[0,v_{2}]) =C(u1,v1)u1FA(h11)C(u2,v2)u2FA(h22)\displaystyle=\frac{C(u_{1},v_{1})}{u_{1}}F_{A}(h_{11})\frac{C(u_{2},v_{2})}{u_{2}}F_{A}(h_{22})
=(a+b)2u1u2(u1v1)ah11+b(u2v2)ah22+b\displaystyle=\frac{(a+b)^{2}}{u_{1}u_{2}}\left(u_{1}v_{1}\right)^{ah_{11}+b}\left(u_{2}v_{2}\right)^{ah_{22}+b} (6.2)

and

KA(u1,[0,v2])KA(u2,[0,v1])\displaystyle K_{A}(u_{1},[0,v_{2}])\,K_{A}(u_{2},[0,v_{1}]) =C(u1,v2)u1FA(h12)C(u2,v1)u2FA(h21)\displaystyle=\frac{C(u_{1},v_{2})}{u_{1}}F_{A}(h_{12})\frac{C(u_{2},v_{1})}{u_{2}}F_{A}(h_{21})
=a+bu1u2FA(s)(u1v2)A(s)(u2v1)ah21+b\displaystyle=\frac{a+b}{u_{1}u_{2}}F_{A}(s)\left(u_{1}v_{2}\right)^{A(s)}\left(u_{2}v_{1}\right)^{ah_{21}+b} (6.3)

and Lemma 6.5 yields v1a(h11h21)u2a(h22h21)=u1a(h12h11)v2a(h12h22)v_{1}^{a(h_{11}-h_{21})}u_{2}^{a(h_{22}-h_{21})}=u_{1}^{a(h_{12}-h_{11})}v_{2}^{a(h_{12}-h_{22})} which gives

KA(u1,[0,v1])KA(u2,[0,v2])KA(u1,[0,v2])KA(u2,[0,v1])\displaystyle\frac{K_{A}(u_{1},[0,v_{1}])\,K_{A}(u_{2},[0,v_{2}])}{K_{A}(u_{1},[0,v_{2}])\,K_{A}(u_{2},[0,v_{1}])} =\displaystyle= (a+b)FA(s)(u1v1)ah11+b(u2v2)ah22+b(u1v2)A(s)(u2v1)ah21+b\displaystyle\frac{(a+b)}{F_{A}(s)}\frac{\left(u_{1}v_{1}\right)^{ah_{11}+b}\left(u_{2}v_{2}\right)^{ah_{22}+b}}{\left(u_{1}v_{2}\right)^{A(s)}\left(u_{2}v_{1}\right)^{ah_{21}+b}}
=\displaystyle= (a+b)FA(s)(u1ah11+bA(s)v1a(h11h21)u2a(h22h21)v2ah22+bA(s))\displaystyle\frac{(a+b)}{F_{A}(s)}\big{(}u_{1}^{ah_{11}+b-A(s)}v_{1}^{a(h_{11}-h_{21})}u_{2}^{a(h_{22}-h_{21})}v_{2}^{ah_{22}+b-A(s)}\big{)}
=\displaystyle= (a+b)FA(s)(u1ah11+bA(s)u1a(h12h11)v2a(h12h22)v2ah22+bA(s))\displaystyle\frac{(a+b)}{F_{A}(s)}\big{(}u_{1}^{ah_{11}+b-A(s)}u_{1}^{a(h_{12}-h_{11})}v_{2}^{a(h_{12}-h_{22})}v_{2}^{ah_{22}+b-A(s)}\big{)}
=\displaystyle= (a+b)FA(s)(u1v2)ah12+bA(s)\displaystyle\frac{(a+b)}{F_{A}(s)}\,(u_{1}v_{2})^{ah_{12}+b-A(s)}
=\displaystyle= (a+b)FA(s)(u11s)as+bA(s)<1\displaystyle\frac{(a+b)}{F_{A}(s)}\,\big{(}u_{1}^{\frac{1}{s}}\big{)}^{as+b-A(s)}<1

Therefore, the EVC is not MK-TP2. ∎