1 Introduction

In recent years, copulas are hot topics in probability and statistics. By Sklar theorem [16], the importance of copulas comes from two aspects, (1) describing dependence properties of random variables, such as Joe [6], Nelsen [11], Siburg [15], Tasena [17], Shan [14], Wei [20]; and (2) constructing the joint distributions of random variables. In the second direction, there are many papers devoting to the constructions of bivariate copulas, such as Rodríguez-Lallena [12], Kim [7], Durante [4], Mesiar [9], Aguilo [1], Mesiar [10], but few of constructions of multivariate copulas, such as Liebscher [8], Durante [3].

In this paper, we discussed several general methods of constructing multivariate copulas, which are generalizations of some bivariate results. The paper is organized as follows: In Sect. 2, we introduce some necessary definitions and existing results. Several general methods for constructing multivariate copulas are provided in Sect. 3 and their dependence properties are discussed in Sect. 4. Finally, two examples are given in Sect. 5.

2 Definitions and Existing Results

A function \(C: I^n\rightarrow I\) is called an n-copula [11], where \(I=[0, 1]\), if C satisfies the following properties:

  1. (i)

    C is grounded, i.e., for any \(\mathbf u =(u_1, \cdots , u_n)^\prime \in I^n\), if at least one \(u_i=0\), then \(C(\mathbf u )=0\),

  2. (ii)

    One-dimensional marginals of C are uniformly distributed, i.e., for any \(u_i\in I\), \(i=1, \cdots , n\),

    $$C(1, \cdots , 1, u_i, 1, \cdots , 1)=u_i,$$
  3. (iii)

    C is n-increasing, i.e., for any \(\mathbf u \), \(\mathbf v \in I^n\) such that \(\mathbf u \le \mathbf v \), we have

    $$V_C([\mathbf u , \mathbf v ])=\sum sgn(\mathbf a )C(\mathbf a )\ge 0,$$

where the sum is taken over all vertices \(\mathbf a \) of the n-box \([\mathbf u , \mathbf v ]=[u_1, v_1]\times \cdots \times [u_n, v_n]\), and

$$\begin{aligned}sgn(\mathbf a )= {\left\{ \begin{array}{ll} \ \ 1, &{} \text {if}\ a_i=u_i\ \text {for an even number of}\ i^\prime s, \\ -1, &{} \text {if}\ a_i=u_i\ \text {for an odd number of}\ i^\prime s. \end{array}\right. } \end{aligned}$$

Equivalently,

$$V_C([\mathbf u , \mathbf v ])=\varDelta ^\mathbf{v }_\mathbf{u }C(\mathbf t )=\varDelta ^{v_n}_{u_n}\cdots \varDelta ^{v_1}_{u_1}C(\mathbf t ),$$

where \(\varDelta ^{v_k}_{u_k}C(\mathbf t )=C(t_1, \cdots , t_{k-1}, v_k, t_{k+1},\cdots , t_n)-C(t_1, \cdots , t_{k-1}, u_k, t_{k+1},\cdots , t_n)\), \(k=1,\cdots ,n\).

Note that above three conditions ensure that the range of C is I. By Sklar’s theorem [16], any n random variables \(X_1, \cdots , X_n\) can be connected by an n-copula via the equation

$$F(x_1, \cdots , x_n)=C(F_1(x_1), \cdots , F_n(x_n)),$$

where F is the joint distribution function of \({X_1}, \cdots , X_n\), \(F_i\) is the marginal distribution functions of \(X_i\), \(i=1, \cdots , n\). In addition, if \({X_1}, \cdots , X_n\) are continuous, then the copula C is unique.

There are three important functions for n-copulas defined respectively by

$$M_n(\mathbf u )=\min \{u_1, \cdots , u_n\},$$
$$\varPi _n(\mathbf u )=\underset{i=1}{\overset{n}{\prod }}u_i,$$

and

$$W_n(\mathbf u )=\max \{u_1+\cdots +u_n-n+1, 0\},$$

for all \(\mathbf u \in I^n\). Functions \(M_n\) and \(\varPi _n\) are n-copulas for all \(n\ge 2\), but \(W_n\) is not an n-copula for any \(n\ge 3\). \(M_n\) and \(W_n\) are called the Fréchert-Hoeffding upper bound and lower bound of n-copulas respectively since for any n-copula C, we have \(W_n\le C \le M_n\).

Let \(H: I^n\rightarrow \mathbb {R}\) be a function. The functions \(H_{i_1i_2\cdots i_k}: I^k\rightarrow \mathbb {R}\) are called k-dimensional marginals of H defined by

$$H_{i_1i_2\cdots i_k}(u_{i_1}, \cdots , u_{i_k})=H(v_1, \cdots ,\ v_n),$$

where \(v_j=u_{i_l}\) if \(j=i_l\) for some \(l=1, 2, \cdots , k\), otherwise, \(v_j=1\).

Any n-copula C defines a function \(\overline{C}: I^n\rightarrow I\) by

$$\begin{aligned} \overline{C}(\mathbf u )=1+ \underset{k=1}{\overset{n}{\sum }}(-1)^k\underset{1\le i_1<\cdots < i_k\le n}{\sum }C_{i_1i_2\cdots i_k}(u_{i_1}, \cdots , u_{i_k}). \end{aligned}$$
(1)

It is called the survival functionof C. For more details about copulas theory, see Nelsen’s book [11].

Now let’s recall some existing results. In 2004, Rodríguez-Lallena and Úbeda-Flores [12] considered the following family of bivariate copulas,

$$\begin{aligned} C_\theta (u, v)=uv+\theta f(u)g(v), \end{aligned}$$
(2)

where \(f, g: [0, 1]\rightarrow \mathbb {R}\) are two functions, \(\theta \in \mathbb {R}\) is a parameter. This family is a generalization of the well-known bivariate Farlie-Gumble-Morgenstern (or FGM, for short) family,

$$C_\theta (u, v)=uv+\theta uv(1-u)(1-v),$$

where \(u, v \in [0, 1]\) and \(\theta \in [-1, 1]\). In 2011, Kim et al. [7] extended Rodríguez-Lallena and Úbeda-Flores’s work to the family,

$$\begin{aligned} C(u, v)=C^*(u, v)+ \theta f(u)g(v), \end{aligned}$$
(3)

where \(C^*\) is a known bivariate copula, \(f, g: [0, 1]\rightarrow \mathbb {R}\) are two functions, \(\theta \) is a parameter. In 2013 and 2015, Durante et al. [4] and Mesiar et al. [10] considered more general cases,

$$\begin{aligned} C(u, v)=C^*(u, v)+ H(u, v), \end{aligned}$$
(4)

where \(C^*\) is a known bivariate copula, \(H: [0, 1]\times [0, 1]\rightarrow \mathbb {R}\) is a function.

3 Constructions of Multivariate Copulas

The constructions of all above results are adding some perturbation functions to a given bivariate copula. In fact, any n-copula C can be represented by a perturbation of the independent copula \(\varPi _n\) [19]. Based on this idea, we are going to extend these bivariate results to multivariate cases.

Firstly, for any given n-copula \(C^*: I^n\rightarrow [0, 1]\), we consider the construction,

$$\begin{aligned} C(u_1, u_2, \cdots , u_n)=C^*(u_1, u_2, \cdots , u_n)+H(u_1, u_2, \cdots , u_n), \end{aligned}$$
(5)

where \(H: I^n\rightarrow \mathbb {R}\) is a function, called a perturbation function. C is called a perturbation of \(C^*\) by H.

Theorem 1

Let \(C^*\) be an n-copula, \(H: I^n\rightarrow \mathbb {R}\) be a function. C is defined by (5). Then C is an n-copula if and only if H satisfies the following three conditions,

  1. (i)

    \(H(0, u_2, \cdots , u_n)=\cdots =H(u_1, \cdots , u_{n-1}, 0)=0\) for all \((u_1, \cdots , u_n)\in I^n\),

  2. (ii)

    There are \(1\le i<j\le n\) such that

    $$H(u_1, \cdots , u_{i-1}, 1, u_{i+1}, \cdots , u_n)=H(u_1, \cdots , u_{j-1}, 1, u_{j+1}, \cdots , u_n)=0,$$
  3. (iii)

    \(V_{C^*}([\mathbf u , \mathbf v ])+V_H([\mathbf u , \mathbf v ])\ge 0\) for all n-box \([\mathbf u , \mathbf v ]\) in \(I^n\).

Proof

The conditions (i) and (ii) ensure that C is grounded, and its one-dimensional marginals are uniform distributed, respectively. The n-increasing property of C is guaranteed by the condition (iii).       \(\square \)

Next we provide a necessary and sufficient condition on H under which C defined by (5) is an absolutely continuous n-copula.

Theorem 2

Let \(C^*\) be an absolutely continuous n-copula with the density \(c^*\), \(H: I^n\rightarrow \mathbb {R}\) be a non-zero absolutely continuous function with the Radon-Nikodym derivative h with respect to the Lebesgue measure on \(I^n\). C is defined by (5) is an absolutely continuous n-copula if and only if H satisfies the following conditions.

  1. (i)

    \(H(0, u_2, \cdots , u_n)=\cdots =H(u_1, \cdots , u_{n-1}, 0)=0\) for all \((u_i, \cdots , u_n)\in I^n\),

  2. (ii)

    There are \(1\le i<j\le n\) such that

    $$H(u_1, \cdots , u_{i-1}, 1, u_{i+1}, \cdots , u_n)=H(u_1, \cdots , u_{j-1}, 1, u_{j+1}, \cdots , u_n)=0,$$
  3. (iii)

    \(c^*+h\ge 0\) almost surely.

Proof

Firstly, the boundary conditions of copulas are ensured by the condition (i) and (ii).

Next, we show that the condition (iii) is equivalent to the n-increasing property of C. On the one hand, suppose that C is n-increasing. If \(c^*+h\) is not nonnegative almost surely, then there exist \(\mathbf u <\mathbf v \in I^n\) such that \(c^*+h< 0\) on \([\mathbf u , \mathbf v ]\). Note that \(V_C([\mathbf u , \mathbf v ])=\int _{[\mathbf u , \mathbf v ]}(c^*+h)(\mathbf t )d\mathbf t \). So \(V_C([\mathbf u , \mathbf v ])<0\). It contradicts the n-increasing property of C. On the other hand, if \(c^*+h\ge 0\) almost surely, we must have \(V_C([\mathbf u , \mathbf v ])\ge 0\) for all \(\mathbf u \), \(\mathbf v \in I^n\) with \(\mathbf u \le \mathbf v \).       \(\square \)

Now let’s consider a special case of (5) as follows, which are multivariate extensions of the result in [7].

$$\begin{aligned} C(u_1, u_2, \cdots , u_n)=C^*(u_1, u_2, \cdots , u_n)+ \overset{n}{\underset{i=1}{\prod }}f_i(u_i), \end{aligned}$$
(6)

where \(C^*\) is an n-copula, \(f_i:[0,1]\rightarrow \mathbb {R}\) is a function, \(i=1, 2, \cdots , n\).

The following theorem give us a sufficient condition under which C defined by (6) is an n-copula.

Theorem 3

Let \(C^*\) be an n-copula, \(f_i:[0,1]\rightarrow \mathbb {R}\) be a function, \(i=1, 2, \cdots , n\). \(C:[0, 1]^n\rightarrow \mathbb {R}\) is defined by (6) is an n-copula if \({f_1}, \cdots , f_n\) satisfy the following conditions,

  1. (i)

    \(f_1(0)=\cdots =f_n(0)=0\), and there exist at least two functions \(f_i\) and \(f_j\) such that \(f_i(1)=f_j(1)=0\), \(1\le i, j\le n\),

  2. (ii)

    \(f_i\) is absolutely continuous,

  3. (iii)

    \(\min (B)\ge \sup \left\{ - \dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}: \mathbf u , \mathbf v \in [0, 1]^n, \mathbf u <\mathbf v \right\} \), where \(B=\{\alpha _{i_1}\cdots \alpha _{i_k}\beta _{j_1}\cdots \beta _{j_{n-k}}: 1\le k\le n, k\ is\ odd, i_1, \cdots , i_k\ and\ j_1, \cdots , j_{n-k}\ are\ pairwise\ distinct\}\), \(\alpha _i=\inf \{f^{\prime }_i(u_i): u_i\in A_i\}<0\), \(\beta _i=\sup \{f^{\prime }_i(u_i): u_i\in A_i\}>0\), \(A_i=\{u_i \in [0, 1]: f^{\prime }(u_i)\ exists\}\), \(i=1, \cdots , n\), and \(\varDelta (\mathbf u , \mathbf v )=(v_1-u_1)\cdots (v_n-u_n)\).

Proof

Firstly, if there is \(f_i=0\), then \(C=C^*\) is an n-copula. So without loss of generality, we may assume that \(f_i\) is non-zero, \(i=1, \cdots , n\).

Since \(C^*\) is an n-copula, C is grounded and its marginals are uniformly distributed if and only if C satisfies the above condition (i). Next we are going to show that C is n-increasing if C satisfies the conditions (ii) and (iii).

Suppose that C satisfies conditions (ii) and (iii). By Lemma 2.2 in [12], it holds that for any \(\mathbf u \), \(\mathbf v \in I^n\) with \(\mathbf u <\mathbf v \),

$$\begin{aligned} \dfrac{(f_1(v_1)-f_1(u_1))\cdots (f_n(v_n)-f_n(u_n))}{(v_1-u_1)\cdots (v_n-u_n)}\ge -\dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}, \end{aligned}$$

i.e.,

$$V_{C}([\mathbf u , \mathbf v ])=V_{C^*}([\mathbf u , \mathbf v ])+(f_1(v_1)-f_1(u_1))\cdots (f_n(v_n)-f_n(u_n))\ge 0,$$

so C is n-increasing.       \(\square \)

Based on the construction (6), we introduce the following parametric families of n-copulas, which is a multivariate extension of (3).

$$\begin{aligned} C(u_1, u_2, \cdots , u_n)=C^*(u_1, u_2, \cdots , u_n)+ \theta \overset{n}{\underset{i=1}{\prod }}f_i(u_i), \end{aligned}$$
(7)

where \(C^*\) is an n-copula, \(f_i:[0,1]\rightarrow \mathbb {R}\) is a function, \(i=1, 2, \cdots , n\), \(\theta \in \mathbb {R}\).

Corollary 1

Let \(C^*\) be an n-copula, \(f_i:[0,1]\rightarrow \mathbb {R}\) be a function, \(i=1, 2, \cdots , n\). \(C:[0, 1]^n\rightarrow \mathbb {R}\) is defined by (6) is an n-copula if \({f_1}, \cdots , f_n\) and \(\theta \) satisfy the following conditions,

  1. (i)

    \(f_1(0)=\cdots =f_n(0)=0\), and there exist at least two functions \(f_i\) and \(f_j\) such that \(f_i(1)=f_j(1)=0\), \(1\le i, j\le n\),

  2. (ii)

    \(f_i\) is absolutely continuous,

  3. (iii)

    \(\sup \{- \dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}: \mathbf u , \mathbf v \in [0, 1]^n, \mathbf u<\mathbf v \}\dfrac{1}{\max (B^\prime )}\le \theta \le \sup \{- \dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}: \mathbf u , \mathbf v \in [0, 1]^n, \mathbf u <\mathbf v \}\dfrac{1}{\min (B)}\), where B is the same as Theorem 3, \(B^\prime =\{\alpha _{i_1}\cdots \alpha _{i_k}\beta _{j_1}\cdots \beta _{j_{n-k}}: 1\le k\le n,\ k\ is\ even, i_1, \cdots , i_k\ and\ j_1, \cdots , j_{n-k}\ are\ pairwise\ distinct\}\), \(\alpha _i=\inf \{f^{\prime }_i(u_i): u_i\in A_i\}<0\), \(\beta _i=\sup \{f^{\prime }_i(u_i): u_i\in A_i\}>0\), \(A_i=\{u_i \in [0, 1]: f^{\prime }(u_i)\ exists\}\), \(i=1, \cdots , n\), and \(\varDelta (\mathbf u , \mathbf v )=(v_1-u_1)\cdots (v_n-u_n)\).

Remark 1

Conditions in Theorem 3 and Corollary 1 are sufficient but may not be necessary. Consider the Fréchert-Hoeffding upper bound of n-copulas, \(M_n(u_1, \cdots , u_n)=\min \{u_1, \cdots , u_n\}\). For any \(\mathbf u , \mathbf v \in [0, 1]^n\) such that \(\mathbf u <\mathbf v \), it can be shown that

$$V_{M_n}([\mathbf u , \mathbf v ])=\max \{\min \{v_1, \cdots , v_n\}-\max \{u_1, \cdots , u_n\}, 0\}.$$

Thus,

$$\sup \left\{ - \dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}: \mathbf u , \mathbf v \in [0, 1]^n, \mathbf u <\mathbf v \right\} =0.$$

So functions \(f_1, \cdots , f_n\) that satisfy conditions in Theorem 3 or Corollary 1 for \(M_n\) must be zero, i.e., \(f_1=\cdots =f_n=0\).

Next we provide a stronger sufficient condition on \(f_1, \cdots , f_n\) to ensure that C defined by (6) is an n-copula. Example 2.1 in [12] shows that the condition is not necessary.

Theorem 4

Let C be defined by (6). C is an n-copula if \(f_1, \cdots , f_n\) satisfy the following conditions,

(i) \(f_1(0)=\cdots =f_n(0)=0\), and there exist at least two functions \(f_i\) and \(f_j\) such that \(f_i(1)=f_j(1)=0\), \(1\le i, j\le n\),

(ii) \(f_i\) satisfies the Lipschitz condition,

$$|f_i(v)-f_i(u)|\le M_i|v-u|,$$

for all \(u, v \in I\), such that \(M_i>0\), \(i=1, \cdots , n\), and

$$\underset{1}{\overset{n}{\prod }} M_i\le \inf \left\{ \dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}: \mathbf u , \mathbf v \in [0, 1]^n, \mathbf u \le \mathbf v \right\} .$$

Proof

By the condition (i), C is grounded and one-dimensional marginals of C are uniformly distributed. For any \(\mathbf u \), \(\mathbf v \in I^n\) with \(\mathbf u <\mathbf v \), by the condition (ii), we have

$$\begin{aligned} \begin{aligned} -\dfrac{(f_1(v_1)-f_1(u_1))\cdots (f_n(v_n)-f_n(u_n))}{(v_1-u_1)\cdots (v_n-u_n)}&\le \dfrac{|f_1(v_1)-f_1(u_1)|\cdots |f_n(v_n)-f_n(u_n)|}{|v_1-u_1|\cdots |v_n-u_n|}\\&\le \underset{1}{\overset{n}{\prod }} M_i\\&\le \inf \left\{ \dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}: \mathbf u , \mathbf v \in [0, 1]^n, \mathbf u \le \mathbf v \right\} . \end{aligned} \end{aligned}$$

So

$$\dfrac{(f_1(v_1)-f_1(u_1))\cdots (f_n(v_n)-f_n(u_n))}{(v_1-u_1)\cdots (v_n-u_n)}\ge \sup \left\{ - \dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}: \mathbf u , \mathbf v \in [0, 1]^n, \mathbf u \le \mathbf v \right\} .$$

Thus, as the proof of Theorem 3, C is n-increasing.       \(\square \)

4 Properties of New Families

In this section, we are going to study some non-parametric copula-based measures of multivariate association, some dependence concepts for copulas defined in Sect. 3 and some properties of those families.

Firstly, recall that the multivariate generalizations of Kendall’s tau, Spearman’s rho, and Blomqvist’s beta (see [13, 18] for details) are given by

$$\begin{aligned} \tau _n(C)=\dfrac{1}{2^{n-1}-1}\left[ 2^n\int _{I^n} C(\mathbf u )dC(\mathbf u )-1\right] , \end{aligned}$$
(8)
$$\begin{aligned} \rho _n(C)=\dfrac{n+1}{2^{n}-n-1}\left[ 2^{n-1} \left( \int _{I^n} C(\mathbf u )d\varPi _n(\mathbf u )+\int _{I^n} \varPi _n(\mathbf u )dC(\mathbf u )\right) -1\right] , \end{aligned}$$
(9)
$$\begin{aligned} \beta _n(C)=\dfrac{2^{n-1}\left[ C\left( \frac{1}{2}{} \mathbf 1 _n\right) +\overline{C}\left( \frac{1}{2}{} \mathbf 1 _n\right) \right] -1}{2^{n-1}-1}, \end{aligned}$$
(10)

where \(\mathbf 1 _n\) is the vector \((1, \cdots , 1)^\prime \in \mathbb {R}^n\).

Theorem 5

Let C be an n-copula defined by (5), then the Kendall’s tau, Spearman’s rho, and Blomqvist’s beta of C are given by

$$\begin{aligned} \tau _n(C)=\tau _n(C^*)+\tau _n(H)+a_1, \end{aligned}$$
(11)
$$\begin{aligned} \rho _n(C)=\rho _n(C^*)+\rho _n(H)+\dfrac{n+1}{2^{n}-n-1}, \end{aligned}$$
(12)
$$\begin{aligned} \beta _n(C)=\beta _n(C^*)+\beta _n(H)+\dfrac{1-2^{n-1}}{2^n-1}, \end{aligned}$$
(13)

where \(a_1=\dfrac{1}{2^{n-1}-1}\left[ 2^n\int _{I^n}C^*(\mathbf u )dH(\mathbf u )+2^n\int _{I^n}H(\mathbf u )dC^*(\mathbf u )+1\right] \).

Proof

Firstly, by the definition of \(\tau _n\),

$$\begin{aligned} \begin{aligned} \tau _n(C)&=\dfrac{1}{2^{n-1}-1}\left[ 2^n\int _{I^n} C(\mathbf u )dC(\mathbf u )-1\right] \\&=\dfrac{1}{2^{n-1}-1}\left[ 2^n\int _{I^n} C^*(\mathbf u )+ H(\mathbf u )d(C^*(\mathbf u )+ H(\mathbf u ))-1\right] \\&=\dfrac{1}{2^{n-1}-1}[2^n\int _{I^n} C^*(\mathbf u )dC^*(\mathbf u )+2^n\int _{I^n}H(\mathbf u )dH(\mathbf u )\\&\ \ \ \ \ \ +2^n\int _{I^n}C^*(\mathbf u )dH(\mathbf u )+2^n\int _{I^n}H(\mathbf u )dC^*(\mathbf u )-1]\\&=\tau _n(C^*)+\tau _n(H)+a_1. \end{aligned} \end{aligned}$$

Secondly, by the definition of \(\rho _n\),

$$\begin{aligned} \begin{aligned} \rho _n(C)&=\dfrac{n+1}{2^{n}-n-1}\left\{ 2^{n-1} \left[ \int _{I^n} C(\mathbf u )d\varPi _n(\mathbf u )+\int _{I^n} \varPi _n(\mathbf u )dC(\mathbf u )\right] -1\right\} \\&=\dfrac{n+1}{2^{n}-n-1}\left\{ 2^{n-1} \left[ \int _{I^n} C^*(\mathbf u )+ H(\mathbf u )d\varPi _n(\mathbf u )+\int _{I^n} \varPi _n(\mathbf u )d(C^*(\mathbf u )+ H(\mathbf u ))\right] -1\right\} \\&=\dfrac{n+1}{2^{n}-n-1}\{2^{n-1} [\int _{I^n} C^*(\mathbf u )d\varPi _n(\mathbf u )+\int _{I^n}H(\mathbf u )d\varPi _n(\mathbf u )+ \int _{I^n} \varPi _n(\mathbf u )dC^*(\mathbf u )\\&\ \ \ \ \ +\int _{I^n}\varPi _n(\mathbf u )dH(\mathbf u )]-1\}\\&=\rho _n(C^*)+\rho _n(H)+\dfrac{n+1}{2^{n}-n-1}. \end{aligned} \end{aligned}$$

Lastly, by the definition of survival functions, for any \(\mathbf u \in I^n\),

$$\begin{aligned} \begin{aligned} \overline{C}(\mathbf u )&=1+ \underset{k=1}{\overset{n}{\sum }}(-1)^k\underset{1\le i_1<\cdots< i_k\le n}{\sum }C_{i_1i_2\cdots i_k}(u_{i_1}, \cdots , u_{i_k})\\&=1+ \underset{k=1}{\overset{n}{\sum }}(-1)^k\underset{1\le i_1<\cdots< i_k\le n}{\sum }\left[ C^*_{i_1i_2\cdots i_k}(u_{i_1}, \cdots , u_{i_k})+H_{i_1i_2\cdots i_k}(u_{i_1}, \cdots , u_{i_k})\right] \\&=1+ \underset{k=1}{\overset{n}{\sum }}(-1)^k\underset{1\le i_1<\cdots< i_k\le n}{\sum }C^*_{i_1i_2\cdots i_k}(u_{i_1}, \cdots , u_{i_k})\\&\ \ \ \ \ +1+ \underset{k=1}{\overset{n}{\sum }}(-1)^k\underset{1\le i_1<\cdots < i_k\le n}{\sum }H_{i_1i_2\cdots i_k}(u_{i_1}, \cdots , u_{i_k})-1\\&=\overline{C^*}(\mathbf u )+\overline{H}(\mathbf u )-1. \end{aligned} \end{aligned}$$

Thus,

$$\begin{aligned} \begin{aligned} \beta _n(C)&=\dfrac{2^{n-1}\left[ C\left( \frac{1}{2}{} \mathbf 1 _n\right) +\overline{C}\left( \frac{1}{2}{} \mathbf 1 _n\right) \right] -1}{2^{n-1}-1}\\&=\dfrac{2^{n-1}\left[ C^*\left( \frac{1}{2}{} \mathbf 1 _n\right) +H\left( \frac{1}{2}{} \mathbf 1 _n\right) +\overline{C^*}\left( \frac{1}{2}{} \mathbf 1 _n\right) +\overline{H}\left( \frac{1}{2}{} \mathbf 1 _n\right) -1\right] -1}{2^{n-1}-1}\\&=\dfrac{2^{n-1}\left[ C^*\left( \frac{1}{2}{} \mathbf 1 _n\right) +\overline{C^*}\left( \frac{1}{2}{} \mathbf 1 _n\right) \right] -1 +2^{n-1}\left[ H\left( \frac{1}{2}{} \mathbf 1 _n\right) +\overline{H}\left( \frac{1}{2}{} \mathbf 1 _n\right) \right] -1+1-2^{n-1}}{2^{n-1}-1}\\&=\beta _n(C^*)+\beta _n(H)+\dfrac{1-2^{n-1}}{2^n-1}. \end{aligned} \end{aligned}$$

      \(\square \)

Remark 2

In the above theorem, although the perturbation function H is not a copula, we still use m(H) to denote the corresponding values of H, where \(m=\tau _n\), \(\rho _n,\) or \(\beta _n\), and use \(\overline{H}\) to denoted the corresponding function of H defined by (1). The similar notations are used in the following context.

Remark 3

As n increasing, we can see that

$$\tau _n(C)\approx \tau _n(C^*)+\tau _n(H)+2\int _{I^n}C^*(\mathbf u )dH(\mathbf u )+2\int _{I^n}H(\mathbf u )dC^*(\mathbf u ),$$
$$\rho _n(C)\approx \rho _n(C^*)+\rho _n(H),$$

and

$$\beta _n(C)\approx \beta _n(C^*)+\beta _n(H).$$

Corollary 2

Let C be an n-copula defined by (7), then the Kendall’s tau, Spearman’s rho, and Blomqvist’s beta of C are given by

$$\begin{aligned} \tau _n(C)=\tau _n(C^*)+\tau _n(\theta \overset{n}{\underset{i=1}{\prod }}f_i)+a_2, \end{aligned}$$
(14)
$$\begin{aligned} \rho _n(C)=\rho _n(C^*)+\rho _n(\theta \overset{n}{\underset{i=1}{\prod }}f_i)+\dfrac{n+1}{2^{n}-n-1}, \end{aligned}$$
(15)
$$\begin{aligned} \beta _n(C)=\beta _n(C^*)+\beta _n(\theta \overset{n}{\underset{i=1}{\prod }}f_i)+\dfrac{1-2^{n-1}}{2^n-1}, \end{aligned}$$
(16)

where \(a_2=\dfrac{1}{2^{n-1}-1}\left[ 2^n\int _{I^n}\theta C^*(\mathbf u )\overset{n}{\underset{i=1}{\prod }}{f_i}^\prime (u_i)d\mathbf u +2^n\int _{I^n}\theta \overset{n}{\underset{i=1}{\prod }}f_i(u_i)dC^*(\mathbf u )+1\right] \).

In 2013, Tasena et al. [17] defined a measure of multivariate complete dependence as follows. Let C be an n-copula of random variables \(X_1, \cdots , X_n\). Define

$$\delta _i(X_1, \cdots , X_n)=\delta _i(C)=\dfrac{\int (\partial _iC-\pi _iC)^2}{\int \pi _iC(1-\pi _iC)}, $$

where \(\pi _iC: I^{n-1}\rightarrow I\) is defined by

$$\pi _iC(u_1, \cdots , u_{n-1}) =C(u_1, \cdots , u_{i-1}, 1, u_{i}, \cdots , u_{n-1}),\qquad i=1, 2, \cdots , n.$$

By Theorem 3.6 in [17], \(\delta _i\) satisfies following properties,

  1. (i)

    \(0\le \delta _i(C)\le 1\),

  2. (ii)

    \(\delta _i(C)=1\) if and only if \((X_1, \cdots , X_{i-1}, X_{i+1}, \cdots , X_{n})\) is a function of \(X_i\). For details, see [17].

Theorem 6

Let C be an n-copula defined by (5). If

$$H(u_1, \cdots , u_{i-1}, 1, u_{i+1}, \cdots , u_{n_1})=0,$$

then

$$\delta _i(C)=\delta (C^*)-\dfrac{\int 2\partial _iH(\partial _iC^*-\pi _iC^*)-(\partial _iH)^2}{\int \pi _iC^*(1-\pi _iC^*)}.$$

Proof

By the definition,

$$\begin{aligned} \begin{aligned} \delta _i(C)&=\dfrac{\int (\partial _iC-\pi _iC)^2}{\int \pi _iC(1-\pi _iC)}\\&=\dfrac{\int (\partial _i(C^*+H)-\pi _i(C^*+H))^2}{\int \pi _i(C^*+H)[1-\pi _i(C^*+H)]}\\&=\dfrac{\int (\partial _iC^*+\partial _iH-\pi _iC^*-\pi _iH)^2}{\int (\pi _iC^*+\pi _iH)(1-\pi _iC^*-\pi _iH)}. \end{aligned} \end{aligned}$$

If \(\pi _iH(u_1, \cdots , u_{n_1})=H(u_1, \cdots , u_{i-1}, 1, u_{i+1}, \cdots , u_{n_1})=0,\) then

$$\begin{aligned} \begin{aligned} \delta _i(C)&=\dfrac{\int (\partial _iC^*+\partial _iH-\pi _iC^*-\pi _iH)^2}{\int (\pi _iC^*+\pi _iH)(1-\pi _iC^*\pi _iH)}\\&=\dfrac{\int (\partial _iC^*+\partial _iH-\pi _iC^*)^2}{\int \pi _iC^*(1-\pi _iC^*)} \\&=\dfrac{\int \left[ (\partial _iC^*-\pi _iC^*)^2-2\partial _iH(\partial _iC^*-\pi _iC^*)+(\partial _iH)^2\right] }{\int \pi _iC^*(1-\pi _iC^*)}\\&=\dfrac{\int (\partial _iC^*-\pi _iC^*)^2}{\int \pi _iC^*(1-\pi _iC^*)}-\dfrac{\int 2\partial _iH(\partial _iC^*-\pi _iC^*)-(\partial _iH)^2}{\int \pi _iC^*(1-\pi _iC^*)}\\&=\delta (C^*)-\dfrac{\int 2\partial _iH(\partial _iC^*-\pi _iC^*)-(\partial _iH)^2}{\int \pi _iC^*(1-\pi _iC^*)}. \end{aligned} \end{aligned}$$

      \(\square \)

Corollary 3

Let C be an n-copula defined by (7). If \(f_i(1)=0,\) then

$$ \delta _i(C)=\delta _i(C^*)-\dfrac{\int 2\theta f_i^{\prime } \prod \limits _{j\ne i} f_j(\partial _iC^*-\pi _iC^*)-(\theta f_i^{\prime } \prod \limits _{j\ne i} f_j)^2}{\int \pi _iC^*(1-\pi _iC^*)}. $$

Now, let’s recall some dependence concepts of copulas. For details, see [6, 11]. Let \(C_1\) and \(C_2\) be two n-copulas. If \(C_1\ge C_2\) (\(\overline{C}_1\ge \overline{C}_2\) resp.), i.e., \(C_1(\mathbf u )\ge C_2(\mathbf u )\) (\(\overline{C}_1(\mathbf u )\ge \overline{C}_2(\mathbf u )\) resp.) for all \(\mathbf u \in I^n\), then we say that \(C_1\) is more positive lower (upper resp.) orthant dependent (PLOD) (PUOD resp.) than \(C_2\). \(C_1\) is more positive orthant dependent (POD) than \(C_2\) if \(C_1\ge C_2\) and \(\overline{C}_1\ge \overline{C}_2\) hold.

The following results give us some dependence relations between C and \(C^*\). The proof is trivial.

Proposition 1

Let \(C_1\) and \(C_2\) be two n-copulas defined by (5). If they share the same n-copula \(C^*\) and may have different perturbation functions \(H_i\) \(i=1,2\), then

(i) \(C_1\) more PLOD than \(C_2\) if and only if \(H_1\ge H_2\),

(ii) \(C_1\) more PUOD than \(C_2\) if and only if \(\overline{H}_1\ge \overline{H}_2\),

(iii) \(C_1\) more POD than \(C_2\) if and only if \(H_1\ge H_2\) and \(\overline{H}_1\ge \overline{H}_2\).

Proposition 2

Let \(C_1\) and \(C_2\) are two n-copulas defined by (7). If they share the same known n-copula \(C^*\) and may have different perturbation functions \(f_{j1}, \cdots , f_{jn}\), and parameters \(\theta _j\), \(j=1,2\) respectively, then

(i) \(C_1\) more PLOD than \(C_2\) if and only if \(\theta _1\overset{n}{\underset{i=1}{\prod }}f_{1i}\ge \theta _2\overset{n}{\underset{i=1}{\prod }}f_{2i}\),

(ii) \(C_1\) more PUOD than \(C_2\) if and only if \(\overline{\theta _1\overset{n}{\underset{i=1}{\prod }}f_{1i}}\ge \overline{\theta _2\overset{n}{\underset{i=1}{\prod }}f_{2i}}\),

(iii) \(C_1\) more POD than \(C_2\) if and only if \(\theta _1\overset{n}{\underset{i=1}{\prod }}f_{1i}\ge \theta _2\overset{n}{\underset{i=1}{\prod }}f_{2i}\) and \(\overline{\theta _1\overset{n}{\underset{i=1}{\prod }}f_{1i}}\ge \overline{\theta _2\overset{n}{\underset{i=1}{\prod }}f_{2i}}\).

The next theorem give us a property of the construction (6).

Theorem 7

Let \((U_1^*, \cdots , U_n^*)\) and \((U_1, \cdots , U_n)\) be random vectors with uniform marginals on [0, 1] and connected by copulas \(C^*\) and C respectively. C and \(C^*\) satisfy conditions of Theorem 3. Suppose that \(f_i(1)=f_j(1)=0\), \(1\le i<j\le n\).

(i) If there is \(1\le l \le n\) such that \(l\ne i, j\) and \(f_l(1)=0\), then \(P\{U_i<U_j\}=P\{U_i^*<U_j^*\}\),

(ii) If \(f_l(1)\ne 0\) for all \(l\ne i, j\) and \(f_i=f_j\), then \(P\{U_i<U_j\}=P\{U_i^*<U_j^*\}\).

Proof

(i) Let c and \(c^*\) be the densities of C and \(C^*\) respectively, then we have

$$c(\mathbf u )=\dfrac{\partial ^nC(\mathbf u )}{\partial u_1\cdots \partial u_n}=c^*(\mathbf u )+ \overset{n}{\underset{i=1}{\prod }}{f_i}^\prime (u_i).$$

Then

$$\begin{aligned} \begin{aligned} P\{U_i<U_j\}&=\int _0^1\cdots \int _0^{u_j}\cdots \int _0^1c(u_1,\cdots , u_i,\cdots , u_n)du_1\cdots du_i \cdots du_n \\&=\int _0^1\cdots \int _0^{u_j}\cdots \int _0^1c^*(u_1,\cdots , u_i,\cdots , u_n)du_1\cdots du_i \cdots du_n\\&\ \ \ \ +\int _0^1\cdots \int _0^{u_j}\cdots \int _0^1 f^\prime _1(u_i)\cdots f^\prime _i(u_i)\cdots f^\prime _n(u_n) du_1\cdots du_i\cdots du_n\\&=P\{U^*_i<U^*_j\}+ \overset{n}{\underset{k\ne i, j}{\prod }}(f_k(1)-f_k(0))\int _0^1\int _0^{u_j}f^\prime _j(u_j)f^\prime _i(u_i)du_idu_j\\&=P\{U^*_i<U^*_j\}, \end{aligned} \end{aligned}$$

since \(f_l(0)=f_l(1)=0\) and \(l\ne i, j\).

(ii) Similarly, if \(f_l(1)\ne 0\) for all \(l\ne i, j\),

$$\begin{aligned} \begin{aligned} P\{U_i<U_j\}&=P\{U^*_i<U^*_j\}+ \overset{n}{\underset{k\ne i, j}{\prod }}(f_k(1)-f_k(0))\int _0^1\int _0^{u_j}f^\prime _j(u_j)f^\prime _i(u_i)du_idu_j\\&=P\{U^*_i<U^*_j\}+ \overset{n}{\underset{k\ne i, j}{\prod }}f_k(1)\int _0^1\int _0^{u_j}f^\prime _j(u_j)f^\prime _i(u_i)du_idu_j\\&=P\{U^*_i<U^*_j\}+ \overset{n}{\underset{k\ne i, j}{\prod }}f_k(1)\int _0^1f_i(u_j)f^\prime _j(u_j)du_j. \end{aligned} \end{aligned}$$

Since \(f_i=f_j\),

$$\begin{aligned} \begin{aligned} \int _0^1f_i(u_j)f^\prime _j(u_j)du_j&=f_j(1)f_i(1)-f_j(0)f_i(0)-\int _0^1f_j(u_j)f^\prime _i(u_j)du_j\\&=-\int _0^1f_j(u_j)f^\prime _i(u_j)du_j=-\int _0^1f_i(u_j)f^\prime _j(u_j)du_j, \end{aligned} \end{aligned}$$

and hence \(\int _0^1f_i(u_j)f^\prime _j(u_j)du_j=0\). So \(P\{U_i<U_j\}=P\{U_i^*<U_j^*\}.\)       \(\square \)

The following example shows that the converse of the above result (ii) in Theorem 7 may not hold in general. Moreover, it shows that Theorem 3 in [7] is incorrect.

Example 1

Let \((U^*, V^*)\) and (UV) be random vectors with uniform marginals on [0, 1]. Suppose that \((U^*, V^*)\) is connected by the independent copula, i.e., \(C^*(u, v)=uv\), and (UV) is connected by \(C(u, v)=C^*(u, v)+f(u)g(v)\), where \(f(u)=u(1-u)\), \(g(v)=\dfrac{1}{2}v(1-v)\). Then f and g satisfy the conditions in Theorem 3. In fact, C belongs to the bivariate FGM family.

As the proof of the above theorem, we have

$$\begin{aligned} \begin{aligned} P\{U<V\}&=\int _0^1\int _0^v c(u, v)dudv =\int _0^1\int _0^v c^*(u, v)+f^\prime (u)g^\prime (v)dudv \\&=P\{U^*<V^*\}+\int _0^1\int _0^v f^\prime (u)g^\prime (v)dudv=P\{U^*<V^*\}+\int _0^1 f(v)g^\prime (v)dv.\\ \end{aligned} \end{aligned}$$

where

$$\begin{aligned} \begin{aligned} \int _0^1 f(v)g^\prime (v)dv&= \int _0^1 \frac{1}{2}v(1-v)(1-2v)dv= 0. \end{aligned} \end{aligned}$$

Thus \(P\{U<V\}=P\{U^*<V^*\}\), but \(f\ne g\).

5 Examples

In this section, we provide two examples. The given copula \(C^*\) in the first example is the simplest one, the independent copula. To emphasis multivariate and for simplicity, we will only consider 3-copulas, but results could be extended to n-copulas. In the second example, \(C^*\) is nontrivial. Also for simplicity, we will only consider 2-copulas.

Example 2

Let \(C^*\) be the independent 3-copula, i.e., \(C^*(u,v,w)=uvw\). Let \(f(x)=x(1-x^k)\), where \(u, v, w, x \in I\), \(k\in \mathbb {N}\), the set of all positive integers. Consider the 3-copula family,

$$\begin{aligned} \begin{aligned} C(u, v, w)&= C^*(u,v,w)+\theta f(u)f(v)f(w)\\&= uvw+\theta uvw(1-u^k)(1-v^k)(1-w^k), \end{aligned} \end{aligned}$$

where \(\theta \in \mathbb {R}\).

It is clear that f(x) satisfies the conditions (i) and (ii) of Corollary 1. Next we will use the condition (iii) of Corollary 1 to find the range of the parameter \(\theta \) for each k. Firstly, it is easy to see that \(\dfrac{V_{C^*}([\mathbf u , \mathbf v ])}{\varDelta (\mathbf u , \mathbf v )}=1\) for any \(\mathbf u , \mathbf v \in [0, 1]^3\) with \(\mathbf u <\mathbf v \). Secondly, \(f^\prime (x)=1-(k+1)x^k\), so

$$\alpha =\inf \{f^\prime (x): x\in I\}=f^\prime (1)=1-(k+1)=-k,$$

and

$$\beta =\sup \{f^\prime (x): x\in I\}=f^\prime (0)=1.$$

Thus, as the notations in Theorem 3, \(B=\{-k, -k^3\}\), \(B^\prime =\{k^2\}\). So by the condition (iii) of Corollary 1, the range of \(\theta \) is

$$-\dfrac{1}{\max (B^\prime )}\le \theta \le -\dfrac{1}{\min (B)}, $$

i.e.,

$$-\dfrac{1}{k^2}\le \theta \le \dfrac{1}{k^3}.$$

So, we can see that the range of \(\theta \) is shrinking as k increasing. Specifically, if \(k=1\), \(-1\le \theta \le 1\). If \(k=2\), \(-\dfrac{1}{4}\le \theta \le \dfrac{1}{8}.\) If \(k=3\), \(-\dfrac{1}{9}\le \theta \le \dfrac{1}{27}.\)

Next, let’s compute three measures discussed in Sect. 4 for these 3-copulas. By the definition of \(\tau _n\),

$$\tau _3\left( \theta f(u)f(v)f(w)\right) =\dfrac{1}{3}\left[ 8\theta (\dfrac{k+1}{2k+3}-\dfrac{k+2}{k+3}+\dfrac{1}{3})-1\right] .$$
$$\begin{aligned} \begin{aligned} a_2&=\dfrac{1}{3}\left[ 8\int _{I^3}\theta C^*(\mathbf u )\overset{3}{\underset{i=1}{\prod }}{f_i}^\prime (u_i)d\mathbf u +8\int _{I^3}\theta \overset{3}{\underset{i=1}{\prod }}f_i(u_i)dC^*(\mathbf u )+1\right] \\&=\dfrac{1}{3}\left[ -\dfrac{\theta k^3}{(k+2)^3}+\dfrac{\theta k^3}{(k+2)^3}+1\right] = \dfrac{1}{3}. \end{aligned} \end{aligned}$$

So by Corollary 2,

$$\begin{aligned} \begin{aligned} \tau _3(C)&=\tau _3(C^*)+\tau _3(\theta f(u)f(v)f(w))+a_2\\&= 0+\dfrac{1}{3}\left[ 8\theta \left( \dfrac{k+1}{2k+3}-\dfrac{k+2}{k+3}+\dfrac{1}{3}\right) -1\right] +\dfrac{1}{3}\\&= \dfrac{8\theta }{3}\left( \dfrac{k+1}{2k+3}-\dfrac{k+2}{k+3}+\dfrac{1}{3}\right) . \end{aligned} \end{aligned}$$

So the range of \(\tau _3(C)\) is

$$\dfrac{8}{3k^3}\left( \dfrac{k+1}{2k+3}-\dfrac{k+2}{k+3}+\dfrac{1}{3}\right) \le \tau _3(C)\le -\dfrac{8}{3k^2}\left( \dfrac{k+1}{2k+3}-\dfrac{k+2}{k+3}+\dfrac{1}{3}\right) .$$

By the definition of \(\rho _n\),

$$\rho _3(\theta f(u)f(v)f(w))=4\left[ \dfrac{\theta k^3}{8(k+2)^3}-\dfrac{\theta k^3}{8(k+2)^3}\right] -1=-1.$$

So

$$\begin{aligned} \begin{aligned} \rho _3(C)&=\rho _3(C^*)+\rho _3(\theta f(u)f(v)f(w))+\dfrac{3+1}{2^3-3-1}\\&= 0-1+1 = 0. \end{aligned} \end{aligned}$$

By the definition of survival function (1),

$$\overline{\theta \overset{3}{\underset{i=1}{\prod }}f_i}\left( \frac{1}{2}, \frac{1}{2}, \frac{1}{2}\right) =1-\dfrac{\theta }{8}\left( 1-\dfrac{1}{2^k}\right) ^3.$$

So

$$\begin{aligned} \begin{aligned} \beta _3(\theta f(u)f(v)f(w))&=\dfrac{2^{2}\left[ \theta \overset{3}{\underset{i=1}{\prod }}f_i\left( \frac{1}{2}, \frac{1}{2}, \frac{1}{2}\right) +\overline{\theta \overset{3}{\underset{i=1}{\prod }}f_i}\left( \frac{1}{2}, \frac{1}{2}, \frac{1}{2}\right) \right] -1}{2^{2}-1}\\&=\dfrac{4\left[ \dfrac{\theta }{8}\left( 1-\dfrac{1}{2^k}\right) ^3 +1-\dfrac{\theta }{8}\left( 1-\dfrac{1}{2^k}\right) ^3\right] -1}{3}=1. \end{aligned} \end{aligned}$$

Thus,

$$\begin{aligned} \begin{aligned} \beta _3(C)&=\beta _3(C^*)+\beta _3(\theta f(u)f(v)f(w))+\dfrac{1-2^{3-1}}{2^3-1}\\&=0+1-\frac{3}{7}= \frac{4}{7}. \end{aligned} \end{aligned}$$

Lastly, since \(f(u)f(v)f(w)=uvw(1-u^k)(1-v^k)(1-w^k)\ge 0\) for all \((u,v,w)\in I^3\), we have that C is more PLOD than \(\varPi _3\) if and only if \(\theta \ge 0\) and \(\varPi _3\) is more PLOD than C if and only if \(\theta \le 0\).

Remark 4

From the above example, we can see that this 3-copulas family, \(C(u, v, w)= C^*(u, v, w)+\theta uvw(1-u^k)(1-v^k)(1-w^k)\), is interesting. As long as this C is a 3-copula, \(\rho _3(C)\) and \(\beta _3(C)\) are free of \(\theta \). Specifically, we always have \(\rho _3(C)=\rho _3(C^*)\) and \(\beta _3(C)=\beta _3(C^*)+\frac{4}{7}\).

Example 3

Let \(C^*\) be a Frank’s copula [2, 5] defined by

$$C^*(u, v)=\ln \left[ 1+\dfrac{(e^u-1)(e^v-1)}{e-1}\right] .$$

Let

$$H=\theta (1-u)(1-e^u)(1-v)(1-e^v),$$

where \(\theta \ge 0\). Define a bivariate function C by \(C=C^*+H.\) We will use Theorem 2 to find the range of \(\theta \) such that C is a copula.

Firstly, it is easy to see that \(H(0, v)=H(u, 0)=H(1, v)=H(u, 1)=0\).

Secondly, we can find that

$$c^*(u, v)=\dfrac{(e-1)(u+v)}{[e-1+(e^u-1)(e^v-1)]^2},$$

and

$$h(u, v)=\theta (ue^u-1)(ve^v-1).$$

It can be shown that minimum values of \(c=c^*+h\) occur at (0, 1) and (1, 0). So \(c\ge 0\) if and only if \(c(0,1)=c(1,0)=c^*(0, 1)+h(0,1)=\dfrac{1}{e-1}-\theta (e-1)\ge 0\). Thus \(C=C^*+H\) is a copula if \(\theta \le \dfrac{1}{(e-1)^2}\).