Skip to main content
Log in

Estimating non-simplified vine copulas using penalized splines

  • Published:
Statistics and Computing Aims and scope Submit manuscript

Abstract

Vine copulas (or pair-copula constructions) have become an important tool for high-dimensional dependence modeling. Typically, so-called simplified vine copula models are estimated where bivariate conditional copulas are approximated by bivariate unconditional copulas. We present the first nonparametric estimator of a non-simplified vine copula that allows for varying conditional copulas using penalized hierarchical B-splines. Throughout the vine copula, we test for the simplifying assumption in each edge, establishing a data-driven non-simplified vine copula estimator. To overcome the curse of dimensionality, we approximate conditional copulas with more than one conditioning argument by a conditional copula with the first principal component as conditioning argument. An extensive simulation study is conducted, showing a substantial improvement in the out-of-sample Kullback–Leibler divergence if the null hypothesis of a simplified vine copula can be rejected. We apply our method to the famous uranium data and present a classification of an eye state data set, demonstrating the potential benefit that can be achieved when conditional copulas are modeled.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. For instance, choosing \(K=9\) in (5) results in 729 coefficients (see Table 1).

  2. We do not use the asymptotic normal distribution of Vuong’s test because both over-rejection and under-rejection can be severe when the competing models contain a rather large number of parameters, see Shi (2015). Moreover, we use an out-of-sample version of Vuong’s test because the in-sample version does not account for the number of parameters.

  3. Note that \(\beta \) could be increased up to the value of one, so that the rather strong variation for \(\beta =0.6\) is by far not the strongest possible variation.

  4. We also present tables with all results in the supplementary material. Table 1 in the supplementary material shows the simulation results for the three-dimensional non-simplified vine copulas for the scenario case (a), while Table 2 in the supplementary material presents the results for case (b), where the lowest mean of the out-of-sample KL divergence measure for the nonparametric estimations “Test” is set in bold.

  5. Table 5 in the supplementary material contains the simulation results for the three-dimensional normal distribution and five-dimensional normal distribution, where the lowest mean of the out-of-sample KL divergence measure for the nonparametric estimations “Test” is set in bold.

  6. Table 3 in the supplementary material shows the simulation results for the five-dimensional non-simplified vine copulas for the scenario case (a), while Table 4 in the supplementary material presents the results for case (b), where the lowest mean of the out-of-sample KL divergence measure for the nonparametric estimations “Test” is set in bold.

  7. https://archive.ics.uci.edu/ml/datasets/EEG+Eye+State.

  8. For \(l=2\), we get \(\mathcal{I}_2 = \{2,4\}, \tau (2)=(0,0.25,0.5,0.75,1)\), and

    $$\begin{aligned} {B}^{(\tau (2))}_{\mathcal {I}_2}(\mathbf {u}_j)&= B^{(\tau (2))}(\mathbf {u}_j) \begin{pmatrix}0 &{} 0 &{} 0 &{} 1 &{} 0 \\ 0 &{} 1 &{} 0 &{} 0 &{} 0 \end{pmatrix}^\top \\&= \begin{pmatrix} \phi _{0.25}(u_{j,1}) &{} \phi _{0.75}(u_{j,1})\\ \vdots &{} \vdots \\ \phi _{0.25}(u_{j,n}) &{} \phi _{0.75}(u_{j,n}) \end{pmatrix}. \end{aligned}$$

References

  • Aas, K., Berg, D.: Models for construction of multivariate dependence: a comparison study. Eur. J. Finance 15(7–8), 639–659 (2009)

    Article  Google Scholar 

  • Aas, K., Czado, C., Frigessi, A., Bakken, H.: Pair-copula constructions of multiple dependence. Insur. Math. Econ. 44(2), 182–198 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Abegaz, F., Gijbels, I., Veraverbeke, N.: Semiparametric estimation of conditional copulas. J. Multivar. Anal. 110, 43–73 (2012). Special Issue on Copula Modeling and Dependence

    Article  MathSciNet  MATH  Google Scholar 

  • Acar, E.F., Craiu, R.V., Yao, F.: Dependence calibration in conditional copulas: a nonparametric approach. Biometrics 67(2), 445–453 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Acar, E.F., Genest, C., Nešlehová, J.: Beyond simplified pair-copula constructions. J. Multivar. Anal. 110, 74–90 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Bedford, T., Cooke, R.M.: Vines: a new graphical model for dependent random variables. Ann. Stat. 30(4), 1031–1068 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  • Brechmann, E.C., Czado, C., Aas, K.: Truncated regular vines in high dimensions with application to financial data. Can. J Stat. 40(1), 68–85 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Dennis Cook, R., Johnson, M.E.: Generalized burr-pareto-logistic distributions with applications to a uranium exploration data set. Technometrics 28(2), 123–131 (1986)

    Article  MathSciNet  Google Scholar 

  • Dißmann, J., Brechmann, E., Czado, C., Kurowicka, D.: Selecting and estimating regular vine copulae and application to financial returns. Comput. Statist. Data Anal. 59, 52–69 (2013)

    Article  MathSciNet  Google Scholar 

  • Duong, T.: ks: Kernel Smoothing. R package version 1.10.4 (2016)

  • Efron, B.: Selection criteria for scatterplot smoothers. Ann. Statist. 29, 470–504 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  • Eilers, P.H.C., Marx, B.D.: Flexible smoothing with B-splines and penalties. Statist. Sci. 11(2), 89–121 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  • Fermanian, J.-D., Wegkamp, M.H.: Time-dependent copulas. J. Multivar. Anal. 110, 19–29 (2012). Special Issue on Copula Modeling and Dependence

    Article  MathSciNet  MATH  Google Scholar 

  • Fischer, M., Köck, C., Schlüter, S., Weigert, F.: An empirical analysis of multivariate copula models. Quant. Finance 9(7), 839–854 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Forsey, D.R., Bartels, R.H.: Hierarchical B-spline refinement. In SIGGRAPH ’88: Proceedings of the 15th annual conference on Computer graphics and interactive techniques, New York, NY, pp. 205–212. ACM (1988)

  • Forsey, D.R., Bartels, R.H.: Surface fitting with hierarchical splines. ACM Trans. Graph. 14(2), 134–161 (1995)

    Article  Google Scholar 

  • Gijbels, I., Omelka, M., Veraverbeke, N.: Multivariate and functional covariates and conditional copulas. Electron. J. Stat. 6, 1273–1306 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Gijbels, I., Veraverbeke, N., Omelka, M.: Conditional copulas, association measures and their applications. Comput. Stat. Data Anal. 55(5), 1919–1932 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Hall, P., Yao, Q.: Approximating conditional distribution functions using dimension reduction. Ann. Statist. 33(3), 1404–1421 (2005, 06)

  • Hobæk Haff, I., Aas, K., Frigessi, A.: On the simplified pair-copula construction: simply useful or too simplistic? J. Multivar. Anal. 101(5), 1296–1310 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Hurvich, C.M., Tsai, C.-L.: Regression and time series model selection in small samples. Biometrika 76(2), 297–307 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  • Joe, H.: Dependence modeling with copulas. CRC Press, Boca Raton (2015)

    MATH  Google Scholar 

  • Kauermann, G., Schellhase, C.: Flexible pair-copula estimation in D-vines with penalized splines. Stat. Comput. 24(6), 1081–1100 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  • Kauermann, G., Schellhase, C., Ruppert, D.: Flexible copula density estimation with penalized hierarchical B-splines. Scand. J. Stat. 40(4), 685–705 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Killiches, M., Kraus, D., Czado, C.: Examination and visualisation of the simplifying assumption for vine copulas in three dimensions. ArXiv e-prints (2016, February)

  • Kurowicka, D., Joe, H. (eds.): Dependence modeling. World Scientific, Singapore (2011)

    Google Scholar 

  • Kurz, M.S.: pacotest: Testing for Partial Copulas and the Simplifying Assumption in Vine Copulas, R package version 0.2 (2017)

  • Kurz, M.S., Spanhel, F.: Testing the simplifying assumption in high-dimensional vine copulas. In preparation (2016)

  • Lopez-Paz, D., Hernandez-Lobato, J., Ghahramani, Z.: Gaussian process vine copulas for multivariate dependence. In Proceedings of the 30th International Conference on Machine Learning, W&CP 28(2), pp. 10–18. JMLR (2013)

  • Marx, B., Eilers, P.H.C.: Multidimensional penalized signal regression. Technometrics 47, 13–22 (2005)

    Article  MathSciNet  Google Scholar 

  • Nagler, T., Czado, C.: Evading the curse of dimensionality in nonparametric density estimation with simplified vine copulas. J. Multivar. Anal. 151, 69–89 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Nelsen, R.: An introduction to copulas, 2nd edn. Springer, Berlin (2006)

    MATH  Google Scholar 

  • Patton, A.J.: Modelling asymmetric exchange rate dependence. Int. Econ. Rev. 47(2), 527–556 (2006)

  • Rue, H., Martino, S., Chopin, N.: Approximate Bayesian inference for latent Gaussian models by using integrated nested Laplace approximations. J. Roy. Stat. Soc. B 71(2), 319–392 (2009)

  • Ruppert, D., Wand, M., Carroll, R.: Semiparametric Regression. Cambridge University Press, Cambridge (2003)

    Book  MATH  Google Scholar 

  • Schall, R.: Estimation in generalized linear models with random effects. Biometrika 78(4), 719–727 (1991)

    Article  MATH  Google Scholar 

  • Schepsmeier, U., Stoeber, J., Brechmann, E.C., Graeler, B., Nagler, T., Erhardt, T.: VineCopula: Statistical Inference of Vine Copulas. R package version 2.0.5 (2016)

  • Scott, D.W.: The Curse of Dimensionality and Dimension Reduction. Wiley, Hoboken (2008)

    Google Scholar 

  • Shi, X.: A nondegenerate vuong test. Quant. Econ. 6(1), 85–121 (2015)

    Article  MathSciNet  Google Scholar 

  • Sklar, A.: Fonctions de répartition à n dimensions et leurs marges. Publ. Inst. Statist. Univ. Paris 8, 229–231 (1959)

    MathSciNet  MATH  Google Scholar 

  • Spanhel, F., Kurz, M.S.: Simplified vine copula models: Approximations based on the simplifying assumption. ArXiv e-prints (2015)

  • Spanhel, F., Kurz, M.S.: The partial copula: properties and associated dependence measures. Stat. Probab. Lett. 119, 76–83 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Stein, M.L.: A comparison of generalized cross validation and modified maximum likelihood for estimating the parameters of a stochastic process. Ann. Stat. 18, 1139–1157 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  • Stöber, J., Joe, H., Czado, C.: Simplified pair copula constructions–Limitations and extensions. J. Multivar. Anal. 119, 101–118 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  • Vatter, T., Chavez-Demoulin, V.: Generalized additive models for conditional dependence structures. J. Multivar. Anal. 141, 147–167 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Veraverbeke, N., Omelka, M., Gijbels, I.: Estimation of a conditional copula and association measures. Scand. J. Stat. 38(4), 766–780 (2011)

  • Wahba, G.: A comparison of GCV and GML for choosing the smoothing parameter in the generalized spline smoothing problem. Ann. Stat. 13, 1378–1402 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  • Wand, M., Jones, M.C.: Kernel smoothing. Chapman and Hall, London (1995)

    Book  MATH  Google Scholar 

  • Wand, M.P., Ormerod, J.T.: On semiparametric regression with O’Sullivan penalised splines. Aust. N. Z. J. Stat. 50(2), 179–198 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Wood, S.: Fast stable restricted maximum likelihood and marginal likelihood estimation of semiparametric generalized linear models. J. Roy. Stat. Soc. B 73(1), 3–36 (2011)

    Article  MathSciNet  Google Scholar 

  • Zenger, C.: Sparse grids. Notes Numer. Fluid Mech. Multidiscip. Des. 31, 241–251 (1991)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Göran Kauermann (LMU Munich) for discussions at the beginning of this project.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian Schellhase.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 1301 KB)

Supplementary material 2 (pdf 96 KB)

Appendices

Appendix 1: Construction of sparse B-spline density basis

In the following, objects signed with superscript \(\tilde{\ }\) and \(^{(d)}\) are associated with hierarchical B-spline basis functions, whereas objects signed with superscript \(\tilde{\ }\) and \(^{(d,D)}\) are linked to sparse B-spline basis functions. In order to transform \(B^{(\tau (d))}(\mathbf {u}_j)\) in (6) into its hierarchical representation (see Forsey and Bartels 1988, 1995), we define the hierarchical index sets \(\mathcal {I}_0=\{1, 2\}\) and \(\mathcal {I}_l = \{2j\mid j\in \mathbb {N},1\le j \le 2^{l-1}\}\), \(l=1,\dots ,d\). Let \(B^{(\tau (l))}_{\mathcal {I}_l}(\mathbf {u}_j)\) denote the columns \(\mathcal {I}_l\) of \(B^{\tau (l)}(\mathbf {u}_j)\).Footnote 8

Fig. 2
figure 2

a B-spline density basis \(B^{(\tau (2))}(\mathbf {u}_j)\) and corresponding building blocks of the univariate hierarchical B-spline density basis \(B^{(\tau (0))}_{\mathcal {I}_0}(\mathbf {u}_j), B^{(\tau (1))}_{\mathcal {I}_1}(\mathbf {u}_j)\) and \(B^{(\tau (2))}_{\mathcal {I}_2}(\mathbf {u}_j)\) as graphics (b), (c) and (d)

The univariate hierarchical B-spline basis of degree d is defined as

$$\begin{aligned} \tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j) =\left( {B}^{(\tau (0))}_{\mathcal {I}_0}(\mathbf {u}_j), {{B}^{(\tau (1))}_{\mathcal {I}_1}(\mathbf {u}_j)}, \dots , \ {B}^{(\tau (d))}_{\mathcal {I}_d}(\mathbf {u}_j)\right) . \end{aligned}$$
(22)

Figure 2 presents the univariate B-spline basis \(B^{(\tau (d))}(\mathbf {u}_j)\) and the building parts of the corresponding hierarchical basis \(\tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j)\) for \(d=2\). Let \(\tilde{e}\) be a K-dimensional row vector such that its kth element is given by \(\tilde{e}_k:= \min \{l=0,\dots ,d: k\le |\tau (l)|\}\), e.g., \(\tilde{e}= (0,0,1,2,2)\) if \(d=2\). The vector \(\tilde{e}\) denotes the hierarchical level of \(\tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j)\), and its kth element identifies the hierarchical level of the kth column of \(\tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j)\). By construction, \(B^{(\tau (d))}(\mathbf {u}_j)\) and \(\tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j)\) have full rank, i.e., \(B^{(\tau (d))}(\mathbf {u}_j)\tilde{A}=\tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j)\) for some invertible \(K \times K\) matrix \(\tilde{A}\), so that both univariate bases \(\tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j)\) and \(B^{(\tau (d))}(\mathbf {u}_j)\) span the same space. The three-dimensional hierarchical B-spline basis follows as

$$\begin{aligned} \tilde{\varvec{\varPhi }}^{(d)}(\mathbf {u}_1,\mathbf {u}_2,\mathbf {u}_3) :=\bigotimes _{j=1}^3 \tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j) = \bigotimes _{j=1}^3 B^{(\tau (d))}(\mathbf {u}_j) \tilde{A} \end{aligned}$$
(23)

and the corresponding approximation of the conditional copula density is given by

$$\begin{aligned} c_{12|3} (\mathbf {u}_1,\mathbf {u}_2|\mathbf {u}_3; \tilde{\mathbf {b}}^{(d)})=\tilde{\varvec{\varPhi }}^{(d)}(\mathbf {u}_1, \mathbf {u}_2,\mathbf {u}_3)\tilde{\mathbf {b}}^{(d)}, \end{aligned}$$

where \(\tilde{\mathbf {b}}^{(d)} = (\bigotimes _{j=1}^3\tilde{A})^{-1}\mathbf {b} \).

Fig. 3
figure 3

\({\tilde{\varvec{\Phi }}}^{(2,4)}(u_1,u_2)\) is the full tensor product of two univariate B-spline bases, each margin consists of \({B}^{(\tau (0))}_{\mathcal {I}_0}(\mathbf {u}_j)\), \({B}^{(\tau (1))}_{\mathcal {I}_1}(\mathbf {u}_j)\) and \({B}^{(\tau (2))}_{\mathcal {I}_2}(\mathbf {u}_j)\) for \(j=1,2\) The construction principle of the sparse B-spline basis \({\tilde{\varvec{\Phi }}}^{(2,2)}(u_1,u_2)\) is to remove columns from the full tensor product, reducing the number of spline bases from 25 to 17 in this bivariate example for \(d=2\)

Fig. 4
figure 4

(left) Full tensor product \({\mathbf {\Phi _5}}(u_1,u_2)\) consists of \(5^2=25\) basis functions located at each dot. (right) \({\tilde{\varvec{\Phi }}}^{(2,2)}(u_1,u_2)\) consists of 17 basis functions located at each dot

To overcome the exponential increase in the number of spline coefficients, we use a three-dimensional sparse B-spline basis which reduces the dimension by deleting the columns from the full tensor product basis whose cumulated hierarchy level exceeds D, where \(d\le D\le 3d\). The cumulated hierarchy level of the full tensor product basis is defined as follows. For \(\alpha \in \mathbb {R}^{1\times n}\) and \(\beta \in \mathbb {R}^{1\times q}\), define the lth element of \((\alpha \oplus \beta )\in \mathbb {R}^{1\times nq}\) by \((\alpha \oplus \beta )_{l} = \alpha _{\left\lceil \frac{l}{q}\right\rceil }+ \beta _{l-q(\left\lceil \frac{l}{q}\right\rceil -1)}\), where \(\lceil \cdot \rceil \) is the ceil function. Note that the operation \(\oplus \) is associative. Recall that the kth element of \(\tilde{e}\) identifies the hierarchy level of the kth column of \(\tilde{\mathbf {B}}^{(\tau (d))}(\mathbf {u}_j)\), so that the lth element of \(\epsilon = \tilde{e}\oplus \tilde{e}\oplus \tilde{e}\in \mathbb {R}^{1\times {K}^3}\) denotes the cumulated hierarchy level of the lth column of the hierarchical B-spline basis \(\tilde{\varvec{\mathbf {\Phi }}}^{(d)}(\mathbf {u}_1,\mathbf {u}_2,\mathbf {u}_3)\). Define \(\mathcal{O}_D = \{j=1,\ldots , {K_{}}^3 :\epsilon _j\le D \}\), i.e., \(\mathcal{O}_D\) contains the position of the columns of \(\tilde{\varvec{\mathbf {\Phi }}}^{(d)}(\mathbf {u}_1,\mathbf {u}_2,\mathbf {u}_3)\) whose cumulated hierarchy level does not exceed D. Let \(\mathcal{O}_D(j)\) be the jth smallest element of \(\mathcal{O}_D\) and define the orthogonal matrix \(\mathcal {E}(\mathcal {O}_D)\in \mathbb {R}^{{K}^3\times |\mathcal{O}_D|}\) such that its \((\mathcal {O}_D(j),j)\)-entry is one for \(j=1,\dots ,|{\mathcal{O}}_D|,\) and the other entries are zero. The three-dimensional sparse B-spline basis follows as

$$\begin{aligned} \tilde{\varvec{{\Phi }}}^{(d,D)}(\mathbf {u}_1,\mathbf {u}_2,\mathbf {u}_3) = \left[ \bigotimes _{j=1}^3 \tilde{\mathbf {B}}^{(\tau (d))} (\mathbf {u}_j)\right] \mathcal {E}(\mathcal{O}_D), \end{aligned}$$
(24)
Fig. 5
figure 5

Linear and quadratic functions of Kendell’s \(\tau \) for the construction of the conditional copula in higher trees

Fig. 6
figure 6

Box plots of the differences \(KL_{non-par}-KL_{par}\) for three-dimensional non-simplified vine copulas. \(KL_{non-par}\) is the out-of-sample KL divergence of a nonparametric estimation (“SimpA,” “Cond” or “Test”), and \(KL_{par}\) is the out-of-sample KL divergence of the parametric estimation using the VineCopula package. Blue refers to the estimator “Test,” pink corresponds to the estimator “Cond,” and green refers to the estimator “SimpA.” The whiskers cover 95% of the data

where the lower index d is the degree of the univariate hierarchical B-spline basis and the upper index \(D, d\le D\le 3d,\) refers to the maximum cumulated hierarchy level. Only the columns in the hierarchical B-spline basis \({\tilde{\varvec{\Phi }}}^{(d)}(\mathbf {u}_1,\mathbf {u}_2,\mathbf {u}_3)\) whose cumulated hierarchy level does not exceed D constitute the three-dimensional sparse B-spline basis \(\tilde{\varvec{{\Phi }}}^{(d,D)}(\mathbf {u}_1,\mathbf {u}_2,\mathbf {u}_3)\). Figure 3 shows the construction principle for a bivariate sparse B-spline basis with hierarchy level \(d=2\). Figure 4 presents the placements of the knots for the full tensor product of B-splines and the sparse B-spline basis for \(d=2\). The corresponding spline coefficients are given by \(\tilde{\mathbf {b}}^{(d,D)} = \mathcal{E(O_D)}^\top \tilde{\mathbf {b}}^{(d)}\).

Fig. 7
figure 7

Box plots of the differences \(KL_{non-par}-KL_{par}\) for five-dimensional non-simplified vine copulas. \(KL_{non-par}\) is the out-of-sample KL divergence of a nonparametric estimation (“SimpA,” “Cond” or “Test”), and \(KL_{par}\) is the out-of-sample KL divergence of the parametric estimation using the VineCopula package. Blue refers to the estimator “Test,” pink corresponds to the estimator “Cond,” and green refers to the estimator “SimpA.” The whiskers cover 95% of the data

Fig. 8
figure 8

Box plots of the differences \(KL_{non-par}-KL_{par}\) for three- and five-dimensional mixtures of normal distributions. \(KL_{non-par}\) is the out-of-sample KL divergence of a nonparametric estimation (“SimpA,” “Cond” or “Test”), and \(KL_{par}\) is the out-of-sample KL divergence of the parametric estimation using the VineCopula package. Blue refers to the estimator “Test,” pink corresponds to the estimator “Cond,” and green refers to the estimator “SimpA.” The whiskers cover 95% of the data

Fig. 9
figure 9

Conditional copula densities of the three-dimensional mixture of normal distributions for conditional arguments \(0.01, 0.15, 0.29, 0.43, 0.57, 0.71, 0.85, 0.99\) (top left to bottom right)

Appendix 2: Marginal likelihood

The prior (13) is degenerated, which needs to be corrected as follows. For simplicity, we write \(\mathbf {b}:= \tilde{\mathbf {b}^{(d,D)}}\) in this section. We decompose \(\mathbf {b}\) into the two components \(\mathbf {b}_{\sim }\) and \(\mathbf {b}_{\bot }\), respectively, such that \(\mathbf {b}_{\sim }\) is a normally distributed random vector with non-degenerated variance and \(\mathbf {b}_{\bot }\) contains the remaining components treated as parameters, see also Wand and Ormerod (2008). Applying a singular value decomposition, we have

$$\begin{aligned} \tilde{\mathbf {P}^{(D)}}=\tilde{\mathbf {U}} \tilde{\mathbf {\Lambda }} \tilde{\mathbf {U}}^T \end{aligned}$$

where \(\tilde{\mathbf {\Lambda }}\) is a diagonal matrix with positive eigenvalues and \(\tilde{\mathbf {U}} \in \mathbb {R}^{(K+1) \times h}\) are the corresponding eigenvectors with \(K+1\) being the number of elements in \(\mathbf {b}\) and \(h=K+1-4\) being the rank of \(\tilde{\mathbf {P}^{(D)}}\). Extending \(\tilde{\mathbf {U}}\) to an orthogonal basis by \(\check{\mathbf {U}}\) gives \(\mathbf {b}_{\sim }=\tilde{\mathbf {U}}^T\mathbf {b}\). With the a priori assumption \(\mathbf {b}_{\sim } \sim N(0, \lambda ^{-1}\tilde{\mathbf {\Lambda }}^{-1})\) and \(\mathbf {U}=(\tilde{\mathbf {U}}, \check{\mathbf {U}})\) as orthogonal basis, we get \(\mathbf {b}_{\bot }={\check{\mathbf {U}}}^T \mathbf {b}\). Conditioning on \(\mathbf {b}_{\sim }\), we get the mixed model log-likelihood

$$\begin{aligned} l_m(\lambda , \mathbf {b}^{\bot })=\log \int |\lambda \tilde{\mathbf {\Lambda }}|^\frac{1}{2} \exp \left\{ l^{(D)}_p(\mathbf {b}, \lambda )\right\} \text{ d }{\mathbf {b}_{\sim }}. \end{aligned}$$

The integral can be approximated by a Laplace approximation (see also Rue et al. 2009)

$$\begin{aligned} l_m(\lambda ,\mathbf {b}^{\bot })\approx & {} \frac{1}{2} \log |\lambda \tilde{\mathbf {\Lambda }}|+ l_p^{(D)} (\hat{\mathbf {b}}, \lambda )\nonumber \\&-\frac{1}{2} \log |\tilde{\mathbf {U}}^T {\mathbf H}_{pen}^{(d,D)} (\hat{\mathbf {b}},\lambda ) \tilde{\mathbf {U}}| \end{aligned}$$
(25)

where \(\hat{\mathbf {b}}\) denotes the penalized maximum likelihood estimate. We can now differentiate (25) with respect to \(\lambda \) which gives

$$\begin{aligned}&\frac{\partial l_m(\lambda ,\hat{\mathbf {b}}^{\bot })}{\partial \lambda } = - \frac{1}{2} \hat{\mathbf {b}}^T {\tilde{\mathbf {P}}}^{(D)} (\mathbf \lambda ) \hat{\mathbf {b}}^{\bot } + \frac{1}{2 \lambda } \nonumber \\&\quad \cdot \text{ tr } \underbrace{\left\{ \left( \tilde{\mathbf {U}}^\top \mathbf {H}_{pen}^{(d,D)}(\hat{\mathbf {b}},\lambda =0) \tilde{\mathbf {U}} + \lambda \tilde{\mathbf {\Lambda }}\right) ^{-1} \tilde{\mathbf {U}}^\top \mathbf {H}_{pen}^{(d,D)}(\hat{\mathbf {b}},\lambda =0) \tilde{\mathbf {U}} \right\} }_{:=S(\lambda )}.\nonumber \\ \end{aligned}$$
(26)

and \({\mathbf H}_{pen}^{(d,D)}(\hat{\mathbf {b}},\lambda )\) denotes the second-order partial derivative of (10) with respect to \(\hat{\mathbf {b}}\), i.e.,

$$\begin{aligned}&{\mathbf {H}}_{pen}^{(d,D)}(\hat{\mathbf {b}},\mathbf {\lambda })=\\&\quad -\sum _{i=1}^{n} \frac{{\tilde{\varvec{\Phi }}}^{(d,D)} (u_{1,i},u_{2,i},u_{3,i}) {\varvec{\tilde{\varPhi }}^{(d,D)}}^\top (u_{1,i},u_{2,i},u_{3,i})}{c_{12|3}(u_{1,i},u_{2,i}|u_{3,i};\hat{\mathbf {b}})}\\&\quad - \tilde{\mathbf {{{P}}}}^{(D)}(\mathbf \lambda ). \end{aligned}$$
Fig. 10
figure 10

ROC curve for classification of EEG Eye State Data Set

Table 3 Mean elapsed computing time in seconds (standard deviation) for (i) bivariate unconditional copula densities, (ii) bivariate conditional copula densities with one conditioning argument and (iii) the estimation of the whole simplified and non-simplified vine copulas with “Test” estimator (measured for \(N=10\) runs)

Appendix 3: Computing time

We present some computing times, defined as elapsed time on the system, measured by R on a machine with an Intel Core i7-2600 CPU @3.40 Ghz \(\times \) 4 using R.3.3.1 on Linux Mint 17.2 (64-bit). To this end, we take the first ten data sets of the three-dimensional non-simplified vine copulas and the five-dimensional non-simplified vine copulas from our simulation study, each for case (b) with \(\beta =0.6\) for both sample sizes \(n=500\) and \(n=2000\). Computing time is measured for (i) bivariate unconditional copula densities \(c_{12}\) for the first two marginal arguments from the selected data, (ii) bivariate conditional copula densities \(c_{12|3}\) with one conditioning argument and (iii) the estimation of the non-simplified vine copula. The computing times are realized with one starting value of \(\lambda \). Choosing three starting values for each copula estimation as in the simulation study increases the computing time linearly. See Table 3 for the results. The five-dimensional non-simplified vine copulas are computed in approximately 3–7 min, depending on the basis size and sample size, while the computing time for a three-dimensional non-simplified vine copula is less than 1 min. It might seem counterintuitive that the computing time decreases if the sample size increases from \(n=500\) to \(n=2000\) observations. That is because less iteration steps in the quadratic programming are required.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schellhase, C., Spanhel, F. Estimating non-simplified vine copulas using penalized splines. Stat Comput 28, 387–409 (2018). https://doi.org/10.1007/s11222-017-9737-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11222-017-9737-7

Keywords

Navigation