Skip to main content

On ANOVA Decompositions of Kernels and Gaussian Random Field Paths

  • Conference paper
  • First Online:
Monte Carlo and Quasi-Monte Carlo Methods

Part of the book series: Springer Proceedings in Mathematics & Statistics ((PROMS,volume 163))

  • 2656 Accesses

Abstract

The FANOVA (or “Sobol’-Hoeffding”) decomposition of multivariate functions has been used for high-dimensional model representation and global sensitivity analysis. When the objective function f has no simple analytic form and is costly to evaluate, computing FANOVA terms may be unaffordable due to numerical integration costs. Several approximate approaches relying on Gaussian random field (GRF) models have been proposed to alleviate these costs, where f is substituted by a (kriging) predictor or by conditional simulations. Here we focus on FANOVA decompositions of GRF sample paths, and we notably introduce an associated kernel decomposition into \(4^{d}\) terms called KANOVA. An interpretation in terms of tensor product projections is obtained, and it is shown that projected kernels control both the sparsity of GRF sample paths and the dependence structure between FANOVA effects. Applications on simulated data show the relevance of the approach for designing new classes of covariance kernels dedicated to high-dimensional kriging.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 169.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Adler, R., Taylor, J.: Random Fields and Geometry. Springer, Boston (2007)

    MATH  Google Scholar 

  2. Antoniadis, A.: Analysis of variance on function spaces. Statistics 15, 59–71 (1984)

    MathSciNet  MATH  Google Scholar 

  3. Berlinet, A., Thomas-Agnan, C.: Reproducing Kernel Hilbert Spaces in Probability and Statistics. Kluwer Academic Publishers, Boston (2004)

    Google Scholar 

  4. Chastaing, G., Le Gratiet, L.: ANOVA decomposition of conditional Gaussian processes for sensitivity analysis with dependent inputs. J. Stat. Comput. Simul. 85(11), 2164–2186 (2015)

    Article  MathSciNet  Google Scholar 

  5. Durrande, N., Ginsbourger, D., Roustant, O.: Additive covariance kernels for high-dimensional Gaussian process modeling. Ann. Fac. Sci. Toulous. Math. 21, 481–499 (2012)

    Google Scholar 

  6. Durrande, N., Ginsbourger, D., Roustant, O., Carraro, L.: ANOVA kernels and RKHS of zero mean functions for model-based sensitivity analysis. J. Multivar. Anal. 115, 57–67 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  7. Dupuy, D., Helbert, C., Franco, J.: DiceDesign and DiceEval: Two R packages for design and analysis of computer experiments. J. Stat. Softw. 65(11): 1–38 (2015)

    Google Scholar 

  8. Duvenaud, D.: Automatic model construction with Gaussian processes. Ph.D. thesis, Department of Engineering, University of Cambridge (2014)

    Google Scholar 

  9. Duvenaud, D., Nickisch, H., Rasmussen, C.: Additive Gaussian Processes. NIPS conference. (2011)

    Google Scholar 

  10. Efron, B., Stein, C.: The jackknife estimate of variance. Ann. Stat. 9, 586–596 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  11. Franco, J., Dupuy, D., Roustant, O., Damblin, G., Iooss, B.: DiceDesign: Designs of computer experiments. R package version 1.7 (2015)

    Google Scholar 

  12. Gikhman, I.I., Skorokhod, A.V.: The theory of stochastic processes. Springer, Berlin (2004). Translated from the Russian by S. Kotz, Reprint of the 1974 edition

    Google Scholar 

  13. Hoeffding, W.: A class of statistics with asymptotically normal distributions. Ann. Math. Stat. 19, 293–325 (1948)

    Article  MathSciNet  MATH  Google Scholar 

  14. Jan, B., Bect, J., Vazquez, E., Lefranc, P.: approche bayésienne pour l’estimation d’indices de Sobol. In 45èmes Journées de Statistique - JdS 2013. Toulouse, France (2013)

    Google Scholar 

  15. Janon, A., Klein, T., Lagnoux, A., Nodet, M., Prieur, C.: Asymptotic Normality and Efficiency of Two Sobol Index Estimators. Probability And Statistics, ESAIM (2013)

    MATH  Google Scholar 

  16. Kaufman, C., Sain, S.: Bayesian functional ANOVA modeling using Gaussian process prior distributions. Bayesian Anal. 5, 123–150 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  17. Krée, P.: Produits tensoriels complétés d’espaces de Hilbert. Séminaire Paul Krée Vol 1, No. 7 (1974–1975)

    Google Scholar 

  18. Kuelbs, J.: Expansions of vectors in a Banach space related to Gaussian measures. Proc. Am. Math. Soc. 27(2), 364–370 (1971)

    MathSciNet  MATH  Google Scholar 

  19. Kuo, F.Y., Sloan, I.H., Wasilkowski, G.W., Wozniakowski, H.: On decompositions of multivariate functions. Math. Comput. 79, 953–966 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  20. Le Gratiet, L., Cannamela, C., Iooss, B.: A Bayesian approach for global sensitivity analysis of (multi-fidelity) computer codes. SIAM/ASA J. Uncertain. Quantif. 2(1), 336–363 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  21. Lenz, N.: Additivity and ortho-additivity in Gaussian random fields. Master’s thesis, Departement of Mathematics and Statistics, University of Bern (2013). http://hal.archives-ouvertes.fr/hal-01063741

  22. Marrel, A., Iooss, B., Laurent, B., Roustant, O.: Calculations of Sobol indices for the Gaussian process metamodel. Reliab. Eng. Syst. Saf. 94, 742–751 (2009)

    Article  Google Scholar 

  23. Muehlenstaedt, T., Roustant, O., Carraro, L., Kuhnt, S.: Data-driven Kriging models based on FANOVA-decomposition. Stat. Comput. 22(3), 723–738 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  24. Oakley, J., O’Hagan, A.: Probabilistic sensitivity analysis of complex models: a Bayesian approach. J. R. Stat. Soc. 66, 751–769 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  25. Rajput, B.S., Cambanis, S.: Gaussian processes and Gaussian measures. Ann. Math. Stat. 43, 1944–1952 (1972)

    Article  MathSciNet  MATH  Google Scholar 

  26. Rasmussen, C.R., Williams, C.K.I.: Gaussian Processes for Machine Learning. Cambridge, MIT Press (2006)

    Google Scholar 

  27. Saltelli, A., Ratto, M., Andres, T., Campolongo, F., Cariboni, J., Gatelli, D., Saisana, M., Tarantola, S.: Global sensitivity analysis: the primer. Wiley Online Library (2008)

    Google Scholar 

  28. Santner, T., Williams, B., Notz, W.: The design and analysis of computer experiments. Springer, New York (2003)

    Google Scholar 

  29. Sawa, T.: The exact moments of the least squares estimator for the autoregressive model. J. Econom. 8(2), 159–172 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  30. Scheuerer, M.: A comparison of models and methods for spatial interpolation in statistics and numerical analysis. Ph.D. thesis, Georg-August-Universität Göttingen (2009)

    Google Scholar 

  31. Schuhmacher, D.: Distance estimates for poisson process approximations of dependent thinnings. Electron. J. Probab. 10(5), 165–201 (2005)

    Google Scholar 

  32. Sobol’, I.: Multidimensional Quadrature Formulas and Haar Functions. Nauka, Moscow (1969). (In Russian)

    MATH  Google Scholar 

  33. Sobol’, I.: Global sensitivity indices for nonlinear mathematical models and their Monte Carlo estimates. Math. Comput. Simul. 55(1–3), 271–280 (2001)

    Google Scholar 

  34. Steinwart, I., Scovel, C.: Mercer’s theorem on general domains: on the interaction between measures, kernels, and RKHSs. Constr. Approx. 35(3), 363–417 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  35. Talagrand, M.: Regularity of Gaussian processes. Acta Math. 159(1–2), 99–149 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  36. Tarieladze, V., Vakhania, N.: Disintegration of Gaussian measures and average-case optimal algorithms. J. Complex. 23(4–6), 851–866 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  37. Touzani, S.: Response surface methods based on analysis of variance expansion for sensitivity analysis. Ph.D. thesis, Université de Grenoble (2011)

    Google Scholar 

  38. Vapnik, V.: Statistical Learning Theory. Wiley, New York (1998)

    MATH  Google Scholar 

  39. Wahba, G.: Spline Models for Observational Data. Siam, Philadelphia (1990)

    Google Scholar 

  40. Welch, W.J., Buck, R.J., Sacks, J., Wynn, H.P., Mitchell, T.J., Morris, M.D.: Screening, predicting, and computer experiments. Technometrics 34, 15–25 (1992)

    Article  Google Scholar 

Download references

Acknowledgments

The authors would like to thank Dario Azzimonti for proofreading, as well as the editors and an anonymous referee for their valuable comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to David Ginsbourger .

Editor information

Editors and Affiliations

Appendices

Proofs

Theorem 1 (a) The first part and the concrete solution (6) follow directly from the corresponding statements in Sect. 2. Having established (6), it is easily seen that \([T_{{\mathfrak {u}}}\otimes T_{{\mathfrak {v}}}]k =T^{(1)}_{{\mathfrak {u}}} T^{(2)}_{{\mathfrak {v}}} k\) coincides with \(k_{{\mathfrak {u}},{\mathfrak {v}}}\).

(b) Under these conditions Mercer’s theorem applies (see [34] for an overview and recent extensions). So there exist a non-negative sequence \((\lambda _{i})_{i \in \mathbb {N}\backslash \{0\}}\), and continuous representatives \((\phi _{i})_{i \in \mathbb {N}\backslash \{0\}}\) of an orthonormal basis of \(\mathrm {L}^2(\nu )\) such that \( k(\mathbf {x}, \mathbf {y})=\sum _{i=1}^{\infty } \lambda _{i} \phi _{i}(\mathbf {x}) \phi _{i}(\mathbf {y}) \), \(\mathbf {x}, \mathbf {y} \in D\), where the convergence is absolute and uniform. Noting that \(T_{{\mathfrak {u}}}, T_{{\mathfrak {v}}}\) are also bounded as operators on continuous functions, applying \(T^{(1)}_{{\mathfrak {u}}} T^{(2)}_{{\mathfrak {v}}}\) from above yields that

$$\begin{aligned} \sum _{{\mathfrak {u}}\subseteq I} \sum _{{\mathfrak {v}}\subseteq I} \alpha _{{\mathfrak {u}}} \alpha _{{\mathfrak {v}}} k_{{\mathfrak {u}}, {\mathfrak {v}}}(\mathbf {x}, \mathbf {y}) = \sum _{i=1}^{\infty } \lambda _{i} \psi _{i}(\mathbf {x}) \psi _{i} (\mathbf {y}), \end{aligned}$$
(17)

where \(\psi _{i}= \sum _{{\mathfrak {u}}\subseteq I} \alpha _{{\mathfrak {u}}} (T_{{\mathfrak {u}}}\phi _{i})\). Thus the considered function is indeed s.p.d. \(\square \)

Corollary 1 Expand the product \(\prod _{l=1}^{d}(1+k_{l}^{(0)}(x_{l},y_{l}))\) and conclude by uniqueness of the KANOVA decomposition, noting that \(\int \prod _{l\in {\mathfrak {u}}}k_{l}^{(0)}(x_{l},y_{l})\nu _i(\mathrm {d}x_i) = \int \prod _{l\in {\mathfrak {u}}}k_{l}^{(0)}(x_{l},y_{l})\nu _j(\mathrm {d}y_j) = 0\) for any \({\mathfrak {u}}\subseteq I\) and any \(i,j \in {\mathfrak {u}}\). \(\square \)

Theorem 2 Sample path continuity implies product-measurability of Z and \(Z^{({\mathfrak {u}})}\), respectively, as can be shown by an approximation argument; see e.g. Prop. A.D. in [31]. Due to Theorem 3 in [35], the covariance kernel k is continuous, hence \(\int _{D} {\mathbb {E}}|Z_{\mathbf {x}}| \, \nu _{-{\mathfrak {u}}}(\mathrm {d}\mathbf {x_{-{\mathfrak {u}}}}) \le (\int _{D} k(\mathbf {x}, \mathbf {x}) \, \nu _{-{\mathfrak {u}}}(\mathrm {d}\mathbf {x_{-{\mathfrak {u}}}}))^{1/2} < \infty \) for any \({\mathfrak {u}}\subseteq I\) and by Cauchy–Schwarz \(\int _{D} \int _{D} {\mathbb {E}}|Z_{\mathbf {x}}Z_{\mathbf {y}}| \, \nu _{-{\mathfrak {u}}}(\mathrm {d}\mathbf {x_{-{\mathfrak {u}}}}) \nu _{-{\mathfrak {v}}}(\mathrm {d}\mathbf {y_{-{\mathfrak {v}}}}) < \infty \) for any \({\mathfrak {u}},{\mathfrak {v}}\subseteq I\). Replacing f by Z in Formula (2), taking expectations and using Fubini’s theorem yields that \(Z^{({\mathfrak {u}})}\) is centred again. Combining (2), Fubini’s theorem, and (6) yields

$$\begin{aligned} \begin{aligned} {{\mathrm{Cov}}}&(Z^{({\mathfrak {u}})}_{\mathbf {x}}, Z^{({\mathfrak {v}})}_{\mathbf {y}}) \\&= \sum _{{\mathfrak {u}}'\subseteq {\mathfrak {u}}} \sum _{{\mathfrak {v}}'\subseteq {\mathfrak {v}}} (-1)^{\vert {\mathfrak {u}}\vert +\vert {\mathfrak {v}}\vert -\vert {\mathfrak {u}}'\vert -\vert {\mathfrak {v}}'\vert } \overbrace{{{\mathrm{Cov}}}\biggl ( \int Z_{\mathbf {x}} \; \nu _{-{\mathfrak {u}}'}(\mathrm {d}\mathbf {x}_{-{\mathfrak {u}}'}),\, \int Z_{\mathbf {y}} \; \nu _{-{\mathfrak {v}}'}(\mathrm {d}\mathbf {y}_{-{\mathfrak {v}}'}) \biggr )}^{\int {{\mathrm{Cov}}}(Z_{\mathbf {x}},Z_{\mathbf {y}}) \; \nu _{-{\mathfrak {u}}'}(\mathrm {d}\mathbf {x}_{-{\mathfrak {u}}'}) \; \nu _{-{\mathfrak {v}}'}(\mathrm {d}\mathbf {y}_{-{\mathfrak {v}}'}) } \\&= [T_{{\mathfrak {u}}} \otimes T_{{\mathfrak {v}}}] k \, (\mathbf {x}, \mathbf {y}). \\& \end{aligned} \end{aligned}$$
(18)

It remains to show the joint Gaussianity of the \(Z^{({\mathfrak {u}})}\). First note that \(C_b(D,\mathbb {R}^r)\) is a separable Banach space for \(r \in {\mathbb {N}}\setminus \{0\}\). We may and do interprete Z as a random element of \(C_b(D)\), equipped with the \(\sigma \)-algebra \(\mathscr {B}^{D}\) generated by the evaluation maps \([C_b(D) \ni f \mapsto f(\mathbf {x}) \in \mathbb {R}]\). By Theorem 2 in [25] the distribution \({\mathbb {P}}Z^{-1}\) of Z is a Gaussian measure on \(\bigl ( C_b(D),\mathscr {B}(C_b(D)) \bigr )\). Since \(T_{{\mathfrak {u}}}\) is a bounded linear operator \(C_b(D) \rightarrow C_b(D)\), we obtain immediately that the “combined operator” \(\mathfrak {T} :C_b(D) \rightarrow C_b(D,\mathbb {R}^{2^d})\), defined by \((\mathfrak {T}(f))(\mathbf {x}) = (T_{{\mathfrak {u}}}f(\mathbf {x}))_{{\mathfrak {u}}\subseteq I}\), is also bounded and linear. Corollary 3.7 of [36] yields that the image measure \(({\mathbb {P}}Z^{-1}) \mathfrak {T}^{-1}\) is a Gaussian measure on \(C_b(D,\mathbb {R}^{2^d})\). This means that for every bounded linear operator \(\Lambda :C_b(D,\mathbb {R}^{2^d}) \rightarrow \mathbb {R}\) the image measure \((({\mathbb {P}}Z^{-1}) \mathfrak {T}^{-1}) \Lambda ^{-1}\) is a univariate normal distribution, i.e. \(\Lambda (\mathfrak {T}Z)\) is a Gaussian random variable. Thus, for all \(n \in \mathbb {N}\), \(\mathbf {x}^{(i)}\in D\) and \(a_i^{({\mathfrak {u}})} \in \mathbb {R}\), where \(1 \le i \le n\), \(u \subseteq I\), we obtain that \(\sum _{i=1}^n \sum _{{\mathfrak {u}}\subseteq I} a_i^{({\mathfrak {u}})} (T_{{\mathfrak {u}}} Z)_{\mathbf {x}^{(i)}}\) is Gaussian by the fact that \([C_b(D) \ni f \mapsto f(\mathbf {x}) \in \mathbb {R}]\) is continuous (and linear) for every \(\mathbf {x} \in D\). We conclude that \(\mathfrak {T}Z = (Z_{\mathbf {x}}^{({\mathfrak {u}})}, {\mathfrak {u}}\subseteq I)_{\mathbf {x}\in D}\) is a vector-valued GRF. \(\square \)

Corollary 2 (a) If (i) holds, \([T_{{\mathfrak {u}}} \otimes T_{{\mathfrak {u}}}]k= T_{{\mathfrak {u}}}^{(2)} (T_{{\mathfrak {u}}}^{(1)}k)=\mathbf {0}\) by \((T_{{\mathfrak {u}}}^{(1)}k)(\bullet ,\mathbf {y}) = T_{{\mathfrak {u}}}(k(\bullet ,\mathbf {y}))\); thus (ii) holds. (ii) trivially implies (iii). Statement (iii) means that \({{\mathrm{Var}}}(Z^{({\mathfrak {u}})}_{\mathbf {x}}) = 0\), which implies that \(Z^{({\mathfrak {u}})}_{\mathbf {x}} = 0\) a.s., since \(Z^{({\mathfrak {u}})}\) is centred. (iv) follows by noting that \({\mathbb {P}}(Z^{({\mathfrak {u}})}_{\mathbf {x}} = 0)=1\) for all \(\mathbf {x} \in D\) implies \(\mathbb {P}( Z^{({\mathfrak {u}})}=\mathbf {0} ) =1\) by the fact that \(Z^{({\mathfrak {u}})}\) has continuous sample paths and is therefore separable. Finally, (iv) implies (i) because \(T_{{\mathfrak {u}}}(k(\bullet ,\mathbf {y}))={{\mathrm{Cov}}}(Z^{({\mathfrak {u}})}_{{\bullet }},Z_{\mathbf {y}})=\mathbf {0}\); see (18) for the first equality.

(b) For any \(m,n \in \mathbb {N}\) and \(\mathbf {x}_1,\ldots ,\mathbf {x}_m,\mathbf {y}_1,\ldots ,\mathbf {y}_n \in D\) we obtain by Theorem 2 that \(Z^{({\mathfrak {u}})}_{\mathbf {x}_1}, \ldots , Z^{({\mathfrak {u}})}_{\mathbf {x}_m}, Z^{({\mathfrak {v}})}_{\mathbf {y}_1}, \ldots , Z^{({\mathfrak {v}})}_{\mathbf {y}_n}\) are jointly normally distributed. Statement (i) is equivalent to saying that \({{\mathrm{Cov}}}(Z^{({\mathfrak {u}})}_{\mathbf {x}},Z^{({\mathfrak {v}})}_{\mathbf {y}}) = 0\) for all \(\mathbf {x}, \mathbf {y} \in D\). Thus \((Z^{({\mathfrak {u}})}_{\mathbf {x}_1}, \ldots , Z^{({\mathfrak {u}})}_{\mathbf {x}_m})\) and \((Z^{({\mathfrak {v}})}_{\mathbf {y}_1}, \ldots , Z^{({\mathfrak {v}})}_{\mathbf {y}_n})\) are independent. Since the sets

$$\begin{aligned} \{ (f,g) \in \mathbb {R}^D \times \mathbb {R}^D :(f(\mathbf {x}_1),\ldots ,f(\mathbf {x}_m)) \in A, (g(\mathbf {y}_1),\ldots ,g(\mathbf {y}_n)) \in B \} \end{aligned}$$
(19)

with \(m,n \in \mathbb {N}\), \(\mathbf {x}_1,\ldots ,\mathbf {x}_m,\mathbf {y}_1,\ldots ,\mathbf {y}_n \in D\), \(A \in \mathscr {B}(\mathbb {R}^m)\), \(B \in \mathscr {B}(\mathbb {R}^n)\) generate \(\mathscr {B}^D \otimes \mathscr {B}^D\) (and the system of such sets is stable under intersections), statement (ii) follows. The converse direction is straightforward. \(\square \)

Corollary 3 By Remark 2, there is a Gaussian white noise sequence \(\varepsilon =(\varepsilon _{i})_{i\in \mathbb {N}\backslash \{0\}}\) such that \(Z_{\mathbf {x}}=\sum _{i=1}^{\infty } \sqrt{\lambda _{i}} \varepsilon _{i} \phi _{i}(\mathbf {x})\) uniformly with probability 1. From \(Z^{({\mathfrak {u}})}_{\mathbf {x}}=\sum _{i=1}^{\infty } \sqrt{\lambda _{i}} \varepsilon _{i} T_{{\mathfrak {u}}}\phi _{i}(\mathbf {x})\), we obtain \(\Vert Z^{({\mathfrak {u}})} \Vert ^2=Q_{{\mathfrak {u}}}(\varepsilon , \varepsilon )\) with \(Q_{{\mathfrak {u}}}\) as defined in the statement. A similar calculation for the denominator of \(S_{{\mathfrak {u}}}(Z)\) leads to \(\sum _{{\mathfrak {v}}\ne \emptyset } Q_{{\mathfrak {v}}}(\varepsilon , \varepsilon )\).\(\square \)

Additional Examples

Here we give useful expressions to compute the KANOVA decomposition of some tensor product kernels with respect to the uniform measure on \([0,1]^{d}\). For simplicity we denote the 1-dimensional kernels on which they are based by k (corresponding to the notation \(k_i\) in Example 2). The uniform measure on [0, 1] is denoted by \(\lambda \).

Example 5

(Exponential kernel) If \(k(x,y) = \exp \left( - \frac{ \vert x-y \vert }{\theta } \right) \), then:

  • \(\int _0^1 k(., y) d\lambda = \theta \times \left[ 2 - k(0, y) - k(1,y) \right] \)

  • \(\iint _{[0,1]^2} k(.,.) d(\lambda \otimes \lambda ) = 2 \theta (1 - \theta + \theta e^{-1 / \theta } )\)

Example 6

(Matérn kernel, \(\nu =p+\frac{1}{2}\)) Define for \(\nu =p+\frac{1}{2}\) (\(p \in \mathbb {N}\)):

$$ k(x,y) = \frac{p!}{(2p)!} \sum _{i=0}^p \frac{(p+i)!}{i!(p-i)!} \left( \frac{ \vert x - y \vert }{\theta / \sqrt{8\nu }} \right) ^{p-i} \times \exp \left( - \frac{\vert x - y \vert }{\theta / \sqrt{2\nu }} \right) .$$

Then, denoting \(\zeta _p = \frac{\theta }{\sqrt{2\nu }}\), we have:

$$\int _0^1 k(., y) d\lambda = \zeta _p \frac{p!}{(2p)!} \times \left[ 2c_{p,0} - A_p \left( \frac{y}{\zeta _p} \right) - A_p \left( \frac{1 - y}{\zeta _p} \right) \right] ,$$

where \( A_p(u) = \left( \sum _{\ell =0}^p c_{p,\ell } u^\ell \right) e^{-u}\) with \(c_{p,\ell } = \frac{1}{\ell !} \sum _{i=0}^{p-\ell }{\frac{(p+i)!}{i!} 2^{p-i}}.\) This generalizes Example 5, corresponding to \(\nu =1/2\). Also, this result can be written more explicitly for the commonly selected value \(\nu =3/2\) (\(p=1, \zeta _1=\theta / \sqrt{3}\)):

  • \(k(x,y) = \left( 1 + \frac{ \vert x-y \vert }{\zeta _1} \right) \exp \left( - \frac{ \vert x-y \vert }{\zeta _1}\right) \)

  • \(\int _0^1 k(., y) d\lambda = \zeta _1 \times \left[ 4 - A_1 \left( \frac{y}{\zeta _1} \right) - A_1 \left( \frac{1 - y}{\zeta _1} \right) \right] \,\,\) with \(A_1(u) = (2+u)e^{-u}\)

  • \(\iint _{[0,1]^2} k(.,.) d(\lambda \otimes \lambda ) = 2\zeta _1 \left[ 2 - 3 \zeta _1 + (1 + 3 \zeta _1 ) e^{ - 1/\zeta _1 } \right] \)

Similarly, for \(\nu =5/2\) (\(p=2, \zeta _2=\theta / \sqrt{5}\)):

  • \(k(x,y) = \left( 1 + \frac{ \vert x-y \vert }{\zeta _2} + \frac{1}{3} \frac{ (x-y)^2 }{(\zeta _2)^2} \right) \exp \left( - \frac{ \vert x-y \vert }{\zeta _2} \right) \)

  • \(\int _0^1 k(., y) d\lambda = \frac{1}{3} \zeta _2 \times \left[ 16 - A_2 \left( \frac{y}{\zeta _2} \right) - A_2 \left( \frac{1 - y}{\zeta _2} \right) \right] \,\,\) with \(A_2(u) = (8 + 5u + u^2) e^{-u}\)

  • \(\iint _{[0,1]^2} k(.,.) d(\lambda \otimes \lambda ) = \frac{1}{3}\zeta _2 (16 - 30 \,\zeta _2) + \frac{2}{3} (1 + 7 \,\zeta _2 + 15 \, (\zeta _2)^2 ) e^{ - 1/\zeta _2 } \)

Example 7

(Gaussian kernel) If \(k(x,y) = \exp \left( - \frac{1}{2} \frac{(x-y)^2}{\theta ^2} \right) \), then

  • \(\int _0^1 k(., y) d\lambda = \theta \sqrt{2\pi } \times \left[ \varPhi \left( \frac{1-y}{\theta } \right) + \varPhi \left( \frac{y}{\theta } \right) - 1 \right] \)

  • \(\iint _{[0,1]^2} k(.,.) d(\lambda \otimes \lambda ) = 2 (e^{-1/(2 \theta ^2)} -1 ) + \theta \sqrt{2 \pi } \times \left( 2 \varPhi \left( \frac{1}{\theta } \right) - 1 \right) \)

where \(\varPhi \) denotes the cdf of the standard normal distribution.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer International Publishing Switzerland

About this paper

Cite this paper

Ginsbourger, D., Roustant, O., Schuhmacher, D., Durrande, N., Lenz, N. (2016). On ANOVA Decompositions of Kernels and Gaussian Random Field Paths. In: Cools, R., Nuyens, D. (eds) Monte Carlo and Quasi-Monte Carlo Methods. Springer Proceedings in Mathematics & Statistics, vol 163. Springer, Cham. https://doi.org/10.1007/978-3-319-33507-0_15

Download citation

Publish with us

Policies and ethics