Skip to main content
Log in

Nonparametric estimation of probabilistic sensitivity measures

  • Published:
Statistics and Computing Aims and scope Submit manuscript

Abstract

Computer experiments are becoming increasingly important in scientific investigations. In the presence of uncertainty, analysts employ probabilistic sensitivity methods to identify the key-drivers of change in the quantities of interest. Simulation complexity, large dimensionality and long running times may force analysts to make statistical inference at small sample sizes. Methods designed to estimate probabilistic sensitivity measures at relatively low computational costs are attracting increasing interest. We first, propose new estimators based on a one-sample design and building on the idea of placing piecewise constant Bayesian priors on the conditional distributions of the output given each input, after partitioning the input space. We then present two alternatives, based on Bayesian non-parametric density estimation, which bypass the need for predefined partitions. Quantification of uncertainty in the estimation process through is possible without requiring additional simulator evaluations via Bootstrap in the simplest proposal, or from the posterior distribution over the sensitivity measures, when the entire inferential procedure is Bayesian. The performance of the proposed methods is compared to that of traditional point estimators in a series of numerical experiments comprising synthetic but challenging simulators, as well as a realistic application.

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
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Antoniano-Villalobos, I., Wade, S., Walker, S.G.: A Bayesian nonparametric regression model with, normalized weights: a study of hippocampal atrophy in in Alzheimer’s disease. J. Am. Stat. Assoc. 109(506), 477–490 (2014)

    MathSciNet  MATH  Google Scholar 

  • Archer, G., Saltelli, A., Sobol, I.: Sensitivity measures, ANOVA-like techniques and the use of bootstrap. J. Stat. Comput. Simul. 58(2), 99–120 (1997)

    MATH  Google Scholar 

  • Barrientos, A.F., Jara, A., Quintana, F.A.: On the support of MacEachern’s dependent Dirichlet processes and extensions. Bayesian Anal. 7, 277–310 (2012)

    MathSciNet  MATH  Google Scholar 

  • Baucells, M., Borgonovo, E.: Invariant probabilistic sensitivity analysis. Manag. Sci. 59(11), 2536–2549 (2013)

    Google Scholar 

  • Benoumechiara, N., Elie-Dit-Cosaque, K.: Shapley effects for sensitivity analysis with dependent inputs: bootstrap and kriging-based algorithms. arXiv preprint arXiv:1801.03300 (2018)

  • Blackwell, D., MacQueen, J.B.: Ferguson distributions via Pólya urn schemes. Ann. Stat. 1(2), 353–355 (1973)

    MATH  Google Scholar 

  • Blatman, G., Sudret, B.: Efficient computation of global sensitivity indices using sparse polynomial chaos expansions. Reliab. Eng. Syst. Saf. 95(11), 1216–1229 (2010)

    Google Scholar 

  • Borgonovo, E.: A new uncertainty importance measure. Reliab. Eng. Syst. Saf. 92(6), 771–784 (2007)

    Google Scholar 

  • Borgonovo, E., Hazen, G., Plischke, E.: A common rationale for global sensitivity measures and their estimation. Risk Anal. 36(10), 1871–1895 (2016)

    Google Scholar 

  • Borgonovo, E., Iooss, B.: Moment-independent and reliability-based importance measures. In: Ghanem, R., Higdon, D., Owhadi, H. (eds.) Handbook of Uncertainty Quantification, pp. 1265–1287. Springer, Berlin (2017)

    Google Scholar 

  • Borgonovo, E., Tarantola, S., Plischke, E., Morris, M.D.: Transformation and invariance in the sensitivity analysis of computer experiments. J. R. Stat. Soc. B 76(5), 925–947 (2014)

    MathSciNet  MATH  Google Scholar 

  • Camerlenghi, F., Lijoi, A., Orbanz, P., Prünster, I.: Distribution theory for hierarchical processes. Ann. Stat. 47, 67–92 (2019)

    MathSciNet  MATH  Google Scholar 

  • Camerlenghi, F., Lijoi, A., Prünster, I.: Bayesian prediction with multiple-samples information. J. Multivar. Anal. 156, 18–28 (2017)

    MathSciNet  MATH  Google Scholar 

  • Castaings, W., Borgonovo, E., Tarantola, S., Morris, M.D.: Sampling strategies in density-based sensitivity analysis. Environ. Modell. Softw. 38, 13–26 (2012)

    Google Scholar 

  • Chastaing, G., Gamboa, F., Prieur, C.: Generalized Hoeffding–Sobol decomposition for dependent variables: application to sensitivity analysis. Electron. J. Stat. 6, 2420–2448 (2012)

    MathSciNet  MATH  Google Scholar 

  • Da Veiga, S.: Global sensitivity analysis with dependence measures. J. Stat. Comput. Simul. 85(7), 1283–1305 (2015)

    MathSciNet  Google Scholar 

  • Da Veiga, S., Wahl, F., Gamboa., F.: Local polynomial estimation for sensitivity analysis on models with correlated inputs. Technometrics 51(4), 452–463 (2009)

    MathSciNet  Google Scholar 

  • De Lozzo, M., Marrel, A.: Estimation of the derivative-based global sensitivity measures using a Gaussian process metamodel. SIAM-ASA J. Uncertain. Quantif. 4(1), 708–738 (2016)

    MathSciNet  MATH  Google Scholar 

  • Dunson, D.B., Park, J.H.: Kernel stick-breaking processes. Biometrika 95(2), 307–323 (2008)

    MathSciNet  MATH  Google Scholar 

  • Dunson, D.B., Rodriguez, A.: Nonparametric Bayesian models through probit stick-breaking processes. Bayesian Anal. 6, 145–178 (2011)

    MathSciNet  MATH  Google Scholar 

  • Escobar, M.D., West, M.: Bayesian density estimation and inference using mixtures. J. Am. Stat. Assoc. 90(430), 577–588 (1995)

    MathSciNet  MATH  Google Scholar 

  • European Commission: Impact Assessment Guidelines (2009)

  • Favaro, S., Lijoi, A., Prünster, I.: On the stick-breaking representation of normalized inverse Gaussian priors. Biometrika 99(3), 663–674 (2012)

    MathSciNet  MATH  Google Scholar 

  • Ferguson, T.S.: A Bayesian analysis of some nonparametric problems. Ann. Stat. 1(2), 209–230 (1973)

    MathSciNet  MATH  Google Scholar 

  • Ferguson, T.S.: Bayesian density estimation by mixtures of normal distributions. In: Rizvi, M.H., Rustagi, J.S., Siegmund, D. (eds.) Recent Advances in Statistics, pp. 287–302. Academic Press, Cambridge (1983)

    Google Scholar 

  • Freedman, D., Diaconis, P.: On the histogram as a density estimator: L2 theory. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete 57(4), 453–476 (1981)

    MathSciNet  MATH  Google Scholar 

  • Gamboa, F., Janon, A., Klein, T., Lagnoux, A., Prieur, C.: Statistical inference for Sobol pick-freeze Monte Carlo method. Statistics 50(4), 881–902 (2016)

    MathSciNet  MATH  Google Scholar 

  • Gamboa, F., Klein, T., Lagnoux, A.: Sensitivity analysis based on Cramér von Mises distance. SIAM/ASA J. Uncertain. Quantif. 6(2), 522–548 (2018)

    MathSciNet  MATH  Google Scholar 

  • Ghanem, R., Higdon, D., Owhadi, H. (eds.): Handbook of Uncertainty Quantification. Springer International Publishing, Berlin (2016)

    Google Scholar 

  • Ghosal, S., van der Vaart, A.: Fundamentals of Nonparametric Bayesian Inference. Cambridge Series in Statistical and Probabilistic Mathematics. Cambridge University Press, Cambridge (2017)

    Google Scholar 

  • Griffin, J.E., Steel, M.F.J.: Order-based dependent Dirichlet processes. J. Am. Stat. Assoc. 101(473), 179–194 (2006)

    MathSciNet  MATH  Google Scholar 

  • Hastie, T.J., Tibshirani, R.J.: Generalized Additive Models. Chapman & Hall/CRC Monographs on Statistics and Applied Probability. Chapman and Hall/CRC, Lodon (1990)

    Google Scholar 

  • He, X.: Rotated sphere packing designs. J. Am. Stat. Assoc. 112(520), 1612–1622 (2017)

    MathSciNet  Google Scholar 

  • Helton, J.C., Sallaberry, C.J.: Computational implementation of sampling-based approaches to the calculation of expected dose in performance assessments for the proposed high-level radioactive waste repository at Yucca Mountain. Nevada Reliab. Eng. Syst. Saf. 94(3), 699–721 (2009)

    Google Scholar 

  • Hjort, N.L.: Bayesian nonparametric bootstrap confidence intervals. Tech. rep. Stanford University Laboratory for Computational Statistics (1985)

  • Hjort, N.L.: Bayesian and empirical Bayesian bootstrapping. Tech. rep. Matematisk Institutt, Universitetet i Oslo (1991)

  • Hjort, N.L., Holmes, C., Müller, P., Walker, S.G. (eds.): Bayesian Nonparametrics. Cambridge University Press, Cambridge (2010)

    MATH  Google Scholar 

  • Homma, T., Saltelli, A.: Importance measures in global sensitivity analysis of nonlinear models. Reliab. Eng. Syst. Saf. 52(1), 1–17 (1996)

    Google Scholar 

  • Iman, R.L., Hora, S.C.: A robust measure of uncertainty importance for use in fault tree system analysis. Risk Anal. 10, 401–406 (1990)

    Google Scholar 

  • Ishwaran, H., James, L.F.: Gibbs sampling methods for stick-breaking priors. J. Am. Stat. Assoc. 96(453), 161–173 (2001)

    MathSciNet  MATH  Google Scholar 

  • Janon, A., Klein, T., Lagnoux, A., Nodet, M., Prieur, C.: Asymptotic normality and efficiency of two Sobol index estimators. ESAIM Probab. Stat. 18, 342–364 (2014a)

    MathSciNet  MATH  Google Scholar 

  • Janon, A., Nodet, M., Prieur, C.: Uncertainties assessment in global sensitivity indices estimation from metamodels. Int. J. Uncertain. Quantif. 4(1), 21–36 (2014b)

    MathSciNet  MATH  Google Scholar 

  • Jara, A., Hanson, T., Quintana, F., Müller, P., Rosner, G.: DPpackage: Bayesian semi- and nonparametric modeling in R. J. Stat. Softw. 40(5), 1–30 (2011)

    Google Scholar 

  • Jiménez Rugama, L.A., Gilquin, L.: Reliable error estimation for Sobol’ indices. Stat. Comput. 28(4), 725–738 (2018)

    MathSciNet  MATH  Google Scholar 

  • Kalli, M., Griffin, J.E., Walker, S.G.: Slice sampling mixture models. Stat. Comput. 21(1), 93–105 (2011)

    MathSciNet  MATH  Google Scholar 

  • Kleijnen, J.P.C.: Design and Analysis of Simulation Experiments. Springer, Berlin (2008)

    MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Le Gratiet, L., Marelli, S., Sudret, B.: Metamodel-based sensitivity analysis: polynomial chaos expansions and Gaussian processes. In: Ghanem, R., Higdon, D., Owhadi, H. (eds.) Handbook of Uncertainty Quantification, pp. 1289–1325. Springer, Berlin (2017)

    Google Scholar 

  • Li, G., Rabitz, H.: General formulation of HDMR component functions with independent and correlated variables. J. Math. Chem. 50, 99–130 (2012)

    MathSciNet  MATH  Google Scholar 

  • Lijoi, A., Mena, R.H., Prünster, I.: Controlling the reinforcement in Bayesian non-parametric mixture models. J. R. Stat. Soc. Ser. B (Stat. Methodol.) 69(4), 715–740 (2007)

    MathSciNet  Google Scholar 

  • Lin, C.D., Bingham, D., Sitter, R.R., Tang, B.: A new and flexible method for constructing designs for computer experiments. Ann. Stat. 38(3), 1460–1477 (2010)

    MathSciNet  MATH  Google Scholar 

  • Lo, A.Y.: On a class of Bayesian nonparametric estimates: I. Density estimates. Ann. Stat. 12(1), 351–357 (1984)

    MathSciNet  MATH  Google Scholar 

  • MacEachern, S.N.: Dependent nonparametric processes. In: ASA Proceedings-Section on Bayesian Statistical Science American Statistical Association, pp. 50–55. American Statistical Association, Baltimore (1999)

  • MacEachern, S.N.: Dependent Dirichlet processes. Tech. rep. Department of Statistics, The Ohio State University (2000)

  • Mara, T., Tarantola, S.: Variance-based sensitivity indices for models with dependent inputs. Reliab. Eng. Syst. Saf. 107, 115–121 (2012)

    Google Scholar 

  • Mara, T.A., Joseph, O.R.: Comparison of some efficient methods to evaluate the main effect of computer model factors. J. Stat. Comput. Simul. 78(2), 167–178 (2008)

    MathSciNet  MATH  Google Scholar 

  • Marrel, A., Iooss, B., Veiga, S., Ribatet, M.: Global sensitivity analysis of stochastic computer models with joint metamodels. Stat. Comput. 22(3), 833–847 (2012)

    MathSciNet  MATH  Google Scholar 

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

    Google Scholar 

  • Müller, P., Quintana, F.A.: Nonparametric Bayesian data analysis. Stat. Sci. 19(1), 95–110 (2004)

    MathSciNet  MATH  Google Scholar 

  • Nuclear Energy Agency: PSACOIN Level E Intercomparison. Tech. rep, Organisation for Economic Co-operation and Development, Paris (France) (1989)

  • Oakley, J., O’Hagan, A.: Bayesian inference for the uncertainty distribution of computer model outputs. Biometrika 89(4), 769–784 (2002)

    Google Scholar 

  • Oakley, J., O’Hagan, A.: Probabilistic sensitivity analysis of complex models: a Bayesian approach. J. R. Stat. Soc. B 66(3), 751–769 (2004)

    MathSciNet  MATH  Google Scholar 

  • Papaspiliopoulos, O., Roberts, G.O.: Retrospective Markov chain Monte Carlo methods for Dirichlet process hierarchical models. Biometrika 95(1), 169–186 (2008)

    MathSciNet  MATH  Google Scholar 

  • Pearson, K.: On the General Theory of Skew Correlation and Non-linear Regression, vol. XIV of Mathematical Contributions to the Theory of Evolution. Drapers’ Company Research Memoirs. Dulau & Co, London (1905)

    Google Scholar 

  • Pitman, J.: Random discrete distributions invariant under size-biased permutation. Adv. Appl. Probab. 28(2), 525–539 (1996)

    MathSciNet  MATH  Google Scholar 

  • Pitman, J., Yor, M.: The two-parameter Poisson–Dirichlet distribution derived from a stable subordinator. Ann. Probab. 25(2), 855–900 (1997)

    MathSciNet  MATH  Google Scholar 

  • Plischke, E., Borgonovo, E.: Probabilistic sensitivity measures from empirical cumulative distribution functions. Work in Progress (2017)

  • Plischke, E., Borgonovo, E., Smith, C.L.: Global sensitivity measures from given data. Eur. J. Oper. Res. 226(3), 536–550 (2013)

    MathSciNet  MATH  Google Scholar 

  • Pronzato, L., Müller, W.G.: Design of computer experiments: space filling and beyond. Stat. Comput. 22(3), 681–701 (2012)

    MathSciNet  MATH  Google Scholar 

  • Rahman, S.: The f-sensitivity index. SIAM/ASA J. Uncertain. Quantif. 4(1), 130–162 (2016)

    MathSciNet  MATH  Google Scholar 

  • Ratto, M., Pagano, A., Young, P.: State dependent parameter metamodelling and sensitivity analysis. Comput. Phys. Commun. 177(11), 863–876 (2007)

    Google Scholar 

  • Rényi, A.: On measures of statistical dependence. Acta Math. Acad. Sci. Hung. 10, 441–451 (1959)

    MATH  Google Scholar 

  • Röhlig, K.J., Plischke, E., Bolado Lavin, R., Becker, D.A., Ekstroem, P.A., Hotzel, S.: Lessons learnt from studies on sensitivity analysis techniques in the EU project PAMINA: a benchmark study. In: Reliability, Risk and Safety. Theory and Applications. Proceedings of the ESREL 2009 Annual Conference, vol. 3, pp. 1769–1775. CRC Press, Boca Raton (2009)

  • Sacks, J., Welch, W., Mitchell, T., Wynn, H.: Design and analysis of computer experiments. Stat. Sci. 4(4), 409–423 (1989)

    MathSciNet  MATH  Google Scholar 

  • Saltelli, A.: Making best use of model evaluations to compute sensitivity indices. Comput. Phys. Commun. 145, 280–297 (2002)

    MATH  Google Scholar 

  • Saltelli, A., Tarantola, S.: On the relative importance of input factors in mathematical models: safety assessment for nuclear waste disposal. J. Am. Stat. Assoc. 97(459), 702–709 (2002)

    MathSciNet  MATH  Google Scholar 

  • Saltelli, A., Tarantola, S., Campolongo, F.: Sensitivity analysis as an ingredient of modelling. Stat. Sci. 19(4), 377–395 (2000)

    Google Scholar 

  • Santner, T., Williams, B., Notz, W.: The Design and Analysis of Computer Experiments. Springer, New York (2003)

    MATH  Google Scholar 

  • Santner, T.J., Williams, B., Notz, W.: The Design and Analysis of Computer Experiments, second edn. Springer, New York, NY (2018)

    MATH  Google Scholar 

  • Sethuraman, J.: A constructive definition of Dirichlet priors. Stat. Sin. 4(2), 639–650 (1994)

    MathSciNet  MATH  Google Scholar 

  • Smith, R.L., Tebaldi, C., Nychka, D., Mearns, L.O.: Bayesian modeling of uncertainty in ensembles of climate models. J. Am. Stat. Assoc. 104(485), 97–116 (2009)

    MathSciNet  MATH  Google Scholar 

  • Sobol’, I.M.: Sensitivity estimates for nonlinear mathematical models. Math. Modell. Comput. Exp. 1, 407–414 (1993)

    MathSciNet  MATH  Google Scholar 

  • Storlie, C., Swiler, L., Helton, J., Sallaberry, C.: Implementation and evaluation of nonparametric regression procedures for sensitivity analysis of computationally demanding models. Reliab. Eng. Syst. Saf. 94(11), 1735–1763 (2009)

    Google Scholar 

  • Storlie, C.B., Helton, J.C.: Multiple predictor smoothing methods for sensitivity analysis: description of techniques. Reliab. Eng. Syst. Saf. 93(1), 28–54 (2008)

    Google Scholar 

  • Strong, M., Oakley, J.E.: An efficient method for computing partial expected value of perfect information for correlated inputs. Med. Decis. Mak. 33(6), 755–766 (2013)

    Google Scholar 

  • Strong, M., Oakley, J.E., Chilcott, J.: Managing structural uncertainty in health economic decision models: a discrepancy approach. J. R. Stat. Soc. Ser. C (Appl. Stat.) 61(1), 25–45 (2012)

    MathSciNet  Google Scholar 

  • Teh, Y.W., Jordan, M.I.: Hierarchical Bayesian nonparametric models with applications. In: Hjort, N., Holmes, C., Müller, P., Walker, S. (eds.) Bayesian Nonparametrics: Principles and Practice. Cambridge University Press, Cambridge (2010)

    Google Scholar 

  • Teh, Y.W., Jordan, M.I., Beal, M.J., Blei, D.M.: Hierarchical Dirichlet processes. J. Am. Stat. Assoc. 101(476), 1566–1581 (2006)

    MathSciNet  MATH  Google Scholar 

  • Tissot, J.Y., Prieur, C.: A randomized orthogonal array-based procedure for the estimation of first- and second-order Sobol’ indices. J. Stat. Comput. Simul. 85(7), 1358–1381 (2015)

    MathSciNet  Google Scholar 

  • Tuo, R., Wu, J.: Efficient calibration of imperfect computer codes. Ann. Stat. 43(6), 2331–2352 (2015)

    MATH  Google Scholar 

  • U.S. Environmental Protection Agency: Guidance on the Development, Evaluation, and Application of Environmental Models (2009)

  • Wei, P., Lu, Z., Song, J.: Moment-independent sensitivity analysis using copula. Risk Anal. 34(2), 210–222 (2014)

    Google Scholar 

  • Wong, R.K.W., Storlie, C.B., Lee, T.C.M.: A frequentist approach to computer model calibration. J. R. Stat. Soc. Ser. B Stat. Methodol. 79(2), 635–648 (2017)

    MathSciNet  MATH  Google Scholar 

  • Wood, F., Gasthaus, J., Archambeau, C., James, L., Teh, Y.W.: The sequence memoizer. Commun. Assoc. Comput. Mach. 54(2), 91–98 (2011)

    Google Scholar 

  • Yau, C., Papaspiliopoulos, O., Roberts, G.O., Holmes, C.: Bayesian non-parametric hidden Markov models with applications in genomics. J. R. Stat. Soc. Ser. B (Stat. Methodol.) 73(1), 37–57 (2011)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank the Associate Editor and two anonymous reviewers for their valuable contributions.

Note The code can be downloaded from: https://github.com/LuXuefei/Nonparametric-estimation-of-probabilisticsensitivity-measures, along with the simulated data to reproduce results.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Emanuele Borgonovo.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 Given data estimators for the sensitivity measures in Table 1

The one-sample estimator of \(\eta _i\) used here relies on a plug-in estimator of the inner statistic, based on the output sample mean and variance, \({\bar{y}}\) and \(s^2_y\) respectively, to estimate the marginal mean and variance of Y. The within cluster sample mean \({\bar{y}}^i_m = \dfrac{1}{n^i_m} \sum _{y\in {\mathbf {y}}^i_m} y\) with \({\mathbf {y}}^{i}_{m}=\{y_{j}:x_{j}^{i}\in {\mathcal {X}}^i_m, j=1,2,\ldots ,n\}\) is used to estimate the conditional mean of \(Y|X^i\in {\mathcal {X}}^i_m\). The final expression (see e.g. Strong et al. 2012) takes the form of Eq. (2) with:

$$\begin{aligned} {\widehat{\eta }}_i^\star = \sum _{m=1}^{M}\frac{n^i_m}{n} \frac{({\bar{y}}^i_m - {\bar{y}})^2}{s^2_y}. \end{aligned}$$
(17)

The one-sample estimator for the \(\delta -\)importance introduced by Plischke et al. (2013) can be written as:

$$\begin{aligned} {\widehat{\delta }}_i^{\star } = \sum ^{M}_{m=1} \frac{n^i_m}{n} \int _{{\mathcal {Y}}}|{\hat{f}}^{\star }_{Y}(y)-{\hat{f}}^i_m (y)|\text {d}y, \end{aligned}$$
(18)

where \( \hat{f}^{\star }_{Y} \) and \( {\hat{f}}^i_m \) denote kernel-smoothed histograms of the full output vector \({\mathbf {y}} = (y_1,\ldots y_n) \) and the within cluster output vector \({\mathbf {y}}^{i}_{m}\), respectively. The authors propose a quadrature method for the numerical integration required by the \( L^1\)-norm in the inner operator, but other solutions could be used, producing similar estimators. Because estimates of this type rely on the approximation or estimation of probability density functions, we refer to them as pdf-based estimators.

Plischke and Borgonovo (2017) observe that the kernel-smoothing methods commonly involved in the calculation of pdf-based estimators may induce bias, even at high sample sizes, for simulators with a sparse output. Therefore, they introduce alternative cdf-based estimators which rely on the properties of empirical cumulative distribution functions. Scheffé’s theorem allows one to write the \(L^1\)-distance between two probability density functions in terms of the associated probability functions, as \(\int _{{\mathcal {Y}}}|f_1(y)-f_2(y)|\text {d}y=2({\mathbb {P}}_1(Y\in B)-{\mathbb {P}}_2(Y\in B))\), where B is the set of values for which \(f_1(y)>f_2(y)\). Since B can be written as a union of intervals \((a(t),b(t))_{t=1}^T\), these probabilities can be calculated from the corresponding cumulative distribution functions. Thus, a cdf-based estimator of \(\delta _i\) can be obtained as:

$$\begin{aligned} {{\widehat{\delta }}}_i^\diamond= & {} \sum _{m=1}^M \frac{n^i_m}{n}\,\sum _{t=1}^{T^i_m} \left( \hat{F}^i_m(\hat{b}^i_m(t)) -\hat{F}^i_m(\hat{a}^i_m(t)) \right) \nonumber \\&-\left( \hat{F}_Y(\hat{b}^i_m(t)) -\hat{F}_Y(\hat{a}^i_m(t)) \right) . \end{aligned}$$
(19)

For further details on the estimation of the intervals \((\hat{a}^i_m(t),\hat{b}^i_m(t)) \), we refer to Plischke and Borgonovo (2017).

Since \(\beta _i\) is itself a cdf-based sensitivity measure, the definition of a one-sample cdf-based estimator is straightforward:

$$\begin{aligned} {\widehat{\beta }}_{i}^{\diamond }=\sum _{m=1}^{M}\frac{n^i_m}{n}\max _{j\in \{1,\ldots ,n\}} \left| {\hat{F}}_Y(y_j)-{\hat{F}}^i_m(y_j) \right| , \end{aligned}$$
(20)

where \({\hat{F}}_Y\), and \({\hat{F}}^i_m\) are the empirical cdf’s of \({\mathbf {y}} \) and \({\mathbf {y}}^i_m\), respectively, i.e.:

$$\begin{aligned} \hat{F}_{Y}(y)= & {} \frac{1}{n}\sum _{j=1}^{n}{\mathbb {1}}_{(-\infty ,y_j]}(y);\nonumber \\ {\hat{F}}^i_m(y)= & {} \frac{1}{n^i_m}\sum _{y_j \in {\mathbf {y}}^i_m }\mathbb {1}_{(-\infty ,y_j]}(y), \end{aligned}$$
(21)

and \(\mathbb {1}_A(y)\) denotes the indicator function, taking the value 1 if \(y\in A\) and 0 otherwise.

Recalling that the expected value of a random variable Y can be calculated as the integral of its survival function, \({\mathbb {E}}[Y]=\int _{\mathcal {Y}}(1-F_Y(y))\text {d}y\), a cdf-based one-sample estimator of the variance-based sensitivity measure, \(\eta _{i}\) is given by:

$$\begin{aligned} {\widehat{\eta }}_{i}^{\Diamond } = \sum _{m=1}^{M}\frac{n^i_{m}}{n}\, \frac{\left( \int _{\mathcal {Y}} {\hat{F}}^i_m(y)- \hat{F}_{Y}(y) \text {d}y \right) ^2 }{{{\widehat{\sigma }}}^2_Y}. \end{aligned}$$
(22)

Notice that, since the empirical distribution functions are piece-wise constant, the integral in the above expression reduces to a sum. Plischke and Borgonovo (2017) propose an efficient way to calculate this integral.

1.2 Numerical experiments for the partition selection problem

Fig. 5
figure 5

RMSE of sensitivity measures estimates for \(X^1\) of the 2-input simulator in Eq. (16). Magenta lines correspond to \(M=2.5\root 3 \of {n}\); \(n \in [300,900]\), \(M\in [5,34]\)

Fig. 6
figure 6

RMSE of sensitivity measures estimates for \(X^3\) of the 21-input simulator in Eq. (7). Magenta lines correspond to \(M=2.5\root 3 \of {n}\); \(n \in [300,900]\), \(M\in [5,34]\)

The authors performed several thought experiments on test cases. The results show the difficulty, maybe impossibility, of finding a universally valid rule for linking the partition size M to the sample size n. We report some experiments results.

Assume the analyst wants to find an “optimal ”(in some sense) partition refining strategy, i.e., a relationship that produces the partition size M that minimizes the estimation error at sample size n for the pdf-based point estimators \( {\widehat{\eta }}_i^{\star } \), \( {\widehat{\delta }}_i^{\star } \) and cdf-based point estimator \({\widehat{\beta }}_i^{\diamond } \) [Eqs. (17), (18) and (20)]. We focus on one estimator type for simplicity and also because Borgonovo et al. (2016) propose an heuristic inspired by the rule of histogram partitioning of Freedman-Diaconis Freedman and Diaconis (1981), in which \(M\sim \root 3 \of {n}\).

To evaluate the estimators’ performance at fixed values of M and n, we use the Root Mean Square Error (RMSE):

$$\begin{aligned} \text {RMSE}_i(n) \approx \sqrt{\frac{\sum _{s=1}^{S}\left( {\widehat{\xi }}^{i}_{s}(n)- \xi _i \right) ^2}{S}} \end{aligned}$$

where S is the number of bootstrap replicates. \({\hat{\xi }}_{i,l}\) is the l-th bootstrap replicates of \(\xi _i\).

We estimate the sensitivity measures with sample sizes varying from 300 to 900, and partition sizes covering the natural numbers between 5 and 35. Then we calculate the RMSEs with \(S=100\) bootstrap replicates. Figures 5 and 6 present the heatplot of RMSEs in percentage (\(\hbox {RMSE}_i/\xi _i\cdot 100\%\)). The horizontal axis indicates the sample size, and the vertical axis the partition size. The darker the color of a region in the plot, the lower the estimation error. For example, in Fig. 5a, dark (blue) refers to low RMSE (less than 10%), and light (red) to relative high RMSE (higher than 14%). The magenta line maps n into M using the previously mentioned heuristic function. Figure 5 shows that the proposed heuristic works well on the 2-input simulator [Eq. (16)], with the magenta line falling mainly into dark coloured regions. However, for the 21-input simulator [Eq. (7)] we would incur in high errors at small sample sizes. For instance consider graph a) in Fig. 6. The graph reports the error in the estimates of \(\delta _3\) for the second model. The heuristic would propose values of M at about 20 for all values of n as optimal partition sizes. However, the partition size that minimizes the error is at about \(M=10\) or lower. The different behavior here could also be related to the differences in structure and dimensionality of the models. However, even for the same model, the heatplots differ significantly across the sensitivity measures. For the first simulator [Eq. (16)], the ideal partition size for the variance-based estimator is between 10 to 15 (Fig. 5b), while for mutual information, if falls between 20 to 25 (Fig. 5c).

These results show that aiming at postulating a universally valid heuristic might be a cumbersome task.

1.3 Details on the non-parametric estimators

We present further details regarding the implementation of the Bayesian non-parametric estimation methods in Sects. 3 and 4. Inference on the three selected sensitivity measures \(\eta _i\), \(\beta _i\) and \(\delta _i\) is performed independently for each \(i = 1,\ldots ,k\). Therefore, in order to simplify the notation, we will leave out the index i throughout this appendix, considering its value fixed. Throughout this section, all the integrals are approximated numerically using trapezoidal rule, and all the supremes are approximated by the maximum on a predetermined grid over \({\mathcal {Y}}\).

Partition-dependent Bu and Pu estimation. Note that model 3 is coherent, in the sense that it induces a unique prior over the unconditional distribution of Y, whenever the partitions are equiprobable, that is when \({\mathbb {P}} (X^i\in {\mathcal {X}}^i_m)=\frac{1}{M}\) for all \(i=1,2,\ldots ,k\) and \(m=1,2,\ldots ,M\). In fact,

$$\begin{aligned} {\mathbb {P}}_{Y}(\cdot |{\mathbb {P}}_{1:M}^{i})=\sum \limits _{m=1}^{M}{\mathbb {P}} _{m}^{i}(\cdot ){\mathbb {P}}(X^i\in {\mathcal {X}}^i_m)=\frac{1}{M} \sum \limits _{m=1}^{M}{\mathbb {P}}_{m}^{i}(\cdot )\text {.} \end{aligned}$$

Then, by marginalizing, we obtain

$$\begin{aligned} {\mathbb {P}}_{Y}(\cdot |\alpha G)= & {} \frac{1}{M}\sum \limits _{m=1}^{M}\int {\mathbb {P}}_{m}^{i}(\cdot )\text {d}{{\mathcal {D}}}{{\mathcal {P}}}({\mathbb {P}}_{m}^{i}|\alpha G)\\= & {} \int {\mathbb {P}}(\cdot )\text {d}{{\mathcal {D}}}{{\mathcal {P}}}({\mathbb {P}}|\alpha G), \end{aligned}$$

because \(\int {\mathbb {P}}_{m}^{i}(\cdot )\)d\({{\mathcal {D}}}{{\mathcal {P}}}({\mathbb {P}}_{m}^{i}|\alpha G)\) does not depend on i or m. In other words, a priori, \({\mathbb {P}}_{Y} \overset{}{\sim }{{\mathcal {D}}}{{\mathcal {P}}}(\alpha G)\), so that the prior for the marginal simulator distribution is also a Dirichlet process. This statement alone, however, provides no information on the probabilistic dependence of Y on \(X^i\). Thus, it is not meaningful, by itself, for a sensitivity analysis.

The posterior of the marginal for Y can be obtained as:

$$\begin{aligned} {\mathbb {P}}_{Y}(\cdot |\alpha G,D^i)=\frac{1}{M}\sum \limits _{m=1}^{M}\int {\mathbb {P}}_{m}^{i}(\cdot )\text {d}{{\mathcal {D}}}{{\mathcal {P}}}({\mathbb {P}}_{m}^{i}|(\alpha +n_{m}^{i}) {\widetilde{G}}_{m}^{i}),\nonumber \\ \end{aligned}$$
(23)

which may depend both on i and m. However, the marginal coherence of the model still holds, at least asymptotically. Informally, for an equiprobable partition, \({\mathbb {P}}(X^i\in {\mathcal {X}}^i_m)=1/M\), \(n_{m}^{i}\simeq n/M\) when the sample size n is sufficiently large, so \(\alpha /(\alpha +n_{m}^{i}) \simeq M\alpha /(M\alpha +n)\) and \(n_{m}^{i}/(\alpha +n_{m}^{i})\simeq n/(M\alpha +n)\). Furthermore, \(\sum _m (1/n_{m}^{i})\delta ^{Dirac}_y\simeq \sum _m (M/n)\delta ^{Dirac}_y\). Thus, asymptotically, \({\mathbb {P}}_{Y}(\cdot |\alpha G,D^i)\) does not depend on m or i and \({\mathbb {P}}_{Y}(\cdot |\alpha G,D^i)\overset{}{\sim } {{\mathcal {D}}}{{\mathcal {P}}}((\alpha +n){\widetilde{G}})\), where

$$\begin{aligned} {\widetilde{G}}=\dfrac{\alpha }{\alpha +n}G+\dfrac{n}{\alpha +n} {\widehat{\mathbb {P}}}_n, \end{aligned}$$
(24)

and \(\widehat{{\mathbb {P}}}_n\) denotes the empirical distribution of Y based on the full set of observations, \((y_1,\ldots ,y_n)\). Note that this is the usual posterior corresponding to the DP prior on \({\mathbb {P}}_Y\).

For the numerical experiments in Sects. 3 and 4.3, the input data, \({\mathbf {x}}\), are obtained by transforming a 21-dimensional standard Gaussian sample generated via a Halton Quasi random sequence via the Cholesky decomposition of the covariance matrix. Specifically, MATLAB© functions haltonset and chol are used. The mass parameter, \( \alpha \), for the DP prior is set equal to \( 0.1\,n/M\) throughout. The base measure, G, is a Normal distribution with hyper-parameters fixed via an empirical approach, based on the available sample \({\mathbf {y}}\). \({\mathbf {y}}_m^i\). Overall, these choices centre the prior distribution for \(Y|X^i\in {\mathcal {X}}_m^i\) roughly around the marginal distribution of Y, thus favouring, a priori, independence between the Y and \(X^i\), with a precision proportional to the number of observations in each partition set. In practical applications, prior information elicited from experts may be expressed through different choices of \(\alpha \) and G. Algorithm 1 details the inferential procedure. Note that the calculations of \(\eta ^{C,s}\), \(\delta ^{C,s}\) and \(\beta ^{C,s}\) are equivalent to the pdf-based estimators in Eqs. (17)–(20) but with the enriched samples. Alternatively, the cdf-based estimators in Eqs. (22) and (19) could be used for \(\eta ^{C,s}\) and \(\delta ^{C,s}\).

figure a

Partition-free BNJ estimation. Following the proposal in Jara et al. (2011), we choose G to be a Normal-Inverse Wishart distribution

$$\begin{aligned}&(\mu _\ell , \varSigma _\ell )| (m_1, \gamma , \psi _1){\mathop {\sim }\limits ^{iid}}{\mathcal {N}}\left( \mu _{\ell } \vert m_1, \frac{1}{\gamma } \varSigma \right) IW(\varSigma _{\ell } \vert 4, \psi _1),\\&\ell =1,2,\ldots \end{aligned}$$

where \({\mathcal {N}}(\cdot |m,A)\) denotes a bivariate normal distribution with mean m and covariance matrix A, and \(IW(\cdot |4, \psi )\) denotes an Inverse-Wishart distribution with mean \(\psi ^{-1}\). A hyper-prior is assigned to the parameters of the base measure, with hyperparameters determined empirically:

$$\begin{aligned}&\gamma \sim \text {Gamma} \left( \cdot |0.5, 0.5\right) ,\; m_1 \vert (m_2, s_2) \sim {\mathcal {N}}(\cdot |m_2, s_2),\\&\psi _1 \vert (s_2) \sim IW (\cdot |4, s_2^{-1}), \end{aligned}$$

where \(\text {Gamma} (\cdot |a_1,a_2)\) denotes the Gamma distribution with mean \(a_1/a_2\). Inference is achieved through the function DPdensity from the DPpackage in R. The output is a MCMC posterior sample \({\underline{\theta }}^s = ({\underline{w}}^s,{\underline{\mu }}^s,{\underline{\varSigma }}^s)\), \(s=1,\ldots ,S\). In practice, the number \(J_s\) of components with non-zero weights is finite, thus we have

$$\begin{aligned} {\underline{w}}^s= & {} (w_1^s, \dots , w_{J_s}^s), \quad \nonumber \\ {\underline{\mu }}^s= & {} (\mu _1^s, \dots , \mu _{J_s}^s), \quad {\underline{\varSigma }}^s = (\varSigma _1^s, \dots , \varSigma _{J_s}^s), \nonumber \\ \text {with} \qquad \mu _{\ell }^s= & {} \left[ {\begin{array}{c} \mu _{1,\ell }^s \\ \mu _{2,\ell }^s \\ \end{array} } \right] , \quad \varSigma _{\ell }^s = \left[ {\begin{array}{cc} \sigma _{1,\ell }^s &{} \sigma _{3,\ell }^s \\ \sigma _{3,\ell }^s &{} \sigma _{2,\ell }^s\\ \end{array} } \right] . \end{aligned}$$
(25)

Given the posterior realizations, the corresponding joint density can be obtained:

$$\begin{aligned} f^{BNJ,s}_{X,Y} (x,y|{\underline{\theta }}^s) = \sum _{\ell =1}^{J_s} w_\ell ^s \cdot {\mathcal {N}} (x,y|\mu _\ell ^s,\varSigma _\ell ^s). \end{aligned}$$
(26)

By the properties of the bivariate Normal distribution, the marginal and conditional distributions, \(f^{BNJ,s}_{Y}\) and \(f^{BNJ,s}_{Y\vert X^i}\) recpectively, are also mixtures of Normal distributions:

$$\begin{aligned} f^{BNJ,s}_{Y} (y \vert {\underline{\theta }}^s)= & {} \sum _{\ell =1}^{J_s} w_\ell ^s \cdot {\mathcal {N}} (y|\mu _{2,\ell }^s,\sigma _{2,\ell }^s), \nonumber \\ f^{BNJ,s}_{Y\vert x} (y \vert x,{\underline{\theta }}^s )= & {} \sum _{\ell =1}^{J_s} w_\ell ^s \cdot {\mathcal {N}} \left( \cdot |\nu _{2,\ell }^s,\tau _{2,\ell }^s \right) \end{aligned}$$
(27)

where \(\nu _{2,\ell }^s = \mu _{2,\ell }^s + \sigma _{3,\ell }^s (x - \mu _{1,\ell }^s )/\sigma _{1,\ell }^s\) and \(\tau _{2,\ell }^s = \sigma _{2,\ell }^s - (\sigma _{3,\ell }^s )^2/\sigma _{1,\ell }^s\). Clearly, the corresponding cdfs, \(F^{BNJ,s}_{Y}\) and \(F^{BNJ,s}_{Y \vert X}\), as well as the marginal mean and variance can be calculated trivially. In particular,

$$\begin{aligned} \mu _{Y}^s :={\mathbb {E}}[Y \vert {\underline{\theta }}^s]= & {} \sum _{\ell =1}^{J_s} w_{\ell }^s \mu _{2,\ell }^s, \nonumber \\ V_Y^s :={\mathbb {V}}[Y \vert {\underline{\theta }}^s]= & {} \sum _{\ell =1}^{J_s} w_{\ell }^s \left( \sigma _{2,\ell }^s + \left( \mu _{Y}^s - \mu _{2,\ell }^s \right) ^2 \right) . \end{aligned}$$
(28)

For the numerical experiments in Sects. 3 and 4.3, the mass parameter is fixed to \(\alpha = 1\), and the remaining hyperparameters are set to \(m_2 = (\mu _{X}, {{\bar{y}}})\) and \(s_2 = \text {diag}(\sigma _{X}^2, s_y^2)\). The Pseudo-code of BNJ estimation for \(\eta , \delta \) and \(\beta \) is illustrated in Algorithm 2.

figure b

Partition-free BNC estimation. Following the proposal of Antoniano-Villalobos et al. (2014), we fix \(\alpha =1\) and choose \({\mathcal {K}}(x|\psi _\ell )\) to be a Normal kernel, with \( \psi _\ell =(\mu _{\ell }, \tau )\). The base measure G is given by:

$$\begin{aligned}&({\mathbf {b}}_\ell ,\sigma _{\ell },\mu _\ell ){\mathop {\sim }\limits ^{iid}}\\&{\mathcal {N}}\left( {\mathbf {b}}_\ell \mid {\mathbf {b}}_0, \sigma _{\ell } C^{-1} \right) \text {Gamma} (\sigma _{\ell }^{-1}\mid 1,1) {\mathcal {N}}\\&\quad \left( \mu _\ell \mid \mu _0, \left( \tau /10 \right) ^{-1} \right) , \end{aligned}$$

where \({\mathbf {b}}_\ell = (a_\ell ,b_\ell )\) and \(\tau \sim \text {Gamma} (\cdot \mid 1,1)\). The hyperparameters are chosen empirically. As an illustration, consider the 21-input simulator. Figure 7 shows the scatter-plot of \(({\mathbf {x}}^3,{\mathbf {y}})\) and the corresponding convex hull, i.e. the smallest convex set containing all points. In this case, we fix \({\mathbf {b}}_0 = (-1.5, -5.5)\) and \(C^{-1} = \text {diag}(43^2, 11^2)\), in order to allow each local linear component to lie between the blue and red lines in the figure, which represent the main behaviour of the data. We use the MATLAB© subroutine provided by Antoniano-Villalobos et al. (2014) to generate an MCMC posterior sample \(({\underline{\theta }}^s,{\underline{\psi }}^s)=({\underline{a}}^s, {\underline{b}}^s, {\underline{\sigma }}^s, {\underline{\omega }}^s, {\underline{\mu }}^s, \tau ^s)\), \(s=1\ldots S\), where

Fig. 7
figure 7

Scatter plot of Y and \(X^3\) for the 21-input simulator. Red lines constitute the convex hull of \(\{(x^3_1, y_1),\ldots ,(x^3_n, y_n)\}\). The bold blue and red lines are used for prior specification

$$\begin{aligned} {\underline{a}}^s= & {} (a_1^s, \dots , a_{J_s}^s), \nonumber \\ {\underline{b}}^s= & {} (b_1^s, \dots , b_{J_s}^s), \quad {\underline{\sigma }}^s = (\sigma _1^s, \dots , \sigma _{J_s}^s),\nonumber \\ {\underline{\omega }}^s= & {} (\omega _1^s, \dots , \omega _{J_s}^s), \quad {\underline{\mu }}^s = (\mu _1^s, \dots , \mu _{J_s}^s). \end{aligned}$$
(29)

Given the a posterior realization \(({\underline{\theta }}^s,{\underline{\psi }}^s))\), a conditional density can be obtained from Eqs. (14) and (15):

$$\begin{aligned} f^{BNC,s}_{Y|X} (y|x,{\underline{\theta }}^s,{\underline{\psi }}^s) = \sum _{\ell =1}^{J_s} w_\ell ^s(x) {\mathcal {N}}\left( y|a_\ell ^s+b_\ell ^s x, \sigma _\ell ^{s} \right) . \end{aligned}$$
(30)

The corresponding marginal pdf \(f^{BNC,s}_{Y}\) of Y is obtained by integrating with respect to the true \(f_X\):

$$\begin{aligned} f^{BNC,s}_{Y}(y \vert {\underline{\theta }}^s) \approx \int _{{\mathcal {X}}} f^{BNC,s}_{Y|X} f_{X} \text {d}x. \end{aligned}$$
(31)

Clearly, the corresponding marginal and conditional cdfs, \( F^{BNC,s}_{Y|X}\) and \(F^{BNC,s}_{Y}\), respectively can be obtained trivially. In particular, posterior realizations of the marginal mean and variance of Y are given by

$$\begin{aligned} \mu _Y^s:= & {} {\mathbb {E}}[Y\vert {\underline{\theta }}^s,{\underline{\psi }}^s] \approx \int _{{\mathcal {Y}}} y f^{BNC,s}_{Y} \text {d}y, \\ V_{Y}^s:= & {} {\mathbb {V}}[Y \vert {\underline{\theta }}^s,{\underline{\psi }}^s] \approx \int _{{\mathcal {Y}}} \left( y - \mu _Y^s \right) ^2 f^{BNC,s}_{Y} \text {d}y \end{aligned}$$

The Pseudo-code of BNC estimation for \(\eta , \delta \) and \(\beta \) is illustrated in Algorithm 3.

figure c

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Antoniano-Villalobos, I., Borgonovo, E. & Lu, X. Nonparametric estimation of probabilistic sensitivity measures. Stat Comput 30, 447–467 (2020). https://doi.org/10.1007/s11222-019-09887-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11222-019-09887-9

Keywords

Navigation