Skip to main content
Log in

Kriging with external drift in a Birnbaum–Saunders geostatistical model

  • Original Paper
  • Published:
Stochastic Environmental Research and Risk Assessment Aims and scope Submit manuscript

Abstract

Spatial models to describe dependent georeferenced data are applied in different fields and, particularly, are used to analyze earth and environmental data. Most of these applications are carried out under Gaussian spatial models. The Birnbaum–Saunders distribution is a unimodal and positively skewed model which has received considerable attention in several areas, including earth and environmental sciences. In addition, theoretical arguments have been provided to justify its usage in the data modeling from these sciences, at least in the same settings where the lognormal distribution can be employed. This paper presents kriging with external drift based on a Birnbaum–Saunders spatial model. The maximum likelihood method is considered to estimate its parameters. The results obtained in the paper are illustrated by an experimental data set related to agricultural management. Specifically, in this illustration, the spatial variability of magnesium content in the soil as a function of calcium content is analyzed.

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
Fig. 5

Similar content being viewed by others

References

  • Allard D, Naveau P (2007) A new spatial skew-normal random field model. Commun Stat Theor Methods 36:1821–1834

    Article  Google Scholar 

  • Assumpção RAB, Uribe-Opazo MA, Galea M (2014) Analysis of local influence in geostatistics using Student-t distribution. J Appl Stat 41:2323–2341

    Article  Google Scholar 

  • Bishop T, McBratney A (2001) A comparison of predictions methods for creation of the creation of field-extent soil property maps. Geoderma 103:149–160

    Article  Google Scholar 

  • Cambardella C, Moorman T, Novak J, Parkin T, Karlen D, Turco R, Konopka A (1994) Field-scale variability of soil properties in central Iowa soils. Soil Sci Soc Am J 58:1501–1511

    Article  Google Scholar 

  • Caro-Lopera F, Leiva V, Balakrishnan N (2012) Connection between the Hadamard and matrix products with an application to matrix-variate Birnbaum–Saunders distributions. J Multivar Anal 104:126–139

    Article  Google Scholar 

  • Chilès JP, Delfiner P (2012) Geostatistics: modeling spatial uncertainty. Wiley, New York

    Book  Google Scholar 

  • Cressie N (2015) Statistics for spatial data. Wiley, New York

    Google Scholar 

  • De Bastiani F, Cysneiros AHMA, Uribe-Opazo MA, Galea M (2015) Influence diagnostics in elliptical spatial linear models. TEST 24:322–340

    Article  Google Scholar 

  • Diggle PJ, Ribeiro PJ (2007) Model-based geoestatistics. Springer, New York

    Google Scholar 

  • Dunn P, Smyth G (1996) Randomized quantile residuals. J Comput Graph Stat 5:236–244

    Google Scholar 

  • Ferreira M, Gomes MI, Leiva V (2012) On an extreme value version of the Birnbaum–Saunders distribution. REVSTAT Stat J 10:181–210

    Google Scholar 

  • Garcia-Papani F, Uribe-Opazo MA, Leiva V, Aykroyd RG (2018) Birnbaum–Saunders spatial regression models: diagnostics and application to chemical data. Chemom Intell Lab Syst (in press)

  • Garcia-Papani F, Leiva V, Uribe-Opazo MA, Aykroyd RG (2017) Birnbaum–Saunders spatial modelling and diagnostics applied to agricultural engineering data. Stoch Environ Res Risk Assess 31:105–124

    Article  Google Scholar 

  • Goovaerts P (1997) Geostatistics for natural resources evaluation. Oxford University Press, Oxford

    Google Scholar 

  • Hengl T, Heuvelink G, Stein A (2003) Comparison of kriging with external drift and regression-kriging. In: Technical report, International Institute for Geo-information Science and Earth Observation (ITC), Enschede, The Netherlands

  • Hengl T, Heuvelink G, Stein A (2004) A generic framework for spatial prediction of soil variables based on regression-kriging. Geoderma 120:75–93

    Article  Google Scholar 

  • Hu Y, Jia Z, Cheng J, Zhao Z, Chen F (2016) Spatial variability of soil arsenic and its association with soil nitrogen in intensive farming systems. J Soils Sedim 16:169–176

    Article  CAS  Google Scholar 

  • Johnson NL, Kotz S, Balakrishnan N (1994) Continuous univariate distributions, vol 1. Wiley, New York

    Google Scholar 

  • Johnson NL, Kotz S, Balakrishnan N (1995) Continuous univariate distributions, vol 2. Wiley, New York

    Google Scholar 

  • Journel AG (1980) The lognormal approach to predicting local distributions of selective mining unit grades. J Int Assoc Math Geol 12:285–303

    Article  Google Scholar 

  • Journel AG, Huijbregts CJ (1978) Mining geostatistics. Academic Press, New York

    Google Scholar 

  • Lange K (2001) Numerical analysis for statisticians. Springer, New York

    Google Scholar 

  • Leiva V (2016) The Birnbaum–Saunders distribution. Academic Press, New York

    Google Scholar 

  • Leiva V, Athayde E, Azevedo C, Marchant C (2011) Modeling wind energy flux by a Birnbaum–Saunders distribution with unknown shift parameter. J Appl Stat 38:2819–2838

    Article  Google Scholar 

  • Leiva V, Ferreira M, Gomes MI, Lillo C (2016) Extreme value Birnbaum–Saunders regression models applied to environmental data. Stoch Environ Res Risk Assess 30:1045–1058

    Article  Google Scholar 

  • Leiva V, Marchant C, Ruggeri F, Saulo H (2015) A criterion for environmental assessment using Birnbaum–Saunders attribute control charts. Environmetrics 26:463–476

    Article  Google Scholar 

  • Leiva V, Sanhueza A, Angulo JM (2009) A length-biased version of the Birnbaum–Saunders distribution with application in water quality. Stoch Environ Res Risk Assess 23:299–307

    Article  Google Scholar 

  • Leiva V, Santos-Neto M, Cysneiros FJA, Barros M (2014) Birnbaum–Saunders statistical modelling: a new approach. Stat Modell 14:21–48

    Article  Google Scholar 

  • Leiva V, Saulo H (2017) Environmental applications based on Birnbaum–Saunders models. In: Adhikari A, Adhikari MR, Chaubey YP (eds) Mathematical and statistical applications in life sciences and engineering. Springer, Singapore, pp 283–304

    Chapter  Google Scholar 

  • Lillo C, Leiva V, Nicolis O, Aykroyd RG (2018) L-moments of the Birnbaum–Saunders distribution and its extreme value version: estimation, goodness of fit and application to earthquake data. J Appl Stat 45:187–209

    Article  Google Scholar 

  • Lopes AS (1998) International soil fertility manual. Potafos, Piracicaba (in Portuguese)

    Google Scholar 

  • Marchant C, Leiva V, Cavieres MF, Sanhueza A (2013) Air contaminant statistical distributions with application to PM10 in Santiago, Chile. Rev Environ Contam Toxicol 223:1–31

    CAS  Google Scholar 

  • Marchant C, Leiva V, Cysneiros FJA (2016a) A multivariate log-linear model for Birnbaum–Saunders distributions. IEEE Trans Reliab 65:816–827

    Article  Google Scholar 

  • Marchant C, Leiva V, Cysneiros FJA, Liu S (2018) Robust multivariate control charts based on Birnbaum–Saunders distributions. J Stat Comput Simul 88:182–202

    Article  Google Scholar 

  • Marchant C, Leiva V, Cysneiros FJA, Vivanco JF (2016b) Diagnostics in multivariate generalized Birnbaum–Saunders regression models. J Appl Stat 43:2829–2849

    Article  Google Scholar 

  • Mardia K, Marshall R (1984) Maximum likelihood estimation of models for residual covariance in spatial regression. Biometrika 71:135–146

    Article  Google Scholar 

  • Nocedal J, Wright S (1999) Numerical optimization. Springer, New York

    Book  Google Scholar 

  • Pan J, Fei Y, Foster P (2014) Case-deletion diagnostics for linear mixed models. Technometrics 56:269–281

    Article  Google Scholar 

  • Pawlowsky-Glahn V, Egozcue JJ (2016) Spatial analysis of compositional data: a historical review. J Geochem Explor 164:28–32

    Article  CAS  Google Scholar 

  • Pawlowsky-Glahn V, Egozcue JJ, Tolosana-Delgado R (2015) Modeling and analysis of compositional data. Wiley, New York

    Google Scholar 

  • Podlaski R (2008) Characterization of diameter distribution data in near-natural forests using the Birnbaum–Saunders distribution. Can J For Res 18:518–527

    Article  Google Scholar 

  • R Core Team (2016) R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria

  • Rendu JM (1979) Normal and lognormal estimation. Math Geol 11:407–422

    Article  Google Scholar 

  • Rimstad K, Omre H (2014) Skew-Gaussian random fields. Spat Stat 10:43–62

    Article  Google Scholar 

  • Santana L, Vilca F, Leiva V (2011) Influence analysis in skew-Birnbaum–Saunders regression models and applications. J Appl Stat 38:1633–1649

    Article  Google Scholar 

  • Saulo H, Leiva V, Ziegelmann FA, Marchant C (2013) A nonparametric method for estimating asymmetric densities based on skewed Birnbaum–Saunders distributions applied to environmental data. Stoch Environ Res Risk Assess 27:1479–1491

    Article  Google Scholar 

  • Severini TA (2000) Likelihood methods in statistics. Oxford University Press, Oxford

    Google Scholar 

  • Soares A (2000) Geostatistics for earth and environmental sciences. IST Press, Lisboa (in Portuguese)

    Google Scholar 

  • Tolosana-Delgado R, Pawlowsky-Glahn V (2007) Kriging regionalized positive variables revisited: sample space and scale considerations. Math Geol 39:529–558

    Article  Google Scholar 

  • Uribe-Opazo MA, Borssoi JA, Galea M (2012) Influence diagnostics in Gaussian spatial linear models. J Appl Stat 39:615–630

    Article  Google Scholar 

  • Vilca F, Sanhueza A, Leiva V, Christakos G (2010) An extended Birnbaum–Saunders model and its application in the study of environmental quality in Santiago, Chile. Stoch Environ Res Risk Assess 24:771–782

    Article  Google Scholar 

  • Webster R, Oliver M (2009) Geostatistics for environmental scientists. Wiley, Chichester

    Google Scholar 

  • Wolter KM (2007) Introduction to variance estimation. Springer, New York

    Google Scholar 

  • Xia J, Zeephongsekul P, Packer D (2011) Spatial and temporal modelling of tourist movements using semi-Markov processes. Tour Manag 51:844–851

    Article  Google Scholar 

  • Zhang H, Zimmerman DL (2005) Towards reconciling two asymptotic frameworks in spatial statistics. Biometrika 92:921–936

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank the Editors and referees for their constructive comments on an earlier version of this manuscript which resulted in this improved version. This research work was partially supported by CNPq and CAPES grants from the Brazilian government, and by FONDECYT 1160868 grant from the Chilean government.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Víctor Leiva.

Appendix: multivariate BS and log-BS distributions

Appendix: multivariate BS and log-BS distributions

A continuous random variable \(T_i > 0\) has a BS distribution with shape (\(\alpha _i>0\)) and scale (\(\varrho _i >0\)) parameters, denoted by \(T_i \sim \text{BS}(\alpha_i, \varrho_i)\), for \(i = 1, \ldots ,n\), if (see details in Leiva 2016, pp. 18–20)

$$\begin{aligned} Z_i= \frac{1}{\alpha _i}\left( \sqrt{{T_i}/{\varrho _i}}-\sqrt{{\varrho _i}/{T_i}}\right) \sim \text {N}(0,\,1), \end{aligned}$$
(16)

such that

$$\begin{aligned} T_i = \varrho _i\left( {\alpha _i Z_i}/{2}+\sqrt{\left( {\alpha _i Z_i}/{2}\right) ^2+1}\right) ^2. \end{aligned}$$

Let the random vector \(\underline{T} = (T_1,\ldots ,T_n)^\top \in {\mathbb {R}}^{n}_{+}\) have a BS\(_n\) distribution, with shape vector \(\underline{\alpha } = (\alpha _1, \ldots , \alpha _n)^\top \in {\mathbb {R}}^{n}_{+}\), scale vector \(\underline{\varrho }=(\varrho _1,\ldots ,\varrho _n)^\top \in {\mathbb {R}}^{n}_{+}\), and association matrix \({\varvec{{\varSigma }}} \in {\mathbb {R}}^{n\times n}_{+}\), with \(\text{rk}({\varvec{{\varSigma}}}) = n\) and \(\text{rk}({\varvec{{\varSigma}}})\) being the rank of the matrix \({\varvec{{\varSigma }}}\). This is denoted by

$$\begin{aligned} \underline{T} \sim \text{BS}_n(\underline{\alpha}, \underline{\varrho}, {\varvec{{\varSigma}}}). \end{aligned}$$

Note that \({\varvec{{\varSigma }}} = (\sigma _{ij})\) is the \(n\times n\) variance-covariance matrix of the random vector \(\underline{Z} = (Z_1, \ldots, Z_n)^\top \sim \text{N}_n(\underline{0}_{n\times 1}, {\varvec{{\varSigma}}})\), with \(\sigma _{ij} = 1\), if \(i=j\), whereas \(\sigma_{ij} = \text{Cor}(Z_i, Z_j)\), if \(i\ne j\), that is, the coefficient of correlation between \(Z_i\) and \(Z_j\) as given in (16), for \(i,j=1,\ldots ,n\). Thus, if \(\underline{T} \sim \text{BS}_n(\underline{\alpha}, \underline{\varrho}, {\varvec{{\varSigma}}})\), then its probability density function is given by

$$\begin{aligned} f_{\underline{T}}(\underline{t}; \underline{\alpha }, \underline{\varrho}, {\varvec{{\varSigma}}}) =& \phi_n\left( \frac{1}{\alpha_1} \left( \sqrt{\frac{t_1}{\varrho_1}} - \sqrt{\frac{\varrho_1}{t_1}}\right), \ldots, \right. \nonumber \\ &\left. \frac{1}{\alpha_n} \left( \sqrt{\frac{t_n}{\varrho_n}} - \sqrt{\frac{\varrho_n}{t_n}}\right); {\varvec{{\varSigma}}}\right) \nonumber \\ &\prod_{i=1}^n \frac{1}{2\alpha_i\varrho_i} \left( \sqrt{\frac{\varrho_i}{t_i}} + \sqrt{\frac{\varrho_i^3}{t_i^3}}\right), \end{aligned}$$
(17)

with \(\underline{t} \in {\mathbb {R}}^{n}_{+}\), where \(\phi _n\) is the n-variate N(0, 1) probability density function. If \(\underline{T} \sim {{\text {BS}}_n}(\underline{\alpha }, \underline{\varrho }, {\varvec{{\varSigma }}})\), properties hold:

(i) For \(i,\,j=1,\ldots ,n\), we have

$$\begin{aligned} {{\text {Cov}}(T_i, T_j)}\approx & {} \varrho _i\varrho _j \left( \alpha _i\alpha _j\left( 2+\frac{3}{4}\alpha _i^2\right. +\,\frac{3}{4}\alpha _j^2+\frac{9}{32} \alpha _i^2\alpha _j^2\right) \\ &\left. \times \rho _{ij}+\frac{1}{2}\alpha _i^2\alpha _j^2\rho _{ij}^2+\frac{3}{16}\alpha _i^2\alpha _j^2\rho _{ij}^3\right) ; \end{aligned}$$

(ii) For the partitions

$$\begin{aligned} \underline{T}= & {} \left( \underline{T_1}^\top ,\underline{T_2}^\top \right) ^\top ,\;\;\underline{\alpha }=\left( \underline{\alpha }_1^\top ,\underline{\alpha }_2^\top \right) ^\top ,\;\;\underline{\varrho }=\left( \underline{\varrho }_1^\top ,\underline{\varrho }_2^\top \right) ^\top ,\\ {\varvec{{\varSigma }}}= & {} \left( \begin{array}{ll}{\varvec{{\varSigma }_{\rm 11}}} &{}\quad {\varvec{{\varSigma }_{\rm 12}}}\\ {\varvec{{\varSigma }_{\rm 21}}} &{}\quad {\varvec{{\varSigma }_{\rm 22}}}\\ \end{array}\right) , \end{aligned}$$

where \(\underline{T_1}, \underline{\alpha }_1,\underline{\varrho }_1\) are \(q\times 1\) vectors, \(\underline{T_2}, \underline{\alpha }_2,\underline{\varrho }_2\) are \(p\times 1\) vectors, \({\varvec{{\varSigma }_{11}}}\) is an \(n_1\times n_1\) matrix, \({\varvec{{\varSigma }_{22}}}\) is an \(n_2\times n_2\) matrix, and \({\varvec{{\varSigma }_{12}}}={\varvec{{\varSigma }_{21}}}^\top\) is an \(n_1\times n_2\) matrix, with \(n_1+n_2=n\); then, \(\underline{T}_l \sim \text{BS}_{n_l}(\underline{\alpha }_l,\underline{\varrho}_l; {\varvec{{\varSigma}}}_{ll})\), for \(l=1,2\);

(iii) If \(c>0\), then \(c\underline{T} \sim \text{BS}_n(\underline{\alpha}, c\underline{\varrho};{\varvec{{\varSigma}}})\) and

$$\begin{aligned} \left( c\underline{T_1}^\top , \underline{T_2}^\top \right) ^\top \sim {{\text {BS}}_n}\left( \underline{\alpha }, \left( c\underline{\varrho }_1^\top ,\underline{\varrho }_2^\top \right) ^\top ;{\varvec{{\varSigma }}}\right) , \end{aligned}$$

whereas an analogous property applies to \(c \underline{T_2}\); and (iv) Considering the notation \(\underline{A}^{-1} = (1/A_i)\), where \(\underline{A} =(A_i)\) is a vector and \(A_i\) its elements, then

$$\begin{aligned} \underline{T}^{-1}= \left( \left( \underline{T_1}^{-1}\right)^\top, \left( \underline{T_2}^{-1}\right)^\top \right)^\top \sim \text{BS}_n(\underline{\alpha}, \underline{\varrho}^{-1}; {\varvec{{\varSigma}}}) \end{aligned}$$

and

$$\begin{aligned} \left( \left( \underline{T_1}^{-1}\right) ^\top, \underline{T_2}^\top \right)^\top \sim \text{BS}_n\left( \underline{\alpha}, \left( \left( \underline{\varrho}_1^{-1}\right)^\top, \underline{\varrho}_2^\top \right)^\top; {\varvec{{\varSigma}}}_1\right)^\top, \end{aligned}$$

with

$$\begin{aligned} {\varvec{{\varSigma }}}_1=\left( \begin{array}{ll} {\varvec{{\varSigma }_{\rm 11}}} &{}\quad -\,{\varvec{{\varSigma }_{\rm 12}}}\\ -\varvec{{\varSigma }}_{\rm 21} &{}\quad {\varvec{{\varSigma }_{\rm 22}}}\\ \end{array}\right) , \end{aligned}$$

which applies analogously to \(\underline{T_2}^{-1}\). Let

$$\begin{aligned} \underline{T} = (T_1, \ldots, T_n)^\top \sim \text{BS}_n(\underline{\alpha}, \underline{\varrho}, {\varvec{{\varSigma}}}), \end{aligned}$$

\(Y_i = \log (T_i)\) and \(\mu _i = \log (\varrho _i)\), for \(i=1,\ldots ,n\), with \(-\,\infty< Y_i < +\,\infty\) and \(-\,\infty<\mu _i < +\,\infty\). Then, the random vector \(\underline{Y}=(Y_1,\ldots ,Y_n)^\top\) has a log-BS\(_n\) distribution, with shape and location vectors, respectively,

$$\begin{aligned} \underline{\alpha } = (\alpha _1,\ldots ,\alpha _n)^\top ,\;\; \underline{\mu } = (\mu _1,\ldots ,\mu _n)^\top , \end{aligned}$$

and association matrix \({\varvec{{\varSigma }}}\) as given in (17). This is denoted by \(\underline{Y} \sim \text{log-BS}_n(\underline{\alpha}, \underline{\mu}, {\varvec{{\varSigma}}})\) and its probability density function is defined as

$$\begin{aligned} f_{\underline{Y}}(\underline{y}; \underline{\alpha }, \underline{\varrho }, {\varvec{{\varSigma }}} )=& {} \phi _n\left( \frac{2}{\alpha _1} \sinh \left( \frac{y_1-\mu _1}{2}\right) ,\ldots ,\right. \nonumber \\&\left. \frac{2}{\alpha _n} \sinh \left( \frac{y_n-\mu _n}{2}\right) ;{\varvec{{\varSigma }}}\right) \\&\prod _{i=1}^n \frac{1}{\alpha _i}\cosh \left( \frac{y_i-\mu _i}{2}\right) ,\;\;\underline{y} \in {\mathbb {R}}^{n},\nonumber \end{aligned}$$
(18)

where \(\phi _n\) is given in (17). Let \(\underline{Y} \sim \text{log-BS}_n(\underline{\alpha}, \underline{\mu}, {\varvec{{\varSigma}}})\). Thus, the following properties hold:

  1. (i)

    \(Y_i \sim \text{log-BS}(\alpha_i,\mu_i)\)  , for \(i=1,\ldots ,n\);

  2. (ii)

    If \(\underline{a}=(a_1,\ldots ,a_n)^\top\) is a constant vector, \(\underline{Y}+\underline{a}\sim \text{log-BS}_n(\underline{\alpha}, \underline{\mu} + \underline{a}; {\varvec{{\varSigma}}})\);

  3. (iii)

    \(\text{E}(\underline{Y}) = \underline{\mu}\);

  4. (iv)

    If \(\underline{\alpha }=\alpha \underline{1}_{n\times 1}^\top\) and \(\underline{\mu }=\underline{0}_{n\times 1}\), then \(\text{Cov}(\underline{Y}) \approx \alpha^2{\varvec{{\varSigma}}}\);

  5. (v)

    If \(\underline{Y} \sim \text{log-BS}_n(\alpha \underline{1}_{n\times 1}, \underline{\mu}; {\varvec{{\varSigma}}})\), then

    $$\begin{aligned} \frac{4}{\alpha ^2}\underline{V}^\top {\varvec{{\varSigma }}}^{-1}\underline{V} \end{aligned}$$

    has a \(\chi ^2\) distribution with n degrees of freedom, where \(\underline{V}= (V_1,\ldots ,V_n)^\top\), with

    $$\begin{aligned} V_i=\sinh \left( \frac{y_i-\underline{x}_i^\top {\underline{\varrho }}}{2}\right) ,\quad i=1,\ldots ,n. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Garcia-Papani, F., Leiva, V., Ruggeri, F. et al. Kriging with external drift in a Birnbaum–Saunders geostatistical model. Stoch Environ Res Risk Assess 32, 1517–1530 (2018). https://doi.org/10.1007/s00477-018-1546-9

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00477-018-1546-9

Keywords

Navigation