Skip to main content

Spatial Filter Versus Conventional Spatial Model Specifications: Some Comparisons

  • Chapter
  • First Online:
Book cover Non-standard Spatial Statistics and Spatial Econometrics

Part of the book series: Advances in Geographic Information Science ((AGIS,volume 1))

Abstract

Spatial statistical analysis of geographically distributed counts data has been widely undertaken for many years, with initial analyses involving log-Gaussian approximations because only the normal probability model was first adapted in an implementable form (Ripley, 1990, pp. 9–10) to handle spatial autocorrelation (SA) effects (i.e., similar values tend to cluster on a map, indicating positive self-correlation among observations). In more recent years, linear regression techniques have given way to generalized linear model techniques that account for non-normality (e.g., logistic and Poisson regression), as well as geographic dependence. In very recent years, both linear and generalized linear models have been supplemented with hierarchical Bayesian models, in part to deal with geographic regions having small counts. The objective of this chapter is to furnish a comparison of this variety of principal techniques—both frequentist and Bayesian—available for map analysis with the newly formulated spatial filtering approach.

This material is based upon work supported by the National Science Foundation under Grant No. BCS-9905213. Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author and do not necessarily reflect the views of the National Science Foundation.

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

Notes

  1. 1.

    A HGLM is a GLM (e.g., Poisson, binomial, gamma) with multiple levels. The lowest level posits the probability model for individual observations. Higher levels posit probability models for parameters (e.g., prior distributions).

  2. 2.

    A careful inspection of these data from multiple sources reveals published discrepancies: Cressie (1999), Waller and Gotway (2004) and GeoBUGS enumerate correct lists of geographic neighbors; in contrast, Clayton and Kaldor (1987), Breslow and Clayton (1993), Stern and Cressie (1999)—vis-à-vis Cressie and Guo (1987)—and Lee and Nelder (2001) have the lists of neighbors for Annandale and Tweeddale switched.

  3. 3.

    An offset variable is one whose regression coefficient is known to be, and hence is set equal to, 1.

  4. 4.

    Regardless of the context, regional aggregates with small base populations tend to yield imprecise standardized ratios, whereas regional aggregates with large base populations almost always yield significant results.

  5. 5.

    For the Scottish lip cancer data, latitude and longitude geo-coordinates were retrieved from Waller and Gotway (see http://www.sph.emory.edu/~lwaller/ch9index.htm), and then refined with an ArcView script.

  6. 6.

    The MC is a covariation-based measure that is similar to a Pearson product-moment correlation coefficient, and has an approximate range of (1/\({{{\uplambda }}_{\textrm{n}}}\), 1/\({{{\uplambda }}_{2}}\)), where \({{{\uplambda }}_{2}}\) and \({{{\uplambda }}_{\textrm{n}}}\) respectively are the second largest and smallest eigenvalues of matrix C. Its expected value is –1/(n–1). The GR is a paired comparisons type of index, is inversely related to the MC, has an approximate range of (0, 2), and has an expected value of 1.

  7. 7.

    The non-zero exponent functional form is \({{\textrm{Y}}_{\textrm{N}}} = {{\upalpha }} + {{\upbeta }}{\left[ {({{\textrm{O}}_{\textrm{i}}} + {{\updelta }})/({{\textrm{E}}_{\textrm{i}}} + {{\updelta }})} \right]^\gamma },\) which here yields rounded-off parameter estimates of \( {\hat{\updelta}}\) = 0.10 and \( {\hat{\upgamma}}\) = 0.33 [RESS = 1.48\( \times \)10–2; P(S-W) = 0.726]. Setting \({{\updelta }}\) = 0.5 yields \( {\hat{\upgamma}}\) = –0.10 (RESS = 1.71\( \times \)10–2), which is very close to 0. Setting \({{\updelta }}\) = 0.5 and executing Friendly’s SAS macro boxcox[1].sas (http://www.math.yorku.ca/SCS/sasmac/boxcox.html) yields \( {\hat{\upgamma}}\) = 0.20 (RESS = 2.56\( \times \)10–2), which also is close to 0; setting \({{\updelta }}\) = 0.1 yields \({\hat{\upgamma}}\) = 0.31, which essentially is the same result obtained with the quantile equation. As an aside, the Freeman-Tukey transformation (Cressie, 1991, p. 540) furnishes an inferior result with its translation parameter of 1 [P(S-W) = 0.061]; its optimal translation parameter estimate also is 0.5, which modestly improves its performance here [P(S-W) = 0.131].

  8. 8.

    A translation parameter is added to both the numerator and the denominator because Ei is based upon a sum of the Ois (i.e., the sums of the Eis and the Ois are equal). In the simple case of each regional expected value being calculated with a landscape-wide rate, for example: \({{\textrm{P}}_{\textrm{i}}}\frac{{\sum\limits_{{\textrm{i}} = {1}}^{\textrm{N}} {{(}{{\textrm{O}}_{\textrm{i}}} + {{\updelta )}}} }}{{\sum\limits_{{\textrm{i}} = {\textrm{1}}}^{\textrm{N}} {{{\textrm{P}}_{\textrm{i}}}} }} = {{\textrm{P}}_{\textrm{i}}}\frac{{\sum\limits_{{\textrm{i}} = {\textrm{1}}}^{\textrm{N}} {{{\textrm{O}}_{\textrm{i}}}} }}{{\sum\limits_{{\textrm{i}} = {\textrm{1}}}^{\textrm{N}} {{{\textrm{P}}_{\textrm{i}}}} }} + \frac{{{{\textrm{P}}_{\textrm{i}}}}}{{\sum\limits_{{\textrm{i}} = {\textrm{1}}}^{\textrm{N}} {{{\textrm{P}}_{\textrm{i}}}}}}\textrm{N}{\updelta} \stackrel{{P_{i} \rightarrow P_{i}}}{\longrightarrow} {\updelta}\Rightarrow \frac{{{{\textrm{O}}_{\textrm{i}}} + {{\updelta }}}}{{{{\textrm{E}}_{\textrm{i}}} + {{\updelta }}}}\), for regional “base populations” Pi in the ith of N areal units. 

  9. 9.

    A value of 0.25 for the MC tends to relate to about 5% of the variance in Y being attributable to redundant information arising from latent spatial autocorrelation, given a particular areal unit neighborhood configuration.

  10. 10.

    The Levene test statistic was used to assess homogeneity of variance across groupings because the magnitude of the numbers involved allows them to be treated as though they approximate a continuous random variable. Meanwhile, there is no reason to expect that these sets of numbers conform to normal distributions, eliminating the possibility of using the Bartlett test statistic. The R measure is described in Gelman and Rubin (1992).

  11. 11.

    These computations are based upon the Fisher information matrix.

  12. 12.

    Huffer and Wu (1998, p. 514) note that studying the multivariate behavior of MCMC parameter estimates is a rather complicated and daunting problem, and suggest examining only univariate aspects of the sampling distributions of the individual MCMC estimates (i.e., each parameter estimate separately).

  13. 13.

    Cressie employs the transformation \(\sqrt {\frac{\#\,\,\hbox{of}\,\,\hbox{lip}\,\,\hbox{cancer}\,\,\hbox{cases}}{\#\,\,\hbox{of}\,\, \hbox{males}\,\, \hbox{at}\,\, \hbox{risk}}} \)+\(\sqrt {\frac{\#\,\,\hbox{of}\,\,\hbox{lip}\,\,\hbox{cancer}\,\,\hbox{cases} + 1}{\#\,\,\hbox{of}\,\, \hbox{males}\,\, \hbox{at}\,\, \hbox{risk}}} \).

  14. 14.

    The P(S-W) values for the various models are: 0.922 for the SAR, 0.703 for the Winsorized auto-Poisson, 0.445 for the Poisson spatial filter, 0.067 for the GeoBUGS proper CAR, and 0.575 for the BUGS spatial filter specification.

  15. 15.

    PCAR denotes the proper conditional autoregressive model specification, which restricts the value of the autoregressive parameter to its feasible parameter space.

  16. 16.

    Effective degrees of freedom were calculated in BUGS as parameter estimate pD (see Spiegelhalter et al., 2002).

  17. 17.

    The P(S-W) values for the various models are: 0.401 for the SAR, 0.289 for the Winsorized auto-Poisson, 0.464 for the Poisson spatial filter, 0.092 for the GeoBUGS proper CAR, and 0.926 for the BUGS spatial filter specification.

References

  • Bartlett, M. 1947. The use of transformations, Biometrics, 3: 39–52.

    Article  Google Scholar 

  • Besag, J.E. 1974. Spatial interaction and the statistical analysis of lattice systems, Journal of the Royal Statistical Society B, 36: 192–225.

    Google Scholar 

  • Besag, J., York, J., Mollié, A. 1991. Bayesian image restoration with two applications in spatial statistics, Annals of the Institute of Statistical Mathematics, 43: 1–59.

    Article  Google Scholar 

  • Breslow, N., Clayton, D. 1993. Approximate inference in generalized linear mixed models, Journal of the American Statistical Association, 88: 9–25.

    Article  Google Scholar 

  • Casella, G. 1985. An introduction to empirical Bayes data analysis, The American Statistician, 39: 83–87.

    Article  Google Scholar 

  • Casella, G., George, E. 1992. Explaining the Gibbs sampler, The American Statistician, 46: 167–174.

    Article  Google Scholar 

  • Chinn, S. 1996. Choosing a transformation, Journal of Applied Statistics, 23: 395–404.

    Article  Google Scholar 

  • Clayton, D., Kaldor, J. 1987. Empirical Bayes estimates of age-standardized relative risks for use in disease mapping, Biometrics, 43: 671–681.

    Article  Google Scholar 

  • Clifford, P., Richardson, S., Hémon, D. 1989. Assessing the significance of the correlation between two spatial processes, Biometrics, 45: 123–134.

    Article  Google Scholar 

  • Cressie, N. 1989. Geostatistics, The American Statistician, 43: 197–202.

    Article  Google Scholar 

  • Cressie, N. 1991. Statistics for Spatial Data. New York, NY: Wiley.

    Google Scholar 

  • Cressie, N., Guo, R. 1987. Mapping variables, in Proceedings of the NCGA Conference, Computer Graphics ‘87. McLean, VA: National Computer Graphics Association, III: 521–530.

    Google Scholar 

  • de Jong, P., Sprenger, C., Van Veen, F. 1984. On extreme values of Moran’s I and Geary’s c, Geographical Analysis, 16: 17–24.

    Article  Google Scholar 

  • Dutilleul, P. 1993. Modifying the t test for assessing the correlation between two spatial processes, Biometrics, 49: 305–314.

    Article  Google Scholar 

  • Gelman, A., Rubin, D. 1992. Inference from iterative simulation using multiple sequences (with discussion), Statistical Science, 7: 457–511.

    Article  Google Scholar 

  • Getis, A., Griffith, D. 2002. Comparative spatial filtering in regression analysis, Geographical Analysis, 34: 130–140.

    Google Scholar 

  • Gilks, R., Richardson, S., Spiegelhalter, J. (eds.). 1996. Markov Chain Monte Carlo in Practice. New York, NY: Chapman & Hall.

    Google Scholar 

  • Griffith, D. 2000a. A linear regression solution to the spatial autocorrelation problem, Journal of Geographical Systems, 2: 141–156.

    Article  Google Scholar 

  • Griffith, D. 2002a. A spatial filtering specification for the auto-Poisson model, Statistics and Probability Letters, 58: 245–251.

    Article  Google Scholar 

  • Griffith, D. 2003. Spatial Autocorrelation and Spatial Filtering: Gaining Understanding Through Theory and Scientific Visualization. Berlin: Springer.

    Google Scholar 

  • Griffith, D., Haining, R. 2006. Beyond mule kicks: The Poisson distribution in geographical analysis, Geographical Analysis, 38: 123–139.

    Google Scholar 

  • Haining, R. 1990. Spatial Data Analysis in the Social and Environmental Sciences. Cambridge: Cambridge University Press.

    Google Scholar 

  • Haining, R. 1991. Bivariate correlation and spatial data, Geographical Analysis, 23: 210–227.

    Article  Google Scholar 

  • Hill, E., Allen A., Waller, L. 1999. A comparison of focused score tests and Bayesian hierarchical models for detecting spatial disease clustering, Journal of the National Institute of Public Health, 48: 102–112.

    Google Scholar 

  • Hubbell S., Ahumada, J., Condit, R., Foster, R. 2001. Local neighborhood effects on long-term survival of individual trees in a neotropical forest, Ecological Research, 16: 859–875.

    Article  Google Scholar 

  • Huffer, F., Wu, H. 1998. Markov chain Monte Carlo for autologistic regression models with application to the distribution of plant species, Biometrics, 54: 509–524

    Article  Google Scholar 

  • Kaiser, M., Cressie, N. 1997. Modeling poisson variables with positive spatial dependence, Statistics and Probability Letters, 35: 423–432.

    Article  Google Scholar 

  • Kuehl, O. 1994. Statistical Principles of Research Design and Analysis. Belmont, CA: Duxbury Press.

    Google Scholar 

  • Lee, Y., Nelder, J. 2001. Modelling and analysing correlated non-normal data, Statistical Modeling, 1: 3–16.

    Article  Google Scholar 

  • McCullagh, P., Nelder, J. 1983 (2nd ed., 1989). Generalized Linear Models, 1st ed. London: Chapman & Hall.

    Google Scholar 

  • Mollie, A. 1996. Bayesian mapping of disease. In R. Gilks, S. Richardson, D. Spiegelhalter (eds.), Markov Chain Monte Carlo in Practice. New York, NY: Chapman & Hall, pp. 359–379.

    Google Scholar 

  • Ripley, B. 1990. Gibbsian interaction models. In D.A. Griffith (ed.), Spatial Statistics: Past, Present, and Future. Ann Arbor, MI: Institute of Mathematical Geography, pp. 3–25.

    Google Scholar 

  • Snedecor, G., Cochran, W. 1967. Statistical Methods, Sixth Edition. Ames, IA: Iowa State U. Press.

    Google Scholar 

  • Spiegelhalter, D., Best, N., Carlin, B., Van der Linde, A. 2002. Bayesian measures of model complexity and fit (with discussion), Journal of Royal Statistic Society B, 64: 583–640.

    Article  Google Scholar 

  • Stern, H., Cressie, N. 1999. Inference for extremes in disease mapping. In A. Lawson, A. Biggeri, D. Bohning, E. Lesaffre, J-F. Viel, R. Bertollini (eds.), Disease Mapping and Risk Assessment for Public Health. Chichester: Wiley, pp. 63–84.

    Google Scholar 

  • Thomas, A., Best, N., Lunn, D., Arnold, R., Spiegelhalter, D. 2004. GeoBUGS User Manual, version 1.2. Accessed at http://www.mrc-bsu.cam.ac.uk/bugs/winbugs/geobugs12manual.pdf on 3/4/2005.

  • Tiefelsdorf, M., Boots, B. 1995. The exact distribution of Moran's I, Environment and Planning A, 27: 985–999.

    Article  Google Scholar 

  • Upton, G., Fingleton, B. 1989. Spatial Data Analysis by Example, vol. 2. New York, NY: Wiley.

    Google Scholar 

  • Waller, L., Gotway, C. 2004. Applied Spatial Statistics for Public Health Data. New York, NY: Wiley.

    Book  Google Scholar 

  • Wrigley, N. 1985 (reprinted in 2002 by Blackburn). Categorical Data Analysis for Geographers and Environmental Scientists. Longman: London.

    Google Scholar 

  • Yeo, I.-K., Johnson, R. 2000. A new family of power transformations to improve normality or symmetry, Biometrika, 87: 954–959.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniel A. Griffith .

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer Berlin Heidelberg

About this chapter

Cite this chapter

Griffith, D.A., Paelinck, J.H. (2011). Spatial Filter Versus Conventional Spatial Model Specifications: Some Comparisons. In: Non-standard Spatial Statistics and Spatial Econometrics. Advances in Geographic Information Science, vol 1. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-16043-1_7

Download citation

Publish with us

Policies and ethics