Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-16T03:29:50.573Z Has data issue: false hasContentIssue false

Bayesian geostatistical modelling for mapping schistosomiasis transmission

Published online by Cambridge University Press:  02 June 2009

P. VOUNATSOU*
Affiliation:
Department of Public Health and Epidemiology, Swiss Tropical Institute, P.O. Box, CH-4002 Basel, Switzerland
G. RASO
Affiliation:
Division of Epidemiology and Social Medicine, School of Population Health, The University of Queensland, Public Health Building, Herston Road, Brisbane, Queensland 4006, Australia Molecular Parasitology Laboratory, Queensland Institute of Medical Research, 300 Herston Road, Brisbane, Queensland 4006, Australia
M. TANNER
Affiliation:
Department of Public Health and Epidemiology, Swiss Tropical Institute, P.O. Box, CH-4002 Basel, Switzerland
E. K. N'GORAN
Affiliation:
UFR Biosciences, Université de Cocody-Abidjan, 22 BP 582, Abidjan 22, Côte d'Ivoire Centre Suisse de Recherches Scientifiques, 01 BP 1303, Abidjan 01, Côte d'Ivoire
J. UTZINGER
Affiliation:
Department of Public Health and Epidemiology, Swiss Tropical Institute, P.O. Box, CH-4002 Basel, Switzerland
*
*Corresponding author: Penelope Vounatsou, Department of Public Health and Epidemiology, Swiss Tropical Institute, P.O. Box, CH-4002 Basel, Switzerland. Tel: +41 61 284-8109. Fax: +41 61 284-8105. E-mail: penelope.vounatsou@unibas.ch

Summary

Progress has been made in mapping and predicting the risk of schistosomiasis using Bayesian geostatistical inference. Applications primarily focused on risk profiling of prevalence rather than infection intensity, although the latter is particularly important for morbidity control. In this review, the underlying assumptions used in a study mapping Schistosoma mansoni infection intensity in East Africa are examined. We argue that the assumption of stationarity needs to be relaxed, and that the negative binomial assumption might result in misleading inference because of a high number of excess zeros (individuals without an infection). We developed a Bayesian geostatistical zero-inflated (ZI) regression model that assumes a non-stationary spatial process. Our model is validated with a high-quality georeferenced database from western Côte d'Ivoire, consisting of demographic, environmental, parasitological and socio-economic data. Nearly 40% of the 3818 participating schoolchildren were infected with S. mansoni, and the mean egg count among infected children was 162 eggs per gram of stool (EPG), ranging between 24 and 6768 EPG. Compared to a negative binomial and ZI Poisson and negative binomial models, the Bayesian non-stationary ZI negative binomial model showed a better fit to the data. We conclude that geostatistical ZI models produce more accurate maps of helminth infection intensity than the spatial negative binomial ones.

Type
SECTION 1 ADVOCACY AND DEFINING AREAS IN NEED OF CONTROL
Copyright
Copyright © Cambridge University Press 2009

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Agarwal, D. K., Gelfand, A. E. and Citron-Pousty, S. (2002). Zero-inflated models with application to spatial count data. Environmental and Ecological Statistics 9, 341355.CrossRefGoogle Scholar
Anderson, R. M. and May, R. M. (1985). Helminth infections of humans: mathematical models, population dynamics and control. Advances in Parasitology 24, 1101.Google Scholar
Banerjee, S., Gelfand, A. E., Knight, J. R. and Sirmans, C. F. (2004). Spatial modeling of house prices using normalized distance-weighted sums of stationary processes. Journal of Business and Economic Statistics 22, 206213.CrossRefGoogle Scholar
Beck-Wörner, C., Raso, G., Vounatsou, P., N'Goran, E. K., Rigo, G., Parlow, E. and Utzinger, J. (2007). Bayesian spatial risk prediction of Schistosoma mansoni infection in western Côte d'Ivoire using a remotely-sensed digital elevation model. American Journal of Tropical Medicine and Hygiene 76, 956963.Google Scholar
Bergquist, R., Johansen, M. V. and Utzinger, J. (2009). Diagnostic dilemmas in helminthology: what tools to use and when? Trends in Parasitology 25, 151156.Google Scholar
Booth, M., Vennervald, B. J., Kenty, L., Butterworth, A. E., Kariuki, H. C., Kadzo, H., Ireri, E., Amaganga, C., Kimani, G., Mwatha, J. K., Otedo, A., Ouma, J. H., Muchiri, E. and Dunne, D. W. (2004). Micro-geographical variation in exposure to Schistosoma mansoni and malaria, and exacerbation of splenomegaly in Kenyan school-aged children. BMC Infectious Diseases 4, 13.Google Scholar
Booth, M., Vounatsou, P., N'Goran, E. K., Tanner, M. and Utzinger, J. (2003). The influence of sampling effort and the performance of the Kato-Katz technique in diagnosing Schistosoma mansoni and hookworm co-infections in rural Côte d'Ivoire. Parasitology 127, 525531.CrossRefGoogle ScholarPubMed
Bradley, D. J. (1972). Regulation of parasite populations. A general theory of the epidemiology and control of parasitic infections. Transactions of the Royal Society of Tropical Medicine and Hygiene 66, 697708.Google Scholar
Brooker, S. (2007). Spatial epidemiology of human schistosomiasis in Africa: risk models, transmission dynamics and control. Transactions of the Royal Society of Tropical Medicine and Hygiene 101, 18.CrossRefGoogle ScholarPubMed
Brooker, S., Alexander, N., Geiger, S., Moyeed, R. A., Stander, J., Fleming, F., Hotez, P. J., Correa-Oliveira, R. and Bethony, J. (2006). Contrasting patterns in the small-scale heterogeneity of human helminth infections in urban and rural environments in Brazil. International Journal for Parasitology 36, 11431151.Google Scholar
Brooker, S. and Clements, A. C. A. (2009). Spatial heterogeneity of parasite co-infection: determinants and geostatistical prediction at regional scales. International Journal for Parasitology 39, 591597.Google Scholar
Brooker, S., Gyapong, J. O., Kabatereine, N. B., Stothard, J. R. and Utzinger, J. (2009). Rapid assessment of schistosomiasis and other neglected tropical diseases in the context of integrated control programmes in Africa. Parasitology 136 (in press).Google Scholar
Brooker, S., Kabatereine, N. B., Fleming, F. and Devlin, N. (2008). Cost and cost-effectiveness of nationwide school-based helminth control in Uganda: intra-country variation and effects of scaling-up. Health Policy and Planning 23, 2435.CrossRefGoogle ScholarPubMed
Brooker, S., Rowlands, M., Haller, L., Savioli, L. and Bundy, D. A. P. (2000). Towards an atlas of human helminth infection in sub-Saharan Africa: the use of geographical information systems (GIS). Parasitology Today 16, 303307.CrossRefGoogle ScholarPubMed
Brooker, S. and Utzinger, J. (2007). Integrated disease mapping in a polyparasitic world. Geospatial Health 1, 141146.Google Scholar
Cameron, A. C. and Trivedi, P. K. (1986). Econometric models based on count data: comparisons and applications of some estimators. Journal of Applied Economics 1, 2953.CrossRefGoogle Scholar
Clements, A. C. A., Brooker, S., Nyandindi, U., Fenwick, A. and Blair, L. (2008). Bayesian spatial analysis of a national urinary schistosomiasis questionnaire to assist geographic targeting of schistosomiasis control in Tanzania, East Africa. International Journal for Parasitology 38, 401415.CrossRefGoogle ScholarPubMed
Clements, A. C. A., Lwambo, N. J. S., Blair, L., Nyandindi, U., Kaatano, G., Kinung'hi, S., Webster, J. P., Fenwick, A. and Brooker, S. (2006 a). Bayesian spatial analysis and disease mapping: tools to enhance planning and implementation of a schistosomiasis control programme in Tanzania. Tropical Medicine and International Health 11, 490503.CrossRefGoogle ScholarPubMed
Clements, A. C. A., Moyeed, R. and Brooker, S. (2006 b). Bayesian geostatistical prediction of the intensity of infection with Schistosoma mansoni in East Africa. Parasitology 133, 711719.Google Scholar
Clennon, J. A., King, C. H., Muchiri, E. M. and Kitron, U. (2007). Hydrological modelling of snail dispersal patterns in Msambweni, Kenya and potential resurgence of Schistosoma haematobium transmission. Parasitology 134, 683693.Google Scholar
Cohen, J. E. (1977). Mathematical models of schistosomiasis. Annual Review of Ecology and Systematics 8, 209233.CrossRefGoogle Scholar
Cressie, N. (1991). Statistics for Spatial Data. Wiley & Sons, New York-Chichester-Toronto-Brisbane-Singapore.Google Scholar
Denwood, M. J., Stear, M. J., Matthews, L., Reid, S. W. J., Toft, N. and Innocent, G. T. (2008). The distribution of the pathogenic nematode Nematodirus battus in lambs is zero-inflated. Parasitology 135, 12251235.Google Scholar
de Vlas, S. J. and Gryseels, B. (1992). Underestimation of Schistosoma mansoni prevalences. Parasitology Today 8, 274277.Google Scholar
Diggle, P. J., Moyeed, R. A. and Tawn, J. A. (1998). Model-based geostatistics. Journal of the Royal Statistical Society. Series C, Applied Statistics 47, 299326.CrossRefGoogle Scholar
Dimatteo, I., Genovese, C. R. and Kass, R. E. (2001). Bayesian curve-fitting with free-knot splines. Biometrika 88, 10551071.CrossRefGoogle Scholar
Doumenge, J. P., Mott, K. E., Cheung, C., Villenave, D., Chapuis, O., Perrin, M. F. and Reaud-Thomas, G. (1987). Atlas of the global distribution of schistosomiasis. World Health Organization and Presses Universitaires de Bordeaux (WHO-CEGET-CNRS), Geneva.Google Scholar
Engels, D., Sinzinkayo, E. and Gryseels, B. (1996). Day-to-day egg count fluctuation in Schistosoma mansoni infection and its operational implications. American Journal of Tropical Medicine and Hygiene 54, 319324.CrossRefGoogle ScholarPubMed
Eubank, R. L. (1988). Spline Smoothing and Nonparametric Regression. Decker, New York.Google Scholar
Gelfand, A. E. and Smith, A. F. M. (1990). Sampling based approaches to calculating marginal densities. Journal of the American Statistical Association 85, 398409.CrossRefGoogle Scholar
Gemperli, A., Vounatsou, P., Kleinschmidt, I., Bagayoko, M., Lengeler, C. and Smith, T. (2004). Spatial patterns of infant mortality in Mali; the effect of malaria endemicity. American Journal of Epidemiology 159, 6472.Google Scholar
Gosoniu, L., Vounatsou, P., Sogoba, N., Maire, N. and Smith, T. (2009). Mapping malaria risk in West Africa using a Bayesian nonparametric non-stationary models. Communications in Statistics – Simulation and Computation, doi: 10.1016/j.csda.2009.02.022Google Scholar
Gosoniu, L., Vounatsou, P., Sogoba, N. and Smith, T. (2006). Bayesian modelling of geostatistical malaria risk data. Geospatial Health 1, 127139.Google Scholar
Heilbron, D. (1994). Zero-altered and other regression models for count data with added zeros. Biometrical Journal 36, 531547.Google Scholar
Katz, N., Chaves, A. and Pellegrino, J. (1972). A simple device for quantitative stool thick-smear technique in schistosomiasis mansoni. Revista do Instituto de Medicina Tropical de São Paulo 14, 397400.Google Scholar
Kitron, U., Clennon, J. A., Cecere, M. C., Gürtler, R. E., King, C. H. and Vazquez-Prokopec, G. (2006). Upscale or downscale: applications of fine scale remotely sensed data to Chagas disease in Argentina and schistosomiasis in Kenya. Geospatial Health 1, 4958.CrossRefGoogle ScholarPubMed
Lambert, D. (1992). Zero-inflated Poisson regression with an application to defects in manufacturing. Technometrics 34, 114.Google Scholar
Li, Y. S., Raso, G., Zhao, Z. Y., He, Y. K., Ellis, M. K. and McManus, D. P. (2007). Large water management projects and schistosomiasis control, Dongting Lake region, China. Emerging Infectious Diseases 13, 973979.Google Scholar
Lunn, D. J., Thomas, A., Best, N. and Spiegelhalter, D. (2000). WinBUGS – a Bayesian modelling framework: concepts, structure, and extensibility. Statistics and Computing 10, 325337.Google Scholar
Majumdar, A. and Gelfand, A. E. (2007). Multivariate spatial modeling for geostatistical data using convolved covariance functions. Mathematical Geology 39, 225245.Google Scholar
Mullahy, J. (1986). Specification and testing of some modified count data models. Journal of Econometrics 33, 341365.Google Scholar
Polderman, A. M. (1979). Transmission dynamics of endemic schistosomiasis. Tropical and Geographical Medicine 31, 465475.Google Scholar
Raso, G., Luginbühl, A., Adjoua, C. A., Tian-Bi, N. T., Silué, K. D., Matthys, B., Vounatsou, P., Wang, Y., Dumas, M. E., Holmes, E., Singer, B. H., Tanner, M., N'Goran, E. K. and Utzinger, J. (2004). Multiple parasite infections and their relationship to self-reported morbidity in a community of rural Côte d'Ivoire. International Journal of Epidemiology 33, 10921102.Google Scholar
Raso, G., Matthys, B., N'Goran, E. K., Tanner, M., Vounatsou, P. and Utzinger, J. (2005). Spatial risk prediction and mapping of Schistosoma mansoni infections among schoolchildren living in western Côte d'Ivoire. Parasitology 131, 97108.CrossRefGoogle ScholarPubMed
Raso, G., Vounatsou, P., Gosoniou, L., Tanner, M., N'Goran, E. K. and Utzinger, J. (2006 b). Risk factors and spatial patterns of hookworm infection among schoolchildren in a rural area of western Côte d'Ivoire. International Journal for Parasitology 36, 201210.Google Scholar
Raso, G., Vounatsou, P., McManus, D. P. and Utzinger, J. (2007). Bayesian risk maps for Schistosoma mansoni and hookworm mono-infections in a setting where both parasites co-exist. Geospatial Health 2, 8596.Google Scholar
Raso, G., Vounatsou, P., Singer, B. H., N'Goran, E. K., Tanner, M. and Utzinger, J. (2006 a). An integrated approach for risk profiling and spatial prediction of Schistosoma mansoni-hookworm coinfection. Proceedings of the National Academy of Sciences, USA 103, 69346939.Google Scholar
Rodrigues, J. (2003). Bayesian analysis of zero-inflated distributions. Communication in Statistics-Theory and Methods 32, 281289.Google Scholar
Scott, J. T., Diakhaté, M., Vereecken, K., Fall, A., Diop, M., Ly, A., De Clercq, D., de Vlas, S. J., Berkvens, D., Kestens, L. and Gryseels, B. (2003). Human water contacts patterns in Schistosoma mansoni epidemic foci in northern Senegal change according to age, sex and place of residence, but are not related to intensity of infection. Tropical Medicine and International Health 8, 100108.Google Scholar
Silué, K. D., Raso, G., Yapi, A., Vounatsou, P., Tanner, M., N'Goran, E. K. and Utzinger, J. (2008). Spatially-explicit risk profiling of Plasmodium falciparum infections at a small scale: a geostatistical modelling approach. Malaria Journal 7, 111.Google Scholar
Sogoba, N., Vounatsou, P., Bagayoko, M. M., Doumbia, S., Dolo, G., Gosoniu, L., Traoré, S. F., Touré, Y. T. and Smith, T. (2007). The spatial distribution of Anopheles gambiae sensu stricto and An. arabiensis (Diptera: Culicidae) in Mali. Geospatial Health 1, 213222.CrossRefGoogle Scholar
Steinmann, P., Du, Z. W., Wang, L. B., Wang, X. Z., Jiang, J. Y., Li, L. H., Marti, H., Zhou, X. N. and Utzinger, J. (2008). Extensive multiparasitism in a village of Yunnan province, People's Republic of China, revealed by a suite of diagnostic methods. American Journal of Tropical Medicine and Hygiene 78, 760769.Google Scholar
Steinmann, P., Keiser, J., Bos, R., Tanner, M. and Utzinger, J. (2006). Schistosomiasis and water resources development: systematic review, meta-analysis, and estimates of people at risk. Lancet Infectious Diseases 6, 411425.CrossRefGoogle ScholarPubMed
Stensgaard, A. S., Jørgensen, A., Kabatereine, N. B., Rahbek, C. and Kristensen, T. K. (2006). Modeling freshwater snail habitat suitability and areas of potential snail-borne disease transmission in Uganda. Geospatial Health 1, 93104.CrossRefGoogle ScholarPubMed
Sutherst, R. W. (2004). Global change and human vulnerability to vector-borne diseases. Clinical Microbiology Reviews 17, 136173.Google Scholar
Utzinger, J., Booth, M., N'Goran, E. K., Müller, I., Tanner, M. and Lengeler, C. (2001). Relative contribution of day-to-day and intra-specimen variation in faecal egg counts of Schistosoma mansoni before and after treatment with praziquantel. Parasitology 122, 537544.Google Scholar
Utzinger, J. and de Savigny, D. (2006). Control of neglected tropical diseases: integrated chemotherapy and beyond. PLoS Medicine 3, e112.Google Scholar
Utzinger, J., N'Goran, E. K., Ossey, Y. A., Booth, M., Traoré, M., Lohourignon, K. L., Allangba, A., Ahiba, L. A., Tanner, M. and Lengeler, C. (2000). Rapid screening for Schistosoma mansoni in western Côte d'Ivoire using a simple school questionnaire. Bulletin of the World Health Organization 78, 389398.Google Scholar
Utzinger, J., Vounatsou, P., N'Goran, E. K., Tanner, M. and Booth, M. (2002). Reduction in the prevalence and intensity of hookworm infections after praziquantel treatment for schistosomiasis infection. International Journal for Parasitology 32, 759765.Google Scholar
WHO (2002). Prevention and control of schistosomiasis and soil-transmitted helminthiasis: report of a WHO expert committee. WHO Technical Report Series 912, 157.Google Scholar
Yang, G. J., Vounatsou, P., Zhou, X. N., Tanner, M. and Utzinger, J. (2005 a). A Bayesian-based approach for spatio-temporal modeling of county level Schistosoma japonicum prevalence data in Jiangsu province, China. International Journal for Parasitology 35, 155162.Google Scholar
Yang, G. J., Vounatsou, P., Zhou, X. N., Tanner, M. and Utzinger, J. (2005 b). A potential impact of climate change and water resource development on the transmission of Schistosoma japonicum in China. Parassitologia 47, 127134.Google ScholarPubMed
Yapi, Y. G., Briët, O. J. T., Diabate, S., Vounatsou, P., Akodo, E., Tanner, M. and Teuscher, T. (2005). Rice irrigation and schistosomiasis in savannah and forest areas of Côte d'Ivoire. Acta Tropica 93, 201211.Google Scholar
Zhou, X. N., Yang, G. J., Yang, K., Wang, X. H., Hong, Q. B., Sun, L. P., Malone, J. B., Kristensen, T. K., Bergquist, N. R. and Utzinger, J. (2008). Potential impact of climate change on schistosomiasis transmission in China. American Journal of Tropical Medicine and Hygiene 78, 188194.Google Scholar