Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-jbqgn Total loading time: 0 Render date: 2024-06-20T07:04:49.200Z Has data issue: false hasContentIssue false

2 - Tools for Designing and Evaluating Post-Border Surveillance Systems

Published online by Cambridge University Press:  03 July 2017

Andrew P. Robinson
Affiliation:
University of Melbourne
Terry Walshe
Affiliation:
Australian Institute of Marine Science
Mark A. Burgman
Affiliation:
Imperial College London
Mike Nunn
Affiliation:
Australian Centre for International Agricultural Research
Get access
Type
Chapter
Information
Invasive Species
Risk Assessment and Management
, pp. 17 - 52
Publisher: Cambridge University Press
Print publication year: 2017

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

Acosta, H. & White, P. (2011). Atlas of biosecurity surveillance. Ministry of Agriculture and Forestry, Wellington, New Zealand. Available from www.mpi.govt.nz/mpi-surveillance-guide/atlas.pdfGoogle Scholar
Anderson, D. P., Ramsey, D. S. L., Nugent, G., et al. (2013). A novel approach to assess the probability of disease eradication from a wild-animal reservoir host. Epidemiology and Infection, 141(7), 15091521.Google Scholar
AusVet Animal Health Services. (n.d.). Analysis of complex surveillance systems. Available from http://freedom.ausvet.com.au/pmwiki/pmwiki.php?n=Freedom.FreedomTrailGoogle Scholar
Bacca, T., Lima, E. R., Picanço, M. C., Guedes, R. N. C. & Viana, J. H. M. (2006). Optimum spacing of pheromone traps for monitoring the coffee leaf miner Leucoptera coffeella. Entomologia Experimentalis et Applicata, 119(1), 3945.CrossRefGoogle Scholar
Bailey, L. L., Hines, J. E., Nichols, J. D. & MacKenzie, D. I. (2007). Sampling design trade-offs in occupancy studies with imperfect detection: Examples and software. Ecological Applications, 17(1), 281290.CrossRefGoogle ScholarPubMed
Barrett, S., Whittle, P., Mengersen, K. & Stoklosa, R. (2010). Biosecurity threats: The design of surveillance systems, based on power and risk. Environmental and Ecological Statistics, 17(4), 509519.Google Scholar
Baxter, P. W. J. & Possingham, H. P. (2011). Optimising search strategies for invasive pests: Learn before you leap. Journal of Applied Ecology, 48(1), 8695.Google Scholar
Berec, L., Kean, J. M., Epanchin-Niell, R., Liebhold, A. M. & Haight, R. G. (2015). Designing efficient surveys: Spatial arrangement of sample points for detection of invasive species. Biological Invasions, 17(1), 445459.Google Scholar
Blackwood, J., Hastings, A. & Costello, C. (2010). Cost-effective management of invasive species using linear-quadratic control. Ecological Economics, 69(3), 519527.Google Scholar
Bogich, T. L., Liebhold, A. M. & Shea, K. (2008). To sample or eradicate? A cost minimization model for monitoring and managing an invasive species. Journal of Applied Ecology, 45(4), 11341142.Google Scholar
Bogich, T. & Shea, K. (2008). A state-dependent model for the optimal management of an invasive metapopulation. Ecological Applications, 18(3), 748761.Google Scholar
Brown, J. A., Harris, S. & Timmins, S. (2004). Estimating the maximum interval between repeat surveys. Austral Ecology, 29(6), 631636.CrossRefGoogle Scholar
Buckland, S. T., Anderson, D. R., Burnham, K. P., et al. (2001). Introduction to distance sampling: Estimating abundance of biological populations. Oxford: Oxford University Press.CrossRefGoogle Scholar
Buhle, E. R., Margolis, M. & Ruesink, J. L. (2005). Bang for buck: Cost-effective control of invasive species with different life histories. Ecological Economics, 52(3), 355366.Google Scholar
Bulman, L. S., Kimberley, M. O. & Gadgil, P. D. (1999). Estimation of the efficiency of pest detection surveys. New Zealand Journal of Forestry Science, 29(1), 102115.Google Scholar
Burgman, M. A., McCarthy, M. A., Robinson, A., et al. (2013). Improving decisions for invasive species management: Reformulation and extensions of the Panetta-Lawes eradication graph. Diversity and Distributions, 19(5–6): 603607.CrossRefGoogle Scholar
Byers, J. A. (2009). Modeling distributions of flying insects: Effective attraction radius of pheromone in two and three dimensions. Journal of Theoretical Biology, 256(1), 8189.Google Scholar
Byers, J. A., Anderbrant, O. & Löfqvist, J. (1989). Effective attraction radius: A method for comparing species attractants and determining densities of flying insects. Journal of Chemical Ecology, 15(2), 749765.CrossRefGoogle ScholarPubMed
Cacho, O. J., Hester, S. & Spring, D. (2007). Applying search theory to determine the feasibility of eradicating an invasive population in natural environments. Australian Journal of Agricultural and Resource Economics, 51(4), 425443.Google Scholar
Cacho, O. J. & Pheloung, P. (2007). Weed Search. Available from www.une.edu.au/staff-profiles/business/ocachoGoogle Scholar
Cacho, O., Reeve, I., Tramell, J. & Hester, S. (2012). Valuing community engagement in biosecurity surveillance. Final Report, ACERA 1004 B 2d. Melbourne, Australia: University of Melbourne. Available from http://cebra.unimelb.edu.au/__data/assets/pdf_file/0006/1290516/1004B_OID7_Report.pdfGoogle Scholar
Cacho, O. J., Spring, D., Hester, S. & Mac Nally, R. (2010). Allocating surveillance effort in the management of invasive species: A spatially-explicit model. Environmental Modelling and Software, 25(4), 444454.Google Scholar
Cacho, O. J., Spring, D., Pheloung, P. & Hester, S. (2006). Evaluating the feasibility of eradicating an invasion. Biological Invasions, 8(4), 903917.CrossRefGoogle Scholar
Cacho, O. J., Wise, R. M., Hester, S. M. & Sinden, J. A. (2008). Bioeconomic modeling for control of weeds in natural environments. Ecological Economics, 65(3), 559568.Google Scholar
Cameron, A. R. (1999). Survey toolbox for livestock diseases – A practical manual and software package for active surveillance in developing countries. ACIAR Monograph 54. Canberra, Australia: Australian Centre for International Agricultural Research.Google Scholar
Cameron, A. R. & Baldock, F. C. (1998a). A new probability formula for surveys to substantiate freedom from disease. Preventive Veterinary Medicine, 34(1), 117.Google Scholar
Cameron, A. R. & Baldock, F. C. (1998b). Two-stage sampling in surveys to substantiate freedom from disease. Preventive Veterinary Medicine, 34(1), 1930.CrossRefGoogle ScholarPubMed
Cannon, R. M. (2001). Sense and sensitivity–designing surveys based on an imperfect test. Preventive Veterinary Medicine, 49(3–4), 141163.CrossRefGoogle Scholar
Cannon, R. M. & Roe, R. T. (1982). Livestock disease surveys: A field manual for veterinarians. Bureau of Resource Science, Department of Primary Industry. Canberra, Australia: Australian Government Publishing Service.Google Scholar
Carrasco, L. R., Baker, R., MacLeod, A., Knight, J. D. & Mumford, J. D. (2009). Optimal and robust control of invasive alien species spreading in homogeneous landscapes. Journal of the Royal Society Interface, 7(44), 529540.CrossRefGoogle ScholarPubMed
Carson, H. W., Lass, L. W. & Callihan, R. H. (1995). Detection of yellow hawkweed (Hieracium pratense) with high resolution multispectral digital imagery. Weed Technology, 9(3), 477483.Google Scholar
Chadès, I., Martin, T. G., Nicol, S., et al. (2011). General rules for managing and surveying networks of pests, diseases and endangered species. Proceedings of the National Academy of Sciences of the USA, 108(20), 83238328.CrossRefGoogle ScholarPubMed
Chen, G., Kéry, M., Zhang, J. & Ma, K. (2009). Factors affecting detection probability in plant distribution studies. Journal of Ecology, 97(6), 13831389.Google Scholar
Christy, M. T., Yackel Adams, A. A., Rodda, G. H., Savidge, J. A. & Tyrrell, C. L. (2010). Modelling detection probabilities to evaluate management and control tools for an invasive species. Journal of Applied Ecology, 47(1), 106113.CrossRefGoogle Scholar
DAFF (Department of Agriculture, Fisheries and Forestry). (2004). Report for the importation of bananas from the Philippines. Revised draft import risk analysis. Biosecurity Policy Memorandum 2004/19. Canberra, Australia: Department of Agriculture.Google Scholar
Decision Systems Laboratories. (n.d.). About GeNIe and SMILE. Available from www.bayesfusion.com/Google Scholar
D’Evelyn, S. T., Tarui, N., Burnett, K. & Roumasset, J. A. (2008). Learning-by-catching: Uncertain invasive-species populations and the value of information. Journal of Environmental Management, 89(4), 284292.Google Scholar
East, I. J., Wicks, R. M., Martin, P. A. J., et al. (2013). Use of a multi-criteria analysis framework to inform the design of risk based general surveillance systems for animal disease in Australia. Preventive Veterinary Medicine, 112(3–4), 230247.Google Scholar
Edwards, P. K. & Leung, B. (2009). Re-evaluating eradication of nuisance species: Invasion of the tunicate, Ciona intestinalis. Frontiers in Ecology and the Environment, 7(6), 326332.CrossRefGoogle Scholar
Elith, J. & Leathwick, J. R. (2009). Species distribution models: Ecological explanation and prediction across space and time. Annual Review of Ecology Evolution and Systematics, 40, 677697.Google Scholar
Epanchin-Niell, R. S., Brockerhoff, E. G., Kean, J. M. & Turner, J. A. (2014). Designing cost-efficient surveillance for early detection and control of multiple biological invaders. Ecological Applications, 24(6), 12581274.Google Scholar
Epanchin-Niell, R. S., Haight, R. G., Berec, L., Kean, J. M. & Liebhold, A. M. (2012). Optimal surveillance and eradication of invasive species in heterogeneous landscapes. Ecology Letters, 15(8), 803812.Google Scholar
Epanchin-Neill, R. S. & Hastings, A. (2010). Controlling established invaders: integrating economics and spread dynamics to determine optimal management. Ecology Letters, 13(4), 528541.CrossRefGoogle Scholar
FAO (Food and Agriculture Organization of the United Nations). (2008). Establishment of areas of low pest prevalence for fruit flies (Tephritidae). International Standards for Phytosanitary Measures Publication No. 30. Available from www.acfs.go.th/sps/downloads/ISPM_30.pdfGoogle Scholar
Field, S. A., Tyre, A. J. & Possingham, H. P. (2005). Optimizing allocation of monitoring effort under economic and observational constraints. Journal of Wildlife Management, 69(2), 473482.Google Scholar
Fitzpatrick, M. C., Preisser, E. L., Ellison, A. M. & Elkinton, J. S. (2009). Observer bias and the detection of low-density populations. Ecological Applications, 19(7), 16731679.CrossRefGoogle ScholarPubMed
Fleischer, S. J., Gaylor, M. J. & Edelson, J. V. (1985). Estimating absolute density from relative sampling of Lygus lineolaris (Heteroptera: Miridae) and selected predators in early to mid-season cotton. Environmental Entomology, 14(6), 709717.CrossRefGoogle Scholar
Fox, J. C., Buckley, Y. M., Panetta, F. D., Bourgoin, J. & Pullar, D. (2009). Surveillance protocols for management of invasive plants: Modelling Chilean needle grass (Nassella neesiana) in Australia. Diversity and Distributions, 15(4), 577589.Google Scholar
Froud, K. J., Oliver, T. M., Bingham, P. C., Flynn, A. R. & Rowswell, N. J. (2008). Passive surveillance of new exotic pests and diseases in New Zealand. In Froud, K. J., Popay, A. I. & Zydenbos (eds.), S. M. Surveillance for biosecurity: Pre-border to pest management (pp. 97110). Hastings, New Zealand: New Zealand Plant Protection Society.Google Scholar
Garrard, G. E., Bekessy, S. A., McCarthy, M. A. & Wintle, B. A. (2008). When have we looked hard enough? A novel method for setting survey effort protocols for flora surveys. Austral Ecology, 33(8), 986998.CrossRefGoogle Scholar
Garrard, G., Bekessy, S. & Wintle, B. (2009). Determining necessary survey effort to detect invasive weeds in native vegetation communities. Final Report, ACERA Project 0906. Melbourne, Australia: University of Melbourne.Google Scholar
Garrard, G. E., McCarthy, M. A., Williams, N. S. G., Bekessy, S. A. & Wintle, B. A. (2013). A general model of detectability using species traits. Methods in Ecology and Evolution, 4(1), 4552.Google Scholar
Green, R. H. & Young, R. C. (1993). Sampling to detect rare species. Ecological Applications, 3(2), 351356.Google Scholar
Guichard, S., Kriticos, D. J., Leriche, A., Kean, J. M. & Worner, S. P. (2012). Individual based modelling of moth dispersal to improve biosecurity incursion response. Journal of Applied Ecology, 49(1), 287296.Google Scholar
Harris, S., Brown, J. & Timmons, S. (2001). Weed surveillance – How often to search? Science for Conservation 175. Wellington, New Zealand: Department of Conservation.Google Scholar
Harris, S. & Timmins, S. M. (2009). Estimating the benefit of early control of all newly naturalised plants. Science for Conservation 292. Wellington, New Zealand: Department of Conservation.Google Scholar
Hastings, A., Cuddington, K., Davie, K. F., et al. (2005). The spatial spread of invasions: New developments in theory and evidence. Ecology Letters, 8(1), 91101.Google Scholar
Hastings, A., Hall, R. J. & Taylor, C. M. (2006). A simple approach to optimal control of invasive species. Theoretical Population Biology, 70(4), 431435.Google Scholar
Hauser, C. E. (2009). Where and how much? A spreadsheet that allocates surveillance effort for a weed. Plant Protection Quarterly, 24(3), 9497.Google Scholar
Hauser, C. E., Garrard, G. E. & Moore, J. L. (2015). Estimating detection rates. In Jarrad, F., Low Choy, S. & Mengersen (eds.), K. Biosecurity surveillance: Quantitative approaches (pp. 151166). Wallingford UK: CABI.Google Scholar
Hauser, C. E., Giljohann, K. M., Rigby, M., Herbert, K., Curran, I., Pascoe, C., Williams, N. S. G., Cousens, R. D. & Moore, J. L. (2016a). Practicable methods for delimiting a plant invasion. Diversity and Distributions, 22, 136147.Google Scholar
Hauser, C. E. & McCarthy, M. A. (2009). Streamlining ‘search and destroy’: cost-effective surveillance for invasive species management. Ecology Letters, 12(7), 683692.Google Scholar
Hauser, C. E., Moore, J. L., Giljohann, K. M., Garrard, G. E. & McCarthy, M. A. (2012). Designing a detection experiment: Tricks and trade-offs. In Eldershaw, V. (ed.), Proceedings of the 18th Australasian Weeds Conference, (pp. 267272). Weed Society of Victoria, October 2012.Google Scholar
Hauser, C. E., Veltheim, I., Crase, B. & Guillera-Arroita, G. (2015). Evaluation of a sniffer dog for detecting invasive hawkweeds (Hieracium spp) on the Bogong High Plains. Report to Victorian Department of Economic Development, Jobs, Transport and Resources, Melbourne, Australia.Google Scholar
Hauser, C. E., Weiss, J., Guillera-Arroita, G., McCarthy, M. A., Giljohann, K. M. & Moore, J. L. (2016b). Designing detection experiments: three more case studies. In R. Randall, S. Lloyd, & C. Borger (eds.), Proceedings of the 20th Australasian Weeds Conference. Weeds Society of Western Australia, September 2016, pp. 171–178. http://caws.org.au/awc/2016/awc201611711.pdfGoogle Scholar
Hester, S. M., Brooks, S. J., Cacho, O. J. & Panetta, F. D. (2010b). Applying a simulation model to the management of an infestation of Miconia (Miconia calvescens) in the wet tropics of Australia. Weed Research, 50(3), 269279.Google Scholar
Hester, S. M. & Cacho, O. J. (2012). Optimising search strategies in managing biological invasions: A simulation approach. Human and Ecological Risk Assessment, 18(1), 181199.Google Scholar
Hester, S. M. & Cacho, O. J. (2017). The contribution of passive surveillance to invasive species management. Biological Invasions, 19(3), 737–748.Google Scholar
Hester, S. M., Cacho, O. J., Panetta, F. D. & Hauser, C. E. (2013). Economic aspects of weed risk management. Diversity and Distributions, 19(5–6), 580589.Google Scholar
Hester, S., Hauser, C., Kean, J., Walshe, T. & Robinson, A. (2010a). Post-border surveillance techniques: Review, synthesis and deployment. Milestone Report 1, ACERA 1004. Melbourne, Australia: University of Melbourne. Available from www.acera.unimelb.edu.au/materials/endorsed/1004_final-report.pdfGoogle Scholar
Hester, S., Herbert, K. & Cook, J. (n.d.). MoniTool: A weed eradication and monitoring tool, version 3. Available from http://cebra.unimelb.edu.au/publications/acera_reports/surveillanceGoogle Scholar
Hester, S. M., Sergeant, E., Robinson, A. P. & Schultz, G. (2015). Animal, vegetable, or … ? A case study in using animal-health monitoring design tools to solve a plant-health surveillance problem. In Jarrad, F., Low-Choy, S. & Mengersen, K. (eds.), Biosecurity surveillance: Quantitative approaches (pp. 313333). Wallingford, UK: CABI.Google Scholar
Hood, G. M. (2010). PopTools, version 3.2.5. Available from www.poptools.orgGoogle Scholar
Hood, G. M., Barry, S. C. & Martin, P. A. J. (2009). Alternative methods for computing the sensitivity of complex surveillance systems. Risk Analysis, 29(2), 16861698.Google Scholar
Hung, C. & Sukkarieh, S. (2015). Using robotic aircraft and intelligent surveillance systems for orange hawkweed detection. Plant Protection Quarterly, 30(3), 100102.Google Scholar
Itami, R. & Cotter, M. (1999). Application of Analytical Hierarchy Process to rank issues, projects and sites in integrated catchment management. In Multiple objective decision support systems for managing watersheds and natural resources, 2nd International Conference (MODSS’99), Brisbane.Google Scholar
Jarrad, F., Barrett, S., Murray, J., et al. (2011). Improved design method for biosecurity surveillance and early detection of nonindigenous rats. New Zealand Journal of Ecology, 35(2), 132144.Google Scholar
Joseph, L. N., Elkin, C., Martin, T. G. & Possingham, H. P. (2009). Modeling abundance using N-mixture models: The importance of considering ecological mechanisms. Ecological Applications, 19(3), 631642.Google Scholar
Kean, J. M. (2015). The effective sampling area of traps: estimation and application. In Beresford, R. M., Froud, K. J., Kean, J. M. & Worner (eds.), S. P. The plant protection data toolbox (pp. 6776)., Christchurch, New Zealand: New Zealand Plant Protection Society.Google Scholar
Kean, J. M., Burnip, G. M. & Pathan, A. (2015). Detection survey design for decision making during biosecurity incursions. In Jarrad, F., Low-Choy, S. & Mengersen, K. (eds.), Biosecurity surveillance: Quantitative approaches (pp. 238250). Wallingford, UK: CABI.CrossRefGoogle Scholar
Kean, J. M., Phillips, C. B. & McNeill, M. R. (2008). Surveillance for early detection: Lottery or investment? In Froud, K. J., Popay, A. I. & Zydenbos (eds.), S. M. Surveillance for biosecurity: Pre-border to pest management (pp. 1117)., Hastings, New Zealand: New Zealand Plant Protection Society.Google Scholar
Kean, J. M. & Suckling, D. M. (2005). Estimating the probability of eradication of painted apple moth from Auckland. New Zealand Plant Protection, 58, 711.Google Scholar
Kean, J. M., Suckling, D. M., Sullivan, N. J. et al. (2016). Global eradication and response database. Available from http://b3.net.nz/gerdaGoogle Scholar
Kompas, T. & Che, N. (2003). A practical optimal measure: Papaya fruit fly in Australia. Draft Report to National Office of Animal and Plant Health, Australia (unpublished).Google Scholar
Kompas, T., Che, N., & Ha, P. V. (2006). An optimal surveillance measure against foot-and-mouth disease in the United States. Working Paper 06-11. Canberra, Australia: Crawford School of Economics and Government, Australian National University.Google Scholar
Krebs, C. J. & Boonstra, R. (1984). Trappability estimates for mark-recapture data. Canadian Journal of Zoology, 62(12), 24402444.Google Scholar
Lawes, R. A. & Wallace, J. F. (2008). Monitoring an invasive perennial at the landscape scale with remote sensing. Ecological Management and Restoration, 9(1), 5359.Google Scholar
Leung, B., Cacho, O. & Spring, D. (2010). Searching for non-indigenous species: rapidly delimiting the invasion boundary. Diversity and Distributions, 16(3), 451460.Google Scholar
Leung, B., Finnoff, D., Shogren, J. & Lodge, D. (2005). Managing invasive species: Rules of thumb for rapid assessment. Ecological Economics, 55(1), 2436.Google Scholar
Leung, B., Lodge, D. M., Finnoff, D., et al. (2002). An ounce of prevention or a pound of cure: Bioeconomic risk analysis of invasive species. Proceedings of the Royal Society of London B: Biological Sciences, 269(1508), 24072413.Google Scholar
MacKenzie, D. I., Nichols, J. D., Lachman, G. B., et al. (2002). Estimating site occupancy rates when detection probabilities are less than one. Ecology, 83(8), 22482255.CrossRefGoogle Scholar
MacKenzie, D. I., Nichols, J. D., Sutton, N., Kawanishi, K. & Bailey, L. L. (2005). Improving inferences in population studies of rare species that are detected imperfectly. Ecology, 86(5), 11011113.Google Scholar
MacKenzie, D. I. & Royle, J. A. (2005). Designing occupancy studies: General advice and allocating survey effort. Journal of Applied Ecology, 42(6), 11051114.Google Scholar
MAFBNZ (Ministry of Agriculture and Forestry Biosecurity New Zealand). (2009a). Review of selected cattle identification and tracing systems worldwide: Lessons for the New Zealand National Animal Identification and Tracing (NAIT) Project. MAF Biosecurity New Zealand Information Paper No 2009/03. Available from https://mpi.govt.nz/document-vault/6394Google Scholar
MAFBNZ (Ministry of Agriculture and Forestry Biosecurity New Zealand). (2009b). Biosecurity surveillance strategy 2020. Wellington, New Zealand: MAF Biosecurity New Zealand,Google Scholar
Magarey, R. D., Borchert, D. M., Engle, J. S., et al. (2011). Risk maps for targeting exotic plant pest detection programs in the United States. EPPO Bulletin, 41(1), 4656.Google Scholar
Mangano, P., Hardie, D., Speijers, J., et al. (2011). The capacity of groups within the community to carry out plant pest surveillance detection. The Open Entomology Journal, 5(1), 1523.Google Scholar
Martin, P. A. J., Cameron, A. R., Barford, K., Sergeant, E. S. G. & Greiner, M. (2007b). Demonstrating freedom from disease using multiple complex data sources 2: Case study – classical swine fever in Denmark. Preventive Veterinary Medicine, 79(2–4), 98115.CrossRefGoogle ScholarPubMed
Martin, P. A. J., Cameron, A. R. & Greiner, M. (2007a). Demonstrating freedom from disease using multiple complex data sources 1: A new methodology based on scenario trees. Preventive Veterinary Medicine, 79(2–4), 7197.Google Scholar
Martin, P. A. J., Langstaff, I., Iglesias, R. M., East, I. J., Sergeant, E. S. G. & Garner, M. G. (2015). Assessing the efficacy of general surveillance for detection of incursions of livestock diseases in Australia. Preventive Veterinary Medicine, 121(3–4), 215230.Google Scholar
McCallum, D. A. (2005). A conceptual guide to detection probability for point counts and other count-based survey methods. In USDA Forest Service General Technical Report, pp. 754–761.Google Scholar
McMaugh, T. (2005). Guidelines for surveillance for plant pests in Asia and the Pacific. ACIAR Monograph No. 119. Canberra, Australia: Australian Centre for International Agricultural Research.Google Scholar
Meats, A. (1998). Cartesian methods of locating spot infestations of the papaya fruit fly Bactrocera papayae Drew and Hancock within the trapping grid at Mareeba, Queensland, Australia. General and Applied Entomology, 28, 5760.Google Scholar
Mehta, S. V., Haight, R. G., Homans, F. R., Polasky, S. & Venette, R. C. (2007). Optimal detection and control strategies for invasive species management. Ecological Economics, 61(2–3), 237245.Google Scholar
MLA (Meat and Livestock Australia). (n.d.). National Livestock Identification System. Available from www.mla.com.au/Meat-safety-and-traceability/National-Livestock-Identification-SystemGoogle Scholar
Moody, M. E. & Mack, R. N. (1988). Controlling the spread of plant invasions: the importance of nascent foci. Journal of Applied Ecology, 25(3), 10091021.Google Scholar
Moore, J. L., Hauser, C. E., Bear, J. L., Williams, N. S. G. & McCarthy, M. A. (2011). Estimating detection-effort curves for plants using search experiments. Ecological Applications, 21(2), 601607.CrossRefGoogle ScholarPubMed
Morton, J. and Harris, W. (2008). Weed Spotters guide: A guide for regional bodies to deliver a Weed Spotters network in their regions. CRC for Australian Weed Management. Available from www.wsq.org.au/Publications/CRC%20WS%20regional%20guide_WEB.pdfGoogle Scholar
NAIT (National Animal Identification and Tracing) (n.d.). Annual report 2013. Available from www.nait.co.nz/assets/Annual-reports/Annual-Report-2013.pdfGoogle Scholar
Narumalani, S., Mishra, D. R., Wilson, R., Reece, R. & Kohler, A. (2009). Detecting and mapping four invasive species along the floodplain of North Platte River, Nebraska. Weed Technology, 23(1), 99107.Google Scholar
Ndeffo Mbah, M. L. & Gilligan, C. A. (2010). Balancing detection and eradication for control of epidemics: Sudden oak death in mixed-species stands. PLoS ONE, 5(9), e12317.Google Scholar
Nichols, J. D., Hines, J. E., Sauer, J. R., et al. (2000). A double-observer approach for estimating detection probability and abundance from point counts. The Auk, 117(2), 393408.Google Scholar
Norsys Software Corporation. (n.d.). NeticaTM application. Available from www.norsys.com/netica.htmlGoogle Scholar
OIE (World Organisation for Animal Health). (2015). Terrestrial Animal Health Code – 2015. Available from www.oie.int/eng/normes/Mcode/en_sommaire.htmGoogle Scholar
Palisade Corporation. (2015). User’s guide @RISK: Risk analysis and simulation add-in for Microsoft Excel, version 7, August 2015. Available from www.palisade.com/downloads/documentation/7/EN/RISK7_EN.pdfGoogle Scholar
Panetta, F. D., Cacho, O. J., Hester, S. & Sims-Chilton, N. (2011). Estimating and influencing the duration of weed eradication programmes. Journal of Applied Ecology, 48(4), 980988.Google Scholar
Panetta, F. D. & Lawes, R. (2005). Evaluation of weed eradication programs: the delimitation of extent. Diversity and Distributions, 11(5), 435442.Google Scholar
Panetta, F. D. & Lawes, R. (2007). Evaluation of the Australian branched broomrape (Orobanche ramose) eradication program. Weed Science, 55(6), 644651.Google Scholar
Perry, G. & Vice, D. (2009). Forecasting the risk of brown tree snake dispersal from Guam: A mixed transport-establishment model. Conservation Biology, 23(4), 9921000.Google Scholar
Peterson, J. T. & Bayley, P. B. (2004). A Bayesian approach to estimating presence when a species is undetected. In Thompson, W. L. (ed.), Sampling rare or elusive species (pp. 173188). London: Island Press.Google Scholar
Philippi, T. (2005). Adaptive cluster sampling for estimation of abundances within local populations of low-abundance plants. Ecology, 86(5), 10911100.Google Scholar
QDPIF (Queensland Department of Primary Industries and Fisheries). (2008). National fire ant eradication program progress report 2007–2008. Oxley, Australia: Queensland Department of Primary Industries and Fisheries.Google Scholar
QNRM (Queensland Department of Natural Resources and Mines) (2006). Siam weed national delimiting survey report. Brisbane, Australia: Department of Natural Resources, Mines and Water.Google Scholar
Queensland Government (n.d.) Weedspotters’ network Queensland. Available from www.qld.gov.au/environment/plants-animals/plants/herbarium/weed-spotters/Google Scholar
Ramsey, D. S. L., Parkes, J. & Morrison, S. A. (2009). Quantifying eradication success: The removal of feral pigs from Santa Cruz Island, California. Conservation Biology, 23(2), 449459.Google Scholar
Regan, T. J., McCarthy, M. A., Baxter, P. W. J., Dane Panetta, F. & Possingham, H. P. (2006). Optimal eradication: When to stop looking for an invasive plant. Ecology Letters, 9(7), 759766.Google Scholar
Rew, L. J., Maxwell, B. D., Dougher, F. L. & Aspinall, R. (2006). Searching for a needle in a haystack: Evaluating survey methods for non-indigenous plant species. Biological Invasions, 8(3), 523539.Google Scholar
Rivadeneira, M. M., Hunt, G. & Roy, K. (2009). The use of sighting records to infer species extinctions: An evaluation of different methods. Ecology, 90(5), 12911300.Google Scholar
Rota, C. T., FletcherJr, R. J., Dorazio, R. M. & Betts, M. G. (2009). Occupancy estimation and the closure assumption. Journal of Applied Ecology, 46(6), 11731181.Google Scholar
Rout, T. M., Salomon, Y. & McCarthy, M. A. (2009a). Using sighting records to declare eradication of an invasive species. Journal of Applied Ecology, 46(1), 110117.Google Scholar
Rout, T. M., Thompson, C. J. & McCarthy, M. A. (2009b). Robust decisions for declaring eradication of invasive species. Journal of Applied Ecology, 46(4), 782786.Google Scholar
Royle, J. A., Nichols, J. D., Karanth, K. U. & Gopalaswamy, A. M. (2009). A hierarchical model for estimating density in camera-trap studies. Journal of Applied Ecology, 46(1), 118127.Google Scholar
Samalens, J. C., Rossi, J. P., Guyon, D., et al. (2007) Adaptive roadside sampling for bark beetle damage assessment. Forest Ecology and Management, 253(1–3), 177187.Google Scholar
Sergeant, E. S. G. (2009). Epitools epidemiological calculators. AusVet Animal Health Services and Australian Biosecurity Cooperative Research Centre for Emerging Infectious Disease. Available from http://epitools.ausvet.com.auGoogle Scholar
Shafii, R., Price, W. J., Prather, T. S., Lass, L. W. & Thill, D. C. (2003). Predicting the likelihood of yellow starthistle (Centaurea solstitialis) occurrence using landscape characteristics. Weed Science, 51(5), 748751.Google Scholar
Sharov, A. & Liebhold, A. M. (1998). Bioeconomics of managing the spread of exotic pest species with barrier zones. Ecological Applications, 8(3), 833845.Google Scholar
Sharov, A. A., Liebhold, A. M. & Roberts, E. A. (1998). Optimizing the use of barrier zones to slow the spread of gypsy moth (Lepidoptera: Lymantriidae) in North America. Journal of Economic Entomology, 91(1), 165174.Google Scholar
Sindel, B. S., van der Meulen, A., Coleman, M. & Reeve, I. (2008). Pathway risk analysis for weed spread within Australia: Final report to Land and Water Australia (UNE61), Armidale, Australia: University of New England.Google Scholar
Smith, D. R., Villella, R. F. & LeMarié, D. P. (2003). Application of adaptive cluster sampling to low-density populations of freshwater mussels. Environmental and Ecological Statistics, 10(1), 715.Google Scholar
Smolik, M. G., Dullinger, S., Essl, F., et al. (2010). Integrating species distribution models and interacting particle systems to predict the spread of an invasive alien plant. Journal of Biogeography, 37(3), 411422.Google Scholar
Solow, A., Seymour, A., Beet, A. & Harris, S. (2008). The untamed shrew: On the termination of an eradication programme for an introduced species. Journal of Applied Ecology, 45(2), 424427.Google Scholar
Stevens, P. M. (2008). High risk site surveillance (HRSS) – an example of best practice plant pest surveillance. In Popay, I., Froud, K. & Zydenbos, S. (eds.), Surveillance for biosecurity: Pre-border to pest management (pp. 127134). Christchurch, New Zealand: New Zealand Plant Protection Society.Google Scholar
Suckling, D. M., Stringer, L. D., Kean, J. M., et al. (2015). Spatial analysis of mass trapping: How close is close enough? Pest Management Science, 71(10), 14521461.Google Scholar
Taylor, C. M. & Hastings, A. (2004). Finding optimal control strategies for invasive species: A density-structured model for Spartina alterniflora. Journal of Applied Ecology, 41(6), 10491057.Google Scholar
Thomas, L., Buckland, S. T., Rexstad, E. A., et al. (2010). Distance software: Design and analysis of distance sampling surveys for estimating population size. Journal of Applied Ecology, 47(1), 514.Google Scholar
Thomas, N., Steel, J., King, C., Hunt, T. & Weiss, J. (2007). Tackling weeds on private land initiative: Weed spread pathway risk assessment – stage 2. Melbourne, Australia: Department of Sustainability and Environment, Department of Primary Industries, Victoria.Google Scholar
Thompson, S. K. (1990). Adaptive cluster sampling. Journal of the American Statistical Association, 85(412), 10501059.Google Scholar
Tobin, P. C., Kean, J. M., Suckling, D. M., et al. (2014). Determinants of successful arthropod eradication programs. Biological Invasions, 16(2), 401414.Google Scholar
Turchin, P. & Odendaal, F. J. (1996). Measuring the effective sampling area of a pheromone trap for monitoring population density of southern pine beetle (Coleoptera: Scolytidae). Environmental Entomology, 25(3), 582588.Google Scholar
Tyre, A. J., Tenhumberg, B., Field, S. A., et al. (2003). Improving precision and reducing bias in biological surveys: Estimating false-negative error rates. Ecological Applications, 13(6), 17901801.Google Scholar
UFAW (Universities Federation for Animal Welfare). (2005). Feeding garden birds: Best practice guidelines. Available from www.ufaw.org.uk/shop/publications/product/feeding-garden-birds-best-practice-guidelinesGoogle Scholar
Václavík, T. & Meentemeyer, R. K. (2009). Invasive species distribution modeling (iSDM): Are absence data and dispersal constraints needed to predict actual distributions? Ecological Modelling, 220(23), 32483258.CrossRefGoogle Scholar
van der Kraan, C. & van Deventer, P. (1982). Range of action and interaction of pheromone traps for the summerfruit tortrix moth, Adoxophyes orana (F.v.R.). Journal of Chemical Ecology, 8(10), 12511262.Google Scholar
Wall, C. & Perry, J. N. (1978). Interactions between pheromone traps for the pea moth, Cydia nigricana (F.). Entomologia Experimentalis et Applicata, 24(2), 155162.Google Scholar
Wenger, S. J. & Freeman, M. C. (2008). Estimating species occurrence, abundance, and detection probability using zero-inflated distributions. Ecology, 89(10), 29532959.Google Scholar
Wikle, C. K. & Royle, J. A. (1999). Space-time dynamic design of environmental monitoring networks. Journal of Agricultural, Biological, and Environmental Statistics, 4(4), 489507.Google Scholar
Williams, N. S. G., Hahs, A. K. & Morgan, J. W. (2008). A dispersal-constrained habitat suitability model for predicting invasion of alpine vegetation. Ecological Applications, 18(2), 347359.Google Scholar
Wintle, B. A., McCarthy, M. A., Parris, K. M. & Burgman, M. A. (2004). Precision and bias of methods for estimating point survey detection probabilities. Ecological Applications, 14(3), 703712.Google Scholar
Yackel Adams, A. A., Stanford, J. W., Wiewel, A. S. & Rodda, G. H. (2011). Modelling detectability of kiore (Rattus exulans) on Aguiguan, Mariana Islands, to inform possible eradication and monitoring efforts. New Zealand Journal of Ecology, 35(2), 145152.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×