var bibbase_data = {"data":"\"Loading..\"\n\n
\n\n \n\n \n\n \n \n\n \n\n \n \n\n \n\n \n
\n generated by\n \n \"bibbase.org\"\n\n \n
\n \n\n
\n\n \n\n\n
\n\n Excellent! Next you can\n create a new website with this list, or\n embed it in an existing web page by copying & pasting\n any of the following snippets.\n\n
\n JavaScript\n (easiest)\n
\n \n <script src=\"https://bibbase.org/show?bib=https%3A%2F%2Fapi.zotero.org%2Fusers%2F6914984%2Fcollections%2F57VGWSMY%2Fitems%3Fkey%3D5MgQbhg8lOqLRgZwglJ47R41%26format%3Dbibtex%26limit%3D100&jsonp=1&jsonp=1\"></script>\n \n
\n\n PHP\n
\n \n <?php\n $contents = file_get_contents(\"https://bibbase.org/show?bib=https%3A%2F%2Fapi.zotero.org%2Fusers%2F6914984%2Fcollections%2F57VGWSMY%2Fitems%3Fkey%3D5MgQbhg8lOqLRgZwglJ47R41%26format%3Dbibtex%26limit%3D100&jsonp=1\");\n print_r($contents);\n ?>\n \n
\n\n iFrame\n (not recommended)\n
\n \n <iframe src=\"https://bibbase.org/show?bib=https%3A%2F%2Fapi.zotero.org%2Fusers%2F6914984%2Fcollections%2F57VGWSMY%2Fitems%3Fkey%3D5MgQbhg8lOqLRgZwglJ47R41%26format%3Dbibtex%26limit%3D100&jsonp=1\"></iframe>\n \n
\n\n

\n For more details see the documention.\n

\n
\n
\n\n
\n\n This is a preview! To use this list on your own web site\n or create a new web site from it,\n create a free account. The file will be added\n and you will be able to edit it in the File Manager.\n We will show you instructions once you've created your account.\n
\n\n
\n\n

To the site owner:

\n\n

Action required! Mendeley is changing its\n API. In order to keep using Mendeley with BibBase past April\n 14th, you need to:\n

    \n
  1. renew the authorization for BibBase on Mendeley, and
  2. \n
  3. update the BibBase URL\n in your page the same way you did when you initially set up\n this page.\n
  4. \n
\n

\n\n

\n \n \n Fix it now\n

\n
\n\n
\n\n\n
\n \n \n
\n
\n  \n 2023\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Climate change impacts on bird migration and highly pathogenic avian influenza.\n \n \n \n \n\n\n \n Prosser, D. J; Teitelbaum, C. S.; Yin, S.; Hill, N. J.; and Xiao, X.\n\n\n \n\n\n\n Nature Microbiology, 8: 2223–2225. November 2023.\n \n\n\n\n
\n\n\n\n \n \n \"ClimatePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 8 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{prosser_climate_2023,\n\ttitle = {Climate change impacts on bird migration and highly pathogenic avian influenza},\n\tvolume = {8},\n\turl = {https://www.nature.com/articles/s41564-023-01538-0},\n\tdoi = {https://doi.org/10.1038/s41564-023-01538-0},\n\tabstract = {The unprecedented extent of highly pathogenic avian influenza coincides with intensifying global climate changes that alter host ecology and physiology, and could impact virus evolution and dynamics.},\n\tlanguage = {en},\n\tjournal = {Nature Microbiology},\n\tauthor = {Prosser, Diann J and Teitelbaum, Claire S. and Yin, Shenglai and Hill, Nichola J. and Xiao, Xiangming},\n\tmonth = nov,\n\tyear = {2023},\n\tpages = {2223--2225},\n}\n\n
\n
\n\n\n
\n The unprecedented extent of highly pathogenic avian influenza coincides with intensifying global climate changes that alter host ecology and physiology, and could impact virus evolution and dynamics.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Ecogeographic drivers of the spatial spread of highly pathogenic avian influenza outbreaks in Europe and North America, 2016 - early 2022.\n \n \n \n \n\n\n \n Gass, J.D.; Hill, N.J.; Damodaran, L.; Naumova, E.N.; Nutter, F.B.; and Jonathan A Runstadler\n\n\n \n\n\n\n International Journal of Environmental Research and Public Health, 20: 6030. May 2023.\n \n\n\n\n
\n\n\n\n \n \n \"EcogeographicPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 8 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{gass_jd_ecogeographic_2023,\n\ttitle = {Ecogeographic drivers of the spatial spread of highly pathogenic avian influenza outbreaks in {Europe} and {North} {America}, 2016 - early 2022},\n\tvolume = {20},\n\turl = {https://www.mdpi.com/1660-4601/20/11/6030},\n\tdoi = {https://doi.org/10.3390/ijerph20116030},\n\tabstract = {H5Nx highly pathogenic avian influenza (HPAI) viruses of clade 2.3.4.4 have caused outbreaks in Europe among wild and domestic birds since 2016 and were introduced to North America via wild migratory birds in December 2021. We examined the spatiotemporal extent of HPAI\nviruses across continents and characterized ecological and environmental predictors of virus spread between geographic regions by constructing a Bayesian phylodynamic generalized linear model (phylodynamic-GLM). The findings demonstrate localized epidemics of H5Nx throughout Europe in the first several years of the epizootic, followed by a singular branching point where H5N1 viruses were introduced to North America, likely via stopover locations throughout the North Atlantic. Once in the United States (US), H5Nx viruses spread at a greater rate between US-based regions as compared to prior spread in Europe. We established that geographic proximity is a predictor of virus spread between regions, implying that intercontinental transport across the Atlantic Ocean is relatively rare. An increase in mean ambient temperature over time was predictive of reduced H5Nx\nvirus spread, which may reflect the effect of climate change on declines in host species abundance, decreased persistence of the virus in the environment, or changes in migratory patterns due to ecological alterations. Our data provide new knowledge about the spread and directionality of H5Nx virus dispersal in Europe and the US during an actively evolving intercontinental outbreak,\nincluding predictors of virus movement between regions, which will contribute to surveillance and\nmitigation strategies as the outbreak unfolds, and in future instances of uncontained avian spread of HPAI viruses.},\n\tjournal = {International Journal of Environmental Research and Public Health},\n\tauthor = {{Gass, J.D.} and {Hill, N.J.} and {Damodaran, L.} and {Naumova, E.N.} and {Nutter, F.B.} and {Jonathan A Runstadler}},\n\tmonth = may,\n\tyear = {2023},\n\tpages = {6030},\n}\n\n
\n
\n\n\n
\n H5Nx highly pathogenic avian influenza (HPAI) viruses of clade 2.3.4.4 have caused outbreaks in Europe among wild and domestic birds since 2016 and were introduced to North America via wild migratory birds in December 2021. We examined the spatiotemporal extent of HPAI viruses across continents and characterized ecological and environmental predictors of virus spread between geographic regions by constructing a Bayesian phylodynamic generalized linear model (phylodynamic-GLM). The findings demonstrate localized epidemics of H5Nx throughout Europe in the first several years of the epizootic, followed by a singular branching point where H5N1 viruses were introduced to North America, likely via stopover locations throughout the North Atlantic. Once in the United States (US), H5Nx viruses spread at a greater rate between US-based regions as compared to prior spread in Europe. We established that geographic proximity is a predictor of virus spread between regions, implying that intercontinental transport across the Atlantic Ocean is relatively rare. An increase in mean ambient temperature over time was predictive of reduced H5Nx virus spread, which may reflect the effect of climate change on declines in host species abundance, decreased persistence of the virus in the environment, or changes in migratory patterns due to ecological alterations. Our data provide new knowledge about the spread and directionality of H5Nx virus dispersal in Europe and the US during an actively evolving intercontinental outbreak, including predictors of virus movement between regions, which will contribute to surveillance and mitigation strategies as the outbreak unfolds, and in future instances of uncontained avian spread of HPAI viruses.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Highly pathogenic avian influenza A(H5N1) virus outbreak in New England seals, United States.\n \n \n \n \n\n\n \n Puryear, W.; Sawatzki, K.; Hill, N.; Foss, A.; Stone, J. J.; Doughty, L.; Walk, D.; Gilbert, K.; Murray, M.; Cox, E.; Patel, P.; Mertz, Z.; Ellis, S.; Taylor, J.; Fauquier, D.; Smith, A.; DiGiovanni, R. A.; van de Guchte, A.; Gonzalez-Reiche, A. S.; Khalil, Z.; van Bakel, H.; Torchetti, M. K.; Lenoch, J. B.; Lantz, K.; and Runstadler, J.\n\n\n \n\n\n\n Emerging Infectious Diseases. April 2023.\n \n\n\n\n
\n\n\n\n \n \n \"HighlyPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 4 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{puryear_highly_2023,\n\ttitle = {Highly pathogenic avian influenza {A}({H5N1}) virus outbreak in {New} {England} seals, {United} {States}},\n\turl = {https://wwwnc.cdc.gov/eid/article/29/4/22-1538_article},\n\tdoi = {https://doi.org/10.3201/eid2904.221538},\n\tabstract = {The recent incursion of Highly Pathogenic Avian Influenza A (H5N1) virus into North America and subsequent dissemination of virus across the continent, has had significant adverse impacts on domestic poultry, and has led to widespread mortality in many wild bird species. Here we report the recent spillover of H5N1 into marine mammals in the northeastern United States, with associated mortality on a regional scale. This spillover is coincident with a second wave of H5N1 in sympatric wild birds also experiencing regional mortality events. Viral sequences derived from both seal and avian hosts reveal distinct viral genetic differences between the two waves of infection. Spillover into seals was closely related to virus from the second wave, and one of eight seal-derived sequences had the mammalian adaptation PB2 E627K.},\n\tlanguage = {en},\n\turldate = {2023-02-03},\n\tjournal = {Emerging Infectious Diseases},\n\tauthor = {Puryear, Wendy and Sawatzki, Kaitlin and Hill, Nichola and Foss, Alexa and Stone, Jonathon J. and Doughty, Lynda and Walk, Dominique and Gilbert, Katie and Murray, Maureen and Cox, Elena and Patel, Priya and Mertz, Zak and Ellis, Stephanie and Taylor, Jennifer and Fauquier, Deborah and Smith, Ainsley and DiGiovanni, Robert A. and van de Guchte, Adriana and Gonzalez-Reiche, Ana Silvia and Khalil, Zain and van Bakel, Harm and Torchetti, Mia K. and Lenoch, Julianna B. and Lantz, Kristina and Runstadler, Jonathan},\n\tmonth = apr,\n\tyear = {2023},\n\tdoi = {10.1101/2022.07.29.501155},\n}\n\n
\n
\n\n\n
\n The recent incursion of Highly Pathogenic Avian Influenza A (H5N1) virus into North America and subsequent dissemination of virus across the continent, has had significant adverse impacts on domestic poultry, and has led to widespread mortality in many wild bird species. Here we report the recent spillover of H5N1 into marine mammals in the northeastern United States, with associated mortality on a regional scale. This spillover is coincident with a second wave of H5N1 in sympatric wild birds also experiencing regional mortality events. Viral sequences derived from both seal and avian hosts reveal distinct viral genetic differences between the two waves of infection. Spillover into seals was closely related to virus from the second wave, and one of eight seal-derived sequences had the mammalian adaptation PB2 E627K.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Global dissemination of influenza A virus is driven by wild bird migration through arctic and subarctic zones.\n \n \n \n \n\n\n \n Gass, J. D.; Dusek, R. J.; Hall, J. S.; Hallgrimsson, G. T.; Halldórsson, H. P.; Vignisson, S. R.; Ragnarsdottir, S. B.; Jónsson, J. E.; Krauss, S.; Wong, S.; Wan, X.; Akter, S.; Sreevatsan, S.; Trovão, N. S.; Nutter, F. B.; Runstadler, J. A.; and Hill, N. J.\n\n\n \n\n\n\n Molecular Ecology, 32(1): 198–213. January 2023.\n \n\n\n\n
\n\n\n\n \n \n \"GlobalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n  \n \n 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{gass_global_2023,\n\ttitle = {Global dissemination of influenza {A} virus is driven by wild bird migration through arctic and subarctic zones},\n\tvolume = {32},\n\tissn = {0962-1083, 1365-294X},\n\turl = {https://onlinelibrary.wiley.com/doi/10.1111/mec.16738},\n\tdoi = {10.1111/mec.16738},\n\tlanguage = {en},\n\tnumber = {1},\n\turldate = {2023-02-23},\n\tjournal = {Molecular Ecology},\n\tauthor = {Gass, Jonathon D. and Dusek, Robert J. and Hall, Jeffrey S. and Hallgrimsson, Gunnar Thor and Halldórsson, Halldór Pálmar and Vignisson, Solvi Runar and Ragnarsdottir, Sunna Bjork and Jónsson, Jón Einar and Krauss, Scott and Wong, Sook‐San and Wan, Xiu‐Feng and Akter, Sadia and Sreevatsan, Srinand and Trovão, Nídia S. and Nutter, Felicia B. and Runstadler, Jonathan A. and Hill, Nichola J.},\n\tmonth = jan,\n\tyear = {2023},\n\tpages = {198--213},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2022\n \n \n (4)\n \n \n
\n
\n \n \n \n\n\n
\n \n\n \n \n \n \n \n \n Ecological divergence of wild birds drives avian influenza spillover and global spread.\n \n \n \n \n\n\n \n Hill, N. J.; Bishop, M. A.; Trovao, N. S.; Ineson, K. M.; Schaefer, A. L.; Puryear, W. B.; Zhou, K.; Foss, A. D.; Clark, D. E.; MacKenzie, K. G.; Gass; Borkenhagen, L. K.; Hall, J. S.; and Runstadler, J. A.\n\n\n \n\n\n\n PLoS Pathogens, 18(5): e1010062. May 2022.\n \n\n\n\n
\n\n\n\n \n \n \"EcologicalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 3 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hill_ecological_2022,\n\ttitle = {Ecological divergence of wild birds drives avian influenza spillover and global spread},\n\tvolume = {18},\n\tissn = {1553-7374 (Electronic) 1553-7366 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/35588106},\n\tdoi = {10.1371/journal.ppat.1010062},\n\tabstract = {The diversity of influenza A viruses (IAV) is primarily hosted by two highly divergent avian orders: Anseriformes (ducks, swans and geese) and Charadriiformes (gulls, terns and shorebirds). Studies of IAV have historically focused on Anseriformes, specifically dabbling ducks, overlooking the diversity of hosts in nature, including gull and goose species that have successfully adapted to human habitats. This study sought to address this imbalance by characterizing spillover dynamics and global transmission patterns of IAV over 10 years at greater taxonomic resolution than previously considered. Furthermore, the circulation of viral subtypes in birds that are either host-adapted (low pathogenic H13, H16) or host-generalist (highly pathogenic avian influenza-HPAI H5) provided a unique opportunity to test and extend models of viral evolution. Using Bayesian phylodynamic modelling we uncovered a complex transmission network that relied on ecologically divergent bird hosts. The generalist subtype, HPAI H5 was driven largely by wild geese and swans that acted as a source for wild ducks, gulls, land birds, and domestic geese. Gulls were responsible for moving HPAI H5 more rapidly than any other host, a finding that may reflect their long-distance, pelagic movements and their immuno-naive status against this subtype. Wild ducks, long viewed as primary hosts for spillover, occupied an optimal space for viral transmission, contributing to geographic expansion and rapid dispersal of HPAI H5. Evidence of inter-hemispheric dispersal via both the Pacific and Atlantic Rims was detected, supporting surveillance at high latitudes along continental margins to achieve early detection. Both neutral (geographic expansion) and non-neutral (antigenic selection) evolutionary processes were found to shape subtype evolution which manifested as unique geographic hotspots for each subtype at the global scale. This study reveals how a diversity of avian hosts contribute to viral spread and spillover with the potential to improve surveillance in an era of rapid global change.},\n\tnumber = {5},\n\tjournal = {PLoS Pathogens},\n\tauthor = {Hill, N. J. and Bishop, M. A. and Trovao, N. S. and Ineson, K. M. and Schaefer, A. L. and Puryear, W. B. and Zhou, K. and Foss, A. D. and Clark, D. E. and MacKenzie, K. G. and Gass, Jr., J. D. and Borkenhagen, L. K. and Hall, J. S. and Runstadler, J. A.},\n\tmonth = may,\n\tyear = {2022},\n\tpages = {e1010062},\n}\n\n
\n
\n\n\n
\n The diversity of influenza A viruses (IAV) is primarily hosted by two highly divergent avian orders: Anseriformes (ducks, swans and geese) and Charadriiformes (gulls, terns and shorebirds). Studies of IAV have historically focused on Anseriformes, specifically dabbling ducks, overlooking the diversity of hosts in nature, including gull and goose species that have successfully adapted to human habitats. This study sought to address this imbalance by characterizing spillover dynamics and global transmission patterns of IAV over 10 years at greater taxonomic resolution than previously considered. Furthermore, the circulation of viral subtypes in birds that are either host-adapted (low pathogenic H13, H16) or host-generalist (highly pathogenic avian influenza-HPAI H5) provided a unique opportunity to test and extend models of viral evolution. Using Bayesian phylodynamic modelling we uncovered a complex transmission network that relied on ecologically divergent bird hosts. The generalist subtype, HPAI H5 was driven largely by wild geese and swans that acted as a source for wild ducks, gulls, land birds, and domestic geese. Gulls were responsible for moving HPAI H5 more rapidly than any other host, a finding that may reflect their long-distance, pelagic movements and their immuno-naive status against this subtype. Wild ducks, long viewed as primary hosts for spillover, occupied an optimal space for viral transmission, contributing to geographic expansion and rapid dispersal of HPAI H5. Evidence of inter-hemispheric dispersal via both the Pacific and Atlantic Rims was detected, supporting surveillance at high latitudes along continental margins to achieve early detection. Both neutral (geographic expansion) and non-neutral (antigenic selection) evolutionary processes were found to shape subtype evolution which manifested as unique geographic hotspots for each subtype at the global scale. This study reveals how a diversity of avian hosts contribute to viral spread and spillover with the potential to improve surveillance in an era of rapid global change.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Highly pathogenic avian influenza is an emerging disease threat to wild birds in North America.\n \n \n \n\n\n \n Ramey, A. M.; Hill, N.; Deliberto, T.; Gibbs, S.; Hopkins, M.; Lang, A.; Poulson, R.; Prosser, D.; Sleeman, J.; Stallknecht, D.; and Wan, X.\n\n\n \n\n\n\n Journal of Wildlife Management,1–21. January 2022.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 3 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{ramey_highly_2022,\n\ttitle = {Highly pathogenic avian influenza is an emerging disease threat to wild birds in {North} {America}},\n\tcopyright = {All rights reserved},\n\tdoi = {https://doi.org/10.1002/jwmg.22171},\n\tabstract = {Prior to the emergence of the A/goose/Guangdong/1/1996 (Gs/GD) H5N1 influenza A virus, the long-held and well-supported paradigm was that highly pathogenic (HP) avian influenza (AI) outbreaks were restricted to poultry, the result of cross-species transmission of precursor viruses from wild aquatic birds that subsequently gained pathogenicity in domestic birds.  Therefore, management agencies typically adopted a prevention, control, and eradication strategy that included strict biosecurity for domestic bird production, isolation of infected and exposed flocks, and prompt depopulation.  In most cases, this strategy has proved sufficient for eradicating HPAI.  Since 2002, this paradigm has been challenged with many detections of viral descendants of the Gs/GD lineage among wild birds, most of which have been associated with sporadic mortality events.  Since the emergence and evolution of the genetically distinct clade 2.3.4.4 Gs/GD lineage HPAI viruses in approximately 2010, there has been further increases in the occurrence of HPAI in wild birds and geographic spread through migratory bird movement.  A prominent example is the introduction of clade 2.3.4.4 Gs/GD HPAI viruses from East Asia to North America via migratory birds in autumn 2014 that ultimately led to the largest outbreak of HPAI in the history of the United States.  Given the apparent maintenance of Gs/GD lineage HPAI viruses in a global avian reservoir; bidirectional virus exchange between wild and domestic birds facilitating the continued adaptation of Gs/GD HPAI viruses in wild bird hosts; the current frequency of HPAI outbreaks in wild birds globally, and particularly in Eurasia where Gs/GD HPAI viruses may now be enzootic; and ongoing dispersal of AI viruses from East Asia to North America via migratory birds, HPAI now represents an emerging disease threat to North American wildlife.  This recent paradigm shift implies that management of HPAI in domestic birds alone may no longer be sufficient to eradicate HPAI viruses from a given country or region.  Rather, agencies managing wild birds and their habitats may consider the development and/or adoption of mitigation strategies to minimize introductions to poultry, to reduce negative impacts on wild bird populations, and to diminish adverse effects to stakeholders utilizing wildlife resources.  The main objective of this review is, therefore, to provide information that will assist wildlife managers in developing mitigation strategies or approaches for dealing with outbreaks of Gs/GD HPAI in wild birds in the form of preparedness, surveillance, research, communications, and targeted management actions.  Resultant outbreak response plans and actions may represent meaningful steps of wildlife managers toward the use collaborative/multi-jurisdictional One Health approaches when it comes to the detection, investigation, and mitigation of emerging viruses at the human-domestic animal-wildlife interface.},\n\tjournal = {Journal of Wildlife Management},\n\tauthor = {Ramey, A. M. and Hill, N.J. and Deliberto, T. and Gibbs, S. and Hopkins, M.C. and Lang, A.S. and Poulson, R. and Prosser, D.J. and Sleeman, J.M. and Stallknecht, D.E. and Wan, X.F.},\n\tmonth = jan,\n\tyear = {2022},\n\tkeywords = {a viruses, a(h5n8) virus, avian influenza, bird flu, clade 2.3.4.4, disease, global spread, hemagglutinin gene, highly pathogenic, hong-kong, influenza, migratory waterfowl, north america, outbreak, subtype h5n1, united-states, wildlife},\n\tpages = {1--21},\n}\n\n
\n
\n\n\n
\n Prior to the emergence of the A/goose/Guangdong/1/1996 (Gs/GD) H5N1 influenza A virus, the long-held and well-supported paradigm was that highly pathogenic (HP) avian influenza (AI) outbreaks were restricted to poultry, the result of cross-species transmission of precursor viruses from wild aquatic birds that subsequently gained pathogenicity in domestic birds. Therefore, management agencies typically adopted a prevention, control, and eradication strategy that included strict biosecurity for domestic bird production, isolation of infected and exposed flocks, and prompt depopulation. In most cases, this strategy has proved sufficient for eradicating HPAI. Since 2002, this paradigm has been challenged with many detections of viral descendants of the Gs/GD lineage among wild birds, most of which have been associated with sporadic mortality events. Since the emergence and evolution of the genetically distinct clade 2.3.4.4 Gs/GD lineage HPAI viruses in approximately 2010, there has been further increases in the occurrence of HPAI in wild birds and geographic spread through migratory bird movement. A prominent example is the introduction of clade 2.3.4.4 Gs/GD HPAI viruses from East Asia to North America via migratory birds in autumn 2014 that ultimately led to the largest outbreak of HPAI in the history of the United States. Given the apparent maintenance of Gs/GD lineage HPAI viruses in a global avian reservoir; bidirectional virus exchange between wild and domestic birds facilitating the continued adaptation of Gs/GD HPAI viruses in wild bird hosts; the current frequency of HPAI outbreaks in wild birds globally, and particularly in Eurasia where Gs/GD HPAI viruses may now be enzootic; and ongoing dispersal of AI viruses from East Asia to North America via migratory birds, HPAI now represents an emerging disease threat to North American wildlife. This recent paradigm shift implies that management of HPAI in domestic birds alone may no longer be sufficient to eradicate HPAI viruses from a given country or region. Rather, agencies managing wild birds and their habitats may consider the development and/or adoption of mitigation strategies to minimize introductions to poultry, to reduce negative impacts on wild bird populations, and to diminish adverse effects to stakeholders utilizing wildlife resources. The main objective of this review is, therefore, to provide information that will assist wildlife managers in developing mitigation strategies or approaches for dealing with outbreaks of Gs/GD HPAI in wild birds in the form of preparedness, surveillance, research, communications, and targeted management actions. Resultant outbreak response plans and actions may represent meaningful steps of wildlife managers toward the use collaborative/multi-jurisdictional One Health approaches when it comes to the detection, investigation, and mitigation of emerging viruses at the human-domestic animal-wildlife interface.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Spatiotemporal changes in influenza A virus prevalence among wild waterfowl inhabiting the continental United States throughout the annual cycle.\n \n \n \n \n\n\n \n Kent, C. M.; Ramey, A. M.; Ackerman, J. T.; Bahl, J.; Bevins, S. N.; Bowman, A. S.; Boyce, W. M.; Cardona, C. J.; Casazza, M. L.; Cline, T. D.; E. De La Cruz, S.; Hall, J. S.; Hill, N. J.; Ip, H. S.; Krauss, S.; Mullinax, J. M.; Nolting, J. M.; Plancarte, M.; Poulson, R. L.; Runstadler, J. A.; Slemons, R. D.; Stallknecht, D. E.; Sullivan, J. D.; Takekawa, J. Y.; Webby, R. J.; Webster, R. G.; and Prosser, D. J.\n\n\n \n\n\n\n Scientific Reports, 12(1). July 2022.\n \n\n\n\n
\n\n\n\n \n \n \"SpatiotemporalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{kent_spatiotemporal_2022,\n\ttitle = {Spatiotemporal changes in influenza {A} virus prevalence among wild waterfowl inhabiting the continental {United} {States} throughout the annual cycle},\n\tvolume = {12},\n\tissn = {2045-2322},\n\turl = {https://www.nature.com/articles/s41598-022-17396-5},\n\tdoi = {10.1038/s41598-022-17396-5},\n\tabstract = {Avian influenza viruses can pose serious risks to agricultural production, human health, and wildlife. An understanding of viruses in wild reservoir species across time and space is important to informing surveillance programs, risk models, and potential population impacts for vulnerable species. Although it is recognized that influenza A virus prevalence peaks in reservoir waterfowl in late summer through autumn, temporal and spatial variation across species has not been fully characterized. We combined two large influenza databases for North America and applied spatiotemporal models to explore patterns in prevalence throughout the annual cycle and across the continental United States for 30 waterfowl species. Peaks in prevalence in late summer through autumn were pronounced for dabbling ducks in the genera Anas and Spatula, but not Mareca. Spatially, areas of high prevalence appeared to be related to regional duck density, with highest predicted prevalence found across the upper Midwest during early fall, though further study is needed. We documented elevated prevalence in late winter and early spring, particularly in the Mississippi Alluvial Valley. Our results suggest that spatiotemporal variation in prevalence outside autumn staging areas may also represent a dynamic parameter to be considered in IAV ecology and associated risks.},\n\tnumber = {1},\n\tjournal = {Scientific Reports},\n\tauthor = {Kent, Cody M. and Ramey, Andrew M. and Ackerman, Joshua T. and Bahl, Justin and Bevins, Sarah N. and Bowman, Andrew S. and Boyce, Walter M. and Cardona, Carol J. and Casazza, Michael L. and Cline, Troy D. and E. De La Cruz, Susan and Hall, Jeffrey S. and Hill, Nichola J. and Ip, Hon S. and Krauss, Scott and Mullinax, Jennifer M. and Nolting, Jacqueline M. and Plancarte, Magdalena and Poulson, Rebecca L. and Runstadler, Jonathan A. and Slemons, Richard D. and Stallknecht, David E. and Sullivan, Jeffery D. and Takekawa, John Y. and Webby, Richard J. and Webster, Robert G. and Prosser, Diann J.},\n\tmonth = jul,\n\tyear = {2022},\n}\n\n
\n
\n\n\n
\n Avian influenza viruses can pose serious risks to agricultural production, human health, and wildlife. An understanding of viruses in wild reservoir species across time and space is important to informing surveillance programs, risk models, and potential population impacts for vulnerable species. Although it is recognized that influenza A virus prevalence peaks in reservoir waterfowl in late summer through autumn, temporal and spatial variation across species has not been fully characterized. We combined two large influenza databases for North America and applied spatiotemporal models to explore patterns in prevalence throughout the annual cycle and across the continental United States for 30 waterfowl species. Peaks in prevalence in late summer through autumn were pronounced for dabbling ducks in the genera Anas and Spatula, but not Mareca. Spatially, areas of high prevalence appeared to be related to regional duck density, with highest predicted prevalence found across the upper Midwest during early fall, though further study is needed. We documented elevated prevalence in late winter and early spring, particularly in the Mississippi Alluvial Valley. Our results suggest that spatiotemporal variation in prevalence outside autumn staging areas may also represent a dynamic parameter to be considered in IAV ecology and associated risks.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2021\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Host barriers to SARS-CoV-2 demonstrated by ferrets in a high-exposure domestic setting.\n \n \n \n \n\n\n \n Sawatzki, K.; Hill, N. J.; Puryear, W. B.; Foss, A. D.; Stone, J. J.; and Runstadler, J. A.\n\n\n \n\n\n\n Proc Natl Acad Sci U S A, 118(18). May 2021.\n \n\n\n\n
\n\n\n\n \n \n \"HostPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 4 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{sawatzki_host_2021,\n\ttitle = {Host barriers to {SARS}-{CoV}-2 demonstrated by ferrets in a high-exposure domestic setting},\n\tvolume = {118},\n\tcopyright = {All rights reserved},\n\tissn = {1091-6490 (Electronic) 0027-8424 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/33858941},\n\tdoi = {10.1073/pnas.2025601118},\n\tabstract = {Ferrets (Mustela putorius furo) are mustelids of special relevance to laboratory studies of respiratory viruses and have been shown to be susceptible to severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection and onward transmission. Here, we report the results of a natural experiment where 29 ferrets in one home had prolonged, direct contact and constant environmental exposure to two humans with symptomatic disease, one of whom was confirmed positive for SARS-CoV-2. We observed no evidence of SARS-CoV-2 transmission from humans to ferrets based on viral and antibody assays. To better understand this discrepancy in experimental and natural infection in ferrets, we compared SARS-CoV-2 sequences from natural and experimental mustelid infections and identified two surface glycoprotein Spike (S) mutations associated with mustelids. While we found evidence that angiotensin-converting enzyme II provides a weak host barrier, one mutation only seen in ferrets is located in the novel S1/S2 cleavage site and is computationally predicted to decrease furin cleavage efficiency. These data support the idea that host factors interacting with the novel S1/S2 cleavage site may be a barrier in ferret SARS-CoV-2 susceptibility and that domestic ferrets are at low risk of natural infection from currently circulating SARS-CoV-2. We propose two mechanistically grounded hypotheses for mustelid host adaptation of SARS-CoV-2, with possible effects that require additional investigation.},\n\tnumber = {18},\n\tjournal = {Proc Natl Acad Sci U S A},\n\tauthor = {Sawatzki, K. and Hill, N. J. and Puryear, W. B. and Foss, A. D. and Stone, J. J. and Runstadler, J. A.},\n\tmonth = may,\n\tyear = {2021},\n\tkeywords = {*Host Adaptation, *Mutation, *SARS-CoV-2, *coronavirus, *genetics, *transmission, *virology, Angiotensin-Converting Enzyme 2/physiology, Animals, COVID-19/*transmission, Disease Susceptibility, Ferrets/*virology, Humans, Spike Glycoprotein, Coronavirus/*genetics},\n}\n\n
\n
\n\n\n
\n Ferrets (Mustela putorius furo) are mustelids of special relevance to laboratory studies of respiratory viruses and have been shown to be susceptible to severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection and onward transmission. Here, we report the results of a natural experiment where 29 ferrets in one home had prolonged, direct contact and constant environmental exposure to two humans with symptomatic disease, one of whom was confirmed positive for SARS-CoV-2. We observed no evidence of SARS-CoV-2 transmission from humans to ferrets based on viral and antibody assays. To better understand this discrepancy in experimental and natural infection in ferrets, we compared SARS-CoV-2 sequences from natural and experimental mustelid infections and identified two surface glycoprotein Spike (S) mutations associated with mustelids. While we found evidence that angiotensin-converting enzyme II provides a weak host barrier, one mutation only seen in ferrets is located in the novel S1/S2 cleavage site and is computationally predicted to decrease furin cleavage efficiency. These data support the idea that host factors interacting with the novel S1/S2 cleavage site may be a barrier in ferret SARS-CoV-2 susceptibility and that domestic ferrets are at low risk of natural infection from currently circulating SARS-CoV-2. We propose two mechanistically grounded hypotheses for mustelid host adaptation of SARS-CoV-2, with possible effects that require additional investigation.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Longitudinal analysis of pinnipeds in the northwest Atlantic provides insights on endemic circulation of phocine distemper virus.\n \n \n \n \n\n\n \n Puryear, W.; Sawatzki, K.; Bogomolni, A.; Hill, N.; Foss, A.; Stokholm, I.; Olsen, M. T.; Nielsen, O.; Waltzek, T.; Goldstein, T.; Subramaniam, K.; Rodrigues, T. C. S.; Belaganahalli, M.; Doughty, L.; Becker, L.; Stokes, A.; Niemeyer, M.; Tuttle, A.; Romano, T.; Linhares, M. B.; Fauquier, D.; and Runstadler, J.\n\n\n \n\n\n\n Proceedings of the Royal Society B: Biological Sciences, 288(1962): 20211841. November 2021.\n \n\n\n\n
\n\n\n\n \n \n \"LongitudinalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n  \n \n 2 downloads\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{puryear_longitudinal_2021,\n\ttitle = {Longitudinal analysis of pinnipeds in the northwest {Atlantic} provides insights on endemic circulation of phocine distemper virus},\n\tvolume = {288},\n\turl = {https://doi.org/10.1098/rspb.2021.1841},\n\tdoi = {10.1098/rspb.2021.1841},\n\tnumber = {1962},\n\turldate = {2021-11-10},\n\tjournal = {Proceedings of the Royal Society B: Biological Sciences},\n\tauthor = {Puryear, Wendy and Sawatzki, Kaitlin and Bogomolni, Andrea and Hill, Nichola and Foss, Alexa and Stokholm, Iben and Olsen, Morten Tange and Nielsen, Ole and Waltzek, Thomas and Goldstein, Tracey and Subramaniam, Kuttichantran and Rodrigues, Thais Carneiro Santos and Belaganahalli, Manjunatha and Doughty, Lynda and Becker, Lisa and Stokes, Ashley and Niemeyer, Misty and Tuttle, Allison and Romano, Tracy and Linhares, Mainity Batista and Fauquier, Deborah and Runstadler, Jonathan},\n\tmonth = nov,\n\tyear = {2021},\n\tkeywords = {morbillivirus, seal, unusual mortality event, viral genetics, virology, wildlife disease},\n\tpages = {20211841},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Age and season predict influenza A virus dynamics in urban gulls: consequences for natural hosts in unnatural landscapes.\n \n \n \n \n\n\n \n Ineson, K. M.; Hill, N. J.; Clark, D. E.; MacKenzie, K. G.; Whitney, J. J.; Laskaris, Y.; Ronconi, R. A.; Ellis, J. C.; Giroux, J.; Lair, S.; Stevens, S.; Puryear, W. B.; and Runstadler, J. A.\n\n\n \n\n\n\n Ecological Applications,e02497. November 2021.\n \n\n\n\n
\n\n\n\n \n \n \"AgePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{ineson_age_2021,\n\ttitle = {Age and season predict influenza {A} virus dynamics in urban gulls: consequences for natural hosts in unnatural landscapes},\n\tcopyright = {All rights reserved},\n\turl = {https://esajournals.onlinelibrary.wiley.com/doi/10.1002/eap.2497},\n\tdoi = {10.1002/eap.2497},\n\tabstract = {Gulls are ubiquitous in urban areas due to a growing reliance on anthropogenic feeding sites, which has led to changes in their abundance, distribution, and migration ecology, with implications for disease transmission. Gulls offer a valuable model for testing hypotheses regarding the dynamics of influenza A virus (IAV) - for which gulls are a natural reservoir in urban areas. We sampled sympatric populations of Ring-billed (Larus delawarensis), Herring (L. argentatus), and Great Black-backed Gulls (L. marinus) along the densely populated Atlantic rim of North America to understand how IAV transmission is influenced by drivers such as annual cycle, host species, age, habitat type, and their interplay. We found that horizontal transmission, rather than vertical transmission, played an outsized role in the amplification of IAV due to the convergence of gulls from different breeding grounds and age classes. We detected overlapping effects of age and season in our prevalence model, identifying juveniles during autumn as the primary drivers of the seasonal epidemic in gulls. Gulls accumulated immunity over their lifespan, however short-term fluctuations in seroprevalence were observed, suggesting that migration may impose limits on the immune system to maintain circulating antibodies. We found that gulls in coastal urban habitats had higher viral prevalence than gulls captured inland, correlating with higher richness of waterbird species along the coast, a mechanism supported by our movement data. The peak in viral prevalence in newly fledged gulls that are capable of long-distance movement has important implications for the spread of pathogens to novel hosts during the migratory season as well as for human health as gulls increasingly utilize urban habitats.},\n\tjournal = {Ecological Applications},\n\tauthor = {Ineson, Katherine M. and Hill, Nichola J. and Clark, Daniel E. and MacKenzie, Kenneth G. and Whitney, Jillian J. and Laskaris, Yianni and Ronconi, Robert A. and Ellis, Julie C. and Giroux, Jean-François and Lair, Stéphane and Stevens, Skyler and Puryear, Wendy B. and Runstadler, Jonathan A.},\n\tmonth = nov,\n\tyear = {2021},\n\tkeywords = {avian influenza, gull, immunity, pathogen, prevalence, transmission, urban, virus},\n\tpages = {e02497},\n}\n\n
\n
\n\n\n
\n Gulls are ubiquitous in urban areas due to a growing reliance on anthropogenic feeding sites, which has led to changes in their abundance, distribution, and migration ecology, with implications for disease transmission. Gulls offer a valuable model for testing hypotheses regarding the dynamics of influenza A virus (IAV) - for which gulls are a natural reservoir in urban areas. We sampled sympatric populations of Ring-billed (Larus delawarensis), Herring (L. argentatus), and Great Black-backed Gulls (L. marinus) along the densely populated Atlantic rim of North America to understand how IAV transmission is influenced by drivers such as annual cycle, host species, age, habitat type, and their interplay. We found that horizontal transmission, rather than vertical transmission, played an outsized role in the amplification of IAV due to the convergence of gulls from different breeding grounds and age classes. We detected overlapping effects of age and season in our prevalence model, identifying juveniles during autumn as the primary drivers of the seasonal epidemic in gulls. Gulls accumulated immunity over their lifespan, however short-term fluctuations in seroprevalence were observed, suggesting that migration may impose limits on the immune system to maintain circulating antibodies. We found that gulls in coastal urban habitats had higher viral prevalence than gulls captured inland, correlating with higher richness of waterbird species along the coast, a mechanism supported by our movement data. The peak in viral prevalence in newly fledged gulls that are capable of long-distance movement has important implications for the spread of pathogens to novel hosts during the migratory season as well as for human health as gulls increasingly utilize urban habitats.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A standardized instrument quantifying risk factors associated with bi-directional transmission of SARS-CoV-2 and other zoonotic pathogens: The COVID-19 human-animal interactions survey (CHAIS).\n \n \n \n \n\n\n \n Gass, J. D.; Waite, K. B.; Hill, N. J.; Dalton, K. R.; Sawatzki, K.; Runstadler, J.; and Davis, M. F.\n\n\n \n\n\n\n One Health. December 2021.\n Section: 100422\n\n\n\n
\n\n\n\n \n \n \"APaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{gass_standardized_2021,\n\ttitle = {A standardized instrument quantifying risk factors associated with bi-directional transmission of {SARS}-{CoV}-2 and other zoonotic pathogens: {The} {COVID}-19 human-animal interactions survey ({CHAIS})},\n\tissn = {23527714},\n\turl = {https://www.sciencedirect.com/science/article/pii/S2352771422000544?via%3Dihub},\n\tdoi = {10.1016/j.onehlt.2022.100422},\n\tabstract = {Similar to many zoonotic pathogens which transmit from animals to humans, SARS-CoV-2 (CoV-2), the virus responsible for the COVID-19 pandemic, most likely originated in Rhinolophus bats before spreading among humans globally. Early into the pandemic, reports of CoV-2 diagnoses in animals from various countries emerged. While most CoV-2 positive animals were confirmed to have been in close contact with CoV-2 positive humans, there has been a paucity of published evidence to-date describing risk factors associated with CoV-2 transmission among humans and animals. The COVID-19 Human-Animal Interactions Survey (CHAIS) was developed to provide a standardized instrument describing human-animal interactions during the pandemic and to evaluate behavioral, spatiotemporal, and biological risk factors associated with bi-directional zoonotic transmission of CoV-2 within shared environments, predominantly households with limited information about human-wildlife or human-livestock interactions. CHAIS measures four broad domains of transmission risk: 1) risk and intensity of infection in human hosts, 2) spatial characteristics of shared environments, 3) behaviors and human-animal interactions, and 4) susceptible animal subpopulations. Following the development of CHAIS, with a One Health approach, a multidisciplinary group of experts (n = 20) was invited to review and provide feedback on the survey for content validity. Expert feedback was incorporated into two final survey formats—an extended version and an abridged version for which specific core questions addressing zoonotic and reverse zoonotic transmission were identified. Both versions are modularized, with each section having the capacity to serve as independent instruments, allowing researchers to customize the survey based on context and research-specific needs. Further adaptations for studies seeking to investigate other zoonotic pathogens with similar routes of transmission (i.e. respiratory, direct contact) are also possible. The CHAIS instrument is a standardized human-animal interaction survey developed to provide important data on risk factors that guide transmission of CoV-2, and other similar pathogens, among humans and animals.},\n\tjournal = {One Health},\n\tauthor = {Gass, Jonathon D. and Waite, Kaitlin B. and Hill, Nichola J. and Dalton, Kathryn R. and Sawatzki, Kaitlin and Runstadler, Jonathan and Davis, Meghan F.},\n\tmonth = dec,\n\tyear = {2021},\n\tnote = {Section: 100422},\n}\n\n
\n
\n\n\n
\n Similar to many zoonotic pathogens which transmit from animals to humans, SARS-CoV-2 (CoV-2), the virus responsible for the COVID-19 pandemic, most likely originated in Rhinolophus bats before spreading among humans globally. Early into the pandemic, reports of CoV-2 diagnoses in animals from various countries emerged. While most CoV-2 positive animals were confirmed to have been in close contact with CoV-2 positive humans, there has been a paucity of published evidence to-date describing risk factors associated with CoV-2 transmission among humans and animals. The COVID-19 Human-Animal Interactions Survey (CHAIS) was developed to provide a standardized instrument describing human-animal interactions during the pandemic and to evaluate behavioral, spatiotemporal, and biological risk factors associated with bi-directional zoonotic transmission of CoV-2 within shared environments, predominantly households with limited information about human-wildlife or human-livestock interactions. CHAIS measures four broad domains of transmission risk: 1) risk and intensity of infection in human hosts, 2) spatial characteristics of shared environments, 3) behaviors and human-animal interactions, and 4) susceptible animal subpopulations. Following the development of CHAIS, with a One Health approach, a multidisciplinary group of experts (n = 20) was invited to review and provide feedback on the survey for content validity. Expert feedback was incorporated into two final survey formats—an extended version and an abridged version for which specific core questions addressing zoonotic and reverse zoonotic transmission were identified. Both versions are modularized, with each section having the capacity to serve as independent instruments, allowing researchers to customize the survey based on context and research-specific needs. Further adaptations for studies seeking to investigate other zoonotic pathogens with similar routes of transmission (i.e. respiratory, direct contact) are also possible. The CHAIS instrument is a standardized human-animal interaction survey developed to provide important data on risk factors that guide transmission of CoV-2, and other similar pathogens, among humans and animals.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2020\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Crossroads of highly pathogenic H5N1: overlap between wild and domestic birds in the Black Sea-Mediterranean impacts global transmission.\n \n \n \n \n\n\n \n Hill, N. J.; Smith, L. M.; Muzaffar, S. B.; Nagel, J. L.; Prosser, D. J.; Sullivan, J. D.; Spragens, K. A.; DeMattos, C. A.; DeMattos, C. C.; El Sayed, L.; Erciyas-Yavuz, K.; Davis, C. T.; Jones, J.; Kis, Z.; Donis, R. O.; Newman, S. H; and Takekawa, J. Y.\n\n\n \n\n\n\n Virus Evolution, 7(1). 2020.\n \n\n\n\n
\n\n\n\n \n \n \"CrossroadsPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_crossroads_2020,\n\ttitle = {Crossroads of highly pathogenic {H5N1}: overlap between wild and domestic birds in the {Black} {Sea}-{Mediterranean} impacts global transmission},\n\tvolume = {7},\n\tcopyright = {All rights reserved},\n\tissn = {2057-1577},\n\turl = {https://academic.oup.com/ve/article/7/1/veaa093/6046952},\n\tdoi = {10.1093/ve/veaa093},\n\tnumber = {1},\n\tjournal = {Virus Evolution},\n\tauthor = {Hill, Nichola J. and Smith, Lacy M. and Muzaffar, Sabir B. and Nagel, Jessica L. and Prosser, Diann J. and Sullivan, Jeffery D. and Spragens, Kyle A. and DeMattos, Carlos A. and DeMattos, Cecilia C. and El Sayed, Lu’ay and Erciyas-Yavuz, Kiraz and Davis, C. Todd and Jones, Joyce and Kis, Zoltan and Donis, Ruben O. and Newman, Scott  H and Takekawa, John Y.},\n\tyear = {2020},\n\tkeywords = {Egypt, Turkey, avian influenza, domestic poultry, influenza A virus, satellite telemetry, waterfowl migration, wild-domestic interface},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2019\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Pharmacokinetics and safety of intramuscular meloxicam in Zebra Finches (Taeniopygia guttata).\n \n \n \n \n\n\n \n Miller, K. A.; Hill, N. J.; Carrasco, S. E.; and Patterson, M. M.\n\n\n \n\n\n\n J Am Assoc Lab Anim Sci, 58(5): 589–593. September 2019.\n \n\n\n\n
\n\n\n\n \n \n \"PharmacokineticsPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{miller_pharmacokinetics_2019,\n\ttitle = {Pharmacokinetics and safety of intramuscular meloxicam in {Zebra} {Finches} ({Taeniopygia} guttata)},\n\tvolume = {58},\n\tcopyright = {All rights reserved},\n\tissn = {1559-6109 (Print) 1559-6109 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/31462348},\n\tdoi = {10.30802/AALAS-JAALAS-19-000032},\n\tabstract = {Meloxicam is the most frequently used NSAID in birds; however, its elimination t1/2 is highly variable among species. Because zebra finches that require analgesia could benefit from receiving meloxicam, we performed a pharmacokinetic study involving a single intramuscular dose of 1 or 2 mg/kg. Data analysis showed that Cmax, t1/2, and elimination rate constants were not significantly different between the 2 doses. In contrast, Cmax for 1- and 2-mg/kg doses of meloxicam approached a significant difference, and those for AUC0-infinity were significantly different. Importantly, a plasma concentration of 3500 ng/mL, considered a target level for meloxicam in other avian species, was maintained for approximately 9.5 h in finches that received 2 mg/kg, which was 4 h longer than in birds given 1 mg/kg. Both doses reached low plasma concentrations by 12 h after administration. Subsequently, 8 total doses of 1 or 2 mg/kg were administered to birds at 12-h intervals; these regimens caused no significant changes in select biochemical analytes or the Hct of meloxicam-treated birds. In addition, histopathologic changes for injection sites, kidney, liver, proventriculus, and ventriculus were minimal and similar between control and experimental groups after the multiple doses. These results suggest a 12-h or more frequent dosing interval is likely needed in zebra finches and that meloxicam at 1 or 2 mg/kg IM twice daily for 4 d is safe. The higher dose might provide longer analgesia compared with the lower dose, but a pharmacodynamics evaluation of meloxicam in zebra finches is needed to confirm analgesic efficacy.},\n\tnumber = {5},\n\tjournal = {J Am Assoc Lab Anim Sci},\n\tauthor = {Miller, K. A. and Hill, N. J. and Carrasco, S. E. and Patterson, M. M.},\n\tmonth = sep,\n\tyear = {2019},\n\tpages = {589--593},\n}\n\n
\n
\n\n\n
\n Meloxicam is the most frequently used NSAID in birds; however, its elimination t1/2 is highly variable among species. Because zebra finches that require analgesia could benefit from receiving meloxicam, we performed a pharmacokinetic study involving a single intramuscular dose of 1 or 2 mg/kg. Data analysis showed that Cmax, t1/2, and elimination rate constants were not significantly different between the 2 doses. In contrast, Cmax for 1- and 2-mg/kg doses of meloxicam approached a significant difference, and those for AUC0-infinity were significantly different. Importantly, a plasma concentration of 3500 ng/mL, considered a target level for meloxicam in other avian species, was maintained for approximately 9.5 h in finches that received 2 mg/kg, which was 4 h longer than in birds given 1 mg/kg. Both doses reached low plasma concentrations by 12 h after administration. Subsequently, 8 total doses of 1 or 2 mg/kg were administered to birds at 12-h intervals; these regimens caused no significant changes in select biochemical analytes or the Hct of meloxicam-treated birds. In addition, histopathologic changes for injection sites, kidney, liver, proventriculus, and ventriculus were minimal and similar between control and experimental groups after the multiple doses. These results suggest a 12-h or more frequent dosing interval is likely needed in zebra finches and that meloxicam at 1 or 2 mg/kg IM twice daily for 4 d is safe. The higher dose might provide longer analgesia compared with the lower dose, but a pharmacodynamics evaluation of meloxicam in zebra finches is needed to confirm analgesic efficacy.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Evidence of influenza A in wild Norway rats (Rattus norvegicus) in Boston, Massachusetts.\n \n \n \n \n\n\n \n Cummings, C. O.; Hill, N. J.; Puryear, W. B.; Rogers, B.; Mukherjee, J.; Leibler, J. H.; Rosenbaum, M. H.; and Runstadler, J. A.\n\n\n \n\n\n\n Frontiers in Ecology and Evolution, 7(36). March 2019.\n \n\n\n\n
\n\n\n\n \n \n \"EvidencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{cummings_evidence_2019,\n\ttitle = {Evidence of influenza {A} in wild {Norway} rats ({Rattus} norvegicus) in {Boston}, {Massachusetts}},\n\tvolume = {7},\n\tcopyright = {All rights reserved},\n\tissn = {2296-701X},\n\tshorttitle = {Influenza in wild {Norway} rats},\n\turl = {https://www.frontiersin.org/article/10.3389/fevo.2019.00036 https://fjfsdata01prod.blob.core.windows.net/articles/files/434274/pubmed-zip/.versions/1/.package-entries/fevo-07-00036/fevo-07-00036.pdf?sv=2015-12-11&sr=b&sig=uxdGCrkb1QSh0B43ESyzo2WinRCxdGbQhCvymCpM0Cs%3D&se=2019-09-11T00%3A07%3A16Z&sp=r&rscd=attachment%3B%20filename%2A%3DUTF-8%27%27fevo-07-00036.pdf},\n\tdoi = {10.3389/fevo.2019.00036},\n\tabstract = {Influenza A virus (IAV) is known to circulate among human and animal reservoirs, yet there are few studies that address the potential for urban rodents to carry and shed IAV. Rodents are often used as influenza models in the lab, but the few field studies that have looked for evidence of IAV in rodents have done so primarily in rural areas following outbreaks of IAV in poultry. This study sought to assess the prevalence of IAV recovered from wild Norway rats in a dense urban location (Boston). To do this, we sampled the oronasal cavity, paws, and lungs of Norway rats trapped by the City of Boston’s Inspectional Services from December 2016 to September 2018. All samples were screened by real-time, reverse transcriptase PCR targeting the conserved IAV matrix segment. A total of 163 rats were trapped, 18 of which (11.04\\%) were RT-PCR positive for IAV in either oronasal swabs (9), paw swabs (9), both (2), or lung homogenates (2). A generalized linear model indicated that month and geographic location were correlated with IAV-positive PCR status of rats. A seasonal trend in IAV-PCR status was observed with the highest prevalence occurring in the winter months (December - January) followed by a decline over the course of the year reaching lowest prevalence in September. Sex and weight of rats were not significantly associated with IAV-PCR status, suggesting that rodent demography is not a primary driver of infection. This pilot study provides evidence of the need to further investigate the role that wild rats may play as reservoirs or mechanical vectors for IAV circulation in urban environments across seasons.},\n\tlanguage = {English},\n\tnumber = {36},\n\tjournal = {Frontiers in Ecology and Evolution},\n\tauthor = {Cummings, Charles O. and Hill, Nichola J. and Puryear, Wendy B. and Rogers, Benjamin and Mukherjee, Jean and Leibler, Jessica H. and Rosenbaum, Marieke H. and Runstadler, Jonathan A.},\n\tmonth = mar,\n\tyear = {2019},\n\tkeywords = {Novel host, Orthomyxoviridae, Urban rodents, Wildlife disease, influenza},\n}\n\n
\n
\n\n\n
\n Influenza A virus (IAV) is known to circulate among human and animal reservoirs, yet there are few studies that address the potential for urban rodents to carry and shed IAV. Rodents are often used as influenza models in the lab, but the few field studies that have looked for evidence of IAV in rodents have done so primarily in rural areas following outbreaks of IAV in poultry. This study sought to assess the prevalence of IAV recovered from wild Norway rats in a dense urban location (Boston). To do this, we sampled the oronasal cavity, paws, and lungs of Norway rats trapped by the City of Boston’s Inspectional Services from December 2016 to September 2018. All samples were screened by real-time, reverse transcriptase PCR targeting the conserved IAV matrix segment. A total of 163 rats were trapped, 18 of which (11.04%) were RT-PCR positive for IAV in either oronasal swabs (9), paw swabs (9), both (2), or lung homogenates (2). A generalized linear model indicated that month and geographic location were correlated with IAV-positive PCR status of rats. A seasonal trend in IAV-PCR status was observed with the highest prevalence occurring in the winter months (December - January) followed by a decline over the course of the year reaching lowest prevalence in September. Sex and weight of rats were not significantly associated with IAV-PCR status, suggesting that rodent demography is not a primary driver of infection. This pilot study provides evidence of the need to further investigate the role that wild rats may play as reservoirs or mechanical vectors for IAV circulation in urban environments across seasons.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2017\n \n \n (4)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Reassortment of Influenza A Viruses in Wild Birds in Alaska before H5 Clade 2.3.4.4 Outbreaks.\n \n \n \n \n\n\n \n Hill, N. J.; Hussein, I. T.; Davis, K. R.; Ma, E. J.; Spivey, T. J.; Ramey, A. M.; Puryear, W. B.; Das, S. R.; Halpin, R. A.; Lin, X.; Fedorova, N. B.; Suarez, D. L.; Boyce, W. M.; and Runstadler, J. A.\n\n\n \n\n\n\n Emerg Infect Dis, 23(4): 654–657. April 2017.\n \n\n\n\n
\n\n\n\n \n \n \"ReassortmentPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_reassortment_2017,\n\ttitle = {Reassortment of {Influenza} {A} {Viruses} in {Wild} {Birds} in {Alaska} before {H5} {Clade} 2.3.4.4 {Outbreaks}},\n\tvolume = {23},\n\tcopyright = {All rights reserved},\n\tissn = {1080-6059 (Electronic) 1080-6040 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/28322698},\n\tdoi = {10.3201/eid2304.161668},\n\tabstract = {Sampling of mallards in Alaska during September 2014-April 2015 identified low pathogenic avian influenza A virus (subtypes H5N2 and H1N1) that shared ancestry with highly pathogenic reassortant H5N2 and H5N1 viruses. Molecular dating indicated reassortment soon after interhemispheric movement of H5N8 clade 2.3.4.4, suggesting genetic exchange in Alaska or surrounds before outbreaks.},\n\tnumber = {4},\n\tjournal = {Emerg Infect Dis},\n\tauthor = {Hill, N. J. and Hussein, I. T. and Davis, K. R. and Ma, E. J. and Spivey, T. J. and Ramey, A. M. and Puryear, W. B. and Das, S. R. and Halpin, R. A. and Lin, X. and Fedorova, N. B. and Suarez, D. L. and Boyce, W. M. and Runstadler, J. A.},\n\tmonth = apr,\n\tyear = {2017},\n\tkeywords = {Alaska, H5 clade 2.3.4.4, agricultural biosecurity, avian, bird migration, ducks, genetic reassortment, infectious diseases, influenza A virus, mallards, poultry, spillover transmission, viruses, wild birds, zoonoses},\n\tpages = {654--657},\n}\n\n
\n
\n\n\n
\n Sampling of mallards in Alaska during September 2014-April 2015 identified low pathogenic avian influenza A virus (subtypes H5N2 and H1N1) that shared ancestry with highly pathogenic reassortant H5N2 and H5N1 viruses. Molecular dating indicated reassortment soon after interhemispheric movement of H5N8 clade 2.3.4.4, suggesting genetic exchange in Alaska or surrounds before outbreaks.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Surveillance for highly pathogenic influenza A viruses in California during 2014-2015 provides insights into viral evolutionary pathways and the spatiotemporal extent of viruses in the Pacific Americas Flyway.\n \n \n \n \n\n\n \n Ramey, A. M.; Hill, N. J.; Cline, T.; Plancarte, M.; De La Cruz, S.; Casazza, M. L.; Ackerman, J. T.; Fleskes, J. P.; Vickers, T. W.; Reeves, A. B.; Gulland, F.; Fontaine, C.; Prosser, D. J.; Runstadler, J. A.; and Boyce, W. M.\n\n\n \n\n\n\n Emerg Microbes Infect, 6(9): e80. September 2017.\n \n\n\n\n
\n\n\n\n \n \n \"SurveillancePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{ramey_surveillance_2017,\n\ttitle = {Surveillance for highly pathogenic influenza {A} viruses in {California} during 2014-2015 provides insights into viral evolutionary pathways and the spatiotemporal extent of viruses in the {Pacific} {Americas} {Flyway}},\n\tvolume = {6},\n\tcopyright = {All rights reserved},\n\tissn = {2222-1751 (Electronic) 2222-1751 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/28874792},\n\tdoi = {10.1038/emi.2017.66},\n\tabstract = {We used surveillance data collected in California before, concurrent with, and subsequent to an outbreak of highly pathogenic (HP) clade 2.3.4.4 influenza A viruses (IAVs) in 2014-2015 to (i) evaluate IAV prevalence in waterfowl, (ii) assess the evidence for spill-over infections in marine mammals and (iii) genetically characterize low-pathogenic (LP) and HP IAVs to refine inference on the spatiotemporal extent of HP genome constellations and to evaluate possible evolutionary pathways. We screened samples from 1496 waterfowl and 1142 marine mammals collected from April 2014 to August 2015 and detected IAV RNA in 159 samples collected from birds (n=157) and pinnipeds (n=2). HP IAV RNA was identified in three samples originating from American wigeon (Anas americana). Genetic sequence data were generated for a clade 2.3.4.4 HP IAV-positive diagnostic sample and 57 LP IAV isolates. Phylogenetic analyses revealed that the HP IAV was a reassortant H5N8 virus with gene segments closely related to LP IAVs detected in mallards (Anas platyrhynchos) sampled in California and other IAVs detected in wild birds sampled within the Pacific Americas Flyway. In addition, our analysis provided support for common ancestry between LP IAVs recovered from waterfowl sampled in California and gene segments of reassortant HP H5N1 IAVs detected in British Columbia, Canada and Washington, USA. Our investigation provides evidence that waterfowl are likely to have played a role in the evolution of reassortant HP IAVs in the Pacific Americas Flyway during 2014-2015, whereas we did not find support for spill-over infections in potential pinniped hosts.},\n\tnumber = {9},\n\tjournal = {Emerg Microbes Infect},\n\tauthor = {Ramey, A. M. and Hill, N. J. and Cline, T. and Plancarte, M. and De La Cruz, S. and Casazza, M. L. and Ackerman, J. T. and Fleskes, J. P. and Vickers, T. W. and Reeves, A. B. and Gulland, F. and Fontaine, C. and Prosser, D. J. and Runstadler, J. A. and Boyce, W. M.},\n\tmonth = sep,\n\tyear = {2017},\n\tpages = {e80},\n}\n\n
\n
\n\n\n
\n We used surveillance data collected in California before, concurrent with, and subsequent to an outbreak of highly pathogenic (HP) clade 2.3.4.4 influenza A viruses (IAVs) in 2014-2015 to (i) evaluate IAV prevalence in waterfowl, (ii) assess the evidence for spill-over infections in marine mammals and (iii) genetically characterize low-pathogenic (LP) and HP IAVs to refine inference on the spatiotemporal extent of HP genome constellations and to evaluate possible evolutionary pathways. We screened samples from 1496 waterfowl and 1142 marine mammals collected from April 2014 to August 2015 and detected IAV RNA in 159 samples collected from birds (n=157) and pinnipeds (n=2). HP IAV RNA was identified in three samples originating from American wigeon (Anas americana). Genetic sequence data were generated for a clade 2.3.4.4 HP IAV-positive diagnostic sample and 57 LP IAV isolates. Phylogenetic analyses revealed that the HP IAV was a reassortant H5N8 virus with gene segments closely related to LP IAVs detected in mallards (Anas platyrhynchos) sampled in California and other IAVs detected in wild birds sampled within the Pacific Americas Flyway. In addition, our analysis provided support for common ancestry between LP IAVs recovered from waterfowl sampled in California and gene segments of reassortant HP H5N1 IAVs detected in British Columbia, Canada and Washington, USA. Our investigation provides evidence that waterfowl are likely to have played a role in the evolution of reassortant HP IAVs in the Pacific Americas Flyway during 2014-2015, whereas we did not find support for spill-over infections in potential pinniped hosts.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Evaluation of best practices for the euthanasia of Zebra Finches (Taeniopygia guttata).\n \n \n \n \n\n\n \n Scott, K. E.; Bracchi, L. A.; Lieberman, M. T.; Hill, N. J.; Caron, T. J.; and Patterson, M. M.\n\n\n \n\n\n\n J Am Assoc Lab Anim Sci, 56(6): 802–806. November 2017.\n \n\n\n\n
\n\n\n\n \n \n \"EvaluationPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{scott_evaluation_2017,\n\ttitle = {Evaluation of best practices for the euthanasia of {Zebra} {Finches} ({Taeniopygia} guttata)},\n\tvolume = {56},\n\tcopyright = {All rights reserved},\n\tissn = {1559-6109 (Print) 1559-6109 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/29256376},\n\tabstract = {Although zebra finches (Taeniopygia guttata) have been used in biomedical research for many years, no published reports are available about euthanizing these small birds. In this study, we compared 5 methods for zebra finch euthanasia: sodium pentobarbital (NaP) given intracoelomically with physical restraint but no anesthesia; isoflurane anesthesia followed by intracoelomic injection of NaP; and CO2 asphyxiation at 20\\%, 40\\%, and 80\\% chamber displacement rates (percentage of chamber volume per minute). Birds undergoing euthanasia were videorecorded and scored by 2 observers for behaviors potentially related to discomfort or distress. Time to recumbency and time until respiratory arrest (RA) were also assessed. RA was achieved faster by using NaP in a conscious bird compared to using isoflurane anesthesia followed by NaP; however, neither method caused behaviors that might affect animal welfare, such as open-mouth breathing, to any appreciable extent. Among the CO2 treatment groups, there was an inverse correlation between the chamber displacement rate used and the duration of open-mouth breathing, onset of head retroflexion, and time to RA. The results demonstrate that the intracoelomic administration of NaP in an awake, restrained zebra finch is a rapid and effective method of euthanasia. If CO2 is used to euthanize these birds, a high displacement rate (for example, 80\\%) will minimize the duration of the procedure and associated behaviors.},\n\tnumber = {6},\n\tjournal = {J Am Assoc Lab Anim Sci},\n\tauthor = {Scott, K. E. and Bracchi, L. A. and Lieberman, M. T. and Hill, N. J. and Caron, T. J. and Patterson, M. M.},\n\tmonth = nov,\n\tyear = {2017},\n\tkeywords = {*Animal Welfare, Animals, Animals, Laboratory, Carbon Dioxide/*administration \\& dosage, Euthanasia, Animal/*methods, Female, Finches/classification/*physiology, Isoflurane/*administration \\& dosage, Male, Pentobarbital/*administration \\& dosage},\n\tpages = {802--806},\n}\n\n
\n
\n\n\n
\n Although zebra finches (Taeniopygia guttata) have been used in biomedical research for many years, no published reports are available about euthanizing these small birds. In this study, we compared 5 methods for zebra finch euthanasia: sodium pentobarbital (NaP) given intracoelomically with physical restraint but no anesthesia; isoflurane anesthesia followed by intracoelomic injection of NaP; and CO2 asphyxiation at 20%, 40%, and 80% chamber displacement rates (percentage of chamber volume per minute). Birds undergoing euthanasia were videorecorded and scored by 2 observers for behaviors potentially related to discomfort or distress. Time to recumbency and time until respiratory arrest (RA) were also assessed. RA was achieved faster by using NaP in a conscious bird compared to using isoflurane anesthesia followed by NaP; however, neither method caused behaviors that might affect animal welfare, such as open-mouth breathing, to any appreciable extent. Among the CO2 treatment groups, there was an inverse correlation between the chamber displacement rate used and the duration of open-mouth breathing, onset of head retroflexion, and time to RA. The results demonstrate that the intracoelomic administration of NaP in an awake, restrained zebra finch is a rapid and effective method of euthanasia. If CO2 is used to euthanize these birds, a high displacement rate (for example, 80%) will minimize the duration of the procedure and associated behaviors.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Detection of Wellfleet Bay Virus Antibodies in Sea Birds of the Northeastern United States.\n \n \n \n \n\n\n \n Ballard, J. R.; Mickley, R.; Brown, J. D.; Hill, N. J.; Runstadler, J. A.; Clark, D. E.; Ellis, J. C.; Mead, D. G.; and Fischer, J. R.\n\n\n \n\n\n\n J Wildl Dis, 53(4): 875–879. June 2017.\n \n\n\n\n
\n\n\n\n \n \n \"DetectionPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{ballard_detection_2017,\n\ttitle = {Detection of {Wellfleet} {Bay} {Virus} {Antibodies} in {Sea} {Birds} of the {Northeastern} {United} {States}},\n\tvolume = {53},\n\tcopyright = {All rights reserved},\n\tissn = {1943-3700 (Electronic) 0090-3558 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/28640712},\n\tdoi = {10.7589/2016-10-237},\n\tabstract = {Wellfleet Bay virus (WFBV) is a recently described orthomyxovirus isolated from the tissues of Common Eiders (Somateria mollissima) collected during recurrent mortality events on Cape Cod, Massachusetts, USA. Coastal Massachusetts is the only location where disease or mortality associated with this virus has been detected in wild birds, and a previous seroprevalence study found a significantly higher frequency of viral exposure in eiders from this location than from other areas sampled in North America. This suggests that coastal Massachusetts is an epicenter of WFBV exposure, but the reason for this is unknown. Opportunistic sampling of sympatric species and testing of banked serum was used to investigate potential host range and spatiotemporal patterns of WFBV exposure. Antibodies were detected in Herring Gulls (Larus argentatus), Ring-billed Gulls (Larus delawarensis), a White-winged Scoter (Melanitta fusca), and a Black Scoter (Melanitta nigra). These findings demonstrate the likely occurrence of fall/winter transmission, expand our understanding of the host range of the virus, and provide further insight into the epidemiology of WFBV in the northeastern US.},\n\tnumber = {4},\n\tjournal = {J Wildl Dis},\n\tauthor = {Ballard, J. R. and Mickley, R. and Brown, J. D. and Hill, N. J. and Runstadler, J. A. and Clark, D. E. and Ellis, J. C. and Mead, D. G. and Fischer, J. R.},\n\tmonth = jun,\n\tyear = {2017},\n\tkeywords = {Black Scoter, Common Eider, Herring Gull, Ring-billed Gull, Wellfleet Bay virus, White-winged Scoter},\n\tpages = {875--879},\n}\n\n
\n
\n\n\n
\n Wellfleet Bay virus (WFBV) is a recently described orthomyxovirus isolated from the tissues of Common Eiders (Somateria mollissima) collected during recurrent mortality events on Cape Cod, Massachusetts, USA. Coastal Massachusetts is the only location where disease or mortality associated with this virus has been detected in wild birds, and a previous seroprevalence study found a significantly higher frequency of viral exposure in eiders from this location than from other areas sampled in North America. This suggests that coastal Massachusetts is an epicenter of WFBV exposure, but the reason for this is unknown. Opportunistic sampling of sympatric species and testing of banked serum was used to investigate potential host range and spatiotemporal patterns of WFBV exposure. Antibodies were detected in Herring Gulls (Larus argentatus), Ring-billed Gulls (Larus delawarensis), a White-winged Scoter (Melanitta fusca), and a Black Scoter (Melanitta nigra). These findings demonstrate the likely occurrence of fall/winter transmission, expand our understanding of the host range of the virus, and provide further insight into the epidemiology of WFBV in the northeastern US.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2016\n \n \n (7)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Prevalence of influenza A virus in live-captured North Atlantic gray seals: a possible wild reservoir.\n \n \n \n \n\n\n \n Puryear, W. B.; Keogh, M.; Hill, N.; Moxley, J.; Josephson, E.; Davis, K. R.; Bandoro, C.; Lidgard, D.; Bogomolni, A.; Levin, M.; Lang, S.; Hammill, M.; Bowen, D.; Johnston, D. W.; Romano, T.; Waring, G.; and Runstadler, J.\n\n\n \n\n\n\n Emerg Microbes Infect, 5(8): e81. 2016.\n \n\n\n\n
\n\n\n\n \n \n \"PrevalencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n\n\n\n
\n
@article{puryear_prevalence_2016,\n\ttitle = {Prevalence of influenza {A} virus in live-captured {North} {Atlantic} gray seals: a possible wild reservoir},\n\tvolume = {5},\n\tcopyright = {All rights reserved},\n\tissn = {2222-1751 (Electronic) 2222-1751 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/27485496 http://www.nature.com/emi/journal/v5/n8/pdf/emi201677a.pdf},\n\tdoi = {10.1038/emi.2016.77},\n\tabstract = {Influenza A virus (IAV) has been associated with multiple unusual mortality events (UMEs) in North Atlantic pinnipeds, frequently attributed to spillover of virus from wild-bird reservoirs. To determine if endemic infection persists outside of UMEs, we undertook a multiyear investigation of IAV in healthy, live-captured Northwest Atlantic gray seals (Halichoerus grypus). From 2013 to 2015, we sampled 345 pups and 57 adults from Cape Cod, MA, USA and Nova Scotia, Canada consistently detecting IAV infection across all groups. There was an overall viral prevalence of 9.0\\% (95\\% confidence interval (CI): 6.4\\%-12.5\\%) in weaned pups and 5.3\\% (CI: 1.2\\%-14.6\\%) in adults, with seroprevalences of 19.3\\% (CI: 15.0\\%-24.5\\%) and 50\\% (CI: 33.7\\%-66.4\\%), respectively. Positive sera showed a broad reactivity to diverse influenza subtypes. IAV status did not correlate with measures of animal health nor impact animal movement or foraging. This study demonstrated that Northwest Atlantic gray seals are both permissive to and tolerant of diverse IAV, possibly representing an endemically infected wild reservoir population.},\n\tnumber = {8},\n\tjournal = {Emerg Microbes Infect},\n\tauthor = {Puryear, W. B. and Keogh, M. and Hill, N.J. and Moxley, J. and Josephson, E. and Davis, K. R. and Bandoro, C. and Lidgard, D. and Bogomolni, A. and Levin, M. and Lang, S. and Hammill, M. and Bowen, D. and Johnston, D. W. and Romano, T. and Waring, G. and Runstadler, J.},\n\tyear = {2016},\n\tkeywords = {Nichola Hill},\n\tpages = {e81},\n}\n\n
\n
\n\n\n
\n Influenza A virus (IAV) has been associated with multiple unusual mortality events (UMEs) in North Atlantic pinnipeds, frequently attributed to spillover of virus from wild-bird reservoirs. To determine if endemic infection persists outside of UMEs, we undertook a multiyear investigation of IAV in healthy, live-captured Northwest Atlantic gray seals (Halichoerus grypus). From 2013 to 2015, we sampled 345 pups and 57 adults from Cape Cod, MA, USA and Nova Scotia, Canada consistently detecting IAV infection across all groups. There was an overall viral prevalence of 9.0% (95% confidence interval (CI): 6.4%-12.5%) in weaned pups and 5.3% (CI: 1.2%-14.6%) in adults, with seroprevalences of 19.3% (CI: 15.0%-24.5%) and 50% (CI: 33.7%-66.4%), respectively. Positive sera showed a broad reactivity to diverse influenza subtypes. IAV status did not correlate with measures of animal health nor impact animal movement or foraging. This study demonstrated that Northwest Atlantic gray seals are both permissive to and tolerant of diverse IAV, possibly representing an endemically infected wild reservoir population.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Reticulate evolution is favored in influenza niche switching.\n \n \n \n \n\n\n \n Ma, E. J.; Hill, N. J.; Zabilansky, J.; Yuan, K.; and Runstadler, J. A.\n\n\n \n\n\n\n Proc Natl Acad Sci U S A, 113(19): 5335–9. May 2016.\n \n\n\n\n
\n\n\n\n \n \n \"ReticulatePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{ma_reticulate_2016,\n\ttitle = {Reticulate evolution is favored in influenza niche switching},\n\tvolume = {113},\n\tcopyright = {All rights reserved},\n\tissn = {1091-6490 (Electronic) 0027-8424 (Linking)},\n\turl = {http://www.pnas.org/content/113/19/5335.abstract},\n\tdoi = {10.1073/pnas.1522921113},\n\tabstract = {Reticulate evolution is thought to accelerate the process of evolution beyond simple genetic drift and selection, helping to rapidly generate novel hybrids with combinations of adaptive traits. However, the long-standing dogma that reticulate evolutionary processes are likewise advantageous for switching ecological niches, as in microbial pathogen host switch events, has not been explicitly tested. We use data from the influenza genome sequencing project and a phylogenetic heuristic approach to show that reassortment, a reticulate evolutionary mechanism, predominates over mutational drift in transmission between different host species. Moreover, as host evolutionary distance increases, reassortment is increasingly favored. We conclude that the greater the quantitative difference between ecological niches, the greater the importance of reticulate evolutionary processes in overcoming niche barriers.},\n\tnumber = {19},\n\tjournal = {Proc Natl Acad Sci U S A},\n\tauthor = {Ma, E. J. and Hill, N. J. and Zabilansky, J. and Yuan, K. and Runstadler, J. A.},\n\tmonth = may,\n\tyear = {2016},\n\tkeywords = {ecology, host switch, influenza, reassortment, reticulate evolution},\n\tpages = {5335--9},\n}\n\n
\n
\n\n\n
\n Reticulate evolution is thought to accelerate the process of evolution beyond simple genetic drift and selection, helping to rapidly generate novel hybrids with combinations of adaptive traits. However, the long-standing dogma that reticulate evolutionary processes are likewise advantageous for switching ecological niches, as in microbial pathogen host switch events, has not been explicitly tested. We use data from the influenza genome sequencing project and a phylogenetic heuristic approach to show that reassortment, a reticulate evolutionary mechanism, predominates over mutational drift in transmission between different host species. Moreover, as host evolutionary distance increases, reassortment is increasingly favored. We conclude that the greater the quantitative difference between ecological niches, the greater the importance of reticulate evolutionary processes in overcoming niche barriers.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Transmission of influenza reflects seasonality of wild birds across the annual cycle.\n \n \n \n \n\n\n \n Hill, N. J.; Ma, E. J.; Meixell, B. W.; Lindberg, M. S.; Boyce, W. M.; and Runstadler, J. A.\n\n\n \n\n\n\n Ecology Letters, 19(8): 915–925. June 2016.\n \n\n\n\n
\n\n\n\n \n \n \"TransmissionPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n  \n \n 1 download\n \n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_transmission_2016,\n\ttitle = {Transmission of influenza reflects seasonality of wild birds across the annual cycle},\n\tvolume = {19},\n\tcopyright = {All rights reserved},\n\tissn = {1461-0248 (Electronic) 1461-023X (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/27324078 http://onlinelibrary.wiley.com/doi/10.1111/ele.12629/abstract},\n\tdoi = {10.1111/ele.12629},\n\tabstract = {Influenza A Viruses (IAV) in nature must overcome shifting transmission barriers caused by the mobility of their primary host, migratory wild birds, that change throughout the annual cycle. Using a phylogenetic network of viral sequences from North American wild birds (2008-2011) we demonstrate a shift from intraspecific to interspecific transmission that along with reassortment, allows IAV to achieve viral flow across successive seasons from summer to winter. Our study supports amplification of IAV during summer breeding seeded by overwintering virus persisting locally and virus introduced from a wide range of latitudes. As birds migrate from breeding sites to lower latitudes, they become involved in transmission networks with greater connectivity to other bird species, with interspecies transmission of reassortant viruses peaking during the winter. We propose that switching transmission dynamics may be a critical strategy for pathogens that infect mobile hosts inhabiting regions with strong seasonality.},\n\tnumber = {8},\n\tjournal = {Ecology Letters},\n\tauthor = {Hill, N. J. and Ma, E. J. and Meixell, B. W. and Lindberg, M. S. and Boyce, W. M. and Runstadler, J. A.},\n\tmonth = jun,\n\tyear = {2016},\n\tkeywords = {Avian influenza, biological rhythms, bird migration, host contact structure, influenza A virus, migratory cycle, seasonality, transmission networks, viral flow, zoonotic disease},\n\tpages = {915--925},\n}\n\n
\n
\n\n\n
\n Influenza A Viruses (IAV) in nature must overcome shifting transmission barriers caused by the mobility of their primary host, migratory wild birds, that change throughout the annual cycle. Using a phylogenetic network of viral sequences from North American wild birds (2008-2011) we demonstrate a shift from intraspecific to interspecific transmission that along with reassortment, allows IAV to achieve viral flow across successive seasons from summer to winter. Our study supports amplification of IAV during summer breeding seeded by overwintering virus persisting locally and virus introduced from a wide range of latitudes. As birds migrate from breeding sites to lower latitudes, they become involved in transmission networks with greater connectivity to other bird species, with interspecies transmission of reassortant viruses peaking during the winter. We propose that switching transmission dynamics may be a critical strategy for pathogens that infect mobile hosts inhabiting regions with strong seasonality.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Genetic characterization of H5N2 influenza viruses isolated from wild birds in Japan suggests multiple reassortment.\n \n \n \n \n\n\n \n Sultan, S.; Bui, V. N.; Hill, N. J.; Hussein, I. T.; Trinh, D. Q.; Inage, K.; Hashizume, T.; Runstadler, J. A.; Ogawa, H.; and Imai, K.\n\n\n \n\n\n\n Arch Virol. August 2016.\n \n\n\n\n
\n\n\n\n \n \n \"GeneticPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{sultan_genetic_2016,\n\ttitle = {Genetic characterization of {H5N2} influenza viruses isolated from wild birds in {Japan} suggests multiple reassortment},\n\tcopyright = {All rights reserved},\n\tissn = {1432-8798 (Electronic) 0304-8608 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/27573808},\n\tdoi = {10.1007/s00705-016-3023-4},\n\tabstract = {Low-pathogenic avian influenza viruses (LPAIVs) of the H5 subtype can mutate to highly pathogenic forms, potentially destabilizing the poultry industry. Wild migratory birds are considered a natural reservoir of LPAIVs capable of dispersing both high- and low-pathogenic forms of the virus. Therefore, surveillance and characterization of AIV in wild birds are essential. Here, we report on the isolation and genetic characterization of 10 AIVs of the H5N2 subtype obtained through surveillance in Hokkaido, Japan, during 2009 and 2011. Full-genome sequencing revealed that the H5 and N2 genes of these isolates are all closely related to each other, belonging to the Eurasian avian-like lineage, but they are unrelated to H5 highly pathogenic strains of clade 2.3.4.4. The internal genes of the isolates were found to be diverse, consistent with our hypothesis that these H5N2 strains have undergone multiple reassortment events. Even though all of the H5N2 isolates were characterized as LPAIV based on the amino acid sequences at the HA cleavage site, this analysis demonstrates a diverse pool of precursors that may seed future outbreaks in poultry and possible human transmissions, suggesting the need for high-quality surveillance.},\n\tjournal = {Arch Virol},\n\tauthor = {Sultan, S. and Bui, V. N. and Hill, N. J. and Hussein, I. T. and Trinh, D. Q. and Inage, K. and Hashizume, T. and Runstadler, J. A. and Ogawa, H. and Imai, K.},\n\tmonth = aug,\n\tyear = {2016},\n}\n\n
\n
\n\n\n
\n Low-pathogenic avian influenza viruses (LPAIVs) of the H5 subtype can mutate to highly pathogenic forms, potentially destabilizing the poultry industry. Wild migratory birds are considered a natural reservoir of LPAIVs capable of dispersing both high- and low-pathogenic forms of the virus. Therefore, surveillance and characterization of AIV in wild birds are essential. Here, we report on the isolation and genetic characterization of 10 AIVs of the H5N2 subtype obtained through surveillance in Hokkaido, Japan, during 2009 and 2011. Full-genome sequencing revealed that the H5 and N2 genes of these isolates are all closely related to each other, belonging to the Eurasian avian-like lineage, but they are unrelated to H5 highly pathogenic strains of clade 2.3.4.4. The internal genes of the isolates were found to be diverse, consistent with our hypothesis that these H5N2 strains have undergone multiple reassortment events. Even though all of the H5N2 isolates were characterized as LPAIV based on the amino acid sequences at the HA cleavage site, this analysis demonstrates a diverse pool of precursors that may seed future outbreaks in poultry and possible human transmissions, suggesting the need for high-quality surveillance.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A point mutation in the polymerase protein PB2 allows a reassortant H9N2 influenza isolate of wild-bird origin to replicate in human cells.\n \n \n \n \n\n\n \n Hussein, I. T.; Ma, E. J.; Hill, N. J.; Meixell, B. W.; Lindberg, M.; Albrecht, R. A.; Bahl, J.; and Runstadler, J. A.\n\n\n \n\n\n\n Infect Genet Evol, 41: 279–88. July 2016.\n \n\n\n\n
\n\n\n\n \n \n \"APaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hussein_point_2016,\n\ttitle = {A point mutation in the polymerase protein {PB2} allows a reassortant {H9N2} influenza isolate of wild-bird origin to replicate in human cells},\n\tvolume = {41},\n\tcopyright = {All rights reserved},\n\tissn = {1567-7257 (Electronic) 1567-1348 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/27101787},\n\tdoi = {10.1016/j.meegid.2016.04.011},\n\tabstract = {H9N2 influenza A viruses are on the list of potentially pandemic subtypes. Therefore, it is important to understand how genomic reassortment and genetic polymorphisms affect phenotypes of H9N2 viruses circulating in the wild bird reservoir. A comparative genetic analysis of North American H9N2 isolates of wild bird origin identified a naturally occurring reassortant virus containing gene segments derived from both North American and Eurasian lineage ancestors. The PB2 segment of this virus encodes 10 amino acid changes that distinguish it from other H9 strains circulating in North America. G590S, one of the 10 amino acid substitutions observed, was present in {\\textasciitilde}12\\% of H9 viruses worldwide. This mutation combined with R591 has been reported as a marker of pathogenicity for human pandemic 2009 H1N1 viruses. Screening by polymerase reporter assay of all the natural polymorphisms at these two positions identified G590/K591 and S590/K591 as the most active, with the highest polymerase activity recorded for the SK polymorphism. Rescued viruses containing these two polymorphic combinations replicated more efficiently in MDCK cells and they were the only ones tested that were capable of establishing productive infection in NHBE cells. A global analysis of all PB2 sequences identified the K591 signature in six viral HA/NA subtypes isolated from several hosts in seven geographic locations. Interestingly, introducing the K591 mutation into the PB2 of a human-adapted H3N2 virus did not affect its polymerase activity. Our findings demonstrate that a single point mutation in the PB2 of a low pathogenic H9N2 isolate could have a significant effect on viral phenotype and increase its propensity to infect mammals. However, this effect is not universal, warranting caution in interpreting point mutations without considering protein sequence context.},\n\tjournal = {Infect Genet Evol},\n\tauthor = {Hussein, I. T. and Ma, E. J. and Hill, N. J. and Meixell, B. W. and Lindberg, M. and Albrecht, R. A. and Bahl, J. and Runstadler, J. A.},\n\tmonth = jul,\n\tyear = {2016},\n\tkeywords = {H9n2, Influenza, Pb2, Polymorphisms, Viral polymerase},\n\tpages = {279--88},\n}\n\n
\n
\n\n\n
\n H9N2 influenza A viruses are on the list of potentially pandemic subtypes. Therefore, it is important to understand how genomic reassortment and genetic polymorphisms affect phenotypes of H9N2 viruses circulating in the wild bird reservoir. A comparative genetic analysis of North American H9N2 isolates of wild bird origin identified a naturally occurring reassortant virus containing gene segments derived from both North American and Eurasian lineage ancestors. The PB2 segment of this virus encodes 10 amino acid changes that distinguish it from other H9 strains circulating in North America. G590S, one of the 10 amino acid substitutions observed, was present in ~12% of H9 viruses worldwide. This mutation combined with R591 has been reported as a marker of pathogenicity for human pandemic 2009 H1N1 viruses. Screening by polymerase reporter assay of all the natural polymorphisms at these two positions identified G590/K591 and S590/K591 as the most active, with the highest polymerase activity recorded for the SK polymorphism. Rescued viruses containing these two polymorphic combinations replicated more efficiently in MDCK cells and they were the only ones tested that were capable of establishing productive infection in NHBE cells. A global analysis of all PB2 sequences identified the K591 signature in six viral HA/NA subtypes isolated from several hosts in seven geographic locations. Interestingly, introducing the K591 mutation into the PB2 of a human-adapted H3N2 virus did not affect its polymerase activity. Our findings demonstrate that a single point mutation in the PB2 of a low pathogenic H9N2 isolate could have a significant effect on viral phenotype and increase its propensity to infect mammals. However, this effect is not universal, warranting caution in interpreting point mutations without considering protein sequence context.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n A bird's eye view of influenza A virus transmission: challenges with characterizing both sides of a co-evolutionary dynamic.\n \n \n \n \n\n\n \n Hill, N. J.; and Runstadler, J. A.\n\n\n \n\n\n\n Integr Comp Biol, 56(2): 304–16. June 2016.\n \n\n\n\n
\n\n\n\n \n \n \"APaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hill_birds_2016,\n\ttitle = {A bird's eye view of influenza {A} virus transmission: challenges with characterizing both sides of a co-evolutionary dynamic},\n\tvolume = {56},\n\tcopyright = {All rights reserved},\n\tissn = {1557-7023 (Electronic) 1540-7063 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/27252222},\n\tdoi = {10.1093/icb/icw055},\n\tabstract = {In nature, wild birds and influenza A viruses (IAV) are continually co-evolving, locked into a back-and-forth of resistance and conquest that has approached a stable equilibrium over time. This co-evolutionary relationship between bird host and IAV may appear stable at the organismal level, but is highly dynamic at the molecular level manifesting in a constant trade-off between transmissibility and virulence of the virus. Characterizing both sides of the host-virus dynamic has presented a challenge for ecologists and virologists alike, despite the potential for this approach to provide insights into which conditions destabilize the equilibrium state resulting in outbreaks or mortality of hosts in extreme cases. The use of different methods that are either host-centric or virus-centric has made it difficult to reconcile the disparate fields of host ecology and virology for investigating and ultimately predicting wild bird-mediated transmission of IAV. This review distills some of the key lessons learned from virological and ecological studies and explores the promises and pitfalls of both approaches. Ultimately, reconciling ecological and virological approaches hinges on integrating scales for measuring host-virus interactions. We argue that prospects for finding common scales for measuring wild bird-influenza dynamics are improving due to advances in genomic sequencing, host-tracking technology and remote sensing data, with the unit of time (months, year, or seasons) providing a starting point for crossover.},\n\tnumber = {2},\n\tjournal = {Integr Comp Biol},\n\tauthor = {Hill, N. J. and Runstadler, J. A.},\n\tmonth = jun,\n\tyear = {2016},\n\tpages = {304--16},\n}\n\n
\n
\n\n\n
\n In nature, wild birds and influenza A viruses (IAV) are continually co-evolving, locked into a back-and-forth of resistance and conquest that has approached a stable equilibrium over time. This co-evolutionary relationship between bird host and IAV may appear stable at the organismal level, but is highly dynamic at the molecular level manifesting in a constant trade-off between transmissibility and virulence of the virus. Characterizing both sides of the host-virus dynamic has presented a challenge for ecologists and virologists alike, despite the potential for this approach to provide insights into which conditions destabilize the equilibrium state resulting in outbreaks or mortality of hosts in extreme cases. The use of different methods that are either host-centric or virus-centric has made it difficult to reconcile the disparate fields of host ecology and virology for investigating and ultimately predicting wild bird-mediated transmission of IAV. This review distills some of the key lessons learned from virological and ecological studies and explores the promises and pitfalls of both approaches. Ultimately, reconciling ecological and virological approaches hinges on integrating scales for measuring host-virus interactions. We argue that prospects for finding common scales for measuring wild bird-influenza dynamics are improving due to advances in genomic sequencing, host-tracking technology and remote sensing data, with the unit of time (months, year, or seasons) providing a starting point for crossover.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Ecosystem interactions underlie the spread of avian Influenza A viruses with pandemic potential.\n \n \n \n \n\n\n \n Bahl, J.; Pham, T. T.; Hill, N. J.; Hussein, I. T.; Ma, E. J.; Easterday, B. C.; Halpin, R. A.; Stockwell, T. B.; Wentworth, D. E.; Kayali, G.; Krauss, S.; Schultz-Cherry, S.; Webster, R. G.; Webby, R. J.; Swartz, M. D.; Smith, G. J.; and Runstadler, J. A.\n\n\n \n\n\n\n PLoS Pathog, 12(5): e1005620. May 2016.\n \n\n\n\n
\n\n\n\n \n \n \"EcosystemPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bahl_ecosystem_2016,\n\ttitle = {Ecosystem interactions underlie the spread of avian {Influenza} {A} viruses with pandemic potential},\n\tvolume = {12},\n\tcopyright = {All rights reserved},\n\tissn = {1553-7374 (Electronic) 1553-7366 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/27166585 http://www.ncbi.nlm.nih.gov/pmc/articles/PMC4864295/pdf/ppat.1005620.pdf},\n\tdoi = {10.1371/journal.ppat.1005620},\n\tabstract = {Despite evidence for avian influenza A virus (AIV) transmission between wild and domestic ecosystems, the roles of bird migration and poultry trade in the spread of viruses remain enigmatic. In this study, we integrate ecosystem interactions into a phylogeographic model to assess the contribution of wild and domestic hosts to AIV distribution and persistence. Analysis of globally sampled AIV datasets shows frequent two-way transmission between wild and domestic ecosystems. In general, viral flow from domestic to wild bird populations was restricted to within a geographic region. In contrast, spillover from wild to domestic populations occurred both within and between regions. Wild birds mediated long-distance dispersal at intercontinental scales whereas viral spread among poultry populations was a major driver of regional spread. Viral spread between poultry flocks frequently originated from persistent lineages circulating in regions of intensive poultry production. Our analysis of long-term surveillance data demonstrates that meaningful insights can be inferred from integrating ecosystem into phylogeographic reconstructions that may be consequential for pandemic preparedness and livestock protection.},\n\tnumber = {5},\n\tjournal = {PLoS Pathog},\n\tauthor = {Bahl, J. and Pham, T. T. and Hill, N. J. and Hussein, I. T. and Ma, E. J. and Easterday, B. C. and Halpin, R. A. and Stockwell, T. B. and Wentworth, D. E. and Kayali, G. and Krauss, S. and Schultz-Cherry, S. and Webster, R. G. and Webby, R. J. and Swartz, M. D. and Smith, G. J. and Runstadler, J. A.},\n\tmonth = may,\n\tyear = {2016},\n\tpages = {e1005620},\n}\n
\n
\n\n\n
\n Despite evidence for avian influenza A virus (AIV) transmission between wild and domestic ecosystems, the roles of bird migration and poultry trade in the spread of viruses remain enigmatic. In this study, we integrate ecosystem interactions into a phylogeographic model to assess the contribution of wild and domestic hosts to AIV distribution and persistence. Analysis of globally sampled AIV datasets shows frequent two-way transmission between wild and domestic ecosystems. In general, viral flow from domestic to wild bird populations was restricted to within a geographic region. In contrast, spillover from wild to domestic populations occurred both within and between regions. Wild birds mediated long-distance dispersal at intercontinental scales whereas viral spread among poultry populations was a major driver of regional spread. Viral spread between poultry flocks frequently originated from persistent lineages circulating in regions of intensive poultry production. Our analysis of long-term surveillance data demonstrates that meaningful insights can be inferred from integrating ecosystem into phylogeographic reconstructions that may be consequential for pandemic preparedness and livestock protection.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2015\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Genetic characterization of a rare H12N3 avian influenza virus isolated from a green-winged teal in Japan.\n \n \n \n\n\n \n Bui, V. N.; Ogawa, H.; Hussein, I. T.; Hill, N. J.; AboElkhair, M.; Sultan, S.; Ma, E.; Saito, K.; Watanabe, Y.; and Runstadler, J. A\n\n\n \n\n\n\n Virus Genes, 50: 316–320. 2015.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{bui_genetic_2015,\n\ttitle = {Genetic characterization of a rare {H12N3} avian influenza virus isolated from a green-winged teal in {Japan}},\n\tvolume = {50},\n\tcopyright = {All rights reserved},\n\tissn = {0920-8569},\n\tjournal = {Virus Genes},\n\tauthor = {Bui, Vuong Nghia and Ogawa, Haruko and Hussein, Islam TM and Hill, Nichola J. and AboElkhair, Mohammed and Sultan, Serageldeen and Ma, Eric and Saito, Keisuke and Watanabe, Yukiko and Runstadler, Jonathan A},\n\tyear = {2015},\n\tpages = {316--320},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Stable isotopes suggest low site fidelity in Bar-headed Geese (Anser indicus) in Mongolia: implications for disease transmission.\n \n \n \n \n\n\n \n Bridge, E. S.; Kelly, J. F.; Xiao, X.; Batbayar, N.; Natsagdorj, T.; Hill, N. J.; Takekawa, J. Y.; Hawkes, L. A.; Bishop, C. M.; Butler, P. J.; and Newman, S. H.\n\n\n \n\n\n\n Waterbirds, 38(2): 123–132. June 2015.\n \n\n\n\n
\n\n\n\n \n \n \"StablePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{bridge_stable_2015,\n\ttitle = {Stable isotopes suggest low site fidelity in {Bar}-headed {Geese} ({Anser} indicus) in {Mongolia}: implications for disease transmission},\n\tvolume = {38},\n\tcopyright = {All rights reserved},\n\tissn = {1524-4695 (Print) 1524-4695 (Linking)},\n\turl = {https://www.ncbi.nlm.nih.gov/pubmed/27695389},\n\tdoi = {10.1675/063.038.0201},\n\tabstract = {Population connectivity is an important consideration in studies of disease transmission and biological conservation, especially with regard to migratory species. Determining how and when different subpopulations intermingle during different phases of the annual cycle can help identify important geographical regions or features as targets for conservation efforts and can help inform our understanding of continental-scale disease transmission. In this study, stable isotopes of hydrogen and carbon in contour feathers were used to assess the degree of molt-site fidelity among Bar-headed Geese (Anser indicus) captured in north-central Mongolia. Samples were collected from actively molting Bar-headed Geese (n = 61), and some individual samples included both a newly grown feather (still in sheath) and an old, worn feather from the bird's previous molt (n = 21). Although there was no difference in mean hydrogen isotope ratios for the old and new feathers, the isotopic variance in old feathers was approximately three times higher than that of the new feathers, which suggests that these birds use different and geographically distant molting locations from year to year. To further test this conclusion, online data and modeling tools from the isoMAP website were used to generate probability landscapes for the origin of each feather. Likely molting locations were much more widespread for old feathers than for new feathers, which supports the prospect of low molt-site fidelity. This finding indicates that population connectivity would be greater than expected based on data from a single annual cycle, and that disease spread can be rapid even in areas like Mongolia where Bar-headed Geese generally breed in small isolated groups.},\n\tnumber = {2},\n\tjournal = {Waterbirds},\n\tauthor = {Bridge, E. S. and Kelly, J. F. and Xiao, X. and Batbayar, N. and Natsagdorj, T. and Hill, N. J. and Takekawa, J. Y. and Hawkes, L. A. and Bishop, C. M. and Butler, P. J. and Newman, S. H.},\n\tmonth = jun,\n\tyear = {2015},\n\tkeywords = {Anser indicus, Bar-headed Goose, annual cycle, avian influenza, carbon, connectivity, deuterium, epidemiology, feather isotopes, molt},\n\tpages = {123--132},\n}\n\n
\n
\n\n\n
\n Population connectivity is an important consideration in studies of disease transmission and biological conservation, especially with regard to migratory species. Determining how and when different subpopulations intermingle during different phases of the annual cycle can help identify important geographical regions or features as targets for conservation efforts and can help inform our understanding of continental-scale disease transmission. In this study, stable isotopes of hydrogen and carbon in contour feathers were used to assess the degree of molt-site fidelity among Bar-headed Geese (Anser indicus) captured in north-central Mongolia. Samples were collected from actively molting Bar-headed Geese (n = 61), and some individual samples included both a newly grown feather (still in sheath) and an old, worn feather from the bird's previous molt (n = 21). Although there was no difference in mean hydrogen isotope ratios for the old and new feathers, the isotopic variance in old feathers was approximately three times higher than that of the new feathers, which suggests that these birds use different and geographically distant molting locations from year to year. To further test this conclusion, online data and modeling tools from the isoMAP website were used to generate probability landscapes for the origin of each feather. Likely molting locations were much more widespread for old feathers than for new feathers, which supports the prospect of low molt-site fidelity. This finding indicates that population connectivity would be greater than expected based on data from a single annual cycle, and that disease spread can be rapid even in areas like Mongolia where Bar-headed Geese generally breed in small isolated groups.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2014\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Prevalence and molecular identification of nematode and dipteran parasites in an Australian alpine grasshopper (Kosciuscola tristis).\n \n \n \n \n\n\n \n Umbers, K. D.; Byatt, L. J.; Hill, N. J.; Bartolini, R. J.; Hose, G. C.; Herberstein, M. E.; and Power, M. L.\n\n\n \n\n\n\n PLoS One, 10(4): e0121685. 2014.\n \n\n\n\n
\n\n\n\n \n \n \"PrevalencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{umbers_prevalence_2014,\n\ttitle = {Prevalence and molecular identification of nematode and dipteran parasites in an {Australian} alpine grasshopper ({Kosciuscola} tristis)},\n\tvolume = {10},\n\tcopyright = {All rights reserved},\n\tissn = {1932-6203 (Electronic) 1932-6203 (Linking)},\n\turl = {http://journals.plos.org/plosone/article?id=10.1371/journal.pone.0121685},\n\tdoi = {10.1371/journal.pone.0121685},\n\tabstract = {In alpine Australia, Orthoptera are abundant, dominant herbivores, important prey species, and hosts for parasites and parasitoids. Despite the central role of orthopterans in alpine ecosystems, the impact of parasites on orthopteran populations is under-explored. In this study we describe the relationship between parasite prevalence and host sex, body size and year of collection. We accessed an existing, preserved collection of 640 Kosciuscola tristis collected from across its range between 2007 and 2011. Upon dissection we collected juvenile parasites and used molecular tools to identify them to three families (Nematoda; Mermithidae, and Arthropoda: Diptera: Tachinidae and Sarcophagidae). The prevalence of nematodes ranged from 3.5\\% to 25.0\\% and dipterans from 2.4\\% to 20.0\\%. Contrary to predictions, we found no associations between parasite prevalence and grasshopper sex or size. Although there was an association between prevalence of both nematodes and dipterans with year of collection, this is likely driven by a small sample size in the first year. Our results provide a foundation for future studies into parasite prevalence within the alpine environment and the abiotic factors that might influence these associations.},\n\tnumber = {4},\n\tjournal = {PLoS One},\n\tauthor = {Umbers, K. D. and Byatt, L. J. and Hill, N. J. and Bartolini, R. J. and Hose, G. C. and Herberstein, M. E. and Power, M. L.},\n\tyear = {2014},\n\tpages = {e0121685},\n}\n\n
\n
\n\n\n
\n In alpine Australia, Orthoptera are abundant, dominant herbivores, important prey species, and hosts for parasites and parasitoids. Despite the central role of orthopterans in alpine ecosystems, the impact of parasites on orthopteran populations is under-explored. In this study we describe the relationship between parasite prevalence and host sex, body size and year of collection. We accessed an existing, preserved collection of 640 Kosciuscola tristis collected from across its range between 2007 and 2011. Upon dissection we collected juvenile parasites and used molecular tools to identify them to three families (Nematoda; Mermithidae, and Arthropoda: Diptera: Tachinidae and Sarcophagidae). The prevalence of nematodes ranged from 3.5% to 25.0% and dipterans from 2.4% to 20.0%. Contrary to predictions, we found no associations between parasite prevalence and grasshopper sex or size. Although there was an association between prevalence of both nematodes and dipterans with year of collection, this is likely driven by a small sample size in the first year. Our results provide a foundation for future studies into parasite prevalence within the alpine environment and the abiotic factors that might influence these associations.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Ectoparasite infestation patterns, haematology and serum biochemistry of urban-dwelling common brushtail possums.\n \n \n \n \n\n\n \n Webster, K. N.; Hill, N. J.; Burnett, L.; and Deane, E. M.\n\n\n \n\n\n\n Wildlife Biology, 20(4): 206–216. August 2014.\n \n\n\n\n
\n\n\n\n \n \n \"EctoparasitePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{webster_ectoparasite_2014,\n\ttitle = {Ectoparasite infestation patterns, haematology and serum biochemistry of urban-dwelling common brushtail possums},\n\tvolume = {20},\n\tcopyright = {All rights reserved},\n\tissn = {0909-6396},\n\turl = {http://dx.doi.org/10.2981/wlb.00027},\n\tdoi = {10.2981/wlb.00027},\n\tnumber = {4},\n\turldate = {2014-12-14},\n\tjournal = {Wildlife Biology},\n\tauthor = {Webster, Koa N. and Hill, Nichola J. and Burnett, Leslie and Deane, Elizabeth M.},\n\tmonth = aug,\n\tyear = {2014},\n\tpages = {206--216},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Bird migration and avian influenza: A comparison of hydrogen stable isotopes and satellite tracking methods.\n \n \n \n \n\n\n \n Bridge, E. S.; Kelly, J. F.; Xiao, X.; Takekawa, J. Y.; Hill, N. J.; Yamage, M.; Haque, E. U.; Islam, M. A.; Mundkur, T.; Yavuz, K. E.; Leader, P.; Leung, C. Y. H.; Smith, B.; Spragens, K. A.; Vandegrift, K. J.; Hosseini, P. R.; Saif, S.; Mohsanin, S.; Mikolon, A.; Islam, A.; George, A.; Sivananinthaperumal, B.; Daszak, P.; and Newman, S. H.\n\n\n \n\n\n\n Ecological Indicators, 45(0): 266–273. 2014.\n \n\n\n\n
\n\n\n\n \n \n \"BirdPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{bridge_bird_2014,\n\ttitle = {Bird migration and avian influenza: {A} comparison of hydrogen stable isotopes and satellite tracking methods},\n\tvolume = {45},\n\tcopyright = {All rights reserved},\n\tissn = {1470-160X},\n\turl = {http://www.sciencedirect.com/science/article/pii/S1470160X14001770},\n\tdoi = {10.1016/j.ecolind.2014.04.027},\n\tabstract = {Satellite-based tracking of migratory waterfowl is an important tool for understanding the potential role of wild birds in the long-distance transmission of highly pathogenic avian influenza. However, employing this technique on a continental scale is prohibitively expensive. This study explores the utility of stable isotope ratios in feathers in examining both the distances traveled by migratory birds and variation in migration behavior. We compared the satellite-derived movement data of 22 ducks from 8 species captured at wintering areas in Bangladesh, Turkey, and Hong Kong with deuterium ratios (deltaD) of these and other individuals captured at the same locations. We derived likely molting locations from the satellite tracking data and generated expected isotope ratios based on an interpolated map of deltaD in rainwater. Although deltaD was correlated with the distance between wintering and molting locations, surprisingly, measured deltaD values were not correlated with either expected values or latitudes of molting sites. However, population-level parameters derived from the satellite-tracking data, such as mean distance between wintering and molting locations and variation in migration distance, were reflected by means and variation of the stable isotope values. Our findings call into question the relevance of the rainfall isotope map for Asia for linking feather isotopes to molting locations, and underscore the need for extensive ground truthing in the form of feather-based isoscapes. Nevertheless, stable isotopes from feathers could inform disease models by characterizing the degree to which regional breeding populations interact at common wintering locations. Feather isotopes also could aid in surveying wintering locations to determine where high-resolution tracking techniques (e.g. satellite tracking) could most effectively be employed. Moreover, intrinsic markers such as stable isotopes offer the only means of inferring movement information from birds that have died as a result of infection. In the absence of feather based-isoscapes, we recommend a combination of isotope analysis and satellite-tracking as the best means of generating aggregate movement data for informing disease models.},\n\tnumber = {0},\n\tjournal = {Ecological Indicators},\n\tauthor = {Bridge, Eli S. and Kelly, Jeffrey F. and Xiao, Xiangming and Takekawa, John Y. and Hill, Nichola J. and Yamage, Mat and Haque, Enam Ul and Islam, Mohammad Anwarul and Mundkur, Taej and Yavuz, Kiraz Erciyas and Leader, Paul and Leung, Connie Y. H. and Smith, Bena and Spragens, Kyle A. and Vandegrift, Kurt J. and Hosseini, Parviez R. and Saif, Samia and Mohsanin, Samiul and Mikolon, Andrea and Islam, Ausrafal and George, Acty and Sivananinthaperumal, Balachandran and Daszak, Peter and Newman, Scott H.},\n\tyear = {2014},\n\tkeywords = {Connectivity, Deuterium, Disease modeling, Disease vector, Epidemiology, Geographical indicators, Waterfowl},\n\tpages = {266--273},\n}\n\n
\n
\n\n\n
\n Satellite-based tracking of migratory waterfowl is an important tool for understanding the potential role of wild birds in the long-distance transmission of highly pathogenic avian influenza. However, employing this technique on a continental scale is prohibitively expensive. This study explores the utility of stable isotope ratios in feathers in examining both the distances traveled by migratory birds and variation in migration behavior. We compared the satellite-derived movement data of 22 ducks from 8 species captured at wintering areas in Bangladesh, Turkey, and Hong Kong with deuterium ratios (deltaD) of these and other individuals captured at the same locations. We derived likely molting locations from the satellite tracking data and generated expected isotope ratios based on an interpolated map of deltaD in rainwater. Although deltaD was correlated with the distance between wintering and molting locations, surprisingly, measured deltaD values were not correlated with either expected values or latitudes of molting sites. However, population-level parameters derived from the satellite-tracking data, such as mean distance between wintering and molting locations and variation in migration distance, were reflected by means and variation of the stable isotope values. Our findings call into question the relevance of the rainfall isotope map for Asia for linking feather isotopes to molting locations, and underscore the need for extensive ground truthing in the form of feather-based isoscapes. Nevertheless, stable isotopes from feathers could inform disease models by characterizing the degree to which regional breeding populations interact at common wintering locations. Feather isotopes also could aid in surveying wintering locations to determine where high-resolution tracking techniques (e.g. satellite tracking) could most effectively be employed. Moreover, intrinsic markers such as stable isotopes offer the only means of inferring movement information from birds that have died as a result of infection. In the absence of feather based-isoscapes, we recommend a combination of isotope analysis and satellite-tracking as the best means of generating aggregate movement data for informing disease models.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2013\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Connecting the study of wild influenza with the potential for pandemic disease.\n \n \n \n \n\n\n \n Runstadler, J.; Hill, N.; Hussein, I. T.; Puryear, W.; and Keogh, M.\n\n\n \n\n\n\n Infect Genet Evol, 17: 162–187. March 2013.\n Edition: 2013/04/02\n\n\n\n
\n\n\n\n \n \n \"ConnectingPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{runstadler_connecting_2013,\n\ttitle = {Connecting the study of wild influenza with the potential for pandemic disease},\n\tvolume = {17},\n\tcopyright = {All rights reserved},\n\tissn = {1567-7257 (Electronic) 1567-1348 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/23541413},\n\tdoi = {10.1016/j.meegid.2013.02.020},\n\tabstract = {Continuing outbreaks of pathogenic (H5N1) and pandemic (SOIVH1N1) influenza have underscored the need to understand the origin, characteristics, and evolution of novel influenza A virus (IAV) variants that pose a threat to human health. In the last 4-5years, focus has been placed on the organization of large-scale surveillance programs to examine the phylogenetics of avian influenza virus (AIV) and host-virus relationships in domestic and wild animals. Here we review the current gaps in wild animal and environmental surveillance and the current understanding of genetic signatures in potentially pandemic strains.},\n\tlanguage = {Eng},\n\tjournal = {Infect Genet Evol},\n\tauthor = {Runstadler, J. and Hill, N.J. and Hussein, I. T. and Puryear, W. and Keogh, M.},\n\tmonth = mar,\n\tyear = {2013},\n\tnote = {Edition: 2013/04/02},\n\tpages = {162--187},\n}\n\n
\n
\n\n\n
\n Continuing outbreaks of pathogenic (H5N1) and pandemic (SOIVH1N1) influenza have underscored the need to understand the origin, characteristics, and evolution of novel influenza A virus (IAV) variants that pose a threat to human health. In the last 4-5years, focus has been placed on the organization of large-scale surveillance programs to examine the phylogenetics of avian influenza virus (AIV) and host-virus relationships in domestic and wild animals. Here we review the current gaps in wild animal and environmental surveillance and the current understanding of genetic signatures in potentially pandemic strains.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Cryptosporidium from a free-ranging marsupial host: bandicoots in urban Australia.\n \n \n \n \n\n\n \n Dowle, M.; Hill, N. J.; and Power, M. L.\n\n\n \n\n\n\n Vet Parasitol, 198(1-2): 197–200. November 2013.\n Edition: 2013/09/24\n\n\n\n
\n\n\n\n \n \n \"CryptosporidiumPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{dowle_cryptosporidium_2013,\n\ttitle = {Cryptosporidium from a free-ranging marsupial host: bandicoots in urban {Australia}},\n\tvolume = {198},\n\tcopyright = {All rights reserved},\n\tissn = {1873-2550 (Electronic) 0304-4017 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/24054949 http://www.sciencedirect.com/science/article/pii/S0304401713004809},\n\tdoi = {10.1016/j.vetpar.2013.08.017},\n\tabstract = {Expansion of human settlement has increased the interface between people and bandicoots with implications for the emergence and spread of zoonotic parasites. The host status of bandicoots inhabiting suburban areas and their potential role in Cryptosporidium transmission remains unresolved. Our study aimed to determine the prevalence and identity of Cryptosporidium in two sympatric bandicoot species. Cryptosporidium signatures were detected in twelve bandicoot faecal samples (n=98) through amplification of the 18S rRNA. Phylogenetic inference placed the isolates in a clade with Cryptosporidium parvum, a species with a broad host range and zoonotic potential, or loosely related to Cryptosporidium hominis. However, the identity of the bandicoot isolates was not fully resolved and whether they were infected or simply passively transmitting oocysts is unknown. This study revealed that free-ranging bandicoots of northern Sydney were shedding Cryptosporidium oocysts at a prevalence of 12.2\\% (95\\% CI [6.76, 20.8]), similar to marsupial species that act as reservoirs for Cryptosporidium. Our findings expand the range of hosts known to shed Cryptosporidium in urban areas.},\n\tnumber = {1-2},\n\tjournal = {Vet Parasitol},\n\tauthor = {Dowle, M. and Hill, N. J. and Power, M. L.},\n\tmonth = nov,\n\tyear = {2013},\n\tnote = {Edition: 2013/09/24},\n\tkeywords = {Bandicoot, Cryptosporidium, Emerging pathogen, Urban wildlife},\n\tpages = {197--200},\n}\n\n
\n
\n\n\n
\n Expansion of human settlement has increased the interface between people and bandicoots with implications for the emergence and spread of zoonotic parasites. The host status of bandicoots inhabiting suburban areas and their potential role in Cryptosporidium transmission remains unresolved. Our study aimed to determine the prevalence and identity of Cryptosporidium in two sympatric bandicoot species. Cryptosporidium signatures were detected in twelve bandicoot faecal samples (n=98) through amplification of the 18S rRNA. Phylogenetic inference placed the isolates in a clade with Cryptosporidium parvum, a species with a broad host range and zoonotic potential, or loosely related to Cryptosporidium hominis. However, the identity of the bandicoot isolates was not fully resolved and whether they were infected or simply passively transmitting oocysts is unknown. This study revealed that free-ranging bandicoots of northern Sydney were shedding Cryptosporidium oocysts at a prevalence of 12.2% (95% CI [6.76, 20.8]), similar to marsupial species that act as reservoirs for Cryptosporidium. Our findings expand the range of hosts known to shed Cryptosporidium in urban areas.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2012\n \n \n (5)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Role of bird movements in the epidemiology of West Nile and avian influenza virus.\n \n \n \n\n\n \n Muzaffar, S.; Hill, N.; Takekawa, J.; Perry, W. M.; Smith, L.; and Boyce, W.\n\n\n \n\n\n\n Human-Wildlife Interactions, 6(1): 72–88. 2012.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{muzaffar_role_2012,\n\ttitle = {Role of bird movements in the epidemiology of {West} {Nile} and avian influenza virus},\n\tvolume = {6},\n\tcopyright = {All rights reserved},\n\tabstract = {Avian influenza virus (AIV) is influenced by site fidelity and movements of bird hosts.\nWe examined the movement ecology of American crows (Corvus brachyrhynchos) as potential\nhosts for West Nile virus (WNV) and greater white-fronted geese (Anser albifrons frontalis) as\npotential hosts for AIVs. Research was based on radio-telemetry studies conducted in the\nCentral Valley of California, USA. While crows were restricted to a small area of only a few square\nkilometers, the distribution of the geese encompassed the northern Central Valley. The crows\nused 1.5 to 3.5 different roosting areas monthly from February through October, revealing lower\nroost fi delity than the geese that used 1.1 to 1.5 roosting areas each month from November\nthrough March. The crows moved a mean distance of 0.11 to 0.49 km/month between their\nroosting sites and 2.5 to 3.9 km/month between roosting and feeding sites. In contrast, the geese\nmoved 4.2 to 19.3 km/month between roosting areas, and their feeding range varied from 13.2\nto 19.0 km/month. Our comparison of the ecological characteristics of bird movements suggests\nthat the limited local movements of crows coupled with frequent turnover of roosts may result\nin persistence of focal areas for WNV infection. In contrast, widespread areas used by geese\nwill provide regular opportunities for intermixing of AIVs over a much greater geographic area.},\n\tnumber = {1},\n\tjournal = {Human-Wildlife Interactions},\n\tauthor = {Muzaffar, S.B. and Hill, N.J. and Takekawa, J.Y. and Perry, W. M. and Smith, L.S. and Boyce, W.M.},\n\tyear = {2012},\n\tpages = {72--88},\n}\n\n
\n
\n\n\n
\n Avian influenza virus (AIV) is influenced by site fidelity and movements of bird hosts. We examined the movement ecology of American crows (Corvus brachyrhynchos) as potential hosts for West Nile virus (WNV) and greater white-fronted geese (Anser albifrons frontalis) as potential hosts for AIVs. Research was based on radio-telemetry studies conducted in the Central Valley of California, USA. While crows were restricted to a small area of only a few square kilometers, the distribution of the geese encompassed the northern Central Valley. The crows used 1.5 to 3.5 different roosting areas monthly from February through October, revealing lower roost fi delity than the geese that used 1.1 to 1.5 roosting areas each month from November through March. The crows moved a mean distance of 0.11 to 0.49 km/month between their roosting sites and 2.5 to 3.9 km/month between roosting and feeding sites. In contrast, the geese moved 4.2 to 19.3 km/month between roosting areas, and their feeding range varied from 13.2 to 19.0 km/month. Our comparison of the ecological characteristics of bird movements suggests that the limited local movements of crows coupled with frequent turnover of roosts may result in persistence of focal areas for WNV infection. In contrast, widespread areas used by geese will provide regular opportunities for intermixing of AIVs over a much greater geographic area.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Pinning down a polymorphic parasite: new genetic and morphological descriptions of Eimeria macropodis from the Tammar wallaby (Macropus eugenii).\n \n \n \n \n\n\n \n Hill, N. J.; Richter, C.; and Power, M. L.\n\n\n \n\n\n\n Parasitology international, 61(3): 461–465. September 2012.\n Edition: 2012/04/04\n\n\n\n
\n\n\n\n \n \n \"PinningPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hill_pinning_2012,\n\ttitle = {Pinning down a polymorphic parasite: new genetic and morphological descriptions of {Eimeria} macropodis from the {Tammar} wallaby ({Macropus} eugenii)},\n\tvolume = {61},\n\tcopyright = {All rights reserved},\n\tissn = {1873-0329 (Electronic) 1383-5769 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/22469916},\n\tdoi = {10.1016/j.parint.2012.03.003},\n\tabstract = {Identification of the protozoan parasite, Eimeria has traditionally relied on oocyst morphology, host range and life-cycle attributes. However, it is increasingly recognized that Eimeria species can vary in size and shape across their host range, an attribute known as 'polymorphism' that presents a unique challenge for identification. Advances in molecular tools hold promise for characterising Eimeria that may otherwise be misclassified based on morphology. Our study used morphologic and molecular traits of the oocyst life stage to identify a polymorphic parasite, Eimeria macropodis in a captive Tammar wallaby (Macropus eugenii) population in Australia. Molecular characterization highlighted the need to use multiple genetic markers (18S SSU and cytochrome c oxidase subunit I) to accurately identify E. macropodis owing to heterozygous alleles at the 18S SSU locus. This study provided an opportunity to assess the utility and shortcomings of morphologic and molecular techniques for 'pinning down' a polymorphic species. Moreover, our study was able to place E. macropodis in an evolutionary context and enhance resolution of the under-studied marsupial clade.},\n\tlanguage = {eng},\n\tnumber = {3},\n\tjournal = {Parasitology international},\n\tauthor = {Hill, N. J. and Richter, C. and Power, M. L.},\n\tmonth = sep,\n\tyear = {2012},\n\tnote = {Edition: 2012/04/04},\n\tpages = {461--465},\n}\n\n
\n
\n\n\n
\n Identification of the protozoan parasite, Eimeria has traditionally relied on oocyst morphology, host range and life-cycle attributes. However, it is increasingly recognized that Eimeria species can vary in size and shape across their host range, an attribute known as 'polymorphism' that presents a unique challenge for identification. Advances in molecular tools hold promise for characterising Eimeria that may otherwise be misclassified based on morphology. Our study used morphologic and molecular traits of the oocyst life stage to identify a polymorphic parasite, Eimeria macropodis in a captive Tammar wallaby (Macropus eugenii) population in Australia. Molecular characterization highlighted the need to use multiple genetic markers (18S SSU and cytochrome c oxidase subunit I) to accurately identify E. macropodis owing to heterozygous alleles at the 18S SSU locus. This study provided an opportunity to assess the utility and shortcomings of morphologic and molecular techniques for 'pinning down' a polymorphic species. Moreover, our study was able to place E. macropodis in an evolutionary context and enhance resolution of the under-studied marsupial clade.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Migration strategy affects avian influenza dynamics in mallards (Anas platyrhynchos).\n \n \n \n\n\n \n Hill, N.; Takekawa, J.; Ackerman, J.; Hobson, K.; Herring, G.; Cardona, C.; Runstadler, J.; and Boyce, W.\n\n\n \n\n\n\n Molecular Ecology, 21(24): 5986–5999. 2012.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_migration_2012,\n\ttitle = {Migration strategy affects avian influenza dynamics in mallards ({Anas} platyrhynchos)},\n\tvolume = {21},\n\tcopyright = {All rights reserved},\n\tdoi = {10.1111/j.1365-294X.2012.05735.x},\n\tnumber = {24},\n\tjournal = {Molecular Ecology},\n\tauthor = {Hill, N.J. and Takekawa, J.Y. and Ackerman, J.T. and Hobson, K.A. and Herring, G. and Cardona, C.J. and Runstadler, J.A. and Boyce, W.M.},\n\tyear = {2012},\n\tkeywords = {*Animal Migration, *Genetic Variation, Animals, Animals, Wild/virology, California/epidemiology, Ducks/*virology, Feathers, Gene Flow, Influenza A virus/*genetics, Influenza in Birds/epidemiology/*virology, Phylogeny, Prevalence, Seasons},\n\tpages = {5986--5999},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Eco-virological approach for assessing the role of wild birds in the spread of avian influenza H5N1 along the Central Asian Flyway.\n \n \n \n \n\n\n \n Newman, S. H.; Hill, N. J.; Spragens, K. A.; Janies, D.; Voronkin, I. O.; Prosser, D. J.; Yan, B.; Lei, F.; Batbayar, N.; Natsagdorj, T.; Bishop, C. M.; Butler, P. J.; Wikelski, M.; Balachandran, S.; Mundkur, T.; Douglas, D. C.; and Takekawa, J. Y.\n\n\n \n\n\n\n PLoS ONE, 7(2): e30636. 2012.\n Edition: 2012/02/22\n\n\n\n
\n\n\n\n \n \n \"Eco-virologicalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{newman_eco-virological_2012,\n\ttitle = {Eco-virological approach for assessing the role of wild birds in the spread of avian influenza {H5N1} along the {Central} {Asian} {Flyway}},\n\tvolume = {7},\n\tcopyright = {All rights reserved},\n\tissn = {1932-6203 (Electronic) 1932-6203 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/22347393},\n\tdoi = {10.1371/journal.pone.0030636},\n\tabstract = {A unique pattern of highly pathogenic avian influenza (HPAI) H5N1 outbreaks has emerged along the Central Asia Flyway, where infection of wild birds has been reported with steady frequency since 2005. We assessed the potential for two hosts of HPAI H5N1, the bar-headed goose (Anser indicus) and ruddy shelduck (Tadorna tadorna), to act as agents for virus dispersal along this 'thoroughfare'. We used an eco-virological approach to compare the migration of 141 birds marked with GPS satellite transmitters during 2005-2010 with: 1) the spatio-temporal patterns of poultry and wild bird outbreaks of HPAI H5N1, and 2) the trajectory of the virus in the outbreak region based on phylogeographic mapping. We found that biweekly utilization distributions (UDs) for 19.2\\% of bar-headed geese and 46.2\\% of ruddy shelduck were significantly associated with outbreaks. Ruddy shelduck showed highest correlation with poultry outbreaks owing to their wintering distribution in South Asia, where there is considerable opportunity for HPAI H5N1 spillover from poultry. Both species showed correlation with wild bird outbreaks during the spring migration, suggesting they may be involved in the northward movement of the virus. However, phylogeographic mapping of HPAI H5N1 clades 2.2 and 2.3 did not support dissemination of the virus in a northern direction along the migration corridor. In particular, two subclades (2.2.1 and 2.3.2) moved in a strictly southern direction in contrast to our spatio-temporal analysis of bird migration. Our attempt to reconcile the disciplines of wild bird ecology and HPAI H5N1 virology highlights prospects offered by both approaches as well as their limitations.},\n\tlanguage = {eng},\n\tnumber = {2},\n\tjournal = {PLoS ONE},\n\tauthor = {Newman, S. H. and Hill, N. J. and Spragens, K. A. and Janies, D. and Voronkin, I. O. and Prosser, D. J. and Yan, B. and Lei, F. and Batbayar, N. and Natsagdorj, T. and Bishop, C. M. and Butler, P. J. and Wikelski, M. and Balachandran, S. and Mundkur, T. and Douglas, D. C. and Takekawa, J. Y.},\n\tyear = {2012},\n\tnote = {Edition: 2012/02/22},\n\tpages = {e30636},\n}\n\n
\n
\n\n\n
\n A unique pattern of highly pathogenic avian influenza (HPAI) H5N1 outbreaks has emerged along the Central Asia Flyway, where infection of wild birds has been reported with steady frequency since 2005. We assessed the potential for two hosts of HPAI H5N1, the bar-headed goose (Anser indicus) and ruddy shelduck (Tadorna tadorna), to act as agents for virus dispersal along this 'thoroughfare'. We used an eco-virological approach to compare the migration of 141 birds marked with GPS satellite transmitters during 2005-2010 with: 1) the spatio-temporal patterns of poultry and wild bird outbreaks of HPAI H5N1, and 2) the trajectory of the virus in the outbreak region based on phylogeographic mapping. We found that biweekly utilization distributions (UDs) for 19.2% of bar-headed geese and 46.2% of ruddy shelduck were significantly associated with outbreaks. Ruddy shelduck showed highest correlation with poultry outbreaks owing to their wintering distribution in South Asia, where there is considerable opportunity for HPAI H5N1 spillover from poultry. Both species showed correlation with wild bird outbreaks during the spring migration, suggesting they may be involved in the northward movement of the virus. However, phylogeographic mapping of HPAI H5N1 clades 2.2 and 2.3 did not support dissemination of the virus in a northern direction along the migration corridor. In particular, two subclades (2.2.1 and 2.3.2) moved in a strictly southern direction in contrast to our spatio-temporal analysis of bird migration. Our attempt to reconcile the disciplines of wild bird ecology and HPAI H5N1 virology highlights prospects offered by both approaches as well as their limitations.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Cross-seasonal patterns of avian influenza virus in breeding and wintering migratory birds: a flyway perspective.\n \n \n \n \n\n\n \n Hill, N. J.; Takekawa, J. Y.; Cardona, C. J.; Meixell, B. W.; Ackerman, J. T.; Runstadler, J. A.; and Boyce, W. M.\n\n\n \n\n\n\n Vector Borne and Zoonotic Diseases, 12(3): 243–253. October 2012.\n Edition: 2011/10/15\n\n\n\n
\n\n\n\n \n \n \"Cross-seasonalPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hill_cross-seasonal_2012,\n\ttitle = {Cross-seasonal patterns of avian influenza virus in breeding and wintering migratory birds: a flyway perspective},\n\tvolume = {12},\n\tcopyright = {All rights reserved},\n\tissn = {1557-7759 (Electronic) 1530-3667 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/21995264},\n\tdoi = {10.1089/vbz.2010.0246},\n\tabstract = {Abstract The spread of avian influenza viruses (AIV) in nature is intrinsically linked with the movements of wild birds. Wild birds are the reservoirs for the virus and their migration may facilitate the circulation of AIV between breeding and wintering areas. This cycle of dispersal has become widely accepted; however, there are few AIV studies that present cross-seasonal information. A flyway perspective is critical for understanding how wild birds contribute to the persistence of AIV over large spatial and temporal scales, with implications for how to focus surveillance efforts and identify risks to public health. This study characterized spatio-temporal infection patterns in 10,389 waterfowl at two important locations within the Pacific Flyway-breeding sites in Interior Alaska and wintering sites in California's Central Valley during 2007-2009. Among the dabbling ducks sampled, the northern shoveler (Anas clypeata) had the highest prevalence of AIV at both breeding (32.2\\%) and wintering (5.2\\%) locations. This is in contrast to surveillance studies conducted in other flyways that have identified the mallard (Anas platyrhynchos) and northern pintail (Anas acuta) as hosts with the highest prevalence. A higher diversity of AIV subtypes was apparent at wintering (n=42) compared with breeding sites (n=17), with evidence of mixed infections at both locations. Our study suggests that wintering sites may act as an important mixing bowl for transmission among waterfowl in a flyway, creating opportunities for the reassortment of the virus. Our findings shed light on how the dynamics of AIV infection of wild bird populations can vary between the two ends of a migratory flyway.},\n\tlanguage = {Eng},\n\tnumber = {3},\n\tjournal = {Vector Borne and Zoonotic Diseases},\n\tauthor = {Hill, N. J. and Takekawa, J. Y. and Cardona, C. J. and Meixell, B. W. and Ackerman, J. T. and Runstadler, J. A. and Boyce, W. M.},\n\tmonth = oct,\n\tyear = {2012},\n\tnote = {Edition: 2011/10/15},\n\tpages = {243--253},\n}\n\n
\n
\n\n\n
\n Abstract The spread of avian influenza viruses (AIV) in nature is intrinsically linked with the movements of wild birds. Wild birds are the reservoirs for the virus and their migration may facilitate the circulation of AIV between breeding and wintering areas. This cycle of dispersal has become widely accepted; however, there are few AIV studies that present cross-seasonal information. A flyway perspective is critical for understanding how wild birds contribute to the persistence of AIV over large spatial and temporal scales, with implications for how to focus surveillance efforts and identify risks to public health. This study characterized spatio-temporal infection patterns in 10,389 waterfowl at two important locations within the Pacific Flyway-breeding sites in Interior Alaska and wintering sites in California's Central Valley during 2007-2009. Among the dabbling ducks sampled, the northern shoveler (Anas clypeata) had the highest prevalence of AIV at both breeding (32.2%) and wintering (5.2%) locations. This is in contrast to surveillance studies conducted in other flyways that have identified the mallard (Anas platyrhynchos) and northern pintail (Anas acuta) as hosts with the highest prevalence. A higher diversity of AIV subtypes was apparent at wintering (n=42) compared with breeding sites (n=17), with evidence of mixed infections at both locations. Our study suggests that wintering sites may act as an important mixing bowl for transmission among waterfowl in a flyway, creating opportunities for the reassortment of the virus. Our findings shed light on how the dynamics of AIV infection of wild bird populations can vary between the two ends of a migratory flyway.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2011\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Wild bird migration across the Qinghai-Tibetan Plateau: A transmission route for highly pathogenic H5N1.\n \n \n \n\n\n \n Prosser, D. J.; Cui, P.; Takekawa, J. Y.; Tang, M.; Hou, Y.; Collins, B. M.; Yan, B.; Hill, N. J.; Li, T.; Li, Y.; Lei, F.; Guo, S.; Xing, Z.; He, Y.; Zhou, Y.; Douglas, D. C.; Perry, W. M.; and Newman, S. H.\n\n\n \n\n\n\n PLoS ONE, 6(3). 2011.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{prosser_wild_2011,\n\ttitle = {Wild bird migration across the {Qinghai}-{Tibetan} {Plateau}: {A} transmission route for highly pathogenic {H5N1}},\n\tvolume = {6},\n\tcopyright = {All rights reserved},\n\tdoi = {10.1371/journal.pone.0017622},\n\tabstract = {Background: Qinghai Lake in central China has been at the center of debate on whether wild birds play a role in circulation of highly pathogenic avian influenza virus H5N1. In 2005, an unprecedented epizootic at Qinghai Lake killed more than 6000 migratory birds including over 3000 bar-headed geese (Anser indicus). H5N1 subsequently spread to Europe and Africa, and in following years has re-emerged in wild birds along the Central Asia flyway several times. Methodology/Principal Findings: To better understand the potential involvement of wild birds in the spread of H5N1, we studied the movements of bar-headed geese marked with GPS satellite transmitters at Qinghai Lake in relation to virus outbreaks and disease risk factors. We discovered a previously undocumented migratory pathway between Qinghai Lake and the Lhasa Valley of Tibet where 93\\% of the 29 marked geese overwintered. From 2003-2009, sixteen outbreaks in poultry or wild birds were confirmed on the Qinghai-Tibet Plateau, and the majority were located within the migratory pathway of the geese. Spatial and temporal concordance between goose movements and three potential H5N1 virus sources (poultry farms, a captive bar-headed goose facility, and H5N1 outbreak locations) indicated ample opportunities existed for virus spillover and infection of migratory geese on the wintering grounds. Their potential as a vector of H5N1 was supported by rapid migration movements of some geese and genetic relatedness of H5N1 virus isolated from geese in Tibet and Qinghai Lake. Conclusions/Significance: This is the first study to compare phylogenetics of the virus with spatial ecology of its host, and the combined results suggest that wild birds play a role in the spread of H5N1 in this region. However, the strength of the evidence would be improved with additional sequences from both poultry and wild birds on the Qinghai-Tibet Plateau where H5N1 has a clear stronghold.},\n\tnumber = {3},\n\tjournal = {PLoS ONE},\n\tauthor = {Prosser, D. J. and Cui, P. and Takekawa, J. Y. and Tang, M. and Hou, Y. and Collins, B. M. and Yan, B. and Hill, N. J. and Li, T. and Li, Y. and Lei, F. and Guo, S. and Xing, Z. and He, Y. and Zhou, Y. and Douglas, D. C. and Perry, W. M. and Newman, S. H.},\n\tyear = {2011},\n}\n\n
\n
\n\n\n
\n Background: Qinghai Lake in central China has been at the center of debate on whether wild birds play a role in circulation of highly pathogenic avian influenza virus H5N1. In 2005, an unprecedented epizootic at Qinghai Lake killed more than 6000 migratory birds including over 3000 bar-headed geese (Anser indicus). H5N1 subsequently spread to Europe and Africa, and in following years has re-emerged in wild birds along the Central Asia flyway several times. Methodology/Principal Findings: To better understand the potential involvement of wild birds in the spread of H5N1, we studied the movements of bar-headed geese marked with GPS satellite transmitters at Qinghai Lake in relation to virus outbreaks and disease risk factors. We discovered a previously undocumented migratory pathway between Qinghai Lake and the Lhasa Valley of Tibet where 93% of the 29 marked geese overwintered. From 2003-2009, sixteen outbreaks in poultry or wild birds were confirmed on the Qinghai-Tibet Plateau, and the majority were located within the migratory pathway of the geese. Spatial and temporal concordance between goose movements and three potential H5N1 virus sources (poultry farms, a captive bar-headed goose facility, and H5N1 outbreak locations) indicated ample opportunities existed for virus spillover and infection of migratory geese on the wintering grounds. Their potential as a vector of H5N1 was supported by rapid migration movements of some geese and genetic relatedness of H5N1 virus isolated from geese in Tibet and Qinghai Lake. Conclusions/Significance: This is the first study to compare phylogenetics of the virus with spatial ecology of its host, and the combined results suggest that wild birds play a role in the spread of H5N1 in this region. However, the strength of the evidence would be improved with additional sequences from both poultry and wild birds on the Qinghai-Tibet Plateau where H5N1 has a clear stronghold.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rapid diagnosis of avian influenza virus in wild birds: use of a portable rRT-PCR and freeze-dried reagents in the field.\n \n \n \n \n\n\n \n Takekawa, J. Y.; Hill, N. J.; Schultz, A. K.; Iverson, S. A.; Cardona, C. J.; Boyce, W. M.; and Dudley, J. P.\n\n\n \n\n\n\n Journal of Visualized Experiments, (54): e2829. 2011.\n \n\n\n\n
\n\n\n\n \n \n \"RapidPaper\n  \n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{takekawa_rapid_2011,\n\ttitle = {Rapid diagnosis of avian influenza virus in wild birds: use of a portable {rRT}-{PCR} and freeze-dried reagents in the field},\n\tcopyright = {All rights reserved},\n\tissn = {1940-087X},\n\turl = {http://www.jove.com/details.stp?id=2829},\n\tabstract = {Wild birds have been implicated in the spread of highly pathogenic avian influenza (HPAI) of the H5N1 subtype, prompting surveillance along migratory flyways. Sampling of wild birds for avian influenza virus (AIV) is often conducted in remote regions, but results are often delayed because of the need to transport samples to a laboratory equipped for molecular testing. Real-time reverse transcriptase polymerase chain reaction (rRT-PCR) is a molecular technique that offers one of the most accurate and sensitive methods for diagnosis of AIV. The previously strict lab protocols needed for rRT-PCR are now being adapted for the field. Development of freeze-dried (lyophilized) reagents that do not require cold chain, with sensitivity at the level of wet reagents has brought on-site remote testing to a practical goal. Here we present a method for the rapid diagnosis of AIV in wild birds using an rRT-PCR unit (Ruggedized Advanced Pathogen Identification Device or RAPID, Idaho Technologies, Salt Lake City, UT) that employs lyophilized reagents (Influenza A Target 1 Taqman; ASAY-ASY-0109, Idaho Technologies). The reagents contain all of the necessary components for testing at appropriate concentrations in a single tube: primers, probes, enzymes, buffers and internal positive controls, eliminating errors associated with improper storage or handling of wet reagents. The portable unit performs a screen for Influenza A by targeting the matrix gene and yields results in 2-3 hours. Genetic subtyping is also possible with H5 and H7 primer sets that target the hemagglutinin gene. The system is suitable for use on cloacal and oropharyngeal samples collected from wild birds, as demonstrated here on the migratory shorebird species, the western sandpiper (Calidrus mauri) captured in Northern California. Animal handling followed protocols approved by the Animal Care and Use Committee of the U.S. Geological Survey Western Ecological Research Center and permits of the U.S. Geological Survey Bird Banding Laboratory. The primary advantage of this technique is to expedite diagnosis of wild birds, increasing the chances of containing an outbreak in a remote location. On-site diagnosis would also prove useful for identifying and studying infected individuals in wild populations. The opportunity to collect information on host biology (immunological and physiological response to infection) and spatial ecology (migratory performance of infected birds) will provide insights into the extent to which wild birds can act as vectors for AIV over long distances.},\n\tnumber = {54},\n\tjournal = {Journal of Visualized Experiments},\n\tauthor = {Takekawa, John Y. and Hill, Nichola J. and Schultz, Annie K. and Iverson, Samuel A. and Cardona, Carol J. and Boyce, Walter M. and Dudley, Joseph P.},\n\tyear = {2011},\n\tkeywords = {H5N1, Immunology, active surveillance, avian influenza, lyophilized reagents, migratory birds},\n\tpages = {e2829},\n}\n\n
\n
\n\n\n
\n Wild birds have been implicated in the spread of highly pathogenic avian influenza (HPAI) of the H5N1 subtype, prompting surveillance along migratory flyways. Sampling of wild birds for avian influenza virus (AIV) is often conducted in remote regions, but results are often delayed because of the need to transport samples to a laboratory equipped for molecular testing. Real-time reverse transcriptase polymerase chain reaction (rRT-PCR) is a molecular technique that offers one of the most accurate and sensitive methods for diagnosis of AIV. The previously strict lab protocols needed for rRT-PCR are now being adapted for the field. Development of freeze-dried (lyophilized) reagents that do not require cold chain, with sensitivity at the level of wet reagents has brought on-site remote testing to a practical goal. Here we present a method for the rapid diagnosis of AIV in wild birds using an rRT-PCR unit (Ruggedized Advanced Pathogen Identification Device or RAPID, Idaho Technologies, Salt Lake City, UT) that employs lyophilized reagents (Influenza A Target 1 Taqman; ASAY-ASY-0109, Idaho Technologies). The reagents contain all of the necessary components for testing at appropriate concentrations in a single tube: primers, probes, enzymes, buffers and internal positive controls, eliminating errors associated with improper storage or handling of wet reagents. The portable unit performs a screen for Influenza A by targeting the matrix gene and yields results in 2-3 hours. Genetic subtyping is also possible with H5 and H7 primer sets that target the hemagglutinin gene. The system is suitable for use on cloacal and oropharyngeal samples collected from wild birds, as demonstrated here on the migratory shorebird species, the western sandpiper (Calidrus mauri) captured in Northern California. Animal handling followed protocols approved by the Animal Care and Use Committee of the U.S. Geological Survey Western Ecological Research Center and permits of the U.S. Geological Survey Bird Banding Laboratory. The primary advantage of this technique is to expedite diagnosis of wild birds, increasing the chances of containing an outbreak in a remote location. On-site diagnosis would also prove useful for identifying and studying infected individuals in wild populations. The opportunity to collect information on host biology (immunological and physiological response to infection) and spatial ecology (migratory performance of infected birds) will provide insights into the extent to which wild birds can act as vectors for AIV over long distances.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2010\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Waterfowl ecology and avian influenza in California: do host traits inform us about viral occurrence?.\n \n \n \n\n\n \n Hill, N.; Takekawa, J.; Cardona, C.; Ackerman, J.; Schultz, A.; Spragens, K.; and Boyce, W.\n\n\n \n\n\n\n Avian Dis., 54: 426–432. 2010.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hill_waterfowl_2010,\n\ttitle = {Waterfowl ecology and avian influenza in {California}: do host traits inform us about viral occurrence?},\n\tvolume = {54},\n\tcopyright = {All rights reserved},\n\tabstract = {We examined whether host traits influenced the occurrence of avian influenza virus (AIV) in Anatidae (ducks, geese, swans) at wintering sites in California’s Central Valley. In total, 3487 individuals were sampled at Sacramento National Wildlife Refuge and Conaway Ranch Duck Club during the hunting season of 2007–08. Of the 19 Anatidae species sampled,\nprevalence was highest in the northern shoveler (5.09\\%), followed by the ring-necked duck (2.63\\%), American wigeon (2.57\\%), bufflehead (2.50\\%), greater white-fronted goose (2.44\\%), and cinnamon teal (1.72\\%). Among host traits, density of lamellae (filtering plates) of dabbling ducks was significantly associated with AIV prevalence and the number of subtypes shed by the host, suggesting that feeding methods may influence exposure to viral particles.},\n\tjournal = {Avian Dis.},\n\tauthor = {Hill, N.J. and Takekawa, J.Y. and Cardona, C.J. and Ackerman, J.T. and Schultz, A.K. and Spragens, K.A. and Boyce, W.M.},\n\tyear = {2010},\n\tpages = {426--432},\n}\n\n
\n
\n\n\n
\n We examined whether host traits influenced the occurrence of avian influenza virus (AIV) in Anatidae (ducks, geese, swans) at wintering sites in California’s Central Valley. In total, 3487 individuals were sampled at Sacramento National Wildlife Refuge and Conaway Ranch Duck Club during the hunting season of 2007–08. Of the 19 Anatidae species sampled, prevalence was highest in the northern shoveler (5.09%), followed by the ring-necked duck (2.63%), American wigeon (2.57%), bufflehead (2.50%), greater white-fronted goose (2.44%), and cinnamon teal (1.72%). Among host traits, density of lamellae (filtering plates) of dabbling ducks was significantly associated with AIV prevalence and the number of subtypes shed by the host, suggesting that feeding methods may influence exposure to viral particles.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Victims and vectors: highly pathogenic avian influenza H5N1 and the ecology of wild birds.\n \n \n \n\n\n \n Takekawa, J.; Prosser, D. J.; Newman, S. H.; Muzaffar, S.; Hill, N.; Yan, B; Xiao, X.; Lei, F.; Li, T.; Schwarzbach, S. E.; and Howell, J.\n\n\n \n\n\n\n Avian Biol Res, 3(2): 51–73. 2010.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{takekawa_victims_2010,\n\ttitle = {Victims and vectors: highly pathogenic avian influenza {H5N1} and the ecology of wild birds},\n\tvolume = {3},\n\tcopyright = {All rights reserved},\n\tnumber = {2},\n\tjournal = {Avian Biol Res},\n\tauthor = {Takekawa, J.Y. and Prosser, D. J. and Newman, S. H. and Muzaffar, S.B. and Hill, N.J. and Yan, B and Xiao, X. and Lei, F. and Li, T. and Schwarzbach, S. E. and Howell, J.A.},\n\tyear = {2010},\n\tpages = {51--73},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Field detection of avian influenza virus in wild birds: evaluation of a portable rRT-PCR system and freeze-dried reagents.\n \n \n \n \n\n\n \n Takekawa, J. Y.; Iverson, S. A.; Schultz, A. K.; Hill, N. J.; Cardona, C. J.; Boyce, W. M.; and Dudley, J. P.\n\n\n \n\n\n\n J Virol Methods, 166: 92–97. March 2010.\n Edition: 2010/03/09\n\n\n\n
\n\n\n\n \n \n \"FieldPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{takekawa_field_2010,\n\ttitle = {Field detection of avian influenza virus in wild birds: evaluation of a portable {rRT}-{PCR} system and freeze-dried reagents},\n\tvolume = {166},\n\tcopyright = {All rights reserved},\n\tissn = {1879-0984 (Electronic) 0166-0934 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=20206650},\n\tdoi = {10.1016/j.jviromet.2010.02.029},\n\tabstract = {Wild birds have been implicated in the spread of highly pathogenic avian influenza (HPAIV) of the H5N1 subtype, prompting surveillance along migratory flyways. Sampling of wild birds is often conducted in remote regions, but results are often delayed because of limited local analytical capabilities, difficulties with sample transportation and permitting, or problems keeping samples cold in the field. In response to these challenges, the performance of a portable real-time, reverse transcriptase-polymerase chain reaction (rRT-PCR) unit (RAPID((R)), Idaho Technologies, Salt Lake City, UT) that employed lyophilized reagents (Influenza A Target 1 Taqman; ASAY-ASY-0109, Idaho Technologies) was compared to virus isolation combined with real-time RT-PCR conducted in a laboratory. This study included both field- and experimental-based sampling. Field samples were collected from migratory shorebirds captured in northern California, while experimental samples were prepared by spiking fecal material with an H6N2 AIV isolate. Results indicated that the portable rRT-PCR unit had equivalent specificity to virus isolation with no false positives, but sensitivity was compromised at low viral titers. Use of portable rRT-PCR with lyophilized reagents may expedite surveillance results, paving the way to a better understanding of wild bird involvement in HPAIV H5N1 transmission.},\n\tlanguage = {Eng},\n\tjournal = {J Virol Methods},\n\tauthor = {Takekawa, J. Y. and Iverson, S. A. and Schultz, A. K. and Hill, N. J. and Cardona, C. J. and Boyce, W. M. and Dudley, J. P.},\n\tmonth = mar,\n\tyear = {2010},\n\tnote = {Edition: 2010/03/09},\n\tpages = {92--97},\n}\n\n
\n
\n\n\n
\n Wild birds have been implicated in the spread of highly pathogenic avian influenza (HPAIV) of the H5N1 subtype, prompting surveillance along migratory flyways. Sampling of wild birds is often conducted in remote regions, but results are often delayed because of limited local analytical capabilities, difficulties with sample transportation and permitting, or problems keeping samples cold in the field. In response to these challenges, the performance of a portable real-time, reverse transcriptase-polymerase chain reaction (rRT-PCR) unit (RAPID((R)), Idaho Technologies, Salt Lake City, UT) that employed lyophilized reagents (Influenza A Target 1 Taqman; ASAY-ASY-0109, Idaho Technologies) was compared to virus isolation combined with real-time RT-PCR conducted in a laboratory. This study included both field- and experimental-based sampling. Field samples were collected from migratory shorebirds captured in northern California, while experimental samples were prepared by spiking fecal material with an H6N2 AIV isolate. Results indicated that the portable rRT-PCR unit had equivalent specificity to virus isolation with no false positives, but sensitivity was compromised at low viral titers. Use of portable rRT-PCR with lyophilized reagents may expedite surveillance results, paving the way to a better understanding of wild bird involvement in HPAIV H5N1 transmission.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2009\n \n \n (1)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Absence of Ross River virus amongst Common brushtail possums (Trichosurus vulpecula) from metropolitan Sydney, Australia.\n \n \n \n \n\n\n \n Hill, N. J.; Power, M. L.; and Deane, E. M.\n\n\n \n\n\n\n European Journal of Wildlife Research, 55(3): 313–316. June 2009.\n \n\n\n\n
\n\n\n\n \n \n \"AbsencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_absence_2009,\n\ttitle = {Absence of {Ross} {River} virus amongst {Common} brushtail possums ({Trichosurus} vulpecula) from metropolitan {Sydney}, {Australia}},\n\tvolume = {55},\n\tcopyright = {All rights reserved},\n\tissn = {1612-4642},\n\turl = {://WOS:000266476800016},\n\tdoi = {10.1007/s10344-008-0238-z},\n\tabstract = {Ross River virus (RRV) is a mosquito-transmitted Alphavirus emerging in urban centres throughout Australia. The Common brushtail possum (Trichosurus vulpecula), a native marsupial that has successfully adapted to human settlement, has been implicated as a maintenance reservoir for RRV. In the present study, RRV exposure was assessed amongst 72 urban-adapted possums from Northern Sydney and ten possums from a woodland area, remote from urbanisation. Serological screening was performed using an enzyme-linked immunosorbent assay to detect RRV antibodies in possum sera. Findings indicated that both possum populations from urban and woodland habitats were negative for the presence of RRV antibodies. Lack of exposure to RRV highlights that the host status of possums is contingent upon factors other than their abundance and proximity to human settlement. In view of the potential for climate change to favour transmission of mosquito-borne disease in Australia, identification of wildlife populations entirely absent of RRV may prove useful for monitoring the predicted spread of the virus.},\n\tlanguage = {English},\n\tnumber = {3},\n\tjournal = {European Journal of Wildlife Research},\n\tauthor = {Hill, N. J. and Power, M. L. and Deane, E. M.},\n\tmonth = jun,\n\tyear = {2009},\n\tkeywords = {Arbovirus, Environmental Sciences \\& Ecology, Mosquito, Serology, Urban wildlife, Zoology, arboviruses, climate, disease, epidemics, infection, mosquitos, new-south-wales, outbreaks, queensland},\n\tpages = {313--316},\n}\n\n
\n
\n\n\n
\n Ross River virus (RRV) is a mosquito-transmitted Alphavirus emerging in urban centres throughout Australia. The Common brushtail possum (Trichosurus vulpecula), a native marsupial that has successfully adapted to human settlement, has been implicated as a maintenance reservoir for RRV. In the present study, RRV exposure was assessed amongst 72 urban-adapted possums from Northern Sydney and ten possums from a woodland area, remote from urbanisation. Serological screening was performed using an enzyme-linked immunosorbent assay to detect RRV antibodies in possum sera. Findings indicated that both possum populations from urban and woodland habitats were negative for the presence of RRV antibodies. Lack of exposure to RRV highlights that the host status of possums is contingent upon factors other than their abundance and proximity to human settlement. In view of the potential for climate change to favour transmission of mosquito-borne disease in Australia, identification of wildlife populations entirely absent of RRV may prove useful for monitoring the predicted spread of the virus.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2008\n \n \n (2)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n \n Prevalence and genetic characterization of Cryptosporidium isolates from common brushtail possums (Trichosurus vulpecula) adapted to urban settings.\n \n \n \n \n\n\n \n Hill, N. J.; Deane, E. M.; and Power, M. L.\n\n\n \n\n\n\n Appl Environ Microbiol, 74(17): 5549–55. September 2008.\n Edition: 2008/07/22\n\n\n\n
\n\n\n\n \n \n \"PrevalencePaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_prevalence_2008,\n\ttitle = {Prevalence and genetic characterization of {Cryptosporidium} isolates from common brushtail possums ({Trichosurus} vulpecula) adapted to urban settings},\n\tvolume = {74},\n\tcopyright = {All rights reserved},\n\tissn = {1098-5336 (Electronic)},\n\turl = {http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=18641156 http://www.ncbi.nlm.nih.gov/pmc/articles/PMC2546644/pdf/0809-08.pdf},\n\tdoi = {10.1128/AEM.00809-08},\n\tabstract = {The common brushtail possum (Trichosurus vulpecula) is one of the most abundant native marsupials in urban Australia, having successfully adapted to utilize anthropogenic resources. The habituation of possums to food and shelter available in human settlements has facilitated interaction with people, pets, and zoo animals, increasing the potential for transmission of zoonotic Cryptosporidium pathogens. This study sought to examine the identity and prevalence of Cryptosporidium species occurring in possums adapted to urban settings compared to possums inhabiting remote woodlands far from urban areas and to characterize the health of the host in response to oocyst shedding. Findings indicated that both populations were shedding oocysts of the same genotype (brushtail possum 1 [BTP1]) that were genetically and morphologically distinct from zoonotic species and genotypes and most closely related to Cryptosporidium species from marsupials. The urban population was shedding an additional five Cryptosporidium isolates that were genetically distinct from BTP1 and formed a sister clade with Cryptosporidium parvum and Cryptosporidium hominis. Possums that were shedding oocysts showed no evidence of pathogenic changes, including elevated levels of white blood cells, diminished body condition (body mass divided by skeletal body length), or reduced nutritional state, suggesting a stable host-parasite relationship typical of Cryptosporidium species that are adapted to the host. Overall, Cryptosporidium occurred with a higher prevalence in possums from urban habitat (11.3\\%) than in possums from woodland habitat (5.6\\%); however, the host-specific nature of the genotypes may limit spillover infection in the urban setting. This study determined that the coexistence of possums with sympatric populations of humans, pets, and zoo animals in the urban Australian environment is unlikely to present a threat to public health safety.},\n\tlanguage = {eng},\n\tnumber = {17},\n\tjournal = {Appl Environ Microbiol},\n\tauthor = {Hill, N. J. and Deane, E. M. and Power, M. L.},\n\tmonth = sep,\n\tyear = {2008},\n\tnote = {Edition: 2008/07/22},\n\tkeywords = {Analysis of Variance, Animals, Animals, Wild/parasitology, Australia/epidemiology, Cloning, Molecular, Cryptosporidiosis/*epidemiology/transmission/veterinary, Cryptosporidium/*genetics, DNA, Protozoan/genetics, Disease Reservoirs/parasitology/veterinary, Feces/parasitology, Host-Parasite Interactions, Molecular Sequence Data, Oocysts/parasitology, Parasite Egg Count, Phylogeny, Polymerase Chain Reaction, Prevalence, RNA, Ribosomal, 18S/genetics, Sensitivity and Specificity, Sequence Alignment, Species Specificity, Statistics, Nonparametric, Trichosurus/*parasitology, Urbanization},\n\tpages = {5549--55},\n}\n\n
\n
\n\n\n
\n The common brushtail possum (Trichosurus vulpecula) is one of the most abundant native marsupials in urban Australia, having successfully adapted to utilize anthropogenic resources. The habituation of possums to food and shelter available in human settlements has facilitated interaction with people, pets, and zoo animals, increasing the potential for transmission of zoonotic Cryptosporidium pathogens. This study sought to examine the identity and prevalence of Cryptosporidium species occurring in possums adapted to urban settings compared to possums inhabiting remote woodlands far from urban areas and to characterize the health of the host in response to oocyst shedding. Findings indicated that both populations were shedding oocysts of the same genotype (brushtail possum 1 [BTP1]) that were genetically and morphologically distinct from zoonotic species and genotypes and most closely related to Cryptosporidium species from marsupials. The urban population was shedding an additional five Cryptosporidium isolates that were genetically distinct from BTP1 and formed a sister clade with Cryptosporidium parvum and Cryptosporidium hominis. Possums that were shedding oocysts showed no evidence of pathogenic changes, including elevated levels of white blood cells, diminished body condition (body mass divided by skeletal body length), or reduced nutritional state, suggesting a stable host-parasite relationship typical of Cryptosporidium species that are adapted to the host. Overall, Cryptosporidium occurred with a higher prevalence in possums from urban habitat (11.3%) than in possums from woodland habitat (5.6%); however, the host-specific nature of the genotypes may limit spillover infection in the urban setting. This study determined that the coexistence of possums with sympatric populations of humans, pets, and zoo animals in the urban Australian environment is unlikely to present a threat to public health safety.\n
\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Do free-ranging common brushtail possums (Trichosurus vulpecula) play a role in the transmission of Toxoplasma gondii within a zoo environment?.\n \n \n \n \n\n\n \n Hill, N. J.; Dubey, J. P.; Vogelnest, L.; Power, M. L.; and Deane, E. M.\n\n\n \n\n\n\n Vet Parasitol, 152(3-4): 202–9. April 2008.\n Edition: 2008/02/19\n\n\n\n
\n\n\n\n \n \n \"DoPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n  \n \n abstract \n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_free-ranging_2008,\n\ttitle = {Do free-ranging common brushtail possums ({Trichosurus} vulpecula) play a role in the transmission of {Toxoplasma} gondii within a zoo environment?},\n\tvolume = {152},\n\tcopyright = {All rights reserved},\n\tissn = {0304-4017 (Print) 0304-4017 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=18281157},\n\tdoi = {10.1016/j.vetpar.2008.01.002},\n\tabstract = {To investigate the possible role of common brushtail possums (Trichosurus vulpecula) in the transmission of Toxoplasma gondii within a zoo environment, a serological survey of a free-ranging population resident within Taronga Zoo, Sydney, Australia was undertaken using the modified agglutination test (MAT). For comparison, the seroprevalence of T. gondii antibodies was also assessed in a possum population inhabiting a felid-free, non-urban woodland habitat. Six of 126 possums (4.8\\%) from the zoo population had antibodies to T. gondii with a MAT titre of 25 or higher, while in contrast, all of the 17 possums from woodland were seronegative. These observations suggest that possums were at a higher risk of exposure to the parasite as a consequence of co-existing with domestic, stray and captive felids associated with urbanisation. Screening of captive felids at the zoo indicated 16 of 23 individuals (67\\%) and all 6 species were seropositive for T. gondii, implicating them as a possible source of the parasite within the zoo setting. In addition captive, non-felid carnivores including the chimpanzee (Pan troglodytes), saltwater crocodile (Crocodylus porosus), dingo (Canis lupis) and leopard seal (Hydrurga leptonyx) were tested for the presence of T. gondii antibodies as these species predate and are a leading cause of death amongst zoo possums. In total, 5 of 23 individuals (22\\%) were seropositive, representing 2 of the 4 carnivorous species; the dingo and chimpanzee. These data suggest that carnivory was not a highly efficient pathway for the transmission of T. gondii and the free-ranging possum population posed minimal threat to the health of zoo animals.},\n\tlanguage = {eng},\n\tnumber = {3-4},\n\tjournal = {Vet Parasitol},\n\tauthor = {Hill, N. J. and Dubey, J. P. and Vogelnest, L. and Power, M. L. and Deane, E. M.},\n\tmonth = apr,\n\tyear = {2008},\n\tnote = {Edition: 2008/02/19},\n\tkeywords = {Agglutination Tests/veterinary, Animals, Animals, Wild/parasitology, Animals, Zoo/parasitology, Antibodies, Protozoan/*blood, Australia/epidemiology, Disease Reservoirs/parasitology/veterinary, Female, Male, Seroepidemiologic Studies, Species Specificity, Toxoplasma/immunology, Toxoplasmosis, Animal/epidemiology/*transmission, Trichosurus/*parasitology},\n\tpages = {202--9},\n}\n\n
\n
\n\n\n
\n To investigate the possible role of common brushtail possums (Trichosurus vulpecula) in the transmission of Toxoplasma gondii within a zoo environment, a serological survey of a free-ranging population resident within Taronga Zoo, Sydney, Australia was undertaken using the modified agglutination test (MAT). For comparison, the seroprevalence of T. gondii antibodies was also assessed in a possum population inhabiting a felid-free, non-urban woodland habitat. Six of 126 possums (4.8%) from the zoo population had antibodies to T. gondii with a MAT titre of 25 or higher, while in contrast, all of the 17 possums from woodland were seronegative. These observations suggest that possums were at a higher risk of exposure to the parasite as a consequence of co-existing with domestic, stray and captive felids associated with urbanisation. Screening of captive felids at the zoo indicated 16 of 23 individuals (67%) and all 6 species were seropositive for T. gondii, implicating them as a possible source of the parasite within the zoo setting. In addition captive, non-felid carnivores including the chimpanzee (Pan troglodytes), saltwater crocodile (Crocodylus porosus), dingo (Canis lupis) and leopard seal (Hydrurga leptonyx) were tested for the presence of T. gondii antibodies as these species predate and are a leading cause of death amongst zoo possums. In total, 5 of 23 individuals (22%) were seropositive, representing 2 of the 4 carnivorous species; the dingo and chimpanzee. These data suggest that carnivory was not a highly efficient pathway for the transmission of T. gondii and the free-ranging possum population posed minimal threat to the health of zoo animals.\n
\n\n\n
\n\n\n\n\n\n
\n
\n\n
\n
\n  \n 2007\n \n \n (3)\n \n \n
\n
\n \n \n
\n \n\n \n \n \n \n \n Human–possum conflict in urban Sydney, Australia: public perceptions and implications for species management.\n \n \n \n\n\n \n Hill, N. J.; Carbery, K. A.; and Deane, E. M.\n\n\n \n\n\n\n Human Dimensions of Wildlife, 12(2): 101–113. 2007.\n \n\n\n\n
\n\n\n\n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{hill_humanpossum_2007,\n\ttitle = {Human–possum conflict in urban {Sydney}, {Australia}: public perceptions and implications for species management},\n\tvolume = {12},\n\tcopyright = {All rights reserved},\n\tissn = {1087-1209 1533-158X},\n\tdoi = {10.1080/10871200701195928},\n\tnumber = {2},\n\tjournal = {Human Dimensions of Wildlife},\n\tauthor = {Hill, Nichola J. and Carbery, Kelly A. and Deane, Elizabeth M.},\n\tyear = {2007},\n\tpages = {101--113},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n Isolation of the mite Myocoptes musculinus Koch from the Spinifex Hopping mouse (Notomys alexis).\n \n \n \n\n\n \n Old, J. M.; Hill, N.; and Deane, E. M.\n\n\n \n\n\n\n Laboratory Animals, 41: 292–295. 2007.\n \n\n\n\n
\n\n\n\n \n\n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n\n\n\n
\n
@article{old_isolation_2007,\n\ttitle = {Isolation of the mite {Myocoptes} musculinus {Koch} from the {Spinifex} {Hopping} mouse ({Notomys} alexis)},\n\tvolume = {41},\n\tcopyright = {All rights reserved},\n\tjournal = {Laboratory Animals},\n\tauthor = {Old, J. M. and Hill, N.J. and Deane, E. M.},\n\tyear = {2007},\n\tpages = {292--295},\n}\n\n
\n
\n\n\n\n
\n\n\n
\n \n\n \n \n \n \n \n \n Rodentolepis fraterna: the cause of mortality in a new host, the Spinifex hopping mouse (Notomys alexis).\n \n \n \n \n\n\n \n Hill, N. J.; Rose, K.; Deane, E. M.; and Old, J. M.\n\n\n \n\n\n\n Aust Vet J, 85(1-2): 62–4. January 2007.\n \n\n\n\n
\n\n\n\n \n \n \"RodentolepisPaper\n  \n \n\n \n \n doi\n  \n \n\n \n link\n  \n \n\n bibtex\n \n\n \n\n \n\n \n \n \n \n \n \n \n\n  \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n \n\n\n\n
\n
@article{hill_rodentolepis_2007,\n\ttitle = {Rodentolepis fraterna: the cause of mortality in a new host, the {Spinifex} hopping mouse ({Notomys} alexis)},\n\tvolume = {85},\n\tcopyright = {All rights reserved},\n\tissn = {0005-0423 (Print) 0005-0423 (Linking)},\n\turl = {http://www.ncbi.nlm.nih.gov/pubmed/17300462},\n\tdoi = {10.1111/j.1751-0813.2006.00095.x},\n\tnumber = {1-2},\n\tjournal = {Aust Vet J},\n\tauthor = {Hill, N. J. and Rose, K. and Deane, E. M. and Old, J. M.},\n\tmonth = jan,\n\tyear = {2007},\n\tkeywords = {*Murinae/parasitology, Animals, Animals, Laboratory, Hymenolepiasis/mortality/parasitology/*veterinary, Hymenolepis/isolation \\& purification, Immunohistochemistry/veterinary, Rodent Diseases/*mortality/parasitology},\n\tpages = {62--4},\n}\n\n
\n
\n\n\n\n
\n\n\n\n\n\n
\n
\n\n\n\n\n
\n\n\n \n\n \n \n \n \n\n
\n"}; document.write(bibbase_data.data);