Skip to main content
Emerging Infectious Diseases logoLink to Emerging Infectious Diseases
. 2018 Dec;24(12):2184–2194. doi: 10.3201/eid2412.180382

Terrestrial Bird Migration and West Nile Virus Circulation, United States

Daniele Swetnam 1,2,3,4,5,6, Steven G Widen 1,2,3,4,5,6, Thomas G Wood 1,2,3,4,5,6, Martin Reyna 1,2,3,4,5,6, Lauren Wilkerson 1,2,3,4,5,6, Mustapha Debboun 1,2,3,4,5,6, Dreda A Symonds 1,2,3,4,5,6, Daniel G Mead 1,2,3,4,5,6, Barry J Beaty 1,2,3,4,5,6, Hilda Guzman 1,2,3,4,5,6, Robert B Tesh 1,2,3,4,5,6, Alan DT Barrett 1,2,3,4,5,6,
PMCID: PMC6256381  PMID: 30457531

Abstract

Host migration and emerging pathogens are strongly associated, especially with regard to zoonotic diseases. West Nile virus (WNV), a mosquitoborne pathogen capable of causing severe, sometimes fatal, neuroinvasive disease in humans, is maintained in highly mobile avian hosts. Using phylogeographic approaches, we investigated the relationship between WNV circulation in the United States and the flight paths of terrestrial birds. We demonstrated southward migration of WNV in the eastern flyway and northward migration in the central flyway, which is consistent with the looped flight paths of many terrestrial birds. We also identified 3 optimal locations for targeted WNV surveillance campaigns in the United States—Illinois, New York, and Texas. These results illustrate the value of multidisciplinary approaches to surveillance of infectious diseases, especially zoonotic diseases.

Keywords: West Nile virus, bird migration, phylogeography, emerging pathogens, viruses, terrestrial birds, United States, vector-borne infections, zoonoses


West Nile virus (WNV) is a mosquitoborne virus that can cause severe and even fatal disease in humans. After WNV introduction into New York, NY, USA, its geographic range expanded quickly, reaching the West Coast in 2003. Previous studies have shown that the spread of WNV occurred faster than could be explained by contiguous diffusion (14) and demonstrated that its expansion occurred heterogeneously, consisting of contiguous diffusion and long distance translocations (2,5). Since then, phylogeographic studies have reported frequent mixing of WNV strains from local and distant locations. The most notable exception is California, where several genetic studies have shown limited movement into and out of the state (6,7).

The rapid expansion of WNV in the United States probably cannot be attributed to the movement of humans because humans are dead-end hosts. However, in nature, WNV is maintained in an enzootic transmission cycle involving mosquito vectors and highly mobile avian reservoirs. Hyalomma marginatum ticks have also been implicated in the transmission of WNV (8).

Although evidence of WNV infection has been identified in many species of birds, deaths and disease among birds vary greatly, ranging from asymptomatic to fatal infections; peak viremia potentially reaches >1012 PFU/mL (9). WNV RNA has been detected in bird spleen and kidneys as long as 36 weeks after infection (10) and in brains of Nestor notabilis kea up to 72 months after infection (11).

Although phylogenetic evidence of geographic clustering by location is limited, a recent study reported that WNV isolates clustered according to avian flyway (12). Because birds are the primary reservoirs for WNV, this finding was not surprising, but it is relevant because bird migration has also been implicated in the movement of influenza A virus (13), Borrelia burgdorferi (Lyme disease agent) (14), other pathogenic organisms (1), and even invasive invertebrate organisms (15). Several serologic studies (e.g., ELISA, plaque reduction neutralization test) have been used to determine the direction of WNV movement within the Atlantic, Mississippi, and Pacific flyways and demonstrated WNV in birds migrating southward, whereas evidence of the virus in birds during northward migration is limited (16,17).

Studies of virus movement associated with avian hosts in the United States have concentrated on the migration of waterfowl and excluded terrestrial birds, largely because the migratory patterns of waterfowl have been thoroughly characterized by banding studies. However, passerine birds, the primary reservoir for WNV, are terrestrial birds, not waterfowl. Terrestrial birds and waterfowl fly along similar but distinct flyways. Although waterfowl follow regular paths bounded by mountains and rivers, terrestrial birds often follow looped routes that enable them to maximize tail winds, avoid head winds, and correlate with seasonal fluctuations in food availability (18,19). Although looped migration paths have been described for several species of birds (Selasphorus rufus hummingbirds [20], Circus aeruginosus western marsh harriers [21], Falco eleonorae Eleonora’s falcons [22], Cuculus canorus common cuckoos [23]), the general flyways of terrestrial birds have been inadequately studied. However, in 2014, La Sorte et al. provided a general description of terrestrial bird flyways in North America (18). They defined 3 flyways: the single distinct Western flyway and 2 overlapping flyways, the Central and Eastern flyways. A similar 3-flyways system (Pacific, Central, and Atlantic flyways) has been described for waterfowl (24); however, most studies have relied on the more common 4-flyways system (Pacific, Central, Mississippi, and Atlantic flyways). In this study, we used phylogeographic approaches to investigate the relationship between WNV circulation in the United States and the flight paths of terrestrial birds.

Methods

Generation of Alignments

We identified all unique sequences of natural and laboratory WNV strains by using the Virus Variation Resource (25). Virus sequences meeting the following criteria were included in this study: 1) the nucleotide sequence spanned the complete open reading frame, 2) the sequence was derived from natural isolates and not laboratory strains, 3) the sequence was unique (i.e., all sequences differed by >1 nt), and 4) the sequence contained no degenerate nucleotides. All sequences were manually aligned in BioEdit version 7.1.3 (http://www.mbio.ncsu.edu/BioEdit/bioedit.html) or MEGA7 (https://www.megasoftware.net/), and noncoding regions were removed when necessary (i.e., the open reading frame was used for analyses).

Isolation of Viral RNA and Next-Generation Sequencing

We obtained additional WNV isolates from the World Reference Center for Emerging Viruses and Arboviruses at the University of Texas Medical Branch at Galveston (Galveston, TX, USA) (26). Isolates were originally collected from Virginia, Georgia, Texas, and Colorado. We extracted viral RNA from the supernatant of infected Vero cells by using a QIAamp Viral RNA Mini Kit (QIAGEN, Germantown, MD, USA) according to the manufacturer’s instructions. We generated libraries with a TruSeq RNA version 2 kit (Illumina, San Diego, CA, USA) and samples sequenced by the University of Texas Medical Branch at Galveston Next Generation Sequencing Core on an Illumina 1500 Hi-Seq platform. Adaptor sequences and poor quality reads (Q score <20) were removed with Trimmomatic (27). Reads were aligned with Bowtie2 (28) under the sensitive local parameter against the prototypical strain of WNV (NY99 flamingo 382–99, GenBank accession no. AF196835). Consensus sequences were generated by using SAMtools (29).

Phylogeny

To evaluate temporal structure, we generated a time-naive phylogeny (i.e., a maximum-likelihood phylogeny) to enable determination of the patristic distance between all isolates on the phylogeny. We generated maximum-likelihood trees with RAxML-HPC Black Box on Cyberinfrastructure for Phylogenetic Research version 3.3 (30) and determined automatic halting by bootstrapping. We determined the root-to-tip distance, which is a phylogenetic measure of genetic distance, for each isolate of the maximum-likelihood phylogenies by using TempEst (formerly Path-o-gen) (31). We evaluated the correlation (Pearson method) between root-to-tip distance and collection date in R (https://www.r-project.org/).

We used a Bayesian Markov chain Monte Carlo (MCMC) approach to infer phylogeographic relationships and selected the most appropriate phylogenetic model by using standard path sampling and stepping-stone approaches. XML files were generated in BEAUti version 1.8.3 or 1.8.4 and run with BEAST version 1.8.4 (32) on Cyberinfrastructure for Phylogenetic Research (30). We used the GTR+Γ+I (general time reversible with gamma rate distribution and invariable sites) model to infer nucleotide substitution frequencies, an uncorrelated lognormal clock model to infer the mutation rate, and a Bayesian Skyline tree prior to model changes in population size. The evolution rate mean was restricted to 10−4 through 9 × 10−4 substitutions/site/year, consistent with previously reported rates for WNV evolution (7,33).

We ran trees with an MCMC length of 100 million and sampled every 5,000 steps. Log files were reviewed in Tracer (http://tree.bio.ed.ac.uk/software/tracer/) to determine burn-in, which ranged from 5% through 10%. We ran multiple independent MCMC chains until effective sample size values exceeded 200. Log and tree files were combined in LogCombiner version 1.8, and a maximum clade credibility tree was generated in TreeAnnotator (32). Locations were inferred by using ancestral state reconstruction with an asymmetric discrete trait substitution model (34).

Analysis of Migration

After the XML files were generated in BEAUti, we manually edited them to enable counting of all Markov jumps (MJ) (which described the relative magnitude of migration out of source locations and into sink locations) for 2001 through 2009 (35). This method for evaluating migration, first described by Minin and Suchard (35), has been used to characterize migration of several major pathogens including rabies virus (36), dengue virus (37), HIV (38), influenza virus (39,40), and Rift Valley fever virus (41).

As expected for an emerging zoonotic disease, the annual West Nile neurologic disease (WNND) incidence and sample collection efforts varied dramatically among states over time, adding substantial complexity to the model. To mitigate the effects of inconsistent sampling and to confirm the observed results, we applied a stricter inclusion criterion to confirm the results obtained by using the full dataset. The sequences were randomly down-sampled such that the number of sequences used correlated (p<0.05 by Pearson method) with the incidence of WNND reported to the Centers for Disease Control and Prevention (CDC) in a particular year (the most accurate record of relative WNV activity). Our analysis ensured that the dataset was representative of the WNV activity of each region in a particular year.

We calculated incidence by using the number of WNND cases reported to CDC from each state during each year and dividing that number by the estimated population of each state. The population estimates were obtained from the Time Series of Intercensal State Population Estimates available at the Population Division of the US Census Bureau (42). States with insufficient sequences to represent the WNND incidence were excluded. Down-sampling was undertaken in at least duplicate to ensure that reduction in sample size and diversity did not remove important relationships.

Results

Sequence Collection

All previously published sequences of natural WNV isolates collected in the United States were obtained from GenBank on January 1, 2016. The number of WNV sequences varied substantially over time and among locations, which presented statistical challenges. In particular, although GenBank has >900 WNV open reading frames, most come from a few states where laboratories were actively involved in WNV surveillance and research: California, New York, and Texas. The ability to compare multiple isolates over multiple years was critical to the analysis. Only a few states had sufficient numbers of WNV sequences available in GenBank to enable analysis for multiple consecutive years: New York, Connecticut, Illinois, North Dakota, South Dakota, Texas, and California.

To mitigate the influence of sampling bias, we obtained additional WNV isolates from the World Reference Center for Emerging Viruses and Arboviruses for 3 states and sequences to support the analysis: Virginia (n = 39), Georgia (n = 20), and Colorado (n = 31) (Table 1). Given that previous studies have demonstrated limited WNV movement into or out of California (6,7), we did not include isolates from California in the analysis. Similarly, because of the proximity of New York and Connecticut, we chose New York to represent WNV in the Northeast because Connecticut is a small state. Last, to ensure that each location was represented across a similar time frame, we included only isolates collected during 2001–2009 in the Bayesian phylogeny and migration analysis. Table 2 shows the states and availability of yearly isolates.

Table 1. Summary of isolates sequenced in study of terrestrial bird migration and West Nile virus circulation, United States.

Isolate Phylogeny code GenBank accession no. State Year
Laco_3008 CO03C MH170226 CO 2003
AIDL-M-015 CO03D MH170228 CO 2003
LACO-3041 CO03E MH170231 CO 2003
Laco_3038 CO03F MH170234 CO 2003
AIDL-M-012 CO03G MH170237 CO 2003
Laco_3020 CO03H MH170238 CO 2003
Laco_3022 CO03I MH170254 CO 2003
laco_3030 CO03J MH170256 CO 2003
CO1862 CO04E MH170227 CO 2004
CO_2572 CO04F MH170246 CO 2004
DB_4218 CO04G MH170248 CO 2004
DB_4217 CO04H MH170233 CO 2004
CO_06–7390 CO06A MH170229 CO 2006
CO_06–608 CO06B MH170232 CO 2006
CO_06–10725 CO06C MH170235 CO 2006
CO_07–11032 CO06D MH170236 CO 2006
CO_06–10723 CO06E MH170239 CO 2006
CO_06–584 CO06F MH170243 CO 2006
CO_06–10716 CO06G MH170249 CO 2006
CO_07–8779 CO07C MH170230 CO 2007
CO_07–10970 CO07D MH170241 CO 2007
GT_02566 CO07E MH170242 CO 2007
CO_07–8778 CO07F MH170244 CO 2007
CO_07–11027 CO07G MH170251 CO 2007
CO_07–9340 CO07H MH170252 CO 2007
CO_08–13382 CO08A MH170240 CO 2008
CO_08–13386 CO08B MH170245 CO 2008
CO-13363 CO08C MH170247 CO 2008
CO_08–13401 CO08D MH170250 CO 2008
CO_08–13787 CO08E MH170253 CO 2008
CO_08–13410 CO08F MH170255 CO 2008
DES_566–01 GA01C MH170263 GA 2001
DES_107–01 GA01D MH170273 GA 2001
DES_1476–01 GA01E MH170276 GA 2001
DES_1191–02 GA02C MH170274 GA 2002
DES_160–02 GA02D MH170275 GA 2002
DES_1201–02 GA02E MH170264 GA 2002
GA_04–230 GA04A MH170270 GA 2004
GA_Chc_04–1485 GA04B MH170265 GA 2004
GA_05–179 GA05A MH170269 GA 2005
GA_lwn_50_4936 GA05B MH170257 GA 2005
M07–069 GA07A MH170272 GA 2007
M07–087 GA07B MH170258 GA 2007
M07–086 GA07C MH170266 GA 2007
DES_07–53 GA07D MH170267 GA 2007
DES_07–62 GA07E MH170268 GA 2007
DKB_08–0403 GA08A MH170261 GA 2008
DBK_08–0491 GA08B MH170271 GA 2008
FNT_09–199 GA09A MH170259 GA 2009
Lwn_09–846 GA09B MH170262 GA 2009
FNT_09–144 GA09C MH170260 GA 2009
VA_AV_321–00 VA00A MH166882 VA 2000
VA_AV_593 VA00B MH166904 VA 2000
VA_AV_380 VA00C MH166903 VA 2000
VA_AV_573–00 VA00D MH166901 VA 2000
VA_TC_2535–01 VA01A MH166886 VA 2001
VA_B_037–02 VA02A MH166883 VA 2002
VA_BD_37 VA02B MH166905 VA 2002
VA_TC_1500 VA02C MH166887 VA 2002
VA_TC_1500–02 VA02D MH166899 VA 2002
VA_TC_2076 VA02E MH166911 VA 2002
VA_TC_2147 VA02F MH166915 VA 2002
VA_TC_2790–03 VA03C MH166906 VA 2003
VA_TC_4043 VA03D MH166900 VA 2003
VA_TC_3278 VA03E MH166912 VA 2003
VA_1909–04 VA04A MH166884 VA 2004
VA_TC_1597 VA04B MH166919 VA 2004
VA_TC_1155 VA04C MH166913 VA 2004
VA_TC_1272 VA04D MH166917 VA 2004
VA_P_3321–05 VA05A MH166888 VA 2005
VA_P_4209 VA05B MH166895 VA 2005
VA_SN_3082–05 VA05C MH166908 VA 2005
VA_P_4485–06 VA06A MH166889 VA 2006
VA_P_4770–06 VA06B MH166907 VA 2006
VA_SP_5645–06 VA06C MH166890 VA 2006
VA_TC_4177 VA06D MH166910 VA 2006
VA_1660 VA07A MH166891 VA 2007
VA_2327 VA07B MH166894 VA 2007
VA_TC_1368–08 VA08A MH166898 VA 2008
VA_SP_1202–08 VA08B MH166892 VA 2008
VA_TC_2045–08 VA08C MH166921 VA 2008
VA_TC_1732–08 VA08D MH166918 VA 2008
VA_3920 VA09A MH166885 VA 2009
VA_SN_3222–09 VA09B MH166909 VA 2009
VA_SN_5859–09 VA09C MH166896 VA 2009
VA_TC_1732–09 VA09D MH166920 VA 2009
VA_SN_4826–09 VA09E MH166914 VA 2009
VA_2191 VA10A MH166902 VA 2010
VA_TC_1117–10 VA10B MH166897 VA 2010
VA_TC_2020–10 VA10C MH166893 VA 2010
VA_TC_1184–10 VA10D MH166916 VA 2010

Table 2. Years in which West Nile virus sequences were available in study of terrestrial bird migration and West Nile virus circulation, United States*.

Location
1999
2000
2001
2002
2003
2004
2005
2006
2007
2008
2009
2010
2011
2012
2013
2014
NY x x x x x x x x x
VA x x x x x x x x x x x
GA x x x x x x x
IL x x x x x x
TX x x x x x x x x x x x
CO x x x x x x x
ND x x x x x x x
SD x x x x x x x

*x, available; blank cells, not available.

Model Selection

We compared 203 nucleotide substitution models by using the Bayesian and Akaike Information Criteria in JModelTest2 (https://github/com/ddariba/jmodeltest2) and found the GTR+Γ+I model to be the most appropriate. For assessing temporal signature, we used a maximum-likelihood tree with sequences of WNV strains from New York, Virginia, Georgia, Illinois, North Dakota, South Dakota, Texas, and Colorado (n = 379) (Figure 1). We identified a statistically significant positive correlation (r = 0.93, 95% highest posterior density [HPD] = 0.92–0.94; p<2.2 × 10−16) between the root-to-tip distance and the date of isolation in Temp-Est (formerly known as Path-O-gen) (Figure 2). The mutation rate was estimated to be 4.05 × 10−4 substitutions/site/year, and the most recent common ancestor (MRCA) was in 1997. Together these results indicated a strong temporal signal in the dataset. Finally, we evaluated Bayesian tree priors (skyride, skygrid, and skyline) and uncorrelated clock models (lognormal and exponential) by using path-sampling and stepping-stone approaches. The uncorrelated lognormal clock model with the Bayesian skyline tree prior was the most appropriate.

Figure 1.

Figure 1

Maximum-likelihood phylogeny generated with all West Nile virus sequences from New York, Virginia, Georgia, Illinois, North Dakota, South Dakota, Texas, and Colorado (n = 379) in study of terrestrial bird migration and West Nile virus circulation, United States. Sequence names include the 2-letter state abbreviation to indicate the origin of isolation, followed by the year. Multiple isolates collected from the same state within the same year are differentiated by letter. GenBank accession numbers are provided for all taxa that were not sequenced in this study. Scale bar indicates nucleotide substitutions per site.

Figure 2.

Figure 2

Analysis of correlation between virus isolation date and genetic diversity in study of terrestrial bird migration and West Nile virus circulation, United States. Root-to-tip distances of all sequences were determined for each isolate by using the maximum-likelihood tree shown in Figure 1 (https://wwwnc.cdc.gov/EID/article/24/12/18-0382-F1.htm) and plotted against the year. Dots are colored by location of isolation. The correlation between the root-to-tip distance and year of isolation was determined with linear regression shown in blue. 95% CIs are shown in gray. The equation of the linear regression line was used to estimate the year of the most recent common ancestor (MRCA) and the mutation rate (m): y = mx + MRCA.

Phylogeographic Analysis for the United States

Analysis of all WNV sequences collected from New York, Virginia, Georgia, Illinois, North Dakota, South Dakota, Texas, and Colorado during 2001–2009 provided estimates of the introduction date of the MRCA and mean evolution rate that were consistent with the estimates of the root-to-tip distance analysis (Table 3; Figure 3). The date of MRCA introduction was estimated as 1997, and the average evolution rate was 3.92 × 10−4 substitutions/site/year.

Table 3. Statistical support for phylogeny in study of terrestrial bird migration and West Nile virus circulation, United States*.

Variable Mean ESS 95% HPD interval
Posterior −49722.50 1370 −49803.21 to −49640.55
Prior −3987.09 1110 −4051.92 to −3916.56
Likelihood −45735.41 2179 −45780.00 to −45691.43
MRCA 11.92 3119 10.82 to 13.08
UCLD.mean 3.92 × 10−4 1604 3.55 × 10−4 to 4.49 × 10−4

*ESS, effective sample size; HPD, highest posterior density; MRCA, most recent common ancestor (years before 2009); UCLD.mean, evolution rate inferred with an uncorrelated clock model with lognormal distribution.

Figure 3.

Figure 3

Bayesian phylogeny of West Nile virus isolates collected in representative regions along the Eastern and Central flyways between 2001 and 2009, United States. Maximum-clade credibility tree was obtained by using a Bayesian approach. The location of each isolate and the inferred location of each ancestor are depicted by color.

We used MJ between reconstructed ancestral states to estimate the magnitude of relative migration out of, or into, each of the 8 regions (Table 4, 5). Frequent migration (>2 MJ) was detected from Illinois to Colorado (8.38 MJ), Illinois to Georgia (8.23 MJ), Illinois to North Dakota (10.43 MJ), Illinois to New York (29.97 MJ), Illinois to South Dakota (6.69 MJ), Illinois to Texas (22.87 MJ), Illinois to Virginia (11.45 MJ), New York to Colorado (4.36 MJ), New York to Georgia (7.04 MJ), New York to South Dakota (2.18 MJ), New York to Texas (4.56 MJ), New York to Virginia (4.24 MJ), Texas to Colorado (9.78 MJ), Texas to North Dakota (5.18 MJ), Texas to South Dakota (7.56 MJ), and Virginia to Georgia (3.62 MJ).

Table 4. Source and sink analysis in study of terrestrial bird migration and West Nile virus circulation, United States*.

Source
Sink
Markov jumps, mean
ESS
95% HPD interval
CO ND 1.368 8603 0–3
GA IL 1.31 31284 1–3
IL CO 8.376 3311 1–14
IL GA 8.226 3046 3–13
IL ND 10.43 15423 6–14
IL NY 29.965 1355 20–40
IL SD 6.691 3635 1–11
IL TX 22.872 4545 14–30
IL VA 11.449 1631 6–16
NY CO 4.362 2143 0–8
NY GA 7.039 2170 2–11
NY IL 1.012 3023 0–4
NY SD 2.177 2740 0–5
NY TX 4.564 2149 0–9
NY VA 4.24 1171 0–8
SD CO 1.328 3964 0–5
TX CO 9.775 6039 3–16
TX ND 5.177 7617 2–9
TX NY 1.606 9410 0–4
TX SD 7.557 5113 2–12
VA CO 1.16 9807 0–3
VA GA 3.616 14054 2–6

*Mean number of Markov jjumps detected between each source (origin) and sink (destination) location indicates the minimum number of migration events observed from each source to each sink. Only Markov jumps >2 are shown. A summary of all Markov jumps is shown in Table 5. *ESS, effective sample size; HPD, highest posterior density.

Table 5. Markov jump analysis results from study of terrestrial bird migration and West Nile virus circulation, United States*.

Source
Sink
Markov jumps
ESS
95% HPD interval
CO GA 0.164 15683 0–1
CO IL 0.111 33920 0–1
CO ND 1.368 8603 0–3
CO NY 0.261 16508 0–1
CO SD 0.486 8222 0–2
CO TX 0.337 5909 0–2
CO VA 0.165 24157 0–1
GA CO 0.314 13738 0–2
GA IL 1.31 31284 1–3
GA ND 0.122 26193 0–1
GA NY 0.821 6736 0–3
GA SD 0.719 8787 0–2
GA TX 0.313 11570 0–2
GA VA 0.334 10761 0–2
IL CO 8.376 3311 1–14
IL GA 8.226 3046 3–13
IL ND 10.43 15423 6–14
IL NY 29.965 1355 20–40
IL SD 6.691 3635 1–11
IL TX 22.872 4545 14–30
IL VA 11.449 1631 6–16
ND CO 0.618 8832 0–3
ND GA 0.147 23908 0–1
ND IL 0.144 25515 0–1
ND NY 0.288 12855 0–2
ND SD 0.183 20239 0–1
ND TX 0.521 9813 0–2
ND VA 0.255 15630 0–1
NY CO 4.362 2143 0–8
NY GA 7.039 2170 2–11
NY IL 1.012 3023 0–4
NY ND 0.448 7434 0–2
NY SD 2.177 2740 0–5
NY TX 4.564 2149 0–9
NY VA 4.24 1171 0–8
SD CO 1.328 3964 0–5
SD GA 0.567 10905 0–2
SD IL 0.143 26368 0–1
SD ND 0.301 8901 0–2
SD NY 0.236 19732 0–1
SD TX 0.93 3517 0–4
SD VA 0.116 32792 0–1
TX CO 9.775 6039 3–16
TX GA 0.669 6777 0–3
TX IL 0.441 10249 0–2
TX ND 5.177 7617 2–9
TX NY 1.606 9410 0–4
TX SD 7.557 5113 2–12
TX VA 0.296 13706 0–2
VA CO 1.16 9807 0–3
VA GA 3.616 14054 2–6
VA IL 0.261 14878 0–1
VA ND 0.194 20523 0–1
VA NY 0.96 4048 0–3
VA SD 0.176 23197 0–1
VA TX 0.247 15068 0–1

*ESS, effective sample size; HPD, highest posterior density.

Overall, 3 major sources of WNV circulation (New York, Illinois, and Texas) seemed to be the origin of 88.5% of the total migration events observed (Table 4; Figure 4). Southward and westward movements were detected along the East Coast, but only northward movement was observed within the central United States. A notable exception was observed in Illinois, where evidence of WNV movement in all directions was demonstrated.

Figure 4.

Figure 4

Summary of source/sink analysis in study of terrestrial bird migration and West Nile virus circulation, United States. Minimum number of migration events detected from A) the Eastern flyway, B) Illinois, and C) the Central flyway. Only events that occurred at least twice are depicted. Red arrows, northward migration; black arrows, southward migration; green arrow, lateral migration; dotted arrows, migration that could not be confirmed by incident-controlled downsampling because of an insufficient number of sequences.

Incidence-Controlled Phylogeny

To mitigate the effects of inconsistent sampling, we applied a stricter inclusion criterion to ensure that the dataset was representative of WNV activity in each region in a particular year. In this approach, the sequences were randomly down-sampled by using the sample command in R, such that the number of sequences was proportional to the incidence of WNND reported to CDC (Table 6). Illinois, North Dakota, and South Dakota were not included in the down-sampled datasets because there were insufficient sequences to represent WNND incidence in these states. To ensure that reduction in sample size and diversity did not remove important relationships, the down-sampling was independently performed twice.

Table 6. Incidence-controlled down-sampling strategy used in study of of terrestrial bird migration and WNV circulation, United States*.

Location, year
WNV incidence
Sequences available
Sequences used
GA
2001 7.16 × 10−7 5 3
2002 3.29 × 10−6 11 6
2003 3.13 × 10−6 1 1
2004 1.60 × 10−6 2 2
2005 1.01 × 10−6 1 1
2006 2.18 × 10−7 0 0
2007 2.46 × 10−6 5 5
2008 4.21 × 10−7 3 3
2009
4.16 × 10−7
3
3
NY
2001 6.81 × 10−7 10 3
2002 3.55 × 10−6 10 6
2003 2.97 × 10−6 15 5
2004 3.65 × 10−7 10 3
2005 1.57 × 10−6 7 4
2006 8.37 × 10−7 5 3
2007 8.36 × 10−7 9 3
2008 1.67 × 10−6 18 4
2009
3.11 × 10−7
2
2
VA
2001 0 6 2
2002 2.20 × 10−6 10 5
2003 2.58× 10−6 6 5
2004 6.69 × 10−7 4 3
2005 0 3 2
2006 0 4 2
2007 3.87 × 10−7 2 2
2008 0 4 2
2009 6.31 × 10−7 5 3

*The numbers of West Nile virus sequences available and of sequences used in the down-sampled dataset are summarized. WNV, West Nile virus.

According to the 2 incidence-controlled datasets, the MRCA was ≈1997 in both down-sampling exercises (95% HPD 1996.00–1998.52 and 95% HPD 1995.7–1998.25), and the overall mutation rates were estimated to be 4.02 × 10−4 and 3.83 × 10−4 substitutions/site/year (Table 7; Figure 5). As with the full dataset, the Markov analysis demonstrated that New York and Texas were strong sources of WNV circulation. Significant movement (mean >2 MJ) was detected from Texas to Colorado (20.42 and 20.44 MJ); Texas to New York (12.36 and 11.77 MJ); Texas to Georgia (8.28 and 9.55 MJ); Texas to Virginia (7.14 and 7.732 MJ); New York to Georgia (6.1 and 5.38 MJ); New York to Virginia (4.95 and 3.65 MJ); New York to Colorado (4.04 and 2.66 MJ); New York to Texas (2.66 and 2.73 MJ); Virginia to Georgia (1.55 and 3.62 MJ); and, in dataset 2 only, Virginia to Colorado (1.31 MJ) (Figure 6).

Table 7. Statistical support for the incidence-controlled phylogenies determined in study of terrestrial bird migration and West Nile virus circulation, United States*.

Variable
Dataset 1
Dataset 2
Mean
ESS
95% HPD interval
Mean
ESS
95% HPD interval
Posterior −36338.60 1385 −36394.36 to −36280.37 −36798.90 1803 −36853.19 to −36740.97
Prior −3141.96 1187 −3189.54 to −3091.38 −3153.08 1387 −3201.84 to −3104.50
Likelihood −33196.64 1367 −33229.23 to −33165.79 −33645.82 2115 −33678.80 to −33614.39
MRCA 11.66 4720 10.47 to 12.10 11.97 6985 10.75 to 13.30
UCLD.mean 4.02 × 10−4 2336 3.53 × 10−4 to 4.55 × 10−4 3.83 × 10−4 1749 3.31 × 10−4 to 4.39 × 10−4

*ESS, effective sample size; HPD, highest posterior density; MRCA, most recent common ancestor (years before 2009); UCLD.mean, evolution rate inferred with an uncorrelated clock model with lognormal distribution.

Figure 5.

Figure 5

Incidence-controlled phylogeny of Eastern and Central flyways, United States. Sequences were down-sampled such that the number of sequences was proportional to the annual incidence of West Nile neurologic disease incidence for each location between 2001 and 2009. Down-sampling was undertaken twice (A and B) to ensure that the reduction in sequences did not result in a substantial loss of diversity. Illinois, North Dakota, and South Dakota were not included in the incidence-control analysis because too few sequences were available to support down-sampling. Bayesian approaches were used to generate maximum-clade credibility trees. Scale bars indicate nucleotide substitutions per site.

Figure 6.

Figure 6

Summary of Markov jump analysis performed on the incident-controlled phylogeny. A, B) The results of the Markov jump analysis for each down-sampled dataset are summarized as box plots. Box tops indicate third quartiles, box bottoms indicate first quartiles; horizontal bars within boxes indicate medians; error bars indicate maximums and minimums. Red, northward movement; teal, southward movement; purple, movement that is neither north nor south; dotted arrows, movement that was not observed in the incident-controlled down-sampling because of an insufficient number of sequences. C, D) Movement originating in the eastern and central United States. Only Markov jumps that occurred >2 times are depicted.

Together, the MJ analyses of the incidence-controlled dataset and the full dataset illustrate a consistent pattern of WNV circulation. All southward movement originated in the eastern United States (New York and Virginia), and most of the northward movement originated in the central United States (Texas) (Figure 7).

Figure 7.

Figure 7

Model summarizing the general patterns of West Nile virus movement in the United States. Red, northward movement; teal, southward movement; dotted arrows, relationships that could not be confirmed in incident-controlled datasets because of an insufficient number of sequences.

Discussion

In recent years, emerging zoonotic diseases caused by Ebola, Zika, Nipah, Middle Eastern respiratory syndrome, and influenza A viruses have become major public health problems, devastating communities and costing millions for public health interventions. Decisive, evidence-based approaches are critical for managing emerging infectious diseases, but effective and efficient countermeasures will be possible only after the relationships between pathogens and their hosts have been thoroughly characterized.

Bird migration has been implicated in the movement of a variety of pathogens (1). In particular, characterization of the relationship between avian influenza virus movement and waterfowl migration has supported surveillance and early warning programs (1,43). However, studies of avian hosts in the Americas have mainly concentrated on the migration of waterfowl to the exclusion of terrestrial birds because waterfowl are easily tracked with banding; thus, their migration has been thoroughly characterized.

The introduction and subsequent spread of WNV into the Americas underscores the invasive potential of emerging pathogens in the New World, as has been recently exemplified by Zika virus, another mosquitoborne flavivirus. Dramatic variations in the location, timing, and intensity of WNV strain collection and sequencing has left the field with a limited understanding of virus circulation patterns and no reliable way of predicting the geographic spread of WNV outbreaks. We have addressed this knowledge gap by characterizing the movement of WNV with regard to the migratory patterns of its natural hosts, terrestrial birds. We compiled 379 virus sequences for analysis, including 289 previously reported sequences from New York, Virginia, Georgia, Illinois, North Dakota, South Dakota, Texas, and Colorado, plus 90 novel sequences from Virginia, Georgia, and Colorado.

Phylogeographic analysis revealed that 3 locations— New York, Illinois, and Texas—accounted for 88.5% of the total WNV MJ inferred. Because New York is the presumed original introduction point for WNV into the United States, its role as a major source of WNV movement was expected. However, 74.2% of the observed MJ originated in Illinois and Texas only. Of note, North Dakota and South Dakota, which are 2 of the states with the highest annual WNND incidence, seem to be strong sinks for WNV moving out of Illinois and Texas.

The contributions of Illinois and Texas to WNV circulation are not surprising because both locations are situated at major convergence points between the Eastern and Central flyways. In the case of Texas, birds from both flyways may avoid long-distance flights across the Gulf of Mexico by traveling along the circa-Gulf route that follows the Gulf Coast through Texas into Mexico. In the case of Illinois, seasonal shifts in terrestrial bird migration routes ensure that Illinois supports birds from the Eastern and Central flyways during annual migrations.

Of note, although mosquito and WNV activity occurs earlier in the southern than in the northern United States, southward migration was detected along the East Coast during our sampling period, 2002–2009, indicating that the southeastern United States is probably a dead end for WNV circulation. Indeed, low-level transmission probably occurs during the winter in warmer locations such as Florida, Texas, and Louisiana. This possibility is supported by isolations of WNV from mosquitoes and birds during December and January in Harris County, Texas (44), and suggests that ecologic factors not related to mosquito abundance and WNV activity drive WNV movement along the East Coast. Instead, movement of WNV into the northeastern United States (New York) from Illinois and Texas was observed (in the incident-controlled analysis). These results suggest that introduction of WNV into the northeastern United States originated from the central United States.

Overall, we have defined the pattern of WNV circulation in the United States (Figure 7) and demonstrated looped virus movement patterns in the Eastern and Central flyways that are bridged by Illinois, a region shared between the 2 flyways. This specific pattern correlates with the looped migration patterns of terrestrial birds. Although other geographic regions may contribute to virus movement, there were insufficient virus sequences available from other states to incorporate into this analysis. Thus, on the basis of available information, 3 of the 8 locations considered (New York, Illinois, and Texas) seem to be the preferred sites for efficiently monitoring ongoing WNV evolution.

As new WNV sequences become available, similar phylogeographic methods can be used to develop more detailed information about WNV circulation in the United States. For example, on the East Coast, WNV circulation occurs southward, so surveillance efforts in the Northeast are likely to be more informative than surveillance in the Southeast. Conversely, WNV in the central United States travels northward, so surveillance in the south-central United States is more likely than surveillance in the north-central United States to be informative. Last, the region of overlap between the Eastern and Central flyways is the most likely location for deriving surveillance information because WNV in this area travels in multiple directions.

Collectively, the results of this study illustrate the value of using multidisciplinary approaches to surveillance of infectious diseases, especially zoonotic diseases. Animal migration is shaped by a delicate balance of ecologic factors and anthropomorphic barriers. Natural and manmade events (e.g., climate change, atmospheric fluctuations, habitat destruction) can drastically alter host behavior, which in turn affects the circulation patterns of infectious agents such as WNV. In this study, we defined the patterns of WNV circulation and key areas for surveillance and correlated them with the migratory patterns of their primary reservoir, terrestrial birds. Although this information does not enable prediction of the size of annual WNV outbreaks, these advancements support the construction of targeted surveillance and vector mitigation strategies to predict the annual flow of WNV strains and to enable public health officials to anticipate changes in WNV circulation resulting from altered bird migration.

Acknowledgments

This study was supported in part by National Institutes of Health contracts N01-AI 25489, N01-AI30027, and HHSN272201000040I/HHSN27200004/D04 to R.B.T. and grant AI 067847 to A.D.T.B.

Biography

Dr. Swetnam is a postdoctoral fellow at the University of California at Davis. Her research interests include the evolution and molecular epidemiology of zoonotic and emerging diseases.

Footnotes

Suggested citation for this article: Swetnam D, Widen SG, Wood TG, Reyna M, Wilkerson L, Debboun M, et al. Terrestrial bird migration and West Nile virus circulation, United States. Emerg Infect Dis. 2018 Dec [date cited]. https://doi.org/10.3201/eid2412.180382

References

  • 1.Reed KD, Meece JK, Henkel JS, Shukla SK. Birds, migration and emerging zoonoses: west nile virus, lyme disease, influenza A and enteropathogens. Clin Med Res. 2003;1:5–12. 10.3121/cmr.1.1.5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Pybus OG, Suchard MA, Lemey P, Bernardin FJ, Rambaut A, Crawford FW, et al. Unifying the spatial epidemiology and molecular evolution of emerging epidemics. Proc Natl Acad Sci U S A. 2012;109:15066–71. 10.1073/pnas.1206598109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Lewis M, Rencławowicz J, van den Driessche P. Traveling waves and spread rates for a West Nile virus model. Bull Math Biol. 2006;68:3–23. 10.1007/s11538-005-9018-z [DOI] [PubMed] [Google Scholar]
  • 4.Maidana NA, Yang HM. Spatial spreading of West Nile Virus described by traveling waves. J Theor Biol. 2009;258:403–17. 10.1016/j.jtbi.2008.12.032 [DOI] [PubMed] [Google Scholar]
  • 5.Duggal NK, Reisen WK, Fang Y, Newman RM, Yang X, Ebel GD, et al. Genotype-specific variation in West Nile virus dispersal in California. Virology. 2015;485:79–85. 10.1016/j.virol.2015.07.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Duggal NK, Bosco-Lauth A, Bowen RA, Wheeler SS, Reisen WK, Felix TA, et al. Evidence for co-evolution of West Nile Virus and house sparrows in North America. PLoS Negl Trop Dis. 2014;8:e3262. 10.1371/journal.pntd.0003262 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Añez G, Grinev A, Chancey C, Ball C, Akolkar N, Land KJ, et al. Evolutionary dynamics of West Nile virus in the United States, 1999–2011: phylogeny, selection pressure and evolutionary time-scale analysis [cited 2013 Dec 13]. http://dx.plos.org/10.1371/journal.pntd.0002245 [DOI] [PMC free article] [PubMed]
  • 8.Formosinho P, Santos-Silva MM. Experimental infection of Hyalomma marginatum ticks with West Nile virus. Acta Virol. 2006;50:175–80. [PubMed] [Google Scholar]
  • 9.Komar N, Langevin S, Hinten S, Nemeth N, Edwards E, Hettler D, et al. Experimental infection of North American birds with the New York 1999 strain of West Nile virus. Emerg Infect Dis. 2003;9:311–22.file:///\\\\\\\\cdc.gov\\\\project\\\\CCID_NCPDCID_DEISS_EIDJ\\\\EID%20Production\\\\Editorial\\\\EDITING\\\\Lynne\\\\2018\\\\Dec18\\\\18-0382\\\\%20PubMed 10.3201/eid0903.020628 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Wheeler SS, Langevin SA, Brault AC, Woods L, Carroll BD, Reisen WK. Detection of persistent west nile virus RNA in experimentally and naturally infected avian hosts. Am J Trop Med Hyg. 2012;87:559–64. 10.4269/ajtmh.2012.11-0654 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Bakonyi T, Gajdon GK, Schwing R, Vogl W, Häbich AC, Thaller D, et al. Chronic West Nile virus infection in kea (Nestor notabilis). Vet Microbiol. 2016;183:135–9. 10.1016/j.vetmic.2015.12.012 [DOI] [PubMed] [Google Scholar]
  • 12.Di Giallonardo F, Geoghegan JL, Docherty DE, McLean RG, Zody MC, Qu J, et al. Fluid spatial dynamics of West Nile virus in the United States: rapid spread in a permissive host environment. J Virol. 2015;90:862–72. 10.1128/JVI.02305-15 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Bahl J, Krauss S, Kühnert D, Fourment M, Raven G, Pryor SP, et al. Influenza a virus migration and persistence in North American wild birds. PLoS Pathog. 2013;9:e1003570. 10.1371/journal.ppat.1003570 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Vollmer SA, Bormane A, Dinnis RE, Seelig F, Dobson ADM, Aanensen DM, et al. Host migration impacts on the phylogeography of Lyme Borreliosis spirochaete species in Europe. Environ Microbiol. 2011;13:184–92. 10.1111/j.1462-2920.2010.02319.x [DOI] [PubMed] [Google Scholar]
  • 15.Muñoz J, Amat F, Green AJ, Figuerola J, Gómez A. Bird migratory flyways influence the phylogeography of the invasive brine shrimp Artemia franciscana in its native American range. PeerJ. 2013;1:e200. 10.7717/peerj.200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Dusek RJ, McLean RG, Kramer LD, Ubico SR, Dupuis AP II, Ebel GD, et al. Prevalence of West Nile virus in migratory birds during spring and fall migration. Am J Trop Med Hyg. 2009;81:1151–8. 10.4269/ajtmh.2009.09-0106 [DOI] [PubMed] [Google Scholar]
  • 17.Reisen WK, Wheeler SS, Garcia S, Fang Y. Migratory birds and the dispersal of arboviruses in California. Am J Trop Med Hyg. 2010;83:808–15. 10.4269/ajtmh.2010.10-0200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.La Sorte FA, Fink D, Hochachka WM, Farnsworth A, Rodewald AD, Rosenberg KV, et al. The role of atmospheric conditions in the seasonal dynamics of North American migration flyways. J Biogeogr. 2014;41:1685–96. 10.1111/jbi.12328 [DOI] [Google Scholar]
  • 19.La Sorte FA, Fink D, Hochachka WM, DeLong JP, Kelling S. Spring phenology of ecological productivity contributes to the use of looped migration strategies by birds. Proc Biol Sci. 2014;281:20140984–20140984. 10.1098/rspb.2014.0984 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Phillips AR. Migrations of Allen’s and other hummingbirds. Condor. 1975;77:196–205. 10.2307/1365790 [DOI] [Google Scholar]
  • 21.Klaassen RHG, Strandberg R, Hake M, Olofsson P, Tøttrup AP, Alerstam T. Loop migration in adult marsh harriers Circus aeruginosus, as revealed by satellite telemetry. J Avian Biol. 2010;41:200–7. 10.1111/j.1600-048X.2010.05058.x [DOI] [Google Scholar]
  • 22.Mellone U, López-López P, Limiñana R, Piasevoli G, Urios V. The trans-equatorial loop migration system of Eleonora’s falcon: differences in migration patterns between age classes, regions and seasons. J Avian Biol. 2013;44 [cited 2018 Oct 5]. http://doi.wiley.com/10.1111/j.1600-048X.2013.00139.x
  • 23.Willemoes M, Strandberg R, Klaassen RHG, Tøttrup AP, Vardanis Y, Howey PW, et al. Narrow-front loop migration in a population of the common cuckoo Cuculus canorus, as revealed by satellite telemetry. PLoS One. 2014;9:e83515. 10.1371/journal.pone.0083515 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Boere GC, Stroud DA. The flyway concept: what it is and what it isn’t. In: Boere GC, Galbraith C., Stroud DA, editors. Waterbirds around the World. Edinburgh: The Stationery Office; 2006. p. 40–4. [Google Scholar]
  • 25.Hatcher EL, Zhdanov SA, Bao Y, Blinkova O, Nawrocki EP, Ostapchuck Y, et al. Virus Variation Resource - improved response to emergent viral outbreaks. Nucleic Acids Res. 2017;45(D1):D482–90.file:///\\\\\\\\Cdc.gov\\\\project\\\\CCID_NCPDCID_DEISS_EIDJ\\\\EID%20Production\\\\Editorial\\\\Dec%2018%20Proofing\\\\%20PubMed 10.1093/nar/gkw1065 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Health UTMB. World Reference Center for Emerging Viruses and Arboviruses [cited 2018 Feb 20]. https://www.utmb.edu/wrceva/home
  • 27.Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30:2114–20. 10.1093/bioinformatics/btu170 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2 [cited 2014 Mar 19]. http://www.pubmedcentral.nih.gov/articlerender.fcgi?artid=3322381&tool=pmcentrez&rendertype=abstract [DOI] [PMC free article] [PubMed]
  • 29.Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. ; 1000 Genome Project Data Processing Subgroup. The Sequence Alignment/Map format and SAMtools. Bioinformatics. 2009;25:2078–9. 10.1093/bioinformatics/btp352 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Miller MA, Pfeiffer W, Schwartz T. Creating the CIPRES Science Gateway for inference of large phylogenetic trees [cited 2018 Oct 5].http://ieeexplore.ieee.org/document/5676129/
  • 31.Rambaut A, Lam TT, Max Carvalho L, Pybus OG. Exploring the temporal structure of heterochronous sequences using TempEst (formerly Path-O-Gen) [cited 2018 Oct 5].http://ve.oxfordjournals.org/lookup/doi/10.1093/ve/vew007 [DOI] [PMC free article] [PubMed]
  • 32.Drummond AJ, Suchard MA, Xie D, Rambaut A. Bayesian phylogenetics with BEAUti and the BEAST 1.7. Mol Biol Evol. 2012;29:1969–73. 10.1093/molbev/mss075 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Gray RR, Veras NMC, Santos LA, Salemi M. Evolutionary characterization of the West Nile Virus complete genome. Mol Phylogenet Evol. 2010;56:195–200. 10.1016/j.ympev.2010.01.019 [DOI] [PubMed] [Google Scholar]
  • 34.Lemey P, Rambaut A, Drummond AJ, Suchard MA. Bayesian phylogeography finds its roots. PLOS Comput Biol. 2009;5:e1000520. 10.1371/journal.pcbi.1000520 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Minin VN, Suchard MA. Counting labeled transitions in continuous-time Markov models of evolution. J Math Biol. 2008;56:391–412. 10.1007/s00285-007-0120-8 [DOI] [PubMed] [Google Scholar]
  • 36.Tohma K, Saito M, Kamigaki T, Tuason LT, Demetria CS, Orbina JRC, et al. Phylogeographic analysis of rabies viruses in the Philippines. Infect Genet Evol. 2014;23:86–94. 10.1016/j.meegid.2014.01.026 [DOI] [PubMed] [Google Scholar]
  • 37.Raghwani J, Rambaut A, Holmes EC, Hang VT, Hien TT, Farrar J, et al. Endemic dengue associated with the co-circulation of multiple viral lineages and localized density-dependent transmission. PLoS Pathog. 2011;7:e1002064. 10.1371/journal.ppat.1002064 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Ward MJ, Lycett SJ, Kalish ML, Rambaut A, Leigh Brown AJ. Estimating the rate of intersubtype recombination in early HIV-1 group M strains. J Virol. 2013;87:1967–73. 10.1128/JVI.02478-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Lemey P, Rambaut A, Bedford T, Faria N, Bielejec F, Baele G, et al. Unifying viral genetics and human transportation data to predict the global transmission dynamics of human influenza H3N2. PLoS Pathog. 2014;10:e1003932. 10.1371/journal.ppat.1003932 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Nelson MI, Viboud C, Vincent AL, Culhane MR, Detmer SE, Wentworth DE, et al. Global migration of influenza A viruses in swine. Nat Commun. 2015;6:6696. 10.1038/ncomms7696 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Samy AM, Peterson AT, Hall M. Phylogeography of Rift Valley Fever Virus in Africa and the Arabian Peninsula. PLoS Negl Trop Dis. 2017;11:e0005226. 10.1371/journal.pntd.0005226 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.US Census Bureau. State intercensal tables: 2000–2010 [cited 2018 Oct 3]. https://www.census.gov/data/tables/time-series/demo/popest/intercensal-2000-2010-state.html
  • 43.Global Consortium for H5N8 and Related Influenza Viruses. Role for migratory wild birds in the global spread of avian influenza H5N8. Science. 2016;354:213–7. 10.1126/science.aaf8852 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Tesh RB, Parsons R, Siirin M, Randle Y, Sargent C, Guzman H, et al. Year-round West Nile virus activity, Gulf Coast region, Texas and Louisiana. Emerg Infect Dis. 2004;10:1649–52. 10.3201/eid1009.040203 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Emerging Infectious Diseases are provided here courtesy of Centers for Disease Control and Prevention

RESOURCES