Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Dec 1.
Published in final edited form as: Curr Opin Virol. 2016 Oct 24;21:132–138. doi: 10.1016/j.coviro.2016.09.007

Dynamics of West Nile virus evolution in mosquito vectors

Nathan D Grubaugh 1,2, Gregory D Ebel 1
PMCID: PMC5384794  NIHMSID: NIHMS817530  PMID: 27788400

Abstract

West Nile virus remains the most common cause of arboviral encephalitis in North America. Since it was introduced, it has undergone adaptive genetic change as it spread throughout the continent. The WNV transmission cycle is relatively tractable in the laboratory. Thus the virus serves as a convenient model system for studying the population biology of mosquito-borne flaviviruses as they undergo transmission to and from mosquitoes and vertebrates. This review summarizes the current knowledge regarding the population dynamics of this virus within mosquito vectors.

Introduction

West Nile virus (WNV; Flavivirus; Flaviviridae) is a single-stranded positive sense RNA virus that exists in transmission cycles mainly involving Culex species mosquitoes and passerine birds. WNV was introduced to the Western Hemisphere in 1999 and was quickly spread throughout the US (reviewed by [1]). Understanding the mechanisms that contribute to rapid emergence and subsequent persistence of WNV almost 20 years later is critical for our understanding of other mosquito-borne outbreaks, such as the recent and ongoing epidemics of chikungunya virus (CHIKV) [2] and Zika virus [3] in the Americas. For example, molecular epidemiology demonstrated that WNV quickly adapted to local mosquito vectors during the invasion process [46], which likely enhanced transmission and facilitated its success [4,7]. CHIKV followed a similar pattern during the Indian Ocean epidemic when it adapted to be more efficiently transmitted by Aedes albopictus [8]. However, the current CHIKV epidemic in the Americas and some local emergences of WNV were not associated with previously observed vector-adaptive mutations [9,10]. What, then are the factors that favor the emergence of adaptive mutations within arbovirus populations? While the answer is not entirely clear, experimental evolution studies of WNV are currently seeking to define these conditions.

WNV exists in nature as genetically diverse populations [11]. Like other RNA viruses, genetic diversity is rapidly formed by error-prone polymerases (~10−4/site/round of replication [1214]), which seem to operate at optimal fidelity [1517]. Collectively, intrahost virus variants influence population fitness [18,19], alter disease outcome [20,21], and provide opportunities for adaptation [22,23]. However, the relationships between viral genetic diversity and phenotype become muddled once the temporal aspects of evolution are included: Viral populations are in constant flux. In general, WNV genetic diversity in mosquitoes is generated by strong diversifying selection [24,25], stochastically rearranged by bottlenecks [26,27], and persist due to weak purifying selection [11,28,29]. This produces greater diversity in mosquitoes than birds [30] and humans [31]. Here we outline the forces of selection and drift that alter WNV populations, microhabitat conditions that can direct the evolutionary pathway, and fitness costs during transmission (Figure 1).

Figure 1.

Figure 1

Dynamics of WNV evolution during mosquito transmission. (a) WNV population genetic diversity can be immediately reduced upon midgut infection through bottlenecks, introducing random genetic drift and founder’s effects. These stochastic events occur during each major anatomical barrier to infection: midgut and salivary gland infection and escape. (b) WNV population genetic diversity can be rapidly restored through negative frequency-dependent selection introduced by RNAi. Essentially, common variants are more likely targeted by RNAi-mediated degradation while rare variants with mismatches between the template RNA loaded into RISC are allowed to replicate, increasing population complexity. (c) The influence of repeated random bottlenecks and RNAi-mediated diversification leads to the formation of unique subpopulations in different mosquito tissues and compartments, including what is expectorated in saliva. Furthermore, these processes influenced by the mosquito species, leading to very different WNV populations transmitted between different vectors. (d) The combined effects of bottlenecks, diversifying selection, and weak purifying selection lead to the accumulation many deleterious mutations into a population. In addition, mosquito-adapted variants are often not as fit in birds. Thus, there are fitness trade-offs in birds, which is predicted to remove many of the WNV produced within mosquitoes. (e) Together, the input WNV population taken up by mosquitoes during bloodfeeding drastically diverges and diversifies during mosquito infection, and weak purifying selection allows for many deleterious mutations to persist. During transmission to birds, strong purifying selection removes many of the variants, decreasing WNV population genetic diversity and maintaining fitness.

Bottlenecks during systemic mosquito infection

Several physical barriers within mosquitoes impede systemic WNV infection and dramatically restructure viral populations. These mainly occur during entry and exit of the midgut and salivary glands (recently reviewed by [32,33]). Briefly, WNV must first infect the posterior portion of the midgut where contents of the bloodmeal are digested and absorbed. The virus must then pass through the basal lamina of the midgut and exit into the hemocoel to infect the hemocytes (invertebrate immune cells [34]), fat bodies, neurons, and muscle tissue [35]. Upon salivary gland infection, mature virions are transported and/or are directly released into an extracellular acinus (a holding place for saliva proteins). The contents of the acinus, including virus, are expectorated during mosquito probing and feeding. In general, Culex mosquitoes can expectorate 104–106 WNV plaque forming units during bloodfeeding [36]. Virus populations that pass through these physiological barriers are subjected to genetic bottlenecks (small effective population sizes) [26,27,37,38]. This process can dramatically alter population demographics through random (i.e. nonselective) selection of only a few viruses that establish infection in the next tissue (genetic drift, founder’s effects [39]). The number of infectious viruses that pass through a bottleneck is associated with the strength of the anatomical barrier; a weak midgut infection barrier tends to impose a weaker bottleneck than a strong barrier [26]. Most bottlenecks have a net negative effect on the virus population because they randomly fix low fitness mutations in a population [40,41]. Furthermore, systematic introductions of deleterious mutations can drive the population towards extinction unless mutation and/or recombination [42] can restore fitness (see Muller’s ratchet [43]). In some cases, however, bottlenecks may be beneficial. For example, high fitness variants may be suppressed and remain at low population frequencies when the effective population size remains large [18]. However, if the variant can survive a bottleneck, it can reach dominance because it encounters fewer competitors.

RNAi-mediated diversification

Following genetic homogenization caused by bottlenecks during systemic spread, WNV populations must rapidly diversify in each tissue to evade the mosquito’s primary antiviral response, RNA interference (RNAi) [4446]. Viral RNA is targeted for degradation by sequence complementarity to a small template RNA loaded into the RNA-induced silencing complex (RISC). Mutant viruses are poorer matches to common RISC-loaded guide strands than are un-mutated viruses and therefore evade silencing. Thus RNAi creates an intracellular milieu that promotes diversification by allowing rare viral haplotypes to replicate quickly until they are no longer rare [24,25]. This gives more genetically diverse populations a competitive advantage in mosquitoes [47].

Weak purifying selection

dN/dS ratios from intra-mosquito WNV populations are consistently greater than 1, suggesting that purifying selection is weak [11,26,29]. This could be directly related to the RNAi response, where selection happens at the nucleotide level and neither synonymous nor nonsynonymous mutations are favored. In addition, coinfection of multiple viral genomes within cells may also decrease purifying selection as they permit the persistence of deleterious mutations through complementation [26,48,49]. Together, we would expect that WNV genetic diversity would increase overtime with continued exposure to RNAi-mediated diversification and weak purifying selection. However, WNV diversity does not increase with longer extrinsic incubation periods during Cx. quinquefasciautus infection [49]. Diversifying selection and weak purifying selection therefore appear to be balanced by other forces that shape WNV populations within mosquitoes.

Variables altering the course of evolution

WNV is composed of as many as eight genetically distinct lineages and sublineages (reviewed by [1]), is an ecological generalist that can infect many mosquito species, and persists in several environments (reviewed by [50]). Therefore, the trajectory of WNV evolution is probably influenced by many different virus-, host-, and environment-dependent factors at a given time. Several of these are known to alter WNV demographics during mosquito infection. First, high frequency variants are more likely to survive bottlenecks [27] and the midgut bottleneck severity is inversely proportional to amount of virus in the bloodmeal [37]. Taken together, a WNV strain that can cause higher viremia in birds (e.g. [51,52]) may be more likely to maintain its diversity during initial mosquito infection. Second, we recently described how the species of mosquito involved in transmission can also strongly influence virus divergence, which may be associated with host susceptibility (virus replication and purifying selection) and the strengths of anatomical barriers (vector competence and bottleneck severity) [26]. These factors can be influenced by both the virus and the vector. For example, an enzootic pairing of Venezuelan equine encephalitis virus (VEEV) and Cx. taeniopus mosquitoes are more likely to maintain viral genetic diversity in than an epizootic pairing because more midgut cells become infected allowing for a larger effective population size (i.e. a weaker bottleneck) [53]. In addition, factors influencing the mosquito immune response, such as mosquito genetics [54] and its microbiome [55,56], can significantly alter vector competence. For example, pre-treating Ae. aegypti with antibiotics prior to dengue virus exposure increases the midgut viral titers due to lower immune activation in the absence of an intact microbiota [57]. In other cases, the presence of certain microbes, such as the endosymbiotic bacterium Wolbachia, can increase host resistance to WNV [58] and other mosquito-borne infections [59,60]. Third, environmental conditions impact virus-vector interactions in several ways. One of the most important of these is temperature. Higher temperatures can increase the ability of Culex mosquitoes to transmit WNV [6163]. Given global climate change and that viruses are constantly emerging into new ecological niches, determining how temperature can drive mosquito-borne virus evolution is of upmost importance. It may be that the positive correlation between temperature and vector competence also helps maintain viral genetic diversity and thus aids avoidance of the negative consequences of bottlenecks. Additionally, the frequency of spontaneous vesiculovirus mutation doubles when the temperature is raised from just 39 to 39.8 °C [64], suggesting that the WNV mutation rates could profoundly change during the 5–10 °C variations in mean temperature during the transmission season [65]. Ultimately, the numerous virus-vector-environment interactions that determine vector competence (e.g. [6668]) may all slightly redirect WNV evolution and impact virus population structure.

Positive selection

Perhaps due to the requirement to cycle in two different hosts and the strong influences of genetic drift in mosquitoes, there is very little evidence for adaptive mosquito-borne virus evolution [6973]. In fact, there are only a few known examples of positive selection the enhance virus replication or transmission within mosquitoes. The alphaviruses VEEV and CHIKV both utilized single amino acid substitutions in the envelope glycoprotein to increase vector competence of Ae. taeniorhyncus [74,75] and Ae. albopictus [8] mosquitoes, respectively. During the early years (2001–2003) of the North American WNV invasion, a locally derived variant (WN02) rapidly displaced the original (NY99) [46]. Again, the WN02 variant was demonstrated to contain a single amino acid substitution in the envelope protein (A159V) that conferred a fitness advantage by requiring a shorter extrinsic incubation period in Culex mosquitoes [4,7]. While the CHIKV E1 glycoprotein mutation (A226V) promoting enhanced infectivity in Ae. albopictus was experimentally reproduced in the laboratory [76], the same has not been demonstrated for the WN02 mutation. Specifically, the key mutation to WNV did not arise after NY99 infection of four species of birds [30] and mosquitoes [26]. One possible explanation for this is the homogeneity of the clone-derived, NY99 input virus population used in these studies. As demonstrated for CHIKV, epistatic interactions of mutations on the same haplotype have resulted in several multistep adaptive pathways [7678]. The inclusion of genetic diversity in the founding WNV population may then allow the virus to follow more natural adaptive pathways, which may include the A159V mutation. Alternatively, adaptation and fitness are context specific. Possibly the conditions used in our experimental evolution studies, such as temperature [62], did not resemble the conditions that led to the NY99 displacement. What seems to be required is that a variant must have a very high fitness value to overcome its competitors within a mosquito. Alternatively, it needs be lucky enough to survive a random bottleneck and arrive in a more favorable environment with less competition. To summarize, we are critically lacking knowledge of the conditions that favor adaptive virus mutations to arise in mosquitoes and how they survive repeated bottlenecks.

Fitness trade-offs during transmission

Fundamental theories of evolution predict that genetic diversity provides viruses with opportunities for rapid selection, and therefore adaptation, during host shifts [79,80]. Mosquitoes can transmit unique and diverse virus subpopulations in their saliva allowing the virus to explore a tremendous amount of sequence space [26,27,49,76]. While the genetic diversity provided to WNV during mosquito-borne transmission may be beneficial in some circumstances [22], in general, there is a fitness trade-off from mosquitoes to birds [81].

Our data suggests that surviving the mosquito environment, including (a) repeated bottlenecks imposing genetic drift, (b) RNAi-mediated diversifying selection, and (c) weak purifying selection, collectively imposes detrimental effects on WNV [26]. We therefore would predict that the WNV population will quickly revert towards the master sequence in birds contributing to the observed slow evolutionary rates of mosquito-borne viruses [82,83].

Conclusions

During mosquito infection, WNV populations change rapidly due to genetic drift and diversifying selection. The WNV genetic diversity thus produced likely provides opportunities for rapid adaptation and the emergence of new virus genotypes, but at the cost of lower relative fitness during transmission. Therefore WNV cycles between periods of genetic expansion in mosquitoes and selective constraint in birds. We hypothesize that periods of rapid transmission, such as during an explosive outbreak, are more likely to produce local adaptation because there are more opportunities for selection. Moreover, we also hypothesize that certain transmission cycles can increase these odds. For example, transmission involving Cx. quinquefasciautus vectors [26] and American robins [30] is more likely to allow the virus to explore new sequence space and select for highly fit variants. While a tremendous amount of progress has been made towards understanding the dynamics of WNV evolution, many questions remain regarding the factors that alter its path.

Highlights.

  • WNV populations encounter several bottlenecks during systemic mosquito infection.

  • Genetic diversity can be rapidly recovered by RNAi-mediated diversifying selection.

  • Weak purifying selection acts to maintain deleterious WNV mutations.

  • Many virus-, vector-, and environment-dependent factors can alter WNV evolution.

  • WNV populations transmitted to vertebrates encounter a fitness trade-off.

Acknowledgments

Research in the Ebel laboratory is supported by NIH under grant number AI067380.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References and recommended reading

  • 1.Pesko KN, Ebel GD. West Nile virus population genetics and evolution. Infect Genet Evol. 2012;12:181–190. doi: 10.1016/j.meegid.2011.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Leparc-Goffart I, Nougairede A, Cassadou S, Prat C, de Lamballerie X. Chikungunya in the Americas. Lancet. 2014;383:514. doi: 10.1016/S0140-6736(14)60185-9. [DOI] [PubMed] [Google Scholar]
  • 3.Faria NR, do Azevedo RS, Kraemer MU, Souza R, Cunha MS, et al. Zika virus in the Americas: Early epidemiological and genetic findings. Science. 2016;352:345–349. doi: 10.1126/science.aaf5036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Ebel GD, Carricaburu J, Young D, Bernard KA, Kramer LD. Genetic and phenotypic variation of West Nile virus in New York, 2000–2003. Am J Trop Med Hyg. 2004;71:493–500. [PubMed] [Google Scholar]
  • 5.Davis CT, Ebel GD, Lanciotti RS, Brault AC, Guzman H, et al. Phylogenetic analysis of North American West Nile virus isolates, 2001–2004: evidence for the emergence of a dominant genotype. Virology. 2005;342:252–265. doi: 10.1016/j.virol.2005.07.022. [DOI] [PubMed] [Google Scholar]
  • 6.Beasley DWC, Davis CT, Guzman H, Vanlandingham DL, da Rosa APAT, et al. Limited evolution of West Nile virus has occurred during its southwesterly spread in the United States. Virology. 2003;309:190–195. doi: 10.1016/s0042-6822(03)00150-8. [DOI] [PubMed] [Google Scholar]
  • 7.Moudy RM, Meola MA, Morin LL, Ebel GD, Kramer LD. A newly emergent genotype of West Nile virus is transmitted earlier and more efficiently by Culex mosquitoes. Am J Trop Med Hyg. 2007;77:365–370. [PubMed] [Google Scholar]
  • 8.Tsetsarkin KA, Vanlandingham DL, Mcgee CE, Higgs S. A single mutation in chikungunya virus affects vector specificity and epidemic potential. PLoS Pathog. 2007;3:1895–1906. doi: 10.1371/journal.ppat.0030201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Duggal NK, D’Anton M, Xiang J, Seiferth R, Day J, et al. Sequence analyses of 2012 West Nile virus isolates from Texas fail to associate viral genetic factors with outbreak magnitude. Am J Trop Med Hyg. 2013;89:205–210. doi: 10.4269/ajtmh.13-0140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Stapleford KA, Moratorio G, Henningsson R, Chen R, Matheus S, et al. Whole-Genome Sequencing Analysis from the Chikungunya Virus Caribbean Outbreak Reveals Novel Evolutionary Genomic Elements. PLoS Negl Trop Dis. 2016;10:e0004402. doi: 10.1371/journal.pntd.0004402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11*.Jerzak G, Bernard KA, Kramer LD, Ebel GD. Genetic variation in West Nile virus from naturally infected mosquitoes and birds suggests quasispecies structure and strong purifying selection. J Gen Virol. 2005;86:2175–2183. doi: 10.1099/vir.0.81015-0. First to demonstrate that natural intrahost WNV populations are genetically diverse. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12*.Acevedo A, Brodsky L, Andino R. Mutational and fitness landscapes of an RNA virus revealed through population sequencing. Nature. 2014;505:686–690. doi: 10.1038/nature12861. Shows the vast amounts of low frequency mutations that exist within viral populations at a given time using a sequencing approach (CirSeq) that can achieve great depth and high accuracy. The structure of WNV populations are likely similar. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Holland J, Spindler K, Horodyski F, Grabau E, Nichol S, et al. Rapid evolution of RNA genomes. Science. 1982;215:1577–1585. doi: 10.1126/science.7041255. [DOI] [PubMed] [Google Scholar]
  • 14.Duffy S, Shackelton LA, Holmes EC. Rates of evolutionary change in viruses: patterns and determinants. Nat Rev Genet. 2008;9:267–276. doi: 10.1038/nrg2323. [DOI] [PubMed] [Google Scholar]
  • 15*.Van Slyke GA, Arnold JJ, Lugo AJ, Griesemer SB, Moustafa IM, et al. Sequence-Specific Fidelity Alterations Associated with West Nile Virus Attenuation in Mosquitoes. PLoS Pathog. 2015;11:e1005009. doi: 10.1371/journal.ppat.1005009. Shows that altering WNV replication fidelity diminishes its ability to establish infection in Cx. quinquefasciautus mosquitoes, demonstrating the vast importance of mutation rates for maintaining fitness. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Rozen-Gagnon K, Stapleford KA, Mongelli V, Blanc H, Failloux AB, et al. Alphavirus mutator variants present host-specific defects and attenuation in mammalian and insect models. PLoS Pathog. 2014;10:e1003877. doi: 10.1371/journal.ppat.1003877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Coffey LL, Beeharry Y, Borderia AV, Blanc H, Vignuzzi M. Arbovirus high fidelity variant loses fitness in mosquitoes and mice. Proc Natl Acad Sci U S A. 2011;108:16038–16043. doi: 10.1073/pnas.1111650108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18**.de la Torre JC, Holland JJ. RNA virus quasispecies populations can suppress vastly superior mutant progeny. J Virol. 1990;64:6278–6281. doi: 10.1128/jvi.64.12.6278-6281.1990. Seminal study demonstrating the relationships between virus population size and competition among variants. Highly fit variants can be suppressed during infection through intracellular complementation, suggesting that there may be some advantages for some WNV variants if they survive a bottleneck. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Ciota AT, Ehrbar DJ, Van Slyke GA, Willsey GG, Kramer LD. Cooperative interactions in the West Nile virus mutant swarm. BMC Evol Biol. 2012;12:58. doi: 10.1186/1471-2148-12-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20**.Vignuzzi M, Stone JK, Arnold JJ, Cameron CE, Andino R. Quasispecies diversity determines pathogenesis through cooperative interactions in a viral population. Nature. 2006;439:344–348. doi: 10.1038/nature04388. The most convincing study to date showing how viruses within a population can interact to alter the course of infection. Together with [21,47] demonstrates that WNV diversity also directly impacts the infection phenotype and population fitness. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21*.Ebel GD, Fitzpatrick KA, Lim PY, Bennett CJ, Deardorff ER, et al. Nonconsensus West Nile virus genomes arising during mosquito infection suppress pathogenesis and modulate virus fitness in vivo. J Virol. 2011;85:12605–12613. doi: 10.1128/JVI.05637-11. See [20] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22*.Stern A, Bianco S, Yeh MT, Wright C, Butcher K, et al. Costs and benefits of mutational robustness in RNA viruses. Cell Rep. 2014;8:1026–1036. doi: 10.1016/j.celrep.2014.07.011. Shows that while mutational robustness can facilitate adaptation, it can also have a severe consequence during transmission if the mutations are deleterious in the new host. Provides additional support for the fitness decreases of WNV recovered from mosquito saliva during avian cell infection [26] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Ciota AT, Ngo KA, Lovelace AO, Payne AF, Zhou Y, et al. Role of the mutant spectrum in adaptation and replication of West Nile virus. Journal of General Virology. 2007;88:865–874. doi: 10.1099/vir.0.82606-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24**.Brackney DE, Schirtzinger EE, Harrison TD, Ebel GD, Hanley KA. Modulation of flavivirus population diversity by RNA interference. J Virol. 2015;89:4035–4039. doi: 10.1128/JVI.02612-14. Together with [25], these papers describe that WNV diversification in mosquitoes is driven by RNAi, a process that is predicted to allow for rare variants to replicated until they are no longer rare (i.e. negative frequency-dependent selection) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25*.Brackney DE, Beane JE, Ebel GD. RNAi targeting of West Nile virus in mosquito midguts promotes virus diversification. PLoS Pathog. 2009;5:e1000502. doi: 10.1371/journal.ppat.1000502. See [24] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26**.Grubaugh ND, Weger-Lucarelli J, Murrieta RA, Fauver JR, Garcia-Luna SM, et al. Genetic Drift during Systemic Arbovirus Infection of Mosquito Vectors Leads to Decreased Relative Fitness during Host Switching. Cell Host Microbe. 2016;19:481–492. doi: 10.1016/j.chom.2016.03.002. Demonstrates that WNV bottlenecks occur at each mosquito anatomical barrier, but genetic diversity rapidly recovers and is maintained by weak purifying selection. The consequence of these actions is that deleterious mutations accumulate, decreasing fitness during transmission. Also, evidence is shown supporting species-dependent impacts on WNV evolution. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Ciota AT, Ehrbar DJ, Van Slyke GA, Payne AF, Willsey GG, et al. Quantification of intrahost bottlenecks of West Nile virus in Culex pipiens mosquitoes using an artificial mutant swarm. Infection Genetics and Evolution. 2012;12:557–564. doi: 10.1016/j.meegid.2012.01.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Jerzak GV, Brown I, Shi PY, Kramer LD, Ebel GD. Genetic diversity and purifying selection in West Nile virus populations are maintained during host switching. Virology. 2008;374:256–260. doi: 10.1016/j.virol.2008.02.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Jerzak GV, Bernard K, Kramer LD, Shi PY, Ebel GD. The West Nile virus mutant spectrum is host-dependant and a determinant of mortality in mice. Virology. 2007;360:469–476. doi: 10.1016/j.virol.2006.10.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Grubaugh ND, Smith DR, Brackney DE, Bosco-Lauth AM, Fauver JR, et al. Experimental evolution of an RNA virus in wild birds: evidence for host-dependent impacts on population structure and competitive fitness. PLoS Pathog. 2015;11:e1004874. doi: 10.1371/journal.ppat.1004874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Grubaugh ND, Massey A, Shives KD, Stenglein MD, Ebel GD, et al. West Nile Virus Population Structure, Injury, and Interferon-Stimulated Gene Expression in the Brain From a Fatal Case of Encephalitis. Open Forum Infect Dis. 2016;3:ofv182. doi: 10.1093/ofid/ofv182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Franz AW, Kantor AM, Passarelli AL, Clem RJ. Tissue Barriers to Arbovirus Infection in Mosquitoes. Viruses. 2015;7:3741–3767. doi: 10.3390/v7072795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Kramer LD, Ciota AT. Dissecting vectorial capacity for mosquito-borne viruses. Curr Opin Virol. 2015;15:112–118. doi: 10.1016/j.coviro.2015.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Lavine MD, Strand MR. Insect hemocytes and their role in immunity. Insect Biochem Mol Biol. 2002;32:1295–1309. doi: 10.1016/s0965-1748(02)00092-9. [DOI] [PubMed] [Google Scholar]
  • 35.Girard YA, Klingler KA, Higgs S. West Nile virus dissemination and tissue tropisms in orally infected Culex pipiens quinquefasciatus. Vector-Borne and Zoonotic Diseases. 2004;4:109–122. doi: 10.1089/1530366041210729. [DOI] [PubMed] [Google Scholar]
  • 36.Styer LM, Kent KA, Albright RG, Bennett CJ, Kramer LD, et al. Mosquitoes inoculate high doses of West Nile virus as they probe and feed on live hosts. PLoS Pathog. 2007;3:1262–1270. doi: 10.1371/journal.ppat.0030132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Forrester NL, Guerbois M, Seymour RL, Spratt H, Weaver SC. Vector-Borne Transmission Imposes a Severe Bottleneck on an RNA Virus Population. PLoS Pathog. 2012;8 doi: 10.1371/journal.ppat.1002897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Forrester NL, Coffey LL, Weaver SC. Arboviral bottlenecks and challenges to maintaining diversity and fitness during mosquito transmission. Viruses. 2014;6:3991–4004. doi: 10.3390/v6103991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Mayr E. Systematics and the origin of species from the viewpoint of a zoologist. New York: Columbia University Press; 1942. p. xiv.p. 334. incl. illus. (incl. maps) tables, diagrs. p. [Google Scholar]
  • 40*.Duarte E, Clarke D, Moya A, Domingo E, Holland J. Rapid fitness losses in mammalian RNA virus clones due to Muller’s ratchet. Proc Natl Acad Sci U S A. 1992;89:6015–6019. doi: 10.1073/pnas.89.13.6015. Together with [41], shows how bottlenecks can randomly fix deleterious mutations into a RNA virus population, gradually causing severe fitness declines. These seminal studies help explain how deleterious mutations are introduced into WNV populations during systemic mosquito infection. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41*.Chao L. Fitness of RNA virus decreased by Muller’s ratchet. Nature. 1990;348:454–455. doi: 10.1038/348454a0. See [40] [DOI] [PubMed] [Google Scholar]
  • 42.Xiao Y, Rouzine IM, Bianco S, Acevedo A, Goldstein EF, et al. RNA Recombination Enhances Adaptability and Is Required for Virus Spread and Virulence. Cell Host Microbe. 2016;19:493–503. doi: 10.1016/j.chom.2016.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Muller HJ. The Relation of Recombination to Mutational Advance. Mutat Res. 1964;1:2–9. doi: 10.1016/0027-5107(64)90047-8. [DOI] [PubMed] [Google Scholar]
  • 44.Sanchez-Vargas I, Scott JC, Poole-Smith BK, Franz AW, Barbosa-Solomieu V, et al. Dengue virus type 2 infections of Aedes aegypti are modulated by the mosquito’s RNA interference pathway. PLoS Pathog. 2009;5:e1000299. doi: 10.1371/journal.ppat.1000299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Campbell CL, Keene KM, Brackney DE, Olson KE, Blair CD, et al. Aedes aegypti uses RNA interference in defense against Sindbis virus infection. BMC Microbiol. 2008;8:47. doi: 10.1186/1471-2180-8-47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Olson KE, Blair CD. Arbovirus-mosquito interactions: RNAi pathway. Curr Opin Virol. 2015;15:119–126. doi: 10.1016/j.coviro.2015.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47*.Fitzpatrick KA, Deardorff ER, Pesko K, Brackney DE, Zhang B, et al. Population variation of West Nile virus confers a host-specific fitness benefit in mosquitoes. Virology. 2010;404:89–95. doi: 10.1016/j.virol.2010.04.029. Shows that WNV population genetic diversity is positively correlated with fitness in mosquitoes but not in birds. This further demonstrates how population diversity can influence fitness (see also [20]) and helps to support that rare variants are selected by RNAi (see also [24]) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Aaskov J, Buzacott K, Thu HM, Lowry K, Holmes EC. Long-term transmission of defective RNA viruses in humans and Aedes mosquitoes. Science. 2006;311:236–238. doi: 10.1126/science.1115030. [DOI] [PubMed] [Google Scholar]
  • 49.Brackney DE, Pesko KN, Brown IK, Deardorff ER, Kawatachi J, et al. West Nile virus genetic diversity is maintained during transmission by Culex pipiens quinquefasciatus mosquitoes. PLoS One. 2011;6:e24466. doi: 10.1371/journal.pone.0024466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Reisen WK. Ecology of West Nile virus in North America. Viruses. 2013;5:2079–2105. doi: 10.3390/v5092079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Brault AC, Langevin SA, Bowen RA, Panella NA, Biggerstaff BJ, et al. Differential virulence of West Nile strains for American crows. Emerg Infect Dis. 2004;10:2161–2168. doi: 10.3201/eid1012.040486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Brault AC, Langevin SA, Ramey WN, Fang Y, Beasley DW, et al. Reduced avian virulence and viremia of West Nile virus isolates from Mexico and Texas. Am J Trop Med Hyg. 2011;85:758–767. doi: 10.4269/ajtmh.2011.10-0439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Gutierrez S, Thebaud G, Smith DR, Kenney JL, Weaver SC. Demographics of Natural Oral Infection of Mosquitos by Venezuelan Equine Encephalitis Virus. J Virol. 2015;89:4020–4022. doi: 10.1128/JVI.03265-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Lambrechts L, Quillery E, Noel V, Richardson JH, Jarman RG, et al. Specificity of resistance to dengue virus isolates is associated with genotypes of the mosquito antiviral gene Dicer-2. Proc Biol Sci. 2013;280:20122437. doi: 10.1098/rspb.2012.2437. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Weiss B, Aksoy S. Microbiome influences on insect host vector competence. Trends Parasitol. 2011;27:514–522. doi: 10.1016/j.pt.2011.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Dennison NJ, Jupatanakul N, Dimopoulos G. The mosquito microbiota influences vector competence for human pathogens. Curr Opin Insect Sci. 2014;3:6–13. doi: 10.1016/j.cois.2014.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Xi Z, Ramirez JL, Dimopoulos G. The Aedes aegypti toll pathway controls dengue virus infection. PLoS Pathog. 2008;4:e1000098. doi: 10.1371/journal.ppat.1000098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Glaser RL, Meola MA. The native Wolbachia endosymbionts of Drosophila melanogaster and Culex quinquefasciatus increase host resistance to West Nile virus infection. PLoS One. 2010;5:e11977. doi: 10.1371/journal.pone.0011977. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Bian G, Xu Y, Lu P, Xie Y, Xi Z. The endosymbiotic bacterium Wolbachia induces resistance to dengue virus in Aedes aegypti. PLoS Pathog. 2010;6:e1000833. doi: 10.1371/journal.ppat.1000833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Moreira LA, Iturbe-Ormaetxe I, Jeffery JA, Lu G, Pyke AT, et al. A Wolbachia symbiont in Aedes aegypti limits infection with dengue, Chikungunya, and Plasmodium. Cell. 2009;139:1268–1278. doi: 10.1016/j.cell.2009.11.042. [DOI] [PubMed] [Google Scholar]
  • 61.Dohm DJ, O’Guinn ML, Turell MJ. Effect of environmental temperature on the ability of Culex pipiens (Diptera: Culicidae) to transmit West Nile virus. J Med Entomol. 2002;39:221–225. doi: 10.1603/0022-2585-39.1.221. [DOI] [PubMed] [Google Scholar]
  • 62.Kilpatrick AM, Meola MA, Moudy RM, Kramer LD. Temperature, viral genetics, and the transmission of West Nile virus by Culex pipiens mosquitoes. PLoS Pathog. 2008;4 doi: 10.1371/journal.ppat.1000092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Danforth ME, Reisen WK, Barker CM. The Impact of Cycling Temperature on the Transmission of West Nile Virus. J Med Entomol. 2016 doi: 10.1093/jme/tjw013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Flamand A. Rhabdovirus genetics. In: Bishop DHL, editor. Rhabdovirus. Boca Raton: CRC Press; 1980. pp. 115–139. [Google Scholar]
  • 65.Fauver JR, Pecher L, Schurich JA, Bolling BG, Calhoon M, et al. Temporal and Spatial Variability of Entomological Risk Indices for West Nile Virus Infection in Northern Colorado: 2006–2013. J Med Entomol. 2015 doi: 10.1093/jme/tjv234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Dickson LB, Sanchez-Vargas I, Sylla M, Fleming K, Black WCt. Vector competence in West African Aedes aegypti Is Flavivirus species and genotype dependent. PLoS Negl Trop Dis. 2014;8:e3153. doi: 10.1371/journal.pntd.0003153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Fansiri T, Fontaine A, Diancourt L, Caro V, Thaisomboonsuk B, et al. Genetic mapping of specific interactions between Aedes aegypti mosquitoes and dengue viruses. Plos Genetics. 2013;9:e1003621. doi: 10.1371/journal.pgen.1003621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Kilpatrick AM, Fonseca DM, Ebel GD, Reddy MR, Kramer LD. Spatial and temporal variation in vector competence of Culex pipiens and Cx. restuans mosquitoes for West Nile virus. Am J Trop Med Hyg. 2010;83:607–613. doi: 10.4269/ajtmh.2010.10-0005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Woelk CH, Holmes EC. Reduced positive selection in vector-borne RNA viruses. Molecular Biology and Evolution. 2002;19:2333–2336. doi: 10.1093/oxfordjournals.molbev.a004059. [DOI] [PubMed] [Google Scholar]
  • 70.Di Giallonardo F, Geoghegan JL, Docherty DE, McLean RG, Zody MC, et al. Fluid Spatial Dynamics of West Nile Virus in the USA: Rapid Spread in a Permissive Host Environment. J Virol. 2015 doi: 10.1128/JVI.02305-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Jenkins GM, Rambaut A, Pybus OG, Holmes EC. Rates of molecular evolution in RNA viruses: a quantitative phylogenetic analysis. Journal of Molecular Evolution. 2002;54:156–165. doi: 10.1007/s00239-001-0064-3. [DOI] [PubMed] [Google Scholar]
  • 72.Weaver SC, Rico-Hesse R, Scott TW. Genetic diversity and slow rates of evolution in New World alphaviruses. Curr Top Microbiol Immunol. 1992;176:99–117. doi: 10.1007/978-3-642-77011-1_7. [DOI] [PubMed] [Google Scholar]
  • 73.Holmes EC. Patterns of intra- and interhost nonsynonymous variation reveal strong purifying selection in dengue virus. J Virol. 2003;77:11296–11298. doi: 10.1128/JVI.77.20.11296-11298.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Smith DR, Arrigo NC, Leal G, Muehlberger LE, Weaver SC. Infection and dissemination of Venezuelan equine encephalitis virus in the epidemic mosquito vector, Aedes taeniorhynchus. Am J Trop Med Hyg. 2007;77:176–187. [PubMed] [Google Scholar]
  • 75.Brault AC, Powers AM, Ortiz D, Estrada-Franco JG, Navarro-Lopez R, et al. Venezuelan equine encephalitis emergence: enhanced vector infection from a single amino acid substitution in the envelope glycoprotein. Proc Natl Acad Sci U S A. 2004;101:11344–11349. doi: 10.1073/pnas.0402905101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76**.Stapleford KA, Coffey LL, Lay S, Borderia AV, Duong V, et al. Emergence and Transmission of Arbovirus Evolutionary Intermediates with Epidemic Potential. Cell Host Microbe. 2014;15:706–716. doi: 10.1016/j.chom.2014.05.008. Experimentally reproduced a naturally occurring CHIKV mutation that helped to fuel an epidemic, and provides evidence that proactive evolution studies are capable of predicting arboviruses that may emerge. [DOI] [PubMed] [Google Scholar]
  • 77*.Tsetsarkin KA, Chen RB, Yun RM, Rossi SL, Plante KS, et al. Multi-peaked adaptive landscape for chikungunya virus evolution predicts continued fitness optimization in Aedes albopictus mosquitoes. Nature Communications. 2014;5 doi: 10.1038/ncomms5084. Shows that arbovirus evolution can occur in multiple steps, whereas one mutation is required for the selection of another. This also suggests that using arbovirus populations with a level of natural diversity may be better for detecting selection by already including the first steps. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Tsetsarkin KA, Weaver SC. Sequential Adaptive Mutations Enhance Efficient Vector Switching by Chikungunya Virus and Its Epidemic Emergence. PLoS Pathog. 2011;7 doi: 10.1371/journal.ppat.1002412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Fisher RA. The genetical theory of natural selection. Oxford: The Clarendon press; 1930. p. xiv.p. 272. [Google Scholar]
  • 80.Wright S. Evolution and the Genetics of Populations. Chicago: University of Chicago Press; 1977. [Google Scholar]
  • 81*.Deardorff ER, Fitzpatrick KA, Jerzak GV, Shi PY, Kramer LD, et al. West Nile virus experimental evolution in vivo and the trade-off hypothesis. PLoS Pathog. 2011;7:e1002335. doi: 10.1371/journal.ppat.1002335. Shows that for WNV, a fitness trade-off occurs from mosquitoes to birds, but not vice versa. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Ciota AT, Kramer LD. Insights into arbovirus evolution and adaptation from experimental studies. Viruses. 2010;2:2594–2617. doi: 10.3390/v2122594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Holmes EC. Error thresholds and the constraints to RNA virus evolution. Trends Microbiol. 2003;11:543–546. doi: 10.1016/j.tim.2003.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES