You are on page 1of 9

Selection-Driven Evolution of Emergent Dengue Virus

Shannon N. Bennett,* Edward C. Holmes, Maritza Chirivella,ৠDania M. Rodriguez,*


Manuela Beltran,§ Vance Vorndam,§ Duane J. Gubler,k and W. Owen McMillan*
*Department of Biology, University of Puerto Rico–Rio Piedras, San Juan, Puerto Rico; Department of Zoology, University of
Oxford, Oxford, England; àDepartment of Microbiology and Medical Zoology, University of Puerto Rico–Ciencias Medicas,
San Juan, Puerto Rico; §Centers for Disease Control and Prevention, San Juan Branch, San Juan, Puerto Rico; kCenters for
Disease Control and Prevention, Fort Collins, Colorado

In the last four decades the incidence of dengue fever has increased 30-fold worldwide, and over half the world’s
population is now threatened with infection from one or more of four co-circulating viral serotypes (DEN-1 through
DEN-4). To determine the role of viral molecular evolution in emergent disease dynamics, we sequenced 40% of the
genome of 82 DEN-4 isolates collected from Puerto Rico over the 20 years since the onset of endemic dengue on the
island. Isolates were derived from years with varying levels of DEN-4 prevalence. Over our sampling period there were
marked evolutionary shifts in DEN-4 viral populations circulating in Puerto Rico; viral lineages were temporally

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


clustered and the most common genotype at a particular sampling time often arose from a previously rare lineage.
Expressed changes in structural genes did not appear to drive this lineage turnover, even though these regions include
primary determinants of viral antigenic properties. Instead, recent dengue evolution can be attributed in part to positive
selection on the nonstructural gene 2A (NS2A), whose functions may include replication efficiency and antigenicity.
During the latest and most severe DEN-4 epidemic in Puerto Rico, in 1998, viruses were distinguished by three amino
acid changes in NS2A that were fixed far faster than expected by drift alone. Our study therefore demonstrates viral
genetic turnover within a focal population and the potential importance of adaptive evolution in viral epidemic
expansion.

Introduction
RNA viruses comprise one of the fastest growing vector (Aedes aegypti), increasing host densities, particu-
categories of emergent diseases (Domingo and Holland larly in urban centers, and global travel have substantially
1997). Although they exhibit remarkable genetic diversity, altered dengue’s epidemiologic landscape (Gubler 1998).
attributable to intrinsically high rates of mutation and Now dengue annually infects an estimated 50 million to
replication as well as large population sizes (Domingo and 100 million people worldwide (WHO 1999), many of
Holland 1997; Drake and Holland 1999), the role of viral whom are exposed to two or more co-circulating DEN
evolution in determining disease dynamics has only been serotypes (hyperendemicity), resulting in frequent large-
described in a few cases (for example, Bush et al. 1999; scale epidemics and more frequent severe disease (Gubler
Zanotto et al. 1999; Manzin et al. 2000; Hatta et al. 2001). 1998).
We examine evolutionary change in dengue (DEN), an Determining the contributing factors to the emer-
acute mosquito-borne RNA virus (genus Flavivirus), over gence of dengue as a global pandemic, particularly the
a 20-year period that has marked the emergence of dengue increasing incidence of DHF and DSS, has proven difficult
in Puerto Rico (PR), a dense urban population whose both because there are no satisfactory models or in vitro
growth rate rivals Asian population centers. The virus, correlates with which to study disease transmissibility or
which causes dengue fever (DF), and the more severe pathogenicity directly (Rothman and Ennis 1999), and
dengue hemorrhagic fever (DHF) and dengue shock because most molecular epidemiologic studies to date have
syndrome (DSS), consists of four antigenically distinct had limited scope (Holmes 1998). Associations have been
serotypes, DEN-1 through DEN-4, that are evolutionarily demonstrated between severe manifestations of dengue
derived from at least three independent introductions into (DHF/DSS) and both host infection history and viral
humans from wild primates in Africa and Southeast Asia genotype. Most notably, secondary infections with heter-
(Wang et al. 2000). There is also abundant genetic ologous serotypes are more likely to develop into DHF/
diversity within each serotype, in the guise of phyloge- DSS than primary infections (Halstead 1988; Thein et al.
netically distinct clusters of sequences often referred to as 1997; Gubler 1998) so that increasing hyperendemicity
‘‘genotypes’’ (reviewed in Holmes and Burch 2000). could account in part for the rise of DHF/DSS. However,
The ongoing expansion of dengue throughout Asia there is also evidence that viral genotype may be
and the South Pacific is being recapitulated in the a contributing factor in determining dengue disease. For
Americas (Gubler 1998). Before the 1950s, people were example, attenuated and virulent strains of DEN-2 were
typically exposed to a single strain (hypoendemicity), and first observed simultaneously in the Tonga epidemics of
epidemics were rare and self-limiting (Gubler 1998). 1974/75 (Gubler et al. 1978), and the introduction of
However, geographic expansion of the primary mosquito
a genetically distinct Asian DEN-2 strain into the
Americas has been associated with an increase in DHF/
Key words: dengue virus, positive selection, epidemiology, phy- DSS (Rico-Hesse et al. 1997; Leitmeyer et al. 1999). More
logeny, maximum likelihood. tentatively, an analysis of selection pressures acting on
E-mail: sbennett@rrpac.upr.clu.edu. dengue virus genomes suggested that genotypes of DEN-2
Mol. Biol. Evol. 20(10):1650–1658. 2003 have selectively determined differences in transmissibility,
DOI: 10.1093/molbev/msg182
Molecular Biology and Evolution, Vol. 20, No. 10, in turn determining their ability to cause epidemics on
Ó Society for Molecular Biology and Evolution 2003; all rights reserved. a global scale (Twiddy et al. 2002).

1650
Selection in Emergent Dengue Virus 1651

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


FIG. 1.—Incidence of dengue virus in Puerto Rico since 1981. Years included in this study are marked on the x-axis with a black bar: * denotes the
sample from Dominica. The rise of dengue in Puerto Rico mirrors the onset of the dengue pandemic in the New World. Before WWII, epidemics in
Puerto Rico were rare, but subsequent years have been marked by frequent epidemics and, since the 1980s, continuous hyperendemic transmission
(Dietz et al. 1996; Gubler 1998). Of the dengue cases reported annually (solid black line, right axis), a subset is submitted to the CDC, isolated, and
identified to serotype (hatched area, left axis, plotted against month/year of isolation). The proportions that were DEN-4 are shaded solid gray. Because
of dengue’s variable etiology, it often goes unreported, and thus the number of recorded cases underrepresents the true number of dengue infections by
up to an estimated factor of 50 to 100 (WHO 1999).

Puerto Rico provides an ideal natural laboratory to island. We sample nearly 40% of the viral genome,
gather a detailed record of viral evolutionary change during including all the structural genes known to be important in
disease expansion. The island has a large urban population viral packaging and host cell entry, as well as a subset of
with high mosquito vector densities and, like many tropical nonstructural genes, from 82 viral isolates collected over
regions, has experienced nearly 20 years of dengue a 20-year period. Thus we expand current knowledge of
epidemics that are becoming increasingly severe (Gubler dengue molecular evolution to include genes never before
1998). Although dengue fever was recorded in Puerto Rico systematically surveyed on this scale (Holmes 1998). We
as early as 1915 (Dietz et al. 1996), continuous trans- assess the role of viral molecular evolution in disease
mission of all four serotypes has only occurred since the dynamics by testing for a viral adaptive basis to the
1980s (Dietz et al. 1996; Gubler 1998). The first epidemic changing patterns of DEN-4 incidence in Puerto Rico.
in Puerto Rico, consisting primarily of DEN-4, was Hence, we ascertain for the first time the role of natural
reported in 1981/82, followed by another DEN-4–domi- selection in DEN-4 evolution in the context of a well-
nated outbreak in 1986, this one marked by high characterized pattern of epidemic outbreaks.
incidences of DHF/DSS (Dietz et al. 1996; fig. 1). DHF/
DSS cases have occurred periodically since the 1980s,
Materials and Methods
reaching record levels in the latest DEN-4 epidemic in
Puerto Rico in 1998. We examined substitution patterns in 82 DEN-4
Taking advantage of Puerto Rico’s turbulent epide- isolates from Puerto Rico and surrounding regions since
miological record, we use a longitudinal phylogenetic the disease was established in 1981/82 (Dietz et al. 1996).
approach to recover the history of evolutionary change in We subsampled viral isolates from the U.S. Centers for
DEN-4 during disease emergence. In the absence of Disease Control and Prevention (CDC) sample bank that
experimental models, phylogenetic analyses within a focal had been collected in Puerto Rico during the years 1982 (n¼
population provide the only method with which to correlate 14), 1986/87 (n ¼ 19), 1992 (n ¼ 15), 1994 (n ¼ 14), and
viral genetic change with epidemic behavior. Herein we 1998 (n ¼ 13) to represent both endemic and epidemic
examine viral evolution in DEN-4 isolates collected from disease conditions (fig. 1). With 13 to 19 isolates per
Puerto Rico since the onset of epidemic dengue on the year-group, we have a 75%–86% chance of sampling rare
1652 Bennett et al.

Table 1
Primers Designed to Amplify and Sequence DEN-4 Gene Regions
Labela Sequence Function Gene/Gene Fragment
90U 59ATCTCTGGAAAAATGAACCAACGAA Amplification Capsid/prMem
842L 59ATAAGCCATAAATCCTGCCAAGAGC Amplification Capsid/prMem
138U 59AATATGCTGAAACGCGAGAGAAACC Sequencing Capsid/prMem
410L 59TACGGTGGGAATCAAGCACAGCAA Sequencing Capsid/prMem
518U 59GACAACAGAGGGGATCAACAAATGC Sequencing Capsid/prMem
736L 59GCTCTTGTTTCCAATCCCATTCCTG Sequencing Capsid/prMem
616U 59CCGAACCTGAAGACATTGATTGCTG Amplification EnvA
1676L 59TTCCAGCACTGTCACATCCTGTCTC Amplification EnvA
486U 59CACGTATAAATGCCCCCTACTGGTC Amplification EnvA
1786L 59GCTGTGTTTCTGCCATCTCTTTGTC Amplification EnvA
686U 59GAGCGGAGAACGGAGACGAGAGAAG Sequencing EnvA
1142U 59AACTACGGCAACAAGATGTCCAACG Sequencing EnvA
1181L 59CTGTTGGTCCTGTTCCTCTTTCAGA Sequencing EnvA

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


1603L 59TGAACCTCTGATGTGTCTGCTCCTG Sequencing EnvA
580U 59ACCCAGAGCGGAGAACGGAGACGAG Sequencing EnvA
803L 59GGGGCGACCAGCATCATTAGGACAA Sequencing EnvA
967U 59GAACTGACTAAGACAACAGCCAAGG Sequencing EnvA
1136L 59AACAAGCCACAGCCATTGCCCCACC Sequencing EnvA
1363U 59CCGGACTATGGAGAACTAACACTCG Sequencing EnvA
1658L 59TGTCCTGCAAACATGTGATTTCCAT Sequencing EnvA
1568U 59GCAATGGTTTTTGAATCTGCCTCTT Amplification EnvB/NS1
2679L 59CCTTCACATCCCCAGCCACTACAGT Amplification EnvB/NS1
1602U 59GCAGGAGCAGACACATCAGAGGTTC Sequencing EnvB/NS1
2114U 59GAAAGGGAGTTCCATTGGCAAGATG Sequencing EnvB/NS1
1985L 59CAAAGGGGTGGATGAGATGATACGC Sequencing EnvB/NS1
2519L 59TCTCGCTGGGGACTCTGGTTGAAAT Sequencing EnvB/NS1
3528U 59TTTGTGGAAGAATGCTTGAGGAGAA Amplification NS2A
4225L 59GCCAGAAGTAAGCCTCCTGCCACCA Amplification NS2A
3592U 59CTCTTTGTGCTATCATCTTGGGAGG Sequencing NS2A
4143L 59AACCCACAGCCATTATGCCCTCGTT Sequencing NS2A
7038U 59CTAATGGGGCTTGGAAAAGGATGGC Amplification NS4B
7769L 59TACAACTTCCCCTTTTGGCTTTACC Amplification NS4B
7106U 59ATGCTATTCTCAAGTGAACCCAACA Sequencing NS4B
7674L 59CTTTCAGGGCAGACTTGGCTTCAGT Sequencing NS4B
10133U 59CACCTGGGCGAAGAACATTCACACG Amplification / sequencing 39NTR
10600L 59CACCAATCCATCTTGCGGCGCTCTG Amplification / sequencing 39NTR
10612L 59TTGGATCAACAACACCAATCCATCT Amplification / sequencing 39NTR
10620L 59AGAACCTGTTGGATCAACAACACCA Amplification / sequencing 39NTR
a
Number indicates genome nucleotide position according to Zhao et al. (1986): U for forward and L for reverse.

alleles (defined as existing at a frequency of 10% in the and envelope: E), a subset of nonstructural genes (NS1,
population) at least once. In addition to 75 Puerto Rican NS2A, and NS4B), and the noncoding 39 NTR region.
isolates, we included seven isolates sampled from outside Amplifications were divided into separate reactions
Puerto Rico during the same period: three originating according to length of the target. Before sequencing, RT-
within the Caribbean basin, two from Central America, PCR products were purified using Qiagen PCR purifica-
and one from Ecuador. Caribbean basin samples included tion kits (Qiagen GmbH). We sequenced both strands of
a 1981 sample from Dominica, Lesser Antilles, believed the amplified products using forward and reverse primers
to represent the introduction of Asian DEN-4 into the (table 1) in standard dye-labeling reactions. Sequence data
Caribbean basin. were collected on an ABI 377 slab-gel automated
All samples have low passage histories, reducing the sequencer (Applied Biosystems), edited, and compiled
risk of artificial selection in vitro: only those samples with Sequencher 3.1.1 (Gene Codes) and aligned against
derived from chronic (generally low) infections were first reference sequences (GenBank number M14931; Zhao
cultured in A6/C36 mosquito cells, for one or, at most, two et al. 1986, Mackow et al. 1987) using Megalign’s clustal
passages, prior to RNA extraction. To further eliminate algorithm (version 3.1.7, Lasergene). We imported aligned
potential biases due to artificial selection, samples were not sequences into PAUP* (Swofford 2001) for phylogenetic
processed in temporal (year) order. We extracted sample analysis.
RNA using QIAamp Viral RNA Mini kits (Qiagen Recombination, reported in all DEN serotypes
GmbH). For each isolate we amplified, using reverse- (Worobey, Rambaut, and Holmes 1999; Tolou et al.
transcriptase polymerase chain reaction (RT-PCR), gene 2001; Uzcategui et al. 2001; Twiddy and Holmes 2003),
regions amounting to 40% of the viral genome (4,016 bp can lead to conflicts in phylogenetic trees. We searched for
of an 11 kbp genome) and including both 59 and 39 ends potential recombinants across the entire phylogeny by
(see table 1 for primer sequences). Amplified regions testing for topological incongruity among Neighbor-
included all the structural genes (capsid: C; membrane: M; Joining (NJ) trees generated using a 500-base sliding
Selection in Emergent Dengue Virus 1653

window. The statistical support for recombination in these time (g). h was estimated from given sampling years (1994
sequences, as well as the locations of the breakpoints, was and 1998) using a coalescent method (program Fluctuate;
determined using a maximum likelihood method (program Kuhner, Yamato, and Felsenstein [1998]); the generation
LARD; Holmes, Worobey, and Rambaut 1999), and then time of dengue virus was taken as 14 days comprising
maximum likelihood (ML) trees were constructed on intrinsic (within human) and extrinsic (within mosquito)
either side of the breakpoints identified. replication times of 7 days duration each (Holmes, Bartley,
The evolutionary relationships among DEN-4 isolates and Garnett 1998). Although direct estimates of l are not
were inferred using a ML method (PAUP* package, available for dengue virus, a synonymous rate of 6.89 3
Swofford [2001]). In all cases trees were estimated using 104 substitutions/site/year was recently estimated for
the best fitting model of nucleotide substitution identified DEN-4 (Twiddy, Holmes, and Rambaut 2003). Given
by Modeltest 3.06 (Posada and Crandall 1998). The model a generation time of 14 days, this is equivalent to a l of
of DNA substitution that best described DEN-4 evolution 2.64 3 105 mutations per site, per generation. Putatively
in Puerto Rico (including the outgroup and six other positively selected amino acid changes were identified as
foreign samples) was the general time-reversible model those that fall on the internal branches of the tree that

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


that includes six substitution rate parameters (A$C ¼ separate sampling times (for example, on the branch
2.0346, A$G ¼ 12.5935, A$T ¼ 1.7144, C$G ¼ leading to the viral isolates sampled in 1998); at the
2.0608, C$T ¼ 31.0427, G$T ¼ 1), with 41.5% of sites population genetic level, mutations that are absent from an
variable and a gamma distribution of among-site rate early time-point yet present in all sequences from a later
variation (4 categories) with a shape parameter (a) of time-point can be assumed to have gone to fixation over
1.020 (substitution model GTR þ I þ ). Phylogenies were the course of the sampling period.
generated under successive rounds of tree-bisection/ Sequences generated by this study can be accessed
reconnection (TBR) branch swapping, updating parameter on GenBank according to accession numbers AY152036
estimates at each round. To assess the support for the through AY152363.
phylogenetic groupings observed we undertook a bootstrap
resampling analysis using 1,000 replicate Neighbor-
Results
Joining trees estimated under the ML substitution model
determined above. Trees were rooted with the 1981 isolate We examined over 4,000 nucleotides from each of
from Dominica, the oldest sequence available. 82 DEN-4 isolates collected in Puerto Rico and surround-
We used two methods to assess the extent of adaptive ing regions over a 20-year period. This included isolates
evolution in DEN-4. First, we examined the relative rates from (1) 1982, representing the first major outbreak of
of nonsynonymous (dN) and synonymous (dS) substitution DEN-4 in Puerto Rico; (2) the second major dengue
across coding portions of the viral genome. To do this, we epidemic on the island in 1986 to 1987, marked by
employed a ML approach to compare models of evolution hyperendemic transmission, and 29 DHF/DSS cases
that allow dN/dS to vary within genes or among lineages of including 3 deaths (Dietz et al. 1996); (3) two years—
the ML tree of the PR sequences (Yang et al. 2000). In 1992, 1994—during which DEN-4 occurred at relatively
particular, we compared models that allow for positive low prevalence; and (4) the most recent DEN-4 epidemic
selection because they incorporate a class of codons where in 1998 (fig 1). The 1998 epidemic marked the first time
dN/dS can be greater than 1 (models M2, M3, M8) with in 12 years that DEN-4 again dominated the epidemio-
those that specify neutral evolution because dN is logical landscape in PR (44% of all positively diagnosed
constrained to be less than dS (models M0, M1, and dengue cases) and was one of the most severe on the
M7). We also used the free ratio (FR) model that allows island (396,000–792,000 infections estimated and a record
each branch of the tree to have a different dN/dS ratio. 59 DHF cases reported, 2.5 standard deviations above the
Models were compared using standard likelihood ratio mean of 16.5; CDC data not shown). DEN-4 viruses
tests. A Bayesian approach was used to identify those circulating in the 1998 outbreak shared on average 98.5%
individual codons most likely subject to positive selection. sequence similarity with those from 1981/82. Over the
This approach calculates the posterior probabilities of dN/ entire study period, only 14% of all nucleotide sites
dS categories for each amino acid site so that sites with the experienced substitutions, of which 26% from translated
highest probabilities of falling into dN/dS category.1 are regions (or 3.6% of all coding sites) resulted in amino acid
most likely to have been under positive selection. All these substitutions.
analyses were undertaken using the CODEML program Our ML phylogenetic analysis of DEN-4 in Puerto
from the PAML package (Yang 1997). Rico revealed a pattern of evolution marked by strong
We also employed a population genetic approach to temporal clustering of isolates by year of sampling (fig. 2).
test for adaptive evolution in dengue virus. According to All early (1982) isolates from Puerto Rico were associated
standard theory, the average time to fixation of neutral with the 1981 isolate from Dominica, and were 0.7%
mutations in a haploid population is ;2Ne, generations. (range: 0.5% to 1%) different from the closest group of
Consequently, if mutations have been fixed much faster subsequent PR isolates from 1986/87. Three of six other
than this, we can conclude that their substitution dynamics foreign isolates shared ancestors with this early introduc-
are dominated by positive selection rather than drift. To tion group as opposed to later PR isolates (El Salvador
calculate 2Ne generations for DEN-4 in Puerto Rico, we 1993, Ecuador 1994, and Mexico 1995, data not shown),
estimated the parameter h (¼ 2Nel), the neutral mutation reflecting the widespread distribution of the introduced
rate per site per generation (l), and the viral generation Asian DEN-4 variant from 1981 (Gubler 1998; Foster et al.
1654 Bennett et al.

2003). Since the initial epidemic in 1982, DEN-4 was sites were identified (posterior probability P . 0.99) in the
virtually absent from Puerto Rico until the 1986 epidemic E, NS1, NS4B, and most notably the NS2A genes, where
(fig. 1; Dietz et al. 1996). All viruses sampled in Puerto a small class of codons (0.9%) had a mean dN/dS ratio of
Rico during and after this re-emergence (1986 onward) fell ;4.6, in no case could a model of codon evolution
into a single lineage defined by four silent nucleotide allowing positive selection conclusively reject all compet-
substitutions and one amino acid substitution in the ing neutral models (table 2; the results for all model
envelope (E) gene (methionine to threonine, aa position comparisons are available from the authors on request).
163; fig. 2). With the exception of a single 1994 isolate, However, the evolution of the nonstructural gene NS2A
two additional silent and two conservative amino acid was striking in that the branch leading to the 1998 cluster
substitutions (isoleucine to valine, envelope aa position of sequences was distinguished exclusively by three
351; lysine to arginine, NS1 aa position 51) occurred in the nonconservative amino acid replacements in NS2A (14Ileu
formation of the re-emergent PR lineage. Within this re- to Thr, 54Val to Thr, 101Pro to Ser; fig. 2), in the absence of
emergent lineage, sublineages were largely temporally any synonymous nucleotide changes. This results in an
ordered. For example, most of the 1987 isolates fell into infinitely large dN/dS ratio along this branch, suggestive of

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


a well-defined temporal cluster, distinguished by five silent positive selection: mean dN/dS for all other internal
changes across coding regions examined (gold in fig. 2). branches of our phylogeny in NS2A were significantly
Similarly, major temporal clusters were formed by all 1992 lower (mean dN/dS ¼ 0.038, P ¼ 0.001, using absolute
(green in fig. 2), most 1994 (blue in fig. 2), and all 1998 number nucleotide changes for observed and expected
isolates (red in fig. 2), respectively. The 1998 year group values). Moreover, these mutations appear to have been
was defined by several silent changes concentrated in the E fixed far more quickly than if they were subject to genetic
gene, and more notably three amino acid replacements in drift alone. Estimated values of h (2Nel) are 0.024 (range
the nonstructural NS2A protein (isoleucine to valine, aa 0.014 to 0.045) and 0.027 (range 0.016 to 0.052) for the
position 14; valine to threonine, aa position 54; and proline viruses sampled from years 1994 and 1998, respectively.
to serine, aa position 101). Assuming a neutral mutation rate of 2.64 3 105 mutations
Although DEN-4 isolates grouped into temporal per site, per generation, effective population sizes (Ne)
clusters, the dominant clade (that which included most of were only 454 and 511 for 1994 and 1998, respectively.
the isolates) from a particular year descended from older Taking the mean Ne value across these two sampling times
isolates that represented minor variants in the previous (482), we obtain an expected fixation time under genetic
sampling period. For example, the 1992 cluster did not drift of 13,496 days (482 3 14 days/generation 3 2) or
descend from the major 1987 cluster, but from contem- ;37 years. However, the observed fixation time for these
poraneous (1987) variants representing only 17% (3 out of mutants is a maximum of 6 years; as these mutations were
19) of the isolates sampled in 1987. Similarly, the 1998 first detected in 1994, we assume that they appeared
cluster descended from a rare 1994 lineage represented by sometime between the 1992 and 1994 epidemics, giving
only 8% (2 out of 27) of the isolates sampled between a maximum of 6 years time difference to the 1998 strains.
1992 and 1994. Indeed, only the major 1994 lineage was
nested within the dominant lineage of the previous
Discussion
sampling period, 1992. This pattern of sequence differ-
ences among DEN-4 isolates from Puerto Rico indicates This longitudinal phylogenetic study of DEN-4 in
phylogenetic shifts in the population of variants between a focal host population examined evolutionary changes
sampling periods. We refer to this numerical shift in the during viral epidemic expansion. Our most striking
population of genotypes present in a given year away from observation was that the evolutionary history of DEN-4
the dominant genotypes of an earlier time as ‘‘lineage in Puerto Rico was characterized by the replacement of
turnover,’’ since it infers a replacement of the most lineages between epidemic years: most isolates from
successful lineage from one sampling period to the next. a given year were closely related, but turnover of the
Finally, we found no evidence for major shifts in common variant occurred between sampling periods. This
topological position among gene regions indicative of pattern of lineage turnover is similar to that seen in some
recombination. other acute RNA viruses. For example, in coxsackie-A
To determine whether positive selection has played virus temporally organized lineages, regardless of geo-
a significant role in DEN-4 evolution and lineage turnover, graphic origin, are equally unrelated to each other (Santti
we examined rates of nonsynonymous (dN) and synony- et al. 2000; Ishiko et al. 2002) suggestive of lineage
mous (dS) substitution in individual viral genes using a ML turnover fueled by virus exchange between spatially
method. Although eight potentially positively selected distinct populations. Phylogenetic evidence also suggests

!
FIG. 2.—Maximum likelihood tree based on 3,543 bp sequences (coding regions) from 75 isolates of DEN-4 from Puerto Rico and one from
Dominica (the outgroup sequence). Six other foreign isolates have been omitted from the figure for simplicity. The same topology was obtained when
phylogenies were constructed including non-coding sequence data (4,016 bp per isolate). Branches are color-coded by year of sample isolation.
Bootstrap support values, shown at nodes, were generated by using 1,000 replicate Neighbor-Joining trees reconstructed under the best-fit model of
nucleotide evolution. Three amino acid changes, in envelope (E) and NS1 (N1) genes, that define the post-introduction Puerto Rican lineage, and three
amino acid changes in the positively selected NS2A (2A) gene that define the 1998 clade, are marked with black bars and the amino acid position within
their respective genes.
Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023
Selection in Emergent Dengue Virus 1655
1656 Bennett et al.

Table 2 artificially low if positive selection has purged genetic


Maximum Ratio of Nonsynonymous to Synonymous diversity. Consequently, adaptive evolution in the NS2A
Substitutions for Each DEN-4 Gene Region Examined in gene may have triggered the 1998 epidemic in Puerto Rico,
This Study and DEN-4 genotypes bearing these NS2A modifications
dN/dSa were also associated with contemporaneous epidemics
Gene Max. dN/dS b
Proportion of Codonsc Pd throughout the Greater and Lesser Antilles (Foster et al.
2003). Conversely, a similar association between amino
Capsid / membrane 0.822 0.167 0.997
Envelope / NS1 2.110 0.017 0.157 acid changes and lineage turnover was not observed
NS2A 4.574 0.009 0.725 between 1987 and 1992, where neither clade was defined
NS4B 1.851 0.014 0.937 by amino acid substitutions. In this case, lineage turnover
a
Values given for the M3 model of codon evolution that allows three classes
may have resulted from drift-sensitive population bottle-
of dN/dS per gene sequence alignment, all of which are estimated from the data. necks, inter-island extinction/recolonization, or selection
b
Highest dN/dS for a set of codons estimated under the M3 model. on other parts of the genome not examined in this study.
c
Proportion of codons with the maximum dN/dS value. Although we examined many more nucleotides than

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


d
Significance value obtained from a likelihood ratio test involving M3 and
previous studies (e.g., Rico-Hesse 1990; Lewis et al.
the neutral codon model M1 (which allows two classes of dN/dS, 0 and 1).
1993; Lanciotti et al. 1994; Lanciotti, Gubler, and Trent
1997; Rico-Hesse et al. 1997, 1998; Singh et al. 1999;
that vesicular stomatitis virus in the United States and Uzcategui et al. 2001; Twiddy et al. 2002), 60% of the
Mexico has experienced lineage shifts since the early dengue genome was not surveyed, including genes known
1980s, following its geographical spread in the Americas to be important in virus replication, such as NS5 (Leitmeyer
(Nichol, Rowe, and Fitch 1993). Finally, there is some et al. 1999), and virus antigenicity, such as NS1 (Mathew
evidence for lineage turnover in human influenza A virus, et al. 1998; Jacobs et al. 2000). Because selection is
although phylogenetic trees from this virus tend to have apparently restricted to very few sites, a complete appre-
a more regular temporal structure, most likely reflecting ciation of the forces driving genetic change in DEN-4 will
the continual selection pressure exerted by neutralizing ultimately require the analysis of full genome sequences.
antibodies (Bush et al. 1999). The apparent positive selection on the NS2A gene is
There are several non-mutually exclusive explana- even more anomalous given the relatively strong con-
tions for the lineage turnover observed in DEN-4 evolution straints acting on other regions of the viral genome. In
in Puerto Rico over the last 20 years, aside from incomplete particular, there was no convincing evidence that changes
sampling. Novel lineages could arise and proliferate in in structural genes, the primary targets of specific
a population through multiple re-introductions, genetic immunity, underlie the evolutionary shifts we observed
drift, and/or selection. However, although introductions in DEN-4 after its re-emergence in 1986. Most positions
from other DEN-4 populations may provide a source of within the structural genes were invariant (table 2), and
variation, evidence suggests that DEN-4 in the Caribbean is very few of the nonsynonymous substitutions in these
characterized by local evolution interrupted occasionally regions occurred at internal nodes. Two amino acid
by gene flow (Foster et al. 2003). There was also no changes in E (positions 163 and 351; see fig. 2) defined
evidence that microgeographic population structure within the DEN-4 that re-emerged in the late 1980s after 3 years
Puerto Rico generated the observed pattern, as virus of undetectable transmission, both occurring within well-
samples were obtained from similar geographic regions characterized structural epitope domains (summarized in
in all cases (data not shown). In addition, the distinct and Roehrig [1997]). The E protein, which enables host cell
persistent pattern of lineage turnover is difficult to explain binding and entry, providing a target for the host immune
by random sampling processes alone, because we would response (Roehrig 1997), is the functional analog of
expect common genotypes to become fixed by genetic drift influenza A’s hemagglutinin (HA) gene, which, in contrast,
more often than rare ones. Indeed, the stochastic nature of appears to be under strong antigenic selection (Bush et al.
the dengue virus life-cycle should favor common variants: 1999). In dengue virus, constraints on the E gene may be
genetic bottlenecks occur at every mosquito feeding event, attributable to its two-host life cycle and resultant multi-
along with seasonal reductions in vector populations cell type tropism (Beaty, Trent, and Roehrig 1988; Strauss
(Gubler 1987), and annual variation in the abundance of and Strauss 1988), so that rates of nucleotide substitution
susceptible human hosts. Instead, the dominant Puerto are lower than those seen in many other RNA viruses
Rican lineage of a given year twice descended from earlier (Weaver, Rico-Hesse, and Scott 1992; Jenkins et al. 2002).
rare genotypes, a pattern that suggests that much of the In addition, positive selection is less likely to occur
lineage turnover is driven by selection on viral genotype. In because of intrinsic negative fitness trade-offs (Woelk and
support of this hypothesis, there was an increase in the rate Holmes 2002). Indeed, substitution patterns across the four
of nonsynonymous substitution (in the absence of any gene regions examined here are consistent with a genome
silent changes in NS2A) on the lineage leading to the 1998 under stabilizing selection, with synonymous changes
epidemic, and these amino acid changes were fixed far greatly outnumbering nonsynonymous changes. Against
more quickly than expected by genetic drift. Moreover, our this conservative background, the amino acid changes in
population genetic estimations for the fixation time of the NS2A that distinguish the 1998 virus samples appear even
NS2A mutants are conservative in that these changes may more conspicuous, and natural selection on nonstructural
have been fixed much faster than the 6 years separating the genes has been described for other viruses and correlated
1992 and 1998 samples, and our estimates of Ne may be with epidemic outbreaks (Knowles et al. 2001).
Selection in Emergent Dengue Virus 1657

Aside from epidemiologic evidence, the phenotypic Sather, and I. Gomez. 1996. The 1986 dengue and dengue
traits targeted by natural selection involving NS2A are hemorrhagic fever epidemic in Puerto Rico: epidemiologic
unclear because we know so little about the gene’s and clinical observations. P. R. Health Sci. J. 15:201–210.
function. Dengue viruses in Puerto Rico may be under Domingo, E., and J. J. Holland. 1997. RNA virus mutations for
fitness and survival. Annu. Rev. Microbiol. 51:151–178.
particularly intense selection to improve replication rate,
Drake, J. W., and J. J. Holland. 1999. Mutation rates among
survival, and, ultimately, transmission rate, because the RNA viruses. Proc. Natl. Acad. Sci. USA 96:13910–13913.
only vector present, urban-specialist A. aegypti, is Falgout, B., and L. Markoff. 1995. Evidence that Flavivirus NS1-
relatively inefficient and requires high viral titers to NS2A cleavage is mediated by a membrane-bound host
acquire infection (up to 106 infectious units/ml blood in protease in the endoplasmic reticulum. J. Virol. 69:7232–
laboratory studies [Gubler 1987; Kuno 1997]). Puerto Rico 7243.
also lacks potential reservoir (primate) hosts, and its vector Foster, J. E., S. N. Bennett, H. Vaughan, V. Vorndam, W. O.
exhibits extremely low levels of vertical transmission, such McMillan, and C. V. F. Carrington. 2003. Molecular
that the disease must cycle directly between mosquitoes evolution and phylogeny of dengue type 4 virus in the
and humans to persist (Gubler 1987, 1998). Alternatively, Caribbean. Virology 306:126–134.

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


Gubler, D. J. 1987. Current research on dengue. Pp. 37–56 in
the selection pressure could relate to survival pressure
K. F. Harris, ed. Current topics in vector research, Vol. 3.
exerted by the human immune system in the guise of Springer-Verlag, New York.
cytotoxic T-lymphocytes (CTLs). Epitopes that elicit ———. 1998. Dengue and dengue hemorrhagic fever. Clin.
human T-cell responses ranging from serotype-specific to Microbiol. Rev. 11:480–496.
cross-reactive have been identified throughout the non- Gubler, D. J., D. Reed, L. Rosen, and J. R. Hitchcock, Jr. 1978.
structural regions of the dengue genome (Loke et al. Epidemiologic, clinical, and virologic observations on dengue
2001), and phylogenetic evidence for positive selection at in the Kingdom of Tonga. Am. J. Trop. Med. Hyg. 27:581–
or near T-cell epitopes has been noted previously (Twiddy, 589.
Woelk, and Holmes 2002). The function of NS2A has been Halstead, S. B. 1988. Pathogenesis of dengue: challenges to
associated with viral replication (Falgout and Markoff molecular biology. Science 239:476–481.
1995; Mackenzie et al. 1998) and the mediation of host Hatta, M., P. Gao, P. Halfmann, and Y. Kawaoka. 2001.
Molecular basis for high virulence of Hong Kong H5N1
immune interactions via NS1 (Rothman et al. 1993; influenza A viruses. Science 293:1840–1842.
Mathew et al. 1998; Jacobs et al. 2000). As the three Holmes, E. C. 1998. Molecular epidemiology of dengue virus—
amino acid substitutions in NS2A that define the 1998 the time for big science. Trop. Med. Int. Health 3:855–856.
cluster were all highly nonconservative changes from Holmes, E. C., L. M. Bartley, and G. P. Garnett. 1998. The
hydrophobic, non-polar residues to polar, uncharged emergence of dengue: past, present and future. Pp. 301–325
amino acids, they would at the very least change the 3- in R. M. Krause, ed. Emerging infections, Academic Press,
dimensional structure of the NS2A protein. To fully New York.
determine the repercussions of observed NS2A modifica- Holmes, E. C., and S. S. Burch. 2000. The causes and
tions on viral extended phenotype, future studies must consequences of genetic variation in dengue virus. Trends
endeavor to characterize the NS2A protein’s structure and Microbiol. 8:74–77.
Holmes, E. C., M. Worobey, and A. Rambaut. 1999. Phyloge-
function, and to survey this gene in phylogenetic studies of netic evidence for recombination in dengue virus. Mol. Biol.
epidemic dengue. Evol. 16:405–409.
Ishiko, H., Y. Shimada, M. Yonaha, O. Hashimoto, A. Hayashi,
Acknowledgments K. Sakae, and N. Takeda. 2002. Molecular diagnosis of
human enteroviruses by phylogeny-based classification by use
We thank M. Worobey for assistance with recombi- of the VP4 sequence. J. Infect. Dis. 185:744–754.
nation analyses, and J. J. Bull, K. A. Hanley, and D. D. Jacobs, M. G., P. J. Robinson, C. Bletchly, J. M. Mackenzie, and
Kapan for invaluable comments on the manuscript. Some P. R. Young. 2000. Dengue virus nonstructural protein 1 is
of the data were acquired in partial fulfillment of M.C.’s expressed in a glycosyl-phosphatidylinositol-linked form that
Master’s degree at the Department of Microbiology and is capable of signal transduction. FASEB J. 14:1603–1610.
Medical Zoology, University of Puerto Rico, and thanks Jenkins, G. M., A. Rambaut, O. G. Pybus, and E. C. Holmes.
are therefore due to her advisory committee. This research 2002. Rates of molecular evolution in RNA viruses:
a quantitative phylogenetic analysis. J. Mol. Evol. 54:156–
was supported by the National Institutes of Health (USA)
165.
through a research project grant and the Research Centers Knowles, N., P. Davies, T. Henry, V. O’Donnell, J. M. Pacheco,
in Minority Institutions program, and by The Royal and P. Mason. 2001. Emergence in Asia of foot and mouth
Society (UK). disease viruses with altered host range: characterization of
alteration in the 3A protein. J. Virol. 75:1551–1556.
Literature Cited Kuhner, M. K., J. Yamato, and J. Felsenstein. 1998. Maximum
likelihood estimation of population growth rates based on the
Beaty, B. J., D. W. Trent, and J. T. Roehrig. 1988. Virus coalescent. Genetics 149:429–434.
variation and evolution. Pp. 59–85 in T. P. Monath, ed. The Kuno, G. 1997. Factors influencing the transmission of dengue
arboviruses: epidemiology and ecology, Vol. 1. CRC Press, viruses. Pp. 61–88 in D. J. Gubler, and G. Kuno, eds. Dengue
Boca Raton, Fla. and dengue hemorrhagic fever. CAB International, New York.
Bush, R. M., C. A. Bender, K. Subbarao, N. J. Cox, and W. M. Lanciotti, R. S., D. J. Gubler, and D. W. Trent. 1997. Molecular
Fitch. 1999. Predicting the evolution of human influenza A. evolution and phylogeny of dengue-4 viruses. J. Gen. Virol.
Science 286:1921–1925. 78:2279–2286.
Dietz, V., D. J. Gubler, S. Ortiz, G. Kuno, A. Casta-Velez, G. E. Lanciotti, R. S., J. G. Lewis, D. J. Gubler, and D. W. Trent. 1994.
1658 Bennett et al.

Molecular evolution and epidemiology of dengue-3 viruses. Strauss, J. H., and E. G. Strauss. 1988. Evolution of RNA
J. Gen. Virol. 75:65–75. viruses. Annu. Rev. Microbiol. 42:657–683.
Leitmeyer, K. C., D. W. Vaughn, D. M. Watts, R. Salas, I. Swofford, D. L. 2001. PAUP*: phylogenetic analysis using
Villalobos de Chacon, C. Ramos, and R. Rico-Hesse. 1999. parsimony (*and other methods). Version 4. Sinauer Asso-
Dengue virus structural differences that correlate with ciates, Sunderland, Mass.
pathogenesis. J. Virol. 73:4738–4747. Thein, S., M. M. Aung, T. N. Shwe, M. Aye, Z. Aung, K. Aye,
Lewis, J. A., G-J. Chang, R. S. Lanciotti, R. M. Kinney, L. W. K. M. Aye, and J. Aaskov. 1997. Risk factors in dengue shock
Mayer, and D. W. Trent. 1993. Phylogenetic relationships of syndrome. Am. J. Trop. Med. Hyg. 56:566–572.
dengue-2 viruses. Virology 197:216–224. Tolou, H. J. G., P. Couissinier-Paris, J.-P. Durand, V. Mercier,
Loke, H., D. B. Bethell, C. X. T. Phuong, M. Dung, J. Schneider, J.-J. de Pina, P. de Micco, F. Billoir, R. N. Charrel, and
N. J. White, N. P. Day, J. Farrar, and A. V. S. Hill. 2001. X. de Lamballerie. 2001. Evidence for recombination in
Strong HLA class-I restricted T cell responses in dengue natural populations of dengue virus type 1 based on the
hemorrhagic fever: a double-edged sword. J. Infect. Dis. analysis of complete genome sequences. J. Gen. Virol.
184:1369–1373. 82:1283–1290.
Mackenzie, J. M., A. A. Kromykh, M. K. Jones, and E. G. Twiddy, S. S., J. F. Farrar, N. V. Chau, B. Wills, E. A. Gould, T.

Downloaded from https://academic.oup.com/mbe/article/20/10/1650/1164131 by guest on 07 June 2023


Westaway. 1998. Subcellular localization and some bio- Gritsun, G. Lloyd, and E. C. Holmes. 2002. Phylogenetic
chemical properties of the flavivirus Kunjin nonstructural relationships and differential selection pressures among
proteins NS2A and NS4A. Virology 245:203–215. genotypes of dengue-2 virus. Virology 298:63–72.
Mackow, E., Y. Makino, B. T. Zhao, Y. M. Zhang, L. Markoff, Twiddy, S. S., and E. C. Holmes. 2003. The extent of
A. Buckler-White, M. Guiler, R. Chanock, and C. J. Lai. homologous recombination in the genus Flavivirus. J. Gen.
1987. The nucleotide sequence of dengue type 4 virus: Virol. 84:429–440.
analysis of genes coding for nonstructural proteins. Virology Twiddy, S. S., E. C. Holmes, and A. Rambaut. 2003. Inferring
159:217–228. the rate and time-scale of dengue virus evolution. Mol. Biol.
Manzin, A., L. Solforosi, M. Debiaggi, F. Zara, E. Tanzi, L. Evol. 20:122–129.
Romano, A. R. Zanetti, and M. Clementi. 2000. Dominant Twiddy, S. S., C. H. Woelk, and E. C. Holmes. 2002.
role of host selective pressure in driving hepatitis C virus Phylogenetic evidence for adaptive evolution of dengue
evolution in perinatal infection. J. Virol. 74:4327–4334. viruses in nature. J. Gen. Virol. 83:1679–1689.
Mathew, A., I. Kurane, S. Green, H. A. F. Stephens, D. W. Uzcategui, N. Y., D. Camacho, G. Comach, E. C. Holmes and
Vaughn, S. Kalayanarooj, S. Suntayakorn, F. A. Ennis, and E. A. Gould. 2001. The molecular epidemiology of Dengue-2
virus in Venezuela: evidence for in situ viral evolution and
A. L. Rothman. 1998. Predominance of HLA-restricted CTL
recombination. J. Gen. Virol. 82:2945–2953.
responses to serotype crossreactive epitopes on nonstructural
Wang, E., H. Ni, X. Renling, A. D. T. Barrett, S. J. Watowich,
proteins after natural dengue virus infections. J. Virol.
D. J. Gubler, and S. C. Weaver. 2000. Evolutionary relation-
72:3999–4004.
ships of endemic/epidemic and sylvatic dengue viruses. J.
Nichol, S. T., J. E. Rowe, and W. M. Fitch. 1993. Punctuated
Virol. 74:3227–3234.
equilibrium and positive Darwinian evolution in vesicular
Weaver, S. C., R. Rico-Hesse, and T. W. Scott. 1992. Genetic
stomatitis virus. Proc. Natl. Acad. Sci. USA 90:10424– diversity and slow rates of evolution in new-world alpha-
10428. viruses. Curr. Top. Microbiol. Immunol. 176:99–117.
Posada, D., and K. A. Crandall. 1998. MODELTEST: testing the Woelk, C. H., and E. C. Holmes. 2002. Reduced positive
model of DNA substitution. Bioinformatics 14:817–818. selection in vector-borne RNA viruses. Mol. Biol. Evol.
Rico-Hesse, R. 1990. Molecular evolution and distribution of 19:2333–2336.
dengue viruses type 1 and 2 in nature. Virology 174:479–493. World Health Organization (WHO). 1999. Strengthening imple-
Rico-Hesse, R., L. M. Harrison, A. Nisalak, D. W. Vaughn, S. mentation of the global strategy for dengue fever/ dengue
Kalayanarooj, S. Greene, A. L. Rothman, and F. A. Ennis. haemorrhagic fever prevention and control: report of the
1998. Molecular evolution of Dengue type 2 virus in informal consultation, WHO, Geneva, 18–20 October 1999
Thailand. Am. J. Trop. Med. Hyg. 58:96–101. (WHO Report WHO/CDS/( DEN)/IC/2000. 1; www.who.int/
Rico-Hesse, R., L. M. Harrison, R. A. Salas, D. Tovar, A. emc-documents/dengue/whocdsdenic20001c.html).
Nisalak, C. Ramos, J. Boshell, M. T. de Mesa, R. M. Worobey, M., A. Rambaut, and E. C. Holmes. 1999. Widespread
Nogueira, and A. T. da Rosa. 1997. Origins of dengue type 2 intra-serotype recombination in natural populations of dengue
viruses associated with increased pathogenicity in the virus. Proc. Natl. Acad. Sci. USA 96:7352–7357.
Americas. Virology 230:244–251. Yang, Z. 1997. PAML, a program package for phylogenetic
Roehrig, J. T., 1997. Immunochemistry of dengue viruses. Pp. analysis by maximum likelihood. Comput. Appl. Biosci.
199–219 in D. J. Gubler, and G. Kuno, eds. Dengue and 13:555–556.
dengue hemorrhagic fever. CAB International, New York. Yang, Z., R. Nielsen, N. Goldman, and A.-M. K. Pedersen. 2000.
Rothman, A. L., and F. A. Ennis. 1999. Immunopathogenesis of Codon substitution models for heterogeneous selection
dengue hemorrhagic fever. Virology 257:1–6. pressure at amino acid sites. Genetics 155:431–449.
Rothman, A. L., I. Kurane, C. J. Lai, M. Bray, B. Falgout, R. Zanotto, P. M. de A., E. G. Kallas, R. F. de Souza, and E. C.
Men, and F. A. Ennis. 1993. Dengue virus protein recognition Holmes. 1999. Genealogical evidence for positive selection in
by virus-specific murine CD8þ cytotoxic T lymphocytes. the nef gene of HIV-1. Genetics 153:1077–1089.
J. Virol. 67:801–806. Zhao, B., E. Mackow, A. Buckler-White, L. Markoff, R. M.
Santti, J., H. Harvala, L. Kinnunen, and T. Hyypiä. 2000. Chanock, C. J. Lai, and Y. Makino. 1986. Cloning full-length
Molecular epidemiology and evolution of coxsackievirus A9. dengue type 4 viral DNA sequences: analysis of genes coding
J. Gen. Virol. 81:1361–1372. for structural proteins. Virology 155:77–88.
Singh, U. B., A. Maitra, S. Broor, A. Rai, S. T. Pasha, and P.
Seth. 1999. Partial nucleotide sequencing and molecular Keith Crandall, Associate Editor
evolution of epidemic causing dengue 2 strains. J. Infect. Dis.
180:959–965. Accepted May 25, 2003

You might also like