Advertisement

Global spread of Zika virus

Zika virus was identified in Uganda in 1947; since then, it has enveloped the tropics, causing disease of varying severity. Lessler et al. review the historical literature to remind us that Zika's neurotropism was observed in mice even before clinical case reports in Nigeria in 1953. What determines the clinical manifestations; how local conditions, vectors, genetics, and wild hosts affect transmission and geographical spread; what the best control strategy is; and how to develop effective drugs, vaccines, and diagnostics are all critical questions that are begging for data.
Science, this issue p. 663

Structured Abstract

BACKGROUND

First discovered in 1947, Zika virus (ZIKV) received little attention until a surge in microcephaly cases was reported after a 2015 outbreak in Brazil. The size of the outbreak and the severity of associated birth defects prompted the World Health Organization (WHO) to declare a Public Health Emergency of International Concern on 1 February 2016. In response, there has been an explosion in research and planning as the global health community has turned its attention to understanding and controlling ZIKV. Still, much of the information needed to evaluate the global health threat from ZIKV is lacking. The global threat posed by any emerging pathogen depends on its epidemiology, its clinical features, and our ability to implement effective control measures. Whether introductions of ZIKV result in epidemics depends on local ecology, population immunity, regional demographics, and, to no small degree, random chance. The same factors determine whether the virus will establish itself as an endemic disease. The burden of ZIKV spread on human health is mediated by its natural history and pathogenesis, particularly during pregnancy, and our ability to control the virus’s spread. In this Review, we examine the empirical evidence for a global threat from ZIKV through the lens of these processes, examining historic and current evidence, as well as parallel processes in closely related viruses.

ADVANCES

Because ZIKV was not recognized as an important disease in humans until recently, it was little studied before the recent crisis. Nevertheless, the limited data from the decades following its discovery provide important clues into ZIKV’s epidemiology and suggest that some populations were at risk for the virus for years in the mid-20th century, although this risk may predominantly have been the result of spillover infections from a sylvatic reservoir. Recent outbreaks on Yap Island (2007) and in French Polynesia (2014) provide the only previous observations of large epidemics and are the basis for the little that we do know about ZIKV’s acute symptoms (e.g., rash, fever, conjunctivitis, and arthralgia), the risk of birth defects, such as microcephaly (estimated to be 1 per 100 in French Polynesia), and the incidence of severe neurological outcomes (e.g., Guillain-Barré is estimated to occur in approximately 2 out of every 10,000 cases). The observation of an association between ZIKV and a surge in microcephaly cases in Brazil and the subsequent declaration of a Public Health Emergency of International Concern by the WHO have rapidly accelerated research into the virus. Small, but very important, studies have begun to identify the substantial risk the virus can pose throughout a pregnancy, and careful surveillance has established that ZIKV can be transmitted sexually. Numerous modeling studies have helped to estimate the potential range of ZIKV and measured its reproductive number R0 (estimates range from 1.4 to 6.6), a key measure of transmissibility in a number of settings. Still, it remains unclear whether the recent epidemic in the Americas is the result of fundamental changes in the virus or merely a chance event.

OUTLOOK

ZIKV research is progressing rapidly, and over the coming months and years our understanding of the virus will undoubtedly deepen considerably. Key questions about the virus’s range, its ability to persist, and its clinical severity will be answered as the current epidemic in the Americas runs its course. Moving forward, it is important that information on ZIKV be placed within the context of its effect on human health and that we remain cognizant of the structure of postinvasion epidemic dynamics as we respond to this emerging threat.
The effect of ZIKV is a function of the local transmission regime and viral pathogenesis.
(A) Many countries cannot maintain ongoing vector-mediated ZIKV transmission and are only at risk from importation by travelers and limited onward transmission (e.g., through sex). (B) If conditions are appropriate, importations can lead to postinvasion epidemics with high incidence across age ranges, after which the virus may go locally extinct or remain endemic. (C) There is evidence of ongoing ZIKV incidence in humans over years (e.g., a 1952 serosurvey in Nigeria), but it is unknown whether this is the result of ongoing circulation in humans or frequent spillover infections from a sylvatic cycle. (D) In other areas, ZIKV appears to have been maintained in animals with few human infections. (E) The majority of infections are asymptomatic, and severe outcomes, such as Guillain-Barré syndrome, are rare. (F) However, there is considerable risk of microcephaly and other fetal sequelae when infection occurs during pregnancy.

Abstract

First discovered in 1947, Zika virus (ZIKV) infection remained a little-known tropical disease until 2015, when its apparent association with a considerable increase in the incidence of microcephaly in Brazil raised alarms worldwide. There is limited information on the key factors that determine the extent of the global threat from ZIKV infection and resulting complications. Here, we review what is known about the epidemiology, natural history, and public health effects of ZIKV infection, the empirical basis for this knowledge, and the critical knowledge gaps that need to be filled.
Originally discovered in 1947, Zika virus (ZIKV) received little attention until a surge in microcephaly cases was reported after a 2015 outbreak in Brazil (1, 2). Prompted by the size of the outbreak and the severity of associated birth defects, the World Health Organization (WHO) declared ZIKV to be a Public Health Emergency of International Concern on 1 February 2016 (3). In response, there has been an explosion in research and planning as the global health community has turned its attention to understanding and controlling ZIKV. Still, much of the information needed to evaluate the global health threat from ZIKV remains unknown.
The global threat posed by any emerging pathogen depends on its epidemiology, its clinical features, and our ability to implement effective control measures (Fig. 1). In an interconnected world, introductions of ZIKV to areas free of the virus may be inevitable. Whether these introductions result in only a few subsequent cases or a major epidemic depends on the local ecology, population immunity, demographics of the region, and random chance. The ability of the virus to transmit in any area can be characterized by its reproductive number R: the number of people we expect to become infected from each case in that area (4). When R is greater than one, an epidemic can occur, and when it is less than one, onward transmission will be limited. When ZIKV successfully invades, the threat may be transient and the virus might become locally extinct, as appears to have been the case in Yap Island and French Polynesia (5, 6), or it may persist endemically, as seems to be the case in parts of Africa (7). There are two ways in which ZIKV can persist in a region: through ongoing transmission in animals (i.e., a sylvatic cycle) with occasional spillover into the human population, or through sustained transmission in humans (8, 9). Whichever scenario emerges, the natural history and pathogenesis of ZIKV will determine its effect on human health, with infection in pregnant women being particularly important (10). Finally, the extent of the global threat from ZIKV is mediated by our ability to control the virus and treat those cases that do occur.
Fig. 1 Factors determining the global risk from ZIKV.
(A) As long as ZIKV circulates anywhere, periodic introductions into ZIKV-free regions will occur. Whether these lead to an epidemic depends on the reproductive number R, a measure of transmission efficiency determined by local ecology and population susceptibility to ZIKV. (B) When R > 1, introductions can result in major epidemics, after which the virus may go locally extinct or become endemic. (C) ZIKV could be maintained endemically either in local nonhuman primates (the sylvatic cycle) or through ongoing human transmission. (D) Most ZIKV infections (75 to 80%) are asymptomatic, and those with symptoms are likely at highest risk for rare neurological complications (6, 63, 92), particularly Guillain-Barré (45). Adverse fetal outcomes, notably microcephaly, may also be more common when the mother is symptomatic. Owing to its association with pregnancy, ZIKV’s health effects depend on the fertility rate and the age distribution of infections. The age distribution mirrors the general population in ZIKV-free (A) and epidemic (B) settings but is a function of the force of infection in endemic settings (C) (4, 45). Appropriate control measures can reduce R, decreasing the probability of successful ZIKV invasion (A) and its subsequent effect [(B) and (C)] [see (116)].
In this review, we examine the empirical evidence for a global threat from ZIKV through the lens of these processes. We review what is known about the natural history and pathogenesis of ZIKV in humans, outline what we know about the ability of ZIKV and similar viruses to invade and persist in diverse settings, and summarize the challenges we face in studying and controlling ZIKV. Finally, we examine what we know about why ZIKV has emerged as a public health threat in the Americas after being known for decades as a rare and mild tropical disease.

A brief history of ZIKV

ZIKV was discovered in the blood of a rhesus monkey in 1947 at the Yellow Fever Research Institute in Entebbe, Uganda (1), and was isolated from Aedes africanus mosquitoes the following year (1). Soon after, multiple serosurveys found evidence of antibodies to ZIKV in human populations throughout Africa (1114), India (15), and Southeast Asia (16, 17) (Fig. 2). It was not initially clear that ZIKV caused clinical disease (13), although early evidence suggested that it was neurotropic in mice (18). Human infection was first confirmed in 1953 in Nigeria (13), and ZIKV was definitively established as pathogenic in humans after later experimental (19) and natural (20) infections led to symptoms of fever and rash.
Fig. 2 Current and potential distribution of ZIKV.
(A) Spread of ZIKV across the globe to date. Countries are colored by the timing of the first indication of local ZIKV transmission by serologic evidence or confirmation of human cases. Solid shading indicates clusters of confirmed cases or seropositivity to ZIKV of >10% in some subpopulations, whereas hatched colors indicate 5 to 10% seropositivity (serosurveys showing <5% seropositivity are not shown). Symbols indicate locations and timings of viral isolations from mosquitoes (triangles) and humans (circles). (B) Map of the global occurrence of the widely distributed ZIKV vectors A. aegypti and A. albopictus. Adapted from (100). (C) Map of the occurrence of dengue, a closely related Aedes-transmitted flavivirus. Adapted from (103). Shaded regions correspond to areas with predicted probability of vector or dengue occurrence of >30%. *Somalia did not report the total percentage of those who were ZIKV seropositive, but there was a small percentage of subjects seropositive to ZIKV and no other flavivirus and a large percentage seropositive to two or more flaviviruses, so Somalia’s data are included.
The globally distributed mosquito A. aegypti was identified as a likely vector for ZIKV transmission in the 1950s after successful transmission of the virus to a mosquito from an infected human volunteer (19). Later experiments confirmed A. aegypti’s ability to transmit ZIKV to mice (21), and ZIKV has since been isolated from several Aedes species (and, in a few cases, other genera) (22), including A. albopictus (2326).
In the decades after its discovery, intermittent serosurveys continued to find evidence of ZIKV infection in humans in Africa (2729), the Indian subcontinent (30), and Southeast Asia (16, 31, 32). Evidence for ZIKV’s continued presence was further bolstered by limited viral isolations from mosquitoes (3338), humans (7, 20, 29, 39, 40), and nonhuman primates (9). However, few clinical cases had been reported in humans before 2007 (20, 29, 31, 40), and ZIKV was considered to be of limited public health importance.
In 2007, the first known major outbreak of ZIKV occurred on Yap Island in the Federated States of Micronesia (6). Although several patients initially tested positive for dengue, the unusual clinical presentation prompted physicians to send serum to the Centers for Disease Control and Prevention (CDC) Arbovirus Diagnostic and Reference Laboratory, where it tested positive for ZIKV (6, 41). During the outbreak, ~73% of the island’s residents were infected with ZIKV, and symptoms were generally mild and short-lived (6).
After the Yap Island outbreak, there were sporadic isolations of ZIKV in residents of and travelers to Southeast Asia (4244), but no other major ZIKV outbreaks were observed until late 2013. From October 2013 to April 2014, French Polynesia experienced a large outbreak of ZIKV, estimated to have infected 66% of the general population (5, 45). A contemporaneous surge in the number of cases of Guillain-Barré syndrome raised concerns of an association with ZIKV (5, 45): A total of 42 cases of Guillain-Barré syndrome were reported from November 2013 to February 2014, compared with three cases in all of 2012. These are the first known instances of neurologic sequelae associated with ZIKV infection. Although not noted at the time, retrospective analyses suggest that there may also have been an increase in microcephaly cases (46). After the French Polynesia outbreak, ZIKV spread throughout the South Pacific, including outbreaks in New Caledonia, the Cook Islands, and Easter Island in 2014 (47).
The earliest confirmed cases of ZIKV infection in the Americas occurred in late 2014 in northeastern Brazil (48). Recent work suggests that the virus may also have been present simultaneously in Haiti (49). Over the following months, the virus spread rapidly throughout Brazil (50), followed by a substantial rise in cases of Guillain-Barré syndrome and microcephaly in affected regions (51), prompting the WHO to declare a Public Health Emergency of International Concern on 1 February 2016 (3). Phylogenetic evidence suggests that the strains that seeded this outbreak are descendants of those that caused outbreaks in the South Pacific, which in turn descended from the Asian lineage of the virus (52).
Since late 2014, ZIKV has spread widely throughout South and Central America and the Caribbean (2). As of June 2016, more than 35 countries throughout the Americas have reported locally circulating ZIKV (53). This includes a large outbreak in Colombia, with more than 65,000 reported cases, numerous reports of potentially associated neurological syndromes, and ZIKV-associated microcephaly cases (5456). As of June 2016, the ZIKV situation continues to evolve, and the global threat ultimately posed by ZIKV remains uncertain.

The natural history and pathogenesis of ZIKV

Transmission and natural history of ZIKV

The primary source of ZIKV infection in humans is from bites of infected mosquitoes (57), although there have also been cases of sexual (5860), perinatal (61), and suspected blood-transfusion transmission (62). Evidence from outbreaks in the South Pacific indicates that a minority of those infected with ZIKV develop clinical illness: During the Yap Island outbreak, 19% of people with serological evidence of recent infection [immunoglobulin M (IgM)–positive] reported ZIKV symptoms (6); in French Polynesia, 26% of ZIKV-positive blood donors who were asymptomatic at the time of donation later reported symptoms (63).
On average, those who do develop ZIKV symptoms will do so 6 days after infection (64), and 95% will do so within 11 days (Fig. 3). Virus has been isolated from blood (13), urine (65, 66), saliva (67), semen (68), amniotic fluid (69), and neurologic tissue (70). Virus can be isolated in blood for an average of 10 days after infection (99% will clear the virus by 24 days) (64), and case reports indicate that virus may remain in urine for 12 or more days after infection (65) and in semen for more than 60 days (59). Pregnancy may affect the length of viral shedding: In one case, a woman remained viremic for at least 10 weeks during pregnancy but cleared the virus within 11 days of termination (71). Antibodies to ZIKV become detectable, on average, 9 days after infection (64). Although the duration of immunity against ZIKV remains unknown, evidence from other flaviviruses suggests that it should be lifelong (72). Mosquitoes become infectious about 10 days after biting an infectious human and likely remain so until death (19).
Fig. 3 Schematic of the course of human and mosquito infection.
Symptoms develop, on average, 6 days (95% range, 3 to 11 days) after ZIKV infection (64). Approximately 9 days (95% range, 4 to 14 days) after infection, antibodies start increasing: The first antibodies detectable will be IgM, which will later decline as IgG antibodies increase, then persist indefinitely (the timing of the IgM/IgG switch is for illustrative purposes only and is not meant to indicate the actual length of IgM persistence). Viremia likely starts to increase before symptoms appear, and the magnitude and length of viremia will shape the risk of infection of susceptible mosquitoes that bite this host. After an incubation period, this infected mosquito will be able to transmit infection to susceptible humans (19). The interval from the initial to the subsequent human infection is the generation time of ZIKV, Tg [for estimates, see (116)].
Unfortunately, many of these distributions are estimated based on fewer than 30 cases. Expansion of this pool of evidence is critical for accurate assessment of surveillance activities and modeling of ZIKV risk.

Clinical illness

ZIKV symptoms are typically nonspecific and mild. Consistent with other reports (73), symptoms reported from 31 confirmed cases on Yap Island included maculopapular rash (90%), subjective fever (65%), arthralgia or arthritis (65%), nonpurulent conjunctivitis (55%), myalgia (48%), headache (45%), retro-orbital pain (39%), edema (19%), and vomiting (10%) (6, 10). Case reports suggest that acute symptoms of ZIKV will typically fully resolve within 1 to 2 weeks of onset (44, 60, 7480). Deaths are rare and have primarily occurred in patients with preexisting comorbidities or who are immunocompromised (81, 82).
Persons infected with ZIKV may be at increased risk for severe neurologic sequelae, notably Guillain-Barré syndrome. Data from French Polynesia suggest a risk of Guillain-Barré of 24 per 100,000 ZIKV infections (5, 45), more than 10 times the annual rate in the United States (1.8 per 100,000) (83). Regardless of cause, Guillain-Barré is associated with considerable morbidity and 3 to 10% mortality (84). Guillain-Barré may be more common in symptomatic ZIKV cases; during the French Polynesia outbreak, 88% of Guillain-Barré cases reported symptoms a median of 6 days before Guillain-Barré onset (5, 45). There have been reports of other neurological sequelae, including meningoencephalitis (85) and acute myelitis (86), although no causal link has been established.

ZIKV in pregnancy

Much of the concern surrounding ZIKV has focused on the link between infection in pregnancy and fetal microcephaly. As of 7 May 2016, 7438 suspected microcephaly cases have been reported in Brazil since ZIKV’s emergence (1326 confirmed out of 4005 investigated), versus fewer than 200 per year before the outbreak (87, 88). Quantifying the risk of microcephaly has been complicated by uncertainty in the number of ZIKV-affected pregnancies, owing to the large fraction of cases that are asymptomatic, a lack of consensus on the definition of microcephaly, and other infectious causes of microcephaly, such as cytomegalovirus and rubella (89). However, in light of multiple epidemiologic studies and the isolation of ZIKV in amniotic fluid and fetal brain tissue, the CDC confirmed a causal link between ZIKV infection during pregnancy and severe birth defects, including microcephaly in April 2016 (90). This conclusion is further supported by the presence of microcephaly and other brain abnormalities in the pups of mice experimentally infected with ZIKV (91).
ZIKV symptoms in pregnant women are similar to the general population (92), but it is unknown if immunosuppression during pregnancy changes the rate at which they occur. Among those who are symptomatic, adverse fetal outcomes appear to be frequent, occurring in 29% (12 out of 42) of symptomatic ZIKV-infected pregnant women in a prospective study in Brazil (92). A second Brazilian study found that 74% (26 out of 35) of mothers of infants with microcephaly reported a rash in the first or second trimester (51). The rate of birth defects in asymptomatic pregnant women is likely lower, but not zero. For example, a Colombian study identified four microcephaly cases with virologic evidence of ZIKV infection, all of which were born to women who did not report symptoms of ZIKV (54). Modeling studies suggest that the overall risk of ZIKV-associated microcephaly in the first trimester is around 1 per 100, regardless of symptoms, and low to negligible thereafter (46, 93).
Although microcephaly was the first fetal abnormality to be recognized, there is increasing evidence that ZIKV may be responsible for other fetal sequelae, such as intracranial calcifications, ventriculomegaly, ocular impairment, brainstem hypoplasia, intrauterine growth restriction (IUGR), and fetal demise (92, 94). Placental pathology has also been reported. Although microcephaly is detectable at birth, other findings may require additional, less routine procedures, such as imaging or autopsy, and thus may be underreported. Brasil et al. found that only one in four fetuses with abnormalities in ZIKV-infected women met the criteria for microcephaly (92), indicating that the total number of ZIKV-affected pregnancies may be four times the number of reported microcephaly cases.
Beyond an association with symptoms, it is unclear what factors increase the risk of adverse pregnancy outcomes after maternal ZIKV infection. For other infections that cause fetal abnormalities, risk is often associated with gestational age at infection. For instance, the risk of birth defects from cytomegalovirus and rubella is highest if infection occurs in the first or early in the second trimester (89). Epidemiologic evidence suggests a similar association with first-trimester ZIKV infection (46, 95). In a prospective study of 88 women, microcephaly and brain abnormalities occurred only in first- and second-trimester infections (92). However, 8 of 12 cases of fetal abnormalities overall occurred in second- and third-trimester infections, and women infected as late as 35 weeks experienced fetal death, IUGR, or anhydramnios [although these outcomes commonly occur in the absence of ZIKV; e.g., in Brazil, 11 fetal deaths occur per 1000 births (96), and IUGR rates range from 5 to 7% in developed countries (97)]. A recent Colombian study suggests little to no risk from infection in the third trimester; among 616 Colombian women with clinical symptoms of ZIKV during the third trimester, none gave birth to infants with microcephaly or other brain abnormalities (7% were still pregnant at the time of reporting) (54).
Adverse outcomes in pregnancy are the most worrisome side effects of ZIKV infection, and research into this association is progressing rapidly. Still, much remains to be learned, particularly about the frequency and spectrum of ZIKV sequelae in pregnancy and how we can assess and reduce risk. ZIKV-related birth defects can have long-standing financial, social, and health effects on affected families and communities (98). Hence, the threat from ZIKV cannot purely be assessed based on immediate clinical outcomes but also must account for its lifelong effects.

The potential range and effect of ZIKV

Transmissibility and potential range of ZIKV

Transmission of ZIKV in a population is a function of local ecology, the natural history of ZIKV, and the population’s susceptibility to infection. The suitability of the local environment for ZIKV transmission and the effect of ZIKV’s natural history are captured by the basic reproductive number R0, the number of secondary infections expected from a single case in a population with no preexisting immunity (e.g., French Polynesia before 2013). R0 is a function of both disease and setting and will vary between locales based on the local environment, human behavior, vector abundance, and, potentially, interactions with other viruses. The combined effect of these factors and susceptibility will be captured by the reproductive number R, which is related to R0 by the equation R = R0 × S, where S is the proportion of the population susceptible to ZIKV. This value, combined with the generation time (the time separating two consecutive infections in a chain of transmission), tells us the speed at which ZIKV will spread in a population. As we consider how to assess the range and effects of ZIKV, we rely both on previous experience with ZIKV and related viruses and on an assessment of factors likely to influence R and R0.
The size of an outbreak after an introduction will depend on R (R0 in a ZIKV-naive population) (99), with small, self-limiting outbreaks becoming more likely as R approaches one, and increasing epidemics with larger Rs. Hence, ZIKV can successfully spread to a new region if R > 1, which requires, among other factors, sufficient density of the vector population. ZIKV has been isolated from multiple Aedes genus mosquitoes (2326, 38), including A. albopictus and A. aegypti, which have a large global range (Fig. 2B) (100). Although ZIKV has been occasionally isolated from or experimentally passed to other genera, including Culex species, there is no current evidence that they contribute substantially to its spread (22, 23, 101). It is unclear whether all areas across the range of these mosquitoes are at risk for ZIKV epidemics. Dengue, a virus that is also transmitted by Aedes mosquitoes, has caused epidemics throughout the Americas (Fig. 2C) but has not achieved sustained transmission in the continental United States, despite widespread vector presence (100, 102, 103). The reasons for this may include not only climate but also differences in built environments and social factors (104), all of which are likely to affect ZIKV transmission.
Several groups have attempted to map ZIKV’s potential global range based on currently available data. These maps have been constructed around combinations of environmental, vector abundance, and socioeconomic factors (105109). There is wide agreement that much of the world’s tropical and subtropical regions are at risk for ZIKV spread, including major portions of the Americas, Africa, Southeast Asia, and the Indian subcontinent, as well as many Pacific islands and Northern Australia. These maps differ notably in the extent of risk projected in the southeastern United States and inland areas of South America and Africa, with Carlson and colleagues suggesting a more limited range (107), particularly in the continental United States, than Messina et al. and Samy et al. (108, 109). These maps are important attempts to refine estimates of ZIKV’s global range beyond those based solely on the distribution of dengue or Aedes mosquitoes but, as noted by the authors, are based on limited evidence and should be refined as we learn more about ZIKV. These analyses are, arguably, best interpreted as an assessment of the risk of initial postinvasion ZIKV epidemics, not its long-term persistence. Whether ZIKV will in fact spread throughout these areas is uncertain; similar viruses have failed to spread to or take hold in areas theoretically at risk (e.g., yellow fever in Southeast Asia) (110).
R0 in ZIKV outbreaks in Yap Island and French Polynesia was estimated to be between 1.8 and 5.8 (111113), corresponding to 73.2 to 99.9% of the at-risk population becoming infected in an uncontrolled outbreak, based on classic epidemic theory (4) [although the true relationship between R0 and final attack rates for ZIKV will be somewhat more complex (99)]. Serosurveys in French Polynesia suggest that 66% of the population was infected (46), which is somewhat lower but not inconsistent with these projections. Preliminary estimates of R0 from Colombia vary by location and range from 1.4 to 6.6 (114, 115). These are similar to R0 estimates presented by Ferguson et al. for 13 countries in the Americas (116) and recent estimates of R0 for Rio de Janeiro (117). These values are consistent with R0 estimates for dengue in similar settings. Of note, all of these are from settings with recently observed endogenous transmission of ZIKV, and R0 will vary widely across settings and is likely to be far lower near the limits and outside of ZIKV’s range.

ZIKV’s potential for endemic circulation

After the initial, postinvasion epidemic of ZIKV, the virus may either go extinct locally or be maintained through endemic human spread or sylvatic transmission (Fig. 1). Early age-stratified serosurveys in Africa and Asia offer some insight into past transmission patterns of ZIKV in these regions and ZIKV’s past dynamics (Fig. 4). Serosurveys in Nigeria, the Central African Republic, and Malaysia are consistent with ongoing ZIKV transmission, common spillover infections from a sylvatic reservoir, or frequent reintroductions from other regions over multiple decades (13, 16, 118). However, these results must be interpreted with caution owing to cross-reactivity with other flaviviruses in serologic tests (22). Up-to-date, age-stratified serosurveys, broadly covering regions where ZIKV has previously been detected, would tell us much about the virus’s ability to persist.
Fig. 4 Age-stratified serosurveys provide important clues to local ZIKV epidemiology.
Results must be interpreted with caution because of the possibility of cross-reactivity with other flavivirus antibodies. (A to C) Ongoing ZIKV transmission, whether from endemic human transmission or a constant risk of zoonotic infection, manifests as a smooth increase in the proportion of the population seropositive with increasing age. This pattern is also consistent with frequent reintroductions leading to periodic outbreaks. If we assume that the risk of ZIKV infection is constant over a lifetime, we can estimate the force of infection (FOI), the proportion of the susceptible population infected each year. Serosurvey results consistent with ongoing transmission include (A) Uburu, Nigeria, 1952 (13); (B) Central African Republic, 1979 (pink, female; cyan, male) (118); and (C) Malaysia, 1953 to 1954 (16). Blue dashed lines and text represent the expected distribution from the estimated FOI. (D and E) In areas without substantial ZIKV transmission, there will be very low levels of seropositivity across age groups and no clear age pattern. Some individuals may still be seropositive due to cross-reactivity in serological assays, infection of travelers, and limited imported cases. Examples include (D) Central Nyanza, Kenya, 1966 to 1968 (121) and (E) Mid-Western Region, Nigeria, 1966 to 1967 (120). (F) Substantial shifts in seropositivity between age groups inconsistent with ongoing transmission suggest past epidemics—e.g., results from a 1966 to 1968 serosurvey in the Malindi district of Kenya are consistent with one or more epidemics of ZIKV occurring 15 to 30 years previously (121). Similar patterns could also occur due to differences in infection risk by age or a sharp reduction in transmission intensity at some point in the past.
More recent evidence of sustained transmission comes from Thailand, where seven samples collected in independent outbreak investigations tested positive for ZIKV infection (43). The broad geographic spread of these cases is consistent with endemic transmission throughout Thailand. Furthermore, occasional but consistent serologic and virologic evidence of ZIKV transmission in humans and mosquitoes from across Africa, India, and Southeast Asia spanning more than 60 years suggests that ZIKV has been persistently present throughout these regions (22) (Fig. 1A). Phylogenetic evidence further supports this supposition, because the African and Asian lineages divided in the 1940s and remain distinct up until the present day (22, 26) (Fig. 5).
Fig. 5 ZIKV phylogenetics.
(A) Maximum likelihood tree of phylogenetic relationships between 43 flaviviruses (numbers indicate support from 1000 ultrafast bootstrap replicates), with antigenic clusters from Calisher et al. indicated by color (162). (B) The phylogenetic relationship between ZIKV strains isolated from throughout the globe. Whole-genome nucleotide sequences were aligned using Clustal Omega (163), and trees were constructed using IQ-TREE (164) under a GTR+G+I evolutionary model.
The evidence supports ZIKV’s ability to persist regionally, but it is unclear whether the human population alone can maintain ZIKV endemically. After an initial postinvasion epidemic, the time until there is a risk of additional epidemics will be driven by the replenishment of susceptibles through births and waning immunity [the latter seems unlikely based on evidence that other flaviviruses provide lifelong immunity to the infecting strain (22)]. For ZIKV to persist in the human population over this period, the population must be large enough to support low levels of transmission between epidemics (4).
However, all countries with evidence of persistent ZIKV transmission have a plausible sylvatic cycle. Patterns of ZIKV isolations in a study of samples from multiple hosts in Senegal spanning 50 years support episodic transmission across species (9); phylogenetic evidence indicates ZIKV passes frequently between nonhuman primates and humans in Africa (26); and numerous studies in Africa and Asia show serologic evidence for ZIKV infection in nonhuman primates (1, 18, 22, 33, 119). Some areas, where there has been serological evidence of long periods of consistent risk of ZIKV infection, are near areas where serological evidence suggests that human populations are largely ZIKV free (e.g., Nigeria versus Kenya) (120, 121)—a pattern more consistent with spillover infections from a sylvatic reservoir than of endemic transmission in humans.
In light of this evidence, it is plausible that the persistence of ZIKV in Africa and Asia may depend on the presence of a sustainable sylvatic cycle. However, it is unclear if the primate population in the Americas could support sylvatic transmission (122) or if such a cycle is necessary for ZIKV to remain endemic. Nonhuman primates are present throughout South and Central America, and ZIKV has recently been isolated from two species in the Ceará State of Brazil (123), suggesting at least the possibility for sustained sylvatic transmission in the region. Further characterization of ZIKV ecology in Asia and Africa and monitoring of the developing situation in the Americas is needed to assess the long-term risk from ZIKV in newly affected regions.
Because the most severe outcomes of ZIKV infection are associated with pregnancy, the risk from endemic ZIKV will depend on the age distribution of those infected. Serosurveys indicating ongoing ZIKV circulation (Fig. 4, A to C) support average ages of infection of 17 (Nigeria, 1952), 29 (Central African Republic, 1979) and 30 years (Malaysia 1953 to 1954) (13, 16, 118). Likewise, R0 estimates from the literature are consistent with average ages of infection ranging between 10 and 38 years in the setting of endemic human-to-human transmission (although human-to-human transmission should not be necessarily assumed in the settings covered in Fig. 4, A to C). These ages suggest that in endemic settings, risk of ZIKV infection may be considerable during childbearing years. Importantly, this information could potentially be used to estimate the expected rate of microcephaly and other birth defects in regions where ZIKV becomes endemic.

Why has ZIKV invaded the Americas now?

Little is known about ZIKV’s introduction into the Americas. Phylogenetic analyses indicate that a virus descended from the French Polynesian ZIKV strain entered Brazil between May and December 2013 (52). Although there has been speculation about introduction during specific sporting events (52, 124), Brazil has more than 6 million visitors per year, providing numerous opportunities for ZIKV introduction. Regardless of how and when ZIKV entered the Americas, the reasons for the size and severity of this outbreak are unclear.
The unprecedented size and effect of the ZIKV epidemic in the Americas may be the natural result of a random introduction into a large population without preexisting immunity. Like the Americas, the populations of Yap Island and French Polynesia were fully susceptible when ZIKV was introduced, and both had large outbreaks infecting more than 65% of their populations (6, 45). However, on these small islands the absolute number of adverse outcomes may have been too low to be noticed initially. Likewise, it is possible that small ZIKV epidemics, and even invasion into Southeast Asia in the mid-1900s, resulted in effects that were unnoticed against the backdrop of other infectious diseases, particularly because small population sizes (compared to Brazil) mean that excess microcephaly cases would likely be in the hundreds (or less) in any given country. Endemic transmission would be even less likely to be noticed, because yearly attack rates would be a tenth again lower (Fig. 4) (116). Still, given the magnitude and severity of the outbreak in the Americas, it seems implausible that, if such outbreaks were occurring, none were observed for over 60 years. Hypothesized changes in the biological and ecological drivers of ZIKV transmission must be carefully assessed, because they will influence how we quantify the risk from ZIKV globally.
Warmer temperatures and rainfall resulting from the 2015–2016 El Niño may have facilitated ZIKV transmission throughout the region (125) and increased the geographic range of Aedes mosquitoes. Warmer temperatures have been associated with more efficient transmission of related flaviviruses (126) and greater production of adult mosquitoes (127, 128). El Niño–associated periods of flooding (which increases mosquito breeding sites) and of droughts (which can increase human-mosquito interactions) may facilitate ZIKV transmission (129, 130). However, it should not be assumed that increased temperature or rainfall will universally promote ZIKV transmission, because climatic changes have complex repercussions across food webs (from plant growth to bird behavior) and the thermal effects on the virus itself are likely to be nonlinear (131). Over a longer time scale, development and urbanization has led to a proliferation of A. aegypti and A. albopictus in densely populated areas, which may have facilitated the rise of dengue in the region and may also have provided conditions that favored ZIKV spread (132).
There is some possibility that immunological interactions with other flaviviruses may be facilitating the spread or pathogenesis of ZIKV in the Americas. In dengue, preexisting antibodies to one serotype are hypothesized to enhance subsequent infections with another serotype through a mechanism known as antibody-dependent enhancement (ADE) (133). ADE may result in increased susceptibility to infection, the likelihood of developing severe disease, and the chances of transmission (134, 135). Evidence from some in vitro experiments and epidemiological studies show both protective and enhancing effects between immunity to Japanese encephalitis and dengue (136, 137), and several in vitro studies have shown enhancement of ZIKV replication in the presence of antibodies to other flaviviruses (138, 139). Dengue has circulated throughout much of Central and South America since it reemerged 30 years ago; hence, it is possible that such interactions are contributing to the current outbreak of severe disease. However, this would raise questions as to why similar interactions have not been seen in the dengue endemic regions of Southeast Asia that also show evidence of ZIKV circulation. Studies that measure preexisting dengue and ZIKV antibodies and track clinical outcomes may help illuminate the issue.
The severity of outcomes in recent outbreaks, compared with past observations of mild disease, has led some to hypothesize that the virus has mutated to be more pathogenic (140). Recent evidence suggests distinct codon preferences between African and Asian ZIKV lineages, although adaptive genetic changes may have an effect on viral replication and titers (141), whereas the genetic diversity of viruses isolated in ZIKV-associated microcephaly cases suggest that recent mutations may not be involved (142). Epidemiologic and laboratory studies are needed to determine whether these changes have had a substantive effect on viral pathogenesis. Until the effect of ZIKV evolution is better understood, we should be careful to balance the need to learn from previous research with the possibility that the virus has fundamentally changed.
Human genetics is known to have a profound effect on the pathogenesis of many infectious diseases (143), and there is some indication that the same could be true for flaviviruses (144, 145). While there is evidence of ancient intermixing between Polynesian and American populations (146), there are no indications of a link between ancestry and severe outcomes from ZIKV at this point. Likewise, genetic variation in A. aegypti is known to affect vector competence to transmit flaviviruses (147); hence, it is possible that changes in the makeup of the vector population also influence ZIKV transmission and account for regional differences in ZIKV effects.

Challenges and research priorities for responding to the ZIKV threat

Surveillance and clinical outcomes

The key challenge in ZIKV surveillance is the proportion of cases that remain asymptomatic and the nonspecificity of ZIKV symptoms (148). Dengue and chikungunya are also transmitted by Aedes mosquitoes, cocirculate with ZIKV, and can have a similar presentation, further complicating surveillance efforts.
Laboratory testing is needed to confirm ZIKV infection. Molecular (reverse transcription polymerase chain reaction) techniques can be used to detect ZIKV in serum, saliva, and urine (67, 149). However, there are frequent cases in which testing of different fluids gives discrepant results, and additional studies are needed to assess diagnostic accuracy (67). The timing of sample collection is crucial; viral RNA is only detectable in serum for 3 to 5 days after symptom onset (~10 days after infection) but may persist longer in other fluids (59, 64, 66).
A highly specific, easily administered antibody test would be a boon to surveillance and patient care. Such a test could be used to estimate underlying ZIKV incidence and thus rates of severe outcomes, confirm infection in studies of ZIKV pathogenesis, and test for immunity to ZIKV early in pregnancy so women can know whether they are at risk. However, serological testing is complicated by potential cross-reactivity with other flaviviruses (22). Newer enzyme-linked immunosorbent assay (ELISA) tests show promise, such as an IgG-ELISA test used in French Polynesia that, despite endemic dengue circulation, found <1% ZIKV seropositivity in blood donors before the outbreak (150).
To assess the risk and determinants of ZIKV-related clinical outcomes, we need studies aimed at measuring the underlying incidence of ZIKV infection, regardless of clinical presentation (e.g., serosurveys), the spectrum of illness and risk factors for severe outcomes (e.g., cohort and case-control studies), and the effect of ZIKV over longer time scales, including the length of immunity.

Ecology and evolution

There has been a high level of global concern surrounding the threat from ZIKV. One reason the concern is so great is that we are unable to accurately assess the global threat from the virus, and differing lines of evidence point to conflicting conclusions. For instance, the range of Aedes mosquitoes and ecological analyses would suggest that much of the continental United States is at risk from ZIKV, whereas recent experience with dengue and chikungunya would suggest that ZIKV is unlikely to persist in this region. To assess the epidemiologic and ecologic factors that drive global risk, there is a need for studies that more accurately assess where ZIKV circulation persists over long periods (e.g., global age-stratified serosurveys) and the ecological determinants of persistence (e.g., reservoirs, critical population size, and vector competence), as well as studies characterizing interactions between ZIKV and other flaviviruses. Across both clinical and ecological studies, it is important to evaluate the effect of host, viral, and mosquito genetics.

Interventions and control

A ZIKV vaccine may be the best way to protect at-risk populations over the long term. Vaccine development has been prioritized by the WHO and other public health agencies, and there are at least 18 active manufacturers and research institutions pursuing early stages of ZIKV vaccine development (151). However, phase 1 clinical studies are not expected to begin until the end of 2016 (151); hence, a vaccine is unlikely to become available in time to change the course of the current outbreak in the Americas.
Without a vaccine or antiviral drugs, the tools at our disposal for reducing ZIKV incidence are based on vector control and limiting ZIKV exposure. We have little direct evidence of the effectiveness of these approaches in controlling ZIKV transmission, but there are decades of experience in controlling dengue and other flaviviruses (152154). Effective vector control is possible: Gorgas virtually eliminated yellow fever from Havana and the Panama Canal region in the early 1900s using crude and draconian methods of vector control (155). Intensive vector control in the 1950s and 1960s, including mass DDT spraying, successfully eliminated A. aegypti from 18 countries in the Americas, substantially reducing dengue incidence (154, 156, 157). Later, Singapore and Cuba implemented successful vector-control programs lasting decades (154, 158, 159). However, all of these efforts ultimately proved to be unsustainable, and A. aegypti and dengue reemerged after their discontinuation (154, 158, 159). Nevertheless, there could be benefits from even short-term elimination, but research is needed to identify sustainable policies that can protect areas from ZIKV and/or other Aedes-borne diseases in the long term.
There is limited evidence for the effectiveness of measures aimed at reducing individual exposure to mosquitoes for dengue control. A meta-analysis suggests that use of screens in houses reduces the odds of dengue incidence by 78%, as does combined community environmental management and use of water-container covers (152). Other interventions—such as indoor residual spraying, repellents, bed nets, and traps—showed no statistically significant effect or a negative effect (insecticide aerosols) (152). However, these results are predominantly based on observational studies, limiting the strength of the evidence they provide. Topical insect repellents and other personal protective measures do reduce mosquito biting (160) and should decrease the risk of ZIKV infection. Some randomized trials have assessed the effect of interventions on mosquito populations with inconsistent results (152, 161), and there have been no well-designed trials assessing the effect of the common, WHO-recommended practice of space spraying or fogging to control dengue transmission (152). Well-designed experimental studies with end points of transmission and disease in humans are needed to better evaluate the effectiveness of interventions aimed at vector control and personal risk reduction.

Conclusion

The rise of ZIKV after its long persistence as a disease of apparently little importance highlights how little we truly understand about the global spread of flaviviruses and other vector-borne diseases. Over the past decades, dengue, chikungunya, West Nile virus, and now ZIKV have emerged or reemerged throughout the globe (2, 145). However, why these viruses have expanded their range, while others (e.g., yellow fever) have failed to invade areas potentially ripe for their spread, remains a mystery. New analytic and molecular tools have greatly expanded our ability to forecast risk and track the spread of these viruses, but a deep understanding of what makes one virus a global threat while another is not remains elusive. Although the important role of random chance and the continuing evolution of viral species may make precise forecasting of emerging pandemics impossible, we can continue to improve the speed with which we assess and respond to emerging threats.
The evidence highlighted in this review is both encouraging and disheartening. On the one hand, the speed with which the global community has collected and disseminated clinical, epidemiologic, and laboratory information on ZIKV after identification of the threat is impressive. But the development of therapeutics and diagnostics is hampered by our ignorance, despite knowing of ZIKV’s existence for more than half a century. Consequently, we have been able to do little to contain the virus’s rapid spread across the Americas. New threats from infectious diseases may emerge from unexpected places, and we need strategies in place that we can roll out to rapidly gain an understanding of the transmission, pathogenesis, and control of previously little-known pathogens to protect global public health.

Acknowledgments

We thank M. Kraemer and O. Brady for sharing the maps of the global probability of occurrence of Aedes and dengue. We also thank N. Reich, J. Konikoff, and J. Williamson for their help with a preliminary systematic review and analysis that laid the groundwork for this Review.

References and Notes

1
Dick G. W. A., Kitchen S. F., Haddow A. J., Zika virus. I. Isolations and serological specificity. Trans. R. Soc. Trop. Med. Hyg. 46, 509–520 (1952). 10.1016/0035-9203(52)90042-4
2
Fauci A. S., Morens D. M., Zika virus in the Americas: Yet another arbovirus threat. N. Engl. J. Med. 374, 601–604 (2016). 10.1056/NEJMp1600297
3
WHO, WHO Director-General summarizes the outcome of the Emergency Committee regarding clusters of microcephaly and Guillain-Barré syndrome (WHO, 2016); available at www.who.int/mediacentre/news/statements/2016/emergency-committee-zika-microcephaly/en/.
4
R. M. Anderson, R. M. May, Infectious Diseases of Humans: Dynamics and Control (Oxford Univ. Press, USA, 1991).
5
Cao-Lormeau V.-M., Roche C., Teissier A., Robin E., Berry A. L., Mallet H. P., Sall A. A., Musso D., Zika virus, French polynesia, South Pacific, 2013. Emerg. Infect. Dis. 20, 1085–1086 (2014). 10.3201/eid2011.141380
6
Duffy M. R., Chen T. H., Hancock W. T., Powers A. M., Kool J. L., Lanciotti R. S., Pretrick M., Marfel M., Holzbauer S., Dubray C., Guillaumot L., Griggs A., Bel M., Lambert A. J., Laven J., Kosoy O., Panella A., Biggerstaff B. J., Fischer M., Hayes E. B., Zika virus outbreak on Yap Island, Federated States of Micronesia. N. Engl. J. Med. 360, 2536–2543 (2009). 10.1056/NEJMoa0805715
7
Monlun E., Zeller H., Le Guenno B., Traoré-Lamizana M., Hervy J. P., Adam F., Ferrara L., Fontenille D., Sylla R., Mondo M., Surveillance of the circulation of arbovirus of medical interest in the region of eastern Senegal. Bull. Soc. Pathol. Exot. 86, 21–28 (1993). 8099299
8
Lloyd-Smith J. O., George D., Pepin K. M., Pitzer V. E., Pulliam J. R., Dobson A. P., Hudson P. J., Grenfell B. T., Epidemic dynamics at the human-animal interface. Science 326, 1362–1367 (2009). 10.1126/science.1177345
9
Althouse B. M., Hanley K. A., Diallo M., Sall A. A., Ba Y., Faye O., Diallo D., Watts D. M., Weaver S. C., Cummings D. A., Impact of climate and mosquito vector abundance on sylvatic arbovirus circulation dynamics in Senegal. Am. J. Trop. Med. Hyg. 92, 88–97 (2015). 10.4269/ajtmh.13-0617
10
Petersen L. R., Jamieson D. J., Powers A. M., Honein M. A., Zika virus. N. Engl. J. Med. 374, 1552–1563 (2016). 10.1056/NEJMra1602113
11
Kokernot R. H., Casaca V. M., Weinbren M. P., McIntosh B. M., Survey for antibodies against arthropod-borne viruses in the sera of indigenous residents of Angola. Trans. R. Soc. Trop. Med. Hyg. 59, 563–570 (1965). 10.1016/0035-9203(65)90159-8
12
Smithburn K. C., Studies on certain viruses isolated in the tropics of Africa and South America; immunological reactions as determined by cross-neutralization tests. J. Immunol. 68, 441–460 (1952). 14946384
13
MacNamara F. N., Zika virus: A report on three cases of human infection during an epidemic of jaundice in Nigeria. Trans. R. Soc. Trop. Med. Hyg. 48, 139–145 (1954). 10.1016/0035-9203(54)90006-1
14
Dick G. W., Epidemiological notes on some viruses isolated in Uganda; Yellow fever, Rift Valley fever, Bwamba fever, West Nile, Mengo, Semliki forest, Bunyamwera, Ntaya, Uganda S and Zika viruses. Trans. R. Soc. Trop. Med. Hyg. 47, 13–48 (1953). 10.1016/0035-9203(53)90021-2
15
Smithburn K. C., Kerr J. A., Gatne P. B., Neutralizing antibodies against certain viruses in the sera of residents of India. J. Immunol. 72, 248–257 (1954). 13163397
16
Pond W. L., Arthropod-borne virus antibodies in sera from residents of south-east Asia. Trans. R. Soc. Trop. Med. Hyg. 57, 364–371 (1963). 10.1016/0035-9203(63)90100-7
17
Smithburn K. C., Neutralizing antibodies against arthropod-borne viruses in the sera of long-time residents of Malaya and Borneo. Am. J. Hyg. 59, 157–163 (1954). 13138582
18
Dick G. W. A., Zika virus. II. Pathogenicity and physical properties. Trans. R. Soc. Trop. Med. Hyg. 46, 521–534 (1952). 10.1016/0035-9203(52)90043-6
19
Bearcroft W. G. C., Zika virus infection experimentally induced in a human volunteer. Trans. R. Soc. Trop. Med. Hyg. 50, 442–448 (1956). 10.1016/0035-9203(56)90090-6
20
Simpson D. I., Zika virus infection in man. Trans. R. Soc. Trop. Med. Hyg. 58, 335–337 (1964). 10.1016/0035-9203(64)90201-9
21
Cornet M., Robin Y., Adam C., Valade M., Calvo M. A., Transmission expérimentale comparée du virus amaril et du virus Zika chez Aedes aegypti L. Cah. ORSTOM Série Entomologie Médicale et Parasitologie 17, 47–53 (1979).
22
Musso D., Gubler D. J., Zika virus. Clin. Microbiol. Rev. 29, 487–524 (2016). 10.1128/CMR.00072-15
23
Vogel G., Mosquito hunters search for Zika vectors. Science 352, 1152–1153 (2016). 10.1126/science.352.6290.1152
24
Wong P.-S. J., Li M.-Z. I., Chong C.-S., Ng L.-C., Tan C.-H., Aedes (Stegomyia) albopictus (Skuse): A potential vector of Zika virus in Singapore. PLOS Negl. Trop. Dis. 7, e2348 (2013). 10.1371/journal.pntd.0002348
25
Grard G., Caron M., Mombo I. M., Nkoghe D., Mboui Ondo S., Jiolle D., Fontenille D., Paupy C., Leroy E. M., Zika virus in Gabon (Central Africa)—2007: A new threat from Aedes albopictus? PLOS Negl. Trop. Dis. 8, e2681 (2014). 10.1371/journal.pntd.0002681
26
Faye O., Freire C. C., Iamarino A., Faye O., de Oliveira J. V., Diallo M., Zanotto P. M., Sall A. A., Molecular evolution of Zika virus during its emergence in the 20th century. PLOS Negl. Trop. Dis. 8, e2636 (2014). 10.1371/journal.pntd.0002636
27
Bowen E. T., Simpson D. I., Platt G. S., Way H., Bright W. F., Day J., Achapa S., Roberts J. M., Large scale irrigation and arbovirus epidemiology, Kano Plain, Kenya. II. Preliminary serological survey. Trans. R. Soc. Trop. Med. Hyg. 67, 702–709 (1973). 10.1016/0035-9203(73)90041-2
28
Rodhain F., Gonzalez J. P., Mercier E., Helynck B., Larouze B., Hannoun C., Arbovirus infections and viral haemorrhagic fevers in Uganda: A serological survey in Karamoja district, 1984. Trans. R. Soc. Trop. Med. Hyg. 83, 851–854 (1989). 10.1016/0035-9203(89)90352-0
29
Fagbami A. H., Zika virus infections in Nigeria: Virological and seroepidemiological investigations in Oyo State. J. Hyg. 83, 213–219 (1979). 10.1017/S0022172400025997
30
Darwish M. A., Hoogstraal H., Roberts T. J., Ahmed I. P., Omar F., A sero-epidemiological survey for certain arboviruses (Togaviridae) in Pakistan. Trans. R. Soc. Trop. Med. Hyg. 77, 442–445 (1983). 10.1016/0035-9203(83)90106-2
31
Olson J. G., Ksiazek T. G., Suhandiman, Triwibowo, Zika virus, a cause of fever in Central Java, Indonesia. Trans. R. Soc. Trop. Med. Hyg. 75, 389–393 (1981). 10.1016/0035-9203(81)90100-0
32
Olson J. G., Ksiazek T. G., Gubler D. J., Lubis S. I., Simanjuntak G., Lee V. H., Nalim S., Juslis K., See R., A survey for arboviral antibodies in sera of humans and animals in Lombok, Republic of Indonesia. Ann. Trop. Med. Parasitol. 77, 131–137 (1983). 6309104
33
McCrae A. W., Kirya B. G., Yellow fever and Zika virus epizootics and enzootics in Uganda. Trans. R. Soc. Trop. Med. Hyg. 76, 552–562 (1982). 10.1016/0035-9203(82)90161-4
34
Lee V. H., Moore D. L., Vectors of the 1969 yellow fever epidemic on the Jos Plateau, Nigeria. Bull. World Health Organ. 46, 669–673 (1972). 4403105
35
Haddow A. J., Williams M. C., Woodall J. P., Simpson D. I., Goma L. K., Twelve isolations of Zika virus from Aedes (Stegomyia) africanus (Theobald) taken in and above a Uganda forest. Bull. World Health Organ. 31, 57–69 (1964).14230895
36
McIntosh B. M., Worth C. B., Kokernot R. H., Isolation of Semliki Forest virus from Aedes (Aedimorphus) argenteopunctatus (Theobald) collected in Portuguese East Africa. Trans. R. Soc. Trop. Med. Hyg. 55, 192–198 (1961). 10.1016/0035-9203(61)90025-6
37
Weinbren M. P., Williams M. C., Zika virus: Further isolations in the Zika area, and some studies on the strains isolated. Trans. R. Soc. Trop. Med. Hyg. 52, 263–268 (1958). 10.1016/0035-9203(58)90085-3
38
Marchette N. J., Garcia R., Rudnick A., Isolation of Zika virus from Aedes aegypti mosquitoes in Malaysia. Am. J. Trop. Med. Hyg. 18, 411–415 (1969). 4976739
39
Fagbami A., Epidemiological investigations on arbovirus infections at Igbo-Ora, Nigeria. Trop. Geogr. Med. 29, 187–191 (1977). 906078
40
Moore D. L., Causey O. R., Carey D. E., Reddy S., Cooke A. R., Akinkugbe F. M., David-West T. S., Kemp G. E., Arthropod-borne viral infections of man in Nigeria, 1964-1970. Ann. Trop. Med. Parasitol. 69, 49–64 (1975). 10.1080/00034983.1975.11686983
41
Lanciotti R. S., Kosoy O. L., Laven J. J., Velez J. O., Lambert A. J., Johnson A. J., Stanfield S. M., Duffy M. R., Genetic and serologic properties of Zika virus associated with an epidemic, Yap State, Micronesia, 2007. Emerg. Infect. Dis. 14, 1232–1239 (2008). 10.3201/eid1408.080287
42
Kwong J. C., Druce J. D., Leder K., Zika virus infection acquired during brief travel to Indonesia. Am. J. Trop. Med. Hyg. 89, 516–517 (2013). 10.4269/ajtmh.13-0029
43
Buathong R., Hermann L., Thaisomboonsuk B., Rutvisuttinunt W., Klungthong C., Chinnawirotpisan P., Manasatienkij W., Nisalak A., Fernandez S., Yoon I. K., Akrasewi P., Plipat T., Detection of Zika virus infection in Thailand, 2012-2014. Am. J. Trop. Med. Hyg. 93, 380–383 (2015). 10.4269/ajtmh.15-0022
44
Fonseca K., Meatherall B., Zarra D., Drebot M., MacDonald J., Pabbaraju K., Wong S., Webster P., Lindsay R., Tellier R., First case of Zika virus infection in a returning Canadian traveler. Am. J. Trop. Med. Hyg. 91, 1035–1038 (2014). 10.4269/ajtmh.14-0151
45
Cao-Lormeau V.-M., Blake A., Mons S., Lastère S., Roche C., Vanhomwegen J., Dub T., Baudouin L., Teissier A., Larre P., Vial A. L., Decam C., Choumet V., Halstead S. K., Willison H. J., Musset L., Manuguerra J. C., Despres P., Fournier E., Mallet H. P., Musso D., Fontanet A., Neil J., Ghawché F., Guillain-Barré Syndrome outbreak associated with Zika virus infection in French Polynesia: A case-control study. Lancet 387, 1531–1539 (2016). 10.1016/S0140-6736(16)00562-6
46
Cauchemez S., Besnard M., Bompard P., Dub T., Guillemette-Artur P., Eyrolle-Guignot D., Salje H., Van Kerkhove M. D., Abadie V., Garel C., Fontanet A., Mallet H. P., Association between Zika virus and microcephaly in French Polynesia, 2013-15: A retrospective study. Lancet 387, 2125–2132 (2016). 10.1016/S0140-6736(16)00651-6
47
Roth A., Mercier A., Lepers C., Hoy D., Duituturaga S., Benyon E., Guillaumot L., Souares Y., Concurrent outbreaks of dengue, chikungunya and Zika virus infections - an unprecedented epidemic wave of mosquito-borne viruses in the Pacific 2012-2014. Euro Surveill. 19, 20929 (2014). 10.2807/1560-7917.ES2014.19.41.20929
48
Campos G. S., Bandeira A. C., Sardi S. I., Zika virus outbreak, Bahia, Brazil. Emerg. Infect. Dis. 21, 1885–1886 (2015). 10.3201/eid2110.150847
49
Lednicky J., Beau De Rochars V. M., El Badry M., Loeb J., Telisma T., Chavannes S., Anilis G., Cella E., Ciccozzi M., Rashid M., Okech B., Salemi M., Morris J. G., Zika virus outbreak in Haiti in 2014: Molecular and clinical data. PLOS Negl. Trop. Dis. 10, e0004687 (2016). 10.1371/journal.pntd.0004687
50
World Health Organization, Zika virus outbreaks in the Americas. Wkly. Epidemiol. Rec. 90, 609–610 (2015). 26552108
51
Schuler-Faccini L., Ribeiro E. M., Feitosa I. M., Horovitz D. D., Cavalcanti D. P., Pessoa A., Doriqui M. J., Neri J. I., Neto J. M., Wanderley H. Y., Cernach M., El-Husny A. S., Pone M. V., Serao C. L., Sanseverino M. T.Brazilian Medical Genetics Society–Zika Embryopathy Task Force, Possible association between Zika virus infection and microcephaly - Brazil, 2015. MMWR Morb. Mortal. Wkly. Rep. 65, 59–62 (2016). 10.15585/mmwr.mm6503e2
52
Faria N. R., Azevedo Rdo. S., Kraemer M. U., Souza R., Cunha M. S., Hill S. C., Thézé J., Bonsall M. B., Bowden T. A., Rissanen I., Rocco I. M., Nogueira J. S., Maeda A. Y., Vasami F. G., Macedo F. L., Suzuki A., Rodrigues S. G., Cruz A. C., Nunes B. T., Medeiros D. B., Rodrigues D. S., Nunes Queiroz A. L., da Silva E. V., Henriques D. F., Travassos da Rosa E. S., de Oliveira C. S., Martins L. C., Vasconcelos H. B., Casseb L. M., Simith Dde. B., Messina J. P., Abade L., Lourenço J., Carlos Junior Alcantara L., de Lima M. M., Giovanetti M., Hay S. I., de Oliveira R. S., Lemos Pda. S., de Oliveira L. F., de Lima C. P., da Silva S. P., de Vasconcelos J. M., Franco L., Cardoso J. F., Vianez-Júnior J. L., Mir D., Bello G., Delatorre E., Khan K., Creatore M., Coelho G. E., de Oliveira W. K., Tesh R., Pybus O. G., Nunes M. R., Vasconcelos P. F., Zika virus in the Americas: Early epidemiological and genetic findings. Science 352, 345–349 (2016). 27013429
53
CDC, All Countries and Territories with Active Zika Virus Transmission; available at www.cdc.gov/zika/geo/active-countries.html.
54
Pacheco O., Beltrán M., Nelson C. A., Valencia D., Tolosa N., Farr S. L., Padilla A. V., Tong V. T., Cuevas E. L., Espinosa-Bode A., Pardo L., Rico A., Reefhuis J., González M., Mercado M., Chaparro P., Martínez Duran M., Rao C. Y., Muñoz M. M., Powers A. M., Cuéllar C., Helfand R., Huguett C., Jamieson D. J., Honein M. A., Ospina Martínez M. L., Zika virus disease in Colombia: Preliminary report. N. Engl. J. Med. NEJMoa1604037 (2016). 10.1056/NEJMoa1604037
55
N. Casey, M. E. Díaz, Colombia reports first cases of microcephaly linked to Zika virus. New York Times (14 April, 2016); available at www.nytimes.com/2016/04/15/world/americas/colombia-reports-first-cases-of-microcephaly-linked-to-zika-virus.html?smid=pl-share.
56
Boletín Epidemiológico Semenal, Semana epidemiológica número 12 de 2016; available at www.ins.gov.co/boletin-epidemiologico/Boletn%20Epidemiolgico/2016%20Bolet%C3%ADn%20epidemiol%C3%B3gico%20semana%2012.pdf.
57
Chan J. F. W., Choi G. K. Y., Yip C. C. Y., Cheng V. C. C., Yuen K.-Y., Zika fever and congenital Zika syndrome: An unexpected emerging arboviral disease. J. Infect. 72, 507–524 (2016). 10.1016/j.jinf.2016.02.011
58
Deckard D. T., Chung W. M., Brooks J. T., Smith J. C., Woldai S., Hennessey M., Kwit N., Mead P., Male-to-male sexual transmission of Zika virus - Texas, January 2016. MMWR Morb. Mortal. Wkly. Rep. 65, 372–374 (2016). 10.15585/mmwr.mm6514a3
59
Venturi G., Zammarchi L., Fortuna C., Remoli M. E., Benedetti E., Fiorentini C., Trotta M., Rizzo C., Mantella A., Rezza G., Bartoloni A., An autochthonous case of Zika due to possible sexual transmission, Florence, Italy, 2014. Euro Surveill. 21, 30148 (2016). 10.2807/1560-7917.ES.2016.21.8.30148
60
Foy B. D., Kobylinski K. C., Chilson Foy J. L., Blitvich B. J., Travassos da Rosa A., Haddow A. D., Lanciotti R. S., Tesh R. B., Probable non-vector-borne transmission of Zika virus, Colorado, USA. Emerg. Infect. Dis. 17, 880–882 (2011). 10.3201/eid1705.101939
61
Besnard M., Lastere S., Teissier A., Cao-Lormeau V., Musso D., Evidence of perinatal transmission of Zika virus, French Polynesia, December 2013 and February 2014. Euro Surveill. 19, 20751 (2014). 10.2807/1560-7917.ES2014.19.13.20751
62
R. Editorial, Brazil reports Zika infection from blood transfusions. Reuters (2016); available at www.reuters.com/article/us-health-zika-brazil-blood-idUSKCN0VD22N.
63
Musso D., Nhan T., Robin E., Roche C., Bierlaire D., Zisou K., Shan Yan A., Cao-Lormeau V. M., Broult J., Potential for Zika virus transmission through blood transfusion demonstrated during an outbreak in French Polynesia, November 2013 to February 2014. Euro Surveill. 19, 20761 (2014). 10.2807/1560-7917.ES2014.19.14.20761
64
J. Lessler et al., Times to Key Events in the Course of Zika Infection and their Implications for Surveillance: A Systematic Review and Pooled Analysis. bioRxiv (2016), p. 041913.
65
Gourinat A.-C., O’Connor O., Calvez E., Goarant C., Dupont-Rouzeyrol M., Detection of Zika virus in urine. Emerg. Infect. Dis. 21, 84–86 (2015). 10.3201/eid2101.140894
66
Rozé B., Najioullah F., Fergé J. L., Apetse K., Brouste Y., Cesaire R., Fagour C., Fagour L., Hochedez P., Jeannin S., Joux J., Mehdaoui H., Valentino R., Signate A., Cabié A., Zika virus detection in urine from patients with Guillain-Barré syndrome on Martinique, January 2016. Euro Surveill. 21, (2016). 10.2807/1560-7917.ES.2016.21.9.30154
67
Musso D., Roche C., Nhan T. X., Robin E., Teissier A., Cao-Lormeau V. M., Detection of Zika virus in saliva. J. Clin. Virol. 68, 53–55 (2015). 10.1016/j.jcv.2015.04.021
68
Musso D., Roche C., Robin E., Nhan T., Teissier A., Cao-Lormeau V. M., Potential sexual transmission of Zika virus. Emerg. Infect. Dis. 21, 359–361 (2015). 10.3201/eid2102.141363
69
Calvet G., Aguiar R. S., Melo A. S., Sampaio S. A., de Filippis I., Fabri A., Araujo E. S., de Sequeira P. C., de Mendonça M. C., de Oliveira L., Tschoeke D. A., Schrago C. G., Thompson F. L., Brasil P., Dos Santos F. B., Nogueira R. M., Tanuri A., de Filippis A. M., Detection and sequencing of Zika virus from amniotic fluid of fetuses with microcephaly in Brazil: A case study. Lancet Infect. Dis. 16, 653–660 (2016). 10.1016/S1473-3099(16)00095-5
70
Martines R. B., Bhatnagar J., Keating M. K., Silva-Flannery L., Muehlenbachs A., Gary J., Goldsmith C., Hale G., Ritter J., Rollin D., Shieh W. J., Luz K. G., Ramos A. M., Davi H. P., Kleber de Oliveria W., Lanciotti R., Lambert A., Zaki S., Notes from the field: Evidence of Zika virus infection in brain and placental tissues from two congenitally infected newborns and two fetal losses—Brazil, 2015. MMWR Morb. Mortal. Wkly. Rep. 65, 159–160 (2016). 10.15585/mmwr.mm6506e1
71
Driggers R. W., Ho C. Y., Korhonen E. M., Kuivanen S., Jääskeläinen A. J., Smura T., Rosenberg A., Hill D. A., DeBiasi R. L., Vezina G., Timofeev J., Rodriguez F. J., Levanov L., Razak J., Iyengar P., Hennenfent A., Kennedy R., Lanciotti R., du Plessis A., Vapalahti O., Zika virus infection with prolonged maternal viremia and fetal brain abnormalities. N. Engl. J. Med. 374, 2142–2151 (2016). 10.1056/NEJMoa1601824
72
Murphy B. R., Whitehead S. S., Immune response to dengue virus and prospects for a vaccine. Annu. Rev. Immunol. 29, 587–619 (2011). 10.1146/annurev-immunol-031210-101315
73
Brasil P., Calvet G. A., Siqueira A. M., Wakimoto M., de Sequeira P. C., Nobre A., Quintana M. S., Mendonça M. C., Lupi O., de Souza R. V., Romero C., Zogbi H., Bressan C. S., Alves S. S., Lourenço-de-Oliveira R., Nogueira R. M., Carvalho M. S., de Filippis A. M., Jaenisch T., Zika virus outbreak in Rio de Janeiro, Brazil: Clinical characterization, epidemiological and virological aspects. PLOS Negl. Trop. Dis. 10, e0004636 (2016). 10.1371/journal.pntd.0004636
74
Gyurech D., Schilling J., Schmidt-Chanasit J., Cassinotti P., Kaeppeli F., Dobec M., False positive dengue NS1 antigen test in a traveller with an acute Zika virus infection imported into Switzerland. Swiss Med. Wkly. 146, w14296 (2016). 26859285
75
Chen L. H., Zika virus infection in a Massachusetts resident after travel to Costa Rica: A case report. Ann. Intern. Med. 164, 574–576 (2016). 10.7326/L16-0075
76
Zammarchi L., Stella G., Mantella A., Bartolozzi D., Tappe D., Günther S., Oestereich L., Cadar D., Muñoz-Fontela C., Bartoloni A., Schmidt-Chanasit J., Zika virus infections imported to Italy: Clinical, immunological and virological findings, and public health implications. J. Clin. Virol. 63, 32–35 (2015). 10.1016/j.jcv.2014.12.005
77
Tappe D., Nachtigall S., Kapaun A., Schnitzler P., Günther S., Schmidt-Chanasit J., Acute Zika virus infection after travel to Malaysian Borneo, September 2014. Emerg. Infect. Dis. 21, 911–913 (2015). 10.3201/eid2105.141960
78
Summers D. J., Acosta R. W., Acosta A. M., Zika virus in an American recreational traveler. J. Travel Med. 22, 338–340 (2015). 10.1111/jtm.12208
79
Zammarchi L., Tappe D., Fortuna C., Remoli M. E., Günther S., Venturi G., Bartoloni A., Schmidt-Chanasit J., Zika virus infection in a traveller returning to Europe from Brazil, March 2015. Euro Surveill. 20, 21153 (2015). 10.2807/1560-7917.ES2015.20.23.21153
80
Maria A. T., Maquart M., Makinson A., Flusin O., Segondy M., Leparc-Goffart I., Le Moing V., Foulongne V., Zika virus infections in three travellers returning from South America and the Caribbean respectively, to Montpellier, France, December 2015 to January 2016. Euro Surveill. 21, 30131 (2016). 10.2807/1560-7917.ES.2016.21.6.30131
81
Sarmiento-Ospina A., Vásquez-Serna H., Jimenez-Canizales C. E., Villamil-Gómez W. E., Rodriguez-Morales A. J., Zika virus associated deaths in Colombia. Lancet Infect. Dis. 16, 523–524 (2016). 10.1016/S1473-3099(16)30006-8
82
Arzuza-Ortega L., Polo A., Pérez-Tatis G., López-García H., Parra E., Pardo-Herrera L. C., Rico-Turca A. M., Villamil-Gómez W., Rodríguez-Morales A. J., Fatal sickle cell disease and Zika virus infection in girl from Colombia. Emerg. Infect. Dis. 22, 925–927 (2016). 10.3201/eid2205.151934
83
Alshekhlee A., Hussain Z., Sultan B., Katirji B., Guillain-Barré syndrome: Incidence and mortality rates in US hospitals. Neurology 70, 1608–1613 (2008). 10.1212/01.wnl.0000310983.38724.d4
84
van Doorn P. A., Ruts L., Jacobs B. C., Clinical features, pathogenesis, and treatment of Guillain-Barré syndrome. Lancet Neurol. 7, 939–950 (2008). 10.1016/S1474-4422(08)70215-1
85
Carteaux G., Maquart M., Bedet A., Contou D., Brugières P., Fourati S., Cleret de Langavant L., de Broucker T., Brun-Buisson C., Leparc-Goffart I., Mekontso Dessap A., Zika virus associated with meningoencephalitis. N. Engl. J. Med. 374, 1595–1596 (2016). 10.1056/NEJMc1602964
86
Mécharles S., Herrmann C., Poullain P., Tran T. H., Deschamps N., Mathon G., Landais A., Breurec S., Lannuzel A., Acute myelitis due to Zika virus infection. Lancet 387, 1481 (2016). 10.1016/S0140-6736(16)00644-9
87
Victora C. G., Schuler-Faccini L., Matijasevich A., Ribeiro E., Pessoa A., Barros F. C., Microcephaly in Brazil: How to interpret reported numbers? Lancet 387, 621–624 (2016). 10.1016/S0140-6736(16)00273-7
88
C. D. E. O. de Emergências Em Saúde Pública Sobre Microcefalias, “Informe Epidemiológico No 25 – Semana Epidemiológica (Se) 18/2016 (01/05 A 07/05/2016) Monitoramento Dos Casos De Microcefalia No Brasil” (Ministerio de Salud, Brasil, 2016).
89
R. M. Kliegman, B. M. D. Stanton, J. St. Geme, N. F. Schor, Nelson Textbook of Pediatrics (Elsevier Health Sciences, 2015).
90
Rasmussen S. A., Jamieson D. J., Honein M. A., Petersen L. R., Zika virus and birth defects: Reviewing the evidence for causality. N. Engl. J. Med. 374, 1981–1987 (2016). 10.1056/NEJMsr1604338
91
Cugola F. R., Fernandes I. R., Russo F. B., Freitas B. C., Dias J. L., Guimarães K. P., Benazzato C., Almeida N., Pignatari G. C., Romero S., Polonio C. M., Cunha I., Freitas C. L., Brandão W. N., Rossato C., Andrade D. G., Faria Dde. P., Garcez A. T., Buchpigel C. A., Braconi C. T., Mendes E., Sall A. A., Zanotto P. M., Peron J. P., Muotri A. R., Beltrão-Braga P. C., The Brazilian Zika virus strain causes birth defects in experimental models. Nature 534, 267–271 (2016).27279226
92
Brasil P., Pereira J. P., Raja Gabaglia C., Damasceno L., Wakimoto M., Ribeiro Nogueira R. M., Carvalho de Sequeira P., Machado Siqueira A., Abreu de Carvalho L. M., Cotrim da Cunha D., Calvet G. A., Neves E. S., Moreira M. E., Rodrigues Baião A. E., Nassar de Carvalho P. R., Janzen C., Valderramos S. G., Cherry J. D., Bispo de Filippis A. M., Nielsen-Saines K., Zika virus infection in pregnant women in Rio de Janeiro - Preliminary report. N. Engl. J. Med. NEJMoa1602412 (2016). 10.1056/NEJMoa1602412
93
Johansson M. A., Mier-Y-Teran-Romero L., Reefhuis J., Gilboa S. M., Hills S. L., Zika and the risk of microcephaly. N. Engl. J. Med. 375, 1–4 (2016). 10.1056/NEJMp1605367
94
de Fatima Vasco Aragao M., van der Linden V., Brainer-Lima A. M., Coeli R. R., Rocha M. A., Sobral da Silva P., Durce Costa Gomes de Carvalho M., van der Linden A., Cesario de Holanda A., Valenca M. M., Clinical features and neuroimaging (CT and MRI) findings in presumed Zika virus related congenital infection and microcephaly: Retrospective case series study. BMJ 353, i1901 (2016). 10.1136/bmj.i1901
95
Kleber de Oliveira W., Cortez-Escalante J., De Oliveira W. T., do Carmo G. M., Henriques C. M., Coelho G. E., Araújo de França G. V., Increase in reported prevalence of microcephaly in infants born to women living in areas with confirmed Zika virus transmission during the first trimester of pregnancy - Brazil, 2015. MMWR Morb. Mortal. Wkly. Rep. 65, 242–247 (2016). 10.15585/mmwr.mm6509e2
96
Barbeiro F. M., Fonseca S. C., Tauffer M. G., Ferreira M. S., Silva F. P., Ventura P. M., Quadros J. I., Fetal deaths in Brazil: A systematic review. Rev. Saude Publica 49, 22 (2015). 10.1590/S0034-8910.2015049005568
97
World Health Organization, Maternal Anthropometry and Pregnancy Outcomes: A WHO Collaborative Study (WHO, 1995).
98
Waitzman N. J., Romano P. S., Scheffler R. M., Estimates of the economic costs of birth defects. Inquiry 31, 188–205 (1994). 8021024
99
A. Perkins, A. Siraj, C. Warren Ruktanonchai, M. Kraemer, A. Tatem, Model-based projections of Zika virus infections in childbearing women in the Americas (2016); http://biorxiv.org/content/early/2016/05/19/039610.
100
Kraemer M. U. G., Sinka M. E., Duda K. A., Mylne A. Q., Shearer F. M., Barker C. M., Moore C. G., Carvalho R. G., Coelho G. E., Van Bortel W., Hendrickx G., Schaffner F., Elyazar I. R., Teng H. J., Brady O. J., Messina J. P., Pigott D. M., Scott T. W., Smith D. L., Wint G. R., Golding N., Hay S. I., The global distribution of the arbovirus vectors Aedes aegypti and Ae. albopictus. eLife 4, e08347 (2015). 10.7554/eLife.08347
101
Diallo D., Sall A. A., Diagne C. T., Faye O., Faye O., Ba Y., Hanley K. A., Buenemann M., Weaver S. C., Diallo M., Zika virus emergence in mosquitoes in southeastern Senegal, 2011. PLOS ONE 9, e109442 (2014). 10.1371/journal.pone.0109442
102
Centers for Disease Control and Prevention (CDC), Locally acquired dengue, Key West, Florida, 2009-2010. MMWR Morb. Mortal. Wkly. Rep. 59, 577–581 (2010). 20489680
103
Bhatt S., Gething P. W., Brady O. J., Messina J. P., Farlow A. W., Moyes C. L., Drake J. M., Brownstein J. S., Hoen A. G., Sankoh O., Myers M. F., George D. B., Jaenisch T., Wint G. R., Simmons C. P., Scott T. W., Farrar J. J., Hay S. I., The global distribution and burden of dengue. Nature 496, 504–507 (2013). 10.1038/nature12060
104
Reiter P., Lathrop S., Bunning M., Biggerstaff B., Singer D., Tiwari T., Baber L., Amador M., Thirion J., Hayes J., Seca C., Mendez J., Ramirez B., Robinson J., Rawlings J., Vorndam V., Waterman S., Gubler D., Clark G., Hayes E., Texas lifestyle limits transmission of dengue virus. Emerg. Infect. Dis. 9, 86–89 (2003). 10.3201/eid0901.020220
105
Bogoch I. I., Brady O. J., Kraemer M. U., German M., Creatore M. I., Kulkarni M. A., Brownstein J. S., Mekaru S. R., Hay S. I., Groot E., Watts A., Khan K., Anticipating the international spread of Zika virus from Brazil. Lancet 387, 335–336 (2016). 10.1016/S0140-6736(16)00080-5
106
Monaghan A. J., Morin C. W., Steinhoff D. F., Wilhelmi O., Hayden M., Quattrochi D. A., Reiskind M., Lloyd A. L., Smith K., Schmidt C. A., Scalf P. E., Ernst K., On the seasonal occurrence and abundance of the Zika virus vector mosquito Aedes aegypti in the contiguous United States. PLOS Curr. 8, (2016). 10.1371/currents.outbreaks.50dfc7f46798675fc63e7d7da563da76
107
C. Carlson, C. Colin, D. Eric, G. Wayne, An ecological assessment of the pandemic threat of Zika virus (2016); http://biorxiv.org/content/early/2016/03/09/040386.
108
Messina J. P., Kraemer M. U., Brady O. J., Pigott D. M., Shearer F. M., Weiss D. J., Golding N., Ruktanonchai C. W., Gething P. W., Cohn E., Brownstein J. S., Khan K., Tatem A. J., Jaenisch T., Murray C. J., Marinho F., Scott T. W., Hay S. I., Mapping global environmental suitability for Zika virus. eLife 5, e15272 (2016). 10.7554/eLife.15272
109
Samy A., Thomas S. M., El Wahed A. A., Cohoon K. P., Peterson A. T., Global map of Zika virus. Mem. Inst. Oswaldo Cruz (2016). 10.1590/0074-02760160149
110
Rogers D. J., Wilson A. J., Hay S. I., Graham A. J., The global distribution of yellow fever and dengue. Adv. Parasitol. 62, 181–220 (2006). 10.1016/S0065-308X(05)62006-4
111
Kucharski A. J., Funk S., Eggo R. M., Mallet H. P., Edmunds W. J., Nilles E. J., Transmission dynamics of Zika virus in island populations: A modelling analysis of the 2013-14 French Polynesia outbreak. PLOS Negl. Trop. Dis. 10, e0004726 (2016). 10.1371/journal.pntd.0004726
112
S. Funk et al., Comparative analysis of dengue and Zika outbreaks reveals differences by setting and virus (2016); http://biorxiv.org/content/early/2016/04/26/043265.
113
Nishiura H., Kinoshita R., Mizumoto K., Yasuda Y., Nah K., Transmission potential of Zika virus infection in the South Pacific. Int. J. Infect. Dis. 45, 95–97 (2016). 10.1016/j.ijid.2016.02.017
114
D. P. Rojas et al., The epidemiology and transmissibility of Zika virus in Girardot and San Andres Island, Colombia (2016); http://biorxiv.org/content/early/2016/04/24/049957.
115
Nishiura H., Mizumoto K., Villamil-Gómez W. E., Rodríguez-Morales A. J., Preliminary estimation of the basic reproduction number of Zika virus infection during Colombia epidemic, 2015-2016. Travel Med. Infect. Dis. 14, 274–276 (2016). 10.1016/j.tmaid.2016.03.016
116
Ferguson N. M., Cucunubá Z. M., Dorigatti I., Nedjati-Gilani G. L., Donnelly C. A., Basáñez M.-G., Nouvellet P., Lessler J., Understanding the invasion dynamics of Zika in Latin America: Implications for policy. Science 353, 353–354 (2016). 10.1126/science.aag0219
117
L. Bastos et al., Zika in Rio de Janeiro: Assessment of basic reproductive number and its comparison with dengue (2016); http://biorxiv.org/content/early/2016/05/25/055475.
118
Saluzzo J. F., Gonzalez J. P., Hervé J. P., Georges A. J., Serological survey for the prevalence of certain arboviruses in the human population of the south-east area of Central African Republic (author’s transl). Bull. Soc. Pathol. Exot. Filiales 74, 490–499 (1981). 6274526
119
Kilbourn A. M., Karesh W. B., Wolfe N. D., Bosi E. J., Cook R. A., Andau M., Health evaluation of free-ranging and semi-captive orangutans (Pongo pygmaeus pygmaeus) in Sabah, Malaysia. J. Wildl. Dis. 39, 73–83 (2003). 10.7589/0090-3558-39.1.73
120
Brès P., Recent data from serological surveys on the prevalence of arbovirus infections in Africa, with special reference to yellow fever. Bull. World Health Organ. 43, 223–267 (1970). 5312522
121
Geser A., Henderson B. E., Christensen S., A multipurpose serological survey in Kenya. 2. Results of arbovirus serological tests. Bull. World Health Organ. 43, 539–552 (1970). 5313066
122
B. Althouse et al., Potential for Zika virus to establish a sylvatic transmission cycle in the Americas (2016); http://biorxiv.org/content/early/2016/04/05/047175.
123
S. Favoretto et al., First detection of Zika virus in neotropical primates in Brazil: a possible new reservoir (2016); http://biorxiv.org/content/early/2016/04/20/049395.
124
Musso D., Zika virus transmission from French Polynesia to Brazil. Emerg. Infect. Dis. 21, 1887 (2015). 10.3201/eid2110.151125
125
Climate Prediction Center, ENSO Diagnostic Discussion; available at www.cpc.ncep.noaa.gov/products/analysis_monitoring/enso_advisory/ensodisc.html.
126
Watts D. M., Burke D. S., Harrison B. A., Whitmire R. E., Nisalak A., Effect of temperature on the vector efficiency of Aedes aegypti for dengue 2 virus. Am. J. Trop. Med. Hyg. 36, 143–152 (1987). 3812879
127
Yang H. M., Macoris M. L. G., Galvani K. C., Andrighetti M. T. M., Wanderley D. M. V., Assessing the effects of temperature on the population of Aedes aegypti, the vector of dengue. Epidemiol. Infect. 137, 1188–1202 (2009). 10.1017/S0950268809002040
128
Alto B. W., Juliano S. A., Precipitation and temperature effects on populations of Aedes albopictus (Diptera: Culicidae): implications for range expansion. J. Med. Entomol. 38, 646–656 (2001). 10.1603/0022-2585-38.5.646
129
South America Summer Forecast: El Nino to Bring Flooding Rain to Argentina, Uruguay and Southeast Brazil. AccuWeather; available at www.accuweather.com/en/weather-news/south-america-summer-forecast-2015-2016/53158136.
130
Pontes R. J., Freeman J., Oliveira-Lima J. W., Hodgson J. C., Spielman A., Vector densities that potentiate dengue outbreaks in a Brazilian city. Am. J. Trop. Med. Hyg. 62, 378–383 (2000). 11037781
131
Morin C. W., Comrie A. C., Ernst K., Climate and dengue transmission: Evidence and implications. Environ. Health Perspect. 121, 1264–1272 (2013). 24058050
132
Ríos-Velásquez C. M., Codeço C. T., Honório N. A., Sabroza P. S., Moresco M., Cunha I. C., Levino A., Toledo L. M., Luz S. L., Distribution of dengue vectors in neighborhoods with different urbanization types of Manaus, state of Amazonas, Brazil. Mem. Inst. Oswaldo Cruz 102, 617–623 (2007). 10.1590/S0074-02762007005000076
133
Halstead S. B., In vivo enhancement of dengue virus infection in rhesus monkeys by passively transferred antibody. J. Infect. Dis. 140, 527–533 (1979). 10.1093/infdis/140.4.527
134
Recker M., Blyuss K. B., Simmons C. P., Hien T. T., Wills B., Farrar J., Gupta S., Immunological serotype interactions and their effect on the epidemiological pattern of dengue. Proc. Biol. Sci. 276, 2541–2548 (2009). 10.1098/rspb.2009.0331
135
Burke D. S., Nisalak A., Johnson D. E., Scott R. M., A prospective study of dengue infections in Bangkok. Am. J. Trop. Med. Hyg. 38, 172–180 (1988). 3341519
136
Halstead S. B., Porterfield J. S., O’Rourke E. J., Enhancement of dengue virus infection in monocytes by flavivirus antisera. Am. J. Trop. Med. Hyg. 29, 638–642 (1980). 6157332
137
Anderson K. B., Gibbons R. V., Thomas S. J., Rothman A. L., Nisalak A., Berkelman R. L., Libraty D. H., Endy T. P., Preexisting Japanese encephalitis virus neutralizing antibodies and increased symptomatic dengue illness in a school-based cohort in Thailand. PLOS Negl. Trop. Dis. 5, e1311 (2011). 10.1371/journal.pntd.0001311
138
L. M. Paul et al., Dengue virus antibodies enhance Zika virus infection (2016); http://biorxiv.org/content/early/2016/04/25/050112.
139
Fagbami A. H., Halstead S. B., Marchette N. J., Larsen K., Cross-infection enhancement among African flaviviruses by immune mouse ascitic fluids. Cytobios 49, 49–55 (1987).3028713
140
Wang L., Valderramos S. G., Wu A., Ouyang S., Li C., Brasil P., Bonaldo M., Coates T., Nielsen-Saines K., Jiang T., Aliyari R., Cheng G., From mosquitos to humans: Genetic evolution of Zika virus. Cell Host Microbe 19, 561–565 (2016). 10.1016/j.chom.2016.04.006
141
C. C. de Melo Freire et al., Spread of the pandemic Zika virus lineage is associated with NS1 codon usage adaptation in humans (2015); http://biorxiv.org/content/early/2015/11/25/032839.
142
Fajardo A., Soñora M., Moreno P., Moratorio G., Cristina J., Bayesian coalescent inference reveals high evolutionary rates and diversification of Zika virus populations. J. Med. Virol. (2016). 10.1002/jmv.24596
143
Hill A. V. S., Evolution, revolution and heresy in the genetics of infectious disease susceptibility. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 840–849 (2012). 10.1098/rstb.2011.0275
144
Brass A. L., Huang I. C., Benita Y., John S. P., Krishnan M. N., Feeley E. M., Ryan B. J., Weyer J. L., van der Weyden L., Fikrig E., Adams D. J., Xavier R. J., Farzan M., Elledge S. J., The IFITM proteins mediate cellular resistance to influenza A H1N1 virus, West Nile virus, and dengue virus. Cell 139, 1243–1254 (2009). 10.1016/j.cell.2009.12.017
145
Mackenzie J. S., Gubler D. J., Petersen L. R., Emerging flaviviruses: The spread and resurgence of Japanese encephalitis, West Nile and dengue viruses. Nat. Med. 10 (Suppl), S98–S109 (2004). 10.1038/nm1144
146
Moreno-Mayar J. V., Rasmussen S., Seguin-Orlando A., Rasmussen M., Liang M., Flåm S. T., Lie B. A., Gilfillan G. D., Nielsen R., Thorsby E., Willerslev E., Malaspinas A. S., Genome-wide ancestry patterns in Rapanui suggest pre-European admixture with Native Americans. Curr. Biol. 24, 2518–2525 (2014). 10.1016/j.cub.2014.09.057
147
Black W. C., Bennett K. E., Gorrochótegui-Escalante N., Barillas-Mury C. V., Fernández-Salas I., de Lourdes Muñoz M., Farfán-Alé J. A., Olson K. E., Beaty B. J., Flavivirus susceptibility in Aedes aegypti. Arch. Med. Res. 33, 379–388 (2002). 10.1016/S0188-4409(02)00373-9
148
Chouin-Carneiro T., Vega-Rua A., Vazeille M., Yebakima A., Girod R., Goindin D., Dupont-Rouzeyrol M., Lourenço-de-Oliveira R., Failloux A. B., Differential susceptibilities of Aedes aegypti and Aedes albopictus from the Americas to Zika virus. PLOS Negl. Trop. Dis. 10, e0004543 (2016). 10.1371/journal.pntd.0004543
149
Faye O., Faye O., Dupressoir A., Weidmann M., Ndiaye M., Alpha Sall A., One-step RT-PCR for detection of Zika virus. J. Clin. Virol. 43, 96–101 (2008). 10.1016/j.jcv.2008.05.005
150
Aubry M., Finke J., Teissier A., Roche C., Broult J., Paulous S., Desprès P., Cao-Lormeau V. M., Musso D., Seroprevalence of arboviruses among blood donors in French Polynesia, 2011-2013. Int. J. Infect. Dis. 41, 11–12 (2015). 10.1016/j.ijid.2015.10.005
151
World Health Organization, Current Zika Product Pipeline (WHO, 2016); available at www.who.int/csr/research-and-development/zika-rd-pipeline.pdf.
152
Bowman L. R., Donegan S., McCall P. J., Is dengue vector control deficient in effectiveness or evidence?: Systematic review and meta-analysis. PLOS Negl. Trop. Dis. 10, e0004551 (2016). 10.1371/journal.pntd.0004551
153
Wilson A. L., Dhiman R. C., Kitron U., Scott T. W., van den Berg H., Lindsay S. W., Benefit of insecticide-treated nets, curtains and screening on vector borne diseases, excluding malaria: A systematic review and meta-analysis. PLOS Negl. Trop. Dis. 8, e3228 (2014). 10.1371/journal.pntd.0003228
154
Achee N. L., Gould F., Perkins T. A., Reiner R. C., Morrison A. C., Ritchie S. A., Gubler D. J., Teyssou R., Scott T. W., A critical assessment of vector control for dengue prevention. PLOS Negl. Trop. Dis. 9, e0003655 (2015). 10.1371/journal.pntd.0003655
155
Patterson R., Dr. William Gorgas and his war with the mosquito. CMAJ 141, 596–597, 599 (1989). 2673502
156
Soper F. L., The elimination of urban yellow fever in the Americas through the eradication of Aedes aegypti. Am. J. Public Health Nations Health 53, 7–16 (1963). 10.2105/AJPH.53.1.7
157
The feasibility of eradicating Aedes aegypti in the Americas. Rev. Panam. Salud Publica 1, 68–72 (1997). 10.1590/S1020-49891997000100023
158
Kouri G., Gustavo K., Reemergence of dengue in Cuba: A 1998 epidemic in Santiago de Cuba. Emerg. Infect. Dis. 4, 85–88 (1998). 10.3201/eid0401.980111
159
Ooi E.-E., Goh K.-T., Gubler D. J., Dengue prevention and 35 years of vector control in Singapore. Emerg. Infect. Dis. 12, 887–893 (2006). 10.3201/eid1206.051210
160
Fradin M. S., Day J. F., Comparative efficacy of insect repellents against mosquito bites. N. Engl. J. Med. 347, 13–18 (2002). 10.1056/NEJMoa011699
161
Lenhart A., Trongtokit Y., Alexander N., Apiwathnasorn C., Satimai W., Vanlerberghe V., Van der Stuyft P., McCall P. J., A cluster-randomized trial of insecticide-treated curtains for dengue vector control in Thailand. Am. J. Trop. Med. Hyg. 88, 254–259 (2013). 10.4269/ajtmh.2012.12-0423
162
Calisher C. H., Karabatsos N., Dalrymple J. M., Shope R. E., Porterfield J. S., Westaway E. G., Brandt W. E., Antigenic relationships between flaviviruses as determined by cross-neutralization tests with polyclonal antisera. J. Gen. Virol. 70, 37–43 (1989). 10.1099/0022-1317-70-1-37
163
Sievers F., Higgins D. G., Clustal Omega, accurate alignment of very large numbers of sequences. Methods Mol. Biol. 1079, 105–116 (2013). 10.1007/978-1-62703-646-7_6
164
Nguyen L.-T., Schmidt H. A., von Haeseler A., Minh B. Q., IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 32, 268–274 (2015). 10.1093/molbev/msu300

(0)eLetters

eLetters is a forum for ongoing peer review. eLetters are not edited, proofread, or indexed, but they are screened. eLetters should provide substantive and scholarly commentary on the article. Embedded figures cannot be submitted, and we discourage the use of figures within eLetters in general. If a figure is essential, please include a link to the figure within the text of the eLetter. Please read our Terms of Service before submitting an eLetter.

Log In to Submit a Response

No eLetters have been published for this article yet.

Information & Authors

Information

Published In

Science
Volume 353 | Issue 6300
12 August 2016

Article versions

You are viewing the most recent version of this article.

Submission history

Published in print: 12 August 2016

Permissions

Request permissions for this article.

Acknowledgments

We thank M. Kraemer and O. Brady for sharing the maps of the global probability of occurrence of Aedes and dengue. We also thank N. Reich, J. Konikoff, and J. Williamson for their help with a preliminary systematic review and analysis that laid the groundwork for this Review.

Authors

Affiliations

Justin Lessler*, [email protected]
Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.
Lelia H. Chaisson
Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.
Lauren M. Kucirka
Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.
Department of Surgery, Johns Hopkins University School of Medicine, Baltimore, MD, USA.
Qifang Bi
Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.
Kyra Grantz
Department of Biology, Emerging Pathogens Institute, University of Florida, Gainesville, FL, USA.
Henrik Salje
Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.
Mathematical Modelling of Infectious Diseases Unit, Institut Pasteur, Paris, France.
Andrea C. Carcelen
Department of International Health, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.
Cassandra T. Ott
Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.
Jeanne S. Sheffield
Department of Gynecology and Obstetrics, Johns Hopkins University School of Medicine, Baltimore, MD, USA.
Neil M. Ferguson
Department of Medicine, School of Public Health, Imperial College London, London, UK.
Derek A. T. Cummings
Department of Biology, Emerging Pathogens Institute, University of Florida, Gainesville, FL, USA.
C. Jessica E. Metcalf
Department of Ecology and Evolutionary Biology, Princeton University, Princeton, NJ, USA.
Office of Population Research, Princeton University, Princeton, NJ, USA.
Isabel Rodriguez-Barraquer
Department of Epidemiology, Johns Hopkins Bloomberg School of Public Health, Baltimore, MD, USA.

Notes

*
Corresponding author. Email: [email protected]
These authors contributed equally to this work.

Metrics & Citations

Metrics

Article Usage

Altmetrics

Citations

Cite as

Export citation

Select the format you want to export the citation of this publication.

Cited by

  1. MODELING AND ANALYSIS OF LOW-LEVEL TRANSMISSION ZIKV DYNAMICS VIA A POISSON POINT PROCESS ON SEXUAL TRANSMISSION ROUTE, Journal of Applied Analysis & Computation, 13, 2, (1044-1069), (2023).https://doi.org/10.11948/20220351
    Crossref
  2. Viral immunity: Basic mechanisms and therapeutic applications—a Keystone Symposia report, Annals of the New York Academy of Sciences, 1521, 1, (32-45), (2023).https://doi.org/10.1111/nyas.14960
    Crossref
  3. Bioactivity of brassica seed meals and its compounds as ecofriendly larvicides against mosquitoes, Scientific Reports, 13, 1, (2023).https://doi.org/10.1038/s41598-023-30563-6
    Crossref
  4. Mice 3D testicular organoid system as a novel tool to study Zika virus pathogenesis, Virologica Sinica, 38, 1, (66-74), (2023).https://doi.org/10.1016/j.virs.2022.10.001
    Crossref
  5. Excito-repellency of Myristica fragrans Houtt. and Curcuma longa L. extracts from Southern Thailand against Aedes aegypti (L.) , PeerJ, 10, (e13357), (2022).https://doi.org/10.7717/peerj.13357
    Crossref
  6. Zika Virus: A Review, Research Journal of Pharmacology and Pharmacodynamics, (171-173), (2022).https://doi.org/10.52711/2321-5836.2022.00029
    Crossref
  7. Brazilian Populations of Aedes aegypti Resistant to Pyriproxyfen Exhibit Lower Susceptibility to Infection with Zika Virus, Viruses, 14, 10, (2198), (2022).https://doi.org/10.3390/v14102198
    Crossref
  8. Seroprevalence of Zika Virus in Amphawa District, Thailand, after the 2016 Pandemic, Viruses, 14, 3, (476), (2022).https://doi.org/10.3390/v14030476
    Crossref
  9. Zika Virus Infection of Sertoli Cells Alters Protein Expression Involved in Activated Immune and Antiviral Response Pathways, Carbohydrate Metabolism and Cardiovascular Disease, Viruses, 14, 2, (377), (2022).https://doi.org/10.3390/v14020377
    Crossref
  10. Lipid Droplets and Their Participation in Zika Virus Infection, International Journal of Molecular Sciences, 23, 20, (12584), (2022).https://doi.org/10.3390/ijms232012584
    Crossref
  11. See more
Loading...

View Options

View options

PDF format

Download this article as a PDF file

Download PDF

Check Access

Log in to view the full text

AAAS ID LOGIN

AAAS login provides access to Science for AAAS Members, and access to other journals in the Science family to users who have purchased individual subscriptions.

Log in via OpenAthens.
Log in via Shibboleth.

More options

Register for free to read this article

As a service to the community, this article is available for free. Login or register for free to read this article.

Purchase this issue in print

Buy a single issue of Science for just $15 USD.

Media

Figures

Multimedia

Tables

Share

Share

Share article link

Share on social media