Skip to main content
eLife logoLink to eLife
. 2016 Apr 19;5:e15272. doi: 10.7554/eLife.15272

Mapping global environmental suitability for Zika virus

Jane P Messina 1,*, Moritz UG Kraemer 1, Oliver J Brady 2, David M Pigott 2,3, Freya M Shearer 2, Daniel J Weiss 1, Nick Golding 4, Corrine W Ruktanonchai 5, Peter W Gething 1, Emily Cohn 6, John S Brownstein 6, Kamran Khan 7,8, Andrew J Tatem 5,9, Thomas Jaenisch 10,11, Christopher JL Murray 3, Fatima Marinho 12, Thomas W Scott 13, Simon I Hay 2,3,*
Editor: Mark Jit14
PMCID: PMC4889326  PMID: 27090089

Abstract

Zika virus was discovered in Uganda in 1947 and is transmitted by Aedes mosquitoes, which also act as vectors for dengue and chikungunya viruses throughout much of the tropical world. In 2007, an outbreak in the Federated States of Micronesia sparked public health concern. In 2013, the virus began to spread across other parts of Oceania and in 2015, a large outbreak in Latin America began in Brazil. Possible associations with microcephaly and Guillain-Barré syndrome observed in this outbreak have raised concerns about continued global spread of Zika virus, prompting its declaration as a Public Health Emergency of International Concern by the World Health Organization. We conducted species distribution modelling to map environmental suitability for Zika. We show a large portion of tropical and sub-tropical regions globally have suitable environmental conditions with over 2.17 billion people inhabiting these areas.

DOI: http://dx.doi.org/10.7554/eLife.15272.001

Research Organism: Human, Virus

eLife digest

Zika virus is transmitted between humans by mosquitoes. The majority of infections cause mild flu-like symptoms, but neurological complications in adults and infants have been found in recent outbreaks.

Although it was discovered in Uganda in 1947, Zika only caused sporadic infections in humans until 2007, when it caused a large outbreak in the Federated States of Micronesia. The virus later spread across Oceania, was first reported in Brazil in 2015 and has since rapidly spread across Latin America. This has led many people to question how far it will continue to spread. There was therefore a need to define the areas where the virus could be transmitted, including the human populations that might be risk in these areas.

Messina et al. have now mapped the areas that provide conditions that are highly suitable for the spread of the Zika virus. These areas occur in many tropical and sub-tropical regions around the globe. The largest areas of risk in the Americas lie in Brazil, Colombia and Venezuela. Although Zika has yet to be reported in the USA, a large portion of the southeast region from Texas through to Florida is highly suitable for transmission. Much of sub-Saharan Africa (where several sporadic cases have been reported since the 1950s) also presents an environment that is highly suitable for the Zika virus. While no cases have yet been reported in India, a large portion of the subcontinent is also suitable for Zika transmission.

Over 2 billion people live in Zika-suitable areas globally, and in the Americas alone, over 5.4 million births occurred in 2015 within such areas. It is important, however, to recognize that not all individuals living in suitable areas will necessarily be exposed to Zika.

We still lack a great deal of basic epidemiological information about Zika. More needs to be known about the species of mosquito that spreads the disease and how the Zika virus interacts with related viruses such as dengue. As such information becomes available and clinical cases become routinely diagnosed, the global evidence base will be strengthened, which will improve the accuracy of future maps.

DOI: http://dx.doi.org/10.7554/eLife.15272.002

Introduction

Zika virus (ZIKV) is an emerging arbovirus carried by mosquitoes of the genus Aedes (Musso et al., 2014). Although discovered in Uganda in 1947 (Dick et al., 1952Dick, 1953) ZIKV was only known to cause sporadic infections in humans in Africa and Asia until 2007 (Lanciotti et al., 2008), when it caused a large outbreak of symptomatic cases on Yap island in the Federated States of Micronesia (FSM), followed by another in French Polynesia in 2013–14 and subsequent spread across Oceania (Musso et al., 2015a). In the 2007 Yap island outbreak, it was estimated that approximately 20% of ZIKV cases were symptomatic. While indigenous transmission of ZIKV to humans was reported for the first time in Latin America in 2015 (Zanluca et al., 2015; World Health Organisation, 2015), recent phylogeographic research estimates that the virus was introduced into the region between May and December 2013 (Faria et al., 2016). This recent rapid spread has led to concern that the virus is following a similar pattern of global expansion to that of dengue and chikungunya (Musso et al., 2015a).

ZIKV has been isolated from 19 different Aedes species (Haddow et al., 2012; Grard et al., 2014), but virus has been most frequently found in Ae. aegypti (Monlun et al., 1992; Marchette et al., 1969; Smithburn, 1954; Pond, 1963; Faye et al., 2008; Foy et al., 2011b; WHO Collaborating Center for Reference and Research on Arboviruses and Hemorrhagic Fever Viruses: Annual Report, 1999). These studies were based upon ancestral African strains of ZIKV, but the current rapid spread of ZIKV in Latin America is indicative of this highly efficient arbovirus vector (Marcondes and Ximenes, 2015). The relatively recent global spread of Ae. albopictus (Benedict et al., 2007; Kraemer et al., 2015c) and the rarity of ZIKV isolations from wild mosquitoes may also partially explain the lower frequency of isolations from Ae. albopictus populations. Whilst virus transmission by Ae. albopictus and other minor vector species has normally resulted in only a small number of cases (Kutsuna et al., 2015; Roiz et al., 2015), these vectors do pose the threat of limited transmission (Grard et al., 2014). The wide geographic distribution of Ae. albopictus combined with the frequent virus introduction via viraemic travellers (McCarthy 2016; Bogoch et al., 2016; Morrison et al., 2008; Scott and Takken, 2012), means the risk for ZIKV infection via this vector must therefore also be considered in ZIKV mapping.

The fact that ZIKV reporting was limited to a few small areas in Africa and Asia until 2007 means that global risk mapping has not, until recently, been a priority (Pigott et al., 2015b). Recent associations with Guillain-Barré syndrome in adults and microcephaly in infants born to ZIKV-infected mothers (World Health Organisation, 2015; Martines et al., 2016) have revealed that ZIKV could lead to more severe complications than the mild rash and flu-like symptoms that characterize the majority of symptomatic cases (Gatherer and Kohl, 2016). Considering these potentially severe complications and the rapid expansion of ZIKV into previously unaffected areas, the global public health community needs information about those areas that are environmentally suitable for transmission of ZIKV to humans. Being a closely related flavivirus to DENV, there is furthermore the potential for antigen-based diagnostic tests to exhibit cross-reactivity when IgM ELISA is used for rapid diagnosis. Although ZIKV-specific serologic assays are being developed by the U.S. Centers for Disease Control, currently the only method of confirming ZIKV infection is by using PCR on acute specimens (Lanciotti et al., 2008, Faye et al., 2008). Awareness of suitability for transmission is essential if proper detection methods are to be employed.

In this paper, we use species distribution modelling techniques that have been useful for mapping other vector-borne diseases such as dengue (Bhatt et al., 2013), Leishmaniasis (Pigott et al., 2014b), and Crimean-Congo Haemorrhagic Fever (Messina et al., 2015b) to map environmental suitability for ZIKV. The environmental niche of a disease can be identified according to a combination of environmental conditions supporting its presence in a particular location, with statistical modelling then allowing this niche to be described quantitatively (Kraemer et al., 2016). Niche modelling uses records of known disease occurrence alongside hypothesized environmental covariates to predict suitability for disease transmission in regions where it has yet to be reported (Elith and Leathwick, 2009). Contemporary high spatial-resolution global data representing a variety of environmental conditions allows for these predictions to be made at a global scale (Hay et al., 2006).

Results

Figure 1A shows the locations of the 323 standardized occurrence records in the final dataset, classified by the following date ranges: (i) up until 2006 (before the outbreak in FSM); (ii) between 2007 (the year of the FSM outbreak) and 2014; and (iii) since 2015, the first reporting of ZIKV in the Americas. This map is accompanied by the graph in Figure 1B, showing the number of reported occurrence locations globally by year. These figures highlight the more sporadic nature of reporting until recent years, with the majority of occurrences in the dataset (63%) coming from the recent 2015–2016 outbreak in Latin America.

Figure 1. (A) Map showing the distribution of the final set of 323 ZIKV occurrence locations entered into the ensemble Boosted Regression Tree modelling procedure.

Locations are classified by year of occurrence to show those which took place (i) prior to the 2007 outbreak in Federated States of Micronesia; (ii) between 2007–2014; and (iii) during the 2015–2016 outbreak; (B) the total number of locations reporting symptomatic ZIKV occurrence in humans globally over time.

DOI: http://dx.doi.org/10.7554/eLife.15272.003

Figure 1.

Figure 1—figure supplement 1. Maps of all covariates entered into the 300 BRT models.

Figure 1—figure supplement 1.

(A) probability of being urban, 2015; (B) enhanced vegetation index; (C) minimum relative humidity; (D) cumulative annual precipitation (mm); (E) temperature suitability for dengue via Ae. aegypti; (F) temperature suitability for dengue via Ae. albopictus

The final map that resulted from the mean of 300 ensemble Boosted Regression Tree (BRT) models is shown in Figure 2A (with greater detail shown for each region in Figures 2B–D). Figure 2—figure supplement 1 shows the distribution of uncertainty based upon the upper and lower prediction quantiles from the 300 models. We restricted our models to make predictions only within areas where i) mosquito vectors (in this case Ae. aegypti) were able to persist and ii) where temperature was sufficient for arboviral replication within the mosquito. The former of these was calculated by taking the Ae. aegypti probability of occurrence (Kraemer et al., 2015c) value that incorporated 90% of all known occurrences (Kraemer et al., 2015b) (giving a threshold value of 0.8 and greater) while the latter was evaluated using a mechanistic mosquito model (Brady et al., 2013; 2014), which identified regions where arboviral transmission could be sustained for at least 355 days (one year minus the human incubation period) in an average year. Figure 3 is a country-level map distinguishing between those countries that are currently reporting ZIKV, those which have reported ZIKV in the past, those which have highly suitable areas for transmission, and those which are unsuitable. Our models predicted high levels of risk for ZIKV in many areas within the tropical and sub-tropical zones. Large portions of the Americas are suitable for transmission, with the largest areas of risk occurring in Brazil, followed by Colombia and Venezuela, all of which have reported high numbers of cases in the 2015–2016 outbreak. In Brazil, where the highest numbers of ZIKV are reported in the ongoing epidemic, the coastal cities in the south as well as large areas of the north are identified to have the highest environmental suitability of ZIKV. The central region of Brazil, on the other hand, has low population densities and smaller mosquito populations, which is reflected in the relatively low suitability for ZIKV transmission seen in the map. Although ZIKV has yet to be reported in the USA, a large portion of the southeast region of the country, including much of Texas through to Florida, is also highly suitable for transmission. Potential risk for ZIKV transmission is high in much of sub-Saharan Africa, with continuous suitability in the Democratic Republic of Congo and surrounding areas and several sporadic case reports in western sub-Saharan countries since the 1950s. Although no symptomatic cases have yet been reported in India, a large portion of this country is at potential risk for ZIKV transmission (over 2 million square kilometres), with environmental suitability extending from its northwest regions through to Bangladesh and Myanmar. The Indochina region, southeast China, and Indonesia all have large areas of environmental suitability as well, extending into Oceania. While only representing less than ten percent of Australia’s total land area, the area shown to be suitable for ZIKV transmission in its northernmost regions is considerable (comprising nearly 250,000 square kilometres).

Figure 2. Maps of (A) global environmental suitability for ZIKV, ranging from 0 (grey) to 1 (red), showing greater detail for (B) the Americas, (C) Africa, and (D) Asia and Oceania.

DOI: http://dx.doi.org/10.7554/eLife.15272.005

Figure 2.

Figure 2—figure supplement 1. Uncertainty around Zika suitability predictions displayed in main manuscript – Figure 2, ranging from less than 0.01 (very little uncertainty) to 0.94 (greatest uncertainty).

Figure 2—figure supplement 1.

Figure 2—figure supplement 2. Effect plots for each covariate entered into the ensemble of 300 BRT models.

Figure 2—figure supplement 2.

(A) minimum relative humidity; (B) cumulative annual precipitation (mm); (C) enhanced vegetation index; (B) probability of being urban (%); (E) temperature suitability for dengue via Ae. aegypti; (F) temperature suitability for dengue via Ae. albopictus.

Figure 2—figure supplement 3. Environmental suitability for Zika virus transmission to humans, not taking into account temperature suitability for dengue via Aedes albopictus.

Figure 2—figure supplement 3.

Covariate effects are as follows: cumulative annual precipitation (67.4%); temperature suitability for dengue via Ae. aegypti (16.9%); probability of being urban, 2015 (8.2%); enhanced vegetation index (5.1%); minimum relative humidity (2.4%).

Figure 2—figure supplement 4. Map showing areas predicted to have greater dengue suitability (from Bhatt et al., 2013, Nature) vs those which are predicted to have greater Zika suitability in the current study.

Figure 2—figure supplement 4.

These values are restricted to areas where both diseases had non-zero predictions.

Figure 3. Status of ZIKV reporting as of 2016 by country, showing countries that are highly environmentally suitable (having a suitable area of more than 10,000 square kilometres) but which have not yet reported symptomatic cases of ZIKV in humans. 'Currently reporting' countries are those having reported cases since 2015.

Figure 3.

DOI: http://dx.doi.org/10.7554/eLife.15272.010

Our models showed ZIKV risk to be particularly influenced by annual cumulative precipitation, contributing 65.0% to the variation in the ensemble of models. The next most important predictor in the model was temperature suitability for DENV transmission via Ae. albopictus, contributing 14.6%. These are followed by urban extents (8.3%), temperature suitability for DENV via Ae. aegypti (5.7%), the Enhanced Vegetation Index (EVI; 3.8%), and minimum relative humidity (2.5%). Effect plots for each covariate are provided in Figure 2—figure supplement 2. Validation statistics indicated high predictive performance of the BRT ensemble mean map evaluated in a 10-fold cross-validation procedure, with area under the receiver operating characteristic (AUC) of 0.829 ( ± 0.121 SD). Due to the uncertainty about Ae. albopictus as a competent vector for ZIKV, we also provide results for an ensemble of models which did not include temperature suitability for dengue via this mosquito species in Figure 2—figure supplement 3.

A threshold environmental suitability value of 0.397 in our final map was determined to incorporate 90% of all ZIKV occurrence locations. This was used to classify each 5 km x 5 km pixel on our final map as suitable or unsuitable for ZIKV transmission to humans. Using high-resolution global population estimates (WorldPop, 2015; SEDAC, 2015), we summed the populations living in Zika-suitable areas and have identified 2.17 billion people globally living within areas that are environmentally suitable for ZIKV transmission. Table 1 shows a breakdown of this figure by major world region, also showing the top four contributing countries to the potential population at risk. Asia has the most people living in areas that are suitable for ZIKV transmission at 1.42 billion, accounted for in large part by those living in India. In Africa, roughly 453 million people are living in areas suitable for ZIKV transmission, the largest proportion of which live in Nigeria. In the Americas, more than 298 million people live in ZIKV-suitable transmission zones, with approximately 40 percent of these people living in Brazil. Within the majority of environmentally suitable areas for ZIKV in the Americas, prolonged year-round transmission is possible. Southern Brazil and Argentina, however, are more likely to see transmission interrupted throughout the year, as is the case with the USA should autochthonous ZIKV transmission occur there. Using high-resolution data on births for the year 2015 (WorldPop, 2015), we also estimate that 5.42 million births will occur in the Americas over the next year within areas and times of environmental suitability for ZIKV transmission.

Table 1.

Population living in areas suitable for ZIKV transmission within each major world region and top four countries contributing to these populations at risk.

DOI: http://dx.doi.org/10.7554/eLife.15272.011

Region/Country Population living in areas suitable for ZIKV transmission (millions)
Africa 452.58
Nigeria 111.97
Democratic Republic of the Congo 68.95
Uganda 33.43
United Republic of Tanzania 22.70
Americas 298.36
Brazil 120.65
Mexico 32.22
Colombia 29.54
Venezuela 22.22
Asia 1,422.13
India 413.19
Indonesia 226.04
China 213.84
Bangladesh 133.29
World 2,173.27

Discussion

A large number of viruses (circa 219) are known to be pathogenic (Woolhouse et al., 2012). Of the 53 species of Flavivirus, 19 are reported to have caused illness in humans (ICTV, 2014). Some flaviviruses, such as DENV, YFV, Japanese encephalitis virus, and West Nile virus, are widespread, causing many thousands of infections each year. The remainder, however, have been recognized as being pathogenic to humans for decades, but have highly focal reported distributions and are only minor contributors to mortality and disability globally (Hay et al., 2013; Murray et al., 2015). As a result, many are of relatively low priority when research and policy interest are considered (Pigott et al., 2015b). The recent spread of ZIKV across the globe highlights the need to reassess our consideration of these other flaviviruses, to gain a better understanding of the factors driving their spread and the potential for geographic expansion beyond their currently limited geographical extents.

Environmental suitability for virus transmission in an area does not necessarily mean that it will arrive and/or establish in that location. Arboviral infections in particular are dependent on a variety of non-environmental factors, with their movement having historically been largely attributed to human mobility from travel, trade, and migration, which introduce the viruses to places where mosquito vectors are already present (Murray et al., 2013; Weaver and Reisen, 2010; Nunes et al., 2015; Gubler and Clark, 1995). The identification of locations with permissible environments for transmission of emerging diseases like ZIKV is crucial, as importation could give rise to subsequent autochthonous cases in these locations (Hennessey et al., 2016; Zanluca et al., 2015). In order to identify places potentially receptive for ZIKV, we assembled the first comprehensive spatial dataset for ZIKV occurrence in humans and compiled a comprehensive set of high-resolution environmental covariates. We then used these data to implement a species distribution modelling approach (Elith and Leathwick, 2009) that has proven useful for mapping other vector-borne diseases (Bhatt et al., 2013; Pigott et al., 2014a; Mylne et al., 2015; Messina et al., 2015b), allowing us to make inferences about environmental suitability for ZIKV transmission in areas where it has yet to be reported or where we are less certain about its presence. How the ongoing epidemic unfolds in terms of case numbers (or incidence) will depend on a range of other factors such as local transmission dynamics, herd immunity, patterns of contact among mosquitoes and infectious and susceptible humans (Stoddard et al., 2013), and mosquito-to-human ratios as recently shown for dengue (Kraemer et al., 2015a) and chikungunya (Salje et al., 2016).

Globally, we predict that over 2.17 billion people live in areas that are environmentally suitable for ZIKV transmission. We also estimate the number of births occurring in the Americas only, as it is the region for which the most accurate high-resolution population data on births exists (Tatem et al., 2014; Sorichetta et al., 2015) and because it is the focus of an ongoing outbreak, which is the largest recorded thus far. In the Americas alone, an estimated 5.42 million births occurred in 2015 within areas and at times that are suitable for ZIKV transmission. It is important to recognize that not all individuals will be exposed to ZIKV. Like with other flaviviruses, a ZIKV outbreak may be temporally and spatially sporadic and, even in the most receptive environments, is unlikely that all of the population will be infected. Furthermore, increasing herd immunity of this likely sterilizing infection will rapidly reduce the size of the susceptible population at risk for infection in subsequent years (Dick et al., 1952) and work is ongoing to predict the likely infection dynamics after establishment. Instead, the estimates are intended as indicators of the total number of individuals or births that may require protection during the first wave of the outbreak. Specifically, these populations should be the focus of efforts to increase awareness and provide guidelines for mitigating personal risk of infection. In future analyses, our estimates could be extended to include ZIKV incidence and the virus’ effect on incidence of associated conditions such as Guillain-Barré syndrome and microcephaly. Before appropriately caveated estimates can be generated, however, more information is needed regarding: (i) the background rate of these conditions due to other causes; (ii) how risk may vary throughout the course of a pregnancy; (iii) the proportion of the population exposed during outbreaks; and (iv) whether or not immunity acquired through a mother’s prior exposure is protective.

For all arboviral diseases, public health education about reducing populations and avoiding contact with mosquito vectors is required in at-risk areas. Specific to ZIKV is the risk of microcephaly in newborns, which has led public health agencies to issue warnings for women who are currently or planning on becoming pregnant in areas suspected to have ongoing ZIKV transmission and the declaration of a Public Health Emergency of International Concern (Heymann et al., 2016). Due to the sensitive nature and implications of these warnings, it is important that levels of risk are rigorously estimated, validated, and updated. Transmission of related arboviral diseases still occurs in many areas we defined as at-risk for ZIKV, which highlights the need for improved vector control outcomes, particularly those targeting Ae. aegypti. Predicted levels of risk for ZIKV transmission are potentially helpful for prioritized allocation of vector control resources, as well as for differential diagnosis and, if a vaccine becomes available, delivery efforts. It should be noted that instances of ZIKV sexual transmission have been reported (Patino-Barbosa et al., 2015; Musso et al., 2015b; Foy et al., 2011a). We did not incorporate secondary modes of transmission into the models we described here, but our map can help inform future discussions about the potential impact of this mode of transmission as its relative importance becomes better understood.

A great deal of basic epidemiological information specific to ZIKV is lacking. As a result, information must be leveraged from our knowledge about transmission of related arboviruses. Previous work has focused on mapping other vector borne diseases that share much of the ecology of Zika, such as DENV (Bhatt et al., 2013) and CHIKV, as well as for its primary vectors, Ae. aegypti and Ae. albopictus (Kraemer et al., 2015c). For this reason, temperature suitability for dengue (Brady et al., 2013, 2014) was entered into the models due to the greater number of field and laboratory studies available for parameterising this metric for DENV. Until more studies related to vector competence and temperature constraints on ZIKV transmission to humans are conducted, this is the most accurate indicator of arboviral disease transmission via Aedes mosquitoes currently available. Indeed, all other covariates in our models could equally be applied to mapping DENV and CHIKV, and ZIKV-specific refinements to modelling covariates will be possible as the disease continues to expand to allow for improvements in future iterations of the map. The relatively smaller amount of occurrence data available for ZIKV (especially prior to recent outbreaks) means that this dataset should also be updated with new information as necessary, leading to a stronger global evidence base and improved accuracy of future maps. Better understanding of ZIKV transmission dynamics will eventually allow for further cartographic refinements to be made, such as the differentiation between endemic- and epidemic-prone areas. Still, all covariates included in the current study have been updated and refined since (Bhatt et al., 2013), and when combined with the most extensive occurrence database available for ZIKV, the resulting map we present here is currently the most accurate depiction of the distribution of environmental suitability for ZIKV. A map highlighting differences in predicted suitability for both diseases is provided in Figure 2—figure supplement 4.

Conclusion

In this study, we produced the first global high spatial-resolution map of environmental suitability for ZIKV transmission to humans using an assembly of known records of ZIKV occurrence and environmental covariates in a species distribution modelling framework. While it is clear that much remains to be understood about ZIKV, this first map serves as a baseline for understanding the change in the geographical distribution of this globally emerging arboviral disease. Knowledge of the potential distribution can encourage more vigilant surveillance in both humans and Aedes mosquito populations, as well as help in the allocation of limited resources for disease prevention. Public health awareness campaigns and advice for mitigation of individual risk can also be focused in the areas we have predicted to be highly suitable for ZIKV transmission, particularly during the first wave of infection in a population. The maps we have presented may also inform existing travel advisories for pregnant women and other travellers. The maps and underlying data are freely available online via figshare (http://www.figshare.com).

Materials and methods

To map environmental suitability for ZIKV transmission to humans, we applied a species distribution modelling approach to establish a multivariate empirical relationship between the probability of ZIKV occurrence and the environmental conditions in locations where the disease has been confirmed. We employed an ensemble boosted regression trees (BRT) methodology (De'ath, 2007; Elith et al., 2008), which required the generation of: (i) a comprehensive compendium of known locations of disease occurrence in humans; (ii) a set of background points representing locations where ZIKV has not yet been reported; and (iii) a set of high-resolution globally gridded environmental and socioeconomic covariates hypothesised to affect ZIKV transmission. The resulting model produces a 5 x 5 km spatial-resolution global map of environmental suitability for ZIKV transmission to humans.

Assembly of the geo-referenced ZIKV occurrence dataset

Information about the locations of ZIKV occurrence in humans was extracted from peer-reviewed literature, case reports, and informal online sources following previously established protocols (Kraemer et al., 2015b; Messina et al., 2014; 2015a). To collate the peer-reviewed dataset, literature searches were undertaken using PubMed (http://www.ncbi.nlm.nih.gov/pubmed) and ISI Web of Science (http://www.webofknowledge.com) search engines using the search term 'Zika'. No language restrictions were placed on these searches; however, only those citations with a full title and abstract were retrieved, resulting in the review of 148 references ranging in publication dates between 1951 and 2015. In-house language skills allowed review of all English, French, Portuguese and Spanish articles for useable location information for human ZIKV occurrence. ProMED-mail (http://www.promedmail.org) was also searched using the term 'Zika', resulting in the review of 139 reports between 27 June 2007 and 18 January 2016. Additionally, the most current database of ZIKV case locations in Brazil was obtained directly from the Brazilian Ministry of Health. From all sources, only laboratory confirmation of symptomatic ZIKV infection in humans was entered into the dataset (mention of suspected cases was not entered). Serological evidence from healthy individuals could represent a past infection, with transmission potentially occurring in a different location to that where the individual currently resides (Darwish et al., 1983), or could be an artefact from possible cross-reactivity with a variety of different viruses (Smithburn et al., 1954). As a result, these less reliable diagnoses of ZIKV were excluded.

All available location information was extracted from each peer-reviewed article and ProMED case report. The site name was used together with all contextual information provided about the site to determine its latitudinal and longitudinal coordinates using Google Maps (https://www.maps.google.com). If the study site could be geo-positioned to a specific place, it was recorded as a point location. If the study site could only be identified at an administrative area level (e.g. province or district), it was recorded as a polygon along with an identifier of its administrative unit. If imported cases were reported with information on the site of infection, they were geo-positioned to this site; if imported cases were reported with no information about the site of infection, they were not entered into the dataset. Informal online data sources were collated automatically by the web-based system HealthMap (http://www.healthmap.org) as described elsewhere (Freifeld et al., 2008). Alerts for ZIKV were obtained from HealthMap for the years 2014–2016, and then manually checked for validity. In total, usable location information was extracted from 110 sources. Information was also collected about the status of symptoms in each reported occurrence, distinguishing between those where symptomatic cases were being reported, versus those where only seroprevalence was detected in healthy individuals.

Due to the potential for multiple independent reports referring to the same cases temporal and spatial standardization was required, as we have described previously in detail for dengue mapping efforts (Messina et al., 2014). In brief, an occurrence was defined as a unique location with one or more confirmed cases of ZIKV occurring within one calendar year (the finest temporal resolution available across all records). Point locations were considered to be overlapping if they lay on the same 5 km x 5 km pixel, and polygon locations were identified by a unique administrative unit code. Furthermore, all polygons whose geographic area was greater than one square decimal degree (approximately 111 square kilometers at the equator) were removed from the dataset to avoid averaging covariate values over very large areas, and only those occurrences comprising symptomatic individuals were retained for modelling purposes to ensure an accurate location of infection. In total, the final occurrence dataset contained 323 unique occurrences to be entered into our BRT modelling procedure. A map of the final set of occurrence locations is provided as Figure 1A.

Generation of the background location dataset

Separate maps of the relative probability of occurrence of Ae. aegypti and Ae. albopictus (Kraemer et al., 2015c) were used to compute a combined metric of the relative probability of vector occurrence, by taking the maximum value from the two layers for all 5 km x 5 km gridded cells globally. The inverse of this combined-Aedes occurrence probability layer (higher values indicating greater certainty of absence) was then used to draw a biased sample of 10,000 background locations. As such, a greater number of background points were sampled in areas where we are more certain that Ae. aegypti or Ae. albopictus do not occur, and therefore where ZIKV is less likely to be transmitted to humans. While it has been demonstrated that predictive accuracy from presence-background species distribution models can be improved by biasing background record locations toward areas with greatest reporting probabilities (Phillips et al., 2009), information on possible reporting biases, or proxies of such spatial bias, are currently unavailable for ZIKV. These 10,000 background locations were combined with the standardized occurrence dataset to serve as comparison data locations in the BRT species distribution modelling procedure. The background locations were weighted such that their total sum was equal to the total number of occurrence locations (n=237; pseudo-absence weighting=0.0237), in order to aid in the discrimination capacity of the model (Barbet-Massin et al., 2012).

Explanatory covariates

A set of six covariates hypothesized to influence the global distribution of ZIKV transmission to humans were used in our models to establish an empirical relationship between ZIKV presence or absence and underlying environmental conditions. These six covariates included: (i) an index of temperature suitability for dengue transmission to humans via Ae. aegypti; (ii) temperature suitability for dengue transmission to humans via Ae. albopictus; (iii) minimum relative humidity; (iv) annual cumulative precipitation; (v) an enhanced vegetation index (EVI); and (vi) urban versus rural habitat type. The underlying hypothesis behind each of the covariates is discussed in more detail below, along with a description of data sources and any processing that was undertaken before entering these covariates into our models. Maps of each covariate layer are provided in the supplementary information in Figure 1—figure supplement 1.

Temperature suitability for dengue transmission to humans

via Ae. aegypti or Ae. albopictus: Temperature affects key physiological processes in Aedes mosquitoes, including age- and temperature-dependent adult female survival, as well as the duration of the extrinsic incubation period (EIP) of arboviruses and the length of the gonotrophic cycle (Brady et al., 2013). While these parameters have yet to be measured experimentally for ZIKV, they have been for the closely related DENV. We obtained temperature data from WorldClim v1.03 (http://www.wordclim.org), which uses historic global meteorological station data from 1961–2005 to interpolate global climate surfaces. MARKSIM software (Jones and Thornton, 2000) was then used to apply the coefficients of 17 Global Climate Models (GCMs) to estimate temperature values for the year 2015. This enabled us to incorporate the quantified effects of temperature on DENV transmission into a cohort simulation model that analysed the cumulative effects of both diurnal and inter-seasonal changes in temperature on DENV transmission within an average year, both for Ae. aegypti and Ae. albopictus separately. The models were then applied to the 2015 temperature data for each 5 km x 5 km grid cell globally. This resulted in maps of temperature suitability for DENV transmission by either Aedes species ranging from 0 (no suitable days) to 1 (365 suitable days). These measures were then used as a proxy for temperature suitability for ZIKV transmission to humans.

Annual cumulative precipitation

Presence of static surface water in natural or man-made containers is a pre-requisite for Aedes oviposition and larval and pupal development. While fine-scale spatial and temporal heterogeneities have been observed between precipitation, vector abundance, and incidence of human DENV infections, there is evidence that areas with greater amounts of precipitation are generally associated with higher DENV infection risk (Chandy et al., 2013; Chowell and Sanchez, 2006; Dom et al., 2013; Pinto et al., 2011; Restrepo et al., 2014; Sang et al., 2014; Sankari et al., 2012; Campbell et al., 2015). Although studies that directly connect levels of precipitation to ZIKV transmission have yet to exist, we assumed for Zika a similar association of precipitation as closely related flaviruses. WorldClim v1.03 precipitation data and MARKSIM software were used as described above for temperature, to estimate annual cumulative precipitation for the year 2015 for each 5 km x 5 km grid cell globally.

Minimum relative humidity

Greater relative humidity has been found to promote DENV propagation in Ae. aegypti mosquitoes in several localized settings (Colón-González et al., 2011; Thu et al., 1998), and has also been found to be an important contributor when predicting DENV risk at a global scale (Hales et al., 2002). Therefore, we again assumed a similar association for ZIKV in the absence of any direct studies, and included the minimum annual relative humidity in our models as a potential limiting factor to ZIKV transmission. Relative humidity (RH) was calculated as a percent of saturation humidity, or the amount of water vapour required to saturate the air given a particular temperature, using the temperature data from WorldClim v1.03 described earlier. The saturation, or 'dew', point (Tdew) was calculated using a tabular relationship (Linacre, 1977). RH was then calculated as follows:

RH=V(Tx)V(Tdew)×100

Where V(Tdew) = 611.21×exp(17.502×T240.97+T) and V(Tx) is the humidity at the given temperature. We then extracted the minimum annual RH for each 5 km x 5 km pixel globally for the year 2015.

Enhanced vegetation index (EVI)

A close association has been shown between local moisture supply, vegetation canopy development, and abundance of mosquito reproduction (Linthicum et al., 1999), with previous studies highlighting the importance of moisture-related measures such as relative humidity to DENV occurrence (Hales et al., 2002). Although resistant to desiccation, both Aedes eggs and adults require moisture to survive (Cox et al., 2007; Sota and Mogi, 1992; Reiskind and Lounibos, 2009; Costa et al., 2010; Luz et al., 2008), with low dry season moisture levels substantially affecting Aedes mortality (Russell et al., 2001; Trpis, 1972; Luz et al., 2008). Vegetation canopy cover has previously been associated with higher Aedes larvae density (Fuller et al., 2009; Troyo et al., 2009; Bisset Lazcano et al., 2006; Barrera et al., 2006) by reducing evaporation from containers, decreasing sub-canopy wind speed, and protecting outdoor habitats from direct sunlight. To account for these factors, we included a 5 km x 5 km resolution measure of the EVI derived from NASA’s Moderate Resolution Imaging Spectrometer (MODIS, Boulder, Colorado) imagery (Wan et al., 2002; Lin, 2012), summarized from gap-filled, 8-day, 1 km x 1 km resolution images acquired globally for years 2000 through 2014 (Weiss et al., 2014) to produce a mean annual EVI layer. This mean EVI product is indicative of amount of photosynthesis taking place in the environment over the course of a year, which is positively correlated with the density of vegetation, and is thus a proxy for the level of moisture available given the relationship between precipitation and vegetative growth.

Urban versus rural habitat type

There is a well-established link between urban areas, some vector borne diseases, and their vectors. In particular, Ae. aegypti is found in close proximity to human dwellings often breeding in artificial containers (Brown et al., 2011; Powell and Tabachnick, 2013; Kraemer et al., 2015c). To identify the relationship between urbanisation and ZIKV presence we adapted probabilistic spatial modelling techniques to predict the spatial distribution of global urban extents at a 5 km x 5 km spatial resolution. We used urban growth rates from the United Nations Population Division (Division, 2014), paired with urban extents measured and tested by the Moderate Resolution Imaging Spectroradiometer Collection 5 (MODIS C5) land-cover product for Asia (Schneider et al., 2009; 2010; 2015). A set of spatial covariate datasets hypothesized to influence the spatial patterns of urban expansion was generated, including the time to travel from each 5 km x 5 km pixel to a major city (Nelson, 2008), the proportion of urbanised land within a buffer of 20 km, human population density (Linard and Tatem, 2012; Stevens et al., 2015; Gaughan et al., 2013), slope (Becker et al., 2009), and distance to water (Arino et al., 2008). A BRT modelling approach was then used to predict areas that would become urban in 2015 (Linard et al., 2013). Outputs were tested against a training dataset comprising points from Asia only, and showed good overall predictive performance (AUC=0.82). The output raster is a 5 km x 5 km gridded surface with urban (1) vs. rural (0) pixels.

Ensemble boosted regression trees approach

The boosted regression tree (BRT) modelling procedure combines regression trees with gradient boosting (Friedman, 2001). In this procedure, an initial regression tree is fitted and iteratively improved upon in a forward stagewise manner (boosting) by minimising the variation in the response not explained by the model at each iteration. It has been shown to fit complicated response functions efficiently, while guarding against over-fitting by use of extensive internal cross-validation. As such, this approach has been successfully employed in the past to map dengue and its Aedes mosquito vectors, as well as other vector-borne diseases (Bhatt et al., 2013; Pigott et al., 2014b; Messina et al., 2015b; Kraemer et al., 2015c). To increase the robustness of model predictions and quantify model uncertainty, we fitted an ensemble (Araújo and New, 2007) of 300 BRT models to separate bootstraps of the data. We then evaluated the central tendency as the mean across all 300 BRT models (Bhatt et al., 2013). Each of the 300 individual models was fitted using the gbm.step subroutine in the dismo package in the R statistical programming environment (Elith et al., 2008). All other tuning parameters of the algorithm were held at their default values (tree complexity= 4, learning rate= 0.005, bag fraction= 0.75, step size= 10, cross-validation folds=10). Each of the 300 models predicts environmental suitability on a continuous scale from 0 to 1, with a final prediction map then being generated by calculating the mean prediction across all models for each 5 km x 5 km pixel. Cross-validation was applied to each model, whereby ten subsets of the data comprising 10% of the presence and background observations were assessed based on their ability to predict the distribution of the other 90% of records using the mean area under the curve (AUC) statistic. This AUC value was then averaged across the ten sub-models and finally across all 300 models in the ensemble in order to derive an overall estimate of goodness-of-fit. Additionally, to avoid AUC inflation due to spatial sorting bias, a pairwise distance sampling procedure was used, resulting in a final AUC which is lower than would be returned by standard procedures but which gives a more realistic quantification of the model’s ability to extrapolate predictions to new regions (Wenger and Olden, 2012). We restricted our models to make predictions only within areas where either Ae. aegypti probability of occurrence (Kraemer et al., 2015c) is more than 0.8 or temperature is conducive to transmission for at least 355 days in an average year. A second ensemble of 300 models was executed which did not take into account temperature suitability for dengue transmission via Ae. albopictus, due to the uncertainty of this species as a competent ZIKV vector. The results of this ensemble of models are provided in Figure 2—figure supplement 3.

Population and births at risk

To calculate the number of people located in an area that is at any level of risk for ZIKV transmission, the global ZIKV environmental suitability map was combined with fine-scale global population surfaces (SEDAC, 2015; WorldPop, 2015). Firstly, the continuous ZIKV environmental suitability map (ranging from 0 to 1) was converted into a binary surface indicating whether there is any risk of transmission. To do this, we carried out a protocol previously used in (Pigott et al., 2015a), choosing a threshold environmental suitability value that encompasses 90% of the ZIKV occurrence point locations. This threshold cut-off of 90% was chosen (rather than 100%) to reflect potential errors or inaccurate locations in the occurrence point dataset. Every 5 km x 5 km pixel in the suitability map with a value above this threshold value was considered at risk for ZIKV transmission. Finally, to estimate the population at risk, we multiplied this binary ZIKV risk map by the global population counts (aligned and aggregated to the same 5 x 5 km grid) for the year 2015 and summed across all cells.

We next estimated the maximum number of births potentially affected by ZIKV in Latin America, as this region is the focus of the recent outbreak and the first to point to a possible association with microcephaly in newborn infants to mothers infected with ZIKV. In order to do this, we first identified the proportion of the year that is suitable for ZIKV transmission within areas that are predicted to be suitable in the binary ZIKV risk map. This proportion was derived from existing temperature suitability models (Brady et al., 2013; 2014), which predict the total number of days within an average year that arbovirus transmission can be sustained in Ae. aegypti, assuming there is a local human reservoir of infection. While the intra-mosquito viral dynamics in this model were parameterised for dengue virus, the limited information currently available on other arboviruses suggests that their dynamics are similar (Lambrechts et al., 2011). Using the resulting 5 km x 5 km map showing the proportion of the year suitable for ZIKV transmission to humans, we then multiplied this by a map (also at a 5 km x 5 km resolution) of the number of births in the Americas for the year 2015, updated from (Tatem et al., 2014; UNFPA, 2014). The resulting map indicates the number of births in the Americas potentially at risk for ZIKV (for 2015), assuming ZIKV currently fully occupies its environmental niche and that births are evenly distributed throughout the year.

Acknowledgements

We thank the Secretariat of Health Surveillance, Ministry of Health of Brazil for providing access to the geographical coordinates of occurrence. JPM, MUGK and TJ receive, and OJB and SIH acknowledge funding from the International research Consortium on Dengue Risk Assessment Management and Surveillance (IDAMS; European Commission 7th Framework Programme (21893)). OJB and SIH are supported by the Bill & Melinda Gates Foundation (OPP1053338). SIH is also funded by a Senior Research Fellowship from the Wellcome Trust (095066), and grants from the Bill & Melinda Gates Foundation (OPP1119467, OPP1106023 and OPP1093011). DMP is also funded by the Bill & Melinda Gates Foundation (OPP1093011). DJW and PWG receive support from the Bill and Melinda Gates Foundation (OPP1068048, OPP1106023). NG is supported by a University of Melbourne McKenzie fellowship. CWR is funded through the University of Southampton’s Economic and Social Research Council’s Doctoral Training Centre. TWS is supported by grants from the National Institutes of Health (P01AI098670) and the Bill and Melinda Gates Foundation (OPP1081737). EC and JSB are supported by the National Library of Medicine of the National Institutes of Health (R01LM010812).

Funding Statement

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Funding Information

This paper was supported by the following grants:

  • European Commission 21893 to Jane P Messina, Moritz UG Kraemer, Thomas Jaenisch.

  • Bill and Melinda Gates Foundation OPP1053338; OPP1119467; OPP1106023; OPP1093011; OPP1081737; OPP1068048 to Oliver J Brady, David M Pigott, Daniel J Weiss, Thomas W Scott, Simon I Hay.

  • Wellcome Trust 095066 to Simon I Hay.

  • University of Southampton Economic and Social Research Council's Doctoral Training Centre to Corrine W Ruktanonchai.

  • National Institutes of Health P01AI098670 to Thomas W Scott.

  • University of Melbourne McKenzie fellowship to Nick Golding.

Additional information

Competing interests

SIH: Reviewing editor, eLife.

The other authors declare that no competing interests exist.

Author contributions

JPM, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

MUGK, Conception and design, Analysis and interpretation of data, Drafting or revising the article.

OJB, Conception and design, Analysis and interpretation of data, Drafting or revising the article.

DMP, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

FMS, Conception and design, Acquisition of data, Drafting or revising the article.

DJW, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

NG, Conception and design, Analysis and interpretation of data, Drafting or revising the article.

CWR, Acquisition of data, Analysis and interpretation of data.

PWG, Conception and design, Acquisition of data, Drafting or revising the article.

EC, Acquisition of data, Analysis and interpretation of data.

JSB, Acquisition of data, Drafting or revising the article.

KK, Analysis and interpretation of data, Drafting or revising the article.

AJT, Acquisition of data, Analysis and interpretation of data.

TJ, Analysis and interpretation of data, Drafting or revising the article.

CJLM, Acquisition of data, Drafting or revising the article.

FM, Acquisition of data, Contributed unpublished essential data or reagents.

TWS, Analysis and interpretation of data, Drafting or revising the article.

SIH, Conception and design, Acquisition of data, Analysis and interpretation of data, Drafting or revising the article.

Additional files

Major datasets

The following datasets were generated:

Jane P Messina, Freya M Shearer,2016,Global compendium of human Zika virus occurrence,http://dx.doi.org/10.6084/m9.figshare.2573629,Publicly available at figshare

Jane P Messina, Moritz UG Kraemer, Oliver J Brady, David M Pigott, Freya M Shearer, Daniel J Weiss, Nick Golding, Corrine W Ruktanonchai, Emily Cohn, John S Brownstein, Kamran Khan, Andrew J Tatem, Thomas Jaenisch, Thomas W Scott, Simon I Hay,2016,Environmental suitability for Zika virus transmission,http://dx.doi.org/10.6084/m9.figshare.2574298,Publicly available at figshare

References

  1. Araújo MB, New M. Ensemble forecasting of species distributions. Trends in Ecology & Evolution. 2007;22:42–47. doi: 10.1016/j.tree.2006.09.010. [DOI] [PubMed] [Google Scholar]
  2. Arino O, Bicheron P, Achard F, Latham J, Witt R, Weber JL. GLOBCOVER the most detailed portrait of earth. ESA Bulletin. 2008:24–31. [Google Scholar]
  3. Barbet-Massin M, Jiguet F, Albert CH, Thuiller W. Selecting pseudo-absences for species distribution models: How, where and how many? Methods in Ecology and Evolution. 2012;3:327–338. doi: 10.1111/j.2041-210X.2011.00172.x. [DOI] [Google Scholar]
  4. Barrera R, Amador M, Clark GG. Use of the pupal survey technique for measuring Aedes aegypti (Diptera: Culicidae) productivity in Puerto Rico. The American Journal of Tropical Medicine and Hygiene. 2006;74:290–302. [PubMed] [Google Scholar]
  5. Becker JJ, Sandwell DT, Smith WHF, Braud J, Binder B, Depner J, Fabre D, Factor J, Ingalls S, Kim S-H, Ladner R, Marks K, Nelson S, Pharaoh A, Trimmer R, Von Rosenberg J, Wallace G, Weatherall P. Global bathymetry and elevation data at 30 arc seconds resolution: Srtm30_plus. Marine Geodesy. 2009;32:355–371. doi: 10.1080/01490410903297766. [DOI] [Google Scholar]
  6. Benedict MQ, Levine RS, Hawley WA, Lounibos LP. Spread of the tiger: Global risk of invasion by the mosquito aedes albopictus. Vector Borne and Zoonotic Diseases. 2007;7:76–85. doi: 10.1089/vbz.2006.0562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bhatt S, Gething PW, Brady OJ, Messina JP, Farlow AW, Moyes CL, Drake JM, Brownstein JS, Hoen AG, Sankoh O, Myers MF, George DB, Jaenisch T, Wint GR, Simmons CP, Scott TW, Farrar JJ, Hay SI. The global distribution and burden of dengue. Nature. 2013;496:504–507. doi: 10.1038/nature12060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Bisset Lazcano JA, Marquetti MC, Portillo R, Rodríguez MM, Suárez S, Leyva M. Ecological factors linked to the presence of Aedes aegypti larvae in highly infested areas of Playa, a municipality belonging to Ciudad de La Habana, Cuba. Revista Panamericana de Salud Pública. 2006;19:379–384. doi: 10.1590/S1020-49892006000600003. [DOI] [PubMed] [Google Scholar]
  9. Bogoch II, Brady OJ, Kraemer MU, German M, Creatore MI, Kulkarni MA, Brownstein JS, Mekaru SR, Hay SI, Groot E, Watts A, Khan K. Anticipating the international spread of zika virus from brazil. Lancet. 2016;387:335–336. doi: 10.1016/S0140-6736(16)00080-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Brady OJ, Golding N, Pigott DM, Kraemer MU, Messina JP, Reiner RC, Scott TW, Smith DL, Gething PW, Hay SI. Global temperature constraints on aedes aegypti and ae. albopictus persistence and competence for dengue virus transmission. Parasites & Vectors. 2014;7:338. doi: 10.1186/1756-3305-7-338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Brady OJ, Johansson MA, Guerra CA, Bhatt S, Golding N, Pigott DM, Delatte H, Grech MG, Leisnham PT, Maciel-de-Freitas R, Styer LM, Smith DL, Scott TW, Gething PW, Hay SI. Modelling adult aedes aegypti and aedes albopictus survival at different temperatures in laboratory and field settings. Parasites & Vectors. 2013;6:351. doi: 10.1186/1756-3305-6-351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Brown JE, McBride CS, Johnson P, Ritchie S, Paupy C, Bossin H, Lutomiah J, Fernandez-Salas I, Ponlawat A, Cornel AJ, Black WC, Gorrochotegui-Escalante N, Urdaneta-Marquez L, Sylla M, Slotman M, Murray KO, Walker C, Powell JR. Worldwide patterns of genetic differentiation imply multiple 'domestications' of aedes aegypti, a major vector of human diseases. Proceedings of the Biological Sciences. 2011;278:2446–2454. doi: 10.1098/rspb.2010.2469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Campbell KM, Haldeman K, Lehnig C, Munayco CV, Halsey ES, Laguna-Torres VA, Yagui M, Morrison AC, Lin CD, Scott TW. Weather regulates location, timing, and intensity of dengue virus transmission between humans and mosquitoes. PLoS Neglected Tropical Diseases. 2015;9:e15272. doi: 10.1371/journal.pntd.0003957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Chandy S, Ramanathan K, Manoharan A, Mathai D, Baruah K. Assessing effect of climate on the incidence of dengue in tamil nadu. Indian Journal of Medical Microbiology. 2013;31:283–286. doi: 10.4103/0255-0857.115640. [DOI] [PubMed] [Google Scholar]
  15. Chowell G, Sanchez F. Climate-based descriptive models of dengue fever: the 2002 epidemic in Colima, Mexico. Journal of Environmental Health. 2006;68:40–44. [PubMed] [Google Scholar]
  16. Colón-González FJ, Lake IR, Bentham G. Climate variability and dengue fever in warm and humid mexico. The American Journal of Tropical Medicine and Hygiene. 2011;84:757–763. doi: 10.4269/ajtmh.2011.10-0609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Costa E, Santos EMM, Correia JC, Albuquerque CMR. Impact of small variations in temperature and humidity on the reproductive activity and survival of aedes aegypti (diptera, culicidae) Rev Bras Entomol. 2010;54:488–493. doi: 10.1590/s0085-56262010000300021. [DOI] [Google Scholar]
  18. Cox J, Grillet ME, Ramos OM, Amador M, Barrera R. Habitat segregation of dengue vectors along an urban environmental gradient. The American Journal of Tropical Medicine and Hygiene. 2007;76:820–826. [PubMed] [Google Scholar]
  19. Darwish MA, Hoogstraal H, Roberts TJ, Ahmed IP, Omar F. A sero-epidemiological survey for certain arboviruses (togaviridae) in pakistan. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1983;77:442–445. doi: 10.1016/0035-9203(83)90106-2. [DOI] [PubMed] [Google Scholar]
  20. De'ath G. Boosted trees for ecological modeling and prediction. Ecology. 2007;88:243–251. doi: 10.1890/0012-9658(2007)88[243:btfema]2.0.co;2. [DOI] [PubMed] [Google Scholar]
  21. Dick GW, Kitchen SF, Haddow AJ. Zika virus. I. isolations and serological specificity. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1952;46:509–520. doi: 10.1016/0035-9203(52)90042-4. [DOI] [PubMed] [Google Scholar]
  22. Dick GW Yellow fever; a problem in epidemiology. British Medical Bulletin. 1953;9:215–235. doi: 10.1093/oxfordjournals.bmb.a074360. [DOI] [PubMed] [Google Scholar]
  23. Dom NC, Ahmad AH, Latif ZA, Ismail R, Pradhan B. Coupling of remote sensing data and environmental-related parameters for dengue transmission risk assessment in subang jaya, malaysia. Geocarto International. 2013;28:258–272. doi: 10.1080/10106049.2012.696726. [DOI] [Google Scholar]
  24. Elith J, Leathwick JR, Hastie T. A working guide to boosted regression trees. The Journal of Animal Ecology. 2008;77:802–813. doi: 10.1111/j.1365-2656.2008.01390.x. [DOI] [PubMed] [Google Scholar]
  25. Elith J, Leathwick JR. Species distribution models: Ecological explanation and prediction across space and time. Annu Rev Ecol Evol S. 2009;40:677–697. doi: 10.1146/annurev.ecolsys.110308.120159. [DOI] [Google Scholar]
  26. Faria NR, Azevedo RS, Kraemer MU, Souza R, Cunha MS, Hill SC, Thézé J, Bonsall MB, Bowden TA, Rissanen I, Rocco IM, Nogueira JS, Maeda AY, Vasami FG, Macedo FL, Suzuki A, Rodrigues SG, Cruz AC, Nunes BT, Medeiros DB, Rodrigues DS, Nunes Queiroz AL, da Silva EV, Henriques DF, Travassos da Rosa ES, de Oliveira CS, Martins LC, Vasconcelos HB, Casseb LM, Simith DB, Messina JP, Abade L, Lourenço J, Alcantara LC, de Lima MM, Giovanetti M, Hay SI, de Oliveira RS, Lemos PS, de Oliveira LF, de Lima CP, da Silva SP, de Vasconcelos JM, Franco L, Cardoso JF, Vianez-Júnior JL, Mir D, Bello G, Delatorre E, Khan K, Creatore M, Coelho GE, de Oliveira WK, Tesh R, Pybus OG, Nunes MR, Vasconcelos PF. Zika virus in the americas: Early epidemiological and genetic findings. Science. 2016;352 doi: 10.1126/science.aaf5036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Faye O, Faye O, Dupressoir A, Weidmann M, Ndiaye M, Alpha Sall A. One-step RT-PCR for detection of zika virus. Journal of Clinical Virology. 2008;43:96–101. doi: 10.1016/j.jcv.2008.05.005. [DOI] [PubMed] [Google Scholar]
  28. Foy BD, Kobylinski KC, Foy JLC, Blitvich BJ, Travassos da Rosa A, Haddow AD, Lanciotti RS, Tesh RB. Probable non–vector-borne transmission of zika virus, colorado, USA. Emerging Infectious Diseases. 2011a;17:880–882. doi: 10.3201/eid1705.101939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Foy BD, Kobylinski KC, Foy JLC, Blitvich BJ, Travassos da Rosa A, Haddow AD, Lanciotti RS, Tesh RB. Probable non–vector-borne transmission of zika virus, colorado, USA. Emerging Infectious Diseases. 2011b;17:880–882. doi: 10.3201/eid1705.101939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Freifeld CC, Mandl KD, Reis BY, Brownstein JS. Healthmap: Global infectious disease monitoring through automated classification and visualization of internet media reports. Journal of the American Medical Informatics Association. 2008;15:150–157. doi: 10.1197/jamia.M2544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Friedman JH. Greedy function approximation: A gradient boosting machine. The Annals of Statistics. 2001;29:1189–1232. doi: 10.1214/aos/1013203451. [DOI] [Google Scholar]
  32. Fuller DO, Troyo A, Beier JC. El niño southern oscillation and vegetation dynamics as predictors of dengue fever cases in costa rica. Environmental Research Letters. 2009;4:014011. doi: 10.1088/1748-9326/4/1/014011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Gaughan AE, Stevens FR, Linard C, Jia P, Tatem AJ. High resolution population distribution maps for southeast asia in 2010 and 2015. PloS One. 2013;8:e15272. doi: 10.1371/journal.pone.0055882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Gatherer D, Kohl A. Zika virus: A previously slow pandemic spreads rapidly through the americas. The Journal of General Virology. 2016;97:269–342. doi: 10.1099/jgv.0.000381. [DOI] [PubMed] [Google Scholar]
  35. Grard G, Caron M, Mombo IM, Nkoghe D, Mboui Ondo S, Jiolle D, Fontenille D, Paupy C, Leroy EM. Zika virus in gabon (central africa)--2007: A new threat from aedes albopictus? PLoS Neglected Tropical Diseases. 2014;8:e15272. doi: 10.1371/journal.pntd.0002681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Gubler DJ, Clark GG. Dengue/dengue hemorrhagic fever: The emergence of a global health problem. Emerging Infectious Diseases. 1995;1:55–57. doi: 10.3201/eid0102.952004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Haddow AD, Schuh AJ, Yasuda CY, Kasper MR, Heang V, Huy R, Guzman H, Tesh RB, Weaver SC. Genetic characterization of zika virus strains: Geographic expansion of the asian lineage. PLoS Neglected Tropical Diseases. 2012;6 doi: 10.1371/journal.pntd.0001477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Hales S, de Wet N, Maindonald J, Woodward A. Potential effect of population and climate changes on global distribution of dengue fever: An empirical model. Lancet. 2002;360:830–834. doi: 10.1016/S0140-6736(02)09964-6. [DOI] [PubMed] [Google Scholar]
  39. Hay SI, Battle KE, Pigott DM, Smith DL, Moyes CL, Bhatt S, Brownstein JS, Collier N, Myers MF, George DB, Gething PW. Global mapping of infectious disease. Philosophical Transactions of the Royal Society B. 2013;368:20120250. doi: 10.1098/rstb.2012.0250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Hay SI, Tatem AJ, Graham AJ, Goetz SJ, Rogers DJ. Global environmental data for mapping infectious disease distribution. Advances in Parasitology. 2006;62:37–77. doi: 10.1016/S0065-308X(05)62002-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Hennessey M, Fischer M, Staples JE. Zika virus spreads to new areas - region of the americas, may 2015-january 2016. Morb Mortal Wkly Rep. 2016;65:55–58. doi: 10.1111/ajt.13743. [DOI] [PubMed] [Google Scholar]
  42. Heymann DL, Hodgson A, Sall AA, Freedman DO, Staples JE, Althabe F, Baruah K, Mahmud G, Kandun N, Vasconcelos PFC, Bino S, Menon KU. Zika virus and microcephaly: Why is this situation a PHEIC? The Lancet. 2016;387:719–721. doi: 10.1016/S0140-6736(16)00320-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. ICTV . Virus taxonomy: 2014 release. 1st Edition 2014. [Google Scholar]
  44. Jones PG, Thornton PK. Marksim: Software to generate daily weather data for Latin America and Africa. Agronomy Journal. 2000;92:445–453. doi: 10.2134/agronj2000.923445x. [DOI] [Google Scholar]
  45. Kraemer MU, Hay SI, Pigott DM, Smith DL, Wint GR, Golding N. Progress and challenges in infectious disease cartography. Trends in Parasitology. 2016;32:19–29. doi: 10.1016/j.pt.2015.09.006. [DOI] [PubMed] [Google Scholar]
  46. Kraemer MUG, Perkins TA, Cummings DAT, Zakar R, Hay SI, Smith DL, Reiner RC. Big city, small world: Density, contact rates, and transmission of dengue across pakistan. Journal of the Royal Society Interface. 2015a;12:20150468. doi: 10.1098/rsif.2015.0468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Kraemer MUG, Sinka ME, Duda KA, Mylne A, Shearer FM, Brady OJ, Messina JP, Barker CM, Moore CG, Carvalho RG, Coelho GE, Van Bortel W, Hendrickx G, Schaffner F, Wint GRW, Elyazar IRF, Teng H-J, Hay SI. The global compendium of aedes aegypti and ae. albopictus occurrence. Scientific Data. 2015b;2:150035. doi: 10.1038/sdata.2015.35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Kraemer MUG, Sinka ME, Duda KA, Mylne AQN, Shearer FM, Barker CM, Moore CG, Carvalho RG, Coelho GE, Van Bortel W, Hendrickx G, Schaffner F, Elyazar IRF, Teng H-J, Brady OJ, Messina JP, Pigott DM, Scott TW, Smith DL, Wint GRW, Golding N, Hay SI. The global distribution of the arbovirus vectors aedes aegypti and ae. albopictus. eLife. 2015c;4:e15272. doi: 10.7554/eLife.08347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Kutsuna S, Kato Y, Moi ML, Kotaki A, Ota M, Shinohara K, Kobayashi T, Yamamoto K, Fujiya Y, Mawatari M, Sato T, Kunimatsu J, Takeshita N, Hayakawa K, Kanagawa S, Takasaki T, Ohmagari N. Autochthonous dengue fever, tokyo, japan, 2014. Emerging Infectious Diseases. 2015;21:517. doi: 10.3201/eid2103.141662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Lambrechts L, Paaijmans KP, Fansiri T, Carrington LB, Kramer LD, Thomas MB, Scott TW. Impact of daily temperature fluctuations on dengue virus transmission by aedes aegypti. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:7460–7465. doi: 10.1073/pnas.1101377108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Lanciotti RS, Kosoy OL, Laven JJ, Velez JO, Lambert AJ, Johnson AJ, Stanfield SM, Duffy MR. Genetic and serologic properties of zika virus associated with an epidemic, yap state, micronesia, 2007. Emerging Infectious Diseases. 2008;14:1232–1241. doi: 10.3201/eid1408.080287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Lin QH. Enhanced vegetation index using moderate resolution imaging spectroradiometers. In 5th International Congress on Image and Signal Processing . 2012:1043–1046. doi: 10.1109/cisp.2012.6470008. [DOI] [Google Scholar]
  53. Linacre ET. Simple formula for estimating evaporation rates in various climates using temperature data alone. Agr Meteorol. 1977;18:409–424. doi: 10.1016/0002-1571(77)90007-3. [DOI] [Google Scholar]
  54. Linard C, Tatem AJ, Gilbert M. Modelling spatial patterns of urban growth in africa. Applied Geography. 2013;44:23–32. doi: 10.1016/j.apgeog.2013.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Linard C, Tatem AJ. Large-scale spatial population databases in infectious disease research. Int J Health Geogr. 2012;11 doi: 10.1186/1476-072x-11-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Linthicum KJ, Anyamba A, Tucker CJ, Kelley PW, Myers MF, Peters CJ. Climate and satellite indicators to forecast rift valley fever epidemics in kenya. Science. 1999;285:397–400. doi: 10.1126/science.285.5426.397. [DOI] [PubMed] [Google Scholar]
  57. Luz C, Tai MH, Santos AH, Silva HH. Impact of moisture on survival of aedes aegypti eggs and ovicidal activity of metarhizium anisopliae under laboratory conditions. MemóRias Do Instituto Oswaldo Cruz. 2008;103:214–215. doi: 10.1590/s0074-02762008000200016. [DOI] [PubMed] [Google Scholar]
  58. Marchette NJ, Garcia R, Rudnick A. Isolation of Zika virus from Aedes aegypti mosquitoes in Malaysia. The American Journal of Tropical Medicine and Hygiene. 1969;18:411–415. doi: 10.4269/ajtmh.1969.18.411. [DOI] [PubMed] [Google Scholar]
  59. Marcondes CB, Ximenes MF. Zika virus in brazil and the danger of infestation by aedes (stegomyia) mosquitoes. Revista Da Sociedade Brasileira De Medicina Tropical. 2015 doi: 10.1590/0037-8682-0220-2015. [DOI] [PubMed] [Google Scholar]
  60. Martines RB, Bhatnagar J, Keating MK, Silva-Flannery L, Muehlenbachs A, Gary J, Goldsmith C, Hale G, Ritter J, Rollin D, Shieh WJ, Luz KG, Ramos AM, Davi HP, 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. Morbidity and Mortality Weekly Report. 2016;65:159–219. doi: 10.15585/mmwr.mm6506e1. [DOI] [PubMed] [Google Scholar]
  61. McCarthy M. First US case of zika virus infection is identified in texas. BMJ. 2016;352:i212. doi: 10.1136/bmj.i212. [DOI] [PubMed] [Google Scholar]
  62. Messina JP, Brady OJ, Pigott DM, Brownstein JS, Hoen AG, Hay SI. A global compendium of human dengue virus occurrence. Scientific Data. 2014;1:140004. doi: 10.1038/sdata.2014.4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Messina JP, Pigott DM, Duda KA, Brownstein JS, Myers MF, George DB, Hay SI. A global compendium of human crimean-congo haemorrhagic fever virus occurrence. Scientific Data. 2015a;2:150016. doi: 10.1038/sdata.2015.16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Messina JP, Pigott DM, Golding N, Duda KA, Brownstein JS, Weiss DJ, Gibson H, Robinson TP, Gilbert M, William Wint GR, Nuttall PA, Gething PW, Myers MF, George DB, Hay SI. The global distribution of crimean-congo hemorrhagic fever. Transactions of the Royal Society of Tropical Medicine and Hygiene. 2015b;109:503–513. doi: 10.1093/trstmh/trv050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Monlun E, Zeller H, Le Guenno B, Traore-Lamizana M, Hervy J, 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. Bulletin de la Société de pathologie exotique. 1992;86:21–28. [PubMed] [Google Scholar]
  66. Morrison AC, Zielinski-Gutierrez E, Scott TW, Rosenberg R. Defining challenges and proposing solutions for control of the virus vector aedes aegypti. PLoS Medicine. 2008;5:e15272. doi: 10.1371/journal.pmed.0050068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Murray CJ, Barber RM, Foreman KJ, Abbasoglu Ozgoren A, Abd-Allah F, Abera SF, Aboyans V, Abraham JP, Abubakar I, Abu-Raddad LJ, Abu-Rmeileh NM, Achoki T, Ackerman IN, Ademi Z, Adou AK, Adsuar JC, Afshin A, Agardh EE, Alam SS, Alasfoor D, Albittar MI, Alegretti MA, Alemu ZA, Alfonso-Cristancho R, Alhabib S, Ali R, Alla F, Allebeck P, Almazroa MA, Alsharif U, Alvarez E, Alvis-Guzman N, Amare AT, Ameh EA, Amini H, Ammar W, Anderson HR, Anderson BO, Antonio CA, Anwari P, Arnlöv J, Arsic Arsenijevic VS, Artaman A, Asghar RJ, Assadi R, Atkins LS, Avila MA, Awuah B, Bachman VF, Badawi A, Bahit MC, Balakrishnan K, Banerjee A, Barker-Collo SL, Barquera S, Barregard L, Barrero LH, Basu A, Basu S, Basulaiman MO, Beardsley J, Bedi N, Beghi E, Bekele T, Bell ML, Benjet C, Bennett DA, Bensenor IM, Benzian H, Bernabé E, Bertozzi-Villa A, Beyene TJ, Bhala N, Bhalla A, Bhutta ZA, Bienhoff K, Bikbov B, Biryukov S, Blore JD, Blosser CD, Blyth FM, Bohensky MA, Bolliger IW, Bora Başara B, Bornstein NM, Bose D, Boufous S, Bourne RR, Boyers LN, Brainin M, Brayne CE, Brazinova A, Breitborde NJ, Brenner H, Briggs AD, Brooks PM, Brown JC, Brugha TS, Buchbinder R, Buckle GC, Budke CM, Bulchis A, Bulloch AG, Campos-Nonato IR, Carabin H, Carapetis JR, Cárdenas R, Carpenter DO, Caso V, Castañeda-Orjuela CA, Castro RE, Catalá-López F, Cavalleri F, Çavlin A, Chadha VK, Chang JC, Charlson FJ, Chen H, Chen W, Chiang PP, Chimed-Ochir O, Chowdhury R, Christensen H, Christophi CA, Cirillo M, Coates MM, Coffeng LE, Coggeshall MS, Colistro V, Colquhoun SM, Cooke GS, Cooper C, Cooper LT, Coppola LM, Cortinovis M, Criqui MH, Crump JA, Cuevas-Nasu L, Danawi H, Dandona L, Dandona R, Dansereau E, Dargan PI, Davey G, Davis A, Davitoiu DV, Dayama A, De Leo D, Degenhardt L, Del Pozo-Cruz B, Dellavalle RP, Deribe K, Derrett S, Des Jarlais DC, Dessalegn M, Dharmaratne SD, Dherani MK, Diaz-Torné C, Dicker D, Ding EL, Dokova K, Dorsey ER, Driscoll TR, Duan L, Duber HC, Ebel BE, Edmond KM, Elshrek YM, Endres M, Ermakov SP, Erskine HE, Eshrati B, Esteghamati A, Estep K, Faraon EJ, Farzadfar F, Fay DF, Feigin VL, Felson DT, Fereshtehnejad SM, Fernandes JG, Ferrari AJ, Fitzmaurice C, Flaxman AD, Fleming TD, Foigt N, Forouzanfar MH, Fowkes FG, Paleo UF, Franklin RC, Fürst T, Gabbe B, Gaffikin L, Gankpé FG, Geleijnse JM, Gessner BD, Gething P, Gibney KB, Giroud M, Giussani G, Gomez Dantes H, Gona P, González-Medina D, Gosselin RA, Gotay CC, Goto A, Gouda HN, Graetz N, Gugnani HC, Gupta R, Gupta R, Gutiérrez RA, Haagsma J, Hafezi-Nejad N, Hagan H, Halasa YA, Hamadeh RR, Hamavid H, Hammami M, Hancock J, Hankey GJ, Hansen GM, Hao Y, Harb HL, Haro JM, Havmoeller R, Hay SI, Hay RJ, Heredia-Pi IB, Heuton KR, Heydarpour P, Higashi H, Hijar M, Hoek HW, Hoffman HJ, Hosgood HD, Hossain M, Hotez PJ, Hoy DG, Hsairi M, Hu G, Huang C, Huang JJ, Husseini A, Huynh C, Iannarone ML, Iburg KM, Innos K, Inoue M, Islami F, Jacobsen KH, Jarvis DL, Jassal SK, Jee SH, Jeemon P, Jensen PN, Jha V, Jiang G, Jiang Y, Jonas JB, Juel K, Kan H, Karch A, Karema CK, Karimkhani C, Karthikeyan G, Kassebaum NJ, Kaul A, Kawakami N, Kazanjan K, Kemp AH, Kengne AP, Keren A, Khader YS, Khalifa SE, Khan EA, Khan G, Khang YH, Kieling C, Kim D, Kim S, Kim Y, Kinfu Y, Kinge JM, Kivipelto M, Knibbs LD, Knudsen AK, Kokubo Y, Kosen S, Krishnaswami S, Kuate Defo B, Kucuk Bicer B, Kuipers EJ, Kulkarni C, Kulkarni VS, Kumar GA, Kyu HH, Lai T, Lalloo R, Lallukka T, Lam H, Lan Q, Lansingh VC, Larsson A, Lawrynowicz AE, Leasher JL, Leigh J, Leung R, Levitz CE, Li B, Li Y, Li Y, Lim SS, Lind M, Lipshultz SE, Liu S, Liu Y, Lloyd BK, Lofgren KT, Logroscino G, Looker KJ, Lortet-Tieulent J, Lotufo PA, Lozano R, Lucas RM, Lunevicius R, Lyons RA, Ma S, Macintyre MF, Mackay MT, Majdan M, Malekzadeh R, Marcenes W, Margolis DJ, Margono C, Marzan MB, Masci JR, Mashal MT, Matzopoulos R, Mayosi BM, Mazorodze TT, Mcgill NW, Mcgrath JJ, Mckee M, Mclain A, Meaney PA, Medina C, Mehndiratta MM, Mekonnen W, Melaku YA, Meltzer M, Memish ZA, Mensah GA, Meretoja A, Mhimbira FA, Micha R, Miller TR, Mills EJ, Mitchell PB, Mock CN, Mohamed Ibrahim N, Mohammad KA, Mokdad AH, Mola GL, Monasta L, Montañez Hernandez JC, Montico M, Montine TJ, Mooney MD, Moore AR, Moradi-Lakeh M, Moran AE, Mori R, Moschandreas J, Moturi WN, Moyer ML, Mozaffarian D, Msemburi WT, Mueller UO, Mukaigawara M, Mullany EC, Murdoch ME, Murray J, Murthy KS, Naghavi M, Naheed A, Naidoo KS, Naldi L, Nand D, Nangia V, Narayan KM, Nejjari C, Neupane SP, Newton CR, Ng M, Ngalesoni FN, Nguyen G, Nisar MI, Nolte S, Norheim OF, Norman RE, Norrving B, Nyakarahuka L, Oh IH, Ohkubo T, Ohno SL, Olusanya BO, Opio JN, Ortblad K, Ortiz A, Pain AW, Pandian JD, Panelo CI, Papachristou C, Park EK, Park JH, Patten SB, Patton GC, Paul VK, Pavlin BI, Pearce N, Pereira DM, Perez-Padilla R, Perez-Ruiz F, Perico N, Pervaiz A, Pesudovs K, Peterson CB, Petzold M, Phillips MR, Phillips BK, Phillips DE, Piel FB, Plass D, Poenaru D, Polinder S, Pope D, Popova S, Poulton RG, Pourmalek F, Prabhakaran D, Prasad NM, Pullan RL, Qato DM, Quistberg DA, Rafay A, Rahimi K, Rahman SU, Raju M, Rana SM, Razavi H, Reddy KS, Refaat A, Remuzzi G, Resnikoff S, Ribeiro AL, Richardson L, Richardus JH, Roberts DA, Rojas-Rueda D, Ronfani L, Roth GA, Rothenbacher D, Rothstein DH, Rowley JT, Roy N, Ruhago GM, Saeedi MY, Saha S, Sahraian MA, Sampson UK, Sanabria JR, Sandar L, Santos IS, Satpathy M, Sawhney M, Scarborough P, Schneider IJ, Schöttker B, Schumacher AE, Schwebel DC, Scott JG, Seedat S, Sepanlou SG, Serina PT, Servan-Mori EE, Shackelford KA, Shaheen A, Shahraz S, Shamah Levy T, Shangguan S, She J, Sheikhbahaei S, Shi P, Shibuya K, Shinohara Y, Shiri R, Shishani K, Shiue I, Shrime MG, Sigfusdottir ID, Silberberg DH, Simard EP, Sindi S, Singh A, Singh JA, Singh L, Skirbekk V, Slepak EL, Sliwa K, Soneji S, Søreide K, Soshnikov S, Sposato LA, Sreeramareddy CT, Stanaway JD, Stathopoulou V, Stein DJ, Stein MB, Steiner C, Steiner TJ, Stevens A, Stewart A, Stovner LJ, Stroumpoulis K, Sunguya BF, Swaminathan S, Swaroop M, Sykes BL, Tabb KM, Takahashi K, Tandon N, Tanne D, Tanner M, Tavakkoli M, Taylor HR, Te Ao BJ, Tediosi F, Temesgen AM, Templin T, Ten Have M, Tenkorang EY, Terkawi AS, Thomson B, Thorne-Lyman AL, Thrift AG, Thurston GD, Tillmann T, Tonelli M, Topouzis F, Toyoshima H, Traebert J, Tran BX, Trillini M, Truelsen T, Tsilimbaris M, Tuzcu EM, Uchendu US, Ukwaja KN, Undurraga EA, Uzun SB, Van Brakel WH, Van De Vijver S, van Gool CH, Van Os J, Vasankari TJ, Venketasubramanian N, Violante FS, Vlassov VV, Vollset SE, Wagner GR, Wagner J, Waller SG, Wan X, Wang H, Wang J, Wang L, Warouw TS, Weichenthal S, Weiderpass E, Weintraub RG, Wenzhi W, Werdecker A, Westerman R, Whiteford HA, Wilkinson JD, Williams TN, Wolfe CD, Wolock TM, Woolf AD, Wulf S, Wurtz B, Xu G, Yan LL, Yano Y, Ye P, Yentür GK, Yip P, Yonemoto N, Yoon SJ, Younis MZ, Yu C, Zaki ME, Zhao Y, Zheng Y, Zonies D, Zou X, Salomon JA, Lopez AD, Vos T, GBD 2013 DALYs and HALE Collaborators Global, regional, and national disability-adjusted life years (dalys) for 306 diseases and injuries and healthy life expectancy (HALE) for 188 countries, 1990-2013: Quantifying the epidemiological transition. Lancet . 2015;386:2145–2191. doi: 10.1016/S0140-6736(15)61340-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Murray NE, Quam MB, Wilder-Smith A. Epidemiology of dengue: Past, present and future prospects. Clinical Epidemiology. 2013;5:299–309. doi: 10.2147/CLEP.S34440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Musso D, Nilles EJ, Cao-Lormeau V-M. Rapid spread of emerging zika virus in the pacific area. Clinical Microbiology and Infection. 2014;20:O595–O596. doi: 10.1111/1469-0691.12707. [DOI] [PubMed] [Google Scholar]
  70. Musso D, Cao-Lormeau VM, Gubler DJ. Zika virus: Following the path of dengue and chikungunya? Lancet. 2015a;386:243–247. doi: 10.1016/S0140-6736(15)61273-9. [DOI] [PubMed] [Google Scholar]
  71. Musso D, Roche C, Robin E, Nhan T, Teissier A, Cao-Lormeau VM. Potential sexual transmission of zika virus. Emerging Infectious Diseases. 2015b;21:359–361. doi: 10.3201/eid2102.141363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Mylne AQ, Pigott DM, Longbottom J, Shearer F, Duda KA, Messina JP, Weiss DJ, Moyes CL, Golding N, Hay SI. Mapping the zoonotic niche of lassa fever in africa. Transactions of the Royal Society of Tropical Medicine and Hygiene. 2015;109:483–492. doi: 10.1093/trstmh/trv047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Nelson A. Travel Time to Major Cities: A Global Map of Accessibility. Ispra, Italy: GEMU-JRCOTE Commission; 2008. [Google Scholar]
  74. Nunes MR, Faria NR, de Vasconcelos JM, Golding N, Kraemer MU, de Oliveira LF, Azevedo RS, da Silva DE, da Silva EV, da Silva SP, Carvalho VL, Coelho GE, Cruz AC, Rodrigues SG, Vianez JL, Nunes BT, Cardoso JF, Tesh RB, Hay SI, Pybus OG, Vasconcelos PF. Emergence and potential for spread of chikungunya virus in brazil. BMC Medicine. 2015;13 doi: 10.1186/s12916-015-0348-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Patiño-Barbosa AM, Medina I, Gil-Restrepo AF, Rodriguez-Morales AJ. Zika: Another sexually transmitted infection? Sexually Transmitted Infections. 2015;91 doi: 10.1136/sextrans-2015-052189. [DOI] [PubMed] [Google Scholar]
  76. Phillips SJ, Dudík M, Elith J, Graham CH, Lehmann A, Leathwick J, Ferrier S. Sample selection bias and presence-only distribution models: Implications for background and pseudo-absence data. Ecological Applications. 2009;19:181–197. doi: 10.1890/07-2153.1. [DOI] [PubMed] [Google Scholar]
  77. Pigott DM, Bhatt S, Golding N, Duda KA, Battle KE, Brady OJ, Messina JP, Balard Y, Bastien P, Pratlong F, Brownstein JS, Freifeld CC, Mekaru SR, Gething PW, George DB, Myers MF, Reithinger R, Hay SI. Global distribution maps of the leishmaniases. eLife. 2014b;3 doi: 10.7554/eLife.02851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Pigott DM, Bhatt S, Golding N, Duda KA, Battle KE, Brady OJ, Messina JP, Balard Y, Bastien P, Pratlong F, Brownstein JS, Freifeld CC, Mekaru SR, Gething PW, George DB, Myers MF, Reithinger R, Hay SI. Global distribution maps of the leishmaniases. eLife. 3 doi: 10.7554/eLife.02851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Pigott DM, Golding N, Mylne A, Huang Z, Weiss DJ, Brady OJ, Kraemer MU, Hay SI. Mapping the zoonotic niche of marburg virus disease in africa. Transactions of the Royal Society of Tropical Medicine and Hygiene. 2015a;109:366–378. doi: 10.1093/trstmh/trv024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Pigott DM, Howes RE, Wiebe A, Battle KE, Golding N, Gething PW, Dowell SF, Farag TH, Garcia AJ, Kimball AM, Krause LK, Smith CH, Brooker SJ, Kyu HH, Vos T, Murray CJL, Moyes CL, Hay SI. Prioritising infectious disease mapping. PLOS Neglected Tropical Diseases. 2015b;9:e15272. doi: 10.1371/journal.pntd.0003756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Pinto E, Coelho M, Oliver L, Massad E. The influence of climate variables on dengue in singapore. International Journal of Environmental Health Research. 2011;21:415–426. doi: 10.1080/09603123.2011.572279. [DOI] [PubMed] [Google Scholar]
  82. Pond WL. ARTHROPOD-borne virus antibodies in sera from residents of south-east asia. Transactions of the Royal Society of Tropical Medicine and Hygiene. 1963;57:364–371. doi: 10.1016/0035-9203(63)90100-7. [DOI] [PubMed] [Google Scholar]
  83. Powell JR, Tabachnick WJ. History of domestication and spread of aedes aegypti--a review. MemóRias Do Instituto Oswaldo Cruz. 2013;108 Suppl 1:11–17. doi: 10.1590/0074-0276130395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Reiskind MH, Lounibos LP. Effects of intraspecific larval competition on adult longevity in the mosquitoes aedes aegypti and aedes albopictus. Medical and Veterinary Entomology. 2009;23:62–68. doi: 10.1111/j.1365-2915.2008.00782.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Restrepo AC, Baker P, Clements AC. National spatial and temporal patterns of notified dengue cases, colombia 2007-2010. Tropical Medicine & International Health. 2014;19:863–871. doi: 10.1111/tmi.12325. [DOI] [PubMed] [Google Scholar]
  86. Roiz D, Boussès P, Simard F, Paupy C, Fontenille D. Autochthonous chikungunya transmission and extreme climate events in southern france. PLoS Neglected Tropical Diseases. 2015;9:e15272. doi: 10.1371/journal.pntd.0003854. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Russell BM, Kay BH, Shipton W. Survival of aedes aegypti (diptera: Culicidae) eggs in surface and subterranean breeding sites during the northern queensland dry season. Journal of Medical Entomology. 2001;38:441–445. doi: 10.1603/0022-2585-38.3.441. [DOI] [PubMed] [Google Scholar]
  88. Salje H, Cauchemez S, Alera MT, Rodriguez-Barraquer I, Thaisomboonsuk B, Srikiatkhachorn A, Lago CB, Villa D, Klungthong C, Tac-An IA, Fernandez S, Velasco JM, Roque VG, Nisalak A, Macareo LR, Levy JW, Cummings D, Yoon IK. Reconstruction of 60 years of chikungunya epidemiology in the philippines demonstrates episodic and focal transmission. The Journal of Infectious Diseases. 2016;213:604–610. doi: 10.1093/infdis/jiv470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Sang S, Yin W, Bi P, Zhang H, Wang C, Liu X, Chen B, Yang W, Liu Q. Predicting local dengue transmission in guangzhou, china, through the influence of imported cases, mosquito density and climate variability. PloS One. 2014;9:e15272. doi: 10.1371/journal.pone.0102755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Sankari T, Hoti SL, Singh TB, Shanmugavel J. Outbreak of dengue virus serotype-2 (DENV-2) of Cambodian origin in Manipur, India - association with meteorological factors. The Indian Journal of Medical Research. 2012;136:649–655. [PMC free article] [PubMed] [Google Scholar]
  91. Schneider A, Friedl MA, Potere D. A new map of global urban extent from MODIS satellite data. Environmental Research Letters. 2009;4:044003. doi: 10.1088/1748-9326/4/4/044003. [DOI] [Google Scholar]
  92. Schneider A, Friedl MA, Potere D. Mapping global urban areas using MODIS 500-m data: New methods and datasets based on ‘urban ecoregions’. Remote Sensing of Environment. 2010;114:1733–1746. doi: 10.1016/j.rse.2010.03.003. [DOI] [Google Scholar]
  93. Schneider A, Mertes CM, Tatem AJ, Tan B, Sulla-Menashe D, Graves SJ, Patel NN, Horton JA, Gaughan AE, Rollo JT, Schelly IH, Stevens FR, Dastur A. A new urban landscape in east–southeast asia, 2000–2010. Environmental Research Letters. 2015;10:034002. doi: 10.1088/1748-9326/10/3/034002. [DOI] [Google Scholar]
  94. Scott TW, Takken W. Feeding strategies of anthropophilic mosquitoes result in increased risk of pathogen transmission. Trends in Parasitology. 2012;28:114–121. doi: 10.1016/j.pt.2012.01.001. [DOI] [PubMed] [Google Scholar]
  95. Socio-economic Data and Applications Center Gridded Population of the World, v4 (GPWv4) 2015.
  96. Smithburn K. Neutralizing antibodies against arthropod-borne viruses in the sera of long-time residents of Malaya and Borneo. American Journal of Hygiene. 1954a;59:157–163. doi: 10.1093/oxfordjournals.aje.a119630. [DOI] [PubMed] [Google Scholar]
  97. Smithburn KC, Taylor RM, Rizk F, Kader A. Immunity to certain arthropod-borne viruses among indigenous residents of Egypt. The American Journal of Tropical Medicine and Hygiene. 1954b;3:9–18. doi: 10.4269/ajtmh.1954.3.9. [DOI] [PubMed] [Google Scholar]
  98. Sorichetta A, Hornby GM, Stevens FR, Gaughan AE, Linard C, Tatem AJ. High-resolution gridded population datasets for latin america and the caribbean in 2010, 2015, and 2020. Scientific Data. 2015;2 doi: 10.1038/sdata.2015.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Sota T, Mogi M. Interspecific variation in desiccation survival time of aedes (stegomyia) mosquito eggs is correlated with habitat and egg size. Oecologia. 1992;90:353–358. doi: 10.1007/BF00317691. [DOI] [PubMed] [Google Scholar]
  100. Stevens FR, Gaughan AE, Linard C, Tatem AJ. Disaggregating census data for population mapping using random forests with remotely-sensed and ancillary data. PloS One. 2015;10:e15272. doi: 10.1371/journal.pone.0107042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Stoddard ST, Forshey BM, Morrison AC, Paz-Soldan VA, Vazquez-Prokopec GM, Astete H, Reiner RC, Vilcarromero S, Elder JP, Halsey ES, Kochel TJ, Kitron U, Scott TW. House-to-house human movement drives dengue virus transmission. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:994–999. doi: 10.1073/pnas.1213349110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Tatem AJ, Campbell J, Guerra-Arias M, de Bernis L, Moran A, Matthews Z. Mapping for maternal and newborn health: The distributions of women of childbearing age, pregnancies and births. International Journal of Health Geographics. 2014;13 doi: 10.1186/1476-072X-13-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Thu HM, Aye KM, Thein S. The effect of temperature and humidity on dengue virus propagation in Aedes aegypti mosquitos. The Southeast Asian Journal of Tropical Medicine and Public Health. 1998;29:280–284. [PubMed] [Google Scholar]
  104. Troyo A, Fuller DO, Calderón-Arguedas O, Solano ME, Beier JC. Urban structure and dengue fever in puntarenas, costa rica. Singapore Journal of Tropical Geography. 2009;30:265–282. doi: 10.1111/j.1467-9493.2009.00367.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Trpis M. Dry season survival of Aedes aegypti eggs in various breeding sites in the Dar es Salaam area, Tanzania. Bulletin of the World Health Organization. 1972;47 [PMC free article] [PubMed] [Google Scholar]
  106. United Nations . World Urbanization Prospects: The 2014 Revision. New York; 2014. [Google Scholar]
  107. UNFPA . The State of the World’s Midwifery 2014: A Universal Pathway. A Woman's Right to Health. UNFPA; 2014. pp. 1–228. [Google Scholar]
  108. Wan ZM, Zhang YL, Zhan QC, Li ZL. The MODIS land-surface temperature products for regional environmental monitoring and global change studies. Int Geosci Remote Se. 2002:3683–3685. doi: 10.1109/igarss.2002.1027290. [DOI] [Google Scholar]
  109. Weaver SC, Reisen WK. Present and future arboviral threats. Antiviral Research. 2010;85:328–345. doi: 10.1016/j.antiviral.2009.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Weiss DJ, Atkinson PM, Bhatt S, Mappin B, Hay SI, Gething PW. An effective approach for gap-filling continental scale remotely sensed time-series. ISPRS Journal of Photogrammetry and Remote Sensing. 2014;98:106–118. doi: 10.1016/j.isprsjprs.2014.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Wenger SJ, Olden JD. Assessing transferability of ecological models: An underappreciated aspect of statistical validation. Methods in Ecology and Evolution. 2012;3:260–267. doi: 10.1111/j.2041-210X.2011.00170.x. [DOI] [Google Scholar]
  112. WHO Collaborating Center for Reference and Research on Arboviruses and Hemorrhagic Fever Viruses: Annual Report . Dakar, Senegal: 1999. p. 143. [Google Scholar]
  113. Woolhouse M, Scott F, Hudson Z, Howey R, Chase-Topping M. Human viruses: Discovery and emergence. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 2012;367:2864–2871. doi: 10.1098/rstb.2011.0354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. World Health Organisation Zika virus infection – Brazil and Colombia. http://www.who.int/csr/don/21-october-2015-zika/en/ 2015
  115. WorldPop . High Resolution Age-Structured Population Distribution Maps. U. O. S. GeoData Institute. 2015. [Google Scholar]
  116. Zanluca C, Melo VC, Mosimann AL, Santos GI, Santos CN, Luz K. First report of autochthonous transmission of zika virus in brazil. MemóRias Do Instituto Oswaldo Cruz. 2015;110:569–572. doi: 10.1590/0074-02760150192. [DOI] [PMC free article] [PubMed] [Google Scholar]
eLife. 2016 Apr 19;5:e15272. doi: 10.7554/eLife.15272.016

Decision letter

Editor: Mark Jit1

In the interests of transparency, eLife includes the editorial decision letter and accompanying author responses. A lightly edited version of the letter sent to the authors after peer review is shown, indicating the most substantive concerns; minor comments are not usually included.

Thank you for submitting your article "Mapping global environmental suitability for Zika virus" for consideration by eLife. Your article has been reviewed by three peer reviewers, one of whom, Mark Jit, is a member of our Board of Reviewing Editors, and another is Jean-Paul Chretien. The evaluation has been overseen by Prabhat Jha as the Senior Editor.

The reviewers have discussed the reviews with one another and the Reviewing Editor has drafted this decision to help you prepare a revised submission.

The reviewers all felt that this was a timely analysis that provides important first estimates that are needed in light of the ongoing spread of Zika virus into new areas via infected travelers. They also commended the manuscript for increasing the epidemiological relevance by translating global temperatures into model-based estimates of suitability for dengue transmission as a surrogate for Zika virus. These are imperfect and based on static, climatic summaries of conditions that vary seasonally, but are a reasonable choice given the gaps in our present knowledge base. The reviewers also appreciated the open sharing of the datasets used.

However, the reviewers had two major concerns about the manuscript:

a) The definition of limits where "either Ae. aegypti probability of occurrence is more than 0.8 or temperature is conducive to transmission via this mosquito for at least 355 days in an average year" seems quite subjective and presumably helped to keep the mapped values from straying too far from known distributional limits. The reviewers would like you to indicate the objective criteria used to define these limits. Since any criteria are unlikely to be completely indisputable, they suggest that you consider providing some measure of uncertainty for the predicted suitability scores e.g. perhaps the distribution of suitability scores across ensembles.

b) Many (and possibly all) of the covariates in the models also apply to dengue – indeed two of them look at suitability for dengue transmission explicitly (due to lack of current understanding about physiological processes governing Zika transmission). Hence it remains unclear to the reviewers why the expected niche of Zika should be fundamentally different from that of dengue.

The reviewers did identify some differences between the Zika map and your corresponding map for dengue (Bhatt et al.), particularly in Africa. These are presumably driven by the dataset used to parameterise the model (based on Zika case reports in the literature). However, the very small (and potentially biased) body of literature on Zika virus prior to recent outbreaks may limit the accuracy of the estimates presented in the paper.

This needs to be carefully discussed: one solution would be to map the difference between your previous map for dengue (Bhatt et al.) and the map for Zika, with a discussion of the drivers for the difference and whether these seem to be justified (e.g. have at least some face validity).

In addition to the major comments, the reviewers also had some minor comments about presentational issues which we recommend you address, but which do not have to be explicitly responded to in your response letter:

a) On the whole the writing is excellent, but the Introduction is rather long. The focus should be sharpened to address the rationale for the present study. Much of the material presented is relevant, such as the connection to precedents of dengue and chikungunya, or the potential roles of various Aedes species in transmission, but it feels like a wandering review and the language should more directly point the reader to the relevance of each section.

b) The paragraph on cross-reactivity between dengue and Zika and newly emerging symptoms of Zika virus seems like a digression with little connection made to the broader purposes of the paper.

c) For vector competence, it would be useful to cite also very recent work emerging on Ae. aegypti and Ae. albopictus, and Cx. quinquefasciatus. Also to point out that all but the most recent vector competence work was done with ancestral African strains of Zika virus, not those circulating in the Americas.

d) Ae. hensilli should be mentioned as a critical vector on Yap.

e) When you state "wild isolations", it should be "isolations from wild mosquitoes" or similar.

f) The final statement in the Results about expected numbers of births needs additional context to understand how it was determined without the need to consult the detailed methods.

eLife. 2016 Apr 19;5:e15272. doi: 10.7554/eLife.15272.017

Author response


The reviewers all felt that this was a timely analysis that provides important first estimates that are needed in light of the ongoing spread of Zika virus into new areas via infected travelers. They also commended the manuscript for increasing the epidemiological relevance by translating global temperatures into model-based estimates of suitability for dengue transmission as a surrogate for Zika virus. These are imperfect and based on static, climatic summaries of conditions that vary seasonally, but are a reasonable choice given the gaps in our present knowledge base. The reviewers also appreciated the open sharing of the datasets used.

However, the reviewers had two major concerns about the manuscript:

a) The definition of limits where "either Ae. aegypti probability of occurrence is more than 0.8 or temperature is conducive to transmission via this mosquito for at least 355 days in an average year" seems quite subjective and presumably helped to keep the mapped values from straying too far from known distributional limits. The reviewers would like you to indicate the objective criteria used to define these limits. Since any criteria are unlikely to be completely indisputable, they suggest that you consider providing some measure of uncertainty for the predicted suitability scores e.g. perhaps the distribution of suitability scores across ensembles.

We thank the reviewers for this suggestion, and agree that our criteria should have been better explained. The final paragraph of the Results section now details the formulation of these criteria and refers to the appropriate sources. Additionally, we have added a map of uncertainty in the suitability scores to the supplementary information and made reference to it within this same paragraph.

b) Many (and possibly all) of the covariates in the models also apply to dengue – indeed two of them look at suitability for dengue transmission explicitly (due to lack of current understanding about physiological processes governing Zika transmission). Hence it remains unclear to the reviewers why the expected niche of Zika should be fundamentally different from that of dengue. The reviewers did identify some differences between the Zika map and your corresponding map for dengue (Bhatt et al.), particularly in Africa. These are presumably driven by the dataset used to parameterise the model (based on Zika case reports in the literature). However, the very small (and potentially biased) body of literature on Zika virus prior to recent outbreaks may limit the accuracy of the estimates presented in the paper. This needs to be carefully discussed: one solution would be to map the difference between your previous map for dengue (Bhatt et al.) and the map for Zika, with a discussion of the drivers for the difference and whether these seem to be justified (e.g. have at least some face validity).

We understand the reviewers’ concerns about the overlap in covariates used for modelling dengue and Zika. We have now addressed this in greater detail in the paragraph in the Discussion. While the covariates are indeed similar and would be well suited for a new iteration of the dengue map, we do not feel it is appropriate to make direct comparisons with the covariates from Bhatt et al. (2013) due to the fact that many refinements have been made to the covariates since that study. Regarding a difference map, while he Bhatt et al. (2013) dengue map is fundamentally different in that it used an evidence consensus layer to constrain predictions to areas with greater certainty of dengue transmission, the current ZIKV map shows environmental suitability for transmission, even in places where we know the disease has yet to be reported. Still, we felt it was useful to provide a difference map as the reviewers recommended, so in order to do so, we subtracted the values of the ZIKV map from those of the DENV map only in places where each had a non-zero prediction. This map is now found in the supporting information (Figure 2 —figure supplement 4), and reference is made to it and the patterns shown in the paper.


Articles from eLife are provided here courtesy of eLife Sciences Publications, Ltd

RESOURCES