Skip to main content

A statistical calibration tool for methods used to sample outdoor-biting mosquitoes

Abstract

Background

Improved methods for sampling outdoor-biting mosquitoes are urgently needed to improve surveillance of vector-borne diseases. Such tools could potentially replace the human landing catch (HLC), which, despite being the most direct option for measuring human exposures, raises significant ethical and logistical concerns. Several alternatives are under development, but detailed evaluation still requires common frameworks for calibration relative to HLC. The aim of this study was to develop and validate a statistical framework for predicting human-biting rates from different exposure-free alternatives.

Methods

We obtained mosquito abundance data (Anopheles arabiensis, Anopheles funestus and Culex spp.) from a year-long Tanzanian study comparing six outdoor traps [Suna Trap (SUN), BG Sentinel (BGS), M-Trap (MTR), M-Trap + CDC (MTRC), Ifakara Tent Trap-C (ITT-C) and Mosquito Magnet-X Trap (MMX)] and HLC. Generalised linear models were developed within a Bayesian framework to investigate associations between the traps and HLC, taking intra- and inter-specific density dependence into account. The best model was used to create a calibration tool for predicting HLC-equivalents.

Results

For An. arabiensis, SUN catches had the strongest correlation with HLC (R2 = 19.4), followed by BGS (R2 = 17.2) and MTRC (R2 = 13.1) catches. The least correlated catch was MMX (R2 = 2.5). For An. funestus, BGS had the strongest correlation with the HLC (R2 = 53.4), followed by MTRC (R2 = 37.4) and MTR (R2 = 37.4). For Culex mosquitoes, the traps most highly correlated with the HLC were MTR (R2 = 45.4) and MTRC (R2 = 44.2). Density dependence, both between and within species, influenced the performance of only BGS traps. An interactive Shiny App calibration tool was developed for this and similar applications.

Conclusion

We successfully developed a calibration tool to assess the performance of different traps for assessing outdoor-biting risk, and established a valuable framework for estimating human exposures based on the trap catches. The performance of candidate traps varied between mosquito taxa; thus, there was no single optimum. Although all the traps tested underestimated the HLC-derived exposures, it was possible to mathematically define their representativeness of the true biting risk, with or without density dependence. The results of this study emphasise the need to aim for a consistent and representative sampling approach, as opposed to simply seeking traps that catch the most mosquitoes.

Graphical Abstract

Introduction

Malaria control primarily relies on the use of insecticide-treated nets (ITNs) and indoor residual spraying (IRS) [1,2,3,4], with both of these control measures providing protection by targeting mosquitoes that are host seeking or resting indoors. Wide-scale use of these tools has yielded significant progress but challenges remain, such as insecticide resistance [5,6,7,8] and outdoor-biting [9,10,11]. Drivers of outdoor mosquito biting may include human behaviours [12,13,14], plasticity in mosquito behaviours (e.g. shifting from feeding indoors to feeding outdoors) [7, 15,16,17] and the effects of some indoor interventions [18, 19]. Sampling mosquito populations is a core component of malaria surveillance activities [20], and the aims of these activities include determining when and where people are most at risk. For the best results, this surveillance should consistently capture the key drivers of biting risk both indoors and outdoors. Unfortunately, representative sampling of mosquito vectors remains a challenge, particularly in outdoor settings.

The main entomological indicators assessed during vector surveillance include human-biting rate (HBR) [16, 21], sporozoite infection prevalence [22, 23], entomological inoculation rate (EIR) [21, 24], time of exposure and proportion of exposure prevented by ITNs [25,26,27,28]. The HBR is a fundamental variable for estimating the transmission of malaria and other mosquito-borne diseases [29]. As defined in the Ross MacDonald model, the HBR is required for estimation of the reproductive rate (R0) of malaria. Both the HBR and sporozoite prevalence are required for calculation of EIR [29], which is calculated as the number of infectious bites a person would be expected to receive in a given location over a given time period. The HBR and EIR are frequently used to estimate the impact of vector control interventions by highlighting how much they reduce exposure [4, 21, 30, 31].

The human landing catch (HLC) has long been the gold standard for direct measurement of human exposure and other key entomological variables. However, this method has several limitations and ethical concerns [32,33,34,35,36] due to its requirement that human volunteers expose parts of their body (usually lower legs) to mosquitoes [26, 37, 38], and this combination of ethical concerns and practical limitations has led to the wide recognition that alternative, exposure-free methods for measuring the HBR are needed [38,39,40,41,42]. Alternatives such as CDC light traps are already widely used for sampling host-seeking mosquitoes indoors [43], but these are unsuitable in outdoor settings. The urgency to identify suitable methods for measuring exposure outdoors is therefore greater [3, 42, 44, 45], especially due to the growing recognition of the importance of outdoor exposure to residual transmission [9, 12, 31].

To date, a number of alternative exposure-free methods have been independently developed and tested in different settings in Africa [3, 25, 38, 40, 42, 45,46,47,48]. Some methods provide a good representation of vector species composition and their biting activities, but underestimate density [3, 41, 45]. Others catch more mosquitoes than the HLC and thus overestimate typical human exposure [40, 49]. Finally, there are traps that are easy to implement, but which provide biased estimates of outdoor exposure by disproportionately sampling endophilic rather than exophilic species [50]. These strengths and weaknesses suggest that different traps are optimal for different surveillance applications. Unfortunately, there are no standardised calibration methods to allow estimation of HLC-equivalent exposure from the range of different outdoor sampling methods. Development of a standardised and validated calibration framework for such prediction would enable the results and methods from different studies to be compared. Such a calibration tool would need to reflect the potential non-linear relationship between trap counts and HLC values; this means that no single conversion ‘value’ between methods may apply across the full range of mosquito densities.

Several studies have indicated that trap performance relative to the HLC is density dependent [43, 51], although it should be noted that density dependence is often considered in terms of “intraspecific” density (e.g. the baseline density of the target vector species [42, 51]) but not the density of all mosquitoes, target vectors or not, that are attracted to the trap. However, the mechanisms that could give rise to intraspecific density dependence in trap performance could also generate dependence, with the overall densities of all mosquitoes attracted to the trap, including other species not of interest. While such interspecific dependence on the wider mosquito community is plausible, this has not been formally evaluated in trap evaluation studies.

The overall aim of this study was to provide an extensive comparison of six exposure-free traps for three vectors (Anopheles arabiensis, Anopheles funestus and Culex spp.). Specifically, we aimed to (i) assess the contribution of intra- and interspecific density dependence to trap performance, and (ii) develop an interactive calibration tool (in the form of a Shiny App) through which the number of a given species caught in an HLC can be predicted from catches made by alternative traps.

Methods

Study area and vector species

Mosquito trapping data were collected from six adjacent villages in the Ulanga and Kilombero districts of south-eastern Tanzania: Kivukoni (8.2135°S, 36.6879°E), Minepa (8.2710°S, 36.6771°E), Mavimba (8.3124°S, 36.6771°E), Milola (8.3306°S, 36.6727°E), Igumbiro (8.3511°S, 36.6725°E) and Lupiro (8.385°S, 36.670°E). Data were collected over 12 months between 2015 and 2016. The valley has relatively high mosquito abundance which peaks at the end of the rainy season. The common vectors of malaria transmission are An. arabiensis and An. funestus [16, 24, 52]. Mosquitoes in the Culex genera are also highly abundant, with some species being potential vectors for arboviruses found in the study area [53, 54].

Data collection

Mosquito sampling was carried out during both the wet and dry seasons, using six different traps for sampling outdoor-biting mosquitoes around human dwellings. The traps were: the Mosquito Magnet trap (MMX) [55], BG-Sentinel trap (BGS) [56], Suna trap (SUN) [3], Ifakara Tent Trap-C (ITT-C) [48], M-Trap (MTR) [57], M-Trap fitted with CDC Light trap (MTRC) (this study) and the HLC [3]. Most of these traps have been extensively described elsewhere, with the exception of the MTR fitted with a CDC light trap (MTRC), which was adapted from the original exposure-free M-Trap designed by Mwangungulu et al. [57]. Briefly, the HLC method involved male volunteers aged between 18 and 35 years who sat on a chair with their legs exposed and collected the mosquitoes that attempted to bite, using the mouth aspirator. Mosquitoes were sampled for 45 min each hour, allowing 15 min for rest. Each sampling village had its own set of volunteers.

In the present study, the original MTR was divided into two compartments made of UV-resistant shade netting: one in which a human volunteer sat to attract mosquitoes and the other section in which mosquito are entered [57]. A CDC light trap was suspended inside the other section of the trap to attract more mosquitos to the light source.

The traps were located at least 100 m apart. We assumed that the distance of 100 m offers sufficient independence between the traps as described by previous authors [58, 59]. Initial trap allocation was random, but their positions were switched over successive sampling nights in a Latin square design. In this way each trap was used in each position once over a 7-night cycle. After completion of each cycle, the study team moved to the next village so that one round of sampling in all six villages was completed over 42 trap-nights. Six rounds of data collection were completed, spanning the wettest and the driest periods of the year (252 trap-nights between April 2015 and April 2016). Mosquito sampling was done overnight from 6 pm to 6 am. The collected mosquitoes were morphologically sorted by taxa. A subsample of An. gambiae senso lato (s.l.) (n = 1405, 26% of total) was analysed by PCR [60] to assess sibling species composition within the complex.

Model fitting

The main goal of our analyses was to create a calibration tool to evaluate outdoor mosquito traps and to validate the tool by comparing the performance of candidate trapping methods relative to HLC (regarded in this study as the “gold standard”). In particular, we wanted to test the shape of the association between the numbers of mosquitoes collected by each trap type with those collected by the HLC. First, we pooled all the hourly collections into a single collection cup per trap per night. Then, for each of the focus mosquito groups (Culex genera, An. arabiensis and An. funestus s.l.), we modelled HLC catches as a function of the catching rate of each alternative trap.

Four general linear models were developed within a Bayesian model fitting framework to allow us to test for linear and non-linear associations through increasing the levels of complexity. The Bayesian approach allowed specific constraints to be placed on the parameters based on biological plausibility; this took the form of priors and uncertainty when converting the counts from alternative traps into HLC-equivalent values in the form of full posteriors.

For any given trap and mosquito group, we defined the response variable \(\left( {N_{i} } \right)\) as the number of female mosquitoes on every \(i{\text{th}}\) sampling night. Preliminary investigation of the data using Poisson likelihood showed over-dispersion for all three mosquito groups. Our final models did not account for other environmental covariates at specific trap locations (e.g. temperature, humidity). We accounted for the over-dispersion by using a negative binomial likelihood model formulated as a Gamma-Poisson mixture distribution [61]:

$$N_{i} \sim {\text{Poisson}}\left( {{\varvec{\theta}}_{{{\varvec{i}} }} {{\mathbf{\uplambda}}}_{i} } \right)$$

with

$${\varvec{\theta}}_{{\varvec{i}}} \sim {\text{Gamma}}\left( {\tau {{\mathbf{\uplambda}}}_{i} ,\tau {{\mathbf{\uplambda}}}_{i} } \right)$$

where the Poisson rate \({{\mathbf{\uplambda}}}_{{\varvec{i}}}\) is defined by the shape of the relationship between \(N_{i}\) and the number of mosquitoes collected with the alternative trap (\(n_{i} ,\) Table 1).

Table 1 Description of models used to investigate the relationships between female mosquito catches by human landing catch and the alternative traps

Since the algebraic form of this relationship is not known, we made three mutually inclusive assumptions with specified mathematical definitions, as follows: (i) that the relationship must start at the origin (i.e. when HLCs catch zero mosquitoes, the other traps will, on average, also collect zero mosquitoes); (ii) that the relationship is positive (i.e. no negative relationships between trap catches); and (iii) that any given trap could potentially suffer from a density effect (i.e. the slope of the relationship is not constant and it can change according to the baseline abundance of mosquitoes, either only of the same mosquito group or of all mosquitoes).

To define \({{\mathbf{\uplambda}}}_{{\varvec{i}}}\) we therefore formulated four possible scenarios to describe the relationship between HLCs and other trapping methods as summarised in Table 1 and Fig. 1. In Model 1, we considered a simple linear relationship between \(N_{i}\) and \(n_{i}\) (Table 1; Fig. 1a). In Model 2, we tested if the efficiency of the alternative trap was dependent on the density of the focal mosquito (e.g. “intra-specific” density dependence) by adding a quadratic term \(n_{i}^{2}\) (Table 1; Fig. 1b). In Model 3, we tested if the captures of a given group by a given trap were dependent on the abundance of the other taxonomic groups (e.g. “inter-specific” density dependence) by adding, an interaction term between \(n_{i}\) and the number of all the females from other mosquito groups collected with the same trap (\(m_{i}\)) (Table 1; Fig. 1c). Model 4 was similar to Model 3, but we considered all the other \(K_{i}\) taxonomic groups separately. Therefore, Model 4 included all the pairwise interaction terms between \(n_{i}\) and the number of females of each \(k{\text{th}}\) mosquito group \((s_{{k_{i} }} )\) (Table 1; Fig. 1d). Our analysis mainly focussed on three mosquito groups, but we collected a higher number of species hence \(K > 3\) (Additional file 1: Table S1).

Fig. 1
figure 1

Illustration of models used to investigate the relationship between number of female mosquitoes collected with human landing catch and six alternative traps. N = Number of female mosquitoes collected with human landing catch; n = number of female mosquitoes collected with a given alternative trap; m = pooled female mosquitoes of all the other species collected with the same alternative trap of n; s1, s2, s3 = number of female mosquitoes of each of the other K species, where K refers to all the species collected we developed the focus 1 (here K = 3). A Model 1 considers a simple linear relationship with n. B Model 2 considers a quadratic term n2. C Model 3 includes an interaction term between n and the number of all females of the other species collected with the same trap (m). D Model 4 considers all the pairwise interaction between n and s1, s2, s3

The analysis was performed in the statistical environment R [62], with Bayesian model fitting to the data done using the program JAGS [63] interfaced within R via the package rjags [64]. For parameters \(\beta_{1}\), \(\beta_{2}\) and \(\beta_{k}\) we used a gamma prior (shape = 0.1, rate = 0.1). The prior for \(\beta_{1}\) was chosen to ensure a positive relationship between \(n_{i}\) and \(N_{i}\) and a positive effect of the quadratic and the interaction terms for \(\beta_{2}\) and \(\beta_{k}\). To achieve convergence, the models were run for up to \(3 x 10^{4}\) iterations. Means of posterior distributions with corresponding credible intervals were obtained for each model coefficient \(\beta\). We compared different models by their deviance information criteria (DIC) and the goodness of fit of each model using pseudo \(R^{2}\) values. Models with the lowest DIC were selected as best. As a further cross-validation, we randomly split the data into a training (75%) and a test (25%) data set, and we calculated the root-mean-square error (RMSE), as the average prediction error by each model.

Interactive calibration tool

We designed a lookup table (Table 3) containing the means of posterior predictions for different combinations of mosquito taxa, trap types and models. This allowed us to predict the expected number of a given mosquito taxa from an HLC (with credible intervals) based on the number caught in the alternative traps. We also developed an interactive online tool, in the form of an R Shiny App [65] to facilitate these evaluations. This tool provides users with an interactive graphical user interface (GUI) to select the number of captured mosquitoes for a group of interest by trap type, and to explore the predicted number of mosquitoes caught in an HLC by method.

Results

The statistical correlations between HLCs and other trapping methods for each of the three mosquito groups are summarised in Table 2. The fit of models varied between trap types and mosquito group, with correlations with the HLC (R2 values) ranging from 0.8 to 53.4% (Table 2). The strength and nature of associations (Models 1–4) varied considerably between mosquito groups and traps; thus, no one single model was best in all cases. We provide an example of a prediction table (Table 3) which describes how mosquito abundance in a HLC can be estimated from catches made by the alternative traps (using Model 1, with intervals grouped by 10). Other model (Models 2–4) outputs/predictions can be easily retrieved from the Shiny App tool. Environmental covariates (temperature and humidity) were dropped during the initial model fitting process as they were not improving the goodness of fit of the model (Models 1–4).

Table 2 Summary (R2, DIC and RMSE values) of models used to investigate the relationship between the numbers of female mosquitoes collected with human landing catch and the six alternative outdoor traps
Table 3 Predicted values for estimating the expected mosquito catches by human landing catch and alternative traps, according to the linear model (Model 1)

Anopheles arabiensis

In most of the models, trap catches of An. arabiensis were only weakly correlated with HLC counts [Table 2 (b)]. SUN was the only alternative trap with a consistent correlation with the HLC of ≥ 17% (R2 > 17). For this trap, the relationship with HLC catches was best described by Model 4 [R2 = 19.4; Table 2 (b)], which incorporates both intra- and interspecific density dependence. However, the DIC values however did not vary much between models of differing complexity [ΔDIC = 1.32; Table 2 (b)]. Overall, SUN consistently underestimated HLC catches (for example 100 mosquitoes collected with SUN corresponded to 194 HLCs [95% credible intervals (CIs): 142–257; Table 3 (b); Fig. 2b).

Fig. 2
figure 2figure 2figure 2

Expected number of female Culex spp. (A) Anopheles arabiensis (B) and Anopheles funestus (C) mosquitoes collected with HLC (y-axis), given the number of females collected with alternative traps (x-axis). Continuous line is the prediction of a Gamma-Poisson model assuming a linear relationship; dashed lines are 95% credible intervals. Abbreviations: HLC, Human landing catch; SUN, Suna trap; BGS, BG-Sentinel trap; ITT-C, Ifakara Tent Trap version C; MMX, Mosquito Magnet trap; MTRC, M-Trap-Trap combined with CDC light source; MTR, M-trapTrap

BGS was the only trap where R2 values substantially increased with model complexity. Here, the most complex model (Model 4), which incorporated intra- and interspecific density dependence, had an R2 value of 17.2% [Table 2 (b)]. For all other trap types, correlations with the HLC were best explained by the simplest linear relationship (Model 1). Collections from BGS also underestimated the number of mosquitoes caught by HLC, and to a larger extent than the SUN [e.g. 100 mosquitoes caught by BGS is equivalent to 423 (95% CIs: 268–629) by HLC; Table 3 (b); Fig. 2b].

The MMX trap had the poorest correlation and was least representative of the HLC (R2 range: 0.8–2.5%), particularly at low densities where it often failed to capture any individuals. This trap therefore also significantly underestimated the catches relative to HLC [for example 100 catches of MMX is equivalent to 325 (95% CI: 187–504); Table 3 (b); Fig. 2b].

Anopheles funestus

There were no major differences between the alternative models when describing associations between HLC and the other traps for collecting An. funestus. Thus, on the basis of parsimony, we concluded that the simple linear model (Model 1) was sufficient to describe these relationships. BGS was the most highly correlated with the HLC (R2 range: 46.6–53.4%). The highest R2 value was from the most complex model (Model 4). However, similar to An. arabiensis, the BGS underestimated the number of An. funestus caught by HLC [Table 3 (c)] while, in contrast, the MTR, MTRC, ITT-C and MMX traps were only moderately correlated with HLC (R2 range: 30.0–37.4%); the SUN trap was the worst performing trap for this species [Table 2 (c)].

In general, predictions obtained with all An. funestus trap models (Models 1–4, for all trap types) were characterised by very large credible intervals (Fig. 2c), meaning that there was insufficient precision to define a useful calibration factor. This large uncertainty amount of HLC-equivalents of trap catches was particularly pronounced at higher An. funestus densities. In that sense, the trap that resulted in a (relatively) narrower prediction was MTR, where for 100 mosquitoes collected, the model would estimate 49 HLC-equivalents, with 95% CIs ranging from 10 to 126 [Table 3 (c)].

Culex species

Overall, there were moderate correlations between the alternative traps and HLC for Culex catches compared to those for the Anopheles groups (Table 2). However, there were no major differences between the tested models (based on ΔDIC estimates); thus, the simplest linear model was adopted based simply on being the thriftiest. Full details of all models are presented in Table 2 (a). Catches from MTRC and MTR traps had the highest correlations with HLCs (R2 range: 44.2–45.4%). On the other hand, MMX and ITT-C were the worst performing traps and significantly underestimated the HLC catches [Table 2 (a)].

Interactive calibration tool

To support detailed assessment and comparison of these and any future trap types for outdoor-sampling, we developed an interactive calibration tool incorporating the key parameters as identified in the analysis above. This tool is designed with simple user interfaces to simplify model inputs and outputs. For example, reporting full conversion tables for Models 3 and 4, which include density dependence, would be challenging since the associated interaction terms would require every possible combination of mosquito group, trap type and catch range. To obtain estimates according to these models, readers can use of our interactive online tool, which is available as an R Shiny App. The coefficients of these models will be updated regularly as additional data are gathered. This tool may be expanded to cover additional geographic regions and mosquito species not currently captured. The tool is hosted by an online server of the “Boyd Orr Centre for Population and Ecosystem Health” (University of Glasgow), and it is freely available at https://boydorr.gla.ac.uk/lucanelli/trapcalibration/.

Discussion

Despite the growing importance of outdoor-biting mosquitoes and their role in malaria transmission in different settings, there are limited methods for sampling outdoors. HLCs remain common and are sometimes considered to be the gold standard, but there are multiple ethical, cost and logistical concerns limiting its application [66, 67]. Multiple alternative tools have therefore been tested as potential HLC replacements in different settings [25, 39, 41, 43, 48, 50, 68]. While most efforts have focused on finding an alternative that catches as many mosquitoes as the HLC, it is now recognised that what matters more is how representative the catches from any specific trap are relative to HLCs. This means that efforts to improve surveillance methods should include not just new traps, but also a statistical tool for assessing their representativeness.

In this study, we therefore developed and validated a statistical framework for predicting credible intervals of HLC-derived exposure rates based on catches from multiple exposure-free alternatives. We have provided extensive comparison and correction factors for the different trapping methods, as well as evidence for the most representative alternative to the HLC. Furthermore, we have translated the results of our modelling approach into an easy-to-use interactive calibration tool that generates the expected means and credible intervals of nightly HBRs (using HLC as a proxy) based on inputs of other trap catches.

Among the several trapping methods that have been proposed for outdoor mosquito sampling of malaria vectors, only a few have been calibrated relative to the HLC [43], and even fewer have been calibrated in the outdoor setting [42, 45]. These traps provide disparate levels of efficacy relative to the HLC, and they rely primarily on two mutually inclusive principles: (i) the substitution of human subjects with human odours and a carbon dioxide source [4, 46]; or (ii) a trap design that protects human volunteers from bites with physical barriers [25, 41, 45, 57]. Many studies have assessed the correlations between mosquito abundance as estimated from the HLC and an alternative trap [3, 38, 40, 41, 45, 48, 51, 69, 70], but only a few provide the relevant quantitative estimates of “accuracy” (i.e. how close the estimates are to the HLC) and precision (i.e. how variable the estimates are) [38, 40, 41, 45, 48]. Furthermore, to our knowledge, none have provided an explicit calibration tool to facilitate rapid predictions of mosquito counts from an alternative trap into an HLC-equivalent. Such a calibration tool would need to reflect the potential non-linear relationship between trap counts and HLC values, which means that no single conversion “value” between methods may apply across the full range of mosquito densities. This hypothesis is backed up by a multi-country study which evaluated the limitation of CDC light traps on African malaria vectors after observing the non-linearity [43].

In general, the overall measure for goodness of fit (R2) for models predicting HLC counts was highest in An. funestus, followed by Culex spp. and An. arabiensis. Despite the higher value of R2 in An. funestus, the wider credible intervals were probably due to the much small sample size of this species (total mosquito caught with HCL: An. funestus = 226, An. arabiensis = 5282, Culex = 7191), although it could also have been affected by other ecological features that were not directly captured with this study (e.g. other environmental conditions apart from humidity and temperature). During the model fitting exercise, temperature and humidity were excluded via the model selection process. The proportion of An. funestus in the study area compared to other species such as An. arabiensis and Culex has been historically low [10, 16, 24] although the former species carries a significant amount of infection compared to other commonly known malaria vectors [24].

The performance of some alternative traps in comparison to the HLC has been shown to be density dependent in several investigations [43, 51] although such density-dependent impacts are usually only considered in terms of “intraspecific” dependencies, such as the baseline density of the target vector species [42, 51], overlooking the larger mosquito community. However, the same mechanisms that cause intraspecific density dependence in trap performance may also cause dependence on the overall densities of all mosquito species lured to the trap, including species that are not of public health importance. While such reliance on the wider mosquito community is plausible, it has yet to be tested in trap evaluation studies. Therefore, the present study and the calibration tool that we developed also included a robust assessment of how density dependence may play a role. Models 3 and 4 included these variables and will allow users to incorporate these as covariates when predicting outdoor-biting rates in their settings of interest.

Overall, this study found little evidence that the relative performance of the trapping methods investigated here is modified by the density of the target mosquito taxa or other members of the mosquito community. Models that incorporate intra- or interspecific density dependence in trap performance did not yield any substantial improvements over those assuming simple linear relationship between mosquito counts in the HLC and the alternative method. This indicates that neither intra- nor interspecific density dependence has a large impact on the relative efficiency of the alternative traps tested here. Given the wide range of trap catches, the calibration tool we developed here allows users to incorporate such density-dependence effects (both within and between species) and to examine if these are applicable in their settings. Previous studies detected (intraspecific) density dependence in the performance of some trapping methods [45, 48, 49], but evidence of density dependence in trap performance can be variable even for the same trapping method. For example, studies investigating the performance of the Mosquito Electrocuting Trap relative to the HLC have detected density dependence in some cases [25, 43], but not others [45].

One limitation of this study is that while the HLC is broadly considered to be the gold standard for collecting host-seeking mosquitoes both indoors and outdoors, we only focussed on traps for outdoor sampling. Although we compared a large number of trap types commonly used in Africa settings, other traps may perform differently and potentially better than some of the candidate traps investigated here [25, 41, 45]. Additional studies including additional alternative traps for indoor and outdoor use would be of further value—with the calibration tool developed here providing a useful framework for their evaluation and comparison. Also based on the results presented here, we recommend that for whatever trap used, the users should generate credible estimates of what the HBRs (as estimated from HLC) could be. Due to the potential variation in trap performance between different ecological settings and mosquito species, we do not yet recommend any one specific trap as the best replacement for the HLC. Instead, we recommend that users consider and define the statistical relationships between a prospective trap and the HLC when planning surveillance and interpreting results. The interactive conversion tool we have developed here can be used for that purpose and is now available online as a Shiny App interface.

Conclusion

Methods for sampling outdoor-biting mosquitoes are urgently needed to improve surveillance of vector-borne diseases. Even if an alternative traps do not catch as many mosquitoes as HLC, it is desirable to define the statistical relationship between them so that credible ranges of actual biting risk can be predicted in units of HLC equivalents. In this study, we successfully evaluated six different outdoor traps and developed a calibration tool to assess their performance relative to the HLC. This tool was validated using data from year-round field collections and enabled a framework for predicting HLC-derived exposure rates representative of individual risk to mosquito biting. The tool incorporates multiple models, including two that allow assessment of effects of both inter- and intra-specific density dependence of the performance of candidate traps. In the specific field trials from which data were obtained here, density dependence between and within mosquito species influenced the performance of only one trap, the BGS, but not any others. An interactive Shiny App calibration tool was developed for this and similar applications. We conclude that this calibration approach provides a valuable framework for assessing human exposure from different outdoor trapping methods. As the performance of candidate traps relative to the HLC varied between mosquito taxa, there was no single optimum. While all the candidate traps underestimated HLC catches, and thus HBRs, the calibration tool created here enables a mathematical definition of the traps relationship as well as model-fitting limits. Further studies of trapping methods and associated evaluation criteria should focus on consistency and representativeness as opposed to simply finding traps that catch as many mosquitoes as HLC.

Availability of data and materials

The R-codes used for analysis and the trap data are available upon reasonable request to the author. The Shiny App is accessible by the link provided in the manuscript.

Abbreviations

DIC:

Deviance information criteria

BGS:

BG-Sentinel

CDC Light Trap:

Centre for Diseases and Control light trap

CIs:

Confidence Intervals

DIC:

Deviance Information Criteria

EIR:

Entomological inoculation rate

GUI:

Graphical User Interface

HBR:

Human Biting Rate

HLC:

Human Landing catches

IRS:

Indoor Residual Spray

ITNs:

Insecticide Treated Nets

ITT-C:

Ifakara Tent Trap type C

JAGS:

Just Another Gibbs Sampler

MMX:

Mosquito Magnet-X trap

MTR:

M-Trap

MTRC:

M-Trap + CDC Light Trap

PCR:

Polymerase Chain Reaction

RMSE:

Root Mean Square Error

SUN:

Suna Trap

References

  1. Pluess B, Tanser FC, Lengeler C, Sharp BL. Indoor residual spraying for preventing malaria (Review). Cochrane Database Syst Rev. 2010;2010(4):CD006657. https://doi.org/10.1002/14651858.CD006657.pub2.

  2. World Health Organization. Achieving and maintaining universal coverage with long-lasting insecticidal nets for malaria control. 2017. https://apps.who.int/iris/handle/10665/259478.

    Google Scholar 

  3. Hiscox A, Otieno B, Kibet A, Mweresa CK, Omusula P, Geier M, et al. Development and optimization of the Suna trap as a tool for mosquito monitoring and control. Malar J. 2014;13:257.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Russell TL, Beebe NW, Cooper RD, Lobo NF, Burkot TR. Successful malaria elimination strategies require interventions that target changing vector behaviours. Malar J. 2013;12:56.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Viana M, Hughes A, Matthiopoulos J, Ranson H, Ferguson HM. Delayed mortality effects cut the malaria transmission potential of insecticide-resistant mosquitoes. Proc Natl Acad Sci USA. 2016;113:8975–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Wondji CS, Coleman M, Kleinschmidt I, Mzilahowa T, Irving H, Ndula M, et al. Impact of pyrethroid resistance on operational malaria control in Malawi. Proc Natl Acad Sci USA. 2012;109:19063–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Sanou A, Nelli L, Guelbéogo WM, Cissé F, Tapsoba M, Ouédraogo P, et al. Insecticide resistance and behavioural adaptation as a response to long-lasting insecticidal net deployment in malaria vectors in the Cascades region of Burkina Faso. Sci Rep. 2021;11:1–14.

    Article  CAS  Google Scholar 

  8. Hemingway J, Hawkes NJ, McCarroll L, Ranson H. The molecular basis of insecticide resistance in mosquitoes. Insect Biochem Mol Biol. 2004;34:653–65.

    Article  CAS  PubMed  Google Scholar 

  9. Russell TL, Govella NJ, Azizi S, Drakeley CJ, Kachur SP, Killeen GF. Increased proportions of outdoor feeding among residual malaria vector populations following increased use of insecticide-treated nets in rural Tanzania. Malar J. 2011;10:80. https://doi.org/10.1186/1475-2875-10-80.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Lwetoijera D, Harris C, Kiware SS, Dongus S, Devine GJ, McCall PJ, et al. Increasing role of Anopheles funestus and Anopheles arabiensis in malaria transmission in the Kilombero Valley, Tanzania. Malar J. 2014;13:331.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Mendis C, Jacobsen JL, Gamage-Mendis A, Bule E, Dgedge M, Thompson R, et al. Anopheles arabiensis and An. funestus are equally important vectors of malaria in Matola coastal suburb of Maputo, southern Mozambique. Med Vet Entomol. 2000;14:171–80.

    Article  CAS  PubMed  Google Scholar 

  12. Monroe A, Mihayo K, Okumu F, Finda M, Moore S, Koenker H, et al. Human behaviour and residual malaria transmission in Zanzibar: findings from in-depth interviews and direct observation of community events. Malar J. 2019;18:220.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Finda MF, Moshi IR, Monroe A, Limwagu AJ, Nyoni AP, Swai JK, et al. Linking human behaviours and malaria vector biting risk in south-eastern Tanzania. PLoS ONE. 2019;14:e0217414.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Monroe A, Asamoah O, Lam Y, Koenker H, Psychas P, Lynch M, et al. Outdoor-sleeping and other night-time activities in northern Ghana: implications for residual transmission and malaria prevention. Malar J. 2015;14:35.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Takken W, Verhulst NO. Host preferences of blood-feeding mosquitoes. Annu Rev Entomol. 2011;58:120928130709004.

    Google Scholar 

  16. Ngowo HS, Kaindoa EW, Matthiopoulos J, Ferguson HM, Okumu FO. Variations in household microclimate affect outdoor-biting behaviour of malaria vectors. Wellcome Open Res. 2017;2:102.

    Article  PubMed  PubMed Central  Google Scholar 

  17. Kreppel KS, Viana M, Main BJ, Johnson PCD, Govella NJ, Lee Y, et al. Emergence of behavioural avoidance strategies of malaria vectors in areas of high LLIN coverage in Tanzania. Sci Rep. 2020;10:1–11.

    Article  CAS  Google Scholar 

  18. Ogoma SB, Ngonyani H, Simfukwe ET, Mseka A, Moore J, Killeen GF. Spatial repellency of transfluthrin-treated hessian strips against laboratory-reared Anopheles arabiensis mosquitoes in a semi-field tunnel cage. Parasit Vectors. 2012;5:1–5.

    Article  Google Scholar 

  19. Ogoma SB, Ngonyani H, Simfukwe ET, Mseka A, Moore J, Maia MF, et al. The mode of action of spatial repellents and their impact on vectorial capacity of Anopheles gambiae sensu stricto. PLoS ONE. 2014;9:e110433.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. World Health Organization. Malaria surveillance, monitoring & evaluation: a reference manual. 2018. https://www.who.int/publications/i/item/9789241565578.

    Google Scholar 

  21. Killeen GF, McKenzie FE, Foy BD, Schieffelin C, Billingsley PF, Beier JC. A simplified model for predicting malaria entomologic inoculation rates based on entomologic and parasitologic parameters relevant to control. Am J Trop Med Hyg. 2000;62:535–44.

    Article  CAS  PubMed  Google Scholar 

  22. Tusting LS, Bousema T, Smith DL, Drakeley C. Measuring changes in Plasmodium falciparum transmission: precision, accuracy and costs of metrics. Adv Parasitol. 2014;84:151-208. https://doi.org/10.1016/B978-0-12-800099-1.00003-X.

  23. Charlwood JD, Smith T, Billingsley PF, Takken W, Lyimo EOK, Meuwissen JHET. Survival and infection probabilities of anthropophagic anophelines from an area of high prevalence of Plasmodium falciparum in humans. Bull Entomol Res. 1997;87:445.

    Article  Google Scholar 

  24. Kaindoa EW, Matowo NS, Ngowo HS, Mkandawile G, Mmbando A, Finda M, et al. Interventions that effectively target Anopheles funestus mosquitoes could significantly improve control of persistent malaria transmission in south-eastern Tanzania. PLoS ONE. 2017;12:e0177807.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Maliti DV, Govella NJ, Killeen GF, Mirzai N, Johnson PCD, Kreppel K, et al. Development and evaluation of mosquito-electrocuting traps as alternatives to the human landing catch technique for sampling host-seeking malaria vectors. Malar J. 2015;14:502.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Govella NJ, Maliti DF, Mlwale AT, Masallu JP, Mirzai N, Johnson PCD, et al. An improved mosquito electrocuting trap that safely reproduces epidemiologically relevant metrics of mosquito human-feeding behaviours as determined by human landing catch. Malar J. 2016;15:465.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Huho B, Briët O, Seyoum A, Sikaala C, Bayoh N, Gimnig J, et al. Consistently high estimates for the proportion of human exposure to malaria vector populations occurring indoors in rural Africa. Int J Epidemiol. 2013;42:235–47.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Seyoum A, Sikaala CH, Chanda J, Chinula D, Ntamatungiro AJ, Hawela M, et al. Human exposure to anopheline mosquitoes occurs primarily indoors, even for users of insecticide-treated nets in Luangwa Valley, South-east Zambia. Parasit Vectors. 2012;5:101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Macdonald G. The epidemiology and control of malaria. Epidemiol Control Malar. 1957;7:577–8.

    Google Scholar 

  30. Bayoh MN, Mathias DK, Odiere MR, Mutuku FM, Kamau L, Gimnig JE, et al. Anopheles gambiae: historical population decline associated with regional distribution of insecticide-treated bed nets in western Nyanza Province, Kenya. Malar J. 2010;9:62.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Killeen GF. Characterizing, controlling and eliminating residual malaria transmission. Malar J. 2014;13:330.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Ndebele P, Musesengwa R. View point: ethical dilemmas in malaria vector research in Africa: making the difficult choice between mosquito, science and humans. Malawi Med J. 2012;24:65–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Service MW. A critical review of procedures for sampling populations of adult mosquitoes. Bull Entomol Res. 1977;67:343–82.

    Article  Google Scholar 

  34. Knols BGJ, de Jong R, Takken W. Differential attractiveness of isolated humans to mosquitoes in Tanzania. Trans R Soc Trop Med Hyg. 1995;89:604–6.

    Article  CAS  PubMed  Google Scholar 

  35. Lindsay SW, Adiamah JH, Miller JE, Pleass RJ, Armstrong JR. Variation in attractiveness of human subjects to malaria mosquitoes (Diptera: Culicidae) in The Gambia. J Med Entomol. 1993;30:368–73.

    Article  CAS  PubMed  Google Scholar 

  36. Achee NL, Youngblood L, Bangs MJ, Lavery JV, James S. Considerations for the use of human participants in vector biology research: a tool for investigators and regulators. Vector-Borne Zoonotic Dis. 2015;15:89–102.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Mboera L. Sampling techniques for adult Afrotropical malaria vectors and their reliability in the estimation of entomological inoculation rate. Tanzan J Health Res. 2006;7:117–24.

    Article  Google Scholar 

  38. Meza FC, Kreppel KS, Maliti DF, Mlwale AT, Mirzai N, Killeen GF, et al. Mosquito electrocuting traps for directly measuring biting rates and host-preferences of Anopheles arabiensis and Anopheles funestus outdoors. Malar J. 2019;18:1–11.

    Article  Google Scholar 

  39. Okumu F, Biswaro L, Mbeleyela E, Killeen GF, Mukabana R, Moore SJ. Using nylon strips to dispense mosquito attractants for sampling the malaria vector Anopheles gambiae s.s. J Med Entomol. 2010;47:274–82.

    Article  CAS  PubMed  Google Scholar 

  40. Abong’O B, Yu X, Donnelly MJ, Geier M, Gibson G, Gimnig J, et al. Host Decoy Trap (HDT) with cattle odour is highly effective for collection of exophagic malaria vectors. Parasit Vectors. 2018;11:533.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Limwagu AJ, Kaindoa EW, Ngowo HS, Hape E, Finda M, Mkandawile G, et al. Using a miniaturized double-net trap (DN-Mini) to assess relationships between indoor–outdoor biting preferences and physiological ages of two malaria vectors, Anopheles arabiensis and Anopheles funestus. Malar J. 2019;18:282.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Govella NJ, Maliti DF, Mlwale AT, Masallu JP, Mirzai N, Johnson PCD, et al. An improved mosquito electrocuting trap that safely reproduces epidemiologically relevant metrics of mosquito human—feeding behaviours as determined by human landing catch. Malar J. 2016;15:1–17.

    Article  Google Scholar 

  43. Briët OJT, Huho BJ, Gimnig JE, Bayoh N, Seyoum A, Sikaala CH, et al. Applications and limitations of centers for disease control and prevention miniature light traps for measuring biting densities of African malaria vector populations: a pooled-analysis of 13 comparisons with human landing catches. Malar J. 2015;14:247.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Kreppel KS, Johnson PCD, Govella NJ, Pombi M, Maliti D, Ferguson HM. Comparative evaluation of the Sticky-Resting-Box-Trap, the standardised resting-bucket-trap and indoor aspiration for sampling malaria vectors. Parasit Vectors. 2015;8:462.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Sanou A, Guelbéogo WM, Nelli L, Toé KH, Zongo S, Ouédraogo P, et al. Evaluation of mosquito electrocuting traps as a safe alternative to the human landing catch for measuring human exposure to malaria vectors in Burkina Faso. Malar J. 2019;18:1–17.

    Article  CAS  Google Scholar 

  46. Matowo NS, Moore J, Mapua S, Madumla EP, Moshi IR, Kaindoa EW, et al. Using a new odour-baited device to explore options for luring and killing outdoor-biting malaria vectors: a report on design and field evaluation of the Mosquito Landing Box. Parasit Vectors. 2013;6:137.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Mwanga EP, Ngowo HS, Mapua SA, Mmbando AS, Kaindoa EW, Kifungo K, et al. Evaluation of an ultraviolet LED trap for catching Anopheles and Culex mosquitoes in south-eastern Tanzania. Parasit Vectors. 2019;12:1–12.

    Article  CAS  Google Scholar 

  48. Govella NJ, Chaki PP, Geissbuhler Y, Kannady K, Okumu F, Charlwood JD, et al. A new tent trap for sampling exophagic and endophagic members of the Anopheles gambiae complex. Malar J. 2009;8:157.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Hawkes FM, Dabiré RK, Sawadogo SP, Torr SJ, Gibson G. Exploiting Anopheles responses to thermal, odour and visual stimuli to improve surveillance and control of malaria. Sci Rep. 2017;7:17283.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Govella NJ, Chaki PP, Mpangile JM, Killeen GF. Monitoring mosquitoes in urban Dar es Salaam: evaluation of resting boxes, window exit traps, CDC light traps, Ifakara tent traps and human landing catches. Parasit Vectors. 2011;4:40.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Mathenge EM, Omweri GO, Irungu LW, Ndegwa PN, Walczak E, Smith TA, et al. Comparative field evaluation of the Mbita trap, the centers for disease control light trap, and the human landing catch for sampling of malaria vectors in western Kenya. Am J Trop Med Hyg. 2004;70:33–7.

    Article  PubMed  Google Scholar 

  52. Kaindoa EW, Mkandawile G, Ligamba G, Kelly-Hope LA, Okumu FO. Correlations between household occupancy and malaria vector biting risk in rural Tanzanian villages: implications for high-resolution spatial targeting of control interventions. Malar J. 2016;15:199.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Sumaye RD, Abatih EN, Thiry E, Amuri M, Berkvens D, Geubbels E. Inter-epidemic acquisition of Rift Valley fever virus in humans in Tanzania. PLoS Negl Trop Dis. 2015;9:e0003536.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Sumaye RD, Geubbels E, Mbeyela E, Berkvens D. Inter-epidemic transmission of Rift Valley fever in livestock in the Kilombero River Valley, Tanzania: a cross-sectional survey. PLoS Negl Trop Dis. 2013;7:e2356.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Qiu YT, Smallegange RC, ter Braak CJF, Spitzen J, Van Loon JJA, Jawara M, et al. Attractiveness of MM-X traps baited with human or synthetic odor to mosquitoes (Diptera: Culicidae) in The Gambia. J Med Entomol. 2007;44:970–83.

    Article  CAS  PubMed  Google Scholar 

  56. Batista EPA, Ngowo HS, Opiyo M, Shubis GK, Meza FC, Okumu FO, et al. Semi-field assessment of the BG-malaria trap for monitoring the African malaria vector, Anopheles arabiensis. PLoS ONE. 2017;12:1–17.

    Article  Google Scholar 

  57. Mwangungulu SP, Sumaye RD, Limwagu AJ, Siria DJ, Kaindoa EW, Okumu FO. Crowdsourcing vector surveillance: using community knowledge and experiences to predict densities and distribution of outdoor-biting mosquitoes in rural Tanzania. PLoS ONE. 2016;11:e0156388.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Okumu FO, Killeen GF, Ogoma S, Biswaro L, Smallegange RC, Mbeyela E, et al. Development and field evaluation of a synthetic mosquito lure that is more attractive than humans. PLoS ONE. 2010;5:e8951.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  59. Okumu FO, Govella NJ, Moore SJ, Chitnis N, Killeen GF. Potential benefits, limitations and target product-profiles of odor-baited mosquito traps for malaria control in Africa. PLoS ONE. 2010;5:e11573. https://doi.org/10.1371/journal.pone.0011573.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Scott JA, Brogdon WG, Collins FH. Identification of single specimens of the Anopheles gambiae complex by the polymerase chain reaction. Am J Trop Med Hyg. 1993;49:520–9.

    Article  CAS  PubMed  Google Scholar 

  61. Greene W. Functional forms for the negative binomial model for count data. Econ Lett. 2008;99:585–90.

    Article  Google Scholar 

  62. R Development Core Team. R: a language and environment for statistical computing. Vienna: R Foundation for Statistical Computing; 2021.

  63. Plummer M. JAGS: a program for analysis of Bayesian graphical models using Gibbs sampling JAGS: just another Gibbs sampler. 2003.

  64. Denwood MJ. runjags: An R package providing interface utilities, model templates, parallel computing methods and additional distributions for MCMC models in JAGS. J Stat Softw. 2016.

  65. Chang W, Cheng J, Allaire JJ, et al. shiny: web application framework for R. R package version 1.3.2. 2019.

  66. Gimnig JE, Walker ED, Otieno P, Kosgei J, Olang G, Ombok M, et al. Incidence of malaria among mosquito collectors conducting human landing catches in western Kenya. Am J Trop Med Hyg. 2013;88:301–8.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Wotodjo AN, Trape JF, Richard V, Doucouré S, Diagne N, Tall A, et al. No difference in the incidence of malaria in human-landing mosquito catch collectors and non-collectors in a Senegalese village with endemic malaria. PLoS ONE. 2015;10:e0126187.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. Tangena J-AA, Thammavong P, Hiscox A, Lindsay SW, Brey PT. The human-baited double net trap: an alternative to human landing catches for collecting outdoor biting mosquitoes in Lao PDR. PLoS ONE. 2015;10:e0138735.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  69. Odetoyinbo JA. Preliminary investigation on the use of a light-trap for sampling malaria vectors in the Gambia. Bull World Health Organ. 1969;40:547–60.

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Davis JR, Hall T, Chee E, Majala A, Minjas J, Shiff CJ. Comparison of sampling anopheline mosquitoes by light-trap and human-bait collections indoors at Bagamoyo, Tanzania. Med Vet Entomol. 1995;9:249–55.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We would like to thank the village leaders and communities of Ulanga districts for allowing us to work in their areas at the same time look after our traps. We acknowledge the support from field volunteers and technician who were involved in the data collection inside the traps which require human presence.

Funding

This work was supported by the Wellcome Trust [102350] grant awarded to FOO. HSN and FOO were supported by the Wellcome Trust [102350] and Howard Hughes Medical Institute (HHMI)-Gates Foundation (Grants: OPP1099295) grants.

Author information

Authors and Affiliations

Authors

Contributions

FOO designed the study and acquire the funds. HSN and AJL were involved in the study design, data collection and field supervision. HSN, LN and JM developed the modelling framework, analysed the data, and were involved in the interpretation of the results. HSN drafted the manuscript. FOO, HMF, JM and LN reviewed the manuscript and provide supervision. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Halfan S. Ngowo.

Ethics declarations

Ethics approval and consent to participate

Ethical approval was obtained from the Ifakara Health Institute’s Institutional review board (IHI/IRB/No: 06-2016), and the Medical Research Coordination Committee of the National Institute for Medical Research in Tanzania (MRCC) (NIMR/HQ/R.8a/Vol.IX/2218). Written informed consent was obtained prior to the start of the experiment for every compound owner where the traps were located. Consent was also obtained from volunteers involved in the collections of mosquitoes using HLC and inside the M-trap fitted with CDC (MTRC). Those involved in HLC were provided with malaria prophylaxis as pre-scribed by licensed medical personnel. Volunteers were regularly checked for malaria parasite. From the beginning of the study to its completion, no volunteers were found positive with Plasmodium.

Consent for publication

Permission to publish was obtained from the Medical Research Coordinating Committee (MRCC) at the National Institute for Medical Research (NIMR) (Ref: No: NIMR/HQ/P.12 VOL XXXIV/8).

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Table S1.

Summary of all other mosquitoes collected for each trap type.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ngowo, H.S., Limwagu, A.J., Ferguson, H.M. et al. A statistical calibration tool for methods used to sample outdoor-biting mosquitoes. Parasites Vectors 15, 293 (2022). https://doi.org/10.1186/s13071-022-05403-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13071-022-05403-7

Keywords