Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Sep 4.
Published in final edited form as: Nature. 2010 Feb 3;464(7285):66–71. doi: 10.1038/nature08834

Odourant reception in the malaria mosquito Anopheles gambiae

Allison Carey 1, Guirong Wang 2, Chih-Ying Su 1, Laurence J Zwiebel 2, John R Carlson 1
PMCID: PMC2833235  NIHMSID: NIHMS171208  PMID: 20130575

Summary

The mosquito Anopheles gambiae is the major vector of malaria in sub-Saharan Africa. It locates its human hosts primarily through olfaction, but little is known about the molecular basis of this process. Here we functionally characterize the Anopheles gambiae Odourant Receptor (AgOr) repertoire. We identify receptors that respond strongly to components of human odour and that may act in the process of human recognition. Some of these receptors are narrowly tuned, and some salient odourants elicit strong responses from only one or a few receptors, suggesting a central role for specific transmission channels in human host-seeking behavior. This analysis of the Anopheles gambiae receptors permits a comparison with the corresponding Drosophila melanogaster odourant receptor repertoire. We find that odourants are differentially encoded by the two species in ways consistent with their ecological needs. Our analysis of the Anopheles gambiae repertoire identifies receptors that may be useful targets for controlling the transmission of malaria.


Mosquitoes transmit many diseases, including malaria, which afflicts hundreds of millions of people each year1. The malaria burden is heaviest in sub-Saharan Africa, where the An. gambiae mosquito is the major vector. An. gambiae relies heavily on olfactory cues to identify its human hosts24, but the molecular basis of host-seeking behavior is unknown.

Insects detect odours via olfactory receptor neurons (ORNs). The odourant specificities of many ORNs are conferred by the expression of individual odourant receptor genes5. A family of 79 AgOr (Anopheles gambiae Odourant receptor) genes has been identified bioinformatically in An. gambiae6, 7. Two of these receptors have been characterized functionally8 using an in vivo heterologous expression system, the “empty neuron” system9, which has also been used to decode the D. melanogaster odourant receptor repertoire1012. These results invited a systematic, functional characterization of the AgOr repertoire and a comparison between the receptor repertoires of these two species, which exhibit different olfactory-driven behaviors. D. melanogaster consumes fruit and is considered a generalist. An. gambiae has evolved an anthropophilic host-seeking olfactory response that allows it to find human bloodmeals4. Little is known about how the odourant receptor repertoires of these species have adapted to meet their distinct ecological requirements.

Functional expression of the AgOr repertoire

To investigate the molecular basis of odour reception in An. gambiae, we amplified the coding regions of 72 AgOr genes from olfactory organ cDNA of adult mosquitoes. We then expressed each AgOr in the “empty neuron,” a mutant ORN in D. melanogaster that lacks its endogenous odourant receptor9. Fifty of the 72 cloned AgOr receptors were functional in the empty neuron, conferring a regular, characteristic, spontaneous firing rate and exhibiting excitatory and/or inhibitory responses to odourant stimuli (Figure 1a). This success rate (69%) is comparable to that for D. melanogaster antennal Or genes (77%) expressed in the empty neuron11.

Figure 1. Functional characterization of the AgOrs.

Figure 1

(a) Extracellular recordings from the empty neuron expressing an AgOr. Top, excitatory response of AgOr2 to 2-methylphenol; middle, excitatory response of AgOr21 to 6-methyl-5-hepten-2-one; bottom, inhibitory response of AgOr1 to 6-methyl-5-hepten-2-one. Action potentials from both neurons housed in the sensillum can be observed and distinguished by amplitude; the larger-amplitude action potential, from the ‘A’ neuron, expresses the AgOr, while the smaller-amplitude action potential, from the ‘B’ neuron, expresses its endogenous D. melanogaster odourant receptor9. (b) Left, odourant response profile of AgOr8 expressed in the empty neuron. Right, odourant response profile of the An. gambiae neuron that houses AgOr8 (adapted from Lu, et al., 2007). All odourants were tested at a dilution of 10−2. The spontaneous firing rate and the responses to the diluent are subtracted from odourant responses in each panel. Of the 28 odourants that inhibited the spontaneous firing rate by 50% or more in the empty neuron47, 21 gave mean responses that were negative in the endogenous An. gambiae neuron. Error bars represent SEM. (c) Heat map of responses of the 50 functional AgOrs to 110 odourants. Response intensity is colour-coded according to the continuous colour scale on the right and represents the mean activity measured over a 0.5 second odourant stimulation period. Receptors, odourants, and numerical values are indicated in Supplementary Table 2. n = 5–6; for odourants that elicit responses ≥100 spikes/second, n = 6. Spontaneous activity and responses to diluent have been subtracted from response values. All odourants were tested at a 10−2 dilution. Odourants containing both a phenol ring and an ester moiety are classified as aromatics; terpenes containing an ester moiety are classified as terpenes.

The empty neuron system previously was demonstrated to be a high-fidelity expression platform for the D. melanogaster Or genes11, referred to here as DmOrs. Since An. gambiae and D. melanogaster are separated by 250 million years of evolution13, we wanted to determine whether the empty neuron is also a faithful expression system for AgOr genes. One of the few AgOrs that has been unequivocally mapped to a specific ORN in the mosquito is AgOr8, which in its endogenous neuron responded to seven- and eight-carbon-chain compounds among a panel of tested odourants14. We expressed AgOr8 in the empty neuron and found that its response profile closely resembled that of the endogenous neuron (Figure 1b). We also generated dose-response curves for two ligands of AgOr8, 1-octen-3-ol and 1-hepten-3-ol (Supplementary Figure 1), and found that the differential sensitivity to these ligands observed in the endogenous neuron14 was maintained in the empty neuron. These results validate the empty neuron as a faithful heterologous expression system for AgOrs.

The 50 functional AgOrs were tested against a chemically diverse panel of 110 odourants, including components of human emanations and oviposition site volatiles (Supplementary Table 1). Fifty-three of the 110 odourants were previously tested against the D. melanogaster antennal receptor repertoire in the empty neuron system12, permitting functional comparisons between the odourant receptor repertoires of the fruit fly and the mosquito.

We tested each of the functional AgOrs against the 110-odourant panel, generating a data set of 5500 odourant-receptor combinations, with each combination tested n≥5 times (Figure 1c and Supplementary Table 2a–d). We found that individual receptors respond to subsets of odourants, and individual odourants activate subsets of receptors, consistent with a combinatorial model of odour coding12, 15, 16. Some receptors gave strong responses (defined as ≥100 spikes/second) to many odourants, while other receptors are more selective (Supplementary Table 2a–d). These differences were visualized by generating a tuning curve for each receptor (Figure 2). The breadths of the tuning curves were quantified according to their kurtosis value, a measure of the “peakedness” of the distribution (Supplementary Tables 3a, b). We found a continuum of AgOr tuning breadths ranging from broad to narrow, consistent with analysis of the D. melanogaster Or repertoire12. We then considered the receptors at each extreme for insight into the molecular basis of odour recognition.

Figure 2. Tuning breadths of receptors.

Figure 2

Tuning curves for the AgOrs that respond strongly (≥100 spikes/second) to at least one odourant on the panel. The 110 odourants are arranged along the X-axis according to the strength of the response they elicit from each receptor. The odourants that elicit the strongest responses are placed near the center of the distribution; those that elicit the weakest responses are placed near the edges. The order of odourants is therefore different for each receptor. The kurtosis value k, a statistical measure of “peakedness,” is located in the upper right corner of each plot. Structures of odourants that elicit strong responses from the most narrowly tuned AgOrs: 2,3-butanedione (AgOr5); 1-octen-3-ol (AgOr8); 2-ethylphenol (AgOr65); indole (AgOr2) are shown above the receptors they activate.

Narrowly tuned receptors for salient odourants

Narrowly tuned receptors have been suggested to be specialist channels that carry information about odourants of high biological relevance17. Consistent with this hypothesis, the most narrowly tuned AgOrs are robustly excited by odourants with high biological salience. Among the receptors that respond strongly to at least one odourant of the panel, the most narrowly tuned are AgOr2, AgOr8, AgOr5, and AgOr65. AgOr2 is narrowly tuned to a small set of aromatics including indole, which was found to constitute nearly 30% of the volatile headspace of human sweat18. AgOr8 responds strongly to 1-octen-3-ol, a human volatile that is a strong attractant for several species of mosquito4. AgOr5 is tuned to 2,3-butanedione, is a metabolic byproduct of human skin microflora19. AgOr65 responds strongly to 2-ethylphenol, which is found in the urine of many animals20, 21. We found no mosquito receptors narrowly tuned to esters or aldehydes, odourants that dominate the headspace of many fruits22, 23. By contrast, among the most narrowly tuned receptors in the fruit fly, the strongest responses are in most cases to esters (DmOr85a, DmOr59b, DmOr67c) or to a terpene that contains an ester group (DmOr82a)12 (Supplementary Table 3b).

Some of the narrowly tuned AgOrs, in addition to responding strongly to odorants with high biological salience, respond with high sensitivity to these odorants. AgOr8 and AgOr2 respond to concentrations of 1-octen-3-ol and indole that range over more than four orders of magnitude and have response thresholds that lie between a 10−7 and a 10−6 dilution (Supplementary Figure 2).

Broadly tuned receptors lie at the other end of the AgOr distribution. It is possible that these receptors act in signaling the presence of odourants but not in specifically identifying or discriminating among them. We found that in An. gambiae, all strong responses to esters and aldehydes are conferred by broadly tuned AgOrs (all have kurtosis values less than the mean; Supplementary Table 2b and Supplementary Table 3a).

Salient odourants that activate specific receptors

“Odourant tuning curves,” the reciprocal of receptor tuning curves, were also generated (Figure 3, Supplementary Table 3c, Supplementary Figure 3). For each odourant we plotted the responses of the 50 receptors along the X-axis, placing the strongest response at the center. Interestingly, the five odourants with the most narrow response distributions are all highly relevant to mosquito ecology. 3-methylindole is an oviposition site volatile that induces egg-laying and ORN responses in Culex mosquitoes24, 25. Indole is another oviposition site volatile26, in addition to being a major component of human emanations that is found both in sweat18 and human breath27. Geranyl acetate and citronellal are emitted from plants that are repellant to An. gambiae28, 29. Dimethylsulfide is emitted in human breath30 and is attractive to the Aedes aegypti mosquito31.

Figure 3. Odourant tuning curves.

Figure 3

Tuning curves for odourants. The responses of the 50 AgOrs are ordered along the X-axis according to the magnitude of the response they generate for each odourant. The receptor with the strongest response is placed at the center of the distribution; those that have the weakest responses are at the edges. The order of receptors is therefore different for each odourant. The kurtosis value is indicated in each graph. Only odourants that generate a strong response (≥100 spikes/second) from at least one receptor are shown. At the top of the panel are the structures of several odourants that generate strong responses from just one or two receptors, shown above the corresponding graph. Shown here are tuning curves for the 12 most narrowly tuned odorants, the 12 most broadly tuned odorants, and 12 representative odorants of intermediate tuning breadth. Tuning curves for 25 additional odourants are shown in Supplementary Figure 3.

D. melanogaster odourant tuning curves were constructed from earlier data12 to identify odourants that likewise strongly activate a small number of receptors (Supplementary Table 3d). In contrast to the An. gambiae tuning curves, many of the odourants with high kurtosis values are esters, the dominant chemical class in fruit emanations23. Together, the mosquito and fruit fly odourant tuning curves support a model in which odourants of particular biological relevance are coded via a small number of channels.

Taken together, these results suggest that odourant and receptor tuning analyses provide complementary avenues to identify receptors and odourants that are important for innate insect behavioral responses. We note with special interest that in An. gambiae, one of the “narrowly tuned” odourants activates one of the narrowly tuned receptors. AgOr2 is strongly activated by indole, the odourant that constitutes almost 30% of the volatile headspace of human sweat18.

Diverse temporal dynamics of receptor responses

DmOrs were previously shown to yield responses that vary widely in their temporal characteristics, and the temporal dynamics were specific to the odourant-receptor combination11,12. To investigate the temporal dynamics of responses from AgOrs, we generated peristimulus time histograms for a number of odourant-receptor combinations (Supplementary Figure 4). As was observed with DmOrs, we find a diversity of temporal dynamics. Some odourants, including the human volatiles 1-octen-3-ol and indole, were capable of generating tonic responses that persisted throughout the three-second analysis period. Linalool oxide also generated a prolonged response. These odourants generated phasic responses from other receptors. The diversity of these responses suggests that temporal features may be a rich source of information about odourant identity.

How the mosquito covers odour space

How is the chemical world represented by the AgOrs? Excitatory activity is not distributed evenly across the odourant panel (Figure 1c, Supplementary Table 2a). For example, 28% of the strong responses were generated by heterocyclics, which constitute only 8% of the odourants in the panel. However, chemical class is only one descriptor of molecular identity. To represent the structural diversity among odourants more fully we adapted a recently developed odourant metric32. This metric is based on an optimized set of 32 molecular descriptors, including functional group, carbon chain length, and other physicochemical properties, which provide the basis of a 32-dimensional coordinate system. Each odourant can be mapped to a unique location in this multidimensional space. Odourants that are structurally similar lie close together in the space, while odourants that are structurally dissimilar lie far apart. To visualize this space we applied principal components analysis (PCA) to project it into two dimensions (Supplementary Figure 5a). As shown by the aromatics, odourants can be of the same chemical class yet map far apart.

Having mapped the odourants of the panel into this chemically defined odour space, we then asked how the AgOr repertoire covers the space. To illustrate the responses elicited by each odourant we generated a bubble plot, in which the location of each bubble indicates odourant identity, and the size of each bubble represents the magnitude of the response to that odourant, summed across all receptors and measured in total spikes/second (Figure 4a).

Figure 4. Distribution of responses across a physicochemical odour space.

Figure 4

(a) Bubble plot of the responses generated by the AgOr repertoire to each of the 110 odourants. Size of the bubble scales with the sum of spikes across all receptors that exhibit at least one response ≥50 spikes/second from at least one odourant on the panel. Each odourant is plotted and colour-coded according to functional group except for 7-octenoic acid and 2-oxohexenoic acid. Shown are the first two principal components of the 32-dimensional physicochemical space (adapted from Haddad, et al., 2008). Descriptors were normalized. Comparison of responses generated by the AgOrs (b) and DmOrs (c) to the set of 53 odourants that were tested against both receptor repertoires. Area of the bubble corresponds to the sum of spikes across all receptors that exhibit at least one response ≥50 spikes/second from at least one odourant on the panel. Responses were normalized to the sum of spikes elicited by all 53 odourants across all AgOrs or all DmOrs that exhibit at least one response ≥50 spikes/second from at least one odourant on the panel. (d) Percent of odourant-receptor combinations that generate strong responses (≥100 spikes/second) within the ester and aromatic classes for An. gambiae and D. melanogaster. Only responsive receptors (those yielding a response ≥50 spikes/second to at least one odourant) were considered.

The AgOr repertoire is sensitive to a broad region of the odour space. However, the responses are not of uniform magnitude across the space. The odourants in the region of the space that is occupied by heterocyclics (purple), for example, elicit greater responses than the odourants in the region that is home to esters (dark green), and much greater responses than those in the region inhabited by carboxylic acids (pink). The AgOr repertoire may have evolved particular sensitivity to certain regions of odour space, such as the region containing aromatics, some of which are major components in human emanations 18, 3339, and heterocyclics, a chemical class that includes volatiles proposed to promote oviposition behavior40. Such enhanced sensitivity could reflect the insect’s investment in detecting and discriminating among chemicals of these classes.

We also investigated the distribution of inhibitory responses across the An. gambiae receptor repertoire, which are not visualized in the odour space described above. As was observed across the D. melanogaster receptor repertoire12, most odourants elicit at least one inhibitory response, and most receptors are inhibited by at least one odourant (Supplementary Table 2a).

Differences between An. gambiae and D. melanogaster odour spaces

We next asked whether the An. gambiae and D. melanogaster Or repertoires differ in their coverage of odour space. As an initial means of addressing this issue, we considered the 53 odourants that were tested against both the AgOr and the DmOr repertoires12 and constructed odour spaces of the type described above for each receptor repertoire.

The two species differed in their relative coverage of odour space. The mosquito allocated greater relative coverage to the aromatics (Figure 4b; dark blue in lower right quadrant). By contrast, the fly devoted greater relative activity to some of the esters (dark green) and one of the two aldehydes that were compared (gray) (Figure 4c, Supplementary Figure 5b). We then compared the two species with respect to the strong responses (≥100 spikes/second). In the mosquito, 15% of the receptor-aromatic combinations yielded strong responses, compared to 7% of the receptor-ester combinations. By contrast, in the fly, 9% of the receptor-aromatic combinations yielded strong responses, compared to 20% of receptor-ester combinations (Figure 4d).

We next considered whether the species differed with respect to another kind of odour space, a biological odour space that relates odourants based on the primary sensory signals they generate. We created a space in which each axis represents the response magnitude (in spikes/second) for one odourant receptor, as described previously12. Odourants that elicit similar patterns of activity across the receptor repertoire map close together. Odourants that are close may be similar in their perceptual qualities, and may be more difficult for the animal to discriminate because they generate similar patterns of ORN activity. Experiments conducted with Drosophila larvae have provided evidence to support a relationship between such odour space distances and perception10.

We constructed such spaces for the mosquito and the fly and depicted them in three dimensions by applying PCA (Figure 5a). Odourants of the same chemical class tend to cluster together (Figure 5a), as previously observed in the fly12. However, some chemical classes were differentially distributed in the odour spaces of the two insects. Esters were more widely distributed in D. melanogaster space than in An. gambiae space, while aromatics were more widely distributed in An. gambiae space (Figures 5b and 5c). To quantify these differences, we calculated the Euclidean distance for every pair of odourants with the same functional group, within the odour space of each species. We found that the mean inter-odourant distance for esters is significantly higher for D. melanogaster than An. gambiae, while the mean inter-odourant distance for aromatics is higher for An. gambiae (p<0.001 for esters, p=0.01 for aromatics, Mann-Whitney); there were no differences in distances among alcohols or ketones, the other groups that could be compared. These results suggest that mosquitoes may be better able than fruit flies to discriminate among aromatics, while fruit flies may be better able to discriminate among esters, perhaps reflecting the biological relevance of these classes of compounds to the animals.

Figure 5. Distribution of odourants in a receptor activity-based odour space.

Figure 5

First three principal components of a receptor activity-based odour space. (a) Left, An. gambiae odour space. Right, D. melanogaster odour space (adapted from Hallem and Carlson, 2006). All odourants tested against a receptor set were considered. Odourants are colour-coded according to functional group. (b) Only esters are shown. Left, An. gambiae odour space; right D. melanogaster odour space. (c) Only aromatics shown. Left, An. gambiae odour space; right D. melanogaster odour space. Mean inter-odourant distance for the set of esters tested against both receptor sets is 254±12 spikes/second for the AgOrs; 353±13 spikes/second for the DmOrs (p<0.05; t-test). Mean inter-odourant distance for the set of aromatics tested against both receptor sets is 406±29 for the AgOrs; 321±17 for the DmOrs (p<0.05; t-test).

Are the functional differences between the mosquito and fruit fly Or repertoires due to a particular clade of mosquito or fly receptors? When the AgOr family was first identified, phylogenetic analysis revealed a clade of An. gambiae odourant receptors with no close D. melanogaster relatives, and a clade of Drosophila odourant receptors with no close An. gambiae relatives6, 7 (see also Supplementary Figure 6). Do these species-specific clades of receptors respond to odourants of a particular kind? We used matrix analysis to generate a dendrogram of receptors (Supplementary Figure 7), and PCA to create a “receptor space,” based on the responses of the receptors to odorants (Supplementary Figure 8). In neither case did we observe clustering of receptors of the species-specific clades. Thus, the observed functional differences between the An. gambiae and D. melanogaster odourant receptor repertoires may reflect evolutionary changes distributed across the receptor repertoires rather than concentrated in a specific branch of the phylogenetic tree.

We note finally that no AgOr showed even a modest response of 50 spikes/second to any carboxylic acid or to any amine in our system. By contrast, some DmOrs respond strongly to certain carboxylic acids12, 41. Recently, a set of variant-ionotropic glutamate receptors that respond to amines and carboxylic acids have been identified in D. melanogaster42. Receptors of this class may detect these compounds in An. gambiae.

Discussion

Here we have functionally characterized the AgOr repertoire of odourant receptors from the mosquito An. gambiae. The olfactory system of An. gambiae allows the insect to locate human blood-meal hosts, thereby facilitating the transmission of malaria. We have identified individual receptors that respond robustly to human volatiles and may be central to the process by which the mosquito identifies its human hosts.

Strikingly, some receptors that respond to human odourants are narrowly tuned and highly sensitive. Reciprocally, some notable odourants elicit strong responses from one or a few receptors. These highly focused relationships between certain odourants and certain receptors suggest a role for specific transmission channels in guiding the animal’s behavior. A recent study of the D. melanogaster antennal lobe documented lateral inhibitory interactions that increased with greater total ORN input43. Odourants that excite few receptors in the mosquito may produce signals that suffer less inhibition and enjoy more saliency. Another study found that ablation of either of two narrowly tuned ORN classes impaired behavioral attraction to their cognate odourants44.

Since our analysis was conducted using the same expression system as our previous study of D. melanogaster odourant receptors, it provided a unique opportunity to compare the Or repertoires of two species that belong to the same order but that exhibit different olfactory-guided behaviors. An outstanding question in the field of olfaction is how an organism’s ecology shapes the function of its odourant receptor repertoire. A full answer to this question requires the functional characterization of entire receptor repertoires. We found that the two species show different coverage of a chemically defined odour space. Certain classes of odourants are differentially distributed in a biologically defined odour space of each species. These differences suggest the evolution of olfactory acuity and discriminatory power consistent with the ecological needs of each species. These evolutionary changes appear to have occurred over the odourant receptor repertoire as a whole, as opposed to having been effected by the emergence of a species-specific receptor clade.

The results may have implications for the control of malaria, one of the world’s most devastating diseases. Screens for activators and inhibitors of selected receptors may identify compounds that attract mosquitoes into traps, interfere with their navigation, or repel them.

Methods summary

AgOr cDNAs were cloned by standard procedures and expressed in the empty neuron as described previously8, 14. Extracellular single-unit physiology was performed as described previously9, 11. Physicochemical odor space was constructed using the set of 32 optimized DRAGON descriptors32, which were normalized.

Methods

Cloning of the AgOrs

Five-day-old laboratory-maintained An. gambiae mosquitoes of the Suakoko strain were cold anesthetized and the antennae, maxillary palpae, and proboscises were dissected by hand on a chill table. RNA was prepared with RNeasy (QIAGEN) according to the manufacturer’s instructions. The RNA preparation was used for oligo dT-primed cDNA synthesis with Superscript II Reverse Transcriptase (Invitrogen) for the generation of templates for subsequent PCR reactions. Negative control samples with no reverse transcriptase were included in each cDNA synthesis and subsequent PCR analysis. PCR was performed with a Mastercycler Gradient (Eppendorf) under the following conditions: 94°C for 5 min; 40 cycles of 94°C for 30 s, 55°C to 60°C for 30 s (annealing temperature varied depending on primer pair), 72°C for 60 s; and 72°C for 7 min. PCR amplification products were separated on a 1.0% agarose gel and were cloned into the Gateway pENTR entry vector (Invitrogen) or pGEM-TEasy (Promega) and verified by sequencing. Sequences of at least two independent clones were obtained for each Or and compared to verify polymorphisms as such rather than PCR errors. Sequence discrepancies were resolved by a second PCR reaction. We amplified 72 AgOrs, in addition to AgOr7, the An. gambiae ortholog of the atypical D. melanogaster odourant receptor gene Or83b45. Four (AgOr37, AgOr40, AgOr52, AgOr58) of the remaining six AgOr genes that were not amplified from adult tissue are expressed in the larval stage46, and the other two may be artifacts of gene annotation.

Drosophila stocks and transgenes

The ab3A mutant flies and Or22a-GAL4 constructs were described previously 9, 11. To generate UAS-Or constructs, the pUAST vector (C. G. Warr) was adapted to generate a Gateway (Invitrogen) compatible destination vector. LR recombination reaction was performed with the Gateway (Invitrogen) pENTR entry clones and the modified pUAST vector. For those amplification products cloned into pGEM-TEasy, subcloning into unadapted pUAST was performed using the restriction enzymes BglII, KpnI, and NotI (New England Biolabs). An exception was AgOr53, which was cloned into pNmyc-UAST vector (C. G. Warr) in frame with the start codon and three copies of the myc tag coding sequence. The resulting protein had an N-terminal myc tag. No other receptors had epitope tags.

Electrophysiology

Extracellular single-unit recordings were performed essentially as described previously 11. Odourant stimuli were prepared in Pasteur pipettes as described previously 11. Chemicals were of the highest purity available (Sigma-Aldrich). Seven-octenoic acid and cis/trans-3-methyl-2-hexenoic acid were synthesized by Richman Chemical (Lower Gywnedd, PA). All chiral chemicals were racemic mixtures with the exception of (+)-carvone, (−)-carvone, (+)-fenchone, (−)-fenchone, and L(+)-lactic acid. Ammonia, cadaverine, putrescine, acetic acid, propanoic acid, butanoic acid, isobutyric acid, isovaleric acid, L(+)-lactic acid, and 2-oxohexenoic acid were diluted in H2O. Hexadecanoic acid, octadecanoic acid, 5α-androsten-3α-ol, were diluted in ethanol. All other odourants were diluted in paraffin oil (Fluka). Liquid odourants were diluted in solvent to a concentration of 10−2 volume/volume, and solid odourants were dissolved, 50 mg in 5 ml solvent.

Stimuli were presented by placing the tip of the pipette through a hole in a tube carrying a purified air stream (32 ml/s) directed at the fly and administering a pulse of charcoal-filtered air (3.2 ml/s) through the pipette containing the odourant. Pulse duration was 500 ms. Stimuli were used for a maximum of four presentations. Responses were quantified by subtracting the number of impulses in 500 ms of unstimulated activity from the number of impulses in the 500 ms following odourant stimulation, after a 150 ms delay to allow the odourant to travel down the airstream. Responses to diluents were also subtracted. For each odourant, each recording was from a separate sensillum, with no more than three sensilla analyzed per fly. Recordings were obtained from flies between 4 and 14 days old.

A panel of six odourants previously tested by E. Hallem against the ab3B neuron were retested in this study to control for possible differences between electrophysiology rigs; none of the responses were significantly different (p ≥0.12 in all cases, t-test).

Data analysis

Principal component analysis (PCA), and hierarchical cluster analysis were performed using PAST, a statistics program (http://folk.uio.no/ohammer/past/) as described previously 12. Physicochemical odour space was constructed using the set of 32 optimized DRAGON descriptors32 (Talete, srl, DRAGON for Windows, version 5.5, 2007, http://www.talete.mi.it/). Descriptors were normalized. The twelve odourants that generate net negative (inhibitory) responses (1-chlorododecane, 3-methyl-2-hexenoic acid, cadaverine, cis-9-octadecanoic acid, delta-decalactone, hexadecanoic acid, nonanoic acid, octadecanoic acid, octanal, octanoic acid, putrescine, and tridecanoic acid) are not shown in the bubble plots of Figure 4. To generate odour spaces and for cluster analyses, we removed from analysis receptors that did not respond to any odourant on the panel with a response ≥50 spikes/second and odourants that did not elicit any responses ≥50 spikes/second (before solvent responses were subtracted) unless otherwise noted. Error bars represent SEM, unless otherwise noted. Phylogenetic analysis was performed with MEGA 4.0.2, using a Neighbor-joining algorithm and 500 replications. Peristimulus time histograms were generated using IGOR Pro 6.0 (Wavemetrics).

Supplementary Material

1
2
3

Acknowledgments

We thank W. van der Goes van Naters and C. Yao for help with electrophysiology, E. Hallem, S. Kreher, J. Salzman and T. Emonet for assistance with data analyses, and T.-W. Koh for comments on the manuscript. We thank P. Graham, Z. Berman, A. Rabin, M. Dillon and E. Kelley-Smith for technical assistance. We thank Y.-T. Qiu for assistance generating Figure 1b and Supplementary Figure 1. This work was funded in part by grants from the Foundation for the National Institutes of Health (NIH) through the Grand Challenges in Global Health Initiative to L.J.Z. and from the NIH to L.Z. and J.R.C.

Footnotes

Author contributions

Electrophysiology and computational analysis were performed by A.C. Molecular cloning was performed by A.C., G.W., and C.-Y.S. A.C. and J.C. wrote the manuscript. All authors contributed to the design and interpretation of the study.

Competing Interests

The authors declare no competing financial interests.

References

  • 1.World Health Organization. World Malaria Report. 2008. [Google Scholar]
  • 2.Zwiebel LJ, Takken W. Olfactory regulation of mosquito-host interactions. Insect Biochem Mol Biol. 2004;34:645–52. doi: 10.1016/j.ibmb.2004.03.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Takken W. The Role of Olfaction in Host-Seeking of Mosquitos - a Review. Insect Science and Its Application. 1991;12:287–295. [Google Scholar]
  • 4.Takken W, Knols BGJ. Odor-mediated behavior of Afrotropical malaria mosquitoes. Annual Review of Entomology. 1999;44:131–157. doi: 10.1146/annurev.ento.44.1.131. [DOI] [PubMed] [Google Scholar]
  • 5.Su CY, Menuz K, Carlson JR. Olfactory perception: receptors, cells, and circuits. Cell. 2009;139:45–59. doi: 10.1016/j.cell.2009.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Hill CA, et al. G protein coupled receptors in Anopheles gambiae. Science. 2002;298:176–178. doi: 10.1126/science.1076196. [DOI] [PubMed] [Google Scholar]
  • 7.Fox AN, Pitts RJ, Robertson HM, Carlson JR, Zwiebel LJ. Candidate odorant receptors from the malaria vector mosquito Anopheles gambiae and evidence of down-regulation in response to blood feeding. Proc Natl Acad Sci U S A. 2001;98:14693–7. doi: 10.1073/pnas.261432998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Hallem EA, Fox AN, Zwiebel LJ, Carlson JR. Olfaction: mosquito receptor for human-sweat odorant. Nature. 2004;427:212–3. doi: 10.1038/427212a. [DOI] [PubMed] [Google Scholar]
  • 9.Dobritsa AA, van der Goes van Naters W, Warr CG, Steinbrecht RA, Carlson JR. Integrating the molecular and cellular basis of odor coding in the Drosophila antenna. Neuron. 2003;37:827–41. doi: 10.1016/s0896-6273(03)00094-1. [DOI] [PubMed] [Google Scholar]
  • 10.Kreher SA, Mathew D, Kim J, Carlson JR. Translation of sensory input into behavioral output via an olfactory system. Neuron. 2008;59:110–24. doi: 10.1016/j.neuron.2008.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Hallem EA, Ho MG, Carlson JR. The molecular basis of odor coding in the Drosophila antenna. Cell. 2004;117:965–79. doi: 10.1016/j.cell.2004.05.012. [DOI] [PubMed] [Google Scholar]
  • 12.Hallem EA, Carlson JR. Coding of odors by a receptor repertoire. Cell. 2006;125:143–60. doi: 10.1016/j.cell.2006.01.050. [DOI] [PubMed] [Google Scholar]
  • 13.Gaunt MW, Miles MA. An insect molecular clock dates the origin of the insects and accords with palaeontological and biogeographic landmarks. Molecular Biology and Evolution. 2002;19:748–761. doi: 10.1093/oxfordjournals.molbev.a004133. [DOI] [PubMed] [Google Scholar]
  • 14.Lu T, et al. Odor coding in the maxillary palp of the malaria vector mosquito Anopheles gambiae. Curr Biol. 2007;17:1533–44. doi: 10.1016/j.cub.2007.07.062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Malnic B, Hirono J, Sato T, Buck LB. Combinatorial receptor codes for odors. Cell. 1999;96:713–23. doi: 10.1016/s0092-8674(00)80581-4. [DOI] [PubMed] [Google Scholar]
  • 16.Saito H, Chi Q, Zhuang H, Matsunami H, Mainland JD. Odor coding by a Mammalian receptor repertoire. Sci Signal. 2009;2:ra9. doi: 10.1126/scisignal.2000016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Wilson RI, Mainen ZF. Early events in olfactory processing. Annu Rev Neurosci. 2006;29:163–201. doi: 10.1146/annurev.neuro.29.051605.112950. [DOI] [PubMed] [Google Scholar]
  • 18.Meijerink J, et al. Identification of olfactory stimulants for Anopheles gambiae from human sweat samples. Journal of Chemical Ecology. 2000;26:1367–1382. [Google Scholar]
  • 19.Verhulst N, et al. Cultured skin microbiota attracts malaria mosquitoes. Malaria Journal. 2009;8:302. doi: 10.1186/1475-2875-8-302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Sun HW, et al. Alcohol, volatile fatty acid, phenol, and methane emissions from dairy cows and fresh manure. Journal of Environmental Quality. 2008;37:615–622. doi: 10.2134/jeq2007.0357. [DOI] [PubMed] [Google Scholar]
  • 21.Gutierrez-Garcia AG, Contreras CM, Mendoza-Lopez MR, Garcia-Barradas O, Cruz-Sanchez JS. Urine from stressed rats increases immobility in receptor rats forced to swim: Role of 2-heptanone. Physiology and Behavior. 2007;91:166–172. doi: 10.1016/j.physbeh.2007.02.006. [DOI] [PubMed] [Google Scholar]
  • 22.Nijssen LMI-V, van CA, Donders JJH, editors. TNO Quality of Life. 2004. [Google Scholar]
  • 23.Morton IDaMAJ., editor. The flavour of fruits. Elsevier Science Publishers; New York: 1990. [Google Scholar]
  • 24.Millar JG, Chaney JD, Mulla MS. Identification of Oviposition Attractants for Culex-Quinquefasciatus from Fermented Bermuda Grass Infusions. Journal of the American Mosquito Control Association. 1992;8:11–17. [PubMed] [Google Scholar]
  • 25.Syed Z, Leal WS. Acute olfactory response of Culex mosquitoes to a human- and bird-derived attractant. Proc Natl Acad Sci U S A. 2009;106:18803–8. doi: 10.1073/pnas.0906932106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Clements AN, Clements AN. The biology of mosquitoes. Chapman & Hall; London ; New York: 1992. [Google Scholar]
  • 27.Kostelc JG, Preti G, Zelson PR, Stoller NH, Tonzetich J. Salivary Volatiles as Indicators of Periodontitis. Journal of Periodontal Research. 1980;15:185–192. doi: 10.1111/j.1600-0765.1980.tb00273.x. [DOI] [PubMed] [Google Scholar]
  • 28.Moore SaDM. In: Insect repellents: Principles, methods, and uses. Debboun M, Frances S, Strickman D, editors. CRC Boca Raton; FL: 2007. pp. 3–30. [Google Scholar]
  • 29.Omolo MO, Okinyo D, Ndiege IO, Lwande W, Hassanali A. Repellency of essential oils of some Kenyan plants against Anopheles gambiae. Phytochemistry. 2004;65:2797–2802. doi: 10.1016/j.phytochem.2004.08.035. [DOI] [PubMed] [Google Scholar]
  • 30.Tangerman A, Meuwesearends MT, Vantongeren JHM. A New Sensitive Assay for Measuring Volatile Sulfur-Compounds in Human Breath by Tenax Trapping and Gas-Chromatography and Its Application in Liver-Cirrhosis. Clinica Chimica Acta. 1983;130:103–110. doi: 10.1016/0009-8981(83)90263-2. [DOI] [PubMed] [Google Scholar]
  • 31.Allan SA, Bernier UR, Kline DL. Evaluation of oviposition substrates and organic infusions on collection of Culex in Florida. Journal of the American Mosquito Control Association. 2005;21:268–273. doi: 10.2987/8756-971X(2005)21[268:EOOSAO]2.0.CO;2. [DOI] [PubMed] [Google Scholar]
  • 32.Haddad R, et al. A metric for odorant comparison. Nat Methods. 2008;5:425–9. doi: 10.1038/nmeth.1197. [DOI] [PubMed] [Google Scholar]
  • 33.Logan JG, et al. Identification of human-derived volatile chemicals that interfere with attraction of Aedes aegypti mosquitoes. Journal of Chemical Ecology. 2008;34:308–322. doi: 10.1007/s10886-008-9436-0. [DOI] [PubMed] [Google Scholar]
  • 34.Curran AM, Ramirez CF, Schoon AA, Furton KG. The frequency of occurrence and discriminatory power of compounds found in human scent across a population determined by SPME-GEMS. Journal of Chromatography B-Analytical Technologies in the Biomedical and Life Sciences. 2007;846:86–97. doi: 10.1016/j.jchromb.2006.08.039. [DOI] [PubMed] [Google Scholar]
  • 35.Curran AM, Rabin SI, Prada PA, Furton KG. Comparison of the volatile organic compounds present in human odor using SPME-GC/MS. J Chem Ecol. 2005;31:1607–19. doi: 10.1007/s10886-005-5801-4. [DOI] [PubMed] [Google Scholar]
  • 36.Cork A, Park KC. Identification of electrophysiologically-active compounds for the malaria mosquito, Anopheles gambiae, in human sweat extracts. Medical and Veterinary Entomology. 1996;10:269–276. doi: 10.1111/j.1365-2915.1996.tb00742.x. [DOI] [PubMed] [Google Scholar]
  • 37.Bernier UR, Kline DL, Schreck CE, Yost RA, Barnard DR. Chemical analysis of human skin emanations: comparison of volatiles from humans that differ in attraction of Aedes aegypti (Diptera: Culicidae) J Am Mosq Control Assoc. 2002;18:186–95. [PubMed] [Google Scholar]
  • 38.Bernier UR, Kline DL, Barnard DR, Schreck CE, Yost RA. Analysis of human skin emanations by gas chromatography/mass spectrometry. 2. Identification of volatile compounds that are candidate attractants for the yellow fever mosquito (Aedes aegypti) Analytical Chemistry. 2000;72:747–756. doi: 10.1021/ac990963k. [DOI] [PubMed] [Google Scholar]
  • 39.Gallagher M, et al. Analyses of volatile organic compounds from human skin. British Journal of Dermatology. 2008;159:780–791. doi: 10.1111/j.1365-2133.2008.08748.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Lindh JM, Kannaste A, Knols BGJ, Faye I, Borg-Karlson AK. Oviposition Responses of Anopheles gambiae s.s. (Diptera: Culicidae) and Identification of Volatiles from Bacteria-Containing Solutions. Journal of Medical Entomology. 2008;45:1039–1049. doi: 10.1603/0022-2585(2008)45[1039:oroags]2.0.co;2. [DOI] [PubMed] [Google Scholar]
  • 41.Kreher SA, Kwon JY, Carlson JR. The molecular basis of odor coding in the Drosophila larva. Neuron. 2005;46:445–56. doi: 10.1016/j.neuron.2005.04.007. [DOI] [PubMed] [Google Scholar]
  • 42.Benton R, Vannice KS, Gomez-Diaz C, Vosshall LB. Variant ionotropic glutamate receptors as chemosensory receptors in Drosophila. Cell. 2009;136:149–62. doi: 10.1016/j.cell.2008.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Olsen SR, Wilson RI. Lateral presynaptic inhibition mediates gain control in an olfactory circuit. Nature. 2008;452:956–60. doi: 10.1038/nature06864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Schlief ML, Wilson RI. Olfactory processing and behavior downstream from highly selective receptor neurons. Nat Neurosci. 2007;10:623–30. doi: 10.1038/nn1881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Larsson MC, et al. Or83b encodes a broadly expressed odorant receptor essential for Drosophila olfaction. Neuron. 2004;43:703–14. doi: 10.1016/j.neuron.2004.08.019. [DOI] [PubMed] [Google Scholar]
  • 46.Xia Y, et al. The molecular and cellular basis of olfactory-driven behavior in Anopheles gambiae larvae. Proc Natl Acad Sci U S A. 2008;105:6433–8. doi: 10.1073/pnas.0801007105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Yao CA, Ignell R, Carlson JR. Chemosensory coding by neurons in the coeloconic sensilla of the Drosophila antenna. J Neurosci. 2005;25:8359–67. doi: 10.1523/JNEUROSCI.2432-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Benton R, Vannice KS, Vosshall LB. An essential role for a CD36-related receptor in pheromone detection in Drosophila. Nature. 2007;450:289–93. doi: 10.1038/nature06328. [DOI] [PubMed] [Google Scholar]
  • 49.Jin X, Ha TS, Smith DP. SNMP is a signaling component required for pheromone sensitivity in Drosophila. Proc Natl Acad Sci U S A. 2008;105:10996–1001. doi: 10.1073/pnas.0803309105. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1
2
3

RESOURCES