You are on page 1of 13

Insect Biochemistry and Molecular Biology 34 (2004) 653–665

www.elsevier.com/locate/ibmb

The molecular basis of insecticide resistance in mosquitoes


Janet Hemingway , Nicola J. Hawkes, Lynn McCarroll, Hilary Ranson
Liverpool School of Tropical Medicine, Pembroke Place, Liverpool L3 5QA, UK

Received 8 March 2004; accepted 18 March 2004

Abstract

Insecticide resistance is an inherited characteristic involving changes in one or more insect gene. The molecular basis of these
changes are only now being fully determined, aided by the availability of the Drosophila melanogaster and Anopheles gambiae
genome sequences. This paper reviews what is currently known about insecticide resistance conferred by metabolic or target site
changes in mosquitoes.
# 2004 Elsevier Ltd. All rights reserved.

Keywords: Esterase; Insecticide resistance; Molecular mechanism; Glutathione S-transferase (GST); GABA; rdl; Sodium channel; kdr;
Carboxylesterase; P450 monooxygenase; Acetylcholinesterase

1. Introduction they are used as indoor residual house sprays to


impregnate bednets, curtains and screens, and in coils,
In this review we discuss the data currently available
mats and aerosols. Pyrethroids are popular due to their
regarding chemical insecticide resistance in mosquitoes.
Most resistance mechanisms can be divided into two very low toxicity in humans, and rapid killing effect on
groups, metabolic (alterations in the levels or activities the insect, although this varies from product to product
of detoxification proteins), and target site (mutations in and overall efficacy of treatment and knock-down
the sodium channel, acetylcholinesterase and GABA properties depend on irritancy and excito-repellancy
receptor genes). Alone or in combination these (Hougard et al., 2002). For example, bifenthrin has a
mechanisms confer resistance, sometimes at an much lower irritancy effect than deltamethrin, thus the
extremely high level, to all of the available classes of mosquito will stay on bifenthrin-treated material for a
insecticides. In addition, many insecticides such as longer period of time and so obtain a higher dose of
DDT and permethrin also influence behavioural chan-
insecticide. However, the well-documented cross-resist-
ges in the insect—for example, by reducing the rate of
ance that occurs between pyrethroids means that as a
mosquito entry into houses, increasing the rate of early
exit from houses and inducing a shift in biting times replacement for permethrin, bifenthrin may not be
(Lines et al., 1987; Miller et al., 1991; Mbogo et al., effective for long (Chandre et al., 1999). Since pyre-
1996; Mathenge et al., 2001). Some mosquitoes have throids are the only option for treating bednets for the
also evolved thicker or altered cuticles, reducing pen- control of malaria at present, it is important that the
etration of the insecticide (Stone and Brown, 1969; efficacy of this class is maintained for as long as poss-
Apperson and Georghiou, 1975). ible. Worryingly, resistance to pyrethroids may be
Pyrethroids account for approximately 25% of the achieved through several of the mechanisms described
world insecticide market. For the public health market,
below. The use of insecticide mixtures or rotational use
of insecticides to delay the development and/or spread

of resistance is one strategy worth consideration
Corresponding author. Tel.: +44-151-705-3261; fax: +44-151-707-
0155. (Curtis et al., 1998) and is currently under investigation
E-mail address: hemingway@liv.ac.uk (J. Hemingway). on a large scale in Mexico (Penilla et al., 1998).
0965-1748/$ - see front matter # 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ibmb.2004.03.018
654 J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665

2. Metabolic resistance mechanisms downstream of esta21 (Fig. 1; Hemingway et al., 2000;


Coleman et al., 2002).
2.1. Carboxylesterases In C. p. pipiens, early estimates of gene copy number
elevated ~250-fold relative to susceptible mosquitoes
2.1.1. Quantitative changes
(Mouches et al., 1990) have subsequently been reduced
Over-production of non-specific carboxylesterases as
to ~20-fold amplification, although whether this reflects
an evolutionary response to organophosphorus and
technical inaccuracies or an actual lessening of the
carbamate insecticide selection pressure has been docu-
resistance level within the laboratory population with
mented in numerous arthropod species including mos-
time is unknown (Guillemaud et al., 1997). In situ
quitoes, cattle ticks, aphids and cockroaches. The most
hybridisations have revealed a tandem arrangement of
widely studied mosquito species demonstrating this
resistance mechanism are members of the Culex genus, the amplified genes in both C. p. pipiens and in C. p.
including Culex pipiens pipiens, C. p. quinquefasciatus molestus (Nance et al., 1990; Tomita et al., 1996).
and C. tritaeniorhynchus (reviewed in Hemingway and In the OP-resistant C. p. quinquefasciatus laboratory
Karunaratne, 1998). These enzymes are B esterases strain PelRR, the esta21 and estb21 genes are physically
with an active site serine residue (Aldridge, 1953). In very close, and oriented head-to-head (Vaughan et al.,
OP-susceptible insects, the active oxon analogues of the 1997). They are thus divergently transcribed. This is also
insecticides act as esterase inhibitors, because they are true of the alleles in the OP-susceptible strain. Despite
poor substrates with a high affinity for the enzymes. linear (1:1) gene co-amplification within the genome,
Esterases from susceptible insects are significantly less estb21 transcript levels are 2–30-fold higher than esta21
reactive with xenobiotics (including insecticides) than transcript levels within individual adult mosquitoes
their counterparts from resistant insects, which seques- (Paton et al., 2000), and on average 3-fold more Estb21
ter the oxon analogues and thus protect the acet- protein is found in a mass homogenate of resistant
ylcholinesterase target site (Karunaratne et al., 1995). insects than Esta21 (Karunaratne, 1994). Functional
The amplified esterases may comprise up to 0.4% of promoter analysis has shown that the b-esterase pro-
total protein, all with rapid sequestration and slow moters are active in vitro and that transcription from
insecticide hydrolysis rates (Karunaratne et al., 1993). the estb21 promoter requires an initiator sequence
The predominant cause of this excessive enzyme syn- approximately 125 bp upstream of the initiating meth-
thesis is amplification of the gene(s) within the genome ionine, but does not involve a TATA-box (Hemingway
(Mouches et al., 1986; Vaughan and Hemingway, 1995; et al., 1998; Hawkes and Hemingway, 2002). Rather, the
Vaughan et al., 1995), although up-regulated transcrip- presence of a site, 28 bp downstream of the initiator,
tion without an underlying gene amplification event perfectly matching the Drosophila downstream pro-
has been reported (Rooker et al., 1996). There are moter element (DPE) appears important (Burke and
many esterase alleles associated with resistance. The Kadonaga, 1996; Hawkes, unpublished data).
most common genotype is the co-amplification of two Recent studies have found an inverse correlation
esterase genes (esta21 and estb21) on an amplicon of between esterase activity in C. p. quinquefasciatus and
approximately 28 kbp, which is reiterated up to 80 filarial burden (McCarroll et al., 2000), suggesting that
times in the genomes of highly resistant C. p. quinque- high levels of esterase are adversely affecting the sur-
fasciatus mosquitoes (Vaughan et al., 1997; Paton et al., vival and/or development of Wuchereria bancrofti filar-
2000). This genotype is present in >90% of resistant C. iae, but the overall effects on transmission are
p. quinquefasciatus. Its selective advantage may arise unknown.
not only from the contributions of two insecticide- There are many reports of enhanced esterase activi-
sequestering esterases, but also from the presence of a ties in other mosquitoes, for example, in permethrin-
third gene with putative detoxification capacity on the resistant An. gambiae (Vulule et al., 1999) and An.
amplicon. A full-length aldehyde oxidase gene lies just albimanus (Brogdon and Barber, 1990), and in resistant

Fig. 1. Schematic of the amplicon from the organophosphate-resistant Culex quinquefasciatus strain PelRR, showing the relative proximities and
transcriptional orientation of three detoxification genes. Arrows indicate the direction of transcription.
J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665 655

Ae. aegypti (Mourya et al., 1993). However the gene(s) a notable exception. The majority of eukaryotic P450s
involved are unknown. Whilst homologs of the co- require the flavoprotein NADPH cytochrome P450
amplified Culex carboxylesterases have been identified reductase, and sometimes cytochrome b5, for activity.
in An. gambiae in the same physical proximity and Diversity is conferred by the existence of multiple P450
orientation (Ranson et al., 2002a), it remains to be seen isoforms, different expression patterns and wide sub-
whether they are involved in resistance, particularly strate spectra (for recent reviews see Scott and Wen,
since the Culex esterases are ineffective against pyre- 2001; Feyereisen, 1999).
throids (Karunaratne et al., 1993). The amplified car- There are many reports demonstrating elevated P450
boxylesterases of some insects (the cattle tick monooxygenase activities in insecticide-resistant mos-
Boophilus microplus and the peach-potato aphid Myzus quitoes, frequently in conjunction with altered activities
persicae) do have activity against both pyrethroids and of other enzymes. For example, Vulule et al. (1999)
OPs (Hernandez et al., 2002; Devonshire and Moores, demonstrated elevated oxidase and esterase levels
1982). However, there is as yet no evidence of this in permethrin-resistant An. gambiae from Kenya.
mechanism in pyrethroid-resistant mosquitoes. Brogdon et al. (1999a,b) have reported oxidase-based
and esterase-based resistance mechanisms alone and in
2.1.2. Qualitative changes combination in permethrin-resistant An. albimanus
In some mosquito species with resistance, elevated from Guatemala. However, the identification of the
carboxylesterase activity involves rapid hydrolysis of particular P450 gene(s) associated with resistance has
the insecticide, rather than increased sequestration. been far from easy in any insect.
This mechanism is almost always found in association In most cases where a link between insecticide resist-
with malathion resistance, and gives a much narrower ance and elevated P450 activity has been shown, the
cross-resistance spectrum (sometimes malathion-spe- Cyp gene belongs to the Cyp6 family. For example,
cific) than the amplified esterase-based mechanism. CYP6D1 is over-produced in pyrethroid-resistant M.
Malathion carboxylesterase resistance has been found domestica due to up-regulated transcription (Kasai and
in An. culicifacies, An. stephensi and An. arabiensis Scott, 2000) and CYP6A1 is similarly associated with
(Herath et al., 1987; Hemingway, 1982a,b, 1983). organophosphate resistance (Sabourault et al., 2001;
Although the genetic alteration(s) generating these Andersen et al., 1994). Drosophila CYP6A2 expressed
qualitative changes have not as yet been identified in in vitro is capable of metabolising diazinon and cyclo-
mosquito populations, data from other arthropods sug- dienes (Dunkov et al., 1997). Amino acid substitutions
gest that only one or two amino acid mutations may be in the resistant Cyp6a2 allele confer increased activity
responsible. Qualitative changes conferring resistance against DDT (Berge et al., 1998). Up-regulation of
to malathion have been reported in carboxylesterase Cyp6g1 orthologs are linked with DDT resistance in
genes from Musca domestica, the sheep blowfly Lucilia field isolates of both Drosophila melanogaster and D.
cuprina and the parasitoid wasp Anisopteromalus simulans (Daborn et al., 2002). Laboratory selection of
calandrae (Zhu et al., 1999; Campbell et al., 1998; D. melanogaster with DDT also results in up-regulation
Claudianos et al., 2002). In each case, the same trypto- of Cyp6g1 (sometimes in conjunction with Cyp12d1),
phan residue is mutated, to leucine in the blowfly and or Cyp6a8 (Brandt et al., 2002; Le Goff et al., 2003).
housefly, and to glycine in the wasp, leading to the The presence of a transposable element at the 50 end of
speculation that a similar mutation in the orthologous Cyp6g1 in DDT-resistant alleles from multiple sites
gene in the above Anopheles species may thus also be worldwide suggests that this genetic event occurred
mutated in malathion resistance. only once. Its selective advantage over other p450 up-
The above carboxylesterase-based resistance regulation genotypes may be due to the wider cross-
mechanisms are not mutually exclusive; both malathion resistance (to imidacloprid, malathion and lufenuron)
carboxylesterase and amplified carboxylesterase conferred (Le Goff et al., 2003; Daborn et al., 2002).
mechanisms have been found in C. tarsalis (Ziegler Prior to the release of the An. gambiae genome (Holt
et al., 1987). et al., 2002), Ranson et al. (2002b) had identified 34
P450 genes. Subsequent analysis of genome data
2.2. P450 monooxygenases revealed a total of 111 P450 monooxygenases in An.
gambiae (Ranson et al., 2002a), arranged mostly in
Cytochrome P450-dependent monooxygenases are an clusters within the genome. Nikou et al. (2003) have
important and diverse family of hydrophobic, heme- recently shown elevated transcript levels (not due to
containing enzymes involved in the metabolism of gene amplification) of an adult-specific Cyp6 P450
numerous endogenous and exogenous compounds. gene, Cyp6z1, in a pyrethroid-resistant strain of An.
This generally results in the detoxification of the sub- gambiae from East Africa. This gene lies within a clus-
strate, although the activation of OP insecticides from ter of 17 P450 genes located within the boundaries of a
the phosphorothionate to the more toxic oxon form is quantitative trait locus on chromosome arm 3R asso-
656 J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665

ciated with permethrin resistance in this strain (Ranson identity are generally assigned to the same class, with
et al., 2004). A second QTL on chromosome arm 2L other properties such as phylogenetic relationships,
coincides with the para voltage-gated sodium channel immunological properties, tertiary structure and the
gene, and Ranson et al. (2000a,b) identified a leucine to ability to form heterodimers also employed. At least six
serine mutation within a codon frequently associated classes of insect GSTs have been identified in An.
with pyrethroid resistance in many arthropods. gambiae (Ranson et al., 2002a), found in several large
A further Cyp6 gene implicated in insecticide resist- clusters on all three chromosomes. The Delta and Epsi-
ance is Cyp6f1 from the mosquito C. p. quinquefascia- lon classes found exclusively in insects are the largest
tus. Kasai et al. (2000) reported slightly elevated levels classes of insect GSTs. Members of both these classes
(~2.5-fold) of Cyp6f1 transcript in a strain with perme- have been implicated in resistance to all the major clas-
thrin resistance. Deltamethrin resistance in a close rela- ses of insecticide. The other insect GSTs belong to the
tive, C. p. pallens, appears to involve up-regulation of Omega, Sigma, Theta and Zeta classes (Tang and Tu,
members of the Cyp4 family (Shen et al., 2003). Cyp4 1994; Wang et al., 1991; Board et al., 1997, 2000;
gene over-expression has also been found in pyre- Huang et al., 1998; Wei et al., 2001; Ranson et al.,
throid-resistant Drosophila (Amichot et al., 1994) and 2002a; Vontas et al., 2002; Ortelli et al., 2003). Despite
other species (Pridgeon et al., 2003; Pittendrigh et al., the availability of full genome sequence (Holt et al.,
1997). 2002), the possibility that further insect GST classes
exist cannot be discounted, as at least three of the
2.3. Glutathione S-transferases An. gambiae GSTs could not be reliably assigned to
any of the pre-designated classes (Ding et al., 2003).
Glutathione S-transferases (GSTs) are found ubiqui- Individual GST subunits are assigned names indicat-
tously in aerobic organisms. Their central role in ing the species they were isolated from and the GST
detoxification and drug resistance pathways in mam- class. They are also given a number that may either
mals has now been established but additional functions reflect the order of discovery or the genomic organis-
are continually being attributed to this complex ation. For example, AgGSTd12 is the 12th member of
enzyme family (see Ranson and Hemingway, 2004). the An. gambiae Delta class of GSTs to be identified.
There are two distantly related classes of GSTs, classi- The wide range of functions performed by GSTs is
fied according to their location within the cell: micro- aided by the extensive nature of the GST supergene
somal or cytosolic. There are only three distinct family in insects and the broad substrate specificities of
microsomal GSTs transcribed in An. gambiae. Their individual enzymes. Dramatic changes in substrate
homology with vertebrate microsomal GSTs suggests a specificity can be achieved by a small number of amino
similar trimeric structure, with a molecular mass of acid substitutions (Ortelli et al., 2003). The primary
approximately 50 kDa. Their tertiary structure in vivo function of GSTs is generally considered to be the
has not been resolved. This class will not be discussed detoxification of both endogenous and xenobiotic com-
further. pounds either directly or by catalysing the secondary
The vast majority of GSTs are cytosolic dimeric pro- metabolism of a vast array of compounds oxidised by
teins comprising two subunits each around 24–28 kDa the cytochrome P450 family. GSTs also play an impor-
in size. Each subunit has two binding sites, the G site tant role in stress physiology and have been implicated
and the H site. The highly conserved G site binds the in intracellular transport and various biosynthetic
tripeptide glutathione and is largely composed of pathways (Wilce and Parker, 1994).
amino acid residues found in the N terminal of the Elevated GST activity has been implicated in resist-
protein. The H site or substrate binding site is more ance to at least four classes of insecticides. Higher
variable in structure and is largely formed from resi- enzyme activity is usually due to an increase in the
dues in the C-terminal (Mannervik, 1985). The active amount of one or more GST enzymes, either as a result
site residue at the N termini of the protein interacts of gene amplification or more commonly through
with and activates the sulphydryl group of glutathione increases in transcriptional rate, rather than qualitative
to generate the catalytically active thiolate anion changes in individual enzymes (for recent review see
(Armstrong, 1991). This nucleophilic thiolate anion is Ranson and Hemingway, 2004).
then capable of attacking substrates bound to the H Resistance due to increases in GST activity was first
site. The active site residue is generally conserved identified in organophosphate (OP) resistance, and
within GST classes but differs between the classes GSTs have now been implicated in OP resistance in
(Wilce and Parker, 1994; Agianian et al., 2003; many insect species (Hayes and Wolf, 1988). Recombi-
Sheehan et al., 2001). nant GST enzymes from the diamondback moth and
Kinetic characteristics have proven a poor criterion housefly have verified the role of these enzymes in OP
for classification of GSTs (Ranson and Hemingway, metabolism (Huang et al., 1998; Wei et al., 2001).
2004). GST sequences that share over 40% sequence Detoxification occurs via an O-dealkylation or O-dear-
J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665 657

ylation reaction. In O-dealkylation, glutathione is con- 2.4. The role of Regulatory Factors
jugated with the alkyl portion of the insecticide (Oppe-
noorth et al., 1979), whereas the reaction of As described above, transcriptional up-regulation
glutathione with the leaving group (Chiang and Sun, underlies many instances of insecticide resistance, yet
the factors responsible for this are largely undeter-
1993) is an O-dearylation reaction. GSTs can also cata-
mined. Amino acid alterations may be important in
lyse the secondary metabolism of OP insecticides. This
some instances, e.g. CYP6A2 (Berge et al., 1998), but
insecticide class is usually applied in its non-insecticidal
cis- and trans-acting factors are also involved. In
phosphorothionate form and activated to the insecti- houseflies, the expression of Cyp6a1 is influenced by a
cidal form by the action of cytochrome P450s in the trans-acting factor (Carino et al., 1994). Sabourault
insect. GSTs that can detoxify the active oxon ana- et al. (2001) found a strong correlation between the
logue have been described in mosquitoes and elevated presence of an ali-esterase mutation (G137D: G119D
activity of these enzymes is a cause of OP resistance in by Torpedo numbering) or of a deletion at this locus
An. subpictus (Hemingway et al., 1991). and the over-production of CYP6A1, with concomitant
Dehydrochlorination of DDT is a major route of diazinon resistance. The authors suggested that
detoxification in insects. The enzyme catalysing this hydrolysis by the wild-type ali-esterase generates a pro-
reaction was not initially recognised as a member of duct that represses Cyp6a1 expression; absence of wild-
the GST family, partly due to the inability to detect type esterase results in transcriptional up-regulation of
even a transient GSH conjugate and also because of this OP-metabolising p450, and potentially other
the lack of DDT dehydrochlorinase activity in some detoxification genes also. Housefly Cyp6d1 expression
purified GSTs. Evidence that DDT dehydrochlorinase also involves cis- and trans-acting regulators (Liu and
is a GST was finally provided when three enzymes with Scott, 1997). In insecticide-resistant Drosophila strains,
CDNB conjugating activity were isolated from a DDT- mutation in a trans-acting repressor permits constitut-
resistant strain of housefly. Two of these were also ive over-expression of Cyp6a8 (Maitra et al., 1996,
2000, 2002).
active with DDT as a substrate (Clark and Shamaan,
Most studies of GSTs suggest that regulation occurs
1984). The DDT dehydrochlorinase reaction proceeds
at the transcriptional level. Several regulatory elements
via a base abstraction of hydrogen, catalysed by the
have been identified in the promoter regions of GSTs
thiolate anion generated in the active site of the GST, that may mediate their induction but the significance of
leading to the elimination of chlorine from DDT, gen- these findings is unclear in the absence of functional
erating DDE. An increased rate of dehydrochlorination studies. In Ae. aegypti, a mutation in a trans-acting
confers resistance to DDT in Ae. aegypti (Grant et al., repressor element is the proposed mechanism for the
1991) and An. gambiae (Prapanthadara et al., 1993). enhanced expression of a Delta class GST in a DDT-
GSTs have no direct role in the metabolism of pyre- resistant strain (Grant and Hammock, 1992). Genetic
throid insecticides but they play a very important role mapping of the major genes controlling GST-based
in conferring resistance to this insecticide class by DDT resistance in An. gambiae also provided tentative
detoxifying lipid peroxidation products induced by pyr- evidence for a trans-acting regulator (Ranson et al.,
ethroids (Vontas et al., 2001). GSTs may also protect 2000a,b) although in this species, mutations in pro-
against pyrethroid toxicity in insects by sequestering moter elements of the Epsilon GST cluster are also
the insecticide (Kostaropoulos et al., 2001). associated with resistance (Ranson et al., 2001).
GSTs play a vital role in the inactivation of toxic The significance of alternative splicing in regulating
products of oxygen metabolism. Reactive oxygen spe- GST expression has not been fully investigated. Four
cies (ROS), including hydrogen peroxide, superoxide alternative transcripts of a Delta class gene and two
anions and hydroxyl radicals, are generated during of a Sigma GST have been detected in An. gambiae (Ran-
son et al., 1998), and it is possible that different inducers
aerobic respiration. These ROS trigger a cascade of
or stress treatments may affect the choice of splice site.
reactions that can be highly damaging to the cells. The
Maitra et al. (2002) have suggested that inefficient spli-
generation of ROS initiates the conversion of poly-
cing may generate different isoforms of CYP6A8.
unsaturated fatty acids to lipid hydroperoxides that are The majority of individual insect GSTs whose
toxic at high concentrations, but some of these have expression profile has been determined are expressed
important signalling functions affecting cell prolifer- constitutively in all life stages. Recent experiments
ation, apoptosis and differentiation at low concentra- found that in An. gambiae, transcripts for all but one
tions (Sawicki et al., 2003). Peroxidase activity has of the GST supergene family were detectable in one-
been detected in insect GSTs from the Delta, Epsilon day-old adult mosquitoes by RT-PCR (Ding et al.,
and Sigma classes (Vontas et al., 2001; Singh et al., 2003). It is apparent from several studies in other
2001; Ortelli et al., 2003). insects that the levels of individual enzymes can fluctu-
658 J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665

ate widely during the life span of an insect and in dif- bamate insecticides. AChE has a key role in the
ferent insect tissues. nervous system, terminating nerve impulses by catalys-
In a number of studies insects have been exposed to ing the hydrolysis of the neurotransmitter acetylcho-
various xenobiotics to determine the effect of these che- line. The insecticides inhibit enzyme activity by
micals on GST expression levels (Ottea and Plapp, covalently phosphorylating or carbamylating the serine
1984; Yu, 1984, 1996). In the majority of cases, a residue within the active site gorge (Corbett, 1974).
model substrate, such as 1-chloro-2,4-dinitrobenzene Studies on An. albimanus from Central America have
(CDNB), is used to assay GST activity from crude shown that an altered AChE is the most common OP/
insect homogenates and hence the results are general carbamate resistance mechanism (Ayad and Geor-
measurements of the activity of a large subset of GSTs ghiou, 1979; Hemingway and Georghiou, 1983), and
and do not reveal much about fluctuations in the level this is true also in Mexican populations of the same
of individual enzymes. A detailed account of the effect mosquito (Penilla et al., 1998). Insensitive AChE has
of dietary compounds, insecticides and laboratory been reported from C. p. pipiens, C. p. quinquefasciatus
inducers on GST expression is provided in two review and C. tritaeniorhynchus, An. nigerimus, An. atroparvus
articles (Clark, 1989; Yu, 1996). Studies using GST and An. sacharovi (Villani and Hemingway, 1987;
inducers have demonstrated that multiple mechanisms Bisset et al., 1990; Hemingway et al., 1985, 1986;
are involved in the regulation of these enzymes. Up- Hemingway, 1982a,b) but in An. gambiae insensitive
regulation of some p450s after exposure to insecticide AChE has only recently been described (N’Guessan
or to phenobarbital has been well documented (e.g. et al., 2003).
Brandt et al., 2002; Sierra-Santoyo et al., 2000; Maitra Across all insect species there are two very distinct
et al., 1996; Carino et al., 1994). Barbie box response types of target site resistance, conferring high carba-
elements have been identified in Drosophila Cyp6a8 and mate and low OP resistance, or high OP with either
Cyp6a2 promoters (Maitra et al., 2002; Dombrowski equivalent or low carbamate resistance (Russell et al.,
et al., 1998). 2004). All the mosquito AChE target site resistances to
The simultaneous over-expression of multiple Cyp date fall into the former category.
genes has been reported in Drosophila and M. domes- In Drosophila there is only one gene encoding AChE,
tica (Amichot et al., 1994; Brandt et al., 2002; Le Goff ace (Fournier et al., 1989), and various point mutations
et al., 2003; Carino et al., 1992). Ding et al. (2003) have been described in resistant strains (Mutero et al.,
identified differential up-regulation of specific An. 1994). Two of these residues are also altered in the sin-
gambiae GST genes within a cluster. Le Goff et al. gle ace gene from resistant M. domestica (Walsh et al.,
(2003) reported up-regulation of GST and Cyp genes in 2001); and one is found in conjunction with a novel
DDT-resistant D. simulans using a D. melanogaster mutation in OP-resistant olive fruit fly, Bactrocera
microarray. Microarrays of whole genomes or of oleae (Vontas et al., 2002). Individual mutations confer
enzyme families (e.g. detoxification enzymes) will be a a low level of insensitivity without dramatically affect-
powerful tool in dissecting resistance (Ranson, pers. ing the rate of neurotransmitter hydrolysis; combina-
commun.; Le Goff et al., 2003). Vulule et al. (1999) tions of mutations generate enhanced resistance. These
found a correlation between elevated p450 and carbox- mutations constrict the neck of the active site gorge,
ylesterase activity levels in a resistant strain of An. limiting insecticide access to the catalytic residues loca-
gambiae, suggesting co-regulation. The availability of ted at the base of the gorge.
genome sequence for An. gambiae (Holt et al., 2002) In C. p. pipiens and in C. tritaeniorhynchus, the pres-
and D. melanogaster (Misra et al., 2002) may reveal ence of at least two AChE genes was inferred from the
similarities within promoters of co-expressed genes that findings that AChE-based resistance mapped to chro-
will lead to the identification of their regulators, and mosome II, whereas the cloned AChE gene (ace-2) is
potentially to novel control methods. sex-linked (Bourguet et al., 1996; Malcolm et al., 1998;
Finally, it should be remembered that most studies Mori et al., 2001). Recently, the sequences of two
on gene regulation rely on the detection of steady state AChE genes (ace-1 and -2) were identified from the An.
mRNA levels. The role of differential stability of tran- gambiae genome (Holt et al., 2002); they show only
scripts is often overlooked. 53% identity. A partial fragment of ace-1 was amplified
by degenerate PCR from C. p. pipiens (Weill et al.,
2002), and tight linkage between this gene and pro-
3. Target site resistance
poxur resistance was demonstrated using an RFLP
3.1. Insensitive acetylcholinesterase within the coding region. However, sequence analysis
of this fragment did not show any likely resistance-
Biochemical assays in several mosquito species have associated mutations. Homologous ace-1 fragments
identified insecticide resistance due to insensitive acet- were also isolated from several other mosquito species,
ylcholinesterase (AChE), the target site of OP and car- including An. gambiae and Ae. aegypti. Subsequently,
J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665 659

the entire ace-1 coding regions from susceptible and 3.2. The GABA receptor mutation
resistant C. p. pipiens strains were compared. A single
amino acid substitution (G119S) was identified, and The target site of cyclodiene insecticides such as diel-
found also in other insensitive AchE strains of C. p. drin is the type A receptor for the neurotransmitter c-
pipiens, C. p. quinquefasciatus and An. gambiae (Weill aminobutyric acid (GABA). Binding of GABA to the
et al., 2003). This glycine residue lies at the base of the receptor elicits rapid gating of an integral chloride-
active site gorge, close to the oxyanion hole. selective ion channel. GABA receptors comprise five
Intriguingly, a different mutation (G119D) at the subunits arranged around the central ion channel. They
oxyanion hole of related carboxylesterases is respon- have a wide distribution in both vertebrates and inver-
sible for OP-specific resistance in L. cuprina and tebrates. Insect resistance to cyclodienes is widespread,
M. domestica (Newcomb et al., 1997; Campbell et al., and the cause of this insensitivity has been most exten-
1998; Claudianos et al., 1999), and linked with p450 sively studied in D. melanogaster (reviewed in Hosie
over-expression (Sabourault et al., 2001), whilst mala- et al., 1997). Mutations at a single codon in the Rdl
thion-specific resistance is conferred by mutations at a (resistance to dieldrin) gene (encoding one receptor
tryptophan residue as discussed earlier. subunit), from an alanine residue to a serine or more
A mutation in the ace-1 ortholog from Japanese C. rarely to a glycine, have been documented in all diel-
tritaeniorhynchus has recently been described, associa- drin-resistant insect species to date (ffrench-Constant
ted with high levels of OP resistance (Nabeshima et al., et al., 1998), although the only mosquito studied thus
2004). The mutation (F455W, corresponding to F331 far is Ae. aegypti (Thompson et al., 1993). This
in Torpedo) involves a different amino acid to that mutation lies within the second transmembrane region
described by Weill et al. (2003), located close to the of the RDL subunit, presumed to be the principal con-
catalytic His within the acyl pocket of the enzyme. stituent of the ion-channel lining, and appears to con-
Replacements at this position have also been described fer both insensitivity to the insecticide and a decreased
in carbamate-resistant Myzus persicae (Nabeshima rate of desensitisation.
et al., 2003) and OP-resistant two-spotted spider mite The phenyl pyrazole insecticide fipronil also targets
Tetranychus urticae (Anazawa et al., 2003). the GABA receptor, and so problems of resistance due
Fig. 2 summarises the known AChE/carboxylester- to mutation may arise, should fipronil be employed in
ase genotype/phenotype relationships; for a compre- vector control programmes (Kolaczinski and Curtis,
hensive discussion of structure–function correlations, 2001). The appearance of such resistance in areas pre-
see Russell et al. (2004). viously treated with cyclodienes may be much quicker

Fig. 2. Genotype–phenotype correlations between carboxy/cholinesterase mutations and organophosphate/carbamate resistance. Torpedo AchE
numbering is used; the catalytic triad residues are denoted by solid circles. Solid lines indicate AchE mutations; diamonds denote ali-esterase
mutations. Based on data from Mutero et al. (1994), Walsh et al. (2001), Vontas et al. (2002), Weill et al. (2002, 2003), Nabeshima et al. (2003,
2004), Anazawa et al. (2003), Newcomb et al. (1997), Campbell et al. (1998), Claudianos et al. (1999, 2002), Zhu et al. (1999), Russell et al. (2004).
660 J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665

than anticipated, since the alanine!serine mutation type nerve insensitivity might not confer similar levels
can persist in Drosophila populations long after insecti- of resistance to all pyrethroids (Amin and Hemingway,
cide selection has been withdrawn, presumably due to 1989).
the lack of a selective disadvantage in resistant flies Cross-resistance studies and single channel studies in
(Hosie et al., 1997) the housefly confirmed that the molecular mechanism
of pyrethroid resistance is due to a functional change
3.3. Mutations in the voltage-gated sodium channel in the sodium channel (Chinn and Narahashi, 1986).
Using genetic mapping techniques, Ranson et al.
Pyrethroid insecticides have a rapid ‘knock-down’ (2000a,b) demonstrated that the sodium channel map-
effect. However, the intensive use of DDT and pyre- ped to a location in the genome corresponding to a
throids has led to the development of knock-down major quantitative trait locus determining resistance to
resistance (kdr) in many insect species (Liu et al., 2000; permethrin, and that a nucleotide substitution in the
Soderlund and Knipple, 2003) including Anopheles and sodium channel was linked to the inheritance of perme-
Aedes. The use of both DDT and pyrethroids in the thrin resistance in a Kenyan strain of An. gambiae. kdr-
control of rice and cotton pests is likely to have con- like mutations in this gene have been linked to changes
tributed significantly to the development of resistance in target sensitivity in a range of insects. A comprehen-
in An. gambiae from West Africa (Elissa et al., 1993; sive review of the molecular biology of knock-down
Martinez-Torres et al., 1998). The molecular mech- resistance to pyrethroids in other insect species has
anism of kdr can be distinguished from metabolic recently been published (Soderlund and Knipple, 2003).
resistance by the use of synergists (Enayati et al., 2003) kdr occurs due to a change in affinity between the
and is well characterised. insecticide and its binding site on the sodium channel,
Pyrethroids and DDT both target the voltage-gated caused by single or multiple substitutions in the sodium
sodium channel, which comprises four domains (I–IV), channel gene. It is possible that a limited number of
each consisting of six transmembrane helices (S1–S6). substitutions in the target site can lead to nerve insensi-
Action potentials are generated across the membrane tivity (Martinez-Torres et al., 1998). In several insect
in order to conduct electrical information throughout species, the most common kdr mutation is a leucine to
the nervous system. In an inactive state, when the phenylalanine (Leu!Phe) substitution in the S6 hydro-
membrane is at the resting potential, the sodium chan- phobic segment of domain II in the sodium channel
nel is closed. However, upon channel activation, the gene, such as that found in pyrethroid-resistant
membrane becomes depolarised. This causes the West African An. gambiae (Williamson et al., 1996;
sodium channel to open by generating a transient Martinez-Torres et al., 1998) and An. stephensi
sodium current. After approximately 1 ms, inactivation (Enayati et al., 2003). However, a second substitution
occurs due to a conformational change in the sodium at the same position (Leu!Ser) has been found in East
channel that blocks the passage of ions across the African An. gambiae (Ranson et al., 2000a,b). Both
membrane. When the membrane potential returns to mutations have been reported in An. sacharovi Favre in
the resting level, the channel closes (Vais et al., 2001). Turkey (Lüleyap et al., 2002).
Pyrethroids modify the gating kinetics of voltage- Cross-resistance and neurophysiological studies
sensitive sodium channels (Lund and Narahashi, 1983) inferred target site resistance in Ae. aegypti
by slowing both the activation and inactivation of the (Hemingway et al., 1989), and a recent molecular study
channel. The pyrethroid class of insecticides can be has now confirmed this (Brengues et al., 2003). Analy-
subdivided into two groups, each of which has a differ- sis of the cDNA sequence of the S6 hydrophobic
ent effect. Type I pyrethroids (e.g. permethrin) lack a domain II of six pyrethroid-resistant Aedes strains
cyano group. Type II pyrethroids (e.g. deltamethrin) showed that the common Leu!Phe substitution found
contain a cyano group in the a-benzylic position. Type in An. gambiae was not found in these strains of Aedes.
II pyrethroids prolong the sodium current during an This may be because Aedes codon usage does not easily
action potential more than Type I pyrethroids. The allow for this substitution without a simultaneous dou-
half activation potential (the membrane potential at ble nucleotide substitution. However, all insects from
which 50% of the available channels are open) is more a Vietnamese pyrethroid-resistant Aedes strain were
negative when using Type I pyrethroids (Vais et al., homozygous for a double nucleotide change, resulting
2001). This results in the channel opening at the resting in a Leu!Trp mutation, slightly upstream of the com-
potential, with consequent depolarisation of the neuro- monly mutated codon. This double mutation includes a
nal membrane and the initiation of repetitive dis- common silent polymorphism found in a range of
charges in motor and sensory axons that causes Aedes strains. All individuals from a Brazilian strain of
paralysis and death. All pyrethroids have a similar pyrethroid-resistant Aedes were either homozygous or
mode of action, but neurophysiological studies on C. p. heterozygous for both of two, non-adjacent, amino
quinquefasciatus from Saudi Arabia indicated that kdr- acid alterations (isoleucine!methionine and glyci-
J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665 661

ne!valine), each of which occur as a result of a single References


nucleotide mutation. These alterations were also found
in mosquitoes from two other Central American coun- Agianian, B., Tucker, P.A., Schouten, A., Leonard, K., Bullard, B.,
Gros, P., 2003. Structure of a Drosophila sigma class glutathione
tries, suggesting geographical dispersal from a single S-transferase reveals a novel active site topography suited for lipid
originating location. Aedes samples from Indonesia and peroxidation products. J. Mol. Biol. 326, 151–165.
Thailand had a mutation of valine !glycine as a result Aldridge, W.N., 1953. Serum esterases. 1. Two types of esterases (A
and B) hydrolysing p-nitrophenyl acetate, propionate and
of a single nucleotide change.
butyrate, and a method for their determination. Biochem. J. 53,
In total, over 20 unique sodium channel amino acid 110–117.
sequence polymorphisms have been identified in associ- Amichot, M., Brun, A., Cuany, A., Helvig, C., Salaun, J.P., Durst,
ation with pyrethroid resistance. All mutations so far F., et al., 1994. Expression study of CYP genes in Drosophila
strains resistant or susceptible to insecticides. In: Lechner, M.C.
identified in mosquito vectors have been found in
(Ed.), Cytochrome p450. In: 8th International Conference. Euro-
domain II of the sodium channel. Super-kdr mutations text-John Libbey, Paris, pp. 689–692.
found in other insects (Soderland and Knipple, 2003) Amin, A.M., Hemingway, J., 1989. Preliminary investigation of the
that are highly resistant to pyrethroids have not yet mechanisms of DDT and pyrethroid resistance in Culex quinque-
fasciatus Say (Diptera: Culicidae) from Saudi Arabia. Bull. Ent.
been identified in mosquitoes. Res. 79, 361–366.
Direct DNA sequence analysis can be used to estab- Anazawa, Y., Tomita, T., Aiki, T., Kozaki, T., Kono, Y., 2003.
lish the presence of mutations in the voltage-gated Sequence of a cDNA encoding acetylcholinesterase from suscep-
sodium channel. However, this is not practical for tible and resistant two-spotted spider mite, Tetranychus urticae.
Insect Biochem. Mol. Biol. 33, 509–514.
large-scale population studies. A simple diagnostic Andersen, J.F., Utermohlen, J.G., Feyereisen, R., 1994. Expression
PCR assay has been developed (Martinez-Torres et al., of house fly CYPA1 and NADPH-cytochrome P450 reductase in
1998) that enables the rapid diagnosis of the common Escherichia coli and reconstitution of an insecticide-metabolising
Leu!Phe substitution in resistant mosquitoes, and spe- P450 system. Biochemistry 33, 2171–2177.
Apperson, C.S., Georghiou, G.P., 1975. Mechanisms of resistance to
cific assays for other mutations have been developed organophosphorus insecticides in Culex tarsalis. J. Econ. Bio-
(Ranson et al., 2000a,b; Enayati et al., 2003). The chem. 68, 63–78.
sodium channel gene has been used as a molecular tool Armstrong, R.N., 1991. Glutathione S-transferases: reaction mech-
anism, structure, and function. Chem. Res. Toxicol. 4, 131–140.
to look at An. gambiae population structure in Mali
Ayad, H., Georghiou, G.P., 1979. Resistance pattern of Anopheles
and Burkina Faso (Fanello et al., 2003a,b; Diabate albimanus Wied. following selection by parathion. Mosquito
et al., 2003). News 39, 121–125.
Berge, J.B., Feyereisen, R., Amichot, M., 1998. Cytochrome p450
monooxygenases and insecticide resistance in insects. Philos.
Trans. Roy. Soc. Lond. B Biol. Sci. 353, 1701–1705.
4. Conclusions Bisset, J., Rodriguez, M.M., Diaz, C., Ortiz, E., Marquetti, M.C.,
Hemingway, J., 1990. The mechanisms of organophosphate and
Resistance to insecticides is one of the most well carbamate resistance in Culex quinquefasciatus (Diptera: Culici-
documented, and rapid, cases of evolutionary adap- dae) from Cuba. Bull. Ent. Res. 80, 245–250.
Board, P.G., Baker, R.T., Chelvanayagam, G., Jermiin, L.S., 1997.
tation to environmental changes. Although only a lim- Zeta, a novel class of glutathione transferases in a range of species
ited number of resistance mechanisms have been from plants to humans. Biochem. J. 328, 929–935.
implicated to date, the diversity within those enzyme Board, P.G., Coggan, M., Chelvanayagam, G., Easteal, S., Jermiin,
families involved in metabolic resistance is likely to L.S., Schulte, G.K., et al., 2000. Identification, characterization,
and crystal structure of the Omega class glutathione transferases.
contribute substantially to resistance to many insecti- J. Biol. Chem. 275, 24798–24806.
cide classes. The advent of post-genomic technology is Bourguet, D., Raymond, M., Fournier, D., Malcolm, C.A., Toutant,
likely to reveal the contributions of hitherto unsuspec- J.P., Arpagaus, M., 1996. Existence of two acetylcholinesterases
ted enzymes. The ability to compare the transcriptome in the mosquito Culex pipiens (Diptera: Culicidae). J. Neurochem.
67, 2115–2123.
or proteome of a resistant mosquito strain with a sus- Brandt, A., Scharf, M., Pedra, J.H.F., Holmes, G., Dean, A.,
ceptible equivalent will substantially enhance our Kreitman, M., et al., 2002. Differential expression and induction
understanding of resistance and its complexity. In the of two Drosophila cytochrome p450 genes near the Rst(2)DDT
locus. Insect Mol. Biol. 11, 337–341.
long term this can only benefit those who suffer from
Brengues, C., Hawkes, N.J., Chandre, F., McCarroll, L., Duchon, S.,
mosquito-borne disease. Guillet, P., et al., 2003. Pyrethroid and DDT cross-resistance in
Aedes aegypti is correlated with novel mutations in the voltage-
gated sodium channel gene. Med. Vet. Ent. 17, 87–94.
Brogdon, W.G., Barber, A.M., 1990. Fenitrothion–deltamethrin
Acknowledgements cross-resistance conferred by esterases in Guatemalan Anopheles
albimanus. Pestic. Biochem. Physiol. 3, 130–139.
We would like to acknowledge anonymous reviewers Brogdon, W.G., McAllister, J.C., Corwin, A.M., Cordon-Rosales, C.,
for their constructive comments. 1999a. Oxidase-based DDT–pyrethroid cross-resistance in
662 J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665

Guatemalan Anopheles albimanus. Pestic. Biochem. Physiol. 64, Ding, Y., Ortelli, F., Rossiter, L., Hemingway, J., Ranson, H., 2003.
101–111. The Anopheles gambiae glutathione transferase superfamily:
Brogdon, W.G., McAllister, J.C., Corwin, A.M., Cordon-Rosales, C., annotation, phylogeny and expression profiles. BMC Genomics 4,
1999b. Independent selection of multiple mechanisms for pyre- 35–45.
throid resistance in Guatemalan Anopheles albimanus (Diptera: Dombrowski, Krishnan, R., Witte, M., Maitra, S., Diesing, C.,
Culicidae). J. Econ. Entomol. 92, 298–302. Waters, L.C., Ganguly, R., 1998. Constitutive and barbital-
Burke, T.W., Kadonaga, J.T., 1996. Drosophila TFIID binds induced expression of the Cyp6a2 allele of a high producer strain
to a conserved downstream basal promoter element that is of CYP6A2 in the genetic background of a low producer strain.
present in many TATA-box deficient promoters. Genes Dev. 10, Gene 221, 69–77.
711–724. Dunkov, B.C., Mocelin, G., Shotkoski, F., ffrench-Constant, R.H.,
Campbell, P.M., Newcomb, R.D., Russell, R.J., Oakeshott, J.G., Feyereisen, R., 1997. The Drosophila cytochrome p450 gene
1998. Two different amino acid substitutions in the ali-esterase, Cyp6a2: structure, chromosomal localisation, heterologous
E3, confer alternative types of organophosphorus insecticide expression and induction by phenobarbital. DNA Cell Biol. 16,
resistance in the sheep blowfly Lucilia cuprina. Insect Biochem. 1345–1356.
Mol. Biol. 28, 139–150. Elissa, N., Mouchet, J., Riviere, F., Meunier, J.Y., Yao, K., 1993.
Carino, F., Koener, J.F., Plapp, Jr., ., F.W., Feyereisen, R., 1992. Resistance of Anopheles gambiae s.s. to pyrethroids in Cote
Expression of the cytochrome p450 gene CYP6A1 in the housefly, d’Ivoire. Ann. Soc. Belge. Med. Trop. 73, 291–294.
Musca domestica. ACS Symp. Ser. 505, 31–40. Enayati, A.A., Vatandoost, H., Ladonni, H., Townson, H.,
Carino, F., Koener, J.F., Plapp, Jr., ., F.W., Feyereisen, R., 1994. Hemingway, J., 2003. Molecular evidence for a kdr-like pyre-
Constitutive overexpression of the cytochrome p450 gene throid resistance mechanism in the malaria vector mosquito
CYP6A1 in a house fly strain with metabolic resistance to insecti- Anopheles stephensi. Med. Vet. Ent. 17, 138–144.
cides. Insect Biochem. Mol. Biol. 24, 411–418. Fanello, C., Carneiro, I., Ilboudo-Snaogo, E., Cuzin-Ouattara, N.,
Chandre, F., Darriet, F., Manga, L., Akogbeto, M., Faye, O., Badolo, A., Curtis, C.F., 2003a. Comparative evaluation of car-
Moucher, J., et al., 1999. Status of pyrethroid resistance in Anoph- bosulfan- and permethrin- impregnated curtains for preventing
eles gambiae sensu lato. Bull. WHO 77, 230–234. house-entry by the malaria vector Anopheles gambiae in Burkina
Chiang, F., Sun, C., 1993. Glutathione transferase isozymes of dia- Faso. Med. Vet. Ent. 17, 333–338.
mondback moth larvae and their role in the degradation of some Fanello, C., Petrarca, V., della Torre, A., Santolamazza, F., Dolos,
organophosphorus insecticides. Pestic. Biochem. Physiol. 45, 7–14. G., Coulibaly, M., et al., 2003b. The pyrethroid knock-down
Chinn, K., Narahashi, T., 1986. Stabilisation of sodium channels as resistance gene in the Anopheles gambiae complex in Mali and fur-
the basis for drug action: past, present and future. J. Physiol. 380, ther indication of incipient speciation within An. gambiae s.s.
191–207. Insect Mol. Biol. 12, 241–245.
Clark, A.G., 1989. The comparative enzymology of the glutathione Feyereisen, R., 1999. Insect P450 enzymes. Ann. Rev. Entomol. 44,
S-transferases from non-vertebrate organisms. Comp. Biochem. 507–533.
Physiol B 92, 419–446. ffrench-Constant, R.H., Pittendrigh, B., Vaughan, A., Anthony, N.,
Clark, A.G., Shamaan, N.A., 1984. Evidence that DDT dehydro- 1998. Why are there so few resistance-associated mutations in
chlorinase from the housefly is a glutathione S-transferase. Pestic. insecticide target genes? Philos. Trans. Roy. Soc. Lond. B 353,
Biochem. Physiol. 22, 249–261. 1685–1693.
Claudianos, C., Russell, R.J., Oakeshott, J.G., 1999. The same amino Fournier, D., Karch, F., Bride, J.M., Hall, L.M.C., Berge, J.-B.,
acid substitution in orthologous esterases confers organopho- Spierer, P., 1989. Drosophila melanogaster acetylcholinesterase
sphate resistance on the house fly and a blowfly. Insect Biochem. gene, structure evolution and mutations. J. Mol. Evol. 210,
Mol. Biol. 29, 675–686. 15–22.
Claudianos, C., Crone, E., Coppin, C., Russell, R., Oakeshott, J., Grant, D.F., Hammock, B.D., 1992. Genetic and molecular evidence
2002. A genomics perspective on mutant aliesterases and metabolic for a trans-acting regulatory locus controlling glutathione S-trans-
resistance to organophosphates. In: Marshall Clark, J., Yama- ferase-2 expression in Aedes aegypti. Mol. Gen. Genet. 234,
guchi, I. (Eds.), Agrochemical Resistance: Extent, Mechanism and 169–176.
Detection. American Chemical Society, Washington, DC. Grant, D.F., Dietze, E.C., Hammock, B.D., 1991. Glutathione S-
Coleman, M.N., Vontas, J.G., Hemingway, J., 2002. Molecular char- transferase isozymes in Aedes aegypti: purification, characteriza-
acterization of the amplified aldehyde oxidase from insecticide tion, and isozyme-specific regulation. Insect Biochem. 21,
resistant Culex quinquefasciatus. Eur. J. Biochem. 269, 768–779. 421–433.
Corbett, J.R., 1974. The Biochemical Mode of Action of Pesticides. Guillemaud, T., Makate, N., Raymond, M., Hirst, B., Callaghan, A.,
Academic Press, London, UK, pp. 102–130. 1997. Esterase gene amplification in Culex pipiens. Insect Mol.
Curtis, C.F., Miller, J.E., Hassan, M.H., Kolaczinski, J.H., Biol. 6, 319–327.
Kasumba, I., 1998. Can anything be done to maintain the effec- Hawkes, N.J., Hemingway, J., 2002. Analysis of the promoters of the
tiveness of pyrethroid impregnated bednets against malaria vec- b-esterase genes associated with insecticide resistance in the filariasis
tors? Philos. Trans. Roy. Soc. B 353, 1769–1775. vector Culex quinquefasciatus. Biochim. Biophys. Acta 1574, 51–62.
Daborn, P.J., Yen, J.L., Bogwitz, M.R., Le Goff, G., Feil, E., Jeffer, Hayes, J.D., Wolf, C.R., 1988. Role of glutathione transferase in
S., et al., 2002. A single P450 allele associated with insecticide drug resistance. In: Sies, H., Ketterer, B. (Eds.), Glutathione con-
resistance in Drosophila. Science 297, 2253–2256. jugation. Academic Press, pp. 315.
Devonshire, A.L., Moores, G.D., 1982. A carboxylesterase with Hemingway, J., 1982a. Genetics of organophosphate and carbamate
broad substrate specificity causes organophosphorus, carbamate resistance in Anopheles atroparvus (Diptera: Culicidae). J. Econ.
and pyrethroid resistance in peach potato aphids (Myzus persi- Entomol. 75, 1055–1058.
cae). Pestic. Biochem. Physiol. 18, 235–246. Hemingway, J., 1982b. The biochemical nature of malathion resist-
Diabate, A., Baldet, T., Chandre, F., Dabire, K.R., Kengne, P., ance in Anopheles stephensi from Pakistan. Pestic. Biochem. Phy-
Guigemaude, T.R., et al., 2003. Kdr mutation, a genetic marker siol. 17, 149–155.
to assess events of introgression between the molecular M & S Hemingway, J., 1983. Biochemical studies on malathion resistance in
forms of Anopheles gambiae in the tropical savannah area of West Anopheles arabiensis from Sudan. Trans. R. Soc. Trop. Med. Hyg.
Africa. J. Med. Ent. 40, 195–198. 77, 477–480.
J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665 663

Hemingway, J., Georghiou, G.P., 1983. Studies on the acet- Karunaratne, S.H.P.P., Jayawardena, K.G.I., Hemingway, J.,
ylcholinesterase of Anopheles albimanus resistant and susceptible Ketterman, A.J., 1993. Characterisation of a B-type esterase
to organophosphate and carbamate insecticides. Pestic. Biochem. involved in insecticide resistance from the mosquito Culex
Physiol. 19, 167–171. quinquefasciatus. Biochem. J. 294, 575–579.
Hemingway, J., Karunaratne, S.H.P.P., 1998. Mosquito carbox- Kasai, S., Scott, J.G., 2000. Over expression of cytochrome P450
ylesterases: a review of the molecular biology and biochemistry of CYP6D1 is associated with monooxygenase-mediated pyrethroid
a major insecticide resistance mechanism. Med. Vet. Entomol. 12, resistance in houseflies from Georgia. Pestic. Biochem. Physiol.
1–12. 68, 34–41.
Hemingway, J., Malcolm, C.A., Kissoon, K.E., Boddington, R.G., Kasai, S., Weerashinghe, I.S., Shono, T., Yamakawa, M., 2000. Mol-
Curtis, C.F., Hill, N., 1985. The biochemistry of insecticide resist- ecular cloning, nucleotide sequence and gene expression of a cyto-
ance in Anopheles sacharovi: comparative studies with a range of chrome P450 (CYP6F1) from the pyrethroid-resistant mosquito,
insecticide susceptible and resistant Anopheles and Culex species. Culex quinquefasciatus Say. Insect Biochem. Mol. Biol. 30,
Pest. Biochem. Physiol. 24, 68–76. 163–171.
Hemingway, J., Smith, C., Jayawardena, K.G.I., Herath, P.R.J., Kolaczinski, J., Curtis, C., 2001. Laboratory evaluation of fipronil, a
1986. Field and laboratory detection of the altered acet- phenylpyrazole insecticide, against adult Anopheles (Diptera: Culi-
ylcholinesterase resistance genes which confer organophosphate cidae) and investigation of its possible cross-resistance with diel-
and carbamate resistance in mosquitoes (Diptera: Culicidae). Bull. drin in Anopheles stephensi. Pest Manag. Sci. 57, 41–45.
Ent. Res. 76, 559–565. Kostaropoulos, I., Papadopoulos, A.I., Metaxakis, A., Boukouvala,
Hemingway, J., Boddington, R.G., Harris, J., Dunbar, S.J., 1989. E., Papadopoulou-Mourkidou, E., 2001. Glutathione S-transfer-
Mechanisms of insecticide resistance in Aedes aegypti from ase in the defence against pyrethroids in insects. Insect Biochem.
Puerto Rico. Bull. Ent. Res. 79, 123–130. Mol. Biol. 31, 313–319.
Hemingway, J., Miyamoto, J., Herath, P.R.J., 1991. A possible novel Le Goff, G., Boundy, S., Daborn, P.J., Yen, J.L., Sofer, L., Lind, R.,
link between organophosphorus and DDT insecticide resistance et al., 2003. Microarray analysis of cytochrome p450 mediated
genes in Anopheles: supporting evidence from fenitrothion metab- insecticide resistance in Drosophila. Insect Biochem. Mol. Biol. 33,
olism studies. Pestic. Biochem. Physiol. 39, 49–56. 201–708.
Hemingway, J., Hawkes, N.J., Prapanthadara, L., Jayawardena, Lines, J.D., Myamba, J., Curtis, C.F., 1987. Experimental hut trials
K.G.I., Ranson, H., 1998. The role of gene splicing, gene amplifi- of permethrin-impregnated mosquito nets and curtains against
cation and regulation in mosquito insecticide resistance. Philos. malaria vectors in Tanzania. Med. Vet. Entomol. 1, 37–51.
Trans. Roy. Soc. Lond. B 353, 1695–1699. Liu, N., Scott, J., 1997. Inheritance of CYP6D1-mediated pyrethroid
Hemingway, J., Coleman, M.N., Paton, M.G., McCarroll, L., resistance in house fly, Musca domestica (L.). Insect Biochem.
Vaughan, A., DeSilva, D., 2000. Aldehyde oxidase is coamplified Mol. Biol. 32, 1–8.
with the world’s most common Culex mosquito insecticide resist- Liu, Z., Valles, S.M., Dong, K., 2000. Novel point mutations in the
ance-associated esterases. Insect Mol. Biol. 9, 93–99. German cockroach para sodium channel gene are associated with
Herath, P.R.J., Hemingway, J., Weerasinghe, I.S., Jayawardena, knockdown resistance (kdr) to pyrethroid insecticides. Insect Bio-
K.G.I., 1987. The detection and characterisation of malathion chem. Mol. Biol. 30, 991–997.
resistance in field populations of Anopheles culicifacies B in Sri Lüleyap, H.U., Alptekin, D., Kasap, H., Kasap, M., 2002. Detection
Lanka. Pestic. Biochem. Physiol. 29, 157–162. of knockdown resistance mutations in Anopheles sacharovi (Dip-
Hernandez, R., Guerrero, F.D., George, J.E., Wagner, G.G., tera: Culicidae) and genetic distance with Anopheles gambiae
2002. Allele frequency and gene expression of a putative car- (Diptera: Culicidae) using cDNA sequencing of the voltage-gated
boxylesterase-encoding gene in a pyrethroid resistant strain of sodium channel gene. J. Med. Ent. 39, 870–874.
the tick Boophilus microplus. Insect Biochem. Mol. Biol. 32, Lund, A.E., Narahashi, T., 1983. Kinetics of sodium channel modifi-
1009–1016. cation as the basis for the variation in the nerve membrane effects
Holt, R.A., Subramanian, G.M., Halpern, A., Sutton, G.G., of pyrethroid and DDT analogs. Pestic. Biochem. Physiol. 20,
Charlab, R., Nusskern, D.R., et al., 2002. The genome 203–216.
sequence of the malaria mosquito Anopheles gambiae. Science Maitra, S., Dombrowski, S.M., Waters, L.C., Ganguly, R., 1996.
298, 129–149. Three second chromosome-linked clustered Cyp6 genes show
Hosie, A.M., Aronstein, K., Sattelle, D.B., ffrench-Constant, R.H., differential constitutive and barbital-induced expression in
1997. Molecular biology of insect neuronal GABA receptors. DDT-resistant and susceptible strains of Drosophila melanogaster.
Trends Neurosci. 20, 578–583. Gene 180, 165–171.
Hougard, J.M., Duchon, S., Zaim, M., Guillet, P., 2002. Bifenthrin: a Maitra, S., Dombrowski, S.M., Basu, M., Raustol, O., Waters, L.C.,
useful pyrethroid insecticide for treatment of mosquito nets. J. Ganguly, R., 2000. Factors on the third chromosome affect the
Med. Entomol. 39, 526–533. level of Cyp6a2 and Cyp6a8 expression in Drosophila melanoga-
Huang, H.S., Hu, N.T., Yao, Y.E., Wu, C.Y., Chiang, S.W., Sun, ster. Gene 248, 147–156.
C.N., 1998. Molecular cloning and heterologous expression of a Maitra, S., Price, C., Ganguly, R., 2002. Cyp6a8 of Drosophila mela-
glutathione S-transferase involved in insecticide resistance from nogaster: gene structure, and sequence and functional analysis of
the diamondback moth Plutella xylostella. Insect Biochem. Mol. the upstream DNA. Insect Biochem. Mol. Biol. 32, 859–870.
Biol. 28, 651–658. Malcolm, C.A., Bourguet, D., Acolillo, A., Rooker, S.J., Garvey,
Karunaratne, S.H.P.P., 1994. Characterisation of multiple variants of C.F., Hall, L.M., et al., 1998. A sex-linked Ace gene, not linked
carboxylesterases which are involved in insecticide resistance in to insensitive acetylcholinesterase-mediated resistance. Insect Mol.
the mosquito Culex quinquefasciatus. PhD Thesis, University of Biol. 7, 107–120.
London. Mannervik, B., 1985. The isoenzymes of glutathione transferase. Adv.
Karunaratne, S.H.P.P., Hemingway, J., Jayawardena, K.G.I., Dassa- Enzymol. Relat. Areas Mol. Biol. 57, 357–417.
nayaka, V., Vaughan, A., 1995. Kinetic and molecular differences Martinez-Torres, D., Chandre, F., Williamson, M.S., Darriet, F.,
in the amplified and non-amplified esterases from insecticide Berge, J.B., Devonshire, A.L., et al., 1998. Molecular characteris-
resistant and susceptible Culex quinquefasciatus mosquitoes. J. ation of pyrethroid knock-down resistance (kdr) in the major
Biol. Chem. 270, 31124–31128. malaria vector Anopheles gambiae s.s. Insect Mol. Biol. 7, 179–184.
664 J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665

Mathenge, E.M., Gimnig, J.E., Kolczak, M., Ombok, M., Irungu, from the malaria vector Anopheles gambiae. Biochem. J. 373,
L.W., Hawley, W.A., 2001. Effect of permethrin-impregnated nets 957–963.
on exiting behaviour, blood feeding success, and time of feeding Ottea, J.A., Plapp, F.W., 1984. Glutathione S-transferases in the house
of malaria mosquitoes (Diptera: Culicidae) in Western Kenya. J. fly: biochemical and genetic changes associated with induction and
Med. Entomol. 38, 531–536. insecticide resistance. Pestic. Biochem. Physiol. 22, 203–208.
Mbogo, C.N.M., Baya, N.M., Ofulla, A.V.O., Githure, J.I., Snow, Paton, M.G., Karunaratne, S.H.P.P., Giakoumaki, E., Roberts, N.,
R.W., 1996. The impact of permethrin-impregnated bednets on Hemingway, J., 2000. Quantitative analysis of gene amplification
malaria vectors of the Kenyan coast. Med. Vet. Entomol. 10, in insecticide resistant Culex mosquitoes. Biochem. J. 346, 17–24.
251–259. Penilla, R.P., Rodriguez, A.D., Hemingway, J., Torres, J.L., Arre-
McCarroll, L., Paton, M.G., Karunaratne, S.H.P.P., Jayasuriya, dondo-Jimenez, J.I., Rodriguez, M.H., 1998. Resistance manage-
H.T.R., Kalpage, K.S.P., Hemingway, J., 2000. Insecticides and ment strategies in malaria vector mosquito control. Baseline data
mosquito-borne disease. Nature 407, 961–962. for a large-scale field trial against Anopheles albimanus in Mexico.
Misra, S., Crosby, M.A., Mungall, C.J., Matthews, B.B., Campbell, Med. Vet. Entomol. 12, 217–233.
K.S., Hradecky, P., et al., 2002. Annotation of the Drosophila Pittendrigh, B., Aronstein, K., Zinkovsky, E., Andreev, O.,
melanogaster euchromatic genome: a systematic review. Genome Campbell, B., Daly, J., et al., 1997. Cytochrome p450 genes from
Biol. 3, 83. Helicoverpa armigera: expression in a pyrethroid-susceptible and -
Mori, A., Tomita, T., Hidoh, O., Kono, Y., Severson, D.W., 2001. resistant strain. Insect Biochem. Mol. Biol. 27, 507–512.
Comparative linkage map development and identification of an Prapanthadara, L., Hemingway, J., Ketterman, A.J., 1993. Partial
autosomal locus for insensitive acetylcholinesterase-mediated purification and characterization of glutathione S-transferases
insecticide resistance in Culex tritaeniorhynchus. Insect Mol. Biol. involved in DDT resistance from the mosquito Anopheles gam-
10, 197–203. biae. Pest. Biochem. Physiol. 47, 119–133.
Mouches, C., Pasteur, N., Berge, J.B., Hyrien, O., Raymond, M., De Pridgeon, J.W., Zhang, L., Liu, N., 2003. Overexpression
Saint Vincent, B.R., et al., 1986. Amplification of an esterase gene of CYP4G19 associated with a pyrethroid-resistant strain of the
is responsible for insecticide resistance in a Californian Culex German cockroach, Blattella germanica (L.). Gene 314, 157–163.
mosquito. Science 233, 778–780. Ranson, H., Hemingway, J., 2004. Insect pharmacology and control:
Mouches, C., Pauplin, Y., Agarwal, M., Lemieux, L., Herzog, M., glutathione S-transferases. In: Gilbert, L.I., Iatrou, K., Gill, S.
Abadon, M., et al., 1990. Characterization of amplification core (Eds.). Elsevier Ltd, Oxford, UK.
and esterase B1 gene responsible for insecticide resistance in Ranson, H., Collins, F.H., Hemingway, J., 1998. The role of alterna-
Culex. Proc. Natl. Acad. Sci. (USA) 87, 2574–2578. tive mRNA splicing in generating heterogeneity within the Anoph-
Mourya, D.T., Hemingway, J., Leake, C.J., 1993. Changes in enzyme eles gambiae class I glutathione S-transferase family. Proc. Natl.
titres with age in four geographical strains of Aedes aegypti and their Acad. Sci. USA 95, 14284–14289.
association with insecticide resistance. Med. Vet. Ent. 7, 11–16. Ranson, H., Jensen, B., Vulule, J.M., Wang, X., Hemingway, J.,
Mutero, A., Pralavorio, M., Bride, J.M., Fournier, D., 1994. Resist- Collins, F.H., 2000a. Identification of a point mutation in the
ance associated point mutations in insecticide-insensitive acet- voltage-gated sodium channel gene of Kenyan Anopheles gambiae
ylcholinesterase. Proc. Natl. Acad. Sci. USA 91, 5922–5926. associated with resistance to DDT and pyrethroids. Insect Mol.
Nabeshima, T., Kozaki, T., Tomita, T., Kono, Y., 2003. An amino Biol. 9, 491–497.
acid substitution on the second acetylcholinesterase in the Ranson, H., Jensen, B., Wang, X., Prapanthadara, L., Hemingway,
pirimicarb-resistant strains of the peach-potato aphid, Myzus J., Collins, F.H., 2000b. Genetic mapping of two loci affecting
persicae. Biochem. Biophys. Res. Commun. 307, 15–22. DDT resistance in the malaria vector Anopheles gambiae. Insect
Nabeshima, T., Mori, A., Kozaki, T., Iwata, Y., Hidoh, O., Harada, Mol. Biol. 9, 499–507.
S., et al., 2004. An amino acid substitution attributable to insecti- Ranson, H., Rossiter, L., Ortelli, F., Jensen, B., Wang, X., Roth,
cide-insensitivity of acetylcholinesterase in a Japanese encephalitis C.W., et al., 2001. Identification of a novel class of insect glu-
vector mosquito, Culex tritaeniorhynchus. Biochem. Biophys. Res. tathione S-transferases involved in resistance to DDT in the
Commun. 313, 794–801. malaria vector Anopheles gambiae. Biochem. J. 359, 295–304.
Nance, E., Heyse, D., Britton-Davian, J., Pasteur, N., 1990. Chromo- Ranson, H., Claudianos, C., Ortelli, F., Abgrall, C., Hemingway, J.,
somal organisation of the amplified esterase B1 gene in organo- Sharakhova, M.V., et al., 2002a. Evolution of supergene families
phosphate-resistant Culex pipiens quinquefasciatus Say (Diptera: associated with insecticide resistance. Science 298, 179–181.
Culicidae). Genome 33, 148–152. Ranson, H., Nikou, D., Hutchinson, M., Wang, X., Roth, C.W.,
Newcomb, R.D., Campbell, P.M., Ollis, D.L., Cheah, E., Russell, Hemingway, J., et al., 2002b. Molecular analysis of multiple cyto-
R.J., Oakeshott, J.D., 1997. A single amino acid substitution con- chrome P450 genes from the malaria vector Anopheles gambiae.
verts a carboxylesterase to an organophosphorus hydrolase and Insect Mol. Biol. 11, 409–418.
confers insecticide resistance on a blowfly. Proc. Natl. Acad. Sci. Ranson, H., Paton, M.G., Jensen, B., McCarroll, L., Hemingway, J.,
USA 94, 7464–7468. Collins, F.H., 2004. Genetic mapping of genes conferring perme-
N’Guessan, R., Darriet, F., Guillet, P., Carnevale, P., Traore- thrin resistance in the malaria vector Anopheles gambiae. Insect
Lamizana, M., Corbel, V., et al., 2003. Resistance to carbosulfan Mol. Biol. Submitted for publication.
in Anopheles gambiae from Ivory Coast, based on reduced sensi- Rooker, S., Guillemaud, T., Berge, J., Pasteur, N., Raymond, M.,
tivity of acetylcholinesterase. Med. Vet. Ent. 17, 19–25. 1996. Coamplification of esterase A and B genes as a single unit
Nikou, D., Ranson, H., Hemingway, J., 2003. An adult-specific in Culex pipiens mosquitoes. Heredity 77, 555–561.
CYP6 P450 gene is over expressed in a pyrethroid-resistant strain Russell, R.J., Claudianos, C., Campbell, P.M., Horne, I., Sutherland,
of the malaria vector, Anopheles gambiae. Gene 318, 91–102. T.D., Oakeshott, J.G., 2004. Two major classes of target site
Oppenoorth, F.J., Van der Pas, L.J.T., Houx, N.W.H., 1979. insensitivity mutations confer resistance to organophosphate and
Glutathione S-transferases and hydrolytic activity in a tetra- carbamate insecticides. Pestic. Biochem. Physiol., in press.
chlorovinphos-resistant strain of housefly and their influence on Sabourault, C., Guzov, V.M., Koener, J.F., Claudianos, C., Plapp,
resistance. Pestic. Biochem. Physiol. 11, 176–188. Jr., ., F.W., Feyereisen, R., 2001. Overproduction of a P450 that
Ortelli, F., Rossiter, L.C., Vontas, J., Ranson, H., Hemingway, J., metabolizes diazinon is linked to a loss-of-function in the chro-
2003. Heterologous expression of four glutathione transferase mosome 2 ali-esterase (MdalphaE7) gene in resistant house flies.
genes genetically linked to a major insecticide resistance locus, Insect Mol. Biol. 10, 609–618.
J. Hemingway et al. / Insect Biochemistry and Molecular Biology 34 (2004) 653–665 665

Sawicki, R., Singh, S.P., Mondal, A.K., Benes, H., Zimniak, P., Villani, F., Hemingway, J., 1987. The detection and interaction of
2003. Cloning, expression and biochemical characterization of one multiple organophosphorus and carbamate insecticide resistance
Epsilon-class (GST-3) and ten Delta-class (GST-1) glutathione S- genes in field populations of Culex pipiens from Italy. Pestic. Bio-
transferases from Drosophila melanogaster, and identification of chem. Physiol. 27, 218–228.
additional nine members of the Epsilon class. Biochem. J. 370, Vontas, J.G., Small, G.J., Hemingway, J., 2001. Glutathione S-trans-
661–669. ferases as antioxidant defence agents confer pyrethroid resistance
Scott, J.G., Wen, Z., 2001. Cytochromes p450 of insects: the tip of in Nilaparvata lugens. Biochem. J. 357, 65–72.
the iceberg. Pest. Manag. Sci. 57, 958–967. Vontas, J.G., Hejazi, M.J., Hawkes, N.J., Cosmidis, N., Loukas, M.,
Sheehan, D., Meade, G., Foley, V.M., Dowd, C.A., 2001. Structure, Hemingway, J., 2002. Resistance-associated point mutations of
function and evolution of glutathione transferases: implications organophosphate insensitive acetylcholinesterase, in the olive fruit
for classification of non-mammalian members of an ancient fly Bactrocera oleae. Insect Mol. Biol. 11, 329–336.
enzyme superfamily. Biochem. J. 360, 1–16. Vulule, J.M., Beach, R.F., Atieli, F.K., McAllister, J.C., Brogdon,
Shen, B., Dong, H.-Q., Tian, H.-S., Ma, L., Li, X.-L., Wu, G.-L., W.G., Roberts, J.M., et al., 1999. Elevated oxidase and esterase
et al., 2003. Cytochrome p450 genes expressed in the deltame- levels associated with permethrin tolerance in Anopheles gambiae
thrin-susceptible and -resistant strains of Culex pipiens pallens. from Kenyan villages using permethrin-impregnated nets. Med.
Pestic. Biochem. Physiol. 75, 19–26. Vet. Entomol. 13, 239–244.
Sierra-Santoya, A., Hernandez, M., Albores, A., Cebrian, M.E., Walsh, S.B., Dolden, T.A., Moores, G.D., Kristensen, M., Lewis, T.,
2000. Sex-dependent regulation of hepatic cytochrome p450 by Devonshire, A.L., et al., 2001. Identification and characterisation
DDT. Toxicol. Sci. 54, 81–87. of mutations in housefly (Musca domestica) acetylcholinesterase
Singh, S.P., Coronella, J.A., Benes, H., Cochrane, B.J., Zimniak, P., involved in insecticide resistance. Biochem. J. 359, 175–181.
2001. Catalytic function of Drosophila melanogaster glutathione S- Wang, J.Y., McCommas, S., Syvanen, M., 1991. Molecular cloning
transferase DmGSTS1-1 (GST-2) in conjugation of lipid peroxi- of a glutathione S-transferase overproduced in an insecticide-
dation end products. Eur. J. Biochem. 268, 2912–2923. resistant strain of the housefly (Musca domestica). Mol. Gen.
Soderlund, D.M., Knipple, D.C., 2003. The molecular biology of Genet. 227, 260–266.
knockdown resistance to pyrethroid insecticides. Insect. Biochem. Wei, S.H., Clark, A.G., Syvanen, M., 2001. Identification and cloning
Mol. Biol. 33, 563–577. of a key insecticide-metabolizing glutathione S-transferase
Stone, B.F., Brown, A.W.A., 1969. Mechanisms of resistance to fen- (MdGST-6A) from a hyper insecticide-resistant strain of the house-
thion in Culex pipiens fatigans Wied. Bull. WHO 40, 401–408. fly Musca domestica. Insect Biochem. Mol. Biol. 31, 1145–1153.
Tang, A.H., Tu, C.P., 1994. Biochemical characterization of Droso- Weill, M., Fort, P., Berthomieu, A., Dubois, M.P., Pasteur, N.,
phila glutathione S-transferases D1 and D21. J. Biol. Chem. 269, Raymond, M., 2002. A novel acetylcholinesterase gene in mos-
27876–27884. quitoes codes for the insecticide target and is non-homologous to
Thompson, M., Steichen, J.C., ffrench-constant, R.H., 1993. Cloning the ace gene in Drosophila. Proc. R. Soc. Lond. B 269, 2007–2016.
and sequencing of the cyclodiene insecticide resistance gene from Weill, M., Lutfalla, G., Mogensen, K., Chandre, F., Berthomieu, A.,
the yellow-fever mosquito Aedes aegypti—conservation of the Berticat, C., et al., 2003. Insecticide resistance in mosquito vec-
gene and resistance-associated mutation with Drosophila. FEBS tors. Nature 423, 136–137.
Lett. 325, 187–190. Wilce, M.C., Parker, M.W., 1994. Structure and function of glu-
Tomita, T., Kono, Y., Shimada, T., 1996. Chromosomal localisation tathione S-transferases. Biochim. Biophys. Acta 1205, 1–18.
of amplified esterase genes in insecticide resistant Culex mos- Williamson, M.S., Martinez-Torres, D., Hick, C.A., Devonshire,
quitoes. Insect Biochem. Mol. Biol. 26, 853–857. A.L., 1996. Identification of mutations in the housefly para-type
Vais, H., Williamson, M.S., Devonshire, A.L., Usherwood, P.N.R., sodium channel gene associated with knock-down resistance (kdr)
2001. The molecular interactions of pyrethroid insecticides with to pyrethroid insecticides. Mol. Gen. Genet. 252, 51–60.
insect and mammalian sodium channels. Pest Manag. Sci. 57, Yu, S.J., 1984. Interactions of allelochemicals with detoxication
877–888. enzymes of insecticide-susceptible and resistant fall army-worms.
Vaughan, A., Hemingway, J., 1995. Mosquito carboxylesterase Est Pestic. Biochem. Physiol. 22, 60–68.
alpha 2(1) (A2). Cloning and sequence of the full-length cDNA Yu, S.J., 1996. Insect glutathione S-transferases. Zoological Stud. 35,
for a major insecticide resistance gene worldwide in the mosquito 9–19.
Culex quinquefasciatus. J. Biol. Chem. 270, 17044–17049. Zhu, Y.C., Dowdy, A.K., Baker, J.E., 1999. Differential mRNA
Vaughan, A., Rodriguez, M., Hemingway, J., 1995. The independent expression levels and gene sequences of a putative carboxylester-
gene amplification of indistinguishable esterase B electromorphs ase-like enzyme from two strains of the parasitoid Anisopter-
from the insecticide resistant mosquito Culex quinquefasciatus. omalus calandrae (Hymenoptera: Pteromalidae). Insect Biochem.
Biochem. J. 305, 651–658. Mol. Biol. 29, 417–425.
Vaughan, A.M., Hawkes, N.J., Hemingway, J., 1997. Co-amplifi- Ziegler, R., Whyard, S., Downe, A.E.R., Wyatt, G.R., Walker,
cation explains linkage disequilibrium of two mosquito esterase V.K., 1987. General esterase, malathion carboxylesterase and
genes in insecticide resistant Culex quinquefasciatus. Biochem. J. malathion resistance in Culex tarsalis. Pest. Biochem. Physiol. 28,
325, 359–365. 279–285.

You might also like