ArticlePDF Available

Identification of a novel class of insect glutathione S-transferases involved in resistance to DDT in the malaria vector Anopheles gambiae

Authors:

Abstract and Figures

The sequence and cytological location of five Anopheles gambiae glutathione S-transferase (GST) genes are described. Three of these genes, aggst1-8, aggst1-9 and aggst1-10, belong to the insect class I family and are located on chromosome 2R, in close proximity to previously described members of this gene family. The remaining two genes, aggst3-1 and aggst3-2, have a low sequence similarity to either of the two previously recognized classes of insect GSTs and this prompted a re-evaluation of the classification of insect GST enzymes. We provide evidence for seven possible classes of insect protein with GST-like subunits. Four of these contain sequences with significant similarities to mammalian GSTs. The largest novel insect GST class, class III, contains functional GST enzymes including two of the A. gambiae GSTs described in this report and GSTs from Drosophila melanogaster, Musca domestica, Manduca sexta and Plutella xylostella. The genes encoding the class III GST of A. gambiae map to a region of the genome on chromosome 3R that contains a major DDT [1,1,1-trichloro-2,2-bis-(p-chlorophenyl)ethane] resistance gene, suggesting that this gene family is involved in GST-based resistance in this important malaria vector. In further support of their role in resistance, we show that the mRNA levels of aggst3-2 are approx. 5-fold higher in a DDT resistant strain than in the susceptible strain and demonstrate that recombinant AgGST3-2 has very high DDT dehydrochlorinase activity.
Content may be subject to copyright.
Biochem. J. (2001) 359, 295–304 (Printed in Great Britain) 295
Identification of a novel class of insect glutathione S-transferases involved
in resistance to DDT in the malaria vector Anopheles gambiae
Hilary RANSON*
1
, Louise ROSSITER*, Federica ORTELLI*, Betty JENSEN, Xuelan WANG, Charles W. ROTH,
Frank H. COLLINSand Janet HEMINGWAY*
*School of Biosciences, Main College, Cardiff University, PO Box 915, Cardiff CF10 3TL, Wales, U.K., Department of Biological Sciences, University of Notre Dame,
PO Box 369, Notre Dame, IN 46556, U.S.A., and Unite
!
de Biochimie et Biologie Mole
!
culaire des Insectes, Institut Pasteur, 75015 Paris, France
The sequence and cytological location of five Anopheles gambiae
glutathione S-transferase (GST) genes are described. Three of
these genes, aggst1-8, aggst1-9 and aggst1-10, belong to the
insect class I family and are located on chromosome 2R, in close
proximity to previously described members of this gene family.
The remaining two genes, aggst3-1 and aggst3-2, have a low
sequence similarity to either of the two previously recognized
classes of insect GSTs and this prompted a re-evaluation of the
classification of insect GST enzymes. We provide evidence for
seven possible classes of insect protein with GST-like subunits.
Four of these contain sequences with significant similarities to
mammalian GSTs. The largest novel insect GST class, class III,
contains functional GST enzymes including two of the A. gambiae
GSTs described in this report and GSTs from Drosophila
INTRODUCTION
Glutathione S-transferases (GSTs) are a major family of detoxifi-
cation enzymes found in most organisms. They help to protect
cells from oxidative stress and chemical toxicants by aiding the
excretion of electrophilic and lipophilic compounds from the cell
(reviewed in [1]). Eukaryotes contain multiple GSTs with differing
catalytic activities to accommodate the wide range of functions
of this enzyme family. Mammalian GSTs have been classified
into eight cytosolic classes (Alpha, Mu, Pi, Theta, Sigma, Zeta,
Kappa and Omega) and a microsomal class [2–7], whereas only
two classes of insect GSTs (classes I and II) have so far been
described [8]. (An alternative nomenclature in which the insect
classes are assigned Greek letters in line with the mammalian
GST classification system has been proposed [9] and is discussed
in the present paper). The insect class I GSTs are encoded by a
large complex gene family. Additional heterogeneity within this
class is introduced by alternative splicing in Anopheles gambiae
and the presence of fusion genes in Musca domestica [10,11].
In Drosophila melanogaster and A. gambiae this gene family
is tightly clustered [10,12], in contrast with the family in
M. domestica, in which the class I GSTs are dispersed throughout
the genome [13]. The class II insect GST family consists of a
single gene in all three species [14,15].
Abbreviations used: BAC, bacterial artificial chromosome ; CDNB, 1-chloro-2,4-dinitrobenzene; DCNB, 1,2-dichloro-4-nitrobenzene; DDE, 1,1-
dichloro-2,2-bis-(p-chlorophenyl)ethane; DDT, 1,1,1-trichloro-2,2-bis-(p-chlorophenyl)ethane; GST, glutathione S-transferase; RACE, rapid amplifi-
cation of cDNA ends.
1
Present address and address for correspondence : Parasite and Vector Biology Division, Liverpool School of Tropical Medicine, Pembroke Place,
Liverpool L3 5QA, U.K.) (e-mail HRanson!liverpool.ac.uk).
The nucleotide sequence data reported will appear in DDBJ, EMBL and GenBank2 Nucleotide Sequence Databases under the accession numbers
AF316635 to AF316638.
melanogaster, Musca domestica, Manduca sexta and Plutella
xylostella. The genes encoding the class III GST of A. gambiae
map to a region of the genome on chromosome 3R that contains
a major DDT [1,1,1-trichloro-2,2-bis-(p-chlorophenyl)ethane]
resistance gene, suggesting that this gene family is involved in
GST-based resistance in this important malaria vector. In further
support of their role in resistance, we show that the mRNA levels
of aggst3-2 are approx. 5-fold higher in a DDT resistant strain
than in the susceptible strain and demonstrate that recombinant
AgGST3-2 has very high DDT dehydrochlorinase activity.
Key words: classification of insect GST enzymes, Drosophila
genome, insecticide resistance.
Interest in insect GSTs is focused on the role of these enzymes
in insecticide resistance. Elevated GST activity has been de-
tected in strains of insects resistant to organophosphates [8] and
organochlorines [16] and this enzyme family has recently been
implicated in resistance to pyrethroid insecticides [17,18].
A. gambiae GSTs are of particular interest because of their
involvement in resistance to DDT [1,1,1-trichloro-2,2-bis-
(p-chlorophenyl)ethane] in this important malaria vector. In the
1950s and 1960s house spraying with DDT was the primary line
of defence against malaria and, although the advent of DDT-
resistant strains of mosquitoes has decreased the effectiveness of
this control measure, this insecticide is still used today for
malaria control in many parts of the world [19]. In A. gambiae,
an increased rate of DDT dehydrochlorination in the resistant
strain is associated with quantitative increases in multiple GST
enzymes [20].
We have studied the A. gambiae class I and class II GST genes
to ascertain their role in conferring DDT resistance. The single
class II GST, aggst2-1 [15] is highly expressed in A. gambiae
larvae but is barely detectable in adult insects. Because DDT
resistance in this species is life-stage specific and the insecticides
are used as adulticides both in the field and for selection of the
resistant strain in the laboratory, the developmental expression
profile discounted a prominent role for aggst2-1 in conferring
# 2001 Biochemical Society
296 H. Ranson and others
DDT resistance. The class I GSTs are expressed at high levels in
both larvae and adults [10]. Several recombinant A. gambiae
GST enzymes are able to metabolize DDT but, by using
antibodies raised against these class I GSTs, we have demon-
strated that these enzymes are not the most important family in
DDT resistance [21]. Furthermore, we have no evidence to
suggest that any of these genes are overexpressed in resistant
mosquitoes (N. Roberts and J. Hemingway, unpublished work).
Hence either A. gambiae contains additional genes encoding
class I GSTs, or further classes of insect GSTs exist. We now
present evidence to support both these hypotheses. We report the
cloning of five novel A. gambiae genes encoding GSTs. Three of
these have been classified as class I GSTs, whereas the remaining
two genes belong to a previously undescribed class, which we
have named class III. We propose that the class III GSTs
represent the major enzyme family conferring resistance to DDT
in the malaria mosquito, A. gambiae.
EXPERIMENTAL
Mosquito strains
The ZAN\U strain of A. gambiae was colonized from a DDT-
resistant field population from Zanzibar, Tanzania, in 1982. This
strain has been maintained under regular adult selection pressure
with DDT. Kisumu is a laboratory insecticide-susceptible strain
originally colonized from Kisumu, western Kenya. The PEST
strain is fixed for the standard chromosome arrangement [22]
and was used to construct the A. gambiae bacterial artificial
chromosome (BAC) library (X. Wang, Z. Ke, A. J. Cornel, D.
Smoller and F. H. Collins, unpublished work).
DNA extraction and sequencing
BAC DNA was isolated with Qiagen Plasmid maxi kits. BAC
sequencing reactions were performed with 1 µg of BAC DNA as
a template and ABI BigDye Terminator chemistry. After electro-
phoresis on an ABI 377 automatic sequencer, contigs were
assembled and the sequences annotated with the LASERGENE
software package (DNAstar, Madison, WI, U.S.A.).
Total RNA was extracted from individual mosquitoes with the
TRI reagent (Sigma), in accordance with the manufacturer’s
instructions. The RNA was treated with DNase to remove any
contaminating genomic DNA and the mRNA was reverse-
transcribed into cDNA by using Superscript II (Gibco BRL) and
an oligo(dT) adapter primer (5h-GACTCGAGTCGACATCGA-
(dT)
"(
-3h).
Genomic DNA was extracted from individual adult mos-
quitoes as described previously [23].
In situ hybridization
BAC clones were physically mapped to polytene chromosomes
prepared from half-gravid ovaries of the PEST strain of
A. gambiae as described previously [24].
Quantification of aggst3-2 mRNA levels
Incorporation of the fluorescent dye SYBR GreenI (Molecular
Probes) into double-stranded PCR products was used to de-
termine the mRNA copy number of aggst3-2 in individual
mosquitoes. An aggst3-2 standard plasmid was constructed by
inserting a 353 bp fragment from the coding region of the
aggst3-2 gene, amplified from ZAN\U cDNA with the primers
3-2f (5h-GTACGATCATCACCGAGAGC-3h) and 3-2r (5h-
CTTCGACTGCTCCAACGGC-3h), into pGEM T-easy vector
(Promega). A control plasmid was constructed by inserting a
partial fragment from the gene encoding ribosomal S7 protein
[25], amplified with primers SPC (5h-GTGCCGGTGCCGA-
AACAGAA-3h) and SPD (5h-AGCACAAACACTCCAATA-
ATCAAG-3h), into the pGEM T-easy vector. These plasmids
were used as template DNA at concentrations ranging from 1 ng
to 10 fg to produce standard curves using a Roche Lightcycler
in accordance with the manufacturer’s recommended protocols.
For quantification of the copy number, approx. 2% of the
cDNA from an individual mosquito was used as a template for
the control (S7) primers and 6 % was used for quantifying
aggst3-2 expression; 40 rounds of amplification were performed
in glass capillaries containing 5 pmol of each primer, 1iSYBR
GreenI mix and a final concentration of 3 mM MgCl
#
. The
amplification cycle was as follows : 95 mC for 1 s, 62 mC for 3 s
and 72 mC for 15 s, with incorporation of fluorescence measured
at 87 mC for aggst3-2 quantification, and 95 mC for 1 s, 60 mC for
3 s and 72 mC for 10 s, with incorporation of fluorescence
measured at 86 mC for S7 quantification. Each sample was
analysed in duplicate in each experiment and the results are
means for two separate experiments. The data were quantified
with LightCycler Software V3 (Roche) and converted into copy
number as described in [26].
Expression of aggst3-2 in vitro
The coding region of aggst3-2 was amplified in a PCR reaction
with ZAN\U cDNA as a template, Pfu polymerase (Stratagene)
and primers that contained the initiation and termination codons
of the gene preceded by BamH1 sites. The single product of
approx. 680 bp was subcloned into T-easy (Promega) and
sequenced to ensure that no errors had been introduced during
amplification. The insert was then isolated by digestion with
BamH1, ligated into the BamH1 site of the pET3a vector
(Novagen) and the resultant expression construct was used to
transform Escherichia coli Origami (DE3)pLysS cells. The orient-
ations of the inserts were determined by restriction digestion ;
colonies containing the insert in both the forward and reverse
orientations were grown at 37 mC to an attenuance (D
'!!
) of 0.6.
Expression of the recombinant protein was induced by the
addition of isopropyl β--thiogalactoside to 0.4 mM and, after
incubation for a further 3 h, the cells were harvested by centri-
fugation for 10 min at 5000 g. After a single round of freeze–
thawing, the cells were resuspended in 50 mM Tris\HCl (pH 8.0)\
2 mM EDTA\0.1 M NaCl. Protein concentration was deter-
mined with Bio-Rad protein reagent [27] ; GST activity was
assayed spectrophotometrically by measuring the conjugation of
GSH to the standard GST substrates 1-chloro-2,4-dinitrobenzene
(CDNB) and 1,2-dichloro-4-nitrobenzene (DCNB) [28].
DDT dehydrochlorinase activity was assayed by incubating
the crude cell extract with 0.1 mM DDT and 10 mM GSH in
0.1 M sodium phosphate buffer, pH 6.5, for 2 h at 30 mC. The
samples were extracted twice with chloroform, air-dried and then
resuspended in propan-2-ol. HPLC analysis of DDT metabolites
was performed as described by Prapanthadara et al. [20], with a
flow rate of 0.6 ml\min. Constructs containing the insert in the
negative orientation were assayed to control for non-enzymic
DDT metabolism. Controls omitting GSH in the incubation
mixture were also included to verify the dependence of the
reaction on GSH.
Phylogenetic analysis of insect GSTs
A search of the GenBank2 database located 21 non-Drosophila
insect GST sequences including seven A. gambiae GSTs. These
were retrieved and the putative amino acid sequences were aligned
# 2001 Biochemical Society
297Glutathione S-transferases and DDT resistance in Anopheles gambiae
against the A. gambiae GST sequences described here, with the
CLUSTAL W program [29] A total of 42 Drosophila sequences
predicted to contain GST protein domains are present in the D.
melanogaster genome [30]. These sequences were retrieved from
FlyBase (http:\\flybase.bio.indiana.edu) to enable us to in-
corporate them into our phylogenetic analysis. Ten of these
sequences are identical to previously submitted Drosophila
GST sequences and were included in our analysis as they appear
in GenBank2 (with the exception of DmGST26 and DmGST22,
which are reported to be pseudogenes [12] and were therefore
excluded). Of the remaining 32 putative Drosophila genes for
GST, four sequences (CG15100, CG4623, CG12304 and
CG11901) were considered unlikely to encode functional GST
enzymes on the basis of their transcript length and\or low degree
of similarity to genes encoding GST; they were therefore
discarded. The annotations of the remaining genes for GST were
studied and most of the amino acid translations were accepted as
published with the following exceptions. (1) Two putative full-
length transcripts were derived from the annotations for
CG12930 and CG6673. The derived amino acid sequences of
both of these were included in the study (denoted by A and B).
(2) The translation of CG1681 seems to be lacking 37 residues
at the N-terminus. A search of the nucleotide sequence of this
annotation identified a putative 5h exon, approx. 3 kb upstream
from the 3h end of the gene, which was joined to the amino acid
translation to produce a putative full-length gene. (3) Anno-
tations CG1702, CG10065 and CG17639 predicted very large
translation products of which only approx. 220 residues showed
significant similarity to GSTs. These translations were therefore
trimmed manually.
After updating the alignment to contain these Drosophila
sequences, evolutionary distances were calculated by using the
Jukes–Cantor algorithm [31]; phylogenetic trees were determined
by the neighbour-joining method [32] with TREECON for
Windows [33]. The amino acid translation of a Rat Kappa GST
(rGSTTK1-1 [5]) was used as an outgroup to root the tree.
RESULTS
Cloning of GSTs
As part of the A. gambiae genome initiative, the insert ends of
each clone from a BAC library have been determined by single-
pass sequencing at Genoscope and the Institut Pasteur
(www.genoscope.cns.fr\externe\English\Projets\ProjetIAK\
AK.html). The resultant sequences were queried against the
GenBank2 database and two BAC clones were identified in which
end sequences had significant similarity to GSTs (04H09 and
28I09). The end sequence of clone 04H09 (sequenced with primer
SP6) was predicted to encode the carboxy region of a GST.
Primers were designed to amplify this partial gene encoding a
GST and used to screen the BAC library for overlapping clones.
Four positive clones were identified (05C11, 06I12, 12H09 and
28A19) and primers designed against the 3h GST sequence in
04H09 were used for partial sequencing of clone 06I12. The
results of this sequencing not only completed the genomic
sequence of the GST gene present in the end sequence of clone
04H09, but also identified an additional gene encoding a GST,
approx. 350 bp downstream from this gene. These A. gambiae
GST genes have low levels of similarity to the two insect GST
classes previously recognized (class I and class II) and were
therefore tentatively assigned to a third class of insect GSTs and
named aggst3-1 and aggst3-2 (see the Discussion).
The end sequence of BAC clone 28I09 (sequenced with primer
SP6) contained the 3h end of one gene for GST (later named
aggst1-9) and the 5h end of a second gene encoding a GST
Figure 1 Alignment of deduced amino acid sequences of the five A.
gambiae genes encoding GST described in this report
Gaps introduced to maximize sequence identity are shown by a horizontal dash. Residues shown
in bold are shared by all class I and class III A. gambiae GSTs. The residues denoted by an
asterisk are shared by all known GSTs [37]. The arrows indicate intron positions (see the text
for further details).
Figure 2 Schematic representation of the A. gambiae genome showing
the location of genes for GSTs and loci associated with resistance to DDT
(rtd1 and rtd2)
Cytological positions are shown to the right of the chromosome and the two major regions of
the genome associated with DDT resistance are shown as solid bars.
(aggst1-8). Primers designed against the end sequence of 28I09
were used to search the BAC library for overlapping clones and
a single positive, 10O16, was identified and used as a template for
obtaining the full length-genomic sequences of genes aggst1-8
and aggst1-9.
3h-Rapid amplification of cDNA ends (RACE) reactions with
primers incorporating the initiator methionine codon of the
putative genes encoding GST were used to verify that these genes
were expressed in A. gambiae. Transcripts of aggst3-1, aggst3-2
# 2001 Biochemical Society
298 H. Ranson and others
0
10
20
30
40
010203040
0
10
20
30
40
010203040
10 15 20 25 30
–2
0
2
4
10 15 20 25 30
–2
0
2
4
–4
0
0.002
0.004
0.006
Kisumu ZAN/U
Copy no. ratio (Aggst3-2 / SP7)
C
AB
Fluorescence
Fluorescence
Cycle number Cycle number
1 ng
100 pg 10 pg 1 pg 100 fg
1 ng
100 pg 10 fg
1 pg
1 fg
Log conc. (pg)
Log conc. (pg)
Crossing Point Crossing Point
1 ng
100 pg
10 pg
1 pg
100 fg
1 ng
100 pg
1 pg
10 fg
1 fg
Figure 3 Quantification of aggst3-2 mRNA expression levels in individual A. gambiae adults
(A, B) Upper panels: SYBR GreenI fluorescence acquisition by PCR products from serially diluted (1 ng to 1 fg) standard plasmids and individual adult mosquito cDNA against cycle number.
Lower panels: standard curves derived from plotting the crossing points against the logarithm of copy number for plasmids S7 and aggst3-2.(A) SP7 standardization control ; (B) aggst3-2. (C)
mRNA copy number of aggst3-2 transcript relative to the number of SP7 transcripts for each individual mosquito.
and aggst1-8 were detected in adult mosquitoes. 3h-RACE with
a primer complementary to aggst1-9 gave a product of the
expected size but with only 36.8 % sequence identity to aggst1-9.
This gene transcript has relatively high similarity to aggst1-5 and
aggst1-6 (42.9 % and 42.4% similarity at the nucleotide level)
and was therefore provisionally classified as a class I GST and
named aggst1-10. The BAC library was screened with primers
specific to aggst1-10 to enable the genomic organization and
physical location of aggst1-10 to be established. As attempts at
3h-RACE for aggst1-9 were unsuccessful we designed PCR
primer pairs specific to this gene and used these in PCR reactions
with fourth-instar larvae, pupae and adult mosquito cDNA as
templates; however, we were unable to detect a transcript in any
of these life stages. This gene might be expressed in earlier life
stages but the possibility that aggst1-9 is a pseudogene cannot be
discounted at this stage.
An amino acid alignment of the five A. gambiae genes for GSTs
identified in this study is shown in Figure 1. The four invariant
residues proposed to be crucial for the correct folding of GST
enzymes [34] are conserved in these, and in all previously
identified, A. gambiae GST sequences (indicated by an asterisk in
Figure 1). In addition to these, a further eight residues are
constant in all A. gambiae GSTs. aggst1-8 and aggst1-10 contain
a single intron within the 5h coding region at an identical position
to the intron in the alternately spliced A. gambiae gene for GST,
aggst1α [10]. aggst3-1 and aggst3-2 also contained an intron at
this position and an additional intron at position 119 (numbers
according to AgGST3-1 sequence) (Figure 1). aggst1-9 is unique
among the newly described A. gambiae GSTs in being intronless.
Physical mapping
The cytological location of the A. gambiae genes encoding GSTs
was determined by in situ hybridization. Figure 2 shows the
positions of all these genes on the polytene chromosomes. Six
genes for GSTs are located on chromosome 2R within divisions
# 2001 Biochemical Society
299Glutathione S-transferases and DDT resistance in Anopheles gambiae
Figure 4 Heterologous expression of AgGST3-2
E. coli cultures containing aggst3-2 expression constructs, prepared with the insert in either the
correct or the reverse orientation within the pET3a vector, were induced with isopropyl
β-D-thiogalactoside. A 10 µl sample of whole cells was separated on a 12% (w/v) resolving,
5 % (w/v) stacking SDS/polyacrylamide gel and stained with Coomassie Blue R250. Left lane,
an E. coli culture containing aggst3-2 in the correct orientation ; middle lane, an E. coli culture
containing aggst3-2 in reverse orientation (negative control) ; right lane, molecular mass
standards (molecular masses indicated at the right). The subunit size of aggst3-2 is predicted
to be 24.9 kDa on the basis of its amino acid translation. A band of this approximate size is
clearly visible in the left lane.
18-19. These include the aggst1α, aggst1β and aggst1-2 genes
described previously [10,35] and the three class I GSTs described
here, namely aggst1-8, aggst1-9 and aggst1-10. Two possible sites
of hybridization for the single class II GST have been reported
[15] but only one of these, on division 38d, was confirmed by
further experiments. The class III GSTs are located on division
33c on chromosome 3R. This position coincides with the location
of one of two major quantitative trait loci associated with DDT
resistance in the ZAN\U strain of A. gambiae [36] (Figure 2).
Figure 5 HPLC analysis of DDT dehydrochlorinase activity by crude cell extracts expressing recombinant AgGST3-2
(A) Control insert in reverse orientation. (B) Insert in correct orientation; 50 µl of cell extract. (C) Insert in correct orientation ; 200 µl of cell extract.
Quantitative analysis of aggst3-2 expression
The co-localization of the A. gambiae class III genes with a major
gene conferring DDT resistance prompted us to study the relative
expression levels of members of this insect class in susceptible
and resistant insects by using real-time PCR technology. From
the standard curves shown in Figures 3(A) and 3(B) it was
possible to extrapolate the fluorescence values obtained with
Kisumu and ZAN\U cDNA and calculate the initial tem-
plate copy number. By dividing the copy number of aggst3-2 by
the copy number of S7 the values for GST expression in each
individual mosquito were standardized for variations in initial
cDNA concentrations so that the relative expression of aggst3-2
between the susceptible and resistant strains could be compared
(Figure 3C). The average ratio of aggst3-2 copy number to S7
copy number in the ZAN\U strain was (4.6p1.72)i10
$
com-
pared with (9.5p4.75)i10
%
in the Kisumu strain, representing
an approximate 5-fold overexpression in the resistant strain.
Expression in vitro
To verify that the A. gambiae class III GSTs encoded catalytically
active enzymes, we expressed aggst3-2 in E. coli (Figure 4) and
measured the CDNB- and DCNB-conjugating activity of the
crude protein homogenates. No CDNB-conjugating activity was
detectable in the control cultures but replicate crude cell extracts
from two separate E. coli cultures expressing recombinant
AgGST3-2 had a mean CDNB-conjugating activity of 2.879p
0.8 µmol\min per mg of crude protein and a mean DCNB-
conjugating activity of 5.74p2.7 µmol\min per mg of crude
protein.
DDT dehydrochlorinase activity of recombinant AgGST3-2
was measured as nmol of 1,1-dichloro-2,2-bis-(p-chlorophenyl)
ethane (DDE) detected by HPLC analysis after incubation of
the crude cell extract with 100 nmol of DDT as described in the
Experimental section. The recovery of DDT\DDE after extrac-
tion and analysis ranged from 43% to 58 %. Representative
results are shown in Figure 5. DDE was undetectable in the
control reactions containing the insert in the negative orientation
# 2001 Biochemical Society
300 H. Ranson and others
Table 1 Pairwise percentage similarities between derived amino acid sequences of A. gambiae GSTs
Sequences shown in bold are described here for the first time.
Similarity (%)
Sequence agGST1-3 agGST1-4 agGST1-5 agGST1-6 agGST1-7 agGST1-8 agGST1-9 agGST1-10 agGST2-1 agGST3-1 agGST3-2
agGST1-2 47.4 46.4 49.3 49.8 36.4 39.2 37.3 27.8 12.9 27.8 30.1
agGST1-3 58.4 61.2 60.8 33.5 34.2 27.9 29.4 11.0 27.9 29.2
agGST1-4 63.2 65.6 39.6 42.4 34.6 29.4 11.1 30.9 32.7
agGST1-5 82.8 43.1 41.1 35.9 35.4 14.8 33.5 37.3
agGST1-6 45.5 43.1 35.4 35.9 14.4 34.9 38.8
agGST1-7 39.9 32.1 29.4 11.5 30.3 34.4
agGST1-8 37.7 29.9 11.9 26.8 29.9
agGST1-9 28.9 16.1 24.5 26.4
agGST1-10 12.8 28.0 27.5
agGST2-1 10.1 11.9
agGST3-1 63.8
Figure 6 Dendrogram illustrating the relationship between insect GSTs
Amino acid sequences were aligned by using CLUSTAL W and the tree was constructed with the neighbour-joining method program from a similarity matrix of pairwise comparisons made by
using the Jukes–Cantor algorithm. Selected bootstrap values from 500 replicate trees are shown (as percentage values) at the dendrogram nodes. The sequences denoted by CG were obtained
from the Drosophila genome annotations (http://flybase.bio.indiana.edu) as described in the text. All other sequences were retrieved from GenBank2 or are described in this study. The tree was
rooted with the rat gene encoding Kappa GST (GenBank2 accession number S83436). Abbreviations : Dm, D. melanogaster ; Md, M. domestica ; Lucil, Lucilia cuprina ; Ag, A. gambiae ; Cv, Culicodies
variipennis ; Ae, Ae. aegypti ; Plx, P. xylostella ; Msex, Ma. sexta.
(Figure 5A) and in the absence of GSH, indicating that DDT
dehydrochlorinase activity in the assay was dependent on both
enzyme and GSH. In the experimental assays expressing recombi-
nant AgGST3-2, the percentage conversion of DDT to DDE was
dependent on the amount of crude extract used in the assay.
For example, in the experiment shown in Figure 5(B), 50 µlof
extract was used ; 24 nmol of DDE and 33 nmol of DDT were
detected by HPLC. When the volume of cell extract was increased
to 200 µl (Figure 5C) 43 nmol of DDE was recovered, repre-
senting a 92% conversion of DDT to DDE. If these values are
expressed as nmol of DDE\µg of protein, values of 12.5 and 5.5
are obtained for the experiments shown in Figures 5(B) and 5(C)
respectively, suggesting that the concentration of DDT was rate-
limiting in the experiment shown in Figure 5(C).
# 2001 Biochemical Society
301Glutathione S-transferases and DDT resistance in Anopheles gambiae
Figure 7 Dendrogram illustrating the relationship between Drosophila GSTs and mammalian representatives from each of the evolutionarily distinct GST
classes
See the legend to Figure 6 for details.
Phylogenetic analysis
Table 1 shows the percentage similarity between the deduced
amino acid sequences of all known A. gambiae GSTs. Previous
classifications have designated GSTs as being members of the
same class if their amino acid sequences are more than 40 %
identical [1]. By this criterion, only aggst1-8 of the five newly
described A. gambiae GSTs would be classified as belonging to
the insect class I family and none of these genes would be
classified as class II. This suggests that the present classification
of insect GSTs into only two classes might need re-evaluating.
We therefore conducted a phylogenetic analysis of all the known
insect GST sequences, including 28 sequences retrieved from the
Drosophila genome database.
Figure 6 shows a phylogenetic tree illustrating the relationships
between these sequences, based on a CLUSTAL W amino acid
alignment. The available insect GST sequences can be split into
at least seven subdivisions on the basis of this phylogeny. Only
three of these contain sequences with confirmed GST activity.
These are the pre-existing classes I and class II plus a third clade,
which we have called class III, in line with the existing no-
menclature for insect GSTs, containing 20 sequences including
aggst3-1 and aggst3-2 and the published genes DmGST-3 from
D. melanogaster, GST-3 from Plutella xylostella and MsGST1
from Manduca sexta [37–39]. The low support for the monophyly
of class I and of class III, indicated by the bootstrap values in
Figure 6, perhaps suggests that these classes should be further
subdivided. We therefore used amino acid distance matrices
(results not shown) to examine the support for this classification.
In this approach, we classified a gene for GST as belonging to a
particular class if it satisfied the following two criteria: (1) at
least 40% sequence similarity to a member of this class from a
different species, and (2) less than 40 % sequence similarity to all
other classes of insect GST. The assignment of genes to classes II
and III is supported by these criteria but ambiguities arose in the
classification of class I. In Figure 6, and in the choice of
nomenclature, we have classified aggst1-9 and aggst1-10 (plus the
putative Drosophila GST, cg10065) as class I GSTs. However,
these sequences show less than 40% sequence similarity to other
members of this group and therefore do not belong to this class
on the basis of previously published criteria [1]. Nevertheless,
because these genes show the greatest levels of identity with class
I GSTs out of all currently known insect GSTs, we propose to
assign these GSTs to class I at present.
Relationship between mammalian and insect GSTs
The remaining four subdivisions shown in Figure 6 consist solely
of sequences with GST-like domains, retrieved from the Droso-
phila genome database, and have not been experimentally verified
as functional GST enzymes. CG4688 and CG11784 have the
greatest similarity to MdGST6A (31.5% and 35.1% respectively)
but, because these levels of similarity are below the arbitrary cut-
off value, they have been classified as belonging to a separate
# 2001 Biochemical Society
302 H. Ranson and others
subdivision in Figure 6. CG9362 and CG9363 are closely related
(62.6% similarity to each other) but have less than 20 % similarity
to any other known insect GST sequence. Similarly, the amino
acid similarities support the existence of a distinct subdivision
containing CG6773A, CG6773B, CG6781, CG6662 and CG6776
and a separate subdivision containing CG1702, CG12930A and
CG12930B; however, the members of both these subdivisions
have very low similarity to previously characterized insect GSTs.
To investigate the relationship between the insect and mam-
malian GST classes, we took a representative member of each
of the eight cytosolic classes of mammalian GST and used
CLUSTAL W to align these sequences with the putative
Drosophila GST sequences. This alignment was used to generate
the phylogenetic tree shown in Figure 7. Previous classifications
have denoted all insect GSTs as belonging to the Theta class [34]
but the phylogeny shown in Figure 7 does not support this
conclusion. Although this tree is not intended to resolve the true
phylogenetic relationship between GST families, it is interesting
to note that representatives from the Theta, Sigma, Omega and
Zeta class are present in Drosophila. For example, the Drosophila
sequences CG9362 and CG9363 possess 56–58% identity with
the human Zeta GST, GSTZ1-1, and these putative insect GSTs
contain N-terminal motifs closely related to the SSCXWR-
VRIAL (single-letter amino acid codes) motif found in Zeta class
GSTs from plants, nematodes and mammals [4] (the predicted
translations for CG9362 and CG9363 contain the motifs SSCS-
WRVRVAL and SSCSWRVRIAM respectively). A further
subdivision of insect GSTs shows high similarity to the Omega-
class human GST GST O1-1. All five putative Drosophila
sequences within this subdivision, which are clustered on
chromosome 3L, also contain a cysteine residue flanked by
phenylalanine and proline residues at the proposed active site [7].
Finally, as noted recently [40], the insect class II enzymes are
phylogenetically related to the Sigma class.
DISCUSSION
Evolutionary relationship between insect GSTs
We have identified five A. gambiae genes encoding GSTs, at least
four of which are actively transcribed in adult mosquitoes. The
fifth gene, aggst1-9, seems to be transcriptionally silent in adult
and larval mosquitoes. It is not known whether this gene is
expressed during earlier life stages or whether it is a pseudogene.
A comparison of amino acid sequence similarities found that
four of the genes described here were below the threshold for
inclusion in either insect class I or class II, suggesting that the
current classification system for insect GSTs is inadequate. This
observation is supported by biochemical data from A. gambiae
that identified at least eight fractions with GST activity, only one
of which was immunologically related to the class I GST family
[20]. The publication of the first draft of the Drosophila genome
[41] prompted us to re-examine the classification of insect GSTs.
Our analysis supported the existence of seven possible classes of
protein with GST-like subunits. The existing class I and II
families were resolved by our analysis and five additional classes
were proposed. Of these five classes, only the largest, which we
have named class III, includes insect proteins with confirmed
GST activity ([38] and the present study). In our choice of
Drosophila sequences to include in the analysis we selected only
those encoding peptides of the approximate size of GST subunits
(approx. 25 kDa) and showing significant similarity throughout
the entire sequence rather than just at the N-terminus. Never-
theless, we acknowledge that some genes might have GST-like
domains but not possess GST activity and likewise that some
proteins might have acquired GST activity as a result of
convergent evolution [42]. Hence we might have inadvertently
included sequences that are not part of the same phylogenetic
gene tree.
A preliminary investigation into the relationship between the
insect GST classes and the previously characterized mammalian
cytosolic GST classes revealed that four of the subgroups of
insect GSTs have significant sequence similarity to confirmed
GST classes (Theta, Sigma, Omega and Zeta). The class I
insect GST subgroup has also been referred to as the Delta class
and the insect class II as Sigma [9], in line with the nomenclature
of the existing GST classes. The insect class III family, described
for the first time here, does not belong to the Delta class or to any
of the other existing GST families on the basis of established
criteria. Therefore, to maintain consistency with the proposed
nomenclature for GST classes [9], the insect class III would
perhaps be more appropriately denoted by a Greek character.
We suggest that hereafter the insect class III family be referred to
as Epsilon (ε).
Of all known A. gambiae genes encoding GSTs, aggst1-2 and
aggst1-9 are unique in that their open reading frames are
uninterrupted by introns, although the presence of introns in the
5h non-coding sequence, as found in aggst1β, MdGST1 and
DmGST1 [10,43], cannot be discounted. The remaining class I
and class III GSTs in A. gambiae all possess an intron at the
identical position in the 5h coding region and aggst3-1 and
aggst3-2 also have a second intron within the centre of the open
reading frame. This contrasts with the situation in Drosophila,in
which none of the class I GSTD genes clustered on chromosome
3R division 87B [12] have introns within their open reading
frames, and furthermore an analysis of the 10 putative class III
GSTs on chromosome 2R division 55C retrieved from the
Drosophila genome database predicts that these genes are also
intronless [41]. However, it is not true that all Drosophila GST
genes are intronless. For example, the class II Drosophila
GST DmGST-2 is interrupted by two introns [14].
The conservation of intron\exon boundaries across the class I
and class III genes in A. gambiae and the absence of introns in the
homologous gene families in Drosophila might suggest that
the duplication events that occurred to produce the class I and
class III lineages occurred after the divergence of the Nematocera
and Brachycera Dipteran suborders. This hypothesis, however,
is not supported when all members of the class I and class III
families are considered. For example, a homologue of the A.
gambiae class I gene aggst1-7 has been identified in Drosophila
(CG17639) and both of these genes possess two introns at
identical sites (H. Ranson and N. Roberts, unpublished work),
perhaps suggesting a common ancestor for these two genes.
Role of A. gambiae GSTs in insecticide resistance
An association between elevated GST activity and insecticide
resistance has been observed in many insect species but there
have been very few reports describing the individual enzymes
involved. GST-2 from the mosquito Aedes aegypti is over-
expressed in a DDT-resistant strain [16] but the ability of this
enzyme to metabolize DDT has not been established and there-
fore the significance of this result is not clear. In addition,
there have been reports that expression of housefly MdGST-3 is
positively correlated with resistance and that recombinant
MdGST-3 is able to degrade the insecticide dimethylparathion
[11,44]. However, the genomic organization of this gene seems
to be extremely complex with a variant gene copy number
in different strains and hence the exact role of this enzyme in
insecticide resistance is difficult to ascertain [11]. To our know-
ledge there have been only two substantiated reports of a direct
# 2001 Biochemical Society
303Glutathione S-transferases and DDT resistance in Anopheles gambiae
relationship between GST overexpression and resistance. The
first is in the diamondback moth, P. xylostella. In this insect,
increased expression of the PxGST3 gene, which encodes an
enzyme capable of degrading organophosphorous insecticides, is
strongly correlated with resistance [38]. The second example is in
D. melanogaster, in which a recombinant GST enzyme, GST D1,
exhibiting DDTase activity was found at elevated levels in a
DDT-resistant strain [45].
With our results from A. gambiae we have shown that members
of the class I family of insect GSTs are not of major importance
in DDT resistance [21] and the expression profile of class II also
discounts a major role for this gene in resistance [15]. To
establish the identity of the GST enzymes responsible for
resistance, we conducted a genome-wide scan to identify regions
of the genome associated with resistance to DDT. We identified
two major loci, the first, rtd1, on chromosome 3R between
divisions 32c and 34c, and the second, rtd2, on chromosome 2L
in close proximity to division 21 [36]. Two of the genes encoding
GST described in this study map to chromosome 3R, division
33c, i.e. in the exact midpoint of the boundaries defined by rtd1,
invoking the hypothesis that this resistance locus is a cis-acting
regulatory element controlling the expression of these class III
GSTs. In support of this we have now shown that aggst3-2 is
overexpressed in the resistant strain and that recombinant
aggst3-2 is very efficient at metabolizing DDT. DDT dehydro-
chlorinase activity has previously been reported for recombi-
nant GSTs from A. gambiae [21] but this is the first definitive
demonstration that a GST with DDT dehydrochlorinase activity
is overexpressed in a DDT-resistant strain of mosquitoes. By
analogy with Drosophila, in which 10 class III genes for GST
are tightly clustered within approx. 14 kb of DNA [41], it is likely
that the class III family in A. gambiae extends beyond the two
members described here. If multiple members of this gene family
are under the control of a common regulatory factor, a mutation
in this factor could account for the elevated activity of several
different GST enzymes observed in our earlier biochemical studies
[21,46].
We acknowledge the work of the Drosophila genome-sequencing community for
generating the data referenced here and the assistance of Dr John Vontas in HPLC
analysis. C. W. R. acknowledges support from the Institut Pasteur, the Centre
National de la Recherche Scientifique and Genoscope-Centre national de se
!
quenc
:
age,
France. This work was partly funded by Wellcome Trust, Royal Society and
Leverhulme fellowships (to H. R.) and a grant from the John D. and Catherine T.
MacArthur Foundation (to F. H. C.).
REFERENCES
1 Hayes, J. D. and Pulford, D. J. (1995) The glutathione S-transferase supergene
family Regulation of GST and the contribution of the isoenzymes to cancer
chemoprotection and drug resistance. Crit. Rev.Biochem. Mol. Biol. 30, 445–600
2 Mannervik, B. (1985) The isoenzymes of glutathione transferase. Adv. Enzymol. 57,
357–417
3 Meyer, D. J., Coles, B., Pemble, S. E., Gilmore, K. S., Fraser, G. M. and Ketterer, B.
(1991) Theta, a new class of glutathione transferases purified from rat and man.
Biochem. J. 274, 409–414
4 Board, P. G., Baker, R. T., Chelvanayagam, G. and Jermiin, L. S. (1998) Zeta, a novel
class of glutathione transferases in a range of species from plants to humans.
Biochem. J. 328, 929–939
5 Pemble, S. E., Wardle, A. F. and Taylor, J. B. (1996) Glutathione S-transferase class
Kappa: characterization by the cloning of rat mitochondrial GST and identification of a
human homologue. Biochem. J. 319, 749–754
6 DeJong, J. L., Morgenstern, R., Jornvall, H., DePierre, J. W. and Tu, C.-P. D. (1988)
Gene expression of rat and human microsomal glutathione S-transferases. J. Biol.
Chem. 263, 8430–8436
7 Board, P. G., Coggan, M., Chelvanayagam, G., Easteal, S., Jermin, L. S., Schulte,
G. K., Danley, D. E., Hoth, L. R., Griffor, M. C., Kamath, A. V. et al. (2000)
Identification, characterization, and crystal structure of the omega class glutathione
transferases. J. Biol. Chem. 275, 24798–24806
8 Fournier, D., Bride, J. M., Poire, M., Berge, J. B. and Plapp, F. W. (1992) Insect
glutathione S-transferases. Biochemical characteristics of the major forms from
houseflies susceptible and resistant to insecticides. J. Biol. Chem. 267, 1840–1845
9 Chelvanayagam, G., Parker, M. W. and Board, P. G. (2001) Fly fishing for GSTs : a
unified nomenclature for mammalian and insect glutathione transferases. Chem.–Biol.
Interact. 133, 256–260
10 Ranson, H., Collins, F. H. and Hemingway, J. (1998) The role of alternative mRNA
splicing in generating heterogeneity within the Anopheles gambiae class I GST family.
Proc. Natl. Acad. Sci. U.S.A. 95, 14284–14289
11 Syvanen, M., Zhou, Z., Wharton, J., Goldsbury, C. and Clark, A. (1996) Heterogeneity
of the glutathione transferase genes encoding enzymes responsible for insecticide
degradation in the housefly. J. Mol. Evol. 43, 236–240
12 Toung, Y.-P. S., Hsieh, T. and Tu, C.-P. D. (1993) The glutathione S-transferase D
genes: a divergently organized, intronless gene family in Drosophila melanogaster.
J. Biol. Chem. 268, 9737–9746
13 Zhou, Z.-H. and Syvanen, M. (1997) A complex glutathione transferase gene family in
the housefly Musca domestica. Mol. Gen. Genet. 256, 187–194
14 Beall, C., Fyrberg, C., Song, S. and Fyrberg, E. (1992) Isolation of a Drosophila gene
encoding glutathione S-transferase. Biochem. Genet. 30, 515–527
15 Reiss, R. A. and James, A. A. (1993) A glutathione S-transferase gene of the vector
mosquito, Anopheles gambiae. Insect Mol. Biol. 2, 25–32
16 Grant, D. F. and Hammock, B. D. (1992) Genetic and molecular evidence for a
trans-acting regulatory locus controlling glutathione S-transferase-2 expression in
Aedes aegypti. Mol. Gen. Genet. 234, 169–176
17 Kostaropoulos, I., Papadopoulos, A. I., Metaxakis, A., Boukouvala, E. and
Papadopoulou-Mourkidou, E. (2001) Glutathione S-transferase in the defense against
pyrethroid insecticides. Insect Mol. Biol. 31, 313–319
18 Vontas, J. G., Small, G. J. and Hemingway, J. (2001) Glutathione S-transferases as
antioxidant defence agents confer pyrethroid resistance in Nilaparvata lugens.
Biochem. J. 357, 65–72
19 Bradley, D. J. (1998) The particular and general: issues of specificity and verticality
in the history of malaria control. Parassitologia 40, 5–10
20 Prapanthadara, L., Hemingway, J. and Ketterman, A. J. (1993) Partial purification and
characterization of glutathione S-transferase involved in DDT resistance from the
mosquito Anopheles gambiae. Pest. Biochem. Physiol. 47, 119–133
21 Ranson, H., Prapanthadara, L. and Hemingway, J. (1997) Cloning and
characterisation of two glutathione S-transferases from a DDT resistant strain of
Anopheles gambiae. Biochem. J. 324, 97–102
22 Mukabayire, O. and Besansky, N. J. (1996) Distribution of T1, Q, Pegasus and
mariner transposable elements on the polytene chromosomes of PEST, a standard
strain of Anopheles gambiae. Chromosoma 104, 585–595
23 Collins, F. H., Mendez, M. A., Rasmussen, M. O., Mehaffey, P. C., Besansky, N. J.
and Finnerty, V. (1987) A ribosomal RNA gene probe differentiates member species
of the Anopheles gambiae complex. Am. J. Trop. Med. Hyg. 37, 37–41
24 Kumar, V. and Collins, F. H. (1994) A technique for nucleic acid in situ hybridization
to polytene chromosomes of mosquitoes in the Anopheles gambiae complex. Insect
Mol. Biol. 3, 41–47
25 Salazar, C. E., Mills-Hamm, D., Kumar, V. and Collins, F. H. (1993) Sequence of a
cDNA from the mosquito Anopheles gambiae encoding a homologue of human
ribosomal protein S7. Nucleic Acids Res. 21, 4147
26 Paton, M. G., Karunaratne, S. H. P. P., Giakoumaki, E., Roberts, N. and
Hemingway, J. (2000) Quantitative analysis of gene amplification in insecticide-
resistant Culex mosquitoes. Biochem. J. 346, 17–24
27 Bradford, M. M. (1976) A rapid and sensitive method for the quantitation of
microgram quantities of protein utilizing the principle of protein–dye binding.
Anal. Biochem. 72, 248–254
28 Habig, W. H., Pabst, M. J. and Jakoby, W. B. (1974) Glutathione S-transferases: The
first enzymatic step in mercapturic acid formation J. Biol. Chem 249, 7130–7139
29 Thompson, J. D., Higgins, D. G. and Gibson, T. J. (1994) CLUSTAL W : improving the
sensitivity of progressive multiple sequence alignment through sequence weighting,
position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 22,
4673–4680
30 Rubin, G. M., Yandell, M. D., Wortman, J. R., Gabor Miklos, G. L., Nelson, C. R.,
Hariharan, I. K., Fortini, M. E., Li, P. W., Apweiler, R., Fleischmann, W. et al. (2000)
Comparative genomics of eukaryotes. Science. 287, 2204–2215
31 Jukes, T. H. and Cantor, C. R. (1969) Evolution of protein molecules. In Mammalian
Protein Metabolism III (Munro, H. N., ed.), pp. 21–132, Academic Press, New York
32 Saitou, N. and Nei, M. (1987) The neighbour-joining method: a new method for
constructing phylogenetic trees. Mol. Biol. Evol. 4, 406–425
33 van de Peer, Y. and De Wachter, R. (1994) TREECON for Windows: a software
package for the construction and drawing of evolutionary trees for the Microsoft
Windows environment. Comput. Appl. Biosci. 10, 569–570
# 2001 Biochemical Society
304 H. Ranson and others
34 Wilce, M. C. J., Board, P. G., Feil, S. C. and Parker, M. W. (1995) Crystal structure of
a theta-class glutathione transferase. EMBO J. 14, 2133–2143
35 Ranson, H., Cornel, A. J., Fournier, D., Vaughan, A. and Hemingway, J. (1997)
Cloning and localisation of a glutathione S-transferase class I gene from Anopheles
gambiae. J. Biol. Chem. 272, 5464–5468
36 Ranson, H., Jensen, B., Wang, X., Prapanthadara, L., Hemingway, J. and Collins,
F. H. (2000) Genetic mapping of two loci affecting DDT resistance in the malaria
vector Anopheles gambiae. Insect Mol. Biol. 9, 499–507
37 Singh, M., Silva, E., Schulze, S., Sinclair, D. A. R., Fitzpatrick, K. A. and Honda,
B. M. (2000) Cloning and characterization of a new theta-class glutathione-S-
transferase (GST) gene gst-3, from Drosophila melanogaster. Gene 247, 167–173
38 Huang, H.-S., Hu, N.-T., Yao, Y.-E., Wu, C.-Y., Chiang, S.-W. and Sun, C.-N. (1998)
Molecular cloning and heterologous expression of a glutathione S-transferase involved
in insecticide resistance from the diamondback moth, Plutella xylostella. Insect
Biochem. Mol. Biol. 28, 651–658
39 Snyder, M. J., Walding, J. K. and Feyereisen, R. (1995) Glutathione S-transferase
from larval Manduca sexta midgut : sequence of two cDNAs and enzyme induction.
Ins. Biochem. Mol. Biol. 25, 455–465
Received 30 May 2001/27 June 2001; accepted 9 August 2001
40 Singh, S. P., Coronella, J. A., Benes, H., Cochrane, B. J. and Zimniak, P. (2001)
Catalytic function of Drosophila melanogaster glutathione S-transferase DmGSTS1–1
(GST-2) in conjugation of lipid peroxidation end products. Eur. J. Biochem. 268,
2912–2923
41 Adams, M. D., Celniker, S. E., Holt, R. A., Evans, C. A., Gocayne, J. D., Amanatides,
P. G., Scherer, S. E., Li, P. W., Hoskins, R. A., Galle, R. F. et al. (2000) The genome
sequence of Drosophila melanogaster. Science 287, 2185–2195
42 Snyder, M. J. and Maddison, D. R. (1997) Molecular phylogeny of glutathione-
S-transferases. DNA Cell Biol. 16, 1373–1384
43 Lougarre, A., Bride, J. M. and Fournier, D. (1999) Is the insect glutatione
S-transferase I gene family intronless? Insect Mol. Biol. 8, 141–143
44 Syvanen, M., Zhou, Z. and Wang, J. (1994) Glutathione transferase gene family from
the housefly Musca domestica. Mol. Gen. Genet 245, 25–31
45 Tang, A. H. and Tu, C.-P. D. (1994) Biochemical characterization of Drosophila
glutathione S-transferases D1 and D21. J. Biol. Chem. 269, 27876–27884
46 Prapanthadara, L., Hemingway, J. and Ketterman, A. J. (1995) DDT-resistance in
Anopheles gambiae Giles from Zanzibar, Tanzania, based on increased
DDT-dehydrochlorinase activity of glutathione S-transferases. Bull. Entomol. Res.
85, 267–274
# 2001 Biochemical Society
... GSTs are involved in multiple biological functions including xenobiotic detoxification, clearance of oxidative stress products, protein transport, modulation of cell proliferation, and the induction of the apoptosis signaling pathway [1,5,6]. GSTs have been implicated to play important roles in insecticide resistance by metabolizing several classes of insecticides including organophosphates, carbamates, and organochlorines such as DDT [7][8][9][10][11][12]. GSTs also partially contribute to resistance to anticancer and anti-malarial drugs [13,14]. ...
Article
Full-text available
The availability of highly sensitive substrates is critical for the development of precise and rapid assays for detecting changes in glutathione S-transferase (GST) activity that are associated with GST-mediated metabolism of insecticides. In this study, six pyrethroid-like compounds were synthesized and characterized as substrates for insect and mammalian GSTs. All of the substrates were esters composed of the same alcohol moiety, 7-hydroxy-4-methylcoumarin, and acid moieties that structurally mimic some commonly used pyrethroid insecticides including cypermethrin and cyhalothrin. CpGSTD1, a recombinant Delta class GST from the mosquito Culex pipiens, metabolized our pyrethroid-like substrates with both chemical and geometric (i.e., the cis-isomers were metabolized at 2-to 5-fold higher rates than the corresponding trans-isomers) preference. A GST preparation from mouse liver also metabolized most of our pyrethroid-like substrates with both chemical and geometric preference but at 10-to 170-fold lower rates. CpGSTD1 and mouse GSTs metabolized CDNB, a general GST substrate, at more than 200-fold higher rates than our novel pyrethroid-like substrates. There was a 10-fold difference in the specificity constant (k cat /K M ratio) of CpGSTD1 for CDNB and those of CpGSTD1 for cis-DCVC and cis-TFMCVC suggesting that cis-DCVC and cis-TFMCVC may be useful for the detection of GST-based metabolism of pyrethroids in mosquitoes.
... It is notable for its detoxification and/or activation of xenobiotic and endogenous components [52]. GSTs is a dimeric protein with significant solubility, playing a crucial role in the metabolism, detoxification and excretion of a various kinds of endo-and exo-genous components [53][54][55]. The main characteristic of GSTs and P450s in insects is their upregulation of transcription, affecting the enhancement of the level of protein production and enzyme activation. ...
Article
Full-text available
This review discusses the application of the population genetic approach in elucidating the deployment of insecticide resistance to mosquito vectors. Although there have been a lot of scientific work describing population genetic research and insecticide resistance, a study focusing on the spread of insecticide resistance using the population genetic approach needs to be done. Population genetics explains how a population is diverse in response to fitness and the cost of environmental factors. Thus, readers can relate this process to how insecticides spread in the population. Additionally, some fundamental mechanisms of insecticide resistance are also covered. As successive reproduction of advantageous phenotypic traits, such as resistance depends on many factors including continuous pressure, recombination rate, migration rate, genetic drift, and so on. Currently, genome-wide association studies involve chromosome-wide SNPs in which recombination hotspots occur or microsatellite flanking region of resistance gene target in which the fixation process can potentially serve as a suitable marker for elucidating the deployment. The information provided in this review to facilitate how the susceptible individual still exists despite the predominance of resistant individuals and how the resistance reverts to the vulnerable state.
... GSTs can be classified into microsomal, mitochondrial and cytosolic GSTs according to their cellular localization (Sheehan et al., 2001). Interestingly, cytoplasmic GSTs in insects are subdivided into six distinct classes, namely delta, epsilon, omega, sigma, theta and zeta, among which delta and epsilon are unique to insect species (Ketterman et al., 2011;Ranson et al., 2001). The potential function of GSTs in insects is closely related to two specific sites on their subunits, that is, the G-site at the N-terminal and the H-site at the C-terminal (Enayati et al., 2005). ...
Article
Zeugodacus cucurbitae Coquillett (Diptera: Tephritidae) is an agriculturally and economically important pest worldwide that has developed resistance to β ‐cypermethrin. Glutathione S ‐transferases (GSTs) have been reported to be involved in the detoxification of insecticides in insects. We have found that both ZcGSTd6 and ZcGSTd10 were up‐regulated by β ‐cypermethrin induction in our previous study, so we aimed to explore their potential relationship with β ‐cypermethrin tolerance in this study. The heterologous expression of ZcGSTd6 and ZcGSTd10 in Escherichia coli showed significantly high activities against 1‐chloro‐2,4‐dinitrobenzene (CDNB). The kinetic parameters of ZcGSTd6 and ZcGSTd10 were determined by Lineweaver–Burk. The V max and K m of ZcGSTd6 were 0.50 μmol/min·mg and 0.3 mM, respectively. The V max and K m of ZcGSTd10 were 1.82 μmol/min·mg and 0.53 mM. The 3D modelling and molecular docking results revealed that β ‐cypermethrin exhibited a stronger bounding to the active site SER‐9 of ZcGSTd10. The sensitivity to β ‐cypermethrin was significantly increased by 18.73% and 27.21%, respectively, after the knockdown of ZcGSTd6 and ZcGSTd10 by using RNA interference. In addition, the inhibition of CDNB at 50% (IC 50 ) and the inhibition constants ( Ki ) of β ‐cypermethrin against ZcGSTd10 were determined as 0.41 and 0.33 mM, respectively. The Ki and IC 50 of β ‐cypermethrin against ZcSGTd6 were not analysed. These results suggested that ZcGSTd10 could be an essential regulator involved in the tolerance of Z. cucurbitae to β ‐cypermethrin.
... The largest family, cytosolic GSTs, is divided into six subclasses: delta, epsilon, omega, sigma, theta, and zeta (Enayati et al., 2005;Frova, 2006). Among them, epsilon and delta classes are insect-specific and are the major GSTs involved in xenobiotic detoxification (RANSON et al., 2001). The honey bee genome contains only 10 putatively functional GST genes and has only one delta class GST (AmGSTD1) and no epsilon GSTs (Claudianos et al., 2006). ...
... The largest family, the cytosolic GSTs, is divided into six subclasses: delta, epsilon, omega, sigma, theta, and zeta [75,77]. Among them, epsilon and delta classes are insect-specific and are the major GSTs involved in xenobiotic detoxification [78]. The honey bee genome contains only 10 putatively functional GST genes and has only one delta class GST (AmGSTD1) and no epsilon GSTs [61]. ...
Preprint
Full-text available
The European honey bee, Apis mellifera , serves as the principle managed pollinator species globally. In recent decades, honey bee populations have been facing serious health threats from combined biotic and abiotic stressors, including diseases, limited nutrition, and agrochemical exposure. Understanding the molecular mechanisms underlying xenobiotic adaptation of A. mellifera is critical, considering its extensive exposure to phytochemicals and agrochemicals present in flowers, propolis, hives, and the environment. In this study, we conducted a comprehensive structural and functional characterization of AmGSTD1, a delta class glutathione S-transferase (GST) enzyme, to unravel its roles in agrochemical detoxification and antioxidative stress responses. Significantly, we determined the 3D structure of a honey bee GST using protein crystallography for the first time, providing new insights into its molecular structure. Our investigations revealed that AmGSTD1 efficiently metabolizes model substrates, including 1-chloro-2,4-dinitrobenzene (CDNB), p-nitrophenyl acetate (PNA), phenylethyl isothiocyanate (PEITC), propyl isothiocyanate (PITC), and the oxidation byproduct 4-hydroxynonenal (4-HNE). Moreover, we discovered that AmGSTD1 exhibits binding affinity with the fluorophore 8-Anilinonaphthalene-1-sulfonic acid (ANS), which can be inhibited with various herbicides, fungicides, insecticides, and their metabolites. These findings highlight the potential contribution of AmGSTD1 in safeguarding honey bee health against various agrochemicals and their metabolites, while also mitigating oxidative stress resulting from exposure to these substances.
... GSTD1 has been implicated in resistance to DDT in both An. gambiae (Ranson et al. 2001) and D. melanogaster (Tang and Tu 1994). Two genes, CYP4H29 and GSTE6, were uniquely overexpressed in samples resistant to lambda-cyhalothrin. ...
Preprint
Full-text available
Aedes aegypti transmits major arboviruses of public health importance, including dengue, chikungunya, Zika, and yellow fever. The use of insecticides represents the cornerstone of vector control, however insecticide resistance in Ae. aegypti has become widespread. Understanding the molecular basis of insecticide resistance in this species is crucial to design effective resistance management strategies. Here, we applied Illumina RNA-Seq to study the gene expression patterns associated with resistance to three widely used insecticides (malathion, alpha-cypermethrin and lambda-cyhalothrin) in Ae. aegypti populations from 2 sites (Manatí and Isabela) in Puerto Rico (PR). Cytochrome P450s were the most over-expressed detoxification genes across all resistant phenotypes. Some detoxification genes (CYP6Z7, CYP28A5, CYP9J2, CYP6Z6, CYP6BB2, CYP6M9, and two CYP9F2 orthologs) were commonly over-expressed in mosquitoes that survived exposure to all three insecticides (independent of geographical origin) while others including CYP6BY1 (malathion), GSTD1 (alpha-cypermethrin), CYP4H29 and GSTE6 (lambda-cyhalothrin) were uniquely overexpressed in mosquitoes that survived exposure to specific insecticides. The gene ontology (GO) terms associated with monooxygenase, iron binding, and passive transmembrane transporter activities were significantly enriched in 4 out of 6 resistant vs susceptible comparisons while serine protease activity was elevated in all insecticide resistant groups relative to the susceptible strain. Interestingly, cuticular-related protein genes (chinase and chitin) were predominantly downregulated, which was also confirmed in the functional enrichment analysis. This RNA-Seq analysis presents a detailed picture of the candidate detoxification genes and other pathways that are potentially associated with pyrethroid and organophosphate resistance in Ae. aegypti populations from PR. These results could inform development of novel molecular tools for detection of resistance-associated gene expression in this important arbovirus vector and guide the design and implementation of resistance management strategies.
... ; https://doi.org/10.1101/2023.06.28.545844 doi: bioRxiv preprint oxidative stress (Pavlidi et al., 2018). Enhanced expression of GSTs has been shown to be a mechanism of resistance to DDT and organophosphates, and is also associated with resistance to pyrethroids in some insects (Huang et al., 1998;Lumjuan et al., 2005;Ranson et al., 2001). In our results, only one GST gene was up-regulated, and all other genes were down-regulated, which indicated the inhibitory effect of DA on GST in B. tabaci. ...
Preprint
Background: Bemisia tabaci is an important agricultural pest that has been causing significant economic losses to crops across the globe. Destruxins are secondary metabolites of entomopathogenic fungi which can be used as a potential biopesticide against B. tabaci. However, little is known about the molecular mechanism regulating the defense response of B. tabaci post destruxin application. Results: In this study, we explored the molecular responses of B. tabaci exposed to destruxin A (DA) using RNA-Seq and differentially expressed gene (DEG) analysis. A total of 1702, 616, and 555 DEGs were identified in B. tabaci after 4, 8, and 12 h of destruxin A treatment. In addition, 40 putative detoxification-related DEGs, including 29 cytochrome P450s (P450s), 5 glutathione S-transferases (GSTs), and 5 carboxylesterases (CarEs) were also identified. Quantitative real-time PCR analysis indicated that the expression profiles of 19 random DEGs were consistent with the RNA-Seq results. Conclusion: These findings serve as valuable information for a better understanding of the interaction and molecular mechanisms involved in the defense response of B. tabaci against DA.
Article
Full-text available
sp. Commiphora ‫ﺎت‬ ‫ﻟﻨﺒ‬ ‫ﻀﺮاء‬ ‫اﻟﺨ‬ ‫واﻷوراق‬ Ficusnitida ‫ﺔ‬ ‫ﺣﯿﻮﯾ‬ ‫ﻰ‬ ‫ﻋﻠ‬ ‫ﻮض‬ ‫ﺑﻌ‬ ‫ﺎت‬ ‫ﯾﺮﻗ‬ Culexquinquefasciatus ‫ﺮ‬ ‫اﻟﻤﺨﺘﺒ‬ ‫ﺮوف‬ ‫ﻇ‬ ‫ﺖ‬ ‫ﺗﺤ‬ ‫ﺪ‬ ‫ﺑﻌ‬ ‫ﺮاءات‬ ‫اﻟﻘ‬ ‫ﺬت‬ ‫وأﺧ‬ ٤٨ ‫ﺔ‬ ‫اﻟﻤﻌﺎﻣﻠ‬ ‫ﻦ‬ ‫ﻣ‬ ‫ﺎﻋﺔ‬ ‫ﺳ‬. ‫ﺼ‬ ‫ﻣﺴﺘﺨﻠ‬ ‫أن‬ ‫ﺎﺋﺞ‬ ‫اﻟﻨﺘ‬ ‫ﺤﺖ‬ ‫أوﺿ‬ ‫ﺎﺋﻲ‬ ‫اﻟﻤ‬ ‫ﺮ‬ ‫اﻟﻤ‬ ‫ﺎت‬ ‫ﻧﺒ‬ ‫ﻤﻎ‬ ‫ﺻ‬ ‫ﺎت‬ ‫ﯿﺘﻮﻧﻲ‬ ‫واﻻﺳ‬ ‫و‬ ‫ﻮض‬ ‫اﻟﺒﻌ‬ ‫ﺎت‬ ‫ﻟﯿﺮﻗ‬ ‫ﻮت‬ ‫ﻣ‬ ‫ﺴﺐ‬ ‫ﻧ‬ ‫ﺖ‬ ‫أﻋﻄ‬ ‫ﻲ‬ ‫اﻟﺒﻨﺰﯾﻨ‬ ‫و‬ ‫ﺎﻧﻮﻟﻲ‬ ‫اﻟﻤﯿﺜ‬ ٩٠ ‫و‬ ٩٠ ‫و‬ ٧١٫٥٧ ‫و‬ ٦٠ % ‫ﻰ‬ ‫ﻋﻠ‬ ‫اﻟﺘﺮﻛﯿﺰ‬ ‫ﻋﻨﺪ‬ ‫اﻟﺘﻮاﻟﻲ‬ ١٠ .% ‫ﺑﯿﻨﻤﺎ‬ ‫ﺎﺋﻲ‬ ‫اﻟﻤ‬ ‫ﺴﺘﺨﻠﺺ‬ ‫اﻟﻤ‬ ‫أﻋﻄﻰ‬ ‫ﺎت‬ ‫ﻧﺒ‬ ‫ﺼﻤﻎ‬ ‫ﻟ‬ ‫ﯿﺘﻮﻧﻲ‬ ‫واﻻﺳ‬ ‫ﺎﻧﻮﻟﻲ‬ ‫واﻟﻤﯿﺜ‬ ‫اﻟﺒﻌﻮض‬ ‫ﻟﯿﺮﻗﺎت‬ ‫ﻣﻮت‬ ‫ﻧﺴﺐ‬ ‫اﻟﻤﺮ‬ 71.57 ‫و‬ ٦٩٫٧٣ ‫و‬ ٦٨٫٨٧ % ‫اﻟﺘﻮا‬ ‫ﻋﻠﻰ‬ ‫ﺰ‬ ‫اﻟﺘﺮﻛﯿ‬ ‫ﻋﻨﺪ‬ ‫ﻟﻲ‬ ٥ .% ‫ـ‬ ‫ﻟ‬ ‫ﺖ‬ ‫اﻟﻤﻤﯿ‬ ‫ﺰ‬ ‫اﻟﺘﺮﻛﯿ‬ ‫ﺎن‬ ‫وﻛ‬ ٥٠ % ‫ﻮض‬ ‫اﻟﺒﻌ‬ ‫ﺎت‬ ‫ﯾﺮﻗ‬ ‫ﻦ‬ ‫ﻣ‬) (LC 50 ‫ـ‬ ‫ﻟ‬ ‫ﺖ‬ ‫اﻟﻤﻤﯿ‬ ‫ﺰ‬ ‫واﻟﺘﺮﻛﯿ‬ ٩٠ % ‫ﻦ‬ ‫ﻣ‬ ‫اﻟﺒﻌﻮض‬ ‫ﯾﺮﻗﺎت‬ (LC 90) ‫ﺔ‬ ‫اﻟﻤﺎﺋﯿ‬ ‫ﺮ‬ ‫اﻟﻤ‬ ‫ﻧﺒﺎت‬ ‫ﺻﻤﻎ‬ ‫ﻟﻤﺴﺘﺨﻠﺼﺎت‬ ‫ﺔ‬ ‫واﻟﻤﯿﺜﺎﻧﻮﻟﯿ‬ ‫ﺔ‬ ‫اﻟﺒﻨﺰﯾﻨﯿ‬ ‫و‬ ‫ﯿﺘﻮﻧﯿﺔ‬ ‫واﻻﺳ‬) ١٫٧٦ ‫و‬ ٤٫٣٣ (‫و‬) ١٫٩٧ ‫و‬ ١٣٫٥٩ (‫و‬) ١٫٦٣ ‫و‬ ٦٫٢٨ (‫و‬) ٥٫٥١ ‫و‬ ٤٦٫٢٧ % (‫ﻰ‬ ‫ﻋﻠ‬ ‫اﻟﺘﻮ‬ ‫اﻟﻲ‬. ‫ﺎت‬ ‫ﻟﻨﺒ‬ ‫ﻀﺮاء‬ ‫اﻟﺨ‬ ‫اﻷوراق‬ ‫ﺴﺘﺨﻠﺺ‬ ‫ﻣ‬ ‫أن‬ ‫ﺎﺋﺞ‬ ‫اﻟﻨﺘ‬ ‫أوﺿﺤﺖ‬ F. nitida ‫ﺴﺘﺨﺪم‬ ‫اﻟﻤ‬ ‫ﺎﻧﻮﻟﻲ‬ ‫اﻟﻤﯿﺜ‬ ‫ﻮض‬ ‫اﻟﺒﻌ‬ ‫ﺎت‬ ‫ﻟﯿﺮﻗ‬ ‫ﻮت‬ ‫ﻣ‬ ‫ﺴﺒﺔ‬ ‫ﻧ‬ ‫ﻰ‬ ‫أﻋﻠ‬ ‫ﻰ‬ ‫أﻋﻄ‬ ‫ﺘﺨﻼص‬ ‫اﻻﺳ‬ ‫ﺪ‬ ‫ﺑﻌ‬ ٥٠٫٠٧ % ‫ﺰ‬ ‫اﻟﺘﺮﻛﯿ‬ ‫ﺪ‬ ‫ﻋﻨ‬ ١٠ % ‫ﯿﻢ‬ ‫ﻗ‬ ‫ﺖ‬ ‫ﻛﺎﻧ‬ ‫و‬ LC 50 ‫و‬ LC 90 ‫ﻲ‬ ‫ﻫ‬ ٧٫٢٨ ‫و‬ ٢٨٫٢٦ % ‫ﻮاﻟﻲ‬ ‫اﻟﺘ‬ ‫ﻰ‬ ‫ﻋﻠ‬. ‫ﺴﺘﺨﻠﺺ‬ ‫اﻟﻤ‬ ‫أن‬ ‫ﺎﺋﺞ‬ ‫اﻟﻨﺘ‬ ‫ﺖ‬ ‫ﺑﯿﻨ‬ ‫ﻟﻨ‬ ‫ﻀﺮاء‬ ‫اﻟﺨ‬ ‫ﻸوراق‬ ‫ﻟ‬ ‫ﺎﺋﻲ‬ ‫اﻟﻤ‬ ‫ﺎت‬ ‫ﺒ‬ F.nitida ‫ﺪ‬ ‫ﺑﻌ‬ ‫ﺴﺘﺨﺪم‬ ‫واﻟﻤ‬ ٤٨ ‫ﺘﺨﻼص‬ ‫اﻻﺳ‬ ‫ﻦ‬ ‫ﻣ‬ ‫ﺎﻋﺔ‬ ‫ﺳ‬ ‫ﻰ‬ ‫إﻟ‬ ‫أدى‬ ‫ﺑﻨﺴﺒﺔ‬ ‫اﻟﺒﻌﻮض‬ ‫ﯾﺮﻗﺎت‬ ‫ﻣﻮت‬ ٧٢٫٤٤ % ‫ﻮت‬ ‫ﻣ‬ ‫ﺴﺒﺔ‬ ‫وﺑﻨ‬ ‫ﺎﻧﻮﻟﻲ‬ ‫اﻟﻤﯿﺜ‬ ‫اﻟﻤﺴﺘﺨﻠﺺ‬ ‫ﺗﻼه‬ ٥٧٫١٧ % ‫ﺪ‬ ‫ﻋﻨ‬ ‫ﺰات‬ ‫اﻟﺘﺮﻛﯿ‬ ١٠ % ‫ﯿﻢ‬ ‫ﻗ‬ ‫ﺖ‬ ‫وﻛﺎﻧ‬ LC 50 ‫و‬ LC 90 ‫ﺎﺋﻲ‬ ‫اﻟﻤ‬ ‫ﺴﺘﺨﻠﺺ‬ ‫ﻟﻠﻤ‬ ٤٫٠٤ ‫و‬ ١٧٫٤٩ % ‫ﻰ‬ ‫ﻋﻠ‬ ‫ﻮاﻟﻲ‬ ‫اﻟﺘ‬. ‫ﻗ‬ ‫ﻀﻲ‬ ‫ﻟﻠﻤ‬ ‫ﺪه‬ ‫واﻋ‬ ‫ﺎ‬ ‫ﻋﻠﯿﻬ‬ ‫ﺼﻞ‬ ‫اﻟﻤﺘﺤ‬ ‫ﺎﺋﺞ‬ ‫اﻟﻨﺘ‬ ‫ﺮ‬ ‫ﺗﻌﺘﺒ‬ ‫ﺄﺛﯿﺮ‬ ‫اﻟﺘ‬ ‫ذات‬ ‫ﺎت‬ ‫اﻟﻨﺒﺎﺗ‬ ‫ﺔ‬ ‫دراﺳ‬ ‫ﻲ‬ ‫ﻓ‬ ‫ﺪﻣﺎ‬ ‫ﺔ‬ ‫ﻟﻠﺒﯿﺌ‬ ‫ﺻﺪﯾﻘﺔ‬ ‫ﺑﺪاﺋﻞ‬ ‫ﺗﻤﺜﻞ‬ ‫واﻟﺘﻲ‬ ‫اﻵﻓﺎت‬ ‫ﻋﻠﻰ‬ ‫اﻟﺒﯿﻮﻟﻮﺟﻲ‬ ‫ﻦ‬ ‫ﻋ‬ ‫ﺪﻻ‬ ‫ﺑ‬ ‫ا‬ ‫ﺔ‬ ‫ﺑﺎﻟﺒﯿﺌ‬ ‫ﻀﺎرة‬ ‫اﻟ‬ ‫ﺔ‬ ‫اﻟﻜﯿﻤﯿﺎﺋﯿ‬ ‫ﺪات‬ ‫ﻟﻤﺒﯿ‬ ‫واﻹﻧﺴﺎن‬. ‫ﻣﻔﺘﺎﺣﯿﺔ‬ ‫ﻛﻠﻤﺎت‬ : ‫ﻣﺴﺘﺨﻠﺼﺎت‬ Commiphorasp. ‫و‬ Ficusnitida ، Culexquinquefasciatus. ‫اﻟﻤﻘﺪﻣﺔ‬ : ‫اﻷ‬ ‫ذات‬ ‫ﺸﺮات‬ ‫اﻟﺤ‬ ‫ﻦ‬ ‫ﻣ‬ ‫ﻮض‬ ‫اﻟﺒﻌ‬ ‫ﺸﺮات‬ ‫ﺣ‬ ‫ﺴﺎن‬ ‫اﻹﻧ‬ ‫ﺪاء‬ ‫أﻋ‬ ‫ﺪ‬ ‫أﻟ‬ ‫ﯿﻤﻦ‬ ‫وﻫ‬ ‫ﺔ‬ ‫واﻟﺒﯿﻄﺮﯾ‬ ‫ﺔ‬ ‫اﻟﻄﺒﯿ‬ ‫ﺔ‬ ‫ﻫﻤﯿ‬ ‫واﻟﺤﯿﻮان‬. ‫ﻋﺎﺋﻠﺔ‬ ‫اﻟﺒﻌﻮض‬ ‫ﺣﺸﺮات‬ ‫ﺗﺘﺒﻊ‬ Culicidae ‫ﺔ‬ ‫اﻷﺟﻨﺤ‬ ‫ﺔ‬ ‫ﺛﻨﺎﺋﯿ‬ ‫ﺔ‬ ‫ﻟﺮﺗﺒ‬ ‫اﻟﺘﺎﺑﻌﺔ‬ Diptera ‫ﺸﺮ‬ ‫وﺗﻨﺘ‬ ‫ﺎﻟﻢ‬ ‫اﻟﻌ‬ ‫ﺎء‬ ‫أﻧﺤ‬ ‫ﻒ‬ ‫ﻣﺨﺘﻠ‬ ‫ﻲ‬ ‫ﻓ‬ ‫ﺔ‬ ‫ﻣﺘﺒﺎﯾﻨ‬ ‫ﺔ‬ ‫ﺣﺮارﯾ‬ ‫ﺎﻃﻖ‬ ‫وﻣﻨ‬ ‫ﺎت‬ ‫ﺑﯿﺌ‬ ‫ﻲ‬ ‫ﻓ‬ ‫ﺎ‬ ‫أﻧﻮاﻋﻬ‬) Okogun et al. 2003 .(‫وﺗﻨ‬ ‫اﻟﺤﻤ‬ ‫ﻣﻨﻬﺎ‬ ‫اﻷﻣﺮاض‬ ‫ﻣﻦ‬ ‫اﻟﻌﺪﯾﺪ‬ ‫اﻟﺒﻌﻮض‬ ‫ﻘﻞ‬ ‫ﺪﻣﺎغ‬ ‫اﻟ‬ ‫ﺎب‬ ‫واﻟﺘﻬ‬ ‫ﺼﺪع‬ ‫اﻟﻤﺘ‬ ‫اﻟﻮادي‬ ‫وﺣﻤﻰ‬ ‫اﻟﺼﻔﺮاء‬ ‫ﻰ‬ ‫ﺎ‬ ‫واﻟﻔﻼرﯾ‬ ‫واﻟﻤﻼرﯾﺎ‬ ‫اﻟﯿﺎﺑﺎﻧﻲ‬ (Gratz and Jany1994). ‫ﺔ‬ ‫اﻟﻤﻨﻘﻮﻟ‬ ‫ﺮاض‬ ‫اﻷﻣ‬ ‫أن‬ ‫ﻦ‬ ‫ﻣ‬ ‫ﺎﻟﺮﻏﻢ‬ ‫ﺑ‬ ‫ﻰ‬ ‫وﻋﻠ‬ ‫ﯿﺲ‬ ‫ﻟ‬ ‫اﻧﻪ‬ ‫اﻻ‬ ‫اﻻﺳﺘﻮاﺋﻲ‬ ‫وﺷﺒﺔ‬ ‫اﻻﺳﺘﻮاﺋﻲ‬ ‫اﻟﻤﻨﺎخ‬ ‫ﻓﻲ‬ ‫ﺻﺤﯿﺔ‬ ‫ﻣﺸﻜﻠﺔ‬ ‫ﺗﻤﺜﻞ‬ ‫اﻟﺤﺎﺿﺮ‬ ‫اﻟﻮﻗﺖ‬ ‫ﻓﻲ‬ ‫ﺑﺎﻟﺒﻌﻮض‬ ‫ﺮا‬ ‫اﻷﻣ‬ ‫ﺮ‬ ‫ﺧﻄ‬ ‫ﻦ‬ ‫ﻋ‬ ‫ﺄى‬ ‫ﺑﻤﻨ‬ ‫ﺎﻟﻢ‬ ‫اﻟﻌ‬ ‫ﻲ‬ ‫ﻓ‬ ‫ﺰء‬ ‫ﺟ‬ ‫ﺎك‬ ‫ﻫﻨ‬ ‫ﻮض‬ ‫اﻟﺒﻌ‬ ‫ﻦ‬ ‫ﻋ‬ ‫ﺌﺔ‬ ‫اﻟﻨﺎﺷ‬ ‫ض‬ (Fardin and Day2002). ‫ﻛﻠﻤﺔ‬ ‫اﺷﺘﻘﺖ‬ Phytochemicals ‫ﺎت‬ ‫ﻟﻠﯿﺮﻗ‬ ‫ﺪات‬ ‫ﻛﻤﺒﯿ‬ ‫ﺗﻌﻤﻞ‬ ‫أن‬ ‫ﯾﻤﻜﻦ‬ ‫اﻟﺘﻲ‬ ‫اﻟﻨﺒﺎﺗﯿﺔ‬ ‫اﻟﻤﺼﺎدر‬ ‫ﻣﻦ‬ ‫ا‬ ‫ﺔ‬ ‫وﻣﺎﻧﻌ‬ ‫ﺸﺮات‬ ‫ﻟﻠﺤ‬ ‫ﺎردة‬ ‫ﻃ‬ ‫ﻣﻮاد‬ ‫او‬ ‫اﻟﺤﺸﺮات‬ ‫ﻟﻨﻤﻮ‬ ‫ﻣﺎﻧﻌﺎت‬ ‫و‬ ‫ﯿﺾ‬ ‫اﻟﺒ‬ ‫ﻊ‬ ‫ﻟﻮﺿ‬ ‫ﺸﺮات‬ ‫ﻟﻠﺤ‬ ‫ﺎت‬ ‫ﺟﺎذﺑ‬ ‫و‬ ‫و‬ ‫اﻧ‬ ‫ﻣﻦ‬ ‫اﻟﺤﺪ‬ ‫ﻓﻲ‬ ‫ﻣﻬﻤﺎ‬ ‫دورا‬ ‫ﺗﻠﻌﺐ‬ ‫أن‬ ‫ﯾﻤﻜﻦ‬ ‫اﻟﻤﺴﺘﺨﻠﺼﺎت‬ ‫ﺗﻠﻚ‬ ‫ﺎﻟﺒﻌﻮض‬ ‫ﺑ‬ ‫اﻷﻣﺮاض‬ ‫ﺘﻘﺎل‬) Babu and Maruugan 1998; Venketachalam and Jebasan 2001 a&b; Mittal and Subbarao 2003; Bagavan et al. 2008; Ghosh et al. 2008; Kannathasan et al.
Article
Micromelalopha troglodyta (Graeser) is an important pest of poplar in China, and glutathione S -transferase (GST) is an important detoxifying enzyme in M. troglodyta . In this paper, three full-length GST genes from M. troglodyta were cloned and identified. These GST genes all belonged to the epsilon class ( MtGSTe1 , MtGSTe2 , and MtGSTe3 ). Furthermore, the expression of these three MtGSTe genes in different tissues, including midguts and fat bodies, and the MtGSTe expression in association with different concentrations of tannic acid, including 0.001, 0.01, 0.1, 1, and 10 mg ml ⁻¹ , were analysed in detail. The results showed that the expression levels of MtGSTe1 , MtGSTe2 , and MtGSTe3 were all the highest in the fourth instar larvae; the expression levels of MtGSTe1 and MtGSTe3 were the highest in fat bodies, while the expression level of MtGSTe2 was the highest in midguts. Furthermore, the expression of MtGSTe mRNA was induced by tannic acid in M. troglodyta . These studies were helpful to clarify the interaction between plant secondary substances and herbivorous insects at a deep level and provided a theoretical foundation for controlling M. troglodyta .
Article
Full-text available
The fly Drosophila melanogaster is one of the most intensively studied organisms in biology and serves as a model system for the investigation of many developmental and cellular processes common to higher eukaryotes, including humans. We have determined the nucleotide sequence of nearly all of the ∼120-megabase euchromatic portion of theDrosophila genome using a whole-genome shotgun sequencing strategy supported by extensive clone-based sequence and a high-quality bacterial artificial chromosome physical map. Efforts are under way to close the remaining gaps; however, the sequence is of sufficient accuracy and contiguity to be declared substantially complete and to support an initial analysis of genome structure and preliminary gene annotation and interpretation. The genome encodes ∼13,600 genes, somewhat fewer than the smaller Caenorhabditis elegansgenome, but with comparable functional diversity.
Article
Full-text available
Two classes of glutathione transferases have been identified and purified from Musca domestica. The first, designated as GST1, migrates as a single band of 28 kDa in SDS-gel electrophoresis, and the second, designated as GST2, migrates as a 32-kDa band. Antisera prepared against each class have no immunological cross-reactivity, and heterodimeric associations between the two classes have not been detected. Each class is composed of several isoforms: GST1 is composed of forms with isoelectric points from 4 to 9, whereas all the forms of GST2 have acidic pI values. Screening of cDNA libraries yielded clones coding for GST1, and the gene was sequenced and expressed in Escherichia coli. The high activity found in an insecticide-resistant strain (Cornell R) is correlated with high level of GST1 transcript.
Article
A protein determination method which involves the binding of Coomassie Brilliant Blue G-250 to protein is described. The binding of the dye to protein causes a shift in the absorption maximum of the dye from 465 to 595 nm, and it is the increase in absorption at 595 nm which is monitored. This assay is very reproducible and rapid with the dye binding process virtually complete in approximately 2 min with good color stability for 1 hr. There is little or no interference from cations such as sodium or potassium nor from carbohydrates such as sucrose. A small amount of color is developed in the presence of strongly alkaline buffering agents, but the assay may be run accurately by the use of proper buffer controls. The only components found to give excessive interfering color in the assay are relatively large amounts of detergents such as sodium dodecyl sulfate, Triton X-100, and commercial glassware detergents. Interference by small amounts of detergent may be eliminated by the use of proper controls.
Article
A comparative analysis of the genomes ofDrosophila melanogaster, Caenorhabditis elegans, and Saccharomyces cerevisiae—and the proteins they are predicted to encode—was undertaken in the context of cellular, developmental, and evolutionary processes. The nonredundant protein sets of flies and worms are similar in size and are only twice that of yeast, but different gene families are expanded in each genome, and the multidomain proteins and signaling pathways of the fly and worm are far more complex than those of yeast. The fly has orthologs to 177 of the 289 human disease genes examined and provides the foundation for rapid analysis of some of the basic processes involved in human disease.
Article
DDT-resistant Anopheles gambiae Giles from Zanzibar, Tanzania, had increased levels of DDT-dehydrochlorination compared to a DDT-susceptible strain. Glutathione S-transferases (GSTs) are responsible for conversion of DDT to DDE in both the susceptible and resistant strains. Sequential column chromatography, including Q-Sepharose, S-hexylglutathione agarose, hydroxylapatite and phenyl Sepharose, allowed the partial purification of seven GSTs. All seven GSTs possessed different degrees of DDTase activity. There was an eight-fold increase in total DDTase activity in the resistant compared to the susceptible enzymes. Characterization with three substrates, 1-chloro-2,4-dinitrobenzene (CDNB), 1,2-dichloro-4-nitrobenzene (DCNB) and DDT, revealed the different substrate specificity for each isolated GST indicating different isoenzymes. GST Va possessed 60% of total DDTase activity suggesting that it contributed most to DDT-metabolism in this insect species. The DDTase activity of the GSTs in both strains of A. gambiae were found to be correlated with the GST activities toward DCNB. Preliminary studies on DDT-resistant and susceptible A. gambiae showed that both DDT-resistance and the increased levels of GST activity were stage specific which suggested that different GSTs may be involved in DDT-resistance in adults and larvae of A. gambiae.
Article
In this study we characterize the involvement of glutathione S-transferase in DDT resistance of the malaria vector, Anopheles gambiae. A comparative study of the seven glutathione S-transferases isolated from a 20-fold DDT-resistant strain and a DDT-susceptible strain of the mosquito was performed using four inhibitors and six substrates. We report that in A. gambiae DDT resistance was apparently conferred by qualitatively different glutathione S-transferases that also appear to be present in higher concentrations. The seven glutathione S-transferases did not immunologically cross-react with Alpha, Mu, or Pi antisera or antiserum against GST-1 from house fly and Drosophila. The results suggest significant differences between the vertebrate and the insect glutathione S-transferases.
Article
A protein determination method which involves the binding of Coomassie Brilliant Blue G-250 to protein is described. The binding of the dye to protein causes a shift in the absorption maximum of the dye from 465 to 595 nm, and it is the increase in absorption at 595 nm which is monitored. This assay is very reproducible and rapid with the dye binding process virtually complete in approximately 2 min with good color stability for 1 hr. There is little or no interference from cations such as sodium or potassium nor from carbohydrates such as sucrose. A small amount of color is developed in the presence of strongly alkaline buffering agents, but the assay may be run accurately by the use of proper buffer controls. The only components found to give excessive interfering color in the assay are relatively large amounts of detergents such as sodium dodecyl sulfate, Triton X-100, and commercial glassware detergents. Interference by small amounts of detergent may be eliminated by the use of proper controls.
Article
We have isolated a Drosophila gene, DmGST-2, that encodes glutathione S-transferase, a homo- or heterodimeric enzyme thought to be involved in detoxification of xenobiotics, including known carcinogens. The encoded protein has a primary sequence that is more similar to mammalian placental and nematode GSTs than that of a previously described Drosophila GST gene, herein referred to as DmGST-1. We provide a physical map of the gene and show that it specifies at least two mRNAs, measuring 1.9 and 1.6 kb, which differ only in the lengths of their 3' untranslated regions. Both of the mRNAs are present during all developmental stages. In situ hybridization of the DmGST-2 gene to larval polytene chromosomes places it within the 53F subdivision of chromosome 2, and Southern blotting to chromosomal DNA indicates that the gene has no close relatives within the Drosophila genome. Our results make possible molecular genetic approaches for further elaborating the function of glutathione S-transferases in insect development and physiology, in the metabolism of plant toxins, and in conferring insecticide resistance.
Article
The amount of glutathione S-transferase-2 (GST-2) protein and enzyme activity in a mutant strain (strain GG) of the yellow fever mosquito (Aedes aegypti) is approximately 25-fold higher than in the wild-type (++) strain. The mode of inheritance of the GG phenotype was studied in F1 and backcross progeny using GST enzyme assays, isozyme-specific antisera, and Northern blot analysis. Enzyme assay of parental and F1 progeny showed that the ++ phenotype was dominant to the GG phenotype. This was true for larvae as well as for all tissues examined in adults in both sexes. Immunoblotting experiments showed that, like the ++ strain, F1 larvae and adults express very low levels of GST-2 protein compared with the GG strain. Northern blotting experiments showed that the steady-state levels of GST-2 mRNA in parental and F1 hybrid larvae closely matched the enzyme activity and immunological data. These results suggest the existence of a trans-acting regulatory locus that acts to repress GST-2 mRNA transcription and/or decrease GST-2 mRNA stability in ++ and F1 hybrids. GST enzyme activity in backcross progeny, however, did not segregate into the two distinct phenotypes (low and high) predicted for a single locus, dominant allele model. Backcross progeny expressed a wide range of GST activity and GST-2 protein amount with no apparent fit to simple Mendelian ratios. These backcross data suggest that additional loci are also involved in regulating GST-2 isozyme expression.(ABSTRACT TRUNCATED AT 250 WORDS)
Article
Glutathione transferases (GSTs) of a novel class, which it is proposed to term Theta, were purified from rat and human liver. Two, named GST 5-5 and GST 12-12, were obtained from the rat, and one, named GST theta, was from the human. Unlike other mammalian GSTs they lack activity towards 1-chloro-2,4-dinitrobenzene and are not retained by GSH affinity matrices. Only GST 5-5 retains full activity during purification, and its activities towards the substrates 1,2-epoxy-3-(p-nitrophenoxy)propane, p-nitrobenzyl chloride, p-nitrophenethyl bromide, cumene hydroperoxide, dichloromethane and DNA hydroperoxide are 185, 86, 67, 42, 11 and 0.03 mumol/min per mg of protein respectively. Earlier preparations of GST 5-5 or GST E were probably a mixture of GST 5-5 and GST 12-12, which was largely inactive, and may also have been contaminated by less than 1% with another GSH peroxidase of far greater activity. Partial analysis of primary structure shows that subunits 5, 12 and theta are related to each other, particularly at the N-terminus, where 25 of 27 residues are identical, but have little relationship to the Alpha, Mu and Pi classes of mammalian GSTs. They do, however, show some relatedness to subunit I of Drosophila melanogaster [Toung, Hsieh & Tu (1990) Proc. Natl. Acad. Sci. U.S.A. 87, 31-35] and the dichloromethane dehalogenase of Methylobacterium DM4 [La Roche & Leisinger (1990) J. Bacteriol, 172, 164-171].