Research Article

Xenobiotic Induced Expression of Epsilon Glutathione S – Transferases Gste2 In Laboratory Strains of Anopheles Arabiensis  

A.M. Yayo1,2 , A. Ado1 , M. Safiyanu3 , B.R. Muhammad4 , F.I. Sambo5 , A. Abubakar2 , J. Hemingway6
1 Centre for Infectious Diseases Research, Bayero University, Kano, Nigeria
2 Dept. of Medical Parasitology and Microbiology, Bayero University, Kano, Nigeria
3 Dept of Biochemistry, Yusufu Maitama Sule, University Kano, Nigeria
4 Dept of Veterinary Parasitology and Entomology, University of Abuja, Abuja, Nigeria
5 Dept of Biological Sciences, Yusufu Maitama Sule, University Kano, Kano, Nigeria
6 Vector Biology Research Group, Liverpool School of Tropical Medicine, England
Author    Correspondence author
Molecular Entomology, 2019, Vol. 10, No. 1   doi: 10.5376/me.2019.10.0001
Received: 13 Nov., 2019    Accepted: 30 Nov., 2019    Published: 20 Dec., 2019
© 2019 BioPublisher Publishing Platform
This is an open access article published under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Preferred citation for this article:

Yayo A.M., Ado A., Safiyanu M., Muhammad B.R., Sambo F.I., Abubakar A., and Hemingway J., 2019, Xenobiotic induced expression of epsilon glutathione S–Transferases Gste2 in laboratory strains of Anopheles arabiensis, Molecular Entomology, 10(1): 1-9 (doi: 10.5376/me.2019.10.0001)

Abstract

Over expression of GSTe2 has been associated with resistance to DDT in An. arabiensis. The expression of the gene was induced by exposing the fourth instar larvae obtained from KGB and MAT strains to xenobiotic inducers. Transcripts levels of GSTe2 in cDNAs of the larvae from each strain was measured by quantitative PCR at various time points post exposure to DDT, permethrin and hydrogen peroxide. Statistical analysis was used to test for differences and compare the expression of the gene between treatments and mosquito strains. The geometric mean expression levels of GSTe2 induced at one hour by DDT in the KGB was 1.028 (CI 0.901~1.172; p < 100) compared to 0.587 (CI 0.475~0.724; p = 0.025) in MAT strain. Higher induction of GSTe2 in KGB was caused by DDT and hydrogen peroxide and with permethrin and DDT in MAT strain. Further research is needed to understand the molecular mechanisms regulating response of An. arabiensis to xenobitics.

Keywords
Xenobiotic; Expression; Glutathione S-Transferases Gste2; Anopheles arabiensis

Introduction                                                                               

Induction refers to the process in which a chemical stimulus enhances the activity of a detoxification system by production of additional enzymes (Tieriere, 1984). Experiments on induction often involves exposing animals reared in the laboratory to xenobiotic inducers and using biochemical or molecular procedures to demonstrate evidence for involvement of a particular detoxification pathway (Ding et al., 2005; Lumjuan et al., 2011; Wen-Wu zhou et al., 2011; Bamigdele et al., 2017). Three large enzyme families, monooxygenases, carboxylesterases and glutathione s –transferases have been associated with detoxification of xenobiotic compounds in biological systems (Hemingway et al., 2004; Ranson et al., 2011; Mashimaya et al., 2014). In insects, these enzyme families are important in metabolism of insecticides and constitute the basis for metabolic resistance mechanisms (Ranson and Hemingway, 2000; Che-mendoza et al., 2009). The insect specific delta and epsilon glutathione s – transferases in particular have been implicated in detoxification of organophosphates (Chiang and Sun, 1993; Huang et al., 1998), pyrethroids (Vontas et al., 2001; Rakotondranaivo et al., 2018) and organochlorines (Prapanthadra et al., 1995; Ranson et al., 2001) insecticides. Evidence for the role of GSTe2 in resistance to permethrin and DDT has been shown in various species of Anopheline mosquitoes including, An. cracens (Wongtrakul et al., 2010), An. funestus (Gregory et al., 2011), An. gambiae (Ortelli et al., 2003; Patience et al., 2016) and An. arabiensis (Yayo et al., 2018a). Also, some GSTs from the Sigma, Delta and Epsilon classes including GSTe2 have been shown to possess peroxidase activity and protect against oxidative stress due to xenobiotic compounds (Singh et al., 2001; Vontas et al., 2002; Giardano et al., 2007; Lumjuan et al., 2011).

 

Exposure to xenobiotic compounds either in the environment or laboratory have caused overexpression of GST genes in various species of mosquitoes including Aedes aegypti, An. stephensi, and An. gambiae (Ding et al., 2005; David et al., 2010; Gregory et al., 2011). Aedes aegypti larvae exposed to the herbicides glyphosate, the carcinogene benzo(a)pyrene and the synthetic insecticides propoxur and permethrin were found to induce differentially four GSTs including GSTx (David et al., 2010). A two – fold change in expression of GSTe2 was reported in the pupae of pyrethroid resistant An. stephensi compared to the susceptible laboratory strains (Gregory et al., 2011). Ding et al. (2005) found significant increases in transcripts of GSTe1 and GSTe2 in larvae of An. gambiae exposed to H2O2.. Fang (2012), has reviewed in detail studies on inducible GSTs.

 

An. arabiensis is the most closely related ecologically and taxonomically to An. gambiae s.s but is genetically distinct and tends to be more adapted to human environment habitually containing xenobiotic contaminants (Muller et al., 2007; Matowo et al., 2014; Cassone et al., 2014). An. arabiensis occurs sympatrically with An. gambiae and is one of the dominant malaria vectors in tropical Africa (Coetzee et al., 2000; Coetzee et al., 2013). Previously, we isolated and characterized eight Epsilon GSTe genes from An. arabiensis MAT strain and demonstrated differential expression of three of GSTe genes in developmental stages. We reported overexpression of the GSTe1 and GSTe2 in DDT selected lines compared to susceptible strains (Yayo et al., 2018a;b). To investigate the inducible expression of GSTe2, we exposed the fourth instar larvae of two laboratory strains of An. arabiensis to sub-lethal concentrations of H2O2, Permethrin and DDT and monitored the expression of the gene by quantitative PCR.

 

1 Materials and Methods

1.1 Mosquito strains and chemicals

Mosquitoes from the parental lines of MAT and KGB laboratory strains of An. arabiensis (Patton) were used in the induction experiments. The parental lines of these strains were selected for the induction studies because both strains have been maintained in the insectary for a least 3 (MAT) and 2 (KGB) years without any insecticidal selection pressure.  When tested for susceptibility to DDT, the adult mosquitoes from the KGB parental line were more tolerant to DDT compared to the MAT strain (Yayo et al., 2018a). However, susceptibility of the larvae of these mosquitoes to permethrin and DDT was not known. The choice of the chemicals for the induction work was based on the need to simulate the type of xenobiotic contaminants in the mosquito’s natural breeding habitats (Muller et al., 2007; Cassone et al., 2014; Kamdem et al., 2014). Hence, H2O2, permethrin and DDT were chosen as representative xenobiotic compounds which An. arabiensis may typically encounter in the environment. Previously, sub-lethal doses of 3mM H2O2 (Ding et al., 2005) and 0.02 mg/L of permethrin (David JP, personal communication) were used to treat larvae from ZAN/U and RSP strains of An. gambiae, therefore these dosages were adopted in this experiment.  In order to determine the sub-lethal dose of DDT, 3 batches each of 25 larvae were exposed, each for one hour, to the following concentrations of DDT: 0.5 mg/L, 0.05 mg/L and 0.005 mg/L. Larval mortalities were determined after 24 hours holding.  The dose 0.05 mg/L, which gave a mortality of >20%, was adopted and used for subsequent induction.

 

1.2 Recruitment of larvae for induction

Rearing activities were harmonised to minimise any compounding effect, possibly due to excess food or overcrowding. Egg batches laid by the adult females were collected from filter paper placed in the egg-cup. The filter paper containing eggs was carefully cut to split the egg-batch into two halves. Each half paper was placed in an egg pot containing 10~20 mL of water to enable the eggs to hatch to 1st instar larvae. About 25~30 1st instar larvae were transferred to shallow trays measuring 28×20 cm with a depth of 6.5 cm, which were only half filled with distilled water. All larvae from the two strains were fed on Tetramin fish food flakes, using grinded flakes for the first instars. Sufficient food was added each time without causing the water to go scummy. Distilled water was squirted into the larval trays on a daily basis to aerate the water and excess food removed using a pipette. The water was changed if there were any signs of cloudiness, along with sluggish swimming behaviour of the larvae. Extreme precaution was taken to avoid contamination between the strains and lines of the An. arabiensis. Pipettes and egg pots were colour-coded for each strain. This scrupulous method of larval rearing was used to generate all the 4th instar larvae which were used for the induction experiments.

 

1.3 Exposure of 4th instar larvae to H2O2, Permethrin and DDT

A total of six hundred 4th instar larvae each taken from MAT and KGB strains were exposed to H2O2, permethin and DDT in a series of separate experiments. In each experiment, a group of 40~50 larvae were immersed for one hour in either 3 mM H2O2, 0.02 mg/L permethrin or 0.05 mg/L DDT. The larvae were then transferred to sterile water to recover. Unexposed larvae of the same life-stage from both strains were used as control. Larval samples in batches of 10 were collected in triplicate at one, two and 24 hour time points post exposure.

 

1.4 Extraction of RNA and cDNA synthesis

RNA extraction and cDNA synthesis were performed as previously described (Yayo et al., 2018a). Total RNA was extracted from each batch of the larval samples obtained after exposure at each of the time points mentioned above. The samples were homogenized in 400 µL of TRI reagent (sigma) and centrifuged at 12,000 g for 10 min at 4°C. The supernatant was transferred to a DEPC- treated centrifuge tube, homogenized for 15 secs with chloroform and was then incubated at 4°C after which the upper aqueous phase containing RNA was transferred to a fresh tube. The RNA was precipitated by adding 200 µL of molecular grade isopropanol and incubated at room temperature for 10 min. After centrifugation at 12,000 g for 10 min at 4°C, the pellet was air- dried and re-suspended in 26 µL of nuclease-free water. The RNA was treated with RQI DNase (Promega) to remove contaminating genomic DNA. cDNA was synthesized from total RNA using superscript TM111RNase H- reverse transcriptase (Invitrogen) and an oligo (dT) adapter primer (5´-GACTCGAGTCGACATCGA (dT) 17 – 3´). One microgram of RNA was mixed with 1 µg of the primer and heated to 65°C for five minutes to dissociate the secondary conformation. The reaction mixture was first chilled on ice for 1 minute and then pre-warmed to 50°C for 2 minutes after addition of 8 µL of 5X first strand buffer (250 mM Tris-HC1 pH 8.3, 375 mM KCL, 15 mM MgCl2), 2 µL of 10 mM of each dNTP and 2 µL of 0.1 M DTT. One half microlitre of Superscript TM111Rnase H- reverse transcriptase was added and the reaction was incubated at 50°C for 90 minutes. The reaction was then heated to 70oc for 15 minutes to inactivate the enzyme. First strand cDNA samples were stored at -20°C.

 

1.5 Plasmid construction and qPCR

Primer sequences specific for An. arabiensis GSTe2 described previously (Yayo et al., 2018b) were used to amplify the full-length cDNA from the KGB and MAT strains. PCR reactions contained 1.5 mM MgCl2, 0.2 mM of each dNTPs, and 0.5 µm of each primer, 1 X reaction buffer (Qiagen) and 1.25 units of Taq DNA polymerase (Hotstar) in a final volume of 25 µL. The cycling parameters were 95°C for 15 minutes, followed by 30 cycles of 95°C for 30 seconds, 50°C for 30 seconds and 72°C for 30 seconds and a final extension of 72°C for 10 minutes. The amplification was conducted using PCR machine PTC-200 (MJ Research). GSTe2 standard plasmid was constructed by insertion PCR products of cDNA from the MAT and KGB strains into the PGEM-T Easy Vector (Promega) according to manufacture’s instructions. The ribosomal protein gene SP7 (accession no. AY 380336) (Salazar et al., 1993), used as an internal control plasmid, was amplified using forward GCA CGT CGT GTT CAT TGC CG  and reverse GAA CAT TAA CGT CAC GGC CAG TCA primers and ligated into the PGEM-T Easy Vector. The concentrations of the plasmids were determined using an ND- 1000 spectrophotometer (Nanodrop Technologies).

 

1.6 Quantitative PCR

The GSTe2 and SP7 standard plasmids were diluted serially to concentrations ranging from 1 ng/µL to 1 fg/µL. To compare the abundance of GSTe2 transcript in the larvae exposed to each of the three xenobiotic inducers and control samples from the MAT and KGB strains, quantitative real time PCR was performed. In each experiment, duplicate reactions were set up for each of the seven standard plasmids and the cDNA samples. Quantitative PCR was performed in 15 µL reactions with a quantitative PCR Kit (Qiagen), which composed of 1 µL of DNA template, 1 × Master Mix, 0.2 mM of each primer. Details of the primer sequences are given in Table 1. The PCR conditions were 95°C for 15min, followed by 35 cycles of 94°C for 20 s, 54°C for 25 s and 72°C for 30 s with the fluorescence read at 82°C. Plasmid DNA standards and negative controls were included in the same plate, for each experiment. Three biological replicate samples from each strain and batches of samples of larvae exposed to each of the xenobiotic were used as templates. A sample was analysed in duplicate in each experiment and results were averaged from three separate experiments. The number of copies of GSTe2 and SP7 was calculated by measuring the incorporation of the SYBR Green 1 into the PCR products using the DNA Engine Opticon (MJ Research). The fluorescence was plotted in relation to the number of cycles and the crossing line was produced to obtain the standard concentration related to the cycle number. A straight line standard curve was obtained by plotting the cycle number against the logarithmic value of the standard concentration. The ribosomal gene SP7 was used to normalize for variation in total cDNA concentration.

 


Table 1 Oligo nucleotide Primer Sequences used in quantitative real time PCR

 

The normalised copy number of GSTe2 was thus determined for each batch of larval samples obtained at one, two and twenty-four hours post exposure to DDT, Permethrin and H2O2 in the separate experiments. One-way analysis of variance was used initially to test for differences between treatments and the mosquito strains. Significant results were in investigated in detail using the Turkey significant difference multiple comparison test (Winter, 1962). Statistical significance was set at the conventional 5% level for all tests.

 

2 Results

2.1 Inducible expression of Epsilon GSTe2 by DDT, permethrin and hydrogen peroxide.

2.1.1 DDT Induced expression

The geometric means levels of GSTe2 expressed at 1, 2 and 24 hours post-exposure to 0.05mg/l DDT for one hour and control are shown individually in Figure 1. At 1 hour post-exposure to DDT, the mean expression of GSTe2 was raised relative to the control larvae by just over one half (57%) MAT strain (p = 0.025) but was more than doubled (108% increase) in the KGB strain (p < 0.001). At 2-hours, the mean GSTe2 expression was now 60% lower than in the control larvae in the MAT strain (p < 0.001); in the KGB strain, however the mean gene expression remained numerically higher than control (by 29% i.e. by just over one quarter), this difference was not statistically significant (p = 0.146). At 24 hours , mean gene expression was 79% lower than in the control larvae in the MAT strain (p < 0.001); in the KGB strain, mean gene expression was lower than control by just 16% - this difference was not statistically significant (p = 0.445).

 


Figure 1 Epsilon GSTe2 expression in control and larvae exposed to 0.05mg/L DDT for one hour – geometric means and 95% confidence intervals. Statistical differences in geometric means of the observations were calculated using Turkey honestly significant multiple tests

 

2.1.2 Permethrin induced expression

The expression levels of GSTe2 in the control and larvae exposed to permethrin are shown in (Figure 2). The mean GSTe2 expression peaked at 2 hours post-exposure, increasing by 132% in the MAT strain and by 146% in the KGB strain (p < 0.001 for both strains).The gene expression was 17% and 15.6% lower in the MAT and the KGB strains respectively at 5-hours post-exposure; this difference relative to the control strain was statistically not significant for the KGB strain (p = 0.360) and for the MAT strain (p = 0.350). At 24 hours post-exposure, mean GSTe2 expression was 9% lower than control in the MAT strain but 16% higher in the KGB strain; however, neither difference was statistically significant (p = 0.846 for MAT and p = 0.700 for KGB).

 


Figure 2 Induction of Epsilon GSTe2 expression at 0, 2, 5 and 24 hours post-exposure to 0.02mg/L permethrin for one hour – geometric means and 95% confidence intervals

 

2.1.3 Hydrogen peroxide induced expression

The geometric means and their 95% confidence intervals for observed levels of GSTe2 expressed at 0 and 1 hour post-exposure to H2O2 are shown individually in (Figure 3).

 

 

Figure 3 Induction of Epsilon GSTe2 expression at 1 hours post-exposure to H2O2 – geometric means and 95% confidence intervals


Hydrogen peroxide produced increase in the expression of GSTe2 at 1 hour post-exposure of (24%) in the MAT strain (24%) and (28%) in the KGB strain (p = 0.029 and p = 0.005 respectively). The relative expression of GSTe2 induced by DDT, Permethrin and Hydrogen peroxide at different time points after exposure compared to control in larvae from the MAT and KGB strains to each of the xenobiotic is summarized on Table 2 and Table 3 respectively.

 


Table 2 (Geometric) Mean GSTe2 expression levels (with 95% confidence limits) in MAT strains

 


Table 3 (Geometric) Mean GSTe2 expression levels (with 95% confidence limits) in KGB strains

 

3 Discussion

We investigated patterns of expression of the epsilon GSTe2 gene induced by DDT, permethrin and hydrogen peroxide in the fourth instar larvae of the parental (non-selected) lines of KGB and MAT laboratory strains of An. arabiensis. The expression of GSTe2 increased significantly, 2.1 fold in the KGB and 1.6 fold in MAT strains one hour post-exposure to DDT. Similar patterns of GSTe2 expression in response to exposure to DDT have been reported in Aedes aegypti (Lumjuan et al., 2005), An. stephensi (Sanil et al., 2014), An. funestus Djouaka et al. (2011) and An. gambiae (Ding et al., 2005). Increase in levels of GSTe2 expression has been consistently associated to DDT resistance and AgGSTe2 have been shown to possess DDT dehydrochlorinase activity (Ranson et al., 2001; Pontes et al., 2016). Recent investigations on structure and function relationship and simulations based on conformational studies indicate higher efficiency and binding affinity for DDT of isoforms of GSTe2 in An. gambiae (Wang et al., 2008; Pontes et al., 2016; Kettman et al., 2011). It remains to be shown if the GSTe2 in An. arabiensis can metabolize DDT and posses the structural characteristics for the function. The expression of GSTe2 decreased after two hours by approximately 42% in the DDT-treated compared to control larvae in MAT strain, but in the KGB strain, the expression remained induced though not significantly (p = 0.146) (Figure 1). This observed difference in the induction pattern of GSTe2 suggests higher intrinsic levels of GST activity in the KGB compared to the MAT strain. Adults from the colony of the parental line of the KGB were more resistant to DDT, resistant ratio 1.5 compared to those from the MAT strain (Yayo et al., 2018a). DDT was a more potent inducer of GSTe2 than permethrin and H2O2 (Table 2; Table 3). The stronger induction of GSTe2 expression by DDT compared to permethrin and the H2O2 may be explained on the basis of substrate specificities (Aravindan et al., 2014). Permethrin and hydrogen peroxide consecutively increased the expression of GSTe2 by 40% and 30% in the KGB strain and in the MAT strain, permethrin increased the expression of the gene by 2.3 fold and hydrogen peroxide by 1.2 fold. Samra et al. (2012) have reported a two-fold increase in GSTe2 expression in Culex pipiense exposed to permethrin. In An. funestus permethrin has caused an increase of 2.55 in the expression of GSTe2 (Djuaka et al., 2011). We cannot explain the reason for the low expression recorded in the KGB strain. A border line significant induction of GSTe2 expression was observed (1.2-fold) in MAT and (1.4-fold) in KGB strains following exposure to 3 mM H2O2 (Table 2). This result is very similar to that reported for H2O2 induced expression of GSTe2 at one hour time point in An. gambiae (Ding et al., 2005). Exposure to DDT, permethrin and hydrogen peroxide can induce oxidative stress in the mosquitoes and the GSTs may contribute to antioxidant defence by direct peroxidase activity or conjugation of the toxic aldehyde 4-hydroxyl nonenol (Vontas et al., 2002; Sawicki et al., 2003; Abdullahi et al., 2004). The GSTe2 in Aedes aegypti and An. gambiae have been shown to possess peroxidase activity suggesting its potential role in defense against oxidative stress induced by xenoboitics (Ortelli et al., 2003; Lumjuan et al., 2005). Ding et al. (2005) attributed the induced expression of GSTe2 by H2O2 in An. gambiae to the presence of transcription factors responsive to oxidative stress in the promoter region of the gene this indicates the need for identifying the orthologous putative promoter elements in An arabiensis GSTe2.

 

The induction of GSTe2 expression by different chemical compounds may suggest its involvement in the detoxification of xenobiotic compounds found in the natural environment of the mosquito larvae. Elevated tolerance of mosquito larvae to chemical insecticides after their exposure to common herbicides has been demonstrated and this indicated the potential of pesticide residues in larval habitats to confer mosquito larvae pre-adaptive advantage to develop insecticide resistance (David et al. 2010). In this respect, a comparative study of molecular mechanisms controlling the induction of individual GST genes between An. arabiensis and An. gambiae may help to understand the ecological differences identified between these important malaria vectors (Coluzzi et al., 1979).

 

4 Conclusion

GSTe2 has been induced differentially by DDT, permethrin and hydrogen peroxide in laboratory strains of An. arabiensis. There is need to conduct further investigations on biochemical and molecular evidence for involvement of AaGSTe2 in metabolism of xenobiotics.

 

Authors’ contributions

Yayo A.M. and Hemingway J. draft the experimental design and conduct the Lab work; Yayo A.M., Hemingway J and Ado A. conduct the literature search; Yayo A.M., Hemingway J., Ado A. and Safiyanu M. review and edit the manuscript; Safiyanu M. Sambo F.I., Abubakar A. conduct the statistical analysis; All the respective authors have read the manuscript and approved its submission to Molecular Entomology Journal.

 

References

Abdullahi M., Ranjbar A., Shadnia S., Nikfar S., and Rezaie A., 2004, Pesticides and oxidative stress: a review- Medical Science Monitor, 10: RA141-147

 

Aravindan V., Muthukumaravel S., and and Gunasekaram K., 2014, Intercation affinity of Delta and Epsilon class glutathione S-transferases (GSTs) to bind with DDT for detoxification and conferring resistance in Anopheles gambiae, a malaria vector, J Vector Borne Disease, 51: 20, 8-15

 

Bamigdele O.B., Ajela J.O., and Olajuyigbe M.F., 2017, An evolution of glutathione transferase associated with Dichlorvos degradation in African palm weevil Rynchophorus phoenicis larva, Cogent Biology, 3: 1286764

https://doi.org/10.1080/23312025.2017.1286764

 

Cassone B.J., Kamdem C., Cheng C., John C.T., Mathew W., 2014, Gene expression divergence between malaria vectors sibling species Anopheles gambiae and Anopheles colluzzi from rural and urban Yaounde Cameroun. Mal. Ecol., 23(9): 2242-2259

https://doi.org/10.1111/mec.12733
PMid:24673723 PMCid:PMC4034741

 

Che-mendoza A., Penilla R.P., Rodríguez D.A., 2009, Insecticide resistance and glutathione S-transferases in mosquitoes: A review, African Journal of Biotechnology, 8: 1386-1397

 

Chiang F.M., and Sun C.N., 1993, Glutathionetransferase isozymes of diamondback moth larvae and their role in the degradation of some organophosphorus insecticides, Pestic. Biochem.,Physiol., 45,7-14

https://doi.org/10.1006/pest.1993.1002

 

Coetzee M., Hunt R., Wilkerson R., Della Torre A., Coulibaly M., and Besansky N., 2013, Anopheles coluzzii and Anopheles amharicus, new members of the Anopheles gambiae complex, Zootaxa, 3619: 246-274

https://doi.org/10.11646/zootaxa.3619.3.2
PMid:26131476

 

Coetzee M., Craig M., leSuer D., 2000, Distribution of African malaria mosquitoes belonging to the Anopheles gambiae complex, Parasit Today, 16: 74-77

https://doi.org/10.1016/S0169-4758(99)01563-X

 

Coluzzi M., Sabatini A., Petrarca V., and Di Deco M.A., 1979, Chromosomal differentiation and adaptation to human environments in the Anopheles gambiae complex, Trans R Soc Trop Med Hyg., 73: 483-497

https://doi.org/10.1016/0035-9203(79)90036-1

 

David J.P., Coissac E., Melodelima C., Poupardin R., Riaz M.A., Chandor-Proust A., and Reynaud S., 2010, Transcriptome response to pollutants and insecticides in the dengue vector Aedes aegypti using next-generation sequencing technology, BMC Genomics, 11: 216

https://doi.org/10.1186/1471-2164-11-216
PMid:20356352 PMCid:PMC2867825

 

Ding Y., Hawkes N., Meredith J., Eggleston P., Hemingway J., and Ranson H., 2005, Characterization of the promoters of Epsilon glutathione transferases in the mosquito Anopheles gambiae and their response to oxidative stress.- Biochemical Journal, 387: 879-888

https://doi.org/10.1042/BJ20041850
PMid:15631620 PMCid:PMC1135021

 

Djouaka R., Irving H., Tukur Z., and Wondji C.S., 2011, Exploring Mechanisms of Multiple Insecticide Resistance in a Population of the Malaria Vector Anopheles funestus in Benin, PLoS One, 6(11): e27760

https://doi.org/10.1371/journal.pone.0027760
PMid:22110757 PMCid:PMC3218031

 

Fang S.M., 2012, Insect glutathione S –transferases: a review of comparative genomic studies and response to xenobiotics, Bulletin of insectology, 65(2): 265-271

 

Giordano G., Afsharinejad Z., Guizzetti M., Vitalone A., Kavanagh T. J., and Costa L.G., 2007, Organophosphorus insecticides chlorpyrifos and diazinon and oxidative stress in neuronal cells in a genetic model of glutathione deficiency- Toxicology and Applied Pharmacology, 219: 181-189

https://doi.org/10.1016/j.taap.2006.09.016
PMid:17084875

 

Gregory R., Darby A.C., Irving H., Coulibaly M.B., Hughes M., Koekemoer L.L., Coetzee M., Ranson H., Hemingway J., Hall N., Wondji C.S., 2011, A de novo expression profiling of Anopheles funestus, malaria vector in Africa, using 454 pyrosequencing, PLoS One, 6: e17418

https://doi.org/10.1371/journal.pone.0017418
PMid:21364769 PMCid:PMC3045460

 

Hemingway J., and Ranson H., 2000, Insecticide resistance in insect vectors of human disease, Annual Review of Entomology, 45: 371-391

https://doi.org/10.1146/annurev.ento.45.1.371
PMid:10761582

 

Huang H.S., Hu N.T., Yao Y.E., Wu C.Y., Chiang S.W., 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 Biochemistry and Molecular Biology, 28: 651-658

https://doi.org/10.1016/S0965-1748(98)00049-6

 

Ketterman A.J., Saisawang C., and Wongsantichon J., 2011, Insect glutathione transferases, Drug Metabolism Reviews, 43: 253-65

https://doi.org/10.3109/03602532.2011.552911
PMid:21323601

 

Lumjuan N., McCarrol L., Prapanthadara L., Hemingway J., and Ranson H., 2005, Elevated activity of an Epsilon class glutathione transferase confers DDT resistance in the dengue vector, Aedes aegypti, Insect Biochem. Mol. Biol., 35: 861-871

https://doi.org/10.1016/j.ibmb.2005.03.008
PMid:15944082

 

Lumjuan N., Rajatileka S., Changsom D., Wicheer J., Leelapat P., Prapanthadara L. A., Somboon P., Lycett G., and Ranson H., 2011, The role of the Aedes aegypti Epsilon glutathione transferases in conferring resistance to DDT and pyrethroid insecticides, Insect Biochemistry and Molecular Biology, 41: 203-209

https://doi.org/10.1016/j.ibmb.2010.12.005
PMid:21195177

 

Marimo P., Hayeshi R., and Muvanganyama S., 2016, Inactivation of Anopheles gambiae glutathione transferase E2 by Epiphyllocoumar: Biochemistry Research International

https://doi.org/10.1155/2016/2516092
PMid:26925266 PMCid:PMC4746303

 

Mashyama S.T., Malabanam M.M., Akiva E., Bhosle R., Branch M.C., Hillerich B., Jagestser K., Kim J., Patskovsky Y., Seiclel R.D., Sterel M., Toro R., Yetting W.W., Almo S.C., Armstrong R.N., Babbit P.C., 2014, Large scale determination of sequence, structure and function relationship in cytosolic glutathione transferases across the biosphere, PLoS Biology, 2, 4: 1-19

https://doi.org/10.1371/journal.pbio.1001843
PMid:24756107 PMCid:PMC3995644

 

Matowo J., Jones M.C., Kabula B., Ranson H., Sten K., 2014, Genetic basis of pyrethroid resistance in populations of Anopheles arabiensis, the primary malaria vector in lower Moshi, north-eastern Tanzania, Parasites and Vectors 7: 274

https://doi.org/10.1186/1756-3305-7-274
PMid:24946780 PMCid:PMC4082164

 

Muller P., Chouaibou M., Pignatelli M., Etang J., Walker E.D., Donelly M.J., Simard F., and Ranson H., 2007, Pyrethroid tolerance is associated elevated expression of antioxidants and agricultural practice in Anopheles arabiensis sampled from an area of cotton fields in Northern Cameroon, Molecular Ecology, 1365-2947

https://doi.org/10.1111/j.1365-294X.2007.03617.x
PMid:18179425

 

Ortelli F., Rossiter L.C., Vontas J., Ranson H., and Hemingway J., 2003, Heterologous expression of four glutathione transferase genes genetically linked to a major insecticide-resistance locus from the malaria vector Anopheles gambiae, Biochem. J., 373: 957-963

https://doi.org/10.1042/bj20030169
PMid:12718742 PMCid:PMC1223529

 

Patience M., Rose H., ad Satrnley M., 2016, Inactivation of Anopheles gambiae Glutathione s-transferases E2 by epiphyllocoumarin, Biochemistry Research International, 10: 1155

https://doi.org/10.1155/2016/2516092

PMid:26925266 PMCid:PMC4746303

 

Pontes J.S., Maig R.T., Lima P.C., Ayre F.S., Soares T.A., 2016, The role of conformational dynamics of glutathione s-transferases in Anopheles gambiae, Journal Braz Chem Soc, 27(9): 1602-1616

 

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

https://doi.org/10.1017/S0007485300034350

 

Ranson H., N’Guessan R., Lines J., Moiroux N., Nkuni Z., and Corbel V., 2011, Pyrethroid resistance in African anopheline mosquitoes: what are the implications for malaria control? Trends in Parasitology, 27(2):  91-98

https://doi.org/10.1016/j.pt.2010.08.004
PMid:20843745

 

Ranson H., Rossiter L., Ortelli F., Jensen B., Wang X., Roth C.W., Collins F.H., and      Hemingway J., 2001, Identification of a novel class of insect glutathione S-transferases involved in resistance to DDT in the malaria vector Anopheles gambiae, Biochem. J., 359: 295-304

https://doi.org/10.1042/bj3590295
PMid:11583575 PMCid:PMC1222147

 

Samra A.I., Kamita S.G., Yao H.W., Cornel A.J., Hammock B.D., 2012, Cloning and characterization     of two glutathione s-transferases from pyrethroid resistant Culex pipiens, Pest Manag Sci, 68 (5): 764-772

https://doi.org/10.1002/ps.2324
PMid:22290868 PMCid:PMC3583349

 

Sanil D., Shetty V., and Shetty N.J., 2014, Differential expression of glutathione s- transferases enzyme in different life stages of various insecticide resistant strains of An. Stephensi, Journal of Vector Borne Disease 51: 97-105

 

Sawicki R., Singh S.P., Mondal A.K., Benes H., Zimniak P., 2003, Cloning, expression and biochemical characterization of one Epsilon-class (GST-3) and ten Delta-class (GST- 1) glutathione S-transferases from Drosophila melanogaster, and identification of additional nine members of the Epsilon class, Biochemical Journal, 370: 661-669

https://doi.org/10.1042/bj20021287
PMid:12443531 PMCid:PMC1223192

 

Tieriere L.C., 1984, Induction of detoxification enzymes in insects, Annual Review of Entomology, 29: 71-88

https://doi.org/10.1146/annurev.ento.29.1.71
PMid:6362551

 

Singh S.P., Cornella J.A., Benes H., Cochrane B.J., and Zimniak P., 2001, Catalytic function of Drosophila melanogaster glutathione s-transferase DMGSTSH (GST-2) in conjugation of lipid peroxidation end products, European Journal of Biochemeistry, 268: 2912-2923

https://doi.org/10.1046/j.1432-1327.2001.02179.x
PMid:11358508

 

Vontas J., Small G.J., Nikou D., Ranson H., Hemingway J., 2002, Purification, molecular cloning and heterologous expression of a glutathione s-transferase involved in insecticide resistance from the rice brown plant-hopper, Nilaparvata lugens, Biochemical Journal, 362: 329-337

https://doi.org/10.1042/bj3620329
PMid:11853540 PMCid:PMC1222392

 

Vontas J., Small G.J., Nikou D., Ranson H., Hemingway J., 2001, Glutathione s-transferases as an antioxidant defence agents confer pyrethroid resistance in Nilaparvata lugens, Biochemical Journal, 357: 65-72

https://doi.org/10.1042/bj3570065
PMid:11415437 PMCid:PMC1221929

 

Wongrakul J., Saengton P., Postril L., Wuottichai N., Prapandathara L., and Kettman A.J., 2010, Expression and characterization of three New Glutathione transferases and Epsilon (AcGSTe2-2), Omega (AcGSTO1-1) and Theta (AcGSTT-1) from Anopheles cracens, Journal of Medical Entomology, 47(2): 162-171

https://doi.org/10.1603/ME09132
PMid:20380296

 

Yayo A.M., Ado A., Habibu U.A., Muhammad B.R., Ebere N., and Heminway J., 2018, Expression patterns of epsilon glutathione s- transferases genes in developmental stages of susceptible and resistant lines of Anopheles arabiensis strains, International Journal of Entomology Research, 3(2): 143-151

 

Yayo A.M., Ado A., Habibu U.A., Muhammad B.R., Bello M.I., and Heminway J., 2018, Isolation of       epsilon class glutathione s-transferase GST genes from DDT resistant laboratory colony of Anopheles arabiensis MAT strains, International Journal of Entomology Research, 3(3): 11-18

Molecular Entomology
• Volume 10
View Options
. PDF(0KB)
. HTML
Associated material
. Readers' comments
Other articles by authors
. A.M. Yayo
. A. Ado
. M. Safiyanu
. B.R. Muhammad
. F.I. Sambo
. A. Abubakar
. J. Hemingway
Related articles
. Xenobiotic
. Expression
. Glutathione S-Transferases Gste2
. Anopheles arabiensis
Tools
. Email to a friend
. Post a comment

http://103.165.243.97/doc/luar-negeri/http://103.165.243.97/doc/sign/slot-thailand/http://103.165.243.97/doc/before_tte/zeus-slot/https://appv2.tanahlautkab.go.id/doc/mpo/https://www.chuka.ac.ke/gates-of-olympus-1000/https://appv2.tanahlautkab.go.id/doc/unsign/http://103.165.243.97/doc/kamboja/http://mysimpeg.gowakab.go.id/mysimpeg/bangkomplit/http://mysimpeg.gowakab.go.id/toto/http://mysimpeg.gowakab.go.id/mysimpeg/maxwin/https://jurnal.jsa.ikippgriptk.ac.id/public/luar/http://103.165.243.97/doc/dana/https://fk.ulm.ac.id/wp-content/uploads/https://fk.ulm.ac.id/wp-content/fonts/https://ppid.cimahikota.go.id/infomugi/slot-luar-negeri/https://www.unjc.cu/sweet-bonanza/http://103.165.243.97/doc/cth/https://ppid.cimahikota.go.id/assets/maxwin-slot/https://ijatr.polban.ac.id/docs/https://loa.tsipil-uii.ac.id/data/thailand/https://loa.tsipil-uii.ac.id/scatter-hitam/https://ijatr.polban.ac.id/toto/https://ppid.cimahikota.go.id/infomugi/https://ppid.cimahikota.go.id/image/slot-gacor-hari-ini/https://mpp.bandung.go.id/mpo/https://mpp.bandung.go.id/assets/css/https://mpp.bandung.go.id/display/nexus/https://simba.cilacapkab.go.id/simba_project/maxwin-slot/https://simba.cilacapkab.go.id/simba_project/sbobet88/https://simba.cilacapkab.go.id/idnslot/https://mpp.bandung.go.id/surat/idnslot/https://fk.ulm.ac.id/wp-content/thailand/https://perijinan.blitarkota.go.id/assets/jp-gacor/https://rdsp.msp.gob.do/api/thailand/https://perijinan.blitarkota.go.id/data/situs-toto/https://perijinan.blitarkota.go.id/assets2/https://mpp.bandung.go.id/attachments/https://mpp.bandung.go.id/images/https://mpp.bandung.go.id/assets/thailand/https://perijinan.blitarkota.go.id/data/toto-slot/https://simba.cilacapkab.go.id/db/toto-slot/https://simba.cilacapkab.go.id/vendor/https://perijinan.blitarkota.go.id/data/depo-10k/https://mpp.bandung.go.id/git/demo/https://mpp.bandung.go.id/api/jp-gacor/https://simba.cilacapkab.go.id/assets/depo-10k/https://simba.cilacapkab.go.id/api/demo/https://simba.cilacapkab.go.id/api/https://sim.stidar.ac.id/img/https://sim.stidar.ac.id/sweet-bonanza/https://sim.stidar.ac.id/slot-anti-rungkad/https://kuesioner.instidla.ac.id/akun-pro-platinum/https://lpm.instidla.ac.id/lucky-neko/https://ijabr.polban.ac.id/mpo/https://ijabr.polban.ac.id/idn/https://kuesioner.instidla.ac.id/wild-bandito/http://kwitansi.instidla.ac.id/demo-lucky-neko/https://perdami.or.id/wp-includes/zeus-slot/https://perdami.or.id/wp-content/slot-kamboja/https://siipbang.katingankab.go.id/storage_old/maxwin/http://kwitansi.instidla.ac.id/slot-garansi-kekalahan/https://bundamediagrup.co.id/wp-includes/idn/http://103.165.243.97/doc/maxwin-slot/https://bundamediagrup.co.id/akun/demo/https://pmb.sttabdisabda.ac.id/aplikasi/http://103.165.243.97/doc/sv388/https://bundamediagrup.co.id/wp-content/akun-pro-kamboja/https://ijabr.polban.ac.id/api/https://ijabr.polban.ac.id/-/pulsa/https://sipirus.sukabumikab.go.id/storage/uploads/jp-thailand/https://sipirus.sukabumikab.go.id/storage/uploads/-/sthai/https://sipirus.sukabumikab.go.id/storage/uploads/-/stoto/https://sipirus.sukabumikab.go.id/storage/uploads/server-kamboja/https://alwasilahlilhasanah.ac.id/starlight-princess-1000/https://www.remap.ugto.mx/pages/slot-luar-negeri-winrate-tertinggi/https://waper.serdangbedagaikab.go.id/public/images/qrcode/slot-dana/https://waper.serdangbedagaikab.go.id/public/images/qrcode/slot-deposit-pulsa/https://waper.serdangbedagaikab.go.id/storage/framework/https://wbs.klaten.go.id/public/assets/http://103.165.243.97/doc/thailand/https://appv2.tanahlautkab.go.id/easy-win/https://tpid.katingankab.go.id/images/https://kuesioner.instidla.ac.id/asset/http://kwitansi.instidla.ac.id/database/http://lms.instidla.ac.id/backup/https://lpm.instidla.ac.id/wp-includes/block-patterns/http://mutu.instidla.ac.id/app/<https://bakesbangpol.katingankab.go.id/uploads/pulsaa style="display:none" href="https://sipirus.sukabumikab.go.id/storage/uploads/kantah/">https://sipirus.sukabumikab.go.id/storage/uploads/kantah/https://sipirus.sukabumikab.go.id/storage/uploads/slot-depo-10k/https://ijabr.polban.ac.id/classes/slot-gacor-gampang-menang/https://ijabr.polban.ac.id/registry/https://ijabr.polban.ac.id/locale/https://lpm.instidla.ac.id/wp-content/uploads/https://sipirus.sukabumikab.go.id/storage/uploads/pembahas/