A comparison of Drosophila melanogaster detoxication gene induction responses for six insecticides, caffeine and phenobarbital

background image

Insect

Biochemistry

and

Molecular

Biology

Insect Biochemistry and Molecular Biology 36 (2006) 934–942

A comparison of Drosophila melanogaster detoxification gene induction

responses for six insecticides, caffeine and phenobarbital

$

Lee Willoughby

a

, Henry Chung

a

, Chris Lumb

a

, Charles Robin

b

,

Philip Batterham

a,

, Phillip J. Daborn

a

a

Centre for Environmental Stress and Adaptation Research (CESAR), Department of Genetics, Bio21 Molecular Science and Biotechnology Institute,

The University of Melbourne, Vic. 3010, Australia

b

Department of Genetics, The University of Melbourne, Vic. 3010, Australia

Received 21 August 2006; received in revised form 12 September 2006; accepted 12 September 2006

Abstract

Modifications of metabolic pathways are important in insecticide resistance evolution. Mutations leading to changes in expression

levels or substrate specificities of cytochrome P450 (P450), glutathione-S-transferase (GST) and esterase genes have been linked to many
cases of resistance with the responsible enzyme shown to utilize the insecticide as a substrate. Many studies show that the substrates of
enzymes are capable of inducing the expression of those enzymes. We investigated if this was the case for insecticides and the enzymes
responsible for their metabolism. The induction responses for P450s, GSTs and esterases to six different insecticides were investigated
using a custom designed microarray in Drosophila melanogaster. Even though these gene families can all contribute to insecticide
resistance, their induction responses when exposed to insecticides are minimal. The insecticides spinosad, diazinon, nitenpyram,
lufenuron and dicyclanil did not induce any P450, GST or esterase gene expression after a short exposure to high lethal concentrations of
insecticide. DDT elicited the low-level induction of one GST and one P450. These results are in contrast to induction responses we
observed for the natural plant compound caffeine and the barbituate drug phenobarbital, both of which highly induced a number of
P450 and GST genes under the same short exposure regime. Our results indicate that, under the insecticide exposure conditions we used,
constitutive over-expression of metabolic genes play more of a role in insect survival than induction of members of these gene families.
r

2006 Elsevier Ltd. All rights reserved.

Keywords: Cytochrome P450; Insecticide; Microarray; Glutathione-S-transferase; Gene expression; Insecticide resistance

1. Introduction

The ability of a substrate to increase the activity of

enzymes that are capable of metabolizing it is a key feature
of many different biological pathways. Rather than
constitutively expressing genes involved in metabolizing a
substrate, transcriptional induction in response to the
substrate is a means of activating gene expression only
when required. This presumably causes less of a general
metabolic burden than the constitutive expression of all
metabolic enzymes, and helps protect the organism from
the activities of promiscuous enzymes. The link between

exogenous compounds acting as induction agents and the
induced enzymes metabolizing them has been established in
mammalian detoxification systems, generally in drug–drug
interaction studies (

Whitlock, 1999

;

Luo et al., 2004

;

Chen

and Raymond, 2006

). The genes induced by drugs encode

enzymes involved in the metabolism of those drugs,
including members of the cytochrome P450 (P450) and
glutathione-S-transferase (GST) families (

Waxman, 1999

;

Whitlock, 1999

;

Denison and Nagy, 2003

;

Luo et al., 2004

;

Ellinger-Ziegelbauer et al., 2005

;

Chen and Raymond,

2006

).

It is well established that members of the P450, GST and

esterase families are important in many instances of
insecticide resistance. Resistance results from genetic
changes leading to either altered expression, or altered
function, of genes in these families leading to increased

ARTICLE IN PRESS

www.elsevier.com/locate/ibmb

0965-1748/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:

10.1016/j.ibmb.2006.09.004

$

Data deposition:

www.ncbi.nlm.nih.gov/geo/

.

Corresponding author. Tel.: +61 38344 2363; fax: +61 39347 5352.

E-mail address:

p.batterham@unimelb.edu.au (P. Batterham).

background image

metabolism or sequestration of insecticides before they can
reach their molecular target (

Field et al., 1988, 1999

;

Tang

and Tu, 1994

;

Newcomb et al., 1997

;

Berge et al., 1998

;

Ranson et al., 2001

;

Coleman et al., 2002

;

Daborn et al.,

2002

;

Vontas et al., 2002

;

Ortelli et al., 2003

;

Bogwitz et al.,

2005

). The P450, GST and esterase gene families are large,

rapidly evolving gene families. Little is known about the
substrate specificity of most of the encoded enzymes,
making the ability to predict which genes have the potential
to be involved in insecticide resistance difficult.

If substrates can induce the expression of enzymes

involved in their metabolism, as has been suggested
specifically for P450s (

Whitlock, 1986

), and demonstrated

in mammalian detoxification systems (

Waxman, 1999

;

Whitlock, 1999

;

Denison and Nagy, 2003

;

Luo et al.,

2004

;

Ellinger-Ziegelbauer et al., 2005

;

Chen and Ray-

mond, 2006

), then the gene induction responses of insect

P450s, GSTs and esterases to insecticides could be used to
identify those enzymes with the capacity to metabolize
insecticides. Insects exhibit induction responses to other
xenobiotics they come into contact with, such as toxic plant
compounds. For example, the black swallowtail butterfly
Papilio polyxenes induces the P450 genes Cyp6B1 and
Cyp6B3 in response to the toxic furanocoumarin com-
pound xanthotoxin produced by plant families such as
Apiaceae and Rutaceae (

Cohen et al., 1992

;

Hung et al.,

1995

). Both CYP6B1 and CYP6B3 are capable of

metabolizing xanthotoxin to varying degrees, enabling
P. polyxenes to use these plants as a food source (

Petersen

et al., 2001

;

Wen et al., 2003

). A similar response occurs in

the cotton bollworm, Helicoverpa zea, where Cyp6B8 and
Cyp321A1 are induced by and are capable of metabolizing
xanthotoxin (

Li et al., 2003

;

Sasabe et al., 2004

). The

cactophilic Drosophila species D. mettleri feeds on the toxic
allelochemical-producing necrotic tissues of the columnar
cacti species, a substrate that is toxic to all but the normal
resident species. Toxic isoquinoline alkaloids of the cactus
highly induce the expression of the D. mettleri P450
Cyp4D10, which has been suggested to be involved in the
metabolism of isoquinoline alkaloids (

Danielson et al.,

1997

).

In this study, we investigated the capacity of insecticides

to induce the expression of P450, GST or esterase genes
potentially involved in their metabolism. In the absence of
a full genome sequence for any pest insect species we
conducted this study in D. melanogaster where a micro-
array containing all of the P450, GST and esterase genes
was constructed to assay the transcriptional response to
insecticide exposure. While D. melanogaster is not generally
considered to be a pest species, field resistance has been
observed for most of the major insecticides used in
agriculture (

Wilson, 2001, 2005

;

Daborn et al., 2002

;

Bogwitz et al., 2005

). In this study insects were exposed to

six chemically distinct insecticides that have been widely
used in the field plus the natural plant compound caffeine,
and the known P450 inducer phenobarbital (PB). One or
more of the metabolic genes has been shown to have the

capacity to confer resistance to five of the six insecticides
(

Daborn et al., 2002

;

Bogwitz et al., 2005

) (Daborn et al.,

unpublished results). If insecticides induce the expression
of genes responsible for their metabolism, induction could
be used to identify genes with the capacity to be involved in
metabolic resistance. However, this study shows that there
is a minimal induction response to insecticide exposure.

2. Materials and methods

2.1. Fly strain

The y; cn bw sp strain of D. melanogaster (Bloomington

Drosophila Stock Center, Indiana University, IN) was used
for all of the induction experiments reported here. This
strain is isochromosomal for all chromosomes and its
genome has been sequenced (

Adams et al., 2000

).

2.2. Exposure to PB, caffeine and insecticides

DDT, spinosad, nitenpyram and diazinon are fast acting

insecticides that target the nervous system. Lufenuron and
dicyclanil are both insect growth regulators (IGRs),
causing larval lethality usually during life stage transitions.
Third instar larvae were exposed to the insecticides
lufenuron, dicyclanil, spinosad, nitenpyram and diazinon
via the food source for 4 h. For lufenuron and dicyclanil,
third instar larvae were exposed to concentrations sig-
nificantly higher than those required to arrest development
at a life stage transition (500 times and 1000 times,
respectively). For spinosad, nitenpyram and diazinon,
third instar larvae were exposed to a concentration that
exceeded LC99. However, in the 4-h exposure period there
was no significant mortality.

Four-day-old adult males were exposed to nitenpyram

and DDT via direct contact for 4 h, at a concentration that
would be lethal after 12 h of exposure. For adult exposures,
males of the y; cn bw sp strain were collected within a 24 h
window after emergence, sorted into groups of 50 males
and then stored at 25 1C for 4 days. Contact exposure was
conducted, whereby the relevant amount of each com-
pound was added to 150 ml of acetone and immediately
transferred to a 30 ml scintillation vial, which was then
rolled until the acetone had evaporated. A total of 1 M PB
(Sigma) solution (dissolved in dH

2

O), 10 ml for microarray

analysis or 2 ml for timecourse was used for each vial; the
dose used for microarray analysis was lethal over 24 h, so a
lower dose was used for time-course analysis. A total of
10 mg DDT (Sigma) (dissolved in acetone) and 10 ml 1 mM
Nitenpyram (Novartis) (dissolved in dH

2

O) were applied to

each vial. A total of 50 males were transferred into each
scintillation vial, which were then sealed with cotton wool
dampened with dH

2

O. After the appropriate time period

(1–24 h for the time-course experiment, or 4 h for all
microarray experiments), the flies were frozen in liquid
nitrogen and stored at 80 1C until RNA extraction.

ARTICLE IN PRESS

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

935

background image

For larval exposures, adults of the y; cn bw sp strain were

allowed to lay eggs for a 6 h period on Petri dishes
containing standard fly food with 2 agar. Petri dishes
were then collected and incubated at 25 1C. Larvae were
allowed to develop on the original Petri dishes until they
were to be exposed, so that all larvae would be 4.5 days old
when frozen after exposure. Fresh Petri dishes containing
standard fly food with 2 agar were prepared for
exposure, containing either the compound or the appro-
priate control. A final concentration of 10 mM PB (Sigma)
for microarray analysis or 1 mM PB for timecourse
analysis was used; the dose used for microarray analysis
was lethal over 24 h, so a lower dose was used for time-
course analysis. Final concentrations of 6 ppm nitenpyram
(Novartis), 6.7 10

2

% dicyclanil (Novartis), 20 ppm

lufenruon (Novartis), 1.3 10

2

% diazinon (Coopers

Dijet), 6 10

5

% spinosad (Success—Dow Agrosciences)

and 1.5 mg/ml caffeine (Sigma) were also used. 50 larvae
were transferred to each Petri dish. After the appropriate
time period (1–24 h for the time-course experiment, or 4 h
for all microarray experiments), the larvae were frozen in
liquid nitrogen and stored at 80 1C until RNA extraction.

2.3. Isolation of RNA and real-time PCR

RNA was extracted from pooled fly samples using Trizol

reagent (Invitrogen). RQ1 DNase (Promega) treatment
was performed and then an additional round of Trizol
extraction was performed. Reverse transcription using 5 mg
of total RNA was conducted with the Superscipt first
strand synthesis kit (Invitrogen). Quantitative real-time
analysis was performed using SYBR green kit (QIAGEN)
on a Rotor Gene-3000 real-time PCR machine (Corbett).
Dilutions of the exposed cDNA sample (1 , 1/10 and
1/50 ) and 1 of the unexposed sample were set up.
Real-time PCR primers; Cyp6a2 AAACGGTGCTGGAG-
GAAC and TTATGACCTGTGTGCCCTTC; Cyp6a8
GGCTGAGGTGGAGGAGGT and CGATGACGAAG-
TTTGGATGA; Cyp6a9 CCCAGCATCAGGACATTCA
and GCTCCACACGGAATCAAAC; Cyp6a21 CATG-
GATTCGCACTGTATG and CGGGAGAACGGTGTA-
CAATC; Cyp12d1 AGGAACACAAGTAAAGGCCAC
and GTCCATTCAAGACCATGTTCC; GstD2 TGTC-
CACTGTCTCCACGTTC and GGAGTCACCTTCTT-
GGCATT; GstD7 TGGCTGATATCGTCATCCTG and
GCATTCTTAAGCCACCTCTCC; RpL11 CGATCCCT-
CCATCGGTATCT and AACCACTTCATGGCATCC-
TC. Real-time PCR validation of microarray results was
conducted on additional biological replicates.

2.4. Microarray generation

A cDNA microarray containing fragments of 186 D.

melanogaster genes was constructed, including 89 P450s, 37
GSTs and 32 esterases. Each cloned DNA fragment was
then amplified by PCR using the T7 and M13rev primers
using Amplitaq Gold (ABI). In addition to the control

genes, additional series of control spots were included;
PCR products from all of the control genes were pooled
and then a 9-point dilution series of this sample was printed
(250, 125, 60, 30, 15, 7, 4, 2, 1 ng/ml). Plasmids were
extracted from the LD cDNA library (BDGP;

www.fruit-

fly.org/EST/

) and digested using either AluI or BstUI

restriction enzymes (Promega). Digests were pooled and a
dilution series was printed in the same manner as described
for the pooled PCR products. The Lucidea Universal
scorecard (Amersham) was included to aid in evaluating
the quality of the experiments. Slides were printed by the
Australian Genome Research Facility (AGRF;

www.agrf.

org.au

) on GAPS II (Corning) slides. Each unique feature

(not the control dilution series) was duplicated 4 times in a
four printing block arrangement. Full details of the slide
are available from the Gene Expression Omnibus (GEO):
GPL4239 (

www.ncbi.nlm.nih.gov/geo

).

2.5. Microarray procedure

Dye-swap microarrays were conducted for each experi-

mental condition, resulting in 16 individual data points for
each gene. Total RNA was extracted from pooled fly
samples (400 flies per condition) using Trizol reagent
(Invitrogen). RNA was purified using RNeasy minikit
(QIAGEN) and concentrated using NaOAc precipitation.
Lucidea Universal scorecard RNA (Amersham) was added
to 60 mg RNA and labeled with either Cy3-dCTP or Cy5-
dCTP (Amersham) using Superscript RT II (Invitrogen).
Hybridization mixture, including labeled cDNA, was
added to the microarray and hybridized for 20 h at 68 1C.
After hybridization, the slides were washed, and were then
imaged using a Genepix 4000B microarray scanner; image
analysis was conducted using the manufacturer’s software.
Data analysis was conducted with LimmaGUI (

http://

bioinf.wehi.edu.au/limmaGUI/

), a graphical interface for

Limma (

Smyth and Speed, 2003

). Data was normalized

with the print-tip loess method and genes were considered
differentially expressed if they had P-values

o0.05 after

multiple comparison correction (Holm correction) and
were greater than 2-fold up or down regulated. A more in
depth description of the experimental approaches used for
the microarray analyses presented in this paper is available
from GEO (

www.ncbi.nlm.nih.gov/geo

): GPL4239 and

GSE5713.

2.6. In situ hybridizations

Cyp12d1 (Primers: AAAAGGAGATCTATGAATA-

CATTGAGCAGTG and AAATGAGCGGCCGCTTA-
TTGTTCGTATCCGTGAATTTG) and GstD2 (Primers:
CAGGCGTAGTTCAGCACTCA and AGTGTGCTT-
CTCCCCTAACA) were amplified by PCR using Taq
DNA polymerase (Promega) and cloned into pGEM-T
Easy (Promega) in both orientations with respect to the
T7 polmerase annealing site. The sense and antisense
constructs were then linearised with SalI (Promega),

ARTICLE IN PRESS

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

936

background image

transcribed with Megascript T7 polymerase (Ambion), and
labeled with digoxigenin-labeled dNTP mix (Roche). In
situ hybridizations were performed on dissected third-
instar larvae using standard techniques (

Tautz and Pfeifle,

1989

).

3. Results

3.1. Phenobarbital induction time course

To identify the best conditions under which to measure

gene induction, a PB exposure time course was conducted
in both third instar larvae and adult males of the y; cn bw
sp strain. PB is known to induce the expression of
numerous P450s in mammals (

Honkakoski et al., 1998

;

Meyer and Hoffmann, 1999

;

Kodama and Negishi, 2006

),

is documented to induce the P450s Cyp6a2 and Cyp6a8 in
D. melanogaster (

Maitra et al., 1996

;

Dunkov et al., 1997

),

and has recently been shown to induce several other P450s
in D. melanogaster (

King-Jones et al., 2006

;

Le Goff et al.,

2006

;

Sun et al., 2006

). Third instar larvae and 4-day old

adult males, were exposed to PB via food source, and
contact exposure, respectively. For each time point over a
24 h period changes in mRNA levels of Cyp6a2 and
Cyp6a8 were quantified using real-time PCR. In both third
instar larvae and adult males, Cyp6a2 and Cyp6a8 showed
a biphasic PB induction response (

Fig. 1

). The observation

of biphasic induction was replicated using Act42A as the
housekeeping control gene (data not shown). This type of
response has not been reported in previous studies of PB

induction in insects, possibly due to the sensitivity of
techniques or experimental design (

Carino et al., 1992

;

Dunkov et al., 1997

) but has been previously described for

Cyp1a6 and Cyp1a1 in response to 3-methylcholanthrene
in eels (

Ogino et al., 1998

). It is possible that genes

responding to PB treatment metabolize, or sequester PB,
temporarily lowering the level of PB within the organism,
thus decreasing the fold induction. As an initial induction
peak for both Cyp6a2 and Cyp6a8 was detected after 4 h of
exposure to PB, 4 h exposure was used to characterize gene
induction for the entire P450, GST and esterase families in
the subsequent microarray experiments.

3.2. Phenobarbital microarray analysis

After a 4 h exposure to PB, changes in P450, GST and

esterase gene expression for both larvae and adults were
determined using microarray analysis. Larval treatment
with PB did not result in changes in the mRNA level of any
of the 32 esterase genes. In contrast, 9 of 37 GST genes and
21 of 89 P450 genes were induced by PB (

Table 1

). Similar

results were obtained after PB exposure in adult males,
with no esterase genes, 6 GSTs and 10 P450 genes induced
(

Table 2

). Considerable overlap in the PB induction

response between third instar larvae and adult males was
detected, with many of the same genes being induced; 10
P450 and 4 GST genes were induced in both adults and
larvae. Microarray results can be seen in full at GEO:
GSE5713 (

http://www.ncbi.nlm.nih.gov/geo

). Quantitative

real-time PCR on biological replicates for a selection of

ARTICLE IN PRESS

0

3

6

9

12

15

0

4

8

12

16

20

24

Cyp6a8 larvae

0

5

10

15

20

25

30

0

4

8

12

16

20

24

0

3

6

9

12

15

0

4

8

12

16

20

24

0

5

10

15

20

25

30

0

4

8

12

16

20

24

Time (hours)

Time (hours)

Time (hours)

Time (hours)

Fold Induction

Fold Induction

Cyp6a2 adult

Cyp6a8 adult

Cyp6a2 larvae

(A)

(B)

(C)

(D)

Fig. 1. Fold induction over time in response to PB exposure. Transcriptional response when larvae are exposed to PB, for Cyp6a2 (A) and for Cyp6a8 (B)
and response when adult males are exposed to PB, for Cyp6a2 (C) and for Cyp6a8 (D), measured at 1 h after commencement of exposure, every hour until
20 h and then again at 24 h. Expression levels determined with quantitative real-time PCR. The fold change at each time point is relative to a similarly
handled unexposed sample.

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

937

background image

differentially expressed genes was successful in confirming
the microarray results (

Tables 1 and 2

).

3.3. Caffeine microarray analysis

Third instar larvae were exposed to caffeine via the food

source, with changes in P450, GST and esterase gene
expression then determined using microarray analysis.
Larval treatment with caffeine did not result in the changes
in mRNA level of any of the 32 esterase genes. However, 5
of 37 GST and 11 of 89 P450 genes were induced by
caffeine, microarray results can be seen in full at GEO:
GSE5713 (

http://www.ncbi.nlm.nih.gov/geo

), with results

confirmed by quantitative real-time PCR on biological
replicates (

Table 3

). The genes induced in response to

caffeine consist of a similar gene set to those induced
by PB.

3.4. Tissue specificity of induction

The tissue specificity of the PB induction response for

Cyp12d1 and GSTD2 was investigated in third instar larvae
using in situ hybridization. Microarray experiments in
whole larvae show Cyp12d1 and GstD2 to be induced 29
fold and 21 fold, respectively, in response to PB, and 10
fold and 1 fold, respectively, in response to caffeine.
Expression of Cyp12d1 is visible throughout the midgut, in
the Malpighian tubules, the fat body and gastric caecae in
unexposed controls (

Fig. 2A

). After exposure to both

caffeine and PB expression is clearly increased in these
same tissues (

Fig. 2B and C

, respectively). Expression of

GstD2 is visible in regions of the midgut, the gastric caecae,
Malpighian tubules, ureters and hindgut in unexposed

ARTICLE IN PRESS

Table 1
Genes differentially expressed in larvae after exposure for 4 h to 10 mM
phenobarbital, with correction for multiple comparisons

Gene

Fold

P-value

Real time

Cytochrome P450

Cyp12d1

30.9

4.39 10

8

31.7

Cyp4ae1

19.0

2.12 10

5

Cyp6a21

18.6

5.23 10

6

15.7

Cyp4d14

17.4

5.28 10

6

Cyp6w1

15.7

1.75 10

5

Cyp6a9

14.8

5.25 10

7

13.1

Cyp6a2

13.5

3.73 10

4

28.3

Cyp6d5

11.9

3.74 10

4

Cyp28a5

11.5

7.94 10

8

Cyp12c1

9.4

1.66 10

4

Cyp6d4

9.2

1.69 10

5

Cyp4e2

7.4

1.33 10

5

Cyp4p1

6.7

1.20 10

5

Cyp6a8

5.7

4.08 10

2

9.5

Cyp9b1

5.2

4.24 10

4

Cyp12b2

4.9

1.19 10

4

Cyp4e3

4.9

6.12 10

4

Cyp6g1

4.6

3.23 10

2

Cyp9b2

3.6

2.96 10

2

Cyp6a23

3.4

3.44 10

4

Cyp12a5

3.0

3.84 10

4

Glutathione-S-transferase

GstD2

21.7

3.63 10

8

32.3

GstD7

15.7

2.35 10

4

25.7

CG17524

10.3

1.36 10-

5

GstD5

7.5

2.03 10

5

CG6776

4.6

7.44 10

5

CG1681

3.8

2.61 10

3

GstD6

3.6

2.72 10

5

GstD4

3.5

2.55 10

4

GstD10

2.3

2.94 10

3

Table 2
Genes differentially expressed in adult males after contact exposure for 4 h
to 10 ml 1 M phenobarbital, with correction for multiple comparisons

Gene

Fold

P-value

Real time

Cytochrome P450

Cyp4ae1

21.8

1.04 10

2

Cyp6a8

21.5

3.97 10

3

21.2

Cyp12d1

16.2

2.89 10

2

15.4

Cyp6a2

15.6

8.68 10

3

21.6

Cyp6a21

12.4

3.92 10

2

17.3

Cyp6w1

9.8

2.94 10

2

Cyp6d5

8.1

7.89 10

2

Cyp4e2

3.9

4.84 10

2

Cyp4p1

3.2

1.95 10

2

Cyp6g1

2.6

1.16 10

2

Glutathione-S-transferase

GstD2

16.9

2.60 10

2

20.8

CG17524

8.6

7.13 10

2

GstD1

3.7

1.76 10

2

GstD5

3.4

4.36 10

4

GstE1

2.6

1.72 10

2

GstE8

2.0

4.19 10

2

Table 3
Genes differentially expressed in larvae after exposure for 4 h to 1.5 mg/ml
caffeine, with correction for multiple comparisons

Gene

Fold

P-value

Real time

Cytochrome P450

Cyp6w1

12.6

1.09 10

6

Cyp12d1

11.4

5.11 10

4

15.7

Cyp6a8

10.9

2.57 10

5

14.6

Cyp6d5

10.6

8.27 10

4

Cyp6a21

4.4

1.20 10

4

6.3

Cyp4e2

3.8

1.88 10

4

Cyp4ae1

3.3

4.08 10

4

Cyp6a2

2.9

1.46 10

2

Cyp6g1

2.5

1.73 10

2

2.5

Cyp6d4

2.5

3.59 10

5

Cyp4d14

2.3

7.28 10

3

Glutathione-S-transferase

CG17524

6.3

3.33 10

5

CG6662

3.3

8.72 10

4

GstE1

2.9

2.71 10

4

GstD7

2.1

4.27 10

4

1.9

GstD2

2.1

3.54 10

3

2.2

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

938

background image

larvae (

Fig. 2D

). After exposure to caffeine expression is

not noticeably increased (

Fig. 2E

). However, after PB

exposure expression is clearly increased in these tissues
(

Fig. 2F

). Induction by both PB and caffeine appears to be

restricted to the tissues in which Cyp12d1 and GstD2 are
normally expressed.

3.5. Induction response to insecticides

The induction response to six different insecticides was

investigated. Gene induction responses to lufenuron,
dicyclanil, spinosad, nitenpyram and diazinon were in-
vestigated in 3rd instar larvae, while induction responses to
DDT and nitenpyram were investigated in adults. Of the
six insecticides tested, the only one to exhibit a gene
induction response was DDT. All other insecticides did not
induce the expression of any P450s, GSTs or esterases.
Exposure to DDT induced 1 of 37 GST genes (GstD2 at 2
fold, P ¼ 1:6 10

4

) and 1 of 89 P450 genes (Cyp12d1 at 3

fold, P ¼ 1:2 10

4

). No members of the esterase gene

family were induced in response to DDT exposure. The
two genes induced in response to DDT are also induced by
PB and caffeine, and at a much higher level. Microarray
results can be seen in full at GEO: GSE5713 (

http://

www.ncbi.nlm.nih.gov/geo

).

4. Discussion

Detoxification pathways have evolved to aid in the

metabolism of potentially toxic chemical compounds an

organism may encounter in its environment. In many
biological systems, substrates of these pathways induce the
expression of the metabolic enzymes involved in their
metabolism (

Whitlock, 1986, 1999

;

Waxman, 1999

;

Deni-

son and Nagy, 2003

;

Luo et al., 2004

;

Ellinger-Ziegelbauer

et al., 2005

;

Chen and Raymond, 2006

). We investigated

the response of D. melanogaster to six chemically distinct
insecticides to determine if insecticides induce detoxifica-
tion enzymes, as it is known that detoxification enzymes
are capable of metabolizing insecticides.

Our results demonstrate that, with the exception of

DDT, the insecticides tested using our exposure regime do
not induce the expression of P450, GST or esterase genes,
even though members of these gene families have
important roles in insecticide resistance and metabolism
in D. melanogaster. The overexpression of Cyp12a4 results
in lufenuron resistance (

Bogwitz et al., 2005

), and the

overexpression of Cyp6g1 confers lufenuron, nitenpyram,
DDT and dicyclanil resistance (

Daborn et al., 2002

)

(Daborn

et

al.,

unpublished

results).

Additionally,

CYP6A2 and GSTD1 are capable of metabolizing DDT
(

Tang and Tu, 1994

;

Amichot et al., 2004

) when expressed

in heterologous systems.

The one insecticide to which we observed an induction

response was DDT, with the induction of a single P450
(Cyp12d1, 3 fold) and a single GST (GstD2, 2 fold). This
induction is very weak compared with the response to PB
and caffeine in terms of the number of genes and fold
induction. Again, here there is no significant relationship
between

induction,

metabolism

and

resistance.

The

ARTICLE IN PRESS

Fig. 2. In situ hybridization of dissected third instar Drosophila melanogaster larvae. Untreated larvae, probed for Cyp12d1 expression pattern (A) and
GstD2 expression pattern (D). Caffeine treated larvae, probed for Cyp12d1 (B) and GstD2 (E). PB treated larvae, probed for Cyp12d1 (C) and GstD2 (F).
The hindgut (Hg), Malpighian tubules (Mt) and gastric caecae (Gc) are labeled. The ureters are is located at the point where the Malpighian tubules
converge. The midgut is located between the ureter and the gastric caecae.

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

939

background image

Cyp6g1, Cyp6a2 and GstD1 genes that have been
implicated in DDT resistance or metabolism (

Tang and

Tu, 1994

;

Daborn et al., 2002

;

Amichot et al., 2004

) are not

inducible by DDT in our experiments. Of the genes that are
induced by DDT, in vivo assays show that purified GstD2
has no detectable DDT-dehydrochlorinase activity (

Tang

and Tu, 1994

), whilst the involvement of Cyp12d1 in DDT

resistance is uncertain. Cyp12d1 is found to be over-
expressed in one DDT resistant strain that also over-
expresses Cyp6g1 (

Brandt et al., 2002

;

Le Goff et al., 2003

).

In summary, our results suggest that induction is not a
valid approach for determining which P450 and GST gene
family members are capable of insecticide detoxification,
since in the vast majority of cases no induction responses
are observed. When induction responses are observed, the
genes induced are different to the genes involved in
insecticide detoxification. This study has focused on a
single species, D. melanogaster, where the availability of a
whole genome sequence has allowed all of the P450, GST
and esterase genes to be analysed. There is no reason to
suspect that we are observing a strain or species-specific
phenomenon, however this possibility remains. It has been
observed that induction responses can be lower in
insecticide resistant strains, for example PB induction of
Cyp6a2 is lower in a DDT resistant strain in comparison to
a wildtype strain (

Brun et al., 1996

;

Amichot et al., 1998

).

This could reflect the nature of the mutation causing
constitutive over-expression of Cyp6a2 in the DDT-
resistant strain.

In stark contrast to the lack of induction observed in

response to insecticides, 16 genes were induced in larvae
responding to caffeine (

Table 2

). Fourteen of these genes

were also induced in response to PB. Cyp12d1 stands out as
a gene that is induced by a broad spectrum of compounds
including PB (

King-Jones et al., 2006

;

Le Goff et al., 2006

;

Sun et al., 2006

; this study), caffeine (this study), DDT (this

study), pyrethrum and piperamides (

Jensen et al., 2006

)

and the herbicide atrazine (

Le Goff et al., 2006

). Gene

induction after exposure to both PB and caffeine has also
been investigated in human cell culture where both
compounds induce detoxification gene family members
(

Honkakoski et al., 1998

;

Waxman, 1999

). Most signifi-

cantly, the induction responses to caffeine and PB are
regulated by separate receptors. PB mediated induction
occurs through the PXR/CAR xenobiotic responsive
nuclear receptors (

Honkakoski et al., 1998

). The induction

response to caffeine occurs through a completely distinct
xenobiotic detecting receptor, a member of the bHLH-PAS
receptor superfamily, the AhR (

Goadstuff et al., 1996

).

Using our data we are unable to reach any solid
conclusions on what is occurring in Drosophila. However,
our results suggest that in insects either two distinct
receptors have evolved the ability to regulate a very similar
set of genes, or that compounds binding to two receptors in
mammals only bind to one receptor in D. melanogaster. If
more than one receptor pathway exists to regulate similar
sets of genes, then these genes may be highly important for

the interaction between D. melanogaster and its environ-
ment. Recently, DHR96, the Drosophila orthologue of the
mammalian PXR and CAR xenobiotic receptors, has been
shown to play a role in the induction response to PB (

King-

Jones et al., 2006

). The role of DHR96 in P450 induction

by caffeine has not been investigated.

The tissue specificity of the caffeine and PB induction

responses were investigated for two induced genes,
Cyp12d1 and GstD2. Basal expression and induction was
detected in the key metabolic tissues, namely sections of the
midgut, and the Malpighian tubules (

Fig. 2

). Notably there

were differences in the basal expression patterns of these
two genes but in each case induction was limited to the
tissues where basal expression was observed. Therefore, PB
and caffeine both induced Cyp12d1 with the same tissue
specificity and they both induced GstD2 with the same
tissue specificity, however, these induction patterns differed
between GstD2 and Cyp12d1. In this instance, the cis-
regulatory elements controlling the expression of these two
genes may not be acting independently; the induction
module may be acting solely to increase the transcriptional
output of the tissue-specific modules.

Our data highlight the unpredictability of detoxification

gene induction. The mammalian receptors regulating
detoxification gene induction have very diverse ligand
structures, so it is difficult to define similarities between
inducers. However, large steroid molecules or small
lipophilic molecules or molecules containing aromatic rings
have been shown to be strong inducers (

Waxman, 1999

;

Goodwin et al., 2002

). While the insecticides tested here are

not inducers at biologically relevant concentrations it is
possible that insecticides that have not been tested may be
strong inducers.

It is tempting to contrast the lack of induction observed

in response to the insecticides we tested, to other toxins
that do induce insect metabolic genes. In particular, a
range of insecticidal plant secondary metabolites induce the
transcription of P450 genes in a variety of insect species
(

Cohen et al., 1992

;

Hung et al., 1995

;

Danielson et al.,

1997

;

Petersen et al., 2001

;

Li et al., 2003

;

Wen et al., 2003

;

Sasabe et al., 2004

). These induction responses have

possibly evolved to cope with the challenge posed by these
metabolites. At present, however, we are far from under-
standing exactly what it is that triggers an induction
response, making such comparisons between compounds
premature. In terms of insecticides, constitutive changes in
the transcription of metabolic genes have been the
predominant evolutionary response to insecticide exposure
(

Field et al., 1988, 1999

;

Ranson et al., 2001

;

Coleman

et al., 2002

;

Daborn et al., 2002

;

Vontas et al., 2002

;

Ortelli

et al., 2003

;

Bogwitz et al., 2005

). In managing field

resistance to insecticides this constitutive transcriptional
regulation needs to be understood. Induction by insecti-
cides will not provide a fast track to identify the metabolic
genes with the capacity to confer resistance. A better
understanding of the substrate specificity of the individual
detoxification enzymes is required.

ARTICLE IN PRESS

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

940

background image

Acknowledgments

We thank Rene´ Feyereisen for providing cytochrome

P450 clones for the microarray. This work is supported by
grants from The Australian Research Council (ARC)
through its funding of the Special Research Centre CESAR
(Centre for Environmental Stress and Adaptation Re-
search), and an ARC APD-CSIRO linkage fellowship
to PJD.

References

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.,
George, R.A., 2000. The genome sequence of Drosophila melanogaster.
Science 287, 2185–2195.

Amichot, M., Brun, A., Cuany, A., De Souza, G., Le Mouel, T., Bride,

J.M., Babault, M., Salaun, J.P., Rahmani, R., Berge, J.B., 1998.
Induction of cytochrome P450 activities in Drosophila melanogaster
strains susceptible or resistant to insecticides. Comp. Biochem. Physiol.
C—Pharmacol. Toxicol. Endocrinol. 121 (1–3), 311–319.

Amichot, A., Tares, S., Brun-Barale, A., Arthaud, L., Bride, J.M., Berge,

J.B., 2004. Point mutations associated with insecticide resistance in the
Drosophila cytochrome P450 Cyp6a2 enable DDT metabolism. Eur. J.
Biochem. 271, 1250–1257.

Berge, J.B., Feyereisen, R., Amichot, M., 1998. Cytochrome P450

monooxygenases and insecticide resistance in insects. Philos. Trans.
R. Soc. London B 353, 1701–1705.

Bogwitz, M.R., Chung, H., Magoc, L., Rigby, S., Wong, W., O’Keefe, M.,

McKenzie, J.A., Batterham, P., Daborn, P.J., 2005. Cyp12a4 confers
lufenuron resistance in a natural population of Drosophila melanoga-
ster. PNAS 102 (36), 12807–12812.

Brandt, A., Scharf, M.E., Pedra, J.H.F., Holmes, G., Dean, A., Kreitman,

M., Pittendrigh, B.R., 2002. Differential expression and induction of
two Drosophila cytochrome P450 genes near the Rst(2)DDT locus.
Insect Mol. Biol. 11, 337–341.

Brun, A., Cuany, A., Le Mouel, T., Berge, J., Amichot, M., 1996.

Inducibility of the Drosophila melanogaster cytochrome P450 gene,
CYP6A2, by phenobarbital in insecticide susceptible or resistant
strains. Insect Biochem. Mol. Biol. 26 (7), 697–703.

Carino, F.A., Koener, J.F., Plapp, F.W.J., Feyereisen, R., 1992.

Expression of the cytochrome P450 gene Cyp6A1 in the house fly,
Musca domestica. ACS Symp. Ser. 505, 31–40.

Chen, J., Raymond, K., 2006. Roles of rifampicin in drug–drug

interactions: underlying molecular mechanisms involved the nuclear
pregnane X receptor. Ann. Clin. Microbiol. Antimicrob. 15, 3.

Cohen, M.B., Schuler, M.A., Berenbaum, M.R., 1992. A host-inducible

cytochrome P450 from a host-specific caterpillar: molecular cloning
and evolution. PNAS 89, 10920–10924.

Coleman, M., Vontas, J.G., Hemingway, J., 2002. Molecular character-

ization of the amplified aldehyde oxidase from insecticide resistant
Culex quinquefasciatus. Eur. J. Biochem. 269, 768–779.

Daborn, P.J., Yen, J.L., Bogwitz, M.R., Le Goff, G., Feil, E., Jeffers, S.,

Tijet, N., Perry, T., Heckel, D., Batterham, P., Feyereisen, R., Wilson,
T.G., ffrench-Constant, R.H., 2002. A single P450 allele associated
with insecticide resistance in Drosophila. Science 297, 2253–2256.

Danielson, P.B., MacIntyre, R.J., Fogleman, J.C., 1997. Molecular

cloning of a family of xenobiotic-inducible drosophilid cytochrome
P450s: evidence for involvement in host–plant allelochemical resis-
tance. PNAS 94, 10797–10802.

Denison, M.S., Nagy, S.R., 2003. Activation of the aryl hydrocarbon

receptor by structurally diverse exogenous and endogenous chemicals.
Annu. Rev. Pharmacol. Toxicol. 43, 309–334.

Dunkov, B.C., Guzov, V.M., Mocelin, G., Shotkoski, F., Brun, A.,

Amichot, M., ffrench-Constant, R.H., Feyereisen, R., 1997. The
Drosophila cytochrome P450 gene Cyp6a2: structure, localization,

heterologous expression, and induction by phenobarbital. DNA Cell
Biol. 16 (11), 1345–1356.

Ellinger-Ziegelbauer, H., Stuart, B., Wahle, B., Bomann, W., Ahr, H.J.,

2005. Comparison of the expression profiles induced by genotoxic and
nongenotoxic carcinogens in the rat liver. Mutat. Res. 575, 61–84.

Field, L.M., Devonshire, A.L., Forde, B.G., 1988. Molecular evidence

that insecticide resistance in peach–potato aphids (Myzus persicae
Sulz.) results from amplification of an esterase gene. Biochem. J. 251,
309–312.

Field, L.M., Blackman, R.L., Tyler-Smith, C., Devonshire, A.L., 1999.

Relationship between amount of esterase and gene copy number in
insecticide-resistant Myzus persicae (Sulzer). Biochem. J. 339, 737–742.

Goadstuff, T., Dreano, Y., Guillos, B., Menez, J.-F., Berthou, F., 1996.

Induction of liver and kidney CYP1A1/CYP1A2 by caffeine in rat.
Biochem. Pharmcol. 52, 1915–1919.

Goodwin, B., Redinbo, M.R., Kliewer, S.A., 2002. Regulation of CYP3A

gene transcription by the pregnane X receptor. Annu. Rev. Pharm.
Toxicol. 42, 1–23.

Honkakoski, P., Zelko, I., Sueyoshi, T., Negishi, M., 1998. The nuclear

orphan receptor CAR-retinoid X receptor heterodimer activates the
phenobarbital-responsive enhancer module of the CYP2B gene. Mol.
Cell Biol. 18, 5652–5658.

Hung, C.-F., Harrison, T.L., Berenbaum, M.R., Schuler, M.A., 1995.

CYP6B3: a second furanocoumarin-inducible cytochrome P450
expressed in Papilio polyxenes. Insect Mol. Biol. 4, 149–160.

Jensen, H.R., Scott, I.M., Sims, S., Trudeau, V.L., Arnason, J.T., 2006.

Gene expression profiles of Drosophila melanogaster exposed to an
insecticidal extract of Piper nigrum. J. Agric. Food Chem. 54,
1289–1295.

King-Jones, K., Horner, M.A., Lam, G., Thummel, C.S., 2006. The

DHR96 nuclear receptor regulates xenobiotic responses in Drosophila.
Cell Metab. 4 (1), 37–48.

Kodama, S., Negishi, M., 2006. Phenobarbital confers its diverse effects

by activating the orphan nuclear receptor car. Drug Metab. Rev. 38
(1–2), 75–87.

Le Goff, G., Boundy, S., Daborn, P.J., Yen, J.L., Sofer, L., Lind, R.,

Sabourault, C., Madi-Ravazzi, L., ffrench-Constant, R.H., 2003.
Microarray analysis of cytochrome P450 mediated insecticide resis-
tance in Drosophila. Insect Biochem. Mol. Biol. 33 (7), 701–708.

Le Goff, G., Hilliou, F., Siegfried, B.D., Boundy, S., Wajnberg, E., Sofer,

L., Audant, P., ffrench-Constant, R.H., Feyereisen, R., 2006.
Xenobiotic response in Drosophila melanogaster: sex dependence of
P450 and GST gene induction. Insect Biochem. Mol. Biol. 36 (8),
674–682.

Li, X., Baudry, J., Berenbaum, M.R., Schuler, M.A., 2003. Structural and

functional divergence of insect CYP6B proteins: from specialist to
generalist cytochrome P450. PNAS 101, 2939–2944.

Luo, G., Guenthner, T., Gan, L.S., Humphreys, W.G., 2004. CYP3A4

induction by xenobiotics: biochemistry, experimental methods and
impact on drug discovery and development. Curr. Drug Metab. 5,
483–505.

Maitra, S., Dombrowski, S.M., Waters, L.C., Ganguly, R., 1996. Three

second chromosome-linked clustered Cyp6 genes show differential
constitutive and barbital-induced expression in DDT-resistant and
susceptible strains of Drosophila melanogaster. Gene 180 (1–2),
165–171.

Meyer, U.A., Hoffmann, K., 1999. Phenobarbital-mediated changes in

gene expression in the liver. Drug Metab. Rev. 31 (2), 365–373.

Newcomb, R.D., Campbell, P.M., Ollis, D.L., Cheah, E., Russell, R.J.,

Oakeshott, J.G., 1997. A single amino acid substitution converts a
carboxylesterase to an organophosphorous hydrolase and confers
insecticide resistance on a blowfly. Proc. Natl. Acad. Sci. USA 94,
7464–7468.

Ogino, Y., Itakura, T., Mitsuo, R., Sato, M., 1998. Induction of two forms

of eel cytochrome P450 1A genes by 3-methylcholanthrene. Mar.
Biotechnol. 1, 342–345.

Ortelli, F., Rossiter, L.C., Vontas, J., Ranson, H., Hemingway, J., 2003.

Heterologous expression of four glutathione transferase genes

ARTICLE IN PRESS

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

941

background image

genetically linked to a major insecticide-resistance locus from the
malaria vector Anopheles gambiae. Biochem. J. 373, 957–963.

Petersen, R.A., Zangerl, A.R., Berenbaum, M.R., Schuler, M.A., 2001.

Expression of CYP6B1 and CYP6B3 cytochrome P450 monoxy-
genases and furanocoumarin metabolism in different tissues of Papiloi
polyxenes (Lepidoptera: Papilionidae). Insect Biochem. Mol. Biol. 31,
679–690.

Ranson, H., Rossiter, L.C., Ortelli, F., Jensen, B., Wang, X., Roth, C.W.,

Collins, F.H., 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.

Sasabe, M., Wen, Z., Berenbaum, M.R., Schuler, M.A., 2004. Molecular

analysis of Cyp312a1, a novel cytochrome P450 involved in
metabolism of plant allelochemicals (furanocoumarins) and insecti-
cides (cypermethrin) in Helicoverpa zea. Gene 338, 163–175.

Smyth, G.K., Speed, T.P., 2003. Normalization of cDNA microarray

data. Methods 31, 265–273.

Sun, W., Margam, V.M., Sun, L., Buczkowski, G., Bennett, G.W.,

Schemerhorn, B., Muir, W.M., Pittendrigh, B.R., 2006. Genome-wide
analysis of phenobarbital-inducible genes in Drosophila melanogaster.
Insect Mol. Biol. 15 (4), 455–464.

Tang, A.H., Tu, C.D., 1994. Biochemical characterization of Drosophila

glutathione-S-transferases D1 and D21. J. Biol. Chem. 269, 27876–27884.

Tautz, D., Pfeifle, C., 1989. A non-radioactive in situ hybridization

method for the localization of specific RNAs in Drosophila embryos
reveals translational control of the segmentation gene hunchback.
Chromosoma 98, 81–85.

Vontas, J.G., Small, G.J., Nikou, D.C., Ranson, H., Hemingway, J., 2002.

Purification, molecular cloning and heterologous expression of a
glutathione-S-transferase involved in insecticide resistance from the
rice brown planthopper, Nilaparvata lugens. Biochem. J. 362, 329–337.

Waxman, D.J., 1999. P450s gene induction by structurally diverse

xenochemicals: central role of nuclear receptors Car, PXR and PPAR.
Arch. Biochem. Biophys. 369, 11–23.

Wen, Z., Pan, L., Berenbaum, M.R., Schuler, M.A., 2003. Metabolism of

linear and angular furanocoumarins by Papilio polyxenex CYP6B1 co-
expressed with NADPH cytochrome P450 reductase. Insect Biochem.
Mol. Biol. 33, 937–947.

Whitlock Jr., J.P., 1986. The regulation of cytochrome P-450 gene

expression. Annu. Rev. Pharmacol. Toxicol. 26, 333–369.

Whitlock Jr., J.P., 1999. Induction of cytochrome P4501A1. Annu. Rev.

Pharmacol. Toxicol. 39, 103–125.

Wilson, T.G., 2001. Resistance of Drosophila to Toxins. Annu. Rev.

Entomol. 46, 545–571.

Wilson, T.G., 2005. Drosophila: sentinel of environmental toxins. Integr.

Comp. Biol. 45, 127–136.

ARTICLE IN PRESS

L. Willoughby et al. / Insect Biochemistry and Molecular Biology 36 (2006) 934–942

942


Document Outline


Wyszukiwarka

Podobne podstrony:
Induction of two cytochrome P450 genes, Cyp6a2 and Cyp6a8 of Drosophila melanogaster by caffeine
The effects of plant flavonoids on mammalian cells implication for inflammation, heart disease, and
Eurocode 3 Part 1 3 2006 UK NA Design of steel structures General rules Supplementary rules for
Correlates of Sleep and Waking in Drosophila melanogaster
Mapowanie genów na przykładzie Drosophila melanogaster(1)
comparison of PRINCE2 against PMBOK
Comparison of Human Language and Animal Communication
A Comparison of two Poems?out Soldiers Leaving Britain
A Comparison of the Fight Scene in?t 3 of Shakespeare's Pl (2)
Drosophila melanogaster evg
Comparison of the Russians and Bosnians
43 597 609 Comparison of Thermal Fatique Behaviour of Plasma Nitriding
Comparision of vp;atile composition of cooperage oak wood
1 3 16 Comparison of Different Characteristics of Modern Hot Work Tool Steels
Instrukcja do hodowlii Drosophila melanogaster, UG, SEM3, GENETYKA

więcej podobnych podstron