SlideShare a Scribd company logo
1 of 6
Download to read offline
Exploring the biochemical properties and remediation
applications of the unusual explosive-degrading
P450 system XplA/B
Rosamond G. Jackson*, Elizabeth L. Rylott*, Diane Fournier†
, Jalal Hawari†
, and Neil C. Bruce*‡
*Center for Novel Agricultural Products, Department of Biology, University of York, P.O. Box 373, York YO10 5YW, United Kingdom; and †Biotechnology
Research Institute, National Research Council of Canada, Montreal, QC, Canada H4P 2R2
Edited by May R. Berenbaum, University of Illinois at Urbana–Champaign, Urbana, IL, and approved September 10, 2007 (received for review May 31, 2007)
Widespread contamination of land and groundwater has resulted
from the use, manufacture, and storage of the military explosive
hexa-hydro-1,3,5-trinitro-1,3,5-triazine (RDX). This contamination
has led to a requirement for a sustainable, low-cost method to
remediate this problem. Here, we present the characterization of
an unusual microbial P450 system able to degrade RDX, consisting
of flavodoxin reductase XplB and fused flavodoxin-cytochrome
P450 XplA. The affinity of XplA for the xenobiotic compound RDX
is high (Kd ‫؍‬ 58 ␮M) and comparable with the Km of other P450s
toward their natural substrates (ranging from 1 to 500 ␮M). The
maximum turnover (kcat) is 4.44 per s, only 10-fold less than the
fastest self-sufficient P450 reported, BM3. Interestingly, the pres-
ence of oxygen determines the final products of RDX degradation,
demonstrating that the degradation chemistry is flexible, but both
pathways result in ring cleavage and release of nitrite. Carbon
monoxide inhibition is weak and yet the nitroaromatic explosive
2,4,6-trinitrotoluene (TNT) is a potent inhibitor. To test the efficacy
of this system for the remediation of groundwater, transgenic
Arabidopsis plants expressing both xplA and xplB were generated.
They are able to remove saturating levels of RDX from liquid
culture and soil leachate at rates significantly faster than those of
untransformed plants and xplA-only transgenic lines, demonstrat-
ing the applicability of this system for the phytoremediation of
RDX-contaminated sites.
cytochrome P450 ͉ phytoremediation ͉ hexa-hydro-1,3,5-trinitro-
1,3,5-triazine
Cytochrome P450s catalyze a diverse range of chemical reac-
tions including hydroxylation, epoxidation, demethylation,
dehalogenation, desaturation, and isomerization (1). As a con-
sequence, P450s are involved in a host of metabolic pathways. In
eukaryotes and prokaryotes, they catalyze critical steps in the
biosynthesis of key metabolites such as steroids and vitamins (2),
fatty acids (3), and lignin (4). P450s have also been established
as detoxification enzymes with activity toward a range of xeno-
biotics. It was thus in keeping with this role that xplA, which was
isolated from Rhodococcus rhodochrous 11Y by growth on the
explosive hexa-hydro-1,3,5-trinitro-1,3,5-triazine (RDX) as sole
nitrogen source, was found to encode a P450 (5). The synthetic
N-NO2 bond of the RDX molecule, which is rare in nature, was
accommodated by this enzyme. Further analysis revealed an
unusual arrangement of subunits that contribute to the different
steps in P450 catalysis.
Cytochrome P450s, as heme-containing enzymes, require
reduction of the heme to activate the catalytic center, a process
involving supply of electrons from NAD(P)H to the P450 via
partnering enzymes. The nature of the redox partners varies.
Class one P450s include bacterial and mitochondrial enzymes
that use an FAD-containing ferredoxin reductase-like protein
and an iron-sulfur ferredoxin-like protein. Class two P450s are
usually bound to the endoplasmic reticulum membrane along
with the partnering enzyme, NADPH cytochrome P450 reduc-
tase, which contains FAD and FMN domains. Although the
P450s and redox partners predominantly exist as separate
polypeptides, examples have come to light where the three
catalytic domains required for activity are fused together, e.g.,
the Bacillal BM3 (6), Rhodococcal RhF (7, 8), and the fungal
CYP505A1 (8). In all classes, despite the variety of forms,
NADPH is usually the source of the electrons, and three electron
transfer domains are involved. With XplA, a different arrange-
ment of subunits is seen, with the second electron transfer step,
a flavodoxin domain, fused to the P450 domain (9). The
organization of the domains is also unusual, with the flavodoxin
domain fused to the N terminus of the P450. The first electron
transfer step has been postulated to be encoded by a reductase,
xplB, adjacent to xplA in the R. rhodochrous genome. XplB has
homology to adrenodoxin reductase (5), which transfers elec-
trons from NADPH to adrenodoxin in a synthetically fused P450
(10), and also transfers electrons to flavodoxin (11).
Interest in XplA and XplB has arisen after the contamination
of land and groundwater with RDX as a result of the widespread
manufacture, use, and disposal of munitions. This contamination
is of concern as RDX is toxic to all classes of organisms tested,
and the Environmental Protection Agency (EPA) classifies
RDX as a priority pollutant. Contamination on military training
ranges is of particular concern. For example, the use of RDX has
been restricted by the EPA at the Massachusetts Military
Reservation of Cape Cod where RDX contamination is threat-
ening drinking water sources (12).
Microorganisms present in soil heavily contaminated with ex-
plosives have been found to degrade RDX, but do not possess
sufficient biomass or metabolic activity to degrade this compound
before it leaches through soils polluting groundwater. Interestingly,
to date, xplA and xplB have been found only in Rhodococcus and
related bacteria isolated from RDX-contaminated soil, suggesting
that the RDX-degrading ability of XplA may have evolved under
this selective pressure. XplA has been recombinantly expressed and
shown to degrade RDX in vitro with a surrogate reductase (9). In
this article, the activity of XplA with its native reductase XplB
shows the ability of the proteins to work as efficient partners to
degrade RDX. Further characterization is undertaken along with
a detailed analysis of the RDX breakdown pathway under anaer-
obic and aerobic conditions. We have previously demonstrated that
expression of XplA in Arabidopsis confers both the ability to remove
Author contributions: R.G.J., E.L.R., J.H., and N.C.B. designed research; R.G.J., E.L.R., and
D.F. performed research; R.G.J., E.L.R., D.F., J.H., and N.C.B. analyzed data; and R.G.J., E.L.R.,
J.H., and N.C.B. wrote the paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
Abbreviations: RDX, hexa-hydro-1,3,5-trinitro-1,3,5-triazine; TNT, trinitrotoluene; NDAB,
4-nitro-2,4, diazabutanal; MEDINA, methylenedinitramine.
‡To whom correspondence should be addressed. E-mail: ncb5@york.ac.uk.
This article contains supporting information online at www.pnas.org/cgi/content/full/
0705110104/DC1.
© 2007 by The National Academy of Sciences of the USA
16822–16827 ͉ PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 www.pnas.org͞cgi͞doi͞10.1073͞pnas.0705110104
RDX from liquid culture and resistance to the phytotoxic effects of
RDX; however, this activity relies on support from endogenous
plant reductases (9). Here, the expression of both xplA and xplB in
Arabidopsis enabled the rapid removal of RDX from liquid culture
and soil leachate, a rate significantly faster than for plants express-
ing xplA alone. These results demonstrate that this technology can
be applied to remediate RDX from contaminated sites.
Results and Discussion
Optimizing Expression and Assay Conditions. The purification of
XplA to homogeneity has been described (9). Soluble expression
and purification of XplB was achieved by using a pGEX vector
where GST is fused to the N terminus of XplB (Fig. 1A).
Cleavage of the GST resulted in the loss of the 72-kDa XplB
band from SDS/PAGE gels, the appearance of several smaller
bands, and Ͻ5% protein recovery, suggesting insolubility of any
cleaved protein (data not shown). The soluble, fused XplB was
able to transfer electrons to XplA for the degradation of RDX,
whereas GST alone had no such activity (Fig. 1B), therefore
fused GST-XplB was used in subsequent studies. Fig. 1C shows
that a 2-fold molar excess of XplB to XplA was the ratio at which
XplA became limiting (as measured by flavin levels). Con-
versely, a 10-fold molar excess of XplA to XplB was the ratio at
which XplB became limiting (Fig. 1D). At lower ratios, the
concentration of both enzymes influenced the rate of activity
toward RDX, as expected for a second-order reaction, suggest-
ing that collision of the two subunits is a major rate-limiting
contribution. The optimal pH for RDX degradation was pH
6.5–7.0 irrespective of the buffer used, and potassium phosphate
at pH 6.8 was used subsequently. Activity was not significantly
affected by ionic strength between 0 and 100 mM NaCl, but
above 100 mM, an inhibitory effect of sodium chloride was seen
(data not shown), thus NaCl was omitted from further assays.
Spectral Analysis of XplA and XplB. XplA had previously been
shown to contain a classic P450 heme and to be able to bind
carbon monoxide when reduced, producing a spectral shift to 450
nm, and a flavin binding domain (9), but the nature of the flavin
was not determined. Purified XplB also possessed a classic flavin
absorbance spectra, and release of the flavin from XplA and
XplB by boiling and subsequent analysis by HPLC showed XplA
to contain predominantly FMN and XplB FAD [supporting
information (SI) Fig. 8].
Reduction of XplA by sodium dithionite causes a character-
istic decrease in the heme 420-nm peak and a shift to a maximum
of 389 nm. The dominating heme spectrum masked any flavin
absorbance. Six nanomoles of XplA was fully reduced by be-
tween 8 and 10 nmol of sodium dithionite, suggesting the
majority of XplA has flavin bound (full reduction of heme and
flavin would take 9 nmol) (Fig. 2 A and B).
On reduction of XplB by NADPH, the flavin absorbance was
completely bleached. No changes were observed between 550 and
600 nm, consistent with a lack of the stabilized semiquinone form,
and implying a two-electron reduction (Fig. 2C). For XplB, the
concentration of NADPH required for full reduction was approx-
Fig. 1. Recombinant expression of XplB and assay of activity with XplA. (A) Protein purification on 10% SDS/PAGE. Lane 1, molecular weight markers; lane 2,
solubilized recombinant protein; lane 3, affinity-purified XplB. (B) Aerobic activity of GST-XplB (0.26 ␮M; filled symbol) and GST (0.71 ␮M; open symbol) with
XplA (0.27 ␮M). (C) Anaerobic activity of 60 nM XplA with various ratios of XplB. (D) Anaerobic activity of 60 nM XplB with various ratios of XplA. Values are
the mean Ϯ SD of triplicates.
Fig. 2. Spectral analysis of the reduction of XplA and XplB. (A) UV-visible
spectra of 6 nmol of XplA (determined by protein concentration), titrated with
the indicated amount of sodium dithionite (nmol) under anaerobic condi-
tions. (B) Difference spectra of the sodium dithionite titration of XplA, gen-
erated by subtraction of the original XplA spectrum from those with sodium
dithionite. (C) Anaerobic titration of 37 nmol XplB (by protein) with the
indicated amount of NADPH (nmol).
Jackson et al. PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 ͉ 16823
APPLIEDBIOLOGICAL
SCIENCES
imately half the protein concentration, indicating that half of the
XplB has flavin bound. Addition of flavin during purification and
varying growth conditions did not improve this level. XplB was not
readily reduced by NADH nor was NADH a successful electron
donor for RDX degradation (data not shown).
RDX Binding and Activity of XplA and XplB. The binding of RDX to
XplA has been examined in two ways. Titration of XplA with RDX
in an anaerobic environment revealed the low spin to high spin
change in the heme often seen on substrate binding and a binding
affinity (Kd) of 57.9 Ϯ 2.8 ␮M for RDX (Fig. 3 A and B). The Km
was calculated to be 83.7 Ϯ 17.8 ␮M, and the maximum turnover
(kcat) of the enzyme was 4.44 Ϯ 0.46 per s using a saturating ratio
of XplB (2-fold molar excess) and anaerobic conditions (Fig. 3C).
The Km and Kd values are similar to each other, and a turnover of
4.44 per s is comparable with the Pseudomonal P450cam toward its
natural substrate camphor (27 per s) (13), perhaps surprising given
the xenobiotic nature of RDX. The number of electron transfer
steps involved in a P450 system often prevent the turnover from
being substantially faster (1).
RDX Breakdown Pathways. Initially, the breakdown products of
RDX degradation were analyzed anaerobically to determine the
amount of NADPH required without losses to uncoupled cleav-
age of oxygen. Formaldehyde and nitrite were measured directly,
whereas RDX and other products were assayed after freezing
the samples. It was surprising to find ratios of nitrite to RDX,
after 70 min, of 1.4:1.0 and formaldehyde to RDX of 1.96:1.00
as it was previously thought that the breakdown pathway would
follow that proposed by Fournier et al. (14) with 2:1 nitrite and
1:1 formaldehyde and production of 4-nitro-2,4, diazabutanal
(NDAB) (Fig. 4A). NDAB was not detected; however, the RDX
breakdown product methylenedinitramine (MEDINA) was,
reaching a ratio of 0.68:1 after 70 min and then decreasing,
possibly because of instability in water (15). The ratio of NADPH
to RDX degraded was 1.26:1.00 at 70 min, suggesting these
compounds are tightly coupled.
When the breakdown pathway under aerobic conditions was
examined, a different picture arose with an increase in the ratio
of nitrite to RDX to 2.49:1.00 after 130 min and a decrease in
the formaldehyde to RDX ratio to 1.4:1.0 (Fig. 4B). NDAB was
detected, and levels continued to increase throughout the time
course; MEDINA was not detected.
The mechanism for denitration of RDX by XplA is not yet
fully known; however, once denitration and hydration have
occurred, whether under aerobic or anaerobic conditions, the
resulting imine intermediate would be highly unstable in water
and spontaneously decompose (16). Under anaerobic condi-
tions, mono-denitration and mono-hydration of RDX followed
by ring cleavage would produce MEDINA (Fig. 5, route A).
Under aerobic conditions, it is proposed that RDX is subjected
to di-denitration–di-hydration before ring cleavage. This mech-
anism, leading to the formation of NDAB (see Fig. 5, route B),
has been described (14). Degradation of RDX with radiolabeled
Fig. 4. Mass balance of RDX breakdown. (A) Anaerobic degradation of RDX
(60 nM XplA and XplB) and analyses were carried out as in Materials and
Methods. Controls with boiled XplA and XplB (not shown) contained the
following levels of analytes (nmol/ml) over the time course: RDX, 100 Ϯ 4.3;
nitrite, 0 Ϯ 3.9; formaldehyde, 0 Ϯ 6.3; and MEDINA, 0–6.3. (B) Aerobic
degradation of RDX. Reactions contained 90 nM XplA and XplB. Controls with
boiled XplA and XplB (not shown) contained RDX (100 Ϯ 2.5), nitrite (0 Ϯ 2),
formaldehyde (0 Ϯ 16), and NDAB (0–2.6) over the time course. Values are the
mean Ϯ SD of triplicates.
Fig. 3. XplA and RDX binding. All analyses were carried out anaerobically.
(A) UV-visible spectral changes in 4.5 nmol XplA on the addition of 0–150 nmol
of RDX (in DMSO). (Inset) Difference spectra generated by subtraction of
nonbound XplA spectrum from the spectra of XplA with RDX. (B) Plot of A391 Ϫ
A425 generated from the difference spectra against RDX concentration. (C)
Initial rates of substrate use at a range of RDX concentrations plotted against
RDX concentration with 60 nM XplA and 120 nM XplB. Values are the mean Ϯ
SD of triplicates.
16824 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0705110104 Jackson et al.
oxygen and another P450 system suggests that direct hydroxy-
lation of the RDX molecule by XplA is unlikely (17). The role
oxygen plays in the mechanism of RDX degradation by XplA,
either within the enzyme or within the reaction environment, is
therefore currently unclear. The pathway presented in Fig. 5 is
based on the analysis of the final degradation products and the
proposed intermediates by analogy from previously published
work (14, 17). Additional work involving labeled RDX and
deuterated solvent may allow precise details of the mechanism
to be determined.
Other examples of P450s catalyzing different reactions depend-
ing on the presence of oxygen are known. For example, under
anaerobic conditions human CYP2A6 and CYP101 both catalyze
reductive reactions toward halogenated substrates, whereas under
aerobic conditions, CYP2A6 catalyzes dehalogenation, (18) and
CYP101 catalyzes a hydroxylation reaction (19).
XplA Activity with a Wider Range of Substrates and Inhibitors. RDX
as a synthetic compound may not be the native substrate for
XplA, so the activity toward a range of other substrates was
tested. The related nitramine explosive, octahydro-1,3,5,7-
tetranitro-1,3,5,7-tetrazocine (HMX), was not transformed at
the aqueous solubility limit for HMX of 15 ␮M, even after
several hours of incubation. Trace levels of XplA activity toward
the nitroaromatic explosive trinitrotoluene (TNT) were found
(turnover was Ͻ1% per h at 100 ␮M), but transformation
products could not be identified. The heme spectrum of XplA
was not altered by the presence of TNT. To test the ability of
XplA to perform classic P450 hydroxylation and demethylation
reactions, established P450 substrates were tested. No hydroxy-
lating activity was detected toward testosterone or paclitaxel, nor
demethylating activity toward 7-ethoxycoumarin or
ethoxyresorufin. However, oxidizing activity was detected to-
ward both methyl tolyl and methyl phenyl sulfides generating
sulfoxide products (data not shown).
Methyl tolyl sulfide was also shown to inhibit RDX catabolism,
as was the P450-specific inhibitor metyrapone. The strongest
inhibition was seen by TNT (Fig. 6), whereas the rate of carbon
monoxide inhibition was less than expected, given the usual high
affinity of P450s for carbon monoxide. This inhibition was not
increased by a higher concentration of carbon monoxide. XplA
may have a lower affinity for CO in the presence of RDX, as seen
with other substrates of P450s (20).
Application of the Enzymes for Phytodegradation of RDX. It has
previously been shown that transgenic plants expressing xplA can
degrade RDX (9), but this activity relies on the availability of
endogenous plant reductases, which may be limiting. Thus, xplB
was transformed into Arabidopsis, and transgenic plant lines
expressing both xplA and xplB were generated. A previously
characterized plant line expressing xplA (XplA-10) (9) was used
to produce five independently transformed lines expressing xplB.
In addition, five independently transformed lines expressing only
xplB were characterized. The results presented here are from
plants homozygous for these transgenes. Quantitative analysis by
real-time PCR showed that xplA was expressed in all five XplAB
lines (Fig. 7A), although expression levels varied from that of the
original parental line, XplA-10. A range in the level of xplB
transcript was seen; with line XplAB-27 exhibiting the highest
levels of both xplA and xplB transcripts (Fig. 7A). These lines
were grown in axenic liquid culture to determine rates of RDX
uptake. As reported (9), line XplA-10 removed all 180 ␮M RDX
from the medium within 5 days; however, the XplAB lines
removed the RDX significantly faster, with lines XplAB-2 and
XplAB-27 removing Ͼ50% of the RDX within 4 h, 30 times
faster than the XplA-10 line (Fig. 7B). The xplB-only lines had
uptake rates similar to those of untransformed, wild-type plants
(Fig. 7C). NDAB and MEDINA levels were not measured in the
liquid culture or soil-grown plants. Liquid culture-grown plants
and water-saturated soil-grown roots are likely to be hypoxic. It
is possible from our characterization that, depending on oxygen
availability, either NDAB or MEDINA is produced by xplA–
expressing plants.
To investigate the ability of the XplAB lines to reduce levels
of RDX in contaminated ground water, 8-week-old plants were
watered with 180 ␮M RDX. After 1 week, the soil was flushed
with water and the level of RDX in the soil leachate was
measured. After this time, the level of RDX in the leachate from
untransformed, wild-type plants was unaltered, whereas leachate
from the XplA-10 line had decreased by 25%. The RDX in the
leachate from lines XplAB-2 and XplAB27 had decreased by
90–97% (Fig. 7D).
Conclusions
XplA and XplB constitute a novel P450 redox system and
together efficiently degrade the xenobiotic RDX. Degradation
follows two different routes dependent on the presence of
Fig. 5. Proposed degradation pathway of RDX under aerobic and anaerobic
conditions. Ring cleavage occurs at ab under anaerobic conditions (route A)
and at cb under aerobic conditions (route B). Compounds in brackets are
hypothetical, and the mechanisms are based on detection of nitrite and
formaldehyde, the final products, and analogy with previous work (14, 17).
Fig. 6. Activity of XplA and XplB with inhibitors. Standard anaerobic activity
assays using 100 ␮M RDX as substrate were carried out in the presence of 100
␮M metyrapone or 100 ␮M TNT, and activity was compared with that with only
RDX (100%). Carbon monoxide (CO) and methyl tolyl sulfide (MTS) inhibition
was tested in standard aerobic conditions using 100 ␮M RDX as substrate.
Values are the mean Ϯ SD of triplicates.
Jackson et al. PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 ͉ 16825
APPLIEDBIOLOGICAL
SCIENCES
oxygen. One mole of MEDINA and nitrite are the dominant
products anaerobically, whileas 1 mol of NDAB and 2 mol of
nitrite are produced in aerobic conditions. With both pathways
resulting in ring cleavage and nitrite release, applications of
these enzymes for bioremediation look encouraging. One of the
biggest concerns of RDX as a pollutant is that it migrates readily
through soil into the groundwater and subsequently contami-
nates drinking water supplies. Here, we show that Arabidopsis
plants expressing xplA and xplB have the ability to effectively
remove RDX from the soil leachate. The studies here illustrate
that these genes, or possibly an xplA–xplB gene fusion, could be
engineered into plant species suited to growth on military
training ranges and used to remediate RDX.
Materials and Methods
RDX was supplied by the Defense Science and Technology
Laboratory at the U.K. Ministry of Defense (Fort Halstead,
Kent, U.K.). MEDINA and NDAB were provided by Ron
Spanggord (SRI International, Menlo Park, CA).
XplA and XplB Expression and Purification. The xplA gene was cloned
and expressed as described (9). The xplB gene was amplified from
pHSX1 by PCR using primers containing overhanging BamH1
sites, 5Ј-GGATCCGACATCATGAGTGAAGTGGAC and 3Ј-
GGATCCGCAGACCGATTCGGCCGG and ligated into
pGEX2T (Merck Chemicals, Nottingham, U.K.), which engineers
an N-terminal GST domain. The construct was sequenced, trans-
formed into Escherichia coli BL21 (DE3) and grown at 20°C in
Luria broth containing 100 ␮g/ml carbenicillin to OD600 (1.0), 1
mM isopropyl ␤-d-thiogalactopyranoside was added, and the cul-
ture was grown for an additional 24 h. The cell pellet was resus-
pended in 10 ml of PBS (140 mM NaCl/15 mM KH2PO4/80 mM
Na2HPO4) containing 0.2 mM PMSF and lysed at 1,500 psi (1 psi ϭ
6.89 kPa) in a French Press (Thermo IEC). The lysate was
centrifuged (10,000 ϫ g for 15 min), and the soluble protein was
purified by using 200 ␮l of 50% glutathione-coupled Sepharose gel
(Amersham, Piscataway, NJ) according to the manufacturer’s
instructions and recovered in glutathione elution buffer (20 mM
reduced glutathione/100 mM Tris⅐HCl, pH 8.5/120 mM NaCl).
Protein assays were carried out with Coomassie Protein Assay
Reagent (Pierce, Rockford, IL) using BSA as reference. Proteins
were analyzed as described (21).
Flavin Content Determination. The flavin content of XplA and XplB
was determined after boiling for 20 min then centrifugation of the
precipitated protein and trichloroacetic acid precipitation (10% of
240 mg/ml), followed by centrifugation and sodium bicarbonate
neutralization. Both methods gave similar results for each protein.
The concentration of released flavin was used to measure XplA and
XplB concentration for activity assays. The released flavins were
identified by HPLC according to ref. 22 and compared with
commercially available standards for identification.
Spectral Analyses. Spectral analyses were performed on a spec-
trophotometer (v560; Jasco, Easton, MD) scanning between 800
and 300 nm at 400 nm/min. Anaerobic analyses were carried out
in an anaerobic chamber (Coy Laboratory Products, Ann Arbor,
MI). For all spectral analyses, enzyme concentrations quoted are
for sample protein. The XplA titration with sodium dithionite in
the anaerobic chamber followed desalting of the protein on a
PD-10 column (GE Healthcare) into 50 mM potassium phos-
phate, pH 6.8 (bubbled with nitrogen before equilibrating in the
chamber for 2 days). For RDX titration, RDX was dissolved in
DMSO and no more than 3 ␮l was added per ml.
Fig. 7. Characterization of XplAB and XplB transgenic Arabidopsis lines. (A) Expression of xplA and xplB in rosette leaves. Quantitative analysis was done by
real-time PCR of xplA and xplB transcript abundance in rosette leaves of the transgenic lines relative to line XplAB-2. ACTIN2 mRNA was used as an internal
reference. Results represent the mean of three independent RNA isolations measured in duplicate from the pooled rosette leaves of 10 plants Ϯ SE. (B and C)
Uptake of RDX from media by Arabidopsis seedlings. Results are the mean Ϯ SE of five replicate flasks, each containing 200 10-day-old seedlings. (D) Levels of
RDX in soil leachate from Arabidopsis plants watered with 180 ␮M RDX. Results are the mean Ϯ SE of five replicate pots. NPC, no plant control.
16826 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0705110104 Jackson et al.
XplA and XplB Activity Assays Toward RDX. Anaerobic conditions. XplA
and XplB were placed on ice in an anaerobic chamber in open
vials for 3 h for the palladium catalyst to remove oxygen. Buffer
was bubbled with nitrogen before equilibration in the chamber.
The reaction mixture contained 60 nM XplA and XplB in 50 mM
potassium phosphate (pH 6.8), 300 ␮M NADPH, and 100 ␮M
RDX in a total volume of 1 ml. Other reagents and solutions
were placed overnight in the anaerobic chamber with a nitrogen
atmosphere, before use. Oxygen levels were monitored with a
model 10 gas analyzer (Coy Laboratory Products) and main-
tained Ͻ1 ppm by using a 5% (vol/vol) hydrogen mix in the
presence of a palladium catalyst.
Aerobic conditions. The reaction mixture contained 175 nM XplA
and 150 nM XplB, 50 mM potassium phosphate (pH 6.8), 300
␮M NADPH, 100 ␮M RDX, 0.72 units Thermoanaerobium
brockii alcohol dehydrogenase (Sigma, St. Louis, MO), and 30 ␮l
isopropanol in a total volume of 1 ml. All reactions were
performed at room temperature (20°C), and assays were stopped
by the addition of 10% trichloroacetic acid (240 mg/ml) or, for
the mass balance experiments, 30-kDa cut-off spin columns
(Microcon YM-30; Amicon). The reaction was initiated by the
addition of RDX.
Analysis of Products. RDX removal was measured by using RP-
HPLC with a HPLC system (2695 Separations Module and 2996
Photodiode Array Detector; Waters) using a Techsphere C18
column (250 ϫ 4.6 mm) under isocratic conditions of 60% water
and 40% acetonitrile at a flow rate of 1 ml/min over 10 min. RDX
elution was monitored at 205 nm, and intergrations were per-
formed with Empower software.
Formaldehyde analysis was carried out according to Nash (23).
Nitrite analysis followed the method of Scheideler and Ninnemann
(24) terminated by removal of proteins with 30-kDa cut-off spin
columns. The concentration of NDAB and MEDINA was deter-
mined by using an HPLC system as reported (25).
Activity Assays Toward Other Substrates and with Inhibitors. Analysis
of octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine was as de-
scribed for RDX. Loss of TNT was analyzed by using a water
mobile phase of 50% MeCN, 50% water over 12 min with
monitoring at 230 nm. Paclitaxel was analyzed by using a
methanol and water gradient from 60–70% MeOH over 20 min
with monitoring at 230 nm. Testosterone was analyzed by HPLC
using a 58–62% MeOH gradient over 8 min. Activity toward
7-ethoxycoumarin was tested by following fluorescence of 7-hy-
droxy coumarin, with excitation at 350 nm and emission at 450
nm adapted from ref. 26 and TLC. Activity toward ethoxyresoru-
fin was determined fluorometrically with excitation at 530 nm
and emission at 590 nm, adapted from ref. 26. Carbon monoxide
inhibition was carried out aerobically by adding 100 or 500 ␮l per
ml of carbon monoxide-saturated buffer.
Plant Transformation Methods. The xplB gene was cloned into the
binary vector pART27 (27) under the control of the CaMV35S
promoter and ocs terminator, and transformed by Agrobacte-
rium-mediated floral dipping into wild-type and xplA-expressing
Arabidopsis thaliana, ecotype Columbia-0 as in ref. 9.
Liquid Culture and Soil Leachate Experiments. Liquid culture exper-
iments were performed as described (9). Soil leachate studies
were carried out on 6-week-old Arabidopsis plants grown under
180 ␮m⅐m2
⅐s Ϫ1
light in a 12-h photoperiod. Plants were grown
in pots containing 30 g of uncontaminated soil (Levingtons F2
compost), five plants per pot. Each pot was flooded with 50 ml
of 180 ␮M RDX, then 7 days later, flushed through with 50 ml
of water. The collected soil leachates were analyzed for RDX
content by using HPLC as described above.
This work was funded by the Strategic Environmental Research and
Development Program of the U.S. Department of Defense, the Bio-
technology and Biological Sciences Research Council, and the U.K.
Ministry of Defense.
1. Guengerich FP (2001) Chem Res Toxicol 14:611–650.
2. Prosser DE, Jones G (2004) Trends Biochem Sci 29:664–673.
3. Green AJ, Rivers SL, Cheeseman M, Reid GA, Quaroni LG, Macdonald ID,
Chapman SK, Munro AW (2001) J Biol Inorg Chem 6:523–533.
4. Schoch G, Goepfert S, Morant M, Hehn A, Meyer D, Ullmann P, Werck-
Reichhart D (2001) J Biol Chem 276:36566–36574.
5. Seth-Smith HMB, Rosser SJ, Basran A, Travis ER, Dabbs ER, Nicklin S, Bruce
NC (2002) Appl Environ Microbiol 68:4764–4771.
6. Munro AW, Lindsay JG, Coggins JR, Kelly SM, Price NC (1996) Biochim
Biophys Acta 1296:127–137.
7. Hunter DJ, Roberts GA, Ost TW, White JH, Muller S, Turner NJ, Flitsch SL,
Chapman SK (2005) FEBS Lett 579:2215–2220.
8. Nakayama N, Takemae A, Shoun H (1996) J Biochem (Tokyo) 119:435–440.
9. Rylott EL, Jackson RG, Edwards J, Womack GL, Seth-Smith HM, Rathbone
DA, Strand SE, Bruce NC (2006) Nat Biotechnol 24:216–219.
10. Cao PR, Bulow H, Dumas B, Bernhardt R (2000) Biochim Biophys Acta
1476:253–264.
11. Zollner A, Nogues I, Heinz A, Medina M, Gomez-Moreno C, Bernhardt R
(2004) Bioelectrochemistry 63:61–65.
12. Clausen J, Robb J, Curry D, Korte N (2004) Environ Pollut 129:13–21.
13. Peterson JA, Mock DM (1975) Adv Exp Med Biol 58:311–324.
14. Fournier D, Halasz A, Spain J, Fiurasek P, Hawari J (2002) Appl Environ
Microbiol 68:166–172.
15. Halasz A, Spain J, Paquet L, Beaulieu C, Hawari J (2002) Environ Sci Technol
36:633–638.
16. Melius CF (1990) in Chemistry and Physics of Energetic Materials: Thermochemical
Modeling, ed Bulusu SN (Kluwer, Dordrecht, The Netherlands), pp 21–49.
17. Bhushan B, Trott S, Spain JC, Halasz A, Paquet L, Hawari J (2003) Appl
Environ Microbiol 69:1347–1351.
18. Spracklin DK, Kharasch ED (1998) Drug Metab Dispos 26:605–607.
19. Koe GS, Vilker VL (2005) Biotechnol Prog 21:1119–1127.
20. Sevrioukova IF, Peterson JA (1995) Arch Biochem Biophys 317:397–404.
21. Sambrook J, Russell DW (2001) Molecular Cloning: A Laboratory Manual (Cold
Spring Harbor Lab Press, Cold Spring, NY).
22. Aliverti A, Curti B, Vanoni M (1999) in Flavoprotein Protocol, eds Chapman
SK, Reid GA (Humana, Totowa, NJ), pp 9–23.
23. Nash T (1953) Biochem J 55:416–421.
24. Scheideler L, Ninnemann H (1986) Anal Biochem 154:29–33.
25. Fournier D, Halasz A, Spain J, Spanggord RJ, Bottaro JC, Hawari J (2004) Appl
Environ Microbiol 70:1123–1128.
26. Werck-Reichhart D, Gabriac B, Teutsch H, Durst F (1990) Biochem J 270:729–735.
27. Gleave AP (1992) Plant Mol Biol 20:1203–1207.
Jackson et al. PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 ͉ 16827
APPLIEDBIOLOGICAL
SCIENCES

More Related Content

What's hot

PBL 490 Publication
PBL 490 PublicationPBL 490 Publication
PBL 490 PublicationJen Klabis
 
mBio-2015-Melnyk-
mBio-2015-Melnyk-mBio-2015-Melnyk-
mBio-2015-Melnyk-Ryan Melnyk
 
Structural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and Rbjl
Structural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and RbjlStructural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and Rbjl
Structural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and RbjlWard Coats
 
J. biol. chem. 1986-fushitani-8414-23
J. biol. chem. 1986-fushitani-8414-23J. biol. chem. 1986-fushitani-8414-23
J. biol. chem. 1986-fushitani-8414-23guestfebc32d
 
Radiation Protection : Phospholipase A
Radiation Protection : Phospholipase ARadiation Protection : Phospholipase A
Radiation Protection : Phospholipase ADmitri Popov
 
Voss et al. - 2006 - Identification and characterization of riproximin,
Voss et al. - 2006 - Identification and characterization of riproximin,Voss et al. - 2006 - Identification and characterization of riproximin,
Voss et al. - 2006 - Identification and characterization of riproximin,Cristina Voss
 
Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...
Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...
Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...IPN
 
Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...Keiji Takamoto
 
BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13
BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13
BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13Juan Barrera
 
IndianaCTSI_Seed2010_Jessany Maldonado
IndianaCTSI_Seed2010_Jessany MaldonadoIndianaCTSI_Seed2010_Jessany Maldonado
IndianaCTSI_Seed2010_Jessany MaldonadoJessany Maldonado
 
Dual-enzyme natural motors incorporating decontamination and propulsion capab...
Dual-enzyme natural motors incorporating decontamination and propulsion capab...Dual-enzyme natural motors incorporating decontamination and propulsion capab...
Dual-enzyme natural motors incorporating decontamination and propulsion capab...Michael Galarnyk
 
Biochemical engeenering J_2016
Biochemical engeenering J_2016Biochemical engeenering J_2016
Biochemical engeenering J_2016Ravindra kumar
 
Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...
Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...
Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...Debanjan Chatterjee
 
Dimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi and
Dimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi andDimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi and
Dimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi andrkkoiri
 

What's hot (20)

PBL 490 Publication
PBL 490 PublicationPBL 490 Publication
PBL 490 Publication
 
Shukla et al
Shukla et alShukla et al
Shukla et al
 
mBio-2015-Melnyk-
mBio-2015-Melnyk-mBio-2015-Melnyk-
mBio-2015-Melnyk-
 
Acta Crystalo D
Acta Crystalo DActa Crystalo D
Acta Crystalo D
 
Structural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and Rbjl
Structural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and RbjlStructural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and Rbjl
Structural Modeling of the Ptf1a C2 Sequence Bound to Mammalian Rbpj and Rbjl
 
J. biol. chem. 1986-fushitani-8414-23
J. biol. chem. 1986-fushitani-8414-23J. biol. chem. 1986-fushitani-8414-23
J. biol. chem. 1986-fushitani-8414-23
 
New world built with toxic chemical is coming up
New world built with toxic chemical is coming upNew world built with toxic chemical is coming up
New world built with toxic chemical is coming up
 
065771.full
065771.full065771.full
065771.full
 
Radiation Protection : Phospholipase A
Radiation Protection : Phospholipase ARadiation Protection : Phospholipase A
Radiation Protection : Phospholipase A
 
Voss et al. - 2006 - Identification and characterization of riproximin,
Voss et al. - 2006 - Identification and characterization of riproximin,Voss et al. - 2006 - Identification and characterization of riproximin,
Voss et al. - 2006 - Identification and characterization of riproximin,
 
Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...
Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...
Nitrate Reductase Complex of Escherichia coli K-12: Participation of Specific...
 
Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...Principles of RNA compaction : insights from equilibrium folding pathway of p...
Principles of RNA compaction : insights from equilibrium folding pathway of p...
 
BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13
BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13
BarreraBasnetDelgadoLamichhaneShifatuShrestha_Report2_4140_S13
 
IndianaCTSI_Seed2010_Jessany Maldonado
IndianaCTSI_Seed2010_Jessany MaldonadoIndianaCTSI_Seed2010_Jessany Maldonado
IndianaCTSI_Seed2010_Jessany Maldonado
 
Dual-enzyme natural motors incorporating decontamination and propulsion capab...
Dual-enzyme natural motors incorporating decontamination and propulsion capab...Dual-enzyme natural motors incorporating decontamination and propulsion capab...
Dual-enzyme natural motors incorporating decontamination and propulsion capab...
 
Biochemical engeenering J_2016
Biochemical engeenering J_2016Biochemical engeenering J_2016
Biochemical engeenering J_2016
 
2008,pce,gonugunta et al
2008,pce,gonugunta et al2008,pce,gonugunta et al
2008,pce,gonugunta et al
 
CMBDictyPaper
CMBDictyPaperCMBDictyPaper
CMBDictyPaper
 
Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...
Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...
Bacteria Induced Cryptic Meroterpenoid Pathway in Pathogenic Aspergillus fumi...
 
Dimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi and
Dimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi andDimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi and
Dimethyl sulfoxide activates tumorn ecrosisfactor p53 mediated apoptosi and
 

Similar to PNAS paper 2007

mBio-2014-Melnyk-
mBio-2014-Melnyk-mBio-2014-Melnyk-
mBio-2014-Melnyk-Ryan Melnyk
 
Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164Dinesh Barawkar
 
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...John Clarkson
 
P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...
P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...
P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...David W. Salzman
 
microbial community structure of polluted river sediments
microbial community structure of polluted river sedimentsmicrobial community structure of polluted river sediments
microbial community structure of polluted river sedimentsEric Ariel Ben-David
 
Pol.j.environ.stud.vol.25.no.2.563 571
Pol.j.environ.stud.vol.25.no.2.563 571Pol.j.environ.stud.vol.25.no.2.563 571
Pol.j.environ.stud.vol.25.no.2.563 571Syed ShahZaib Shah
 
Role of cytochrome p450 in toxicity
Role of cytochrome p450 in toxicityRole of cytochrome p450 in toxicity
Role of cytochrome p450 in toxicityGhalia Nawal
 
10 nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...
10  nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...10  nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...
10 nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...Dheeraj Vasu
 
2016 RBC RETREAT POSTER TEMPLATE_SMP
2016 RBC RETREAT POSTER TEMPLATE_SMP2016 RBC RETREAT POSTER TEMPLATE_SMP
2016 RBC RETREAT POSTER TEMPLATE_SMPSagar M. Patel
 
3aff_monooxygensae.pdf
3aff_monooxygensae.pdf3aff_monooxygensae.pdf
3aff_monooxygensae.pdfssuserc7b94d
 
Quantitative phosphoproteomic analysis reveals common
Quantitative phosphoproteomic analysis reveals commonQuantitative phosphoproteomic analysis reveals common
Quantitative phosphoproteomic analysis reveals commonDiegoLujan14
 

Similar to PNAS paper 2007 (20)

mBio-2014-Melnyk-
mBio-2014-Melnyk-mBio-2014-Melnyk-
mBio-2014-Melnyk-
 
Os saline
Os salineOs saline
Os saline
 
Celis 2015, KpCld
Celis 2015, KpCldCelis 2015, KpCld
Celis 2015, KpCld
 
Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164Nucleic Acids Research (1995) 23, 159-164
Nucleic Acids Research (1995) 23, 159-164
 
acs.biomac.5b00657
acs.biomac.5b00657acs.biomac.5b00657
acs.biomac.5b00657
 
IPWE2016-000048
IPWE2016-000048IPWE2016-000048
IPWE2016-000048
 
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
Probing the chemistries of flavin ring systems of p hydroxybenzoate hydroxyla...
 
P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...
P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...
P68 RNA helicase unwinds the human let-7 microRNA precursor duplex and is req...
 
microbial community structure of polluted river sediments
microbial community structure of polluted river sedimentsmicrobial community structure of polluted river sediments
microbial community structure of polluted river sediments
 
Lipid peroxidation
Lipid peroxidationLipid peroxidation
Lipid peroxidation
 
Pol.j.environ.stud.vol.25.no.2.563 571
Pol.j.environ.stud.vol.25.no.2.563 571Pol.j.environ.stud.vol.25.no.2.563 571
Pol.j.environ.stud.vol.25.no.2.563 571
 
Reactive oxygen species
Reactive oxygen speciesReactive oxygen species
Reactive oxygen species
 
OBC- 2013
OBC- 2013OBC- 2013
OBC- 2013
 
Paquet JEM2011
Paquet JEM2011Paquet JEM2011
Paquet JEM2011
 
Role of cytochrome p450 in toxicity
Role of cytochrome p450 in toxicityRole of cytochrome p450 in toxicity
Role of cytochrome p450 in toxicity
 
10 nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...
10  nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...10  nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...
10 nazir ahmad malla and mudasir bashir 215 plant protein kinases in signal ...
 
2016 RBC RETREAT POSTER TEMPLATE_SMP
2016 RBC RETREAT POSTER TEMPLATE_SMP2016 RBC RETREAT POSTER TEMPLATE_SMP
2016 RBC RETREAT POSTER TEMPLATE_SMP
 
3aff_monooxygensae.pdf
3aff_monooxygensae.pdf3aff_monooxygensae.pdf
3aff_monooxygensae.pdf
 
Respiration
RespirationRespiration
Respiration
 
Quantitative phosphoproteomic analysis reveals common
Quantitative phosphoproteomic analysis reveals commonQuantitative phosphoproteomic analysis reveals common
Quantitative phosphoproteomic analysis reveals common
 

More from Jalal Hawari

MacPherson 1998 EST
MacPherson 1998 ESTMacPherson 1998 EST
MacPherson 1998 ESTJalal Hawari
 
Sheremata 1999 EST
Sheremata 1999 ESTSheremata 1999 EST
Sheremata 1999 ESTJalal Hawari
 
Chang Petroleum HCs Biodeg EST 2011
Chang Petroleum HCs Biodeg EST 2011Chang Petroleum HCs Biodeg EST 2011
Chang Petroleum HCs Biodeg EST 2011Jalal Hawari
 
Pelletier et al 2015
Pelletier et al 2015Pelletier et al 2015
Pelletier et al 2015Jalal Hawari
 
Archaelogical Sci Robins 1983
Archaelogical  Sci  Robins 1983Archaelogical  Sci  Robins 1983
Archaelogical Sci Robins 1983Jalal Hawari
 
Brindlay 1983 JOrganometChem
Brindlay 1983 JOrganometChemBrindlay 1983 JOrganometChem
Brindlay 1983 JOrganometChemJalal Hawari
 
Kanabus 1987 JAmChemSoc
Kanabus 1987 JAmChemSocKanabus 1987 JAmChemSoc
Kanabus 1987 JAmChemSocJalal Hawari
 
Marini Electro 2006
Marini Electro 2006Marini Electro 2006
Marini Electro 2006Jalal Hawari
 
Barker 1980 JChemSocPerkinTrans2
Barker 1980 JChemSocPerkinTrans2Barker 1980 JChemSocPerkinTrans2
Barker 1980 JChemSocPerkinTrans2Jalal Hawari
 
Al-Bashir 1994b WatRes
Al-Bashir 1994b WatResAl-Bashir 1994b WatRes
Al-Bashir 1994b WatResJalal Hawari
 
Al-Bashir 1994a WatRes
Al-Bashir 1994a WatResAl-Bashir 1994a WatRes
Al-Bashir 1994a WatResJalal Hawari
 
Naja 2008 ZVI RDX EST
Naja 2008 ZVI RDX ESTNaja 2008 ZVI RDX EST
Naja 2008 ZVI RDX ESTJalal Hawari
 
Betancourt et al 2007 SPME PHB JChrom
Betancourt et al  2007 SPME PHB JChromBetancourt et al  2007 SPME PHB JChrom
Betancourt et al 2007 SPME PHB JChromJalal Hawari
 
Sheremata 2000 CD TNT EST
Sheremata 2000 CD TNT ESTSheremata 2000 CD TNT EST
Sheremata 2000 CD TNT ESTJalal Hawari
 
Prasher Env Technol 2004
Prasher Env Technol 2004Prasher Env Technol 2004
Prasher Env Technol 2004Jalal Hawari
 
Monteil-Rivera 2012Biores Tech
Monteil-Rivera 2012Biores TechMonteil-Rivera 2012Biores Tech
Monteil-Rivera 2012Biores TechJalal Hawari
 
Hashaikeh 2007 Fuel
Hashaikeh 2007 FuelHashaikeh 2007 Fuel
Hashaikeh 2007 FuelJalal Hawari
 

More from Jalal Hawari (20)

MacPherson 1998 EST
MacPherson 1998 ESTMacPherson 1998 EST
MacPherson 1998 EST
 
Zhao 2010 Plosone
Zhao 2010 PlosoneZhao 2010 Plosone
Zhao 2010 Plosone
 
Sheremata 1999 EST
Sheremata 1999 ESTSheremata 1999 EST
Sheremata 1999 EST
 
Chang Petroleum HCs Biodeg EST 2011
Chang Petroleum HCs Biodeg EST 2011Chang Petroleum HCs Biodeg EST 2011
Chang Petroleum HCs Biodeg EST 2011
 
Pelletier et al 2015
Pelletier et al 2015Pelletier et al 2015
Pelletier et al 2015
 
Archaelogical Sci Robins 1983
Archaelogical  Sci  Robins 1983Archaelogical  Sci  Robins 1983
Archaelogical Sci Robins 1983
 
Brindlay 1983 JOrganometChem
Brindlay 1983 JOrganometChemBrindlay 1983 JOrganometChem
Brindlay 1983 JOrganometChem
 
Kanabus 1987 JAmChemSoc
Kanabus 1987 JAmChemSocKanabus 1987 JAmChemSoc
Kanabus 1987 JAmChemSoc
 
Marini Electro 2006
Marini Electro 2006Marini Electro 2006
Marini Electro 2006
 
Barker 1980 JChemSocPerkinTrans2
Barker 1980 JChemSocPerkinTrans2Barker 1980 JChemSocPerkinTrans2
Barker 1980 JChemSocPerkinTrans2
 
Whyte 1998 AEM
Whyte 1998 AEMWhyte 1998 AEM
Whyte 1998 AEM
 
Al-Bashir 1994b WatRes
Al-Bashir 1994b WatResAl-Bashir 1994b WatRes
Al-Bashir 1994b WatRes
 
Al-Bashir 1994a WatRes
Al-Bashir 1994a WatResAl-Bashir 1994a WatRes
Al-Bashir 1994a WatRes
 
Naja 2008 ZVI RDX EST
Naja 2008 ZVI RDX ESTNaja 2008 ZVI RDX EST
Naja 2008 ZVI RDX EST
 
Betancourt et al 2007 SPME PHB JChrom
Betancourt et al  2007 SPME PHB JChromBetancourt et al  2007 SPME PHB JChrom
Betancourt et al 2007 SPME PHB JChrom
 
Fournier 2002 AEM
Fournier 2002 AEMFournier 2002 AEM
Fournier 2002 AEM
 
Sheremata 2000 CD TNT EST
Sheremata 2000 CD TNT ESTSheremata 2000 CD TNT EST
Sheremata 2000 CD TNT EST
 
Prasher Env Technol 2004
Prasher Env Technol 2004Prasher Env Technol 2004
Prasher Env Technol 2004
 
Monteil-Rivera 2012Biores Tech
Monteil-Rivera 2012Biores TechMonteil-Rivera 2012Biores Tech
Monteil-Rivera 2012Biores Tech
 
Hashaikeh 2007 Fuel
Hashaikeh 2007 FuelHashaikeh 2007 Fuel
Hashaikeh 2007 Fuel
 

PNAS paper 2007

  • 1. Exploring the biochemical properties and remediation applications of the unusual explosive-degrading P450 system XplA/B Rosamond G. Jackson*, Elizabeth L. Rylott*, Diane Fournier† , Jalal Hawari† , and Neil C. Bruce*‡ *Center for Novel Agricultural Products, Department of Biology, University of York, P.O. Box 373, York YO10 5YW, United Kingdom; and †Biotechnology Research Institute, National Research Council of Canada, Montreal, QC, Canada H4P 2R2 Edited by May R. Berenbaum, University of Illinois at Urbana–Champaign, Urbana, IL, and approved September 10, 2007 (received for review May 31, 2007) Widespread contamination of land and groundwater has resulted from the use, manufacture, and storage of the military explosive hexa-hydro-1,3,5-trinitro-1,3,5-triazine (RDX). This contamination has led to a requirement for a sustainable, low-cost method to remediate this problem. Here, we present the characterization of an unusual microbial P450 system able to degrade RDX, consisting of flavodoxin reductase XplB and fused flavodoxin-cytochrome P450 XplA. The affinity of XplA for the xenobiotic compound RDX is high (Kd ‫؍‬ 58 ␮M) and comparable with the Km of other P450s toward their natural substrates (ranging from 1 to 500 ␮M). The maximum turnover (kcat) is 4.44 per s, only 10-fold less than the fastest self-sufficient P450 reported, BM3. Interestingly, the pres- ence of oxygen determines the final products of RDX degradation, demonstrating that the degradation chemistry is flexible, but both pathways result in ring cleavage and release of nitrite. Carbon monoxide inhibition is weak and yet the nitroaromatic explosive 2,4,6-trinitrotoluene (TNT) is a potent inhibitor. To test the efficacy of this system for the remediation of groundwater, transgenic Arabidopsis plants expressing both xplA and xplB were generated. They are able to remove saturating levels of RDX from liquid culture and soil leachate at rates significantly faster than those of untransformed plants and xplA-only transgenic lines, demonstrat- ing the applicability of this system for the phytoremediation of RDX-contaminated sites. cytochrome P450 ͉ phytoremediation ͉ hexa-hydro-1,3,5-trinitro- 1,3,5-triazine Cytochrome P450s catalyze a diverse range of chemical reac- tions including hydroxylation, epoxidation, demethylation, dehalogenation, desaturation, and isomerization (1). As a con- sequence, P450s are involved in a host of metabolic pathways. In eukaryotes and prokaryotes, they catalyze critical steps in the biosynthesis of key metabolites such as steroids and vitamins (2), fatty acids (3), and lignin (4). P450s have also been established as detoxification enzymes with activity toward a range of xeno- biotics. It was thus in keeping with this role that xplA, which was isolated from Rhodococcus rhodochrous 11Y by growth on the explosive hexa-hydro-1,3,5-trinitro-1,3,5-triazine (RDX) as sole nitrogen source, was found to encode a P450 (5). The synthetic N-NO2 bond of the RDX molecule, which is rare in nature, was accommodated by this enzyme. Further analysis revealed an unusual arrangement of subunits that contribute to the different steps in P450 catalysis. Cytochrome P450s, as heme-containing enzymes, require reduction of the heme to activate the catalytic center, a process involving supply of electrons from NAD(P)H to the P450 via partnering enzymes. The nature of the redox partners varies. Class one P450s include bacterial and mitochondrial enzymes that use an FAD-containing ferredoxin reductase-like protein and an iron-sulfur ferredoxin-like protein. Class two P450s are usually bound to the endoplasmic reticulum membrane along with the partnering enzyme, NADPH cytochrome P450 reduc- tase, which contains FAD and FMN domains. Although the P450s and redox partners predominantly exist as separate polypeptides, examples have come to light where the three catalytic domains required for activity are fused together, e.g., the Bacillal BM3 (6), Rhodococcal RhF (7, 8), and the fungal CYP505A1 (8). In all classes, despite the variety of forms, NADPH is usually the source of the electrons, and three electron transfer domains are involved. With XplA, a different arrange- ment of subunits is seen, with the second electron transfer step, a flavodoxin domain, fused to the P450 domain (9). The organization of the domains is also unusual, with the flavodoxin domain fused to the N terminus of the P450. The first electron transfer step has been postulated to be encoded by a reductase, xplB, adjacent to xplA in the R. rhodochrous genome. XplB has homology to adrenodoxin reductase (5), which transfers elec- trons from NADPH to adrenodoxin in a synthetically fused P450 (10), and also transfers electrons to flavodoxin (11). Interest in XplA and XplB has arisen after the contamination of land and groundwater with RDX as a result of the widespread manufacture, use, and disposal of munitions. This contamination is of concern as RDX is toxic to all classes of organisms tested, and the Environmental Protection Agency (EPA) classifies RDX as a priority pollutant. Contamination on military training ranges is of particular concern. For example, the use of RDX has been restricted by the EPA at the Massachusetts Military Reservation of Cape Cod where RDX contamination is threat- ening drinking water sources (12). Microorganisms present in soil heavily contaminated with ex- plosives have been found to degrade RDX, but do not possess sufficient biomass or metabolic activity to degrade this compound before it leaches through soils polluting groundwater. Interestingly, to date, xplA and xplB have been found only in Rhodococcus and related bacteria isolated from RDX-contaminated soil, suggesting that the RDX-degrading ability of XplA may have evolved under this selective pressure. XplA has been recombinantly expressed and shown to degrade RDX in vitro with a surrogate reductase (9). In this article, the activity of XplA with its native reductase XplB shows the ability of the proteins to work as efficient partners to degrade RDX. Further characterization is undertaken along with a detailed analysis of the RDX breakdown pathway under anaer- obic and aerobic conditions. We have previously demonstrated that expression of XplA in Arabidopsis confers both the ability to remove Author contributions: R.G.J., E.L.R., J.H., and N.C.B. designed research; R.G.J., E.L.R., and D.F. performed research; R.G.J., E.L.R., D.F., J.H., and N.C.B. analyzed data; and R.G.J., E.L.R., J.H., and N.C.B. wrote the paper. The authors declare no conflict of interest. This article is a PNAS Direct Submission. Abbreviations: RDX, hexa-hydro-1,3,5-trinitro-1,3,5-triazine; TNT, trinitrotoluene; NDAB, 4-nitro-2,4, diazabutanal; MEDINA, methylenedinitramine. ‡To whom correspondence should be addressed. E-mail: ncb5@york.ac.uk. This article contains supporting information online at www.pnas.org/cgi/content/full/ 0705110104/DC1. © 2007 by The National Academy of Sciences of the USA 16822–16827 ͉ PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 www.pnas.org͞cgi͞doi͞10.1073͞pnas.0705110104
  • 2. RDX from liquid culture and resistance to the phytotoxic effects of RDX; however, this activity relies on support from endogenous plant reductases (9). Here, the expression of both xplA and xplB in Arabidopsis enabled the rapid removal of RDX from liquid culture and soil leachate, a rate significantly faster than for plants express- ing xplA alone. These results demonstrate that this technology can be applied to remediate RDX from contaminated sites. Results and Discussion Optimizing Expression and Assay Conditions. The purification of XplA to homogeneity has been described (9). Soluble expression and purification of XplB was achieved by using a pGEX vector where GST is fused to the N terminus of XplB (Fig. 1A). Cleavage of the GST resulted in the loss of the 72-kDa XplB band from SDS/PAGE gels, the appearance of several smaller bands, and Ͻ5% protein recovery, suggesting insolubility of any cleaved protein (data not shown). The soluble, fused XplB was able to transfer electrons to XplA for the degradation of RDX, whereas GST alone had no such activity (Fig. 1B), therefore fused GST-XplB was used in subsequent studies. Fig. 1C shows that a 2-fold molar excess of XplB to XplA was the ratio at which XplA became limiting (as measured by flavin levels). Con- versely, a 10-fold molar excess of XplA to XplB was the ratio at which XplB became limiting (Fig. 1D). At lower ratios, the concentration of both enzymes influenced the rate of activity toward RDX, as expected for a second-order reaction, suggest- ing that collision of the two subunits is a major rate-limiting contribution. The optimal pH for RDX degradation was pH 6.5–7.0 irrespective of the buffer used, and potassium phosphate at pH 6.8 was used subsequently. Activity was not significantly affected by ionic strength between 0 and 100 mM NaCl, but above 100 mM, an inhibitory effect of sodium chloride was seen (data not shown), thus NaCl was omitted from further assays. Spectral Analysis of XplA and XplB. XplA had previously been shown to contain a classic P450 heme and to be able to bind carbon monoxide when reduced, producing a spectral shift to 450 nm, and a flavin binding domain (9), but the nature of the flavin was not determined. Purified XplB also possessed a classic flavin absorbance spectra, and release of the flavin from XplA and XplB by boiling and subsequent analysis by HPLC showed XplA to contain predominantly FMN and XplB FAD [supporting information (SI) Fig. 8]. Reduction of XplA by sodium dithionite causes a character- istic decrease in the heme 420-nm peak and a shift to a maximum of 389 nm. The dominating heme spectrum masked any flavin absorbance. Six nanomoles of XplA was fully reduced by be- tween 8 and 10 nmol of sodium dithionite, suggesting the majority of XplA has flavin bound (full reduction of heme and flavin would take 9 nmol) (Fig. 2 A and B). On reduction of XplB by NADPH, the flavin absorbance was completely bleached. No changes were observed between 550 and 600 nm, consistent with a lack of the stabilized semiquinone form, and implying a two-electron reduction (Fig. 2C). For XplB, the concentration of NADPH required for full reduction was approx- Fig. 1. Recombinant expression of XplB and assay of activity with XplA. (A) Protein purification on 10% SDS/PAGE. Lane 1, molecular weight markers; lane 2, solubilized recombinant protein; lane 3, affinity-purified XplB. (B) Aerobic activity of GST-XplB (0.26 ␮M; filled symbol) and GST (0.71 ␮M; open symbol) with XplA (0.27 ␮M). (C) Anaerobic activity of 60 nM XplA with various ratios of XplB. (D) Anaerobic activity of 60 nM XplB with various ratios of XplA. Values are the mean Ϯ SD of triplicates. Fig. 2. Spectral analysis of the reduction of XplA and XplB. (A) UV-visible spectra of 6 nmol of XplA (determined by protein concentration), titrated with the indicated amount of sodium dithionite (nmol) under anaerobic condi- tions. (B) Difference spectra of the sodium dithionite titration of XplA, gen- erated by subtraction of the original XplA spectrum from those with sodium dithionite. (C) Anaerobic titration of 37 nmol XplB (by protein) with the indicated amount of NADPH (nmol). Jackson et al. PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 ͉ 16823 APPLIEDBIOLOGICAL SCIENCES
  • 3. imately half the protein concentration, indicating that half of the XplB has flavin bound. Addition of flavin during purification and varying growth conditions did not improve this level. XplB was not readily reduced by NADH nor was NADH a successful electron donor for RDX degradation (data not shown). RDX Binding and Activity of XplA and XplB. The binding of RDX to XplA has been examined in two ways. Titration of XplA with RDX in an anaerobic environment revealed the low spin to high spin change in the heme often seen on substrate binding and a binding affinity (Kd) of 57.9 Ϯ 2.8 ␮M for RDX (Fig. 3 A and B). The Km was calculated to be 83.7 Ϯ 17.8 ␮M, and the maximum turnover (kcat) of the enzyme was 4.44 Ϯ 0.46 per s using a saturating ratio of XplB (2-fold molar excess) and anaerobic conditions (Fig. 3C). The Km and Kd values are similar to each other, and a turnover of 4.44 per s is comparable with the Pseudomonal P450cam toward its natural substrate camphor (27 per s) (13), perhaps surprising given the xenobiotic nature of RDX. The number of electron transfer steps involved in a P450 system often prevent the turnover from being substantially faster (1). RDX Breakdown Pathways. Initially, the breakdown products of RDX degradation were analyzed anaerobically to determine the amount of NADPH required without losses to uncoupled cleav- age of oxygen. Formaldehyde and nitrite were measured directly, whereas RDX and other products were assayed after freezing the samples. It was surprising to find ratios of nitrite to RDX, after 70 min, of 1.4:1.0 and formaldehyde to RDX of 1.96:1.00 as it was previously thought that the breakdown pathway would follow that proposed by Fournier et al. (14) with 2:1 nitrite and 1:1 formaldehyde and production of 4-nitro-2,4, diazabutanal (NDAB) (Fig. 4A). NDAB was not detected; however, the RDX breakdown product methylenedinitramine (MEDINA) was, reaching a ratio of 0.68:1 after 70 min and then decreasing, possibly because of instability in water (15). The ratio of NADPH to RDX degraded was 1.26:1.00 at 70 min, suggesting these compounds are tightly coupled. When the breakdown pathway under aerobic conditions was examined, a different picture arose with an increase in the ratio of nitrite to RDX to 2.49:1.00 after 130 min and a decrease in the formaldehyde to RDX ratio to 1.4:1.0 (Fig. 4B). NDAB was detected, and levels continued to increase throughout the time course; MEDINA was not detected. The mechanism for denitration of RDX by XplA is not yet fully known; however, once denitration and hydration have occurred, whether under aerobic or anaerobic conditions, the resulting imine intermediate would be highly unstable in water and spontaneously decompose (16). Under anaerobic condi- tions, mono-denitration and mono-hydration of RDX followed by ring cleavage would produce MEDINA (Fig. 5, route A). Under aerobic conditions, it is proposed that RDX is subjected to di-denitration–di-hydration before ring cleavage. This mech- anism, leading to the formation of NDAB (see Fig. 5, route B), has been described (14). Degradation of RDX with radiolabeled Fig. 4. Mass balance of RDX breakdown. (A) Anaerobic degradation of RDX (60 nM XplA and XplB) and analyses were carried out as in Materials and Methods. Controls with boiled XplA and XplB (not shown) contained the following levels of analytes (nmol/ml) over the time course: RDX, 100 Ϯ 4.3; nitrite, 0 Ϯ 3.9; formaldehyde, 0 Ϯ 6.3; and MEDINA, 0–6.3. (B) Aerobic degradation of RDX. Reactions contained 90 nM XplA and XplB. Controls with boiled XplA and XplB (not shown) contained RDX (100 Ϯ 2.5), nitrite (0 Ϯ 2), formaldehyde (0 Ϯ 16), and NDAB (0–2.6) over the time course. Values are the mean Ϯ SD of triplicates. Fig. 3. XplA and RDX binding. All analyses were carried out anaerobically. (A) UV-visible spectral changes in 4.5 nmol XplA on the addition of 0–150 nmol of RDX (in DMSO). (Inset) Difference spectra generated by subtraction of nonbound XplA spectrum from the spectra of XplA with RDX. (B) Plot of A391 Ϫ A425 generated from the difference spectra against RDX concentration. (C) Initial rates of substrate use at a range of RDX concentrations plotted against RDX concentration with 60 nM XplA and 120 nM XplB. Values are the mean Ϯ SD of triplicates. 16824 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0705110104 Jackson et al.
  • 4. oxygen and another P450 system suggests that direct hydroxy- lation of the RDX molecule by XplA is unlikely (17). The role oxygen plays in the mechanism of RDX degradation by XplA, either within the enzyme or within the reaction environment, is therefore currently unclear. The pathway presented in Fig. 5 is based on the analysis of the final degradation products and the proposed intermediates by analogy from previously published work (14, 17). Additional work involving labeled RDX and deuterated solvent may allow precise details of the mechanism to be determined. Other examples of P450s catalyzing different reactions depend- ing on the presence of oxygen are known. For example, under anaerobic conditions human CYP2A6 and CYP101 both catalyze reductive reactions toward halogenated substrates, whereas under aerobic conditions, CYP2A6 catalyzes dehalogenation, (18) and CYP101 catalyzes a hydroxylation reaction (19). XplA Activity with a Wider Range of Substrates and Inhibitors. RDX as a synthetic compound may not be the native substrate for XplA, so the activity toward a range of other substrates was tested. The related nitramine explosive, octahydro-1,3,5,7- tetranitro-1,3,5,7-tetrazocine (HMX), was not transformed at the aqueous solubility limit for HMX of 15 ␮M, even after several hours of incubation. Trace levels of XplA activity toward the nitroaromatic explosive trinitrotoluene (TNT) were found (turnover was Ͻ1% per h at 100 ␮M), but transformation products could not be identified. The heme spectrum of XplA was not altered by the presence of TNT. To test the ability of XplA to perform classic P450 hydroxylation and demethylation reactions, established P450 substrates were tested. No hydroxy- lating activity was detected toward testosterone or paclitaxel, nor demethylating activity toward 7-ethoxycoumarin or ethoxyresorufin. However, oxidizing activity was detected to- ward both methyl tolyl and methyl phenyl sulfides generating sulfoxide products (data not shown). Methyl tolyl sulfide was also shown to inhibit RDX catabolism, as was the P450-specific inhibitor metyrapone. The strongest inhibition was seen by TNT (Fig. 6), whereas the rate of carbon monoxide inhibition was less than expected, given the usual high affinity of P450s for carbon monoxide. This inhibition was not increased by a higher concentration of carbon monoxide. XplA may have a lower affinity for CO in the presence of RDX, as seen with other substrates of P450s (20). Application of the Enzymes for Phytodegradation of RDX. It has previously been shown that transgenic plants expressing xplA can degrade RDX (9), but this activity relies on the availability of endogenous plant reductases, which may be limiting. Thus, xplB was transformed into Arabidopsis, and transgenic plant lines expressing both xplA and xplB were generated. A previously characterized plant line expressing xplA (XplA-10) (9) was used to produce five independently transformed lines expressing xplB. In addition, five independently transformed lines expressing only xplB were characterized. The results presented here are from plants homozygous for these transgenes. Quantitative analysis by real-time PCR showed that xplA was expressed in all five XplAB lines (Fig. 7A), although expression levels varied from that of the original parental line, XplA-10. A range in the level of xplB transcript was seen; with line XplAB-27 exhibiting the highest levels of both xplA and xplB transcripts (Fig. 7A). These lines were grown in axenic liquid culture to determine rates of RDX uptake. As reported (9), line XplA-10 removed all 180 ␮M RDX from the medium within 5 days; however, the XplAB lines removed the RDX significantly faster, with lines XplAB-2 and XplAB-27 removing Ͼ50% of the RDX within 4 h, 30 times faster than the XplA-10 line (Fig. 7B). The xplB-only lines had uptake rates similar to those of untransformed, wild-type plants (Fig. 7C). NDAB and MEDINA levels were not measured in the liquid culture or soil-grown plants. Liquid culture-grown plants and water-saturated soil-grown roots are likely to be hypoxic. It is possible from our characterization that, depending on oxygen availability, either NDAB or MEDINA is produced by xplA– expressing plants. To investigate the ability of the XplAB lines to reduce levels of RDX in contaminated ground water, 8-week-old plants were watered with 180 ␮M RDX. After 1 week, the soil was flushed with water and the level of RDX in the soil leachate was measured. After this time, the level of RDX in the leachate from untransformed, wild-type plants was unaltered, whereas leachate from the XplA-10 line had decreased by 25%. The RDX in the leachate from lines XplAB-2 and XplAB27 had decreased by 90–97% (Fig. 7D). Conclusions XplA and XplB constitute a novel P450 redox system and together efficiently degrade the xenobiotic RDX. Degradation follows two different routes dependent on the presence of Fig. 5. Proposed degradation pathway of RDX under aerobic and anaerobic conditions. Ring cleavage occurs at ab under anaerobic conditions (route A) and at cb under aerobic conditions (route B). Compounds in brackets are hypothetical, and the mechanisms are based on detection of nitrite and formaldehyde, the final products, and analogy with previous work (14, 17). Fig. 6. Activity of XplA and XplB with inhibitors. Standard anaerobic activity assays using 100 ␮M RDX as substrate were carried out in the presence of 100 ␮M metyrapone or 100 ␮M TNT, and activity was compared with that with only RDX (100%). Carbon monoxide (CO) and methyl tolyl sulfide (MTS) inhibition was tested in standard aerobic conditions using 100 ␮M RDX as substrate. Values are the mean Ϯ SD of triplicates. Jackson et al. PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 ͉ 16825 APPLIEDBIOLOGICAL SCIENCES
  • 5. oxygen. One mole of MEDINA and nitrite are the dominant products anaerobically, whileas 1 mol of NDAB and 2 mol of nitrite are produced in aerobic conditions. With both pathways resulting in ring cleavage and nitrite release, applications of these enzymes for bioremediation look encouraging. One of the biggest concerns of RDX as a pollutant is that it migrates readily through soil into the groundwater and subsequently contami- nates drinking water supplies. Here, we show that Arabidopsis plants expressing xplA and xplB have the ability to effectively remove RDX from the soil leachate. The studies here illustrate that these genes, or possibly an xplA–xplB gene fusion, could be engineered into plant species suited to growth on military training ranges and used to remediate RDX. Materials and Methods RDX was supplied by the Defense Science and Technology Laboratory at the U.K. Ministry of Defense (Fort Halstead, Kent, U.K.). MEDINA and NDAB were provided by Ron Spanggord (SRI International, Menlo Park, CA). XplA and XplB Expression and Purification. The xplA gene was cloned and expressed as described (9). The xplB gene was amplified from pHSX1 by PCR using primers containing overhanging BamH1 sites, 5Ј-GGATCCGACATCATGAGTGAAGTGGAC and 3Ј- GGATCCGCAGACCGATTCGGCCGG and ligated into pGEX2T (Merck Chemicals, Nottingham, U.K.), which engineers an N-terminal GST domain. The construct was sequenced, trans- formed into Escherichia coli BL21 (DE3) and grown at 20°C in Luria broth containing 100 ␮g/ml carbenicillin to OD600 (1.0), 1 mM isopropyl ␤-d-thiogalactopyranoside was added, and the cul- ture was grown for an additional 24 h. The cell pellet was resus- pended in 10 ml of PBS (140 mM NaCl/15 mM KH2PO4/80 mM Na2HPO4) containing 0.2 mM PMSF and lysed at 1,500 psi (1 psi ϭ 6.89 kPa) in a French Press (Thermo IEC). The lysate was centrifuged (10,000 ϫ g for 15 min), and the soluble protein was purified by using 200 ␮l of 50% glutathione-coupled Sepharose gel (Amersham, Piscataway, NJ) according to the manufacturer’s instructions and recovered in glutathione elution buffer (20 mM reduced glutathione/100 mM Tris⅐HCl, pH 8.5/120 mM NaCl). Protein assays were carried out with Coomassie Protein Assay Reagent (Pierce, Rockford, IL) using BSA as reference. Proteins were analyzed as described (21). Flavin Content Determination. The flavin content of XplA and XplB was determined after boiling for 20 min then centrifugation of the precipitated protein and trichloroacetic acid precipitation (10% of 240 mg/ml), followed by centrifugation and sodium bicarbonate neutralization. Both methods gave similar results for each protein. The concentration of released flavin was used to measure XplA and XplB concentration for activity assays. The released flavins were identified by HPLC according to ref. 22 and compared with commercially available standards for identification. Spectral Analyses. Spectral analyses were performed on a spec- trophotometer (v560; Jasco, Easton, MD) scanning between 800 and 300 nm at 400 nm/min. Anaerobic analyses were carried out in an anaerobic chamber (Coy Laboratory Products, Ann Arbor, MI). For all spectral analyses, enzyme concentrations quoted are for sample protein. The XplA titration with sodium dithionite in the anaerobic chamber followed desalting of the protein on a PD-10 column (GE Healthcare) into 50 mM potassium phos- phate, pH 6.8 (bubbled with nitrogen before equilibrating in the chamber for 2 days). For RDX titration, RDX was dissolved in DMSO and no more than 3 ␮l was added per ml. Fig. 7. Characterization of XplAB and XplB transgenic Arabidopsis lines. (A) Expression of xplA and xplB in rosette leaves. Quantitative analysis was done by real-time PCR of xplA and xplB transcript abundance in rosette leaves of the transgenic lines relative to line XplAB-2. ACTIN2 mRNA was used as an internal reference. Results represent the mean of three independent RNA isolations measured in duplicate from the pooled rosette leaves of 10 plants Ϯ SE. (B and C) Uptake of RDX from media by Arabidopsis seedlings. Results are the mean Ϯ SE of five replicate flasks, each containing 200 10-day-old seedlings. (D) Levels of RDX in soil leachate from Arabidopsis plants watered with 180 ␮M RDX. Results are the mean Ϯ SE of five replicate pots. NPC, no plant control. 16826 ͉ www.pnas.org͞cgi͞doi͞10.1073͞pnas.0705110104 Jackson et al.
  • 6. XplA and XplB Activity Assays Toward RDX. Anaerobic conditions. XplA and XplB were placed on ice in an anaerobic chamber in open vials for 3 h for the palladium catalyst to remove oxygen. Buffer was bubbled with nitrogen before equilibration in the chamber. The reaction mixture contained 60 nM XplA and XplB in 50 mM potassium phosphate (pH 6.8), 300 ␮M NADPH, and 100 ␮M RDX in a total volume of 1 ml. Other reagents and solutions were placed overnight in the anaerobic chamber with a nitrogen atmosphere, before use. Oxygen levels were monitored with a model 10 gas analyzer (Coy Laboratory Products) and main- tained Ͻ1 ppm by using a 5% (vol/vol) hydrogen mix in the presence of a palladium catalyst. Aerobic conditions. The reaction mixture contained 175 nM XplA and 150 nM XplB, 50 mM potassium phosphate (pH 6.8), 300 ␮M NADPH, 100 ␮M RDX, 0.72 units Thermoanaerobium brockii alcohol dehydrogenase (Sigma, St. Louis, MO), and 30 ␮l isopropanol in a total volume of 1 ml. All reactions were performed at room temperature (20°C), and assays were stopped by the addition of 10% trichloroacetic acid (240 mg/ml) or, for the mass balance experiments, 30-kDa cut-off spin columns (Microcon YM-30; Amicon). The reaction was initiated by the addition of RDX. Analysis of Products. RDX removal was measured by using RP- HPLC with a HPLC system (2695 Separations Module and 2996 Photodiode Array Detector; Waters) using a Techsphere C18 column (250 ϫ 4.6 mm) under isocratic conditions of 60% water and 40% acetonitrile at a flow rate of 1 ml/min over 10 min. RDX elution was monitored at 205 nm, and intergrations were per- formed with Empower software. Formaldehyde analysis was carried out according to Nash (23). Nitrite analysis followed the method of Scheideler and Ninnemann (24) terminated by removal of proteins with 30-kDa cut-off spin columns. The concentration of NDAB and MEDINA was deter- mined by using an HPLC system as reported (25). Activity Assays Toward Other Substrates and with Inhibitors. Analysis of octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine was as de- scribed for RDX. Loss of TNT was analyzed by using a water mobile phase of 50% MeCN, 50% water over 12 min with monitoring at 230 nm. Paclitaxel was analyzed by using a methanol and water gradient from 60–70% MeOH over 20 min with monitoring at 230 nm. Testosterone was analyzed by HPLC using a 58–62% MeOH gradient over 8 min. Activity toward 7-ethoxycoumarin was tested by following fluorescence of 7-hy- droxy coumarin, with excitation at 350 nm and emission at 450 nm adapted from ref. 26 and TLC. Activity toward ethoxyresoru- fin was determined fluorometrically with excitation at 530 nm and emission at 590 nm, adapted from ref. 26. Carbon monoxide inhibition was carried out aerobically by adding 100 or 500 ␮l per ml of carbon monoxide-saturated buffer. Plant Transformation Methods. The xplB gene was cloned into the binary vector pART27 (27) under the control of the CaMV35S promoter and ocs terminator, and transformed by Agrobacte- rium-mediated floral dipping into wild-type and xplA-expressing Arabidopsis thaliana, ecotype Columbia-0 as in ref. 9. Liquid Culture and Soil Leachate Experiments. Liquid culture exper- iments were performed as described (9). Soil leachate studies were carried out on 6-week-old Arabidopsis plants grown under 180 ␮m⅐m2 ⅐s Ϫ1 light in a 12-h photoperiod. Plants were grown in pots containing 30 g of uncontaminated soil (Levingtons F2 compost), five plants per pot. Each pot was flooded with 50 ml of 180 ␮M RDX, then 7 days later, flushed through with 50 ml of water. The collected soil leachates were analyzed for RDX content by using HPLC as described above. This work was funded by the Strategic Environmental Research and Development Program of the U.S. Department of Defense, the Bio- technology and Biological Sciences Research Council, and the U.K. Ministry of Defense. 1. Guengerich FP (2001) Chem Res Toxicol 14:611–650. 2. Prosser DE, Jones G (2004) Trends Biochem Sci 29:664–673. 3. Green AJ, Rivers SL, Cheeseman M, Reid GA, Quaroni LG, Macdonald ID, Chapman SK, Munro AW (2001) J Biol Inorg Chem 6:523–533. 4. Schoch G, Goepfert S, Morant M, Hehn A, Meyer D, Ullmann P, Werck- Reichhart D (2001) J Biol Chem 276:36566–36574. 5. Seth-Smith HMB, Rosser SJ, Basran A, Travis ER, Dabbs ER, Nicklin S, Bruce NC (2002) Appl Environ Microbiol 68:4764–4771. 6. Munro AW, Lindsay JG, Coggins JR, Kelly SM, Price NC (1996) Biochim Biophys Acta 1296:127–137. 7. Hunter DJ, Roberts GA, Ost TW, White JH, Muller S, Turner NJ, Flitsch SL, Chapman SK (2005) FEBS Lett 579:2215–2220. 8. Nakayama N, Takemae A, Shoun H (1996) J Biochem (Tokyo) 119:435–440. 9. Rylott EL, Jackson RG, Edwards J, Womack GL, Seth-Smith HM, Rathbone DA, Strand SE, Bruce NC (2006) Nat Biotechnol 24:216–219. 10. Cao PR, Bulow H, Dumas B, Bernhardt R (2000) Biochim Biophys Acta 1476:253–264. 11. Zollner A, Nogues I, Heinz A, Medina M, Gomez-Moreno C, Bernhardt R (2004) Bioelectrochemistry 63:61–65. 12. Clausen J, Robb J, Curry D, Korte N (2004) Environ Pollut 129:13–21. 13. Peterson JA, Mock DM (1975) Adv Exp Med Biol 58:311–324. 14. Fournier D, Halasz A, Spain J, Fiurasek P, Hawari J (2002) Appl Environ Microbiol 68:166–172. 15. Halasz A, Spain J, Paquet L, Beaulieu C, Hawari J (2002) Environ Sci Technol 36:633–638. 16. Melius CF (1990) in Chemistry and Physics of Energetic Materials: Thermochemical Modeling, ed Bulusu SN (Kluwer, Dordrecht, The Netherlands), pp 21–49. 17. Bhushan B, Trott S, Spain JC, Halasz A, Paquet L, Hawari J (2003) Appl Environ Microbiol 69:1347–1351. 18. Spracklin DK, Kharasch ED (1998) Drug Metab Dispos 26:605–607. 19. Koe GS, Vilker VL (2005) Biotechnol Prog 21:1119–1127. 20. Sevrioukova IF, Peterson JA (1995) Arch Biochem Biophys 317:397–404. 21. Sambrook J, Russell DW (2001) Molecular Cloning: A Laboratory Manual (Cold Spring Harbor Lab Press, Cold Spring, NY). 22. Aliverti A, Curti B, Vanoni M (1999) in Flavoprotein Protocol, eds Chapman SK, Reid GA (Humana, Totowa, NJ), pp 9–23. 23. Nash T (1953) Biochem J 55:416–421. 24. Scheideler L, Ninnemann H (1986) Anal Biochem 154:29–33. 25. Fournier D, Halasz A, Spain J, Spanggord RJ, Bottaro JC, Hawari J (2004) Appl Environ Microbiol 70:1123–1128. 26. Werck-Reichhart D, Gabriac B, Teutsch H, Durst F (1990) Biochem J 270:729–735. 27. Gleave AP (1992) Plant Mol Biol 20:1203–1207. Jackson et al. PNAS ͉ October 23, 2007 ͉ vol. 104 ͉ no. 43 ͉ 16827 APPLIEDBIOLOGICAL SCIENCES