ArticlePDF Available

Glutathione transferases in bacteria

Authors:
  • University “G. d' Annunzio” of Chieti - Pescara
REVIEW ARTICLE
Glutathione transferases in bacteria
Nerino Allocati
1
, Luca Federici
1,2
, Michele Masulli
1
and Carmine Di Ilio
1,2
1 Dipartimento di Scienze Biomediche, Universita
`‘G. d’Annunzio’, Chieti, Italy
2 Ce.S.I. Centro Studi per l’Invecchiamento, Universita
`‘G. d’Annunzio’, Chieti, Italy
Introduction
Glutathione transferases (GSTs; EC 2.5.1.18) consti-
tute a protein superfamily that is involved in cellular
detoxification against harmful xenobiotics and endo-
biotics [1–3]. One common feature of all GSTs is their
ability to catalyse nucleophilic attack by the tripeptide
glutathione (GSH) on the electrophilic groups of a
wide range of hydrophobic toxic compounds, thus pro-
moting their excretion from the cell [1]. GSTs are also
involved in several other cell functions, and are capa-
ble of binding a large number of endogenous and
exogenous compounds non-catalytically [1]. GSTs are
widely distributed in nature and are found in both
eukaryotes and prokaryotes. GSTs are divided into
at least four major families of proteins, namely cyto-
solic GSTs, mitochondrial GSTs, microsomal GSTs
and bacterial fosfomycin-resistance proteins [1,4]. The
cytosolic GSTs (cGSTs) have been subgrouped into
numerous divergent classes on the basis of their chemi-
cal, physical and structural properties [1,2]. The mito-
chondrial GSTs, also known as kappa class GSTs, are
soluble enzymes that have been characterized in
eukaryotes [5]. The third GST family comprises mem-
brane-bound transferases called membrane-associated
proteins involved in ecosanoid and glutathione metab-
olism (MAPEG), but these bear no similarity to solu-
ble GSTs [6]. Representatives of all three families are
also present in prokaryotes. The fourth family is found
exclusively in bacteria.
Keywords
bacterial glutathione transferase; chemical
stress; DCM dehalogenase; detoxification;
drug resistance; fosfomycin resistance
proteins; glutathione; HCCA isomerase;
oxidative stress; TCHQ dehalogenase
Correspondence
N. Allocati, Dipartimento di Scienze
Biomediche, Universita’ ‘G. d’Annunzio’, Via
dei Vestini 31, I-66013 Chieti, Italy
Fax: +39 0871 355 5282
Tel: +39 0871 355 5281
E-mail: allocati@unich.it
(Received 25 July 2008, revised 8 October
2008, accepted 14 October 2008)
doi:10.1111/j.1742-4658.2008.06743.x
Bacterial glutathione transferases (GSTs) are part of a superfamily of
enzymes that play a key role in cellular detoxification. GSTs are widely dis-
tributed in prokaryotes and are grouped into several classes. Bacterial
GSTs are implicated in a variety of distinct processes such as the biodegra-
dation of xenobiotics, protection against chemical and oxidative stresses
and antimicrobial drug resistance. In addition to their role in detoxifica-
tion, bacterial GSTs are also involved in a variety of distinct metabolic
processes such as the biotransformation of dichloromethane, the degrada-
tion of lignin and atrazine, and the reductive dechlorination of pentachloro-
phenol. This review article summarizes the current status of knowledge
regarding the functional and structural properties of bacterial GSTs.
Abbreviations
BxGST, Burkholderia xenovorans GST; CDNB, 1-chloro-2,4-dinitrobenzene; cGST, cytosolic GST; DCM, dichloromethane; EcGST,
Escherichia coli GST; GSH, glutathione; GST, glutathione transferase; HCCA, 2-hydroxychromene-2-carboxylic acid; MAPEG, membrane-
associated proteins involved in ecosanoid and glutathione metabolism; OaGST, Ochrobactrum anthropi GST; PmGST, Proteus mirabilis GST;
TCHQ, tetrachlorohydroquinone.
58 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
The first evidence for the presence of GSTs in bacteria
was reported more 25 years ago by Takashi Shishido
who showed the presence of GST activity in a strain of
Escherichia coli [7]. Since then, GSTs have been found
to be broadly distributed in aerobic prokaryotes, includ-
ing human and plant pathogens and soil bacteria. Con-
versely, GSTs have not been identified to date in
anaerobic bacteria [8] or in Archaea, the other domain
of prokaryotes. The absence of the enzyme in these
microorganisms is consistent with the lack of GSH [9].
This review summarizes current knowledge on the
functional and structural properties of bacterial GSTs
and the potential biotechnological applications of these
enzymes.
Classification and phylogenetic
relationships
cGSTs, also termed canonical GSTs, are homo- or hete-
rodimeric enzymes found in aerobic forms of life. cGSTs
metabolize a broad range of electrophilic substrates via
GSH conjugation. They are involved in protecting cells
against oxidative stresses, have peroxidase and isomer-
ase activities and are implicated in the development of
drug resistance [1,2]. In mammalian species they are well
characterized and have been grouped into seven classes:
alpha, mu, pi, sigma, theta, omega and zeta [1]. Several
other classes are restricted to non-mammalian species:
beta, chi, delta, epsilon, lambda, phi and tau [1,2,10]. To
assign a cGST to a class, it is generally accepted that
proteins with > 40% sequence identity belong to the
same class, whereas GSTs of different classes share
< 25% sequence identity. The identity increases if the
N-terminal region only is considered. This region com-
prises part of the active site, with residues that interact
with GSH, and it is evolutionarily conserved [1,2,11].
Besides amino acid sequence identity, immunological
properties, kinetic features as well as similarity of the
crystal structures provide additional supporting data
[1,2].
In bacteria, four different classes of cGSTs have
been identified: beta, chi, theta and zeta [2,10,12,13].
Beta class cGSTs have been purified and characterized
from several bacteria [14–25]. They are able to conju-
gate the model substrate 1-chloro-2,4-dinitrobenzene
(CDNB) and bind to the GSH–affinity matrix. All beta
class enzymes are characterized by the presence of a
cysteine residue at the GSH site [12]. The first cGST of
this class was characterized from a Proteus mirabilis
strain (PmGST). PmGST displayed biochemical and
structural properties that distinguish it from the GSTs
of other families and it was identified as the prototype
of the beta class [14,26–36].
Three other orthologues of the beta class were also
functionally and structurally characterized, from
E. coli (EcGST) and from two soil bacteria, Ochrobac-
trum anthropi (OaGST) and Burkholderia xenovorans
(BxGST, also known as BphK), respectively
[20,21,24,37–47].
Theta class enzymes in bacteria are represented by
two dichloromethane (DCM) dehalogenases produced
by facultative methylotrophic bacteria [48–52]. They
exhibit high amino acid sequence similarity to eukar-
yotic theta class GSTs and also share some properties
of these enzymes such as their reactivity with DCM,
their lack of activity with CDNB and their inability to
bind to GSH affinity matrices [50].
Tetrachlorohydroquinone (TCHQ) dehalogenase
was reported by Anandarajah et al. [53] to belong to
the zeta class on the basis of multiple sequence align-
ments. In particular, this enzyme is characterized, in
the GSH site, by the distinctive motif of zeta class
enzymes including two serine and a cysteine residues.
TCHQ dehalogenase is involved in two steps of the
biodegradation of pentachlorophenol and it also has
isomerase activity [53–56].
Recently a novel class of cGSTs, called chi class,
was proposed [10]. Two cyanobacterial cGSTs have
been purified and characterized from Thermosynechoc-
cus elongatus BP-1 (TeGST) and Synechoccus elongatus
PCC 6301 (SeGST). Although TeGST and SeGST
showed typical structural features of cGSTs, the results
presented here argued against their incorporation into
the beta class. In particular, these enzymes lack cyste-
ines completely indicating a possible different evolu-
tionary pathway for the cyanobacterial GSTs from the
beta class GSTs.
Like eukaryotic organisms, bacteria are character-
ized by multiple GST genes of widely divergent
sequences and unknown function [57]. In the E. coli
genome, in addition to the beta class GST [37] and to
Stringent starvation protein A, a RNA polymerase
with fold similarity to cGSTs [3], six GST homologues
have been identified [58]. The products of two of these
genes, YfcF and YfcG, exhibited GST- and GSH-
dependent peroxidase activities and were involved in
the defence against oxidative stress [59]. Pseudomonads
possess more than 10 GST genes [57]. In P. mirabilis
as well as in Proteus vulgaris three and four different
GSTs were identified, respectively [14,60]. Recently,
genomic analysis of the Gram-negative Shewanella
oneidensis revealed the presence of six GST-like genes
[61]. Two of these GST products showed high
sequence similarities to DCM dehalogenases.
Bacterial 2-hydroxychromene-2-carboxylic acid (HCCA)
isomerase is a GSH-dependent enzyme that is impli-
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 59
cated in the naphtahalene degradation pathway
[5,62,63]. HCCA isomerase shares strong similarity
with kappa class enzymes, has a conserved serine resi-
due in the GSH site and displays catalytic activity
against CDNB [63]. Kappa class GSTs are soluble
dimeric proteins isolated and characterized from mam-
malian mitochondria [1,5]. In humans, hGSTK1 was
also located in the peroxisomal fraction suggesting a
new role for this class [1]. They are structurally distinct
from cGSTs, exhibit GSH activity towards haloarenes
and reduce cumene hydroperoxide and tert-butyl
hydroperoxide [1,5].
The MAPEG family are a wide and ubiquitous
group with diverse biological functions [6,64]. The
eukaryotic MAPEG family comprises at least six mem-
bers: 5-lipoxigenase activating protein (FLAP) and leu-
kotriene C
4
synthase (LTC
4
) involved in leukotriene
biosynthesis, MGST1, MGST2 and MGST3 with GST
and peroxidase activity and prostaglandin E synthase
(PGES) which catalyses the formation of prostaglan-
din E
2
from prostaglandin H
2
[6,64,65]. Recently,
crystallographic studies on different MAPEG members
clearly demonstrated that they are arranged into
trimers [66].
MAPEG members were also identified in several
bacteria such as E. coli,Vibrio cholerae and Synecho-
cystis sp. [6]. Some of them contained multiple MA-
PEG paralogues. They showed distant evolutionarily
relationships compared with eukaryotic mammalian
and non-mammalian forms. Bacterial MAPEG pro-
teins have been grouped in two distinct subfamilies,
one of which comprises E. coli and V. cholerae pro-
teins and the other the Synechocystis sp. protein which
more closely resembles enzymes belonging to the
MGST2/FLAP/LTC
4
subgroup [65]. Genes encoding
MAPEG proteins from E. coli and Synechocystis sp.
have been cloned and overexpressed. Membrane frac-
tions from cells overproducing E. coli MGST as well
as SynMGST showed GST activity with CDNB as sec-
ond substrate [65]. To date, no information about their
physiological role in bacteria is available.
A phylogenetic analysis using representative
prokaryotic and eukaryotic members of the GST
superfamily is shown in Fig. 1. The phylogenetic rela-
tionships among bacterial GSTs are consistent with
their known functional and structural features, show-
ing that they are distributed in different classes. The
beta class comprises the majority of bacterial GSTs
and enzymes belonging to the chi class are closely
related to this class. The two DCM dehalogenases are
clustered together with mammalian theta class
enzymes. TCHQ dehalogenases are closely related to
the zeta class. Agrobacterium tumefaciens GST shows
closer evolutionary relationship with mammalian alpha
class than with bacterial GSTs as reported previously
[67]. Sphingomonas paucimobilis LigF and LigG group
together and are distant from the Sph. paucimobilis
LigE enzyme which is more closely related to the zeta
class enzymes. Finally, Rhodococcus AD45 form a new
branch and it is not grouped with any other class.
Fig. 1. Evolutionary relationship between representative bacterial and eukaryotic GSTs. Multiple sequence alignment was produced by using
CLUSTAL W2 [143] and improved manually. The unrooted phylogenetic trees were constructed by the neighbour-joining method, based on the
distances derived from the Dayhoff matrix, with MEGA 4.0 software [144]. The robustness of the branches was assessed by the bootstrap
method with 1000 replications. Only nodes with a bootstrap value > 25% are reported. The scale bar represents a distance of 0.5 substitu-
tions per site. The sequences have the following accession numbers: epsilon class: Anopheles gambiae (XP_319972), Drosophila melanogas-
ter (CG5164); delta class: An. gambiae (Q93113), Anopheles dirus (AF273039); theta class: D. melanogaster (Q9VG96), An. gambiae
(Q94999), Bos taurus (Q2NL00), Homo sapiens (P30712); Methylobacterium sp. DM4 (P21161); Methylophilus sp. DM11 (P43387); phi class:
Arabidopsis thaliana (P42769), Zea mays GSTF3 (Q9ZP62), Z. mays GSTF1 (P12653), Triticum aestivum (P30111); Sph. paucimobilis LigF
(P30347); Sph. paucimobilis LigG (BAA77216); lambda class: A. thaliana (Q9LZ07), Z. mays (P49248), A. thaliana GSTL2 (Q9M2W2); omega
class: H. sapiens (P78417), Rattus norvegicus (Q9Z339), Caenorhabditis elegans (AAA27959); tau class: Aegilops tauschii (O04941), Z. mays
GSTU1 (Y12862), Z. mays GSTU2 (AJ010439); Agrobacterium tumefaciens (Q8UJG9); alpha class: H. sapiens GSTA1 (P08263), Mus muscu-
lus, GSTA2 (P10648); R. norvegicus (P04904), Gallus gallus GSTA1 (Q08392), G. gallus GSTA2 (Q08393), Bos taurus (Q5E9G0); sigma class:
H. sapiens (O60760), R. norvegicus (O35543), C. elegans (O16116), Manduca sexta (P46429), Ommastrephes sloanei (P46088); mu class:
H. sapiens (P09488), M. musculus (P10649), G. gallus (P20136), Dermatophagoides pteronyssinus (P46419); pi class: Onchocerca volvulus
(P46427), Bufo bufo (P81942), H. sapiens (P09211), R. norvegicus (P04906); Rhodococcus AD45 (AJ249207); chi class: T. elongatus
(NP_680998), S. elongatus (YP_171005), beta class: Haemophilus influenzae (P44521), Xylella fastidiosa (Q9PE18), Xanthomonas campestris
(P45875), O. anthropi (P81065), Magnetospirillum magnetotacticum (ZP_00054555), B. xenovorans LB400 (Q9RAFO), P. mirabilis (P15214),
E. coli (P39100); Sph. paucimobilis LigE (BAA02032); zeta class: H. sapiens (O43708), M. musculus (Q9WVL0), T. aestivum (O04437), A. tha-
liana (Q9ZVQ3); Sphingobium chlorophenolicum (Q03520); Sphingomonas spUG30 (AY057901); rho class: Pleuronectes platessa (X63761),
Pagrus major (AB158412), Micropterus salmoides (AY335905), Branchiostoma belcheri tsingtaunese (AY279519); kappa class: H. sapiens
(Q9Y2Q3), R. norvegicus (P24473), Xenopus tropicalis (AAH87819), G. gallus (XP_416525), Ps. putida (Q51948). MAPEG: H. sapiens
(P10620), B. taurus (Q64L89), D. melanogaster (AAN85305), Tetraodon nigroviridis (CAF97117), Synechocystis sp. (P73795), Acaryochl-
oris marina (YP_001518348), E. coli (P64515), V. cholerae (NP_232423).
Bacterial GSTs N. Allocati et al.
60 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
Because kappa class and MAPEG proteins are evolu-
tionarily distant from cGSTs two separate trees are
presented as insets in the Fig. 1. Bacterial MAPEG
proteins group with eukaryotic members, whereas
Pseudomonas putida HCCA isomerase groups with
mammalian kappa class members. Fosfomycin
resistance proteins show high divergence in primary
sequence and in structure, thus they were not
considered.
This classification is in agreement with the evolu-
tionary model proposed by Frova [11]. According to
this model, the ancestor from which GSTs originated
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 61
is thioredoxin. The first phase of the evolution of
GSTs from thioredoxins saw the addition at the
C-terminus, or the insertion between the thioredoxin
fold, of an all-helical domain to produce two inter-
mediates: bacterial glutaredoxin 2 (GRX2), which
evolved into the cGSTs; and bacterial disulfide-bond-
forming oxidoreductase A (DsbA), which evolved
into the kappa GSTs. In a second phase, cGSTs fur-
ther diversified into different classes. Beta class
enzymes evolved after the dimerization of GRX2.
This model is supported by the ability of beta class
enzymes to act as thiol disulfide oxidoreductases as
well as GSH-conjugating enzymes and by the pres-
ence of a cysteine residue in the GSH site [14,68,69].
Finally, a shift from cysteine to serine chemistry
resulted in the theta class, followed by the zeta class
and then by all others classes. To date, there have
are no clear hypotheses about the evolution of
MAPEG family members [11].
Structures
Most of the bacterial GSTs identified to date belong
to the bacterial-specific beta class and since 1998 the
crystal structures of four representatives of this class
have been determined, i.e. P. mirabilis GST (PmGST)
[12], E. coli GST (EcGST) [37], B. xenovorans BphK
(BxGST) [47] and O. anthropi GST (OaGST) [41].
This, together with extensive site-directed mutagenesis
analysis, has allowed us to dissect in detail the
structural and catalytic properties of beta class
enzymes.
Beta class GSTs are homodimers with a chain length
of 201–203 residues. They display the canonical
GST fold with a thioredoxin-like N-terminal domain
followed by an all-helical C-terminal domain separated
by a short linker (Fig. 2A).
Comparative analysis of the crystal structures reveals
that the overall fold in beta class GSTs is remarkably
well conserved. Root mean square deviations among
equivalent Cas are generally < 1.5 A
˚when protein
monomers are superimposed, even though sequence
identities can be < 35–40%. When beta class monomers
are superimposed onto cGSTs belonging to other classes
rmsd values increase, ranging from 1.85 to 2.67 A
˚; these
values indicate that the canonical fold has been substan-
tially maintained from bacteria to mammals. Also the
monomers’ orientation in the different dimers is remark-
ably conserved (Fig. 2B). For example, when superim-
posing the OaGST dimer to the PmGST, EcGST and
BxGST dimers, rmsd values of 1.614, 1.418 and 1.54 A
˚,
respectively, are obtained [41]. These values are very
close to those obtained by superimposing the monomers
alone suggesting that strict conservation of the relative
orientation of the monomers in the dimer is required for
function.
Although the monomer fold in beta class GSTs
closely resembles that of other cGSTs, the dimer inter-
face is quite different. In fact, in contrast to other
GST classes, there is no open V-shaped interface,
although a close-packed arrangement is found, shaped
by the presence of stabilizing contacts at both the base
and top of the helical pairing that builds the interface
[12]. Furthermore, the interface in the majority of GST
classes is mainly hydrophobic in nature, whereas in
beta class GSTs it is markedly polar. As a conse-
quence, the well-known lock-and-key motif of alpha,
pi and mu class GSTs, comprising a protruding aro-
matic residue of one monomer that fits into a hydro-
phobic pocket in the other, is absent in beta class
GSTs [12,41]. Conversely, other structural motifs origi-
nally identified in mammalian GSTs are conserved in
beta class GSTs indicating their ancient origin. Among
them, an important stabilizing role is played by the
so-called N-capping box and hydrophobic staple motifs
at the N-terminus of the sixth helix in the C-terminal
domain that contribute to the interaction with the N-
terminal thioredoxin-like domain [36]. An additional
structural motif has been identified in beta class GSTs
that appears to be restricted to this class. This motif is
formed by a network of hydrogen bonds, located in
the proximity of the G-site, which zippers the terminal
helix of the C-terminal domain to the starting helix of
the thioredoxin-like domain. Partial disruption of this
motif has been shown to have dramatic consequences
on the both the stability and the functionality of
OaGST [41].
GSH binds to beta class GSTs in an extended fash-
ion consistent with what is observed in other classes.
Several interactions, both polar and hydrophobic, con-
tribute to its stabilization, including a hydrogen bond
with an aspartate residue from the other monomer
[12,41].
All beta class GSTs are characterized by the pres-
ence of a conserved cysteine residue located close to
the GSH sulfydryl group. Notably a mixed disulfide
bond was found in the structure of PmGST, with the
two sulfurs located at a distance of 2.2 A
˚(Fig. 2C)
[12]. This oxidized state, however, was not found in
the structures of OaGST [41] and BxGST [47], where
the average distance between sulfur atoms is 3.3 A
˚,
consistent with GSH being in the thiolate form and
sharing a hydrogen with the cysteine sulfur of the pro-
tein. In PmGST, two other residues are at hydrogen
bond distance from the GSH sulfydryl group, a
histidine (His106) and a serine (Ser9) (Fig. 2C). The
Bacterial GSTs N. Allocati et al.
62 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
histidine is conserved in other beta class GSTs,
whereas Ser9 is not conserved in OaGST, BxGST and
EcGST [37,41,47]. Interestingly, mutagenesis data on
PmGST and EcGST indicated that none of the above-
mentioned residues (Ser9, Cys10 and His106) is strictly
required for the classical conjugating activity or is
solely responsible for the reduced pK
a
of GSH, sug-
gesting that, in contrast with mammalian GSTs, the
stabilization of the GSH thiolate is likely the result of
a network of interactions [31,33,70]. OaGST deviates
partly from this behaviour. In fact, in OaGST the
G-site histidine adopts a different orientation with
respect to the other enzymes and does not bind GSH
[41]. In OaGST, in contrast to what is observed with
PmGST and EcGST, mutation of the conserved cyste-
ine to alanine led to a marked decrease in specific
activity, due to a dramatic loss in affinity for GSH
[42]. Interestingly, it was shown that, in this mutant,
GSH preferentially binds in a non-canonical position,
occupying the H-site [42]. Another beta class enzyme,
BxGST, was also shown to be able to bind a GSH in
the H-site [47]. Remarkably this enzyme was crystal-
lized in the presence of two GSH molecules, one with
the canonical G-site orientation and the other bound
at the H-site (Fig. 2D). In this structure the G- and
H-site GSH sulfurs are close (4.5 A
˚). These struc-
tural data support a mechanism in which once a disul-
fide bond is formed between the G-site GSH and
Cys10 in the reduction of several substrates, a second
GSH enters the H-site and a disulfide bond exchange
takes place with the formation and release of an
oxidized glutathione (GSSG), thus restoring the
enzyme’s functionality [47].
Two bacterial proteins show significant sequence
identity with mitochondrial kappa class GSTs. They
are DsbA and the HCCA isomerase. Kappa class
enzymes are peculiar in that, although they contain a
thioredoxin domain and an all-helical domain, their
topology differs from that of cytosolic enzymes
because the all-helical domain is inserted into the
thioredoxin fold. DsbA is a structural but not a func-
tional homologue of kappa class GSTs because its
active site does not bind GSH and it contains a
CXXC motif to perform redox catalysis [71]. HCCA
isomerase is the fourth enzyme in the naphthalene
catabolic pathway of Ps. putida. It catalyses the
conversion of HCCA, derived from cis-o-hydroxyben-
zeylidene pyruvic acid, to the more stable trans-o-
hydroxybenzeylidene pyruvic acid (Fig. 3I). Recently,
this enzyme has been subjected to extensive structural
and functional characterization that has shed light on
its peculiarities [63]. HCCA isomerase is a bona fide
AB
CD
Fig. 2. Structural studies on beta class
GSTs. (A) Structure of OaGST shown with
the twofold symmetry axis perpendicular to
the page (pdb code: 2NTO). The N-terminal
thioredoxin domain and the C-terminal all
helical domain are highlighted with different
tonalities. GSH molecules are shown in
sticks. (B) Overlay of four different beta
class GSTs dimers shown as Catraces: Oa-
GST (red), BxGST (green, pdb code: 2GDR),
EcGST (magenta, pdb code: 1A0F) and
PmGST (cyan, pdb code: 1PMT). This repre-
sentation highlights the conservation of beta
class fold as well as of the monomers’ ori-
entation in the dimers. (C) GSH binding site
in PmGST. In this crystal structure, GSH
(green carbons) forms a mixed disulfide with
Cys10. His106 and Ser9 are also at hydro-
gen bond distance from the GSH sulfur. (D)
Structure of BxGST in complex with two
GSH molecules (green carbons), one at the
canonical G-site and the other bound at the
hydrophobic substrates binding site (H-site).
The two GSH sulfurs are at close distance
from each other. This figure was prepared
with PYMOL (DeLano Scientific).
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 63
Bacterial GSTs N. Allocati et al.
64 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
kappa class GST: it conserves the same topology as
the mammalian enzyme, it is able to bind GSH,
which is activated through a conserved serine residue,
and it performs the classic conjugating reaction on
CDNB. Its structure superimposes well onto the
mammalian kappa class enzyme with only minor dif-
ferences and overall rmsd values among equivalent
Caof 2.2 A
˚. GSH is bound and stabilized in a
similar manner to mammalian kappa class GSTs but
more tightly, due to two additional interactions in
HCCA isomerase. A possible mechanism for the
HCCA isomerase reaction has been proposed in
which GSH performs a nucleophilic attack on the
HCCA ring and is then released thus behaving as a
cofactor (see later).
Structural genomics initiatives are increasing our
understanding of the protein-fold evolution and often
lead to the functional annotation of proteins whose
role was previously undetectable or merely hypotheti-
cal, based on sequence alignments alone. This is the
case of the Atu5508 gene product of Ag. tumefaciens
whose structure has recently been determined [67].
The protein has a dimeric organization, it binds GSH
and displays the canonical GST fold even though its
sequence identity is < 20% in pairwise comparisons
with any previously characterized cGSTs. The protein
is also a functional GST because it is able to conju-
gate GSH to p-benzyl chloride, a GST substrate.
Although coming from a bacterium, protein-fold anal-
ysis and structure superpositions suggest that this
protein is more closely related to mammalian GSTs
than to other bacterial GSTs. In particular, it lacks
the hallmarks of beta class GSTs, including absence
of the catalytic cysteine and histidine in the G-site
and the presence of a hydrophobic lock-and-key
motif at the dimer interface. For these reasons it has
been proposed that this protein is the prototype of a
new class of bacterial GSTs that includes several close
homologues found analysing a large set of environ-
mental sequences in the environmental sequencing
project [67].
Functions
Bacterial GSTs are specialized in several detoxification
processes. They are able to detoxify a large number of
molecules via GSH conjugation (an example with the
classic substrate CDNB is shown in Fig. 3A). They
have an active role in protecting against oxidative
stress and are involved in the detoxification of antimi-
crobial agents. Some are implicated in the basal
metabolism and supply bacterial cells with carbon
sources. Bacterial GSTs are also involved in the
degradation of several monocyclic aromatic com-
pounds such as toluene, xylenes, phenols and atrazine
[72]. They also take part in the degradation of poly-
cyclic aromatic hydrocarbons, a class of hazardous
chemicals to both environmental and human health
[21,73–78].
Oxidative and xenobiotic stress
Beta class cGSTs are involved in detoxication reactions
against toxic effects of several xenobiotics [39,40,
79–81].
In a modulation study it was reported that, in a
P. mirabilis strain, PmGST contributed to protect the
cells against oxidative stress induced by hydrogen per-
oxide [80]. Increases in the level of mRNA transcrip-
tion and in protein expression levels were observed
when the bacterial cells were exposed to hydrogen per-
oxide. This result was confirmed by the analysis of a
gstB gene knock-out in the same P. mirabilis strain
that was found to be much more sensitive to hydrogen
peroxide than the wild-type strain [80]. Similar results
were also obtained for E. coli GST [59].
The modulation of OaGST in O. anthropi in the
presence of different xenobiotics was also investigated.
Unlike PmGST, hydrogen peroxide did not influence
the induction of the enzyme. Atrazine caused an
increase in the expression of OaGST at low, non-toxic
concentrations, suggesting its involvement in atrazine
metabolism [39]. Instead, phenolic compounds induced
a marked dose-dependent enhancement of the enzy-
matic cellular levels correlated to the toxicity of the
molecules indicating a role of OaGST in cellular
protection [39]. These data were also corroborated by
the preponderant presence of the enzyme in the peri-
plasmic space when bacteria were exposed to 4-chloro-
phenol [40].
However, although OaGST was found in the peri-
plasm, no signal sequence for export by the general
secretory (Sec) pathway was found [40]. Recently, an
alternative pathway has also been described, the twin-
Fig. 3. Reactions catalysed by bacterial GSTs. (A) Canonical conjugation of CDNB as second substrate; (B) dehalogenation of DCM to form-
aldehyde; (C) reductive dehalogenation of TCHQ to trichlorohydroquinone and of trichlorohydroquinone to 2,6-dichlorohydroquinone; (D) cis
trans isomerization of maleylacetoacetate; (E) dehalogenation of 3-chloro-HOPDA; (F) hypothetical metabolic route for atrazine; (G) opening
of epoxide ring and reductive removal of GSH in isoprene metabolism; (H) b-aryl ether cleavage pathway; (I) step 4 of naphthalene catabolic
pathway; (J) opening of the epoxide ring of fosfomycin.
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 65
arginine translocation (Tat) system, which allows the
transfer of folded proteins into periplasmic space [82].
Proteins transported via Tat system possess a distinc-
tive motif (SRRXFLK) at their N-terminus. This motif
was also not present in the OaGST sequence [83]. Sim-
ilarly, no Sec or Tat signals were found in PmGST,
EcGST, BxGST and Atu5508 [83,84]. Therefore,
the mechanism of OaGST transport (and possibly that
of other GSTs) into the periplasm is uncharacterized
at present. Moreover, a prokaryotic GSH transporter
in E. coli has recently been identified [85]. This
strongly suggests the presence of GSH in the peri-
plasm, as hypothesized previously [86], and underlines
that bacterial GSTs also display activity in this
compartment.
PmGST is able to catalyse a glutaredoxin-like reac-
tion using cysteine S-sulfate and hydroxyethyl disulfide
as substrates [68]. The G-site cysteine residue is essen-
tial for this redox activity [68] and the mixed disulfide
observed in the PmGST crystal structure has been
suggested as an intermediate in glutaredoxin-like
reactions [68].
Bacterial GSTs in dehalogenation
Microbial dehalogenases play a key role in the biodeg-
radation of several chlorinated xenobiotics, both ali-
phatic and aromatic [87,88]. Halogenated compounds
are widely used in industry and constitute an impor-
tant group of environmental pollutants.
Bacterial GSTs catalyse different reactions using
GSH as a cofactor, i.e. DCM dehalogenases catalyse
the hydrolytic dechlorination of dichloromethane,
whereas TCHQ dehalogenase catalyses a reductive
dehalogenation reaction.
DCM dehalogenases
DCM dehalogenase is a GSH-dependent enzyme syn-
thesized by a number of facultative methylotrophic
bacteria that are able to utilize DCM as a sole carbon
and energy source [48]. In the first step of degradation,
the enzyme dechlorinates DCM to formaldeyde and
inorganic chloride (Fig. 3B). DCM, a significant envi-
ronmental contaminant, is widely used as an industrial
solvent. The properties of two DCMDs are well docu-
mented in Methylobacterium dichloromethanicum DM4
and in Methylophilus leisingeri DM11 [48–52]. Both de-
halogenases are closely related to eukaryotic theta class
GSTs (Fig. 1). In particular, like theta class GSTs,
DCMDs share a conserved serine residue at the N-ter-
minal domain that is essential for catalysis [89]. Never-
theless, on the basis of other criteria such as N-
terminal amino acid sequences, kinetic and immuno-
logical properties, they have been further subdivided
into two classes. One class is formed by group A
enzymes, including the DCM dehalogenases of Methy-
lobacterium dichloromethanicum DM4. By contrast,
M. leisingeri DM11 is the prototype for group B
enzymes. The most significant difference between the
two groups lies in their kinetic properties. Under con-
ditions of substrate saturation, DM11 is significantly
faster in dechlorination than DM4.
To date, no structure is known for DCMDs. A 3D
homology model of DCMD from the M. leisingeri
DM11 strain has been presented based on alignments
with GST members from the theta and the alpha class
[90].
TCHQ dehalogenase
TCHQ dehalogenase catalyses the reduction of TCHQ
to trichlorohydroquinone and then to dichlorohydro-
quinone during the biodegradation of pentachlorophe-
nol, a xenobiotic compound present in the
environment, utilized primarily as fungicide for wood
preservation [91–94]. The catalytic mechanism of the
reaction was exhaustively studied by Copley et al.
[54,56,95]. TCHQ dehalogenase was purified by
Sphingobium chlorophenolicum, a soil bacterium that
can grow on pentachlorophenol as a sole carbon
source [92,96]. TCHQ dehalogenase utilizes two GSH
molecules to reductively dechlorinate TCHQ to trichlo-
rohydroquinone and then dichlorohydroquinone
(Fig. 3C). First, TCHQ dehalogenase catalyses the
nucleophilic attack of GSH on the substrate forming a
GSH conjugate, and then converts the conjugate to
reduced products. A conserved cysteine is required to
this second step of the reaction through the formation
of a mixed disulfide with GSH [54,97].
TCHQ dehalogenase also has maleylacetoacetate
isomerase activity, and the conserved cysteine is also
required for this activity [53]. Maleylacetoacetate
isomerases are enzymes that catalyse the GSH-depen-
dent cistrans isomerization of maleylacetoacetate to
fumarylacetoacetate during the catabolism of tyrosine
and phenylalanine (Fig. 3D). Although the overall
sequence identity between TCHQ dehalogenase and
maleylacetoacetate isomerase is low, the active site is
highly conserved and in both cases contains a catalytic
cysteine [53].
Because maleylacetoacetate isomerases are part of
an ancient degradation pathway, whereas TCHQ
dehalogenase is a more specialized enzyme, it was
supposed that TCHQ dehalogenases evolved from
maleylacetoacetate isomerases [53].
Bacterial GSTs N. Allocati et al.
66 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
Biphenyl/polychlorinated
biphenyldehalogenation
A cGST characterized from B. xenovorans strain
LB400, a biphenyl/polychlorinated biphenyl-degrading
microorganism, known as BphK or BxGST, is one of
the most extensively investigated bacterial GSTs
[21,43–47]. This enzyme, which belongs to the beta
class, is encoded by a gene (bphK) located within the
bph locus, which specifies the degradation pathway of
biphenyl and chlorobiphenyl compounds. B. xenovo-
rans LB400 is able to attack a broad spectrum of poly-
chlorinated biphenyl congeners and is able to grow on
biphenyl as sole carbon source [21,43]. Mutagenic
studies indicated that although the bphK gene is not
essential for utilization of this carbon source its expres-
sion gives advantage to the strain in the utilization of
biphenyls [43]. Subsequently, it has been established
that BphK catalyses efficiently the dehalogenation of
3-chloro-2-hydroxy-6-oxo-6-phenyl-2,4-dienoate com-
pounds that are produced by the co-metabolism of
polychlorinated biphenyls (Fig. 3E) [46]. The proposed
reaction mechanism is similar to that described for
TCHQ dehalogenase [56]. Although less efficient,
BphK enzyme also has dechlorination activity against
4-chlorobenzoate [44]. A bphK gene is also present in
Sphingomonas yanoikuyae B1 and its expression allows
Sph. yanoikuyae B1 to grow faster on m-toluate
[98,99].
Atrazine metabolism
GSTs are able to detoxify several classes of herbicides
including triazines, a class of compounds to which
atrazine, one of the most widely used herbicides,
belongs [100,101]. GSTs are involved in the first step
of atrazine biodegradation with the removal of the
chlorine atom produced by atrazine–GSH conjugation
[100,101]. Dechlorination is followed by the stepwise
removal of isopropylamine and ethylamine groups by
dealkylation [102]. In bacteria, atrazine can be degra-
dated either by a microbial consortium or by a single
microorganism [103,104]. For instance, in Pseudomonas
ADP, atrazine is metabolized to cyanuric acid by three
enzymatic steps. The first step is performed by atrazine
chlorohydrolase followed by two dealkylations [104].
O. anthropi is a soil bacterium that is able to grow
on atrazine utilizing it as source of carbon [105] and
which expresses a functional beta class GST [24]. In a
modulation study of OaGST in the presence of several
xenobiotics, Favaloro and co-workers also showed
an increase in the levels of enzyme when atrazine
was added to the exponentially growing cells of
O. anthropi, suggesting an involvement of OaGST in
atrazine conjugation with GSH [39]. A proposed atra-
zine degradation pathway by bacterial GSTs is showed
in Fig. 3F.
Isoprene metabolism
Two GSTs were purified from the isoprene-utilizing
bacterium Rhodococcus sp. strain AD45 and their func-
tional properties were characterized [106–108]. Both
enzymes are involved in the metabolism of isoprene, an
atmospheric reactive hydrocarbon that plays a role in
ozone, organic peroxides and carbonic monoxide for-
mation, and their genes are localized in the isoprene
degradation gene cluster [106,108]. The first GST,
encoded by the isoI gene, catalyses the GSH-dependent
ring opening of isoprene monoxide, the primary oxida-
tion product of isoprene (Fig. 3G). The GSH conjugate
1-hydroxy-2-glutathionyl-2-methyl-3-butene is then oxi-
dized in two consecutive steps to 2-glutathionyl-2-
methyl-3-butenoic acid by a dehydrogenase (IsoH). The
way in which isoprene degradation proceeds has not
been fully characterized. A convincing hypothesis is
that 2-glutathionyl-2-methyl-3-butenoic acid could be
converted to the corresponding CoA-thioester by a
racemase expressed by isoG gene, to allow the second
GST, which is encoded by the isoJ gene, to remove the
GSH molecule [108,109]. It is thought that IsoJ might
catalyse the reductive removal of GSH using a second
GSH molecule in a similar fashion as the TCHQ dehal-
ogenase (Fig. 3G) [54,91,108,109]. A relevant difference
from TCHQ dehalogenase is that the GSH conjugate is
formed several steps before its reduction, because the
removal of the GSH from 1-hydroxy-2-glutathionyl-2-
methyl-3-butene is energetically unfeasible [109]. More-
over, IsoI/GST is able to degrade halogen epoxides
such as 1,2-dichloroepoxyethanes and epichlorohydrin
[106,107]. In Rhodococcus AD45 a novel GST, namely
IsoILR1, was recently characterized with activity
towards cis-1,2-dichloroethylene epoxide and epoxypro-
pane [110].
Lignin degradation pathway
In Sph. paucimobilis SYK-6, the role of three tandemly
located genes, ligE,ligF and ligG, involved in the pro-
cess of lignin degradation, a fundamental step for the
earth’s carbon cycle, has been described [111–113].
Sph. paucimobilis is able to degrade a wide variety of
lignin compounds including b-aryl ether. The b-aryl
ether cleavage is a fundamental step in lignin degrada-
tion, because this intermolecular linkage is the most
abundant in lignin. LigE and LigF are enantioselective
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 67
GSTs that cleave the b-aryl ether linkage with con-
sumption of GSH [113]. LigG is instead a GSH lyase
that catalyses the elimination of GSH from the conju-
gate produced by LigF (Fig. 3H) [112,113].
Naphthalene metabolism
HCCA isomerase is a GSH-dependent enzyme, struc-
turally related to mammalian kappa class enzymes,
involved in the naphthalene degradation pathway
[5,62,63]. HCCA isomerase is also involved in naph-
thalene sulfonates and anthracene catabolism
[114,115]. Naphthalene catabolism consists of an upper
pathway and a lower pathway. In the upper pathway,
naphthalene is metabolized to salycilate in six steps.
HCCA isomerase is the fourth enzyme of this pathway
and catalyses cistrans isomerization between HCCA
and trans-o-hydroxybenzylidene pyruvic acid (Fig. 3I).
GSH is thought to be involved in this reaction by its
covalent addition of HCCA. The addition alters the
hybridization state at C7 promoting a rotation around
the C7–C8 double bond [63].
Interaction with antibiotics
Several studies on the interaction of PmGST with dif-
ferent classes of antibiotics have been performed and a
possible role for the enzyme in antibiotic-resistance
proposed [29,116].
First, it was observed that the efficiency of a number
of antimicrobial drugs decreased notably in the pres-
ence of the purified enzyme in the medium culture, as
shown by increased minimal inhibitory concentration
values [16,116]. By contrast, the presence of mamma-
lian GSTs had no effect on antibiotic efficiency [16].
Second, studies on the interaction of PmGST with sev-
eral antibiotics indicated that the enzyme sequesters
antimicrobial drugs with avidity [29]. The effect of 18
different antibiotics as inhibitors of PmGST activity
using CDNB and GSH as substrates, were also mea-
sured. Four of them, namely minocycline, tetracycline,
rifamycin and nitrofurantoin, were strong inhibitors
with IC
50
values between 49 and 140 lm. These drugs
produced a significant decrease in the k
cat
values of the
bacterial enzyme for both substrates [29]. Furthermore,
PmGST displays a protective action against antibiotics
also in vivo [80,117]. An increase in PmGST levels was
observed when bacteria were grown in the presence of
drugs [80]. Moreover, viability tests showed that a gst
null-mutant P. mirabilis strain was more sensitive to
antibiotics than the wild-type bacterium [80]. Finally,
crystallographic data highlighted the presence of a
hydrophobic cavity large enough to bind antibacterial
molecules located at the dimer interface [12]. These
results were strengthened by the preponderant periplas-
mic location of the enzyme in the periplasmic space
[81].
Fosfomycin resistance proteins
Fosfomycin ([1R,2S]-[1,2-epoxypropyl]-phosphonic acid)
is a bactericidal broad-spectrum antibiotic effective
against both Gram-negative and Gram-positive bacte-
ria. Fosfomycin inhibits the enzyme MurA (UDP-
NAG enolpyruvyl transferase), which catalyses the
transfer of enolpyruvate from phosphoenolpyruvate to
uridine diphospho-N-acetylglucosamine, the first com-
mitted step of bacterial cell-wall biosynthesis. Resis-
tance is mainly chromosomal but resistance genes have
also been found on transmissible plasmids [118]. Resis-
tance to fosfomycin can be achieved by several differ-
ent mechanisms, including decreased uptake of the
antibiotic, overexpression or mutation of MurA, and
enzymatic modification of the antibiotic [118,119].
FosA, FosB and FosX represent three mechanisti-
cally distinct classes of enzymes that confer resistance
to fosfomycin by adding GSH, l-cysteine or a hydro-
xyl group, respectively, to the oxirane ring of the
antibiotic, and inactivating it.
FosA was originally identified in strains carrying
fosfomycin-resistance plasmids obtained from clinical
isolates [120–122]. Suarez and co-workers described a
new mechanism of antibiotic resistance due to the
enzymatic modification of fosfomycin [121,122]. Inacti-
vation of fosfomycin occurred by the formation of an
adduct between its carbon 1 atom and the sulfydryl
group of the GSH cysteine resulting in the opening of
the epoxide ring of the antibiotic (Fig. 3J). The reac-
tion was catalysed by an enzyme that was referred to
as a GST. The enzyme was purified and characterized.
It did not bind to the GSH–agarose matrix and did
not catalyse the reaction between GSH and CDNB,
indicating that this protein had different properties
from the canonical GSTs. The enzyme is a homodimer
of 32 kDa and its activity is dependent on the addition
of the Mn
2+
cation [122]. In subsequent studies, Arm-
strong and co-workers demonstrated that FosA is a
metalloglutathione transferase related to glyoxalase I
and extradiol dioxygenases, members of the vicinal
oxygen chelate superfamily [4,123]. In addition, they
showed that each subunit of the homodimer contains a
mononuclear Mn
2+
centre that interacts strongly with
the antibiotic and also that the enzyme requires a
monovalent cation K
+
for optimal catalytic activity
[123]. FosA is also encoded in the bacterial genomes
and the 3D structure of a genomically encoded FosA
Bacterial GSTs N. Allocati et al.
68 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
from the pathogen Ps. aeruginosa was described [124].
The protein fold showed similarity to members of vici-
nal oxygen chelate superfamily and consists of paired
babbb motifs that form a cupped-shaped cavity for the
metal ion-binding site. The K
+
ion is accommodated
in a loop located 6.5 A
˚from the Mn
2+
cation [124].
The structure of FosA in complex with antiobiotic was
also obtained. Plasmid-encoded and genomically
encoded enzymes are very similar in structure. The
crystal structure of FosA from transposon Tn2921
maintains the same basic molecular arrangement
observed in genomically encoded FosA [125]. Recently,
based on structural data, the residues involved in the
binding of both fosfomycin and GSH substrates have
been identified and characterized by mutagenesis
[126,127].
FosB was originally identified in fosfomycin-resis-
tant Staphylococcus strains and the genes were encoded
by plasmids [128–130]. A second type of FosB was
identified in the genome of a Bacillus subtilis strain
and subsequently characterized [131]. The protein is a
dimeric metallothiol transferase related to FosA.
Unlike FosA, FosB utilizes Mg
2+
as metal cofactor.
In addition, FosB uses l-cysteine rather than GSH as
thiol donor and is less efficient than FosA [131]. This
is in agreement with previous studies because the GSH
molecule was not detectable in most of the Gram-posi-
tive bacteria tested, including Bacillus subtilis [132].
FosX is a Mn(II)-dependent fosfomycin specific
epoxide hydrolase. It catalyses the addition of a water
molecule to fosfomycin, thus inactivating it. The
enzyme is genomically encoded, it is found in several
microorganisms and it has been well characterized
in Mesorhizobium loti and Lysteria monocytogenes
[133,134]. The two enzymes showed significant func-
tional differences. Listeria FosX is a good catalyst and
is responsible for high resistance to fosfomycin [133].
By contrast, M. loti FosX produces modest resistance
to the antibiotic and, consistently, its kinetic costants
are lower than those of Listeria FosX. In addition,
M. loti FosX also catalyses the addition of GSH to
the antibiotic even if with low efficiency. It has been
suggested that M. loti FosX may be an intermediate in
the evolution of fosfomycin resistance proteins that
plays some yet to be identified role in the catabolism
of phosphonates. The structure of FosX was also
determined for both enzymes [133,134]. The FosX
structures are closely related and the overlay with
FosA from Ps. aeruginosa [124] shows a large degree
of similarity. Unlike FosA, the enzymes do not contain
aK
+
ion-binding site near the active site. The
most interesting aspect of these structures is the obser-
vation that fosfomycin binds to FosX enzymes in a
different orientation from that observed in the FosA
enzyme.
A new mechanism of fosfomycin inactivation was
described by Garcia et al. [135] in a fosfomycin-resis-
tant strain of Pseudomonas syringae. The bacterium
yielded an enzyme, FosC, that inactivated the anti-
biotic using ATP to phosphorylate fosfomycin in the
presence of Mg
2+
. This finding was corroborated by
sequence alignments highlighting a region of partial
homology between FosC and the Mg-ATP binding
domains of AMP–ATP phosphotransferases. To date,
no relationship between FosC and the other fosfo-
mycin resistance proteins is known.
Potential applications of bacterial
GSTs
As previously described, bacterial GSTs are involved
in several types of chemical transformations and may
represent a versatile tool with a variety of biotechno-
logical applications, for example, in the field of bio-
remediation, an economical alternative to conventional
physicochemical methods to clean up environmentally
contaminated sites. The relative ease of genetic manip-
ulations in bacteria and their ability to grow rapidly
constitute a further advantage.
Several studies have been carried out exploiting the
potential of GSTs using both purified proteins and
microorganism engineering, some of which are summa-
rized here.
An example of protein engineering using the DNA-
shuffling technique was shown by Mannervik et al.
[136]. They hybridized six alpha class GSTs of differing
mammalian origin obtaining chimeric enzymes with
improved catalytic properties and altered substrate
selectivity towards several noxious iodoalkanes.
Another example is in the engineering of fusion pro-
teins with several distinct enzymatic activities. For
example, a trifunctional enzyme with superoxide
dismutase, glutathione peroxidase and glutathione
transferase activities was recently generated [137]. This
recombinant chimeric enzyme was shown to be effec-
tive in scavenging reactive oxygen species, thus show-
ing that this approach may have several applications
in medicine as well as in environmental field.
Another potential application lies in the preparation
of biosensors. These are detection systems widely used
to check contaminated environments that combine a
biological component with a detector element. Biosen-
sors are competitive systems in comparison to conven-
tional methods being inexpensive, easy to use and
characterized by high sensitivity and selectivity. For
example, a mammalian GST was used to develop an
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 69
optical biosensor for detection of captan in contami-
nated waters [138]. Captan is used to control a broad
spectrum of plant pathogenic microorganisms and it is
a strong inhibitor of GSTs [139].
A large number of bacterial species have developed
the ability to degrade xenobiotics previously considered
to be non-degradable [140]. Examples of microbial con-
sortia combining the ability of two or more bacterial
species to metabolize one or more noxious molecules
have been described [141]. An alternative route is to
engineer a single bacterial strain to carry a complete
metabolic pathway that efficiently eliminates environ-
mental toxic compounds [110]. For example, Wood
et al. [110] engineered in E. coli a metabolic pathway to
improve the degradation of chlorinated ethenes, which
constitute a large group of toxic environmental pollu-
tants [142]. The degradation of these molecules is limited
by the accumulation of reactive intermediates epoxides.
The authors constructed a recombinant E. coli strain in
which a toluene ortho-monooxygenase from Burkholde-
ria cepacia G4, obtained by DNA shuffling, and a GST
from Rhodococcus AD45 with activity towards cis-1,2,-
dichloroethylene epoxide and epoxypropane, were
co-expressed [110]. The ability of GST to transform
epoxides in E. coli strain increased the mineralization of
cis-1,2,-dichloroethylene.
The examples provided above highlight the potential
biotechnological applications of using engineered
proteins and bacterial strains. In this respect, bacterial
GSTs, which are characterized by high stability and by
a wide variety of catalysed reactions, undoubtedly,
constitute an effective resource for the future.
Acknowledgements
This work was supported in part by grants from
the Ministero dell’Istruzione, dell’Universita
`e della
Ricerca (MIUR) of Italy.
References
1 Hayes JD, Flanagan JU & Jowsey IR (2005) Glutathi-
one transferases. Annu Rev Pharmacol Toxicol 45,
51–88.
2 Sheehan D, Meade G, Foley VM & Dowd CA (2001)
Structure, function and evolution of glutathione trans-
ferases: implications for classification of non-mamma-
lian members of an ancient enzyme superfamily.
Biochem J 360, 1–16.
3 Oakley AJ (2005) Glutathione transferases: new
functions. Curr Opin Struct Biol 15, 716–723.
4 Armstrong RN (2000) Mechanistic diversity in a metallo-
enzyme superfamily. Biochemistry 39, 13625–13632.
5 Robinson A, Huttley GA, Booth HS & Board PG
(2004) Modelling and bioinformatics studies of the
human kappa-class glutathione transferase predict a
novel third glutathione transferase family with
similarity to prokaryotic 2-hydroxychromene-2-carbox-
ylate isomerases. Biochem J 379, 541–552.
6 Jakobsson PJ, Morgenstern R, Mancini J, Ford-Hutch-
inson A & Persson B (1999) Common structural fea-
tures of MAPEG – a widespread superfamily of
membrane-associated proteins with highly divergent
functions in eicosanoid and glutathione metabolism.
Protein Sci 8, 689–692.
7 Shishido T (1981) Glutathione S-transferase from Esc-
herichia coli.Agric Biol Chem 45, 2951–2953.
8 Piccolomini R, Aceto A, Allocati N, Faraone A & Di
Ilio C (1991) Purification of a GSH-affinity binding
protein from Bacteroides fragilis devoid of glutathione
transferase activity. FEMS Microbiol Lett 82, 101–106.
9 Fahey RC (2001) Novel thiols of prokaryotes. Annu
Rev Microbiol 55, 333–356.
10 Wiktelius E & Stenberg G (2007) Novel class of gluta-
thione transferases from cyanobacteria exhibit high
catalytic activities towards naturally occurring isothio-
cyanates. Biochem J 406, 115–123.
11 Frova C (2006) Glutathione transferases in the genom-
ics era: new insights and perspectives. Biomol Eng 23,
149–169.
12 Rossjohn J, Polekhina G, Feil SC, Allocati N, Masulli
M, Di Ilio C & Parker MW (1998) A mixed disulfide
bond in bacterial glutathione transferase: functional
and evolutionary implications. Structure 15, 721–734.
13 Vuilleumier S (1997) Bacterial glutathione S-transfer-
ases: what are they good for? J Bacteriol 179, 1431–
1441.
14 Di Ilio C, Aceto A, Piccolomini R, Allocati N, Fara-
one A, Cellini L, Ravagnan G & Federici G (1988)
Purification and characterization of three forms of glu-
tathione transferase from Proteus mirabilis.Biochem J
255, 971–975.
15 Iizuka M, Inoue Y, Murata K & Kimura A (1989)
Purification and some properties of glutathione S-
transferase from Escherichia coli B. J Bacteriol 171,
6039–6042.
16 Piccolomini R, Di Ilio C, Aceto A, Allocati N, Fara-
one A, Cellini L, Ravagnan G & Federici G (1989)
Glutathione transferase in bacteria: subunit composi-
tion and antigenic characterization. J Gen Microbiol
135, 3119–3125.
17 Arca P, Garcia P, Hardisson C & Suarez JE (1990)
Purification and study of a bacterial glutathione
S-transferase. FEBS Lett 263, 77–79.
18 Di Ilio C, Aceto A, Piccolomini R, Allocati N, Fara-
one A, Bucciarelli T, Barra D & Federici G (1991)
Purification and characterization of a novel glutathione
Bacterial GSTs N. Allocati et al.
70 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
transferase from Serratia marcescens.Biochim Biophys
Acta 1077, 141–146.
19 Di Ilio C, Aceto A, Allocati N, Piccolomini R, Bucc-
iarelli T, Dragani B, Faraone A, Sacchetta P, Petruzz-
elli R & Federici G (1993) Characterization of
glutathione transferase from Xanthomonas campestris.
Arch Biochem Biophys 305, 110–114.
20 Nishida M, Kong KH, Inoue H & Takahashi K (1994)
Molecular cloning and site-directed mutagenesis of
glutathione S-transferase from Escherichia coli.
The conserved tyrosyl residue near the N-terminus is
not essential for catalysis. J Biol Chem 269, 32536–
32541.
21 Hofer B, Backhaus S & Timmis KN (1994) The biphe-
nyl/polychlorinated biphenyl-degradation locus (bph)
of Pseudomonas sp. LB400 encodes four additional
metabolic enzymes. Gene 144, 9–16.
22 Zablotowicz RM, Hoagland RE, Locke MA & Hickey
WJ (1995) Glutathione S-transferase activity and
metabolism of glutathione conjugates by rhizosphere
bacteria. Appl Environ Microbiol 61, 1054–1060.
23 Jung U, Cho YS, Seong HM, Kim SJ, Kim YC &
Chung AS (1996) Characterization of a novel glutathi-
one S-transferase from Pseudomonas sp DJ77. J Bio-
chem Mol Biol 29, 111–115.
24 Favaloro B, Tamburro A, Angelucci S, De Luca A,
Melino S, Di Ilio C & Rotilio D (1998) Molecular
cloning, expression and site-directed mutagenesis of
glutathione S-transferase from Ochrobactrum anthropi.
Biochem J 335, 573–579.
25 Travensolo RF, Garcia W, Muniz JR, Caruso CS,
Lemos EG, Carrilho E & Arau´ jo AP (2008) Cloning,
expression, purification and characterization of recom-
binant glutathione S-transferase from Xylella fastidi-
osa.Protein Expr Purif 59, 153–160.
26 Mignogna G, Allocati N, Aceto A, Piccolomini R, Di
Ilio C, Barra D & Martini F (1993) The amino acid
sequence of glutathione transferase from Proteus
mirabilis, a prototype of a new class of enzymes. Eur
J Biochem 211, 421–425.
27 Sacchetta P, Aceto A, Bucciarelli T, Dragani B, Santa-
rone S, Allocati N & Di Ilio C (1993) Multiphasic
denaturation of glutathione transferase B1-1 by guan-
idinium chloride. Role of the dimeric structure on the
flexibility of the active site. Eur J Biochem 215, 741–
745.
28 Aceto A, Dragani B, Allocati N, Angelucci S, Bucciar-
elli T, Sacchetta P, Di Ilio C & Martini F (1995) Anal-
ysis by limited proteolysis of domain organization and
GSH-site arrangement of bacterial glutathione transfer-
ase B1-1. Int J Biochem Cell Biol 27, 1033–1041.
29 Perito B, Allocati N, Casalone E, Masulli M, Dragani
B, Polsinelli M, Aceto A & Di Ilio C (1996) Molecular
cloning and expression of a glutathione transferase
gene from Proteus mirabilis.Biochem J 318, 157–162.
30 Allocati N, Aceto A, Cellini L, Masulli M, Dragani B,
Petruzzelli R & Di Ilio C (1997) Effect of anaerobic
environment on the glutathione transferase isoenzy-
matic pattern in Proteus mirabilis.FEMS Microbiol
Lett 147, 157–162.
31 Casalone E, Allocati N, Ceccarelli I, Masulli M, Ross-
john J, Parker MW & Di Ilio C (1998) Site-directed
mutagenesis of the Proteus mirabilis glutathione trans-
ferase B1-1 G-site. FEBS Lett 423, 122–124.
32 Allocati N, Casalone E, Masulli M, Ceccarelli I, Car-
letti E, Parker MW & Di Ilio C (1999) Functional
analysis of the evolutionary conserved praline 53 resi-
due in Proteus mirabilis glutathione transferase B1-1.
FEBS Lett 445, 347–350.
33 Allocati N, Casalone E, Masulli M, Polekhina G, Ross-
john J, Parker MW & Di Ilio C (2000) Evaluation of the
role of two conserved active-site residues in Beta class
glutathione S-transferases. Biochem J 351, 341–346.
34 Allocati N, Masulli M, Casalone E, Santucci S, Faval-
oro B, Parker MW & Di Ilio C (2002) Glutamic acid
65 is an essential residue for catalysis in Proteus mira-
bilis glutathione S-transferase B1-1. Biochem J 363,
189–193.
35 Allocati N, Masulli M, Pietracupa M, Favaloro B,
Federici L & Di Ilio C (2005) Contribution of the two
conserved tryptophan residues to the catalytic and
structural properties of Proteus mirabilis glutathione
S-transferase B1-1. Biochem J 385, 37–43.
36 Allocati N, Masulli M, Pietracupa M, Federici L & Di
Ilio C (2006) Evolutionarily conserved structural motifs
in bacterial glutathione transferase are involved in
protein folding and stability. Biochem J 394, 11–17.
37 Nishida M, Harada S, Noguchi S, Satow Y, Inoue H
& Takahashi K (1998) Three-dimensional structure of
Escherichia coli glutathione S-transferase complexed
with glutathione sulfonate: catalytic roles of Cys10 and
His106. J Mol Biol 281, 135–147.
38 Favaloro B, Melino S, Petruzzelli R, Di Ilio C &
Rotilio D (1998) Purification and characterization of a
novel glutathione transferase from Ochrobac-
trum anthropi.FEMS Microbiol Lett 160, 81–86.
39 Favaloro B, Tamburro A, Trofino MA, Bologna L,
Rotilio D & Heipieper HJ (2000) Modulation of the
glutathione S-transferase in Ochrobactrum anthropi:
function of xenobiotic substrates and other forms of
stress. Biochem J 346, 553–559.
40 Tamburro A, Robuffo I, Heipieper HJ, Allocati N,
Rotilio D, Di Ilio C & Favaloro B (2004) Expression
of glutathione S-transferase and peptide methionine
sulphoxide reductase in Ochrobactrum anthropi is cor-
related to the production of reactive oxygen species
caused by aromatic substrates. FEMS Microbiol Lett
241, 151–156.
41 Federici L, Masulli M, Bonivento D, Di Matteo A,
Gianni S, Favaloro B, Di Ilio C & Allocati N (2007)
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 71
Role of Ser11 in the stabilization of the structure of
Ochrobactrum anthropi.Biochem J 403, 267–274.
42 Allocati N, Federici L, Masulli M, Favaloro B & Di
Ilio C (2008) Cysteine 10 is critical for the activity of
Ochrobactrum anthropi glutathione transferase and its
mutation to alanine causes the preferential binding of
glutathione to the H-site. Proteins 71, 16–23.
43 Bartels F, Backhaus S, Moore ER, Timmis KN &
Hofer B (1999) Occurrence and expression of glutathi-
one S-transferase-encoding bphK genes in Burkholderia
sp. strain LB400 and other biphenyl-utilizing bacteria.
Microbiology 145, 2821–2834.
44 Gilmartin N, Ryan D, Sherlock O & Dowling D
(2003) BphK shows dechlorination activity against 4-
chlorobenzoate, an end product of bph-promoted deg-
radation of PCBs. FEMS Microbiol Lett 222, 251–255.
45 Gilmartin N, Ryan D & Dowling DN (2005) Analysis
of the C-terminal domain of Burkholderia sp. strain
LB400 BphK reveals a conserved motif that affects
catalytic activity. FEMS Microbiol Lett 249, 23–30.
46 Fortin PD, Horsman GP, Yang HM & Eltis LD
(2006) A glutathione S-transferase catalyzes the de-
halogenation of inhibitory metabolites of polychlori-
nated biphenyls. J Bacteriol 188, 4424–4430.
47 Tocheva EI, Fortin PD, Eltis LD & Murphy MEP
(2006) Structures of ternary complexes of BphK, a
bacterial glutathione S-transferase that reductively
dechlorinates polychlorinated biphenyl metabolites.
J Biol Chem 281, 30933–30940.
48 Scholtz R, Wackett LP, Egli C, Cook AM & Leisinger
T (1988) Dichloromethane dehalogenase with improved
catalytic activity isolated from a fast-growing
dichloromethane-utilizing bacterium. J Bacteriol 170,
5698–5704.
49 La Roche S & Leisinger T (1990) Sequence analysis
and expression of the bacterial dichloromethane
dehalogenase structural gene, a member of the
glutathione S-transferase supergene family. J Bacteriol
172, 164–171.
50 Bader R & Leisinger T (1994) Isolation and character-
ization of the Methylophilus sp. strain DM11 gene
encoding dichloromethane dehalogenase/glutathi-
one S-transferase. J Bacteriol 176, 3466–3473.
51 Vuilleumier S, Ivos N, Dean M & Leisinger T
(2001) Sequence variation in dichloromethane
dehalogenase/glutathione S-transferases. Microbiology
147, 611–619.
52 Stourman NV, Rose JH, Vuilleumier S & Armstrong
RN (2003) Catalytic mechanism of dichloromethane
dehalogenase from Methylophilus sp. strain DM11.
Biochemistry 42, 11048–11056.
53 Anandarajah K, Kiefer PM Jr, Donohoe BS & Copley
SD (2000) Recruitment of a double bond isomerase to
serve as a reductive dehalogenase during biodegrada-
tion of pentachlorophenol. Biochemistry 39, 5303–5311.
54 McCarthy DL, Navarrete S, Willett WS, Babbitt PC &
Copley SD (1996) Exploration of the relationship
between tetrachlorohydroquinone dehalogenase and
the glutathione S-transferase superfamily. Biochemistry
35, 14634–14642.
55 Copley SD (2000) Evolution of a metabolic pathway
for degradation of a toxic xenobiotic: the patchwork
approach. Trends Biochem Sci 25, 261–265.
56 Kiefer PM Jr, McCarthy DL & Copley SD (2002) The
reaction catalyzed by tetrachlorohydroquinone dehalo-
genase does not involve nucleophilic aromatic substitu-
tion. Biochemistry 41, 1308–1314.
57 Vuilleumier S & Pagni M (2002) The elusive roles of
bacterial glutathione S-transferases: new lessons from
genomes. Appl Microbiol Biotechnol 58, 138–146.
58 Rife CL, Parsons JF, Xiao G, Gilliland GL & Arm-
strong RN (2003) Conserved structural elements in
glutathione transferase homologues encoded in the
genome of Escherichia coli.Proteins 53, 777–782.
59 Kanai T, Takahashi K & Inoue H (2006) Three
distinct-type glutathione S-transferases from
Escherichia coli important for defense against oxidative
stress. J Biochem 140, 703–711.
60 Hong G, Chien YC & Chien CI (2003) Separation of
glutathione transferase subunits from Proteus vulgaris
by two-dimensional gel electrophoresis. Curr Microbiol
47, 352–354.
61 Remmerie B, Vandenbroucke K, De Smet L, Carpen-
tier W, De Vos D, Stout J, Van Beeumen J & Savvides
SN (2008) Expression, purification, crystallization and
structure determination of two glutathione S-transfer-
ase-like proteins from Shewanella oneidensis.Acta
Crystallogr Sect F Struct Biol Cryst Commun 64, 548–
553.
62 Eaton RW (1994) Organization and evolution of naph-
thalene catabolic pathways: sequence of the DNA
encoding 2-hydroxychromene-2-carboxylate isomerase
and trans-o-hydroxybenzylidenepyruvate hydratase-
aldolase from the NAH7 plasmid. J Bacteriol 176,
7757–7762.
63 Thompson LC, Ladner JE, Codreanu SG, Harp J, Gil-
liland GL & Armstrong RN (2007) 2-Hydroxychrom-
ene-2-carboxylic acid isomerase: a kappa class
glutathione transferase from Pseudomonas putida.Bio-
chemistry 46, 6710–6722.
64 Jakobsson PJ, Morgenstern R, Mancini J, Ford-Hutch-
inson A & Persson B (2000) Membrane-associated
proteins in eicosanoid and glutathione metabolism
(MAPEG). A widespread protein superfamily. Am
J Respir Crit Care Med 161, S20–S24.
65 Bresell A, Weinander R, Lundqvist G, Raza H, Shim-
oji M, Sun TH, Balk L, Wiklund R, Eriksson J, Jans-
son C et al. (2005) Bioinformatic and enzymatic
characterization of the MAPEG superfamily. FEBS J
272, 1688–1703.
Bacterial GSTs N. Allocati et al.
72 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
66 Molina DM, Eshaghi S & Nordlund P (2008) Catalysis
within the lipid bilayer-structure and mechanism of the
MAPEG family of integral membrane proteins. Curr
Opin Struct Biol 18, 1–8.
67 Kosloff M, Han GW, Krishna SS, Schwarzenbacher
R, Fasnacht M, Elsliger MA, Abdubek P, Agarwalla
S, Ambing E, Astakhova T et al. (2006) Comparative
structural analysis of a novel glutathione S-transferase
(Atu5508) from Agrobacterium tumefaciens at 2.0 A
˚
resolution. Proteins 65, 527–537.
68 Caccuri AM, Antonini G, Allocati N, Di Ilio C, De
Maria F, Innocenti F, Parker MW, Masulli M, Lo
Bello M, Turella P et al. (2002) GST B1-1 from Pro-
teus mirabilis: a snapshot of an enzyme in the evolu-
tionary pathway from a redox enzyme to a conjugating
enzyme. J Biol Chem 277, 18777–18784.
69 Caccuri AM, Antonini G, Allocati N, Di Ilio C, Inno-
centi F, De Maria F, Parker MW, Masulli M, Polizio
F, Federici G et al. (2002) Properties and utility of the
peculiar mixed disulfide in the bacterial glutathione
transferase B1-1. Biochemistry 41, 4686–4693.
70 Inoue H, Nishida M & Takahashi K (2000) Effects of
Cys10 mutation to Ala in glutathione transferase from
Escherichia coli.J Org Chem 611, 593–595.
71 Martin JL, Bardwell JC & Kuriyan J (1993) Crystal
structure of the DsbA protein required for disulphide
bond formation in vivo.Nature 365, 464–468.
72 Santos PM, Mignogna G, Heipieper HJ & Zennaro E
(2002) Occurrence and properties of glutathi-
one S-transferases in phenol-degrading Pseudomonas
strains. Res Microbiol 153, 89–98.
73 Kanaly RA & Harayama S (2000) Biodegradation of
high-molecular-weight polycyclic aromatic hydrocar-
bons by bacteria. J Bacteriol 182, 2059–2067.
74 Mueller JG, Chapman PJ, Blattmann BO & Pritchard
PH (1990) Isolation and characterization of a fluo-
ranthene-utilizing strain of Pseudomonas paucimobilis.
Appl Environ Microbiol 56, 1079–1086.
75 Wang Y, Lau PC & Button DK (1996) A marine
oligobacterium harboring genes known to be part of
aromatic hydrocarbon degradation pathways of soil
pseudomonads. Appl Environ Microbiol 62, 2169–2173.
76 Lloyd-Jones G & Lau PC (1997) Glutathione S-trans-
ferase-encoding gene as a potential probe for environ-
mental bacterial isolates capable of degrading
polycyclic aromatic hydrocarbons. Appl Environ Micro-
biol 63, 3286–3290.
77 Xia Y, Min H, Rao G, Lv ZM, Liu J, Ye YF & Duan
XJ (2005) Isolation and characterization of phenan-
threne-degrading Sphingomonas paucimobilis strain
ZX4. Biodegradation 16, 393–402.
78 Cavalca L, Guerrieri N, Colombo M, Pagani S &
Andreoni V (2007) Enzymatic and genetic profiles in
environmental strains grown on polycyclic aromatic
hydrocarbons. Antonie Van Leeuwenhoek 91, 315–325.
79 Tamburro A, Allocati N, Masulli M, Rotilio D, Di Ilio
C & Favaloro B (2001) Bacterial peptide methionine
sulphoxide reductase: co-induction with glutathi-
one S-transferase during chemical stress conditions.
Biochem J 360, 675–681.
80 Allocati N, Favaloro B, Masulli M, Alexeyev MF & Di
Ilio C (2003) Proteus mirabilis glutathione S-transferase
B1-1 is involved in the protective mechanisms against
oxidative and chemical stresses. Biochem J 373, 305–311.
81 Allocati N, Cellini L, Aceto A, Iezzi T, Angelucci A,
Robuffo I & Di Ilio C (1994) Immunogold localization
of glutathione transferase B1-1 in Proteus mirabilis.
FEBS Lett 354, 191–194.
82 Sargent F, Berks BC & Palmer T (2006) Pathfinders and
trailblazers: a prokaryotic targeting system for transport
of folded proteins. FEMS Microbiol Lett 254, 198–207.
83 Bendtsen JD, Nielsen H, Widdick D, Palmer T & Bru-
nak S (2005) Prediction of twin-arginine signal pep-
tides. BMC Bioinform 6, 167.
84 Bendtsen JD, Nielsen H, von Heijne G & Brunak S
(2004) Improved prediction of signal peptides: SignalP
3.0. J Mol Biol 340, 783–795.
85 Pittman MS, Robinson HC & Poole RK (2005) A bac-
terial glutathione transporter (Escherichia coli CydDC)
exports reductant to the periplasm. J Biol Chem 280,
32254–32261.
86 Dartigalongue C, Nikaido H & Raina S (2000) Protein
folding in the periplasm in the absence of primary oxi-
dant DsbA: modulation of redox potential in periplas-
mic space via OmpL porin. EMBO J 19, 5980–5988.
87 Copley SD (1998) Microbial dehalogenases: enzymes
recruited to convert xenobiotic substrates. Curr Opin
Chem Biol 2, 613–617.
88 Janssen DB, Dinkla IJ, Poelarends GJ & Terpstra P
(2005) Bacterial degradation of xenobiotic compounds:
evolution and distribution of novel enzyme activities.
Environ Microbiol 7, 1868–1882.
89 Vuilleumier S & Leisinger T (1996) Protein engineering
studies of dichloromethane dehalogenase/glutathi-
one S-transferase from Methylophilus sp. strain DM11.
Ser12 but not Tyr6 is required for enzyme activity. Eur
J Biochem 239, 410–417.
90 Marsh A & Ferguson DM (1997) Knowledge-based
modeling of a bacterial dichloromethane dehalogenase.
Proteins 28, 217–226.
91 Xun L, Topp E & Orser CS (1992) Purification and
characterization of a tetrachloro-p-hydroquinone
reductive dehalogenase from a Flavobacterium sp.
J Bacteriol 174, 8003–8007.
92 Xun L, Topp E & Orser CS (1992) Glutathione is the
reducing agent for the reductive dehalogenation of tet-
rachloro-p-hydroquinone by extracts from a Flavobac-
terium sp. Biochem Biophys Res Commun 182, 361–366.
93 Orser CS, Dutton J, Lange C, Jablonski P, Xun L &
Hargis M (1993) Characterization of a Flavobacterium
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 73
glutathione S-transferase gene involved reductive
dechlorination. J Bacteriol 175, 2640–2644.
94 Habash MB, Beaudette LA, Cassidy MB, Leung KT,
Hoang TA, Vogel HJ, Trevors JT & Lee H (2002)
Characterization of tetrachlorohydroquinone reductive
dehalogenase from Sphingomonas sp. UG30. Biochem
Biophys Res Commun 299, 634–640.
95 Warner JR & Copley SD (2007) Pre-steady-state
kinetic studies of the reductive dehalogenation cata-
lyzed by tetrachlorohydroquinone dehalogenase. Bio-
chemistry 46, 13211–13222.
96 McCarthy DL, Claude AA & Copley SD (1997) In vivo
levels of chlorinated hydroquinones in a pentachloro-
phenol-degrading bacterium. Appl Environ Microbiol
63, 1883–1888.
97 McCarthy DL, Louie DF & Copley SD (1997) Identifi-
cation of a covalent intermediate between glutathione
and cysteine13 formed during catalysis by tetrachloro-
hydroquinone dehalogenase. J Am Chem Soc 119,
11337–11338.
98 Zylstra GJ & Kim E (1997) Aromatic hydrocarbon
degradation by Sphingomonas yanoikuyae B1. J Ind
Microbiol Biotechnol 19, 408–414.
99 Bae M, Sul WJ, Koh SC, Lee JH, Zylstra GJ, Kim
YM & Kim E (2003) Implication of two glutathi-
one S-transferases in the optimal metabolism of m-tol-
uate by Sphingomonas yanoikuyae B1. Antonie Van
Leeuwenhoek 84, 25–30.
100 Prade L, Huber R & Bieseler B (1998) Structures of
herbicides in complex with their detoxifying enzyme
glutathione S-transferase – explanations for the selec-
tivity of the enzyme in plants. Structure 15, 1445–1452.
101 Labrou NE, Karavangeli M, Tsaftaris A & Clonis YD
(2005) Kinetic analysis of maize glutathione S-transfer-
ase I catalysing the detoxification from chloroacetani-
lide herbicides. Planta 222, 91–97.
102 Abel EL, Opp SM, Verlinde CL, Bammler TK &
Eaton DL (2004) Characterization of atrazine biotrans-
formation by human and murine glutathione S-trans-
ferases. Toxicol Sci 80, 230–238.
103 Smith D, Alvey S & Crowley DE (2005) Cooperative
catabolic pathways within an atrazine-degrading
enrichment culture isolated from soil. FEMS Microbiol
Ecol 53, 265–273.
104 de Souza ML, Seffernick J, Martinez B, Sadowsky MJ
& Wackett LP (1998) The atrazine catabolism genes
atzABC are widespread and highly conserved.
J Bacteriol 180, 1951–1954.
105 Laura D, De Socio G, Frassanito R & Rotilio D
(1996) Effects of atrazine on Ochrobactrum anthropi
membrane fatty acids. Appl Environ Microbiol 62,
2644–2646.
106 van Hylckama Vlieg JE, Kingma J, van den Wijngaard
AJ & Janssen DB (1998) A glutathione S-transferase
with activity towards cis-1,2-dichloroepoxyethane is
involved in isoprene utilization by Rhodococcus sp.
strain AD45. Appl Environ Microbiol 64, 2800–2805.
107 van Hylckama Vlieg JE, Kingma J, Kruizinga W &
Janssen DB (1999) Purification of a glutathione S-
transferase and a glutathione conjugate-specific dehy-
drogenase involved in isoprene metabolism in Rhodo-
coccus sp. strain AD45. J Bacteriol 181, 2094–2101.
108 van Hylckama Vlieg JE, Leemhuis H, Spelberg JH &
Janssen DB (2000) Characterization of the gene cluster
involved in isoprene metabolism in Rhodococcus sp.
strain AD45. J Bacteriol 182, 1956–1963.
109 Fall R & Copley SD (2000) Bacterial sources and sinks
of isoprene, a reactive atmospheric hydrocarbon. Envi-
ron Microbiol 2, 123–130.
110 Rui L, Kwon YM, Reardon KF & Wood TK (2004)
Metabolic pathway engineering to enhance aerobic
degradation of chlorinated ethenes and to reduce
their toxicity by cloning a novel glutathione S-transfer-
ase, an evolved toluene o-monooxygenase, and
gamma-glutamylcysteine synthetase. Environ Microbiol
6, 491–500.
111 Masai E, Katayama Y, Kubota S, Kawai S, Yamasaki
M & Morohoshi N (1993) A bacterial enzyme degrad-
ing the model lignin compound b-etherase is a member
of the glutathione S-transferase superfamily. FEBS
Lett 323, 135–140.
112 Masai E, Katayama Y, Nishikawa S & Fukuda M
(1999) Characterization of Sphingomonas paucimobilis
SYK-6 genes involved in degradation of lignin-related
compounds. J Ind Microbiol Biotechnol 23, 364–373.
113 Masai E, Ichimura A, Sato Y, Miyauchi K, Katayama
Y & Fukuda M (2003) Roles of the enantioselective
glutathione S-transferases in cleavage of b-aryl ether.
J Bacteriol 185, 1768–1775.
114 Keck A, Conradt D, Mahler A, Stolz A, Mattes R &
Klein J (2006) Identification and functional analysis of
the genes for naphthalenesulfonate catabolism by
Sphingomonas xenophaga BN6. Microbiology 152,
1929–1940.
115 Kim E, Zylstra GJ, Freeman JP, Heinze TM, Deck J
& Cerniglia CE (1997) Evidence for the role of 2-hy-
droxychromene-2-carboxylate isomerase in the degra-
dation of anthracene by Sphingomonas yanoikuyae B1.
FEMS Microbiol Lett 153, 479–484.
116 Piccolomini R, Allocati N, Cellini L, Faraone A, Di
Cola D, Sacchetta P & Ravagnan G (1987) Preliminary
studies on the effect of glutathione S-transferase from
Providencia stuartii on the antimicrobial activity of dif-
ferent antibiotics. Chemioterapia 6, 324–328.
117 Allocati N, Favaloro B, Masulli M, Tamburro A, Rot-
ilio D & Di Ilio C (2001) In vivo effect of xenobiotic
compounds on the Proteus mirabilis glutathione trans-
ferase B1-1. Chem Biol Int 133, 261–264.
118 Nilsson AI, Berg OG, Aspevall O, Kahlmeter G &
Andersson DI (2003) Biological costs and mechanisms
Bacterial GSTs N. Allocati et al.
74 FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS
of fosfomycin resistance in Escherichia coli.Antimicrob
Agents Chemother 47, 2850–2858.
119 Baum EZ, Montenegro DA, Licata L, Turchi I, Webb
GC, Foleno BD & Bush K (2001) Identification and
characterization of new inhibitors of the Escherichi-
a coli MurA enzyme. Antimicrob Agents Chemother 45,
3182–3188.
120 Mendoza C, Garcia JM, Llaneza J, Mendez FJ, Har-
disson C & Ortiz JM (1980) Plasmid-determined resis-
tance to fosfomycin in Serratia marcescens.Antimicrob
Agents Chemother 18, 215–219.
121 Arca P, Rico M, Brana AF, Villar CJ, Hardisson C &
Suarez JE (1988) Formation of an adduct between fos-
fomycin and glutathione: a new mechanism of antibi-
otic resistance in bacteria. Antimicrob Agents
Chemother 32, 1552–1556.
122 Arca P, Hardisson C & Suarez JE (1990) Purification
of a glutathione S-transferase that mediates fosfomycin
resistance in bacteria. Antimicrob Agents Chemother 34,
844–848.
123 Bernat BA, Laughlin LT & Armstrong RN (1997)
Fosfomycin resistance protein (FosA) is a manganese
metalloglutathione transferase related to glyoxalase I
and the extradiol dioxygenases. Biochemistry 36, 3050–
3055.
124 Rife CL, Pharrism RE, Newcomer ME & Armstrong
RN (2002) Crystal structure of a genomically encoded
fosfomycin resistance protein (FosA) at 1.19 A
˚resolu-
tion by MAD phasing off the L-III edge of Tl
+
.JAm
Chem Soc 124, 11001–11003.
125 Pakhomova S, Rife CL, Armstrong RN & Newcomer
ME (2004) Structure of fosfomycin resistance protein
FosA from transposon Tn2921. Protein Sci 13, 1260–
1265.
126 Beharry Z & Palzkill T (2005) Functional analysis of
active site residues of the fosfomycin resistance enzyme
FosA from Pseudomonas aeruginosa.J Biol Chem 280,
17786–17791.
127 Rigsby RE, Brown DW, Dawson E, Lybrand TP &
Armstrong RN (2007) A model for glutathione binding
and activation in the fosfomycin resistance protein,
FosA. Arch Biochem Biophys 464, 277–283.
128 Etienne J, Gerbaud G, Courvalin P & Fleurette J
(1989) Plasmid-mediated resistance to fosfomycin in
Staphylococcus epidermidis.FEMS Microbiol Lett 52,
133–137.
129 Zilhao R & Courvalin P (1990) Nucleotide sequence of
the fosB gene conferring fosfomycin resistance in
Staphylococcus epidermidis.FEMS Microbiol Lett 56,
267–272.
130 Etienne J, Gerbaud G, Fleurette J & Courvalin P
(1991) Characterization of staphylococcal plasmids
hybridizing with the fosfomycin resistance gene fosB.
FEMS Microbiol Lett 68, 119–122.
131 Cao M, Bernat BA, Wang Z, Armstrong RN & Hel-
mann JD (2001) FosB, a cysteine-dependent fosfomy-
cin resistance protein under the control of r
W
,an
extracytoplasmic-function factor in Bacillus subtilis.
J Bacteriol 183, 2380–2383.
132 Fahey RC, Brown WC, Adams WB & Worsham MB
(1978) Occurrence of glutathione in bacteria. J Bacte-
riol 133, 1126–1129.
133 Fillgrove KL, Pakhomova S, Newcomer ME &
Armstrong RN (2003) Mechanistic diversity of
fosfomycin resistance in pathogenic microorganisms.
J Am Chem Soc 125, 15730–15731.
134 Fillgrove KL, Pakhomova S, Schaab MR, Newcomer
ME & Armstrong RN (2007) Structure and mechanism
of the genomically encoded fosfomycin resistance pro-
tein, FosX, from Listeria monocytogenes.Biochemistry
46, 8110–8120.
135 Garcı
´a P, Arca P & Sua
´rez EJ (1995) Product of fosC,
a gene from Pseudomonas syringae, mediates fosfomy-
cin resistance by using ATP as cosubstrate. Antimicrob
Agents Chemother 39, 1569–1573.
136 Kurtovic S, Shokeer A & Mannervik B (2008) Diverg-
ing catalytic capacities and selectivity profiles with
haloalkane substrates of chimeric alpha class glutathi-
one transferases. Protein Eng Des Sel 21, 329–341.
137 Yan F, Yang WK, Li XY, Lin TT, Lun YN, Lin F,
Lv SW, Yan GL, Liu JQ, Shen JC et al. (2008) A tri-
functional enzyme with glutathione S-transferase, glu-
tathione peroxidase and superoxide dismutase activity.
Biochim Biophys Acta 1780, 869–872.
138 Choi JW, Kim YK, Song SY, Lee IH & Lee WH
(2003) Optical biosensor consisting of glutathione
S-transferase for detection of captan. Biosens
Bioelectron 18, 1461–1466.
139 Di Ilio C, Sacchetta P, Angelucci S, Bucciarelli T, Pen-
nelli A, Mazzetti AP, Lo Bello M & Aceto A (1996)
Interaction of glutathione transferase P1-1 with captan
and captafol. Biochem Pharmacol 52, 43–48.
140 Timmis KN & Pieper DH (1999) Bacteria designed for
bioremediation. Trends Biotechnol 17, 200–204.
141 Watanabe K (2001) Microorganisms relevant to bio-
remediation. Curr Opin Biotechnol 12, 237–241.
142 Bradley PM & Chapelle FH (1998) Effect of contami-
nant concentration on aerobic microbial mineralization
of DCE and VC in stream-bed sediments. Environ Sci
Technol 32, 553–557.
143 Larkin MA, Blackshields G, Brown NP, Chenna R,
McGettigan PA, McWilliam H, Valentin F, Wallace
IM, Wilm A, Lopez R et al. (2007) Clustal W and
Clustal X version 2.0. Bioinformatics 23, 2947–2948.
144 Tamura K, Dudley J, Nei M & Kumar S (2007)
MEGA4: molecular evolutionary genetics analysis
(MEGA) software version 4.0. Mol Biol Evol 24, 1596–
1599.
N. Allocati et al. Bacterial GSTs
FEBS Journal 276 (2009) 58–75 ª2008 The Authors Journal compilation ª2008 FEBS 75
... Any exposure to heavy metals and UV-B radiation causes change in the ratio of oxidized/reduced glutathione in plant and cyanobacterial cells (Fahey, 2013;Foyer & Noctor, 2011). Glutathione transferases (GSTs) that play a crucial role in the detoxification of oxidative stresses in aerobic prokaryotes and eukaryotes constitute a superfamily of proteins (Allocati et al., 2009). GSTs are dimeric proteins in which each subunit is composed of two domains-the N-terminal domain is involved in the binding of GSH, whereas the C-terminal domain interacts with the electrophilic substrate. ...
... Glutathione displaces the halide from alkyl, aryl, and vinyl halides, and this displacement is catalyzed by enzyme Glutathione S-transferases (GSTs). GSTs have been extensively studied in prokaryotes and have been the subject of several reviews (Allocati et al., 2009;Allocati et al., 2012;Vuilleumier & Pagni, 2002). GSTs have been characterized from α-, β-, and γ-proteobacteria (Allocati et al., 2009), cyanobacteria (Wiktelius & Stenberg, 2007), and Rhodococcus AD4 (van Hylckama Vlieg et al., 1999). ...
... GSTs have been extensively studied in prokaryotes and have been the subject of several reviews (Allocati et al., 2009;Allocati et al., 2012;Vuilleumier & Pagni, 2002). GSTs have been characterized from α-, β-, and γ-proteobacteria (Allocati et al., 2009), cyanobacteria (Wiktelius & Stenberg, 2007), and Rhodococcus AD4 (van Hylckama Vlieg et al., 1999). ...
Chapter
Cyanobacteria are the ancient, monophyletic, and oxygen-producing photosynthetic group of bacteria that are widely distributed in all possible habitats. They are the first organisms to encounter oxidative stress in response to which they have evolved enzymatic and nonenzymatic strategies to combat stress. One of the crucial defenses relies on nonenzymatic antioxidants called glutathione. These are cellular thiols formed by cysteine with the help of ATP-requiring enzymes. Glutathione protects cyanobacteria against abiotic stresses like UV-B irradiation, ROS species, and high light intensity. It also aids in tolerance against cold, heat, metal stress, and nutrient deprivation as well as induces the production of various antioxidants. Glutathione poses various roles, majorly virulence and pathogenesis, biofilm formation and disruption, synthesis of DNA, and protection against methylglyoxal and copper homeostasis. Apart from the major emphasis on glutathione synthesis, metabolism, degradation, and roles, the present book chapter encompasses the source and effects of oxidative stress and its cross-talk with iron homeostasis.
... A variety of halogenated xenobiotics (e.g., therapeutic agents and agrochemicals) are bound by the lipophilic P450 active site to be detoxified through a dehalogenation step [83][84][85]. Prokaryotic glutathione transferases are key enzymes in the cellular detoxification of a broad range of harmful xenobiotics, including aliphatic, aromatic, and heterocyclic molecules with halide groups [86]. These enzymes catalyze glutathione conjugation to electrophilic groups (mainly introduced by P450) of a wide range of hydrophobic toxic compounds, thus promoting their excretion from the cell [87]. ...
... These enzymes catalyze glutathione conjugation to electrophilic groups (mainly introduced by P450) of a wide range of hydrophobic toxic compounds, thus promoting their excretion from the cell [87]. Glutathione transferases detoxify several classes of herbicides including triazines, a class of man-made chemicals that includes atrazine, one of the most widely used chlorinated herbicides [86]. ...
Article
Full-text available
Background Macroalgae, especially reds (Rhodophyta Division) and browns (Phaeophyta Division), are known for producing various halogenated compounds. Yet, the reasons underlying their production and the fate of these metabolites remain largely unknown. Some theories suggest their potential antimicrobial activity and involvement in interactions between macroalgae and prokaryotes. However, detailed investigations are currently missing on how the genetic information of prokaryotic communities associated with macroalgae may influence the fate of organohalogenated molecules. Results To address this challenge, we created a specialized dataset containing 161 enzymes, each with a complete enzyme commission number, known to be involved in halogen metabolism. This dataset served as a reference to annotate the corresponding genes encoded in both the metagenomic contigs and 98 metagenome-assembled genomes (MAGs) obtained from the microbiome of 2 red (Sphaerococcus coronopifolius and Asparagopsis taxiformis) and 1 brown (Halopteris scoparia) macroalgae. We detected many dehalogenation-related genes, particularly those with hydrolytic functions, suggesting their potential involvement in the degradation of a wide spectrum of halocarbons and haloaromatic molecules, including anthropogenic compounds. We uncovered an array of degradative gene functions within MAGs, spanning various bacterial orders such as Rhodobacterales, Rhizobiales, Caulobacterales, Geminicoccales, Sphingomonadales, Granulosicoccales, Microtrichales, and Pseudomonadales. Less abundant than degradative functions, we also uncovered genes associated with the biosynthesis of halogenated antimicrobial compounds and metabolites. Conclusion The functional data provided here contribute to understanding the still largely unexplored role of unknown prokaryotes. These findings support the hypothesis that macroalgae function as holobionts, where the metabolism of halogenated compounds might play a role in symbiogenesis and act as a possible defense mechanism against environmental chemical stressors. Furthermore, bacterial groups, previously never connected with organohalogen metabolism, e.g., Caulobacterales, Geminicoccales, Granulosicoccales, and Microtrichales, functionally characterized through MAGs reconstruction, revealed a biotechnologically relevant gene content, useful in synthetic biology, and bioprospecting applications. 7KFSwtia72NY8FA2HCcej3Video Abstract
... The initial flux of components between the cell and environment and the cytosol-periplasmic space is achieved through numerous transporters. The following processes were affected at the early timepoint: biofilm formation (bapA) [61], sulphate uptake and utilization (cysU, nrtCD) [62], xanthan biosynthesis, glutathione metabolism (gst) [63], enterobactin induced iron uptake (befA) [64], and metal, broad substrate efflux and MDR (yadH, emrA, mexF) [6,65,66]. These changes would prepare cells for oxidative stress, possibly xenobiotic resistance, redox homeostasis via cysteine/sulphated compound metabolism and possible efflux of copper or other metal ions via czc elements. ...
Article
Full-text available
Background Copper-induced gene expression in Xanthomonas campestris pv. campestris (Xcc) is typically evaluated using targeted approaches involving qPCR. The global response to copper stress in Xcc and resistance to metal induced damage is not well understood. However, homologs of heavy metal efflux genes from the related Stenotrophomonas genus are found in Xanthomonas which suggests that metal related efflux may also be present. Methods and Results Gene expression in Xcc strain BrA1 exposed to 0.8 mM CuSO4.5H2O for 15 minutes was captured using RNA-seq analysis. Changes in expression was noted for genes related to general stress responses and oxidoreductases, biofilm formation, protein folding chaperones, heat-shock proteins, membrane lipid profile, multiple drug and efflux (MDR) transporters, and DNA repair were documented. At this timepoint only the cohL (copper homeostasis/tolerance) gene was upregulated as well as a chromosomal czcCBA efflux operon. An additional screen up to 4 hrs using qPCR was conducted using a wider range of heavy metals. Target genes included a cop-containing heavy metal resistance island and putative metal efflux genes. Several efflux pumps, including a copper resistance associated homolog from S. maltophilia, were upregulated under toxic copper stress. However, these pumps were also upregulated in response to other toxic heavy metals. Additionally, the temporal expression of the coh and cop operons was also observed, demonstrating co-expression of tolerance responses and later activation of part of the cop operon. Conclusions Overall, initial transcriptional responses focused on combating oxidative stress, mitigating protein damage and potentially increasing resistance to heavy metals and other biocides. A putative copper responsive efflux gene and others which might play a role in broader heavy metal resistance were also identified. Furthermore, the expression patterns of the cop operon in conjunction with other copper responsive genes allowed for a better understanding of the fate of copper ions in Xanthomonas. This work provides useful evidence for further evaluating MDR and other efflux pumps in metal-specific homeostasis and tolerance phenotypes in the Xanthomonas genus. Furthermore, non-canonical copper tolerance and resistance efflux pumps were potentially identified. These findings have implications for interpreting MIC differences among strains with homologous copLAB resistance genes, understanding survival under copper stress, and resistance in disease management.
... The expression of catA in DBC for all treatments including RE and/or pyrene was possibly due to the absence of transcriptional regulatory gene in the genome of Pseudomonas fragi DBC, as reported for Pseudomonas stutzeri ZWLR2-1, which degrades 2-chloro, nitrobenzene constitutively, and had no NagR-like transcriptional regulator associated with dioxygenase genes 20,21 . The involvement of bacterial GST during the degradation and detoxification of various xenobiotic compounds has been reported 3,22 . Currently, yfcG is one of the two glutathione-S-transferase genes which has been considered as a new class of GST enzyme family (Nu) 19 , that do not act on reduced glutathione, rather more reactive to oxidized glutathione, and possess a unique disulfide bond reductase activity 23 . ...
Article
Full-text available
Pyrene is an extremely hazardous, carcinogenic polycyclic aromatic hydrocarbon (PAH). The plant–microbe interaction between Pseudomonas fragi DBC and Jatropha curcas was employed for biodegradation of pyrene and their transcriptional responses were compared. The genome of P. fragi DBC had genes for PAH degrading enzymes i.e. dioxygenases and dehydrogenases, along with root colonization (trpD, trpG, trpE and trpF), chemotaxis (flhF and flgD), stress adaptation (gshA, nuoHBEKNMG), and detoxification (algU and yfc). The transcriptional expression of catA and yfc that respectively code for catabolic enzyme (catechol-1, 2-dioxygnase) and glutathione-s-transferase for detoxification functions were quantitatively measured by qPCR. The catA was expressed in presence of artificial root exudate with or without pyrene, and glucose confirming the non-selective approach of bacteria, as desired. Pyrene induced 100-fold increase of yfc expression than catA, while there was no expression of yfc in absence of pyrene. The transcriptome of plant roots, in presence of pyrene, with or without P. fragi DBC inoculation was analysed. The P. fragi DBC could upregulate the genes for plant growth, induced the systemic acquired resistance and also ameliorated the stress response in Jatropha roots.
... GSTs are classified based on their cellular location: mitochondrial GSTs, microsomal GSTs, and cytosolic GSTs (Enayati et al., 2005;Hayes et al., 2005;Allocati et al., 2009;Oakley, 2011). The largest family, cytosolic GSTs, is divided into six subclasses: delta, epsilon, omega, sigma, theta, and zeta (Enayati et al., 2005;Frova, 2006). ...
... Although the role of glutathione in the bacterial degradation of microcystins has not yet been confirmed, it is known that the first step in the biodegradation of MC-LR by higher plants Ceratophyllum demersum, invertebrates Daphnia magma, the shellfish Dreissena polymorpha, and in rats is the conjugation of MC with GSH followed by the formation of the cysteine conjugate MC-LR-Cys (Schmidt et al., 2014;Krausfeldt et al., 2019). Given the role of GSH in the bacterial degradation of many xenobiotics (Allocati et al., 2009), it is assumed that glutathione is also involved in the microbial metabolism of microcystin (Mou et al., 2013;Krausfeldt et al., 2019). Our results confirm the previously stated hypothesis about the wide distribution in natural objects of microorganisms capable of decomposing microcystins according to the mlr mechanism, as well as through other biochemical pathways, including the formation of conjugates of microcystin with glutathione and cysteine (Krausfeldt et al., 2019;Salter et al., 2021). ...
... GSTs mainly catalyzes the covalent conjugation of glutathione with toxic electrophilic and hydrophobic substrates, so as to participate in a variety of detoxification metabolic processes in insects [18]. Additionally, 30 GSTs were identified in A. suturalis (Fig. 4a). ...
Article
Full-text available
Background Adelphocoris suturalis (Hemiptera: Miridae) is a notorious agricultural pest, which causes serious economic losses to a diverse range of agricultural crops around the world. The poor understanding of its genomic characteristics has seriously hindered the establishment of sustainable and environment-friendly agricultural pest management through biotechnology and biological insecticides. Results Here, we report a chromosome-level assembled genome of A. suturalis by integrating Illumina short reads, PacBio, 10x Chromium, and Hi-C mapping technologies. The resulting 1.29 Gb assembly contains twelve chromosomal pseudomolecules with an N50 of 1.4 and 120.6 Mb for the contigs and scaffolds, respectively, and carries 20,010 protein-coding genes. The considerable size of the A. suturalis genome is predominantly attributed to a high amount of retrotransposons, especially long interspersed nuclear elements (LINEs). Transcriptomic and phylogenetic analyses suggest that A. suturalis-specific candidate effectors, and expansion and expression of gene families associated with omnivory, insecticide resistance and reproductive characteristics, such as digestion, detoxification, chemosensory receptors and long-distance migration likely contribute to its strong environmental adaptability and ability to damage crops. Additionally, 19 highly credible effector candidates were identified and transiently overexpressed in Nicotiana benthamiana for functional assays and potential targeting for insect resistance genetic engineering. Conclusions The high-quality genome of A. suturalis provides an important genomic landscape for further investigations into the mechanisms of omnivory, insecticide resistance and survival adaptation, and for the development of integrated management strategies.
... GSTs are classified based on their cellular location: mitochondrial GSTs, microsomal GSTs, and cytosolic GSTs [73][74][75][76]. The largest family, the cytosolic GSTs, is divided into six subclasses: delta, epsilon, omega, sigma, theta, and zeta [75,77]. ...
Preprint
Full-text available
The European honey bee, Apis mellifera , serves as the principle managed pollinator species globally. In recent decades, honey bee populations have been facing serious health threats from combined biotic and abiotic stressors, including diseases, limited nutrition, and agrochemical exposure. Understanding the molecular mechanisms underlying xenobiotic adaptation of A. mellifera is critical, considering its extensive exposure to phytochemicals and agrochemicals present in flowers, propolis, hives, and the environment. In this study, we conducted a comprehensive structural and functional characterization of AmGSTD1, a delta class glutathione S-transferase (GST) enzyme, to unravel its roles in agrochemical detoxification and antioxidative stress responses. Significantly, we determined the 3D structure of a honey bee GST using protein crystallography for the first time, providing new insights into its molecular structure. Our investigations revealed that AmGSTD1 efficiently metabolizes model substrates, including 1-chloro-2,4-dinitrobenzene (CDNB), p-nitrophenyl acetate (PNA), phenylethyl isothiocyanate (PEITC), propyl isothiocyanate (PITC), and the oxidation byproduct 4-hydroxynonenal (4-HNE). Moreover, we discovered that AmGSTD1 exhibits binding affinity with the fluorophore 8-Anilinonaphthalene-1-sulfonic acid (ANS), which can be inhibited with various herbicides, fungicides, insecticides, and their metabolites. These findings highlight the potential contribution of AmGSTD1 in safeguarding honey bee health against various agrochemicals and their metabolites, while also mitigating oxidative stress resulting from exposure to these substances.
Article
Full-text available
Actinobacteria are prevalent in the rhizosphere and phyllosphere of diverse plant species where they help to enhance tolerance of plants against biotic and abiotic stresses. Here, we show that the plant hormones jasmonic acid (JA) and methyl jasmonate (MeJA) affect the growth, development, and specialized metabolism of Streptomyces . Exposure of Streptomyces coelicolor to JA or MeJA led to enhanced production of the polyketide antibiotic actinorhodin. JA also exhibited toxicity toward Streptomyces and Streptacidiphilus at higher concentrations, whereby streptomycetes were more tolerant to JA than members of the genus Streptacidiphilus . Tolerance to JA could be linked to its conjugation by the bacteria with glutamine. Additionally, JA conjugates with valine, tyrosine, phenylalanine, and leucine/isoleucine were identified. In contrast to JA, synthetic JA conjugates failed to activate antibiotic production and showed significantly reduced toxicity. Thus, our findings provide insights into a previously unknown defense mechanism deployed by Streptomycetaceae to a plant hormone. The underlying mechanism encompasses the attachment of amino acids to JA, which in turn safeguards the bacteria against the harmful impacts of the plant hormone. This study adds to the growing body of evidence that plant hormones can have a significant impact on members of the plant microbiome by affecting their growth, development, and secondary metabolism. IMPORTANCE Microorganisms that live on or inside plants can influence plant growth and health. Among the plant-associated bacteria, streptomycetes play an important role in defense against plant diseases, but the underlying mechanisms are not well understood. Here, we demonstrate that the plant hormones jasmonic acid (JA) and methyl jasmonate directly affect the life cycle of streptomycetes by modulating antibiotic synthesis and promoting faster development. Moreover, the plant hormones specifically stimulate the synthesis of the polyketide antibiotic actinorhodin in Streptomyces coelicolor . JA is then modified in the cell by amino acid conjugation, thereby quenching toxicity. Collectively, these results provide new insight into the impact of a key plant hormone on diverse phenotypic responses of streptomycetes.
Article
Full-text available
V4 is a Gram-negative, plant growth promoting endophytic bacterium that promotes the growth of tea plants. The appearance of V4 is rod shaped, with average dimensions of 1.34−1.5 × 0.32−0.39 μm and flagellum at both ends. The complete genome contains one circular chromosome and two plasmids. It is 4,697,109 bp in size, and contains 4,189 protein-coding genes, four gene islands and two prophages. Taxonomic classification suggested that V4 was a strain of Erwinia aphidicola. It was possible to find genes involved in plant growth promotion traits present in the genome of V4. Meanwhile, V4 was consistent with plant growth-promoting endophytic bacteria containing key synthetic genes associated with IAA synthesis, and P-solubilization, siderophores. V4 has siderophore biosynthesis genes compared with plant pathogenic bacteria showing stronger survival ability and the ability to interaction with the host plant. In addition, V4 endophytic bacteria possess a higher copy number of genes for flagellar assembly, bacterial chemotaxis and P-pilus assembly indicating stronger colonization and communication ability with host plants compared with five other bacteria in comparative genomic analysis. Analysis of the V4 endophytic bacterium complete genome sequence provides novel insights into the endophytic bacteria-host plant relationship, and suggests many candidate genes for post-genomic experiments.
Article
Full-text available
The glutathione transferases (GSTs; also known as glutathione S-transferases) are major phase II detoxification enzymes found mainly in the cytosol. In addition to their role in catalysing the conjugation of electrophilic substrates to glutathione (GSH), these enzymes also carry out a range of other functions. They have peroxidase and isomerase activities, they can inhibit the Jun N-terminal kinase (thus protecting cells against H(2)O(2)-induced cell death), and they are able to bind non-catalytically a wide range of endogenous and exogenous ligands. Cytosolic GSTs of mammals have been particularly well characterized, and were originally classified into Alpha, Mu, Pi and Theta classes on the basis of a combination of criteria such as substrate/inhibitor specificity, primary and tertiary structure similarities and immunological identity. Non-mammalian GSTs have been much less well characterized, but have provided a disproportionately large number of three-dimensional structures, thus extending our structure-function knowledge of the superfamily as a whole. Moreover, several novel classes identified in non-mammalian species have been subsequently identified in mammals, sometimes carrying out functions not previously associated with GSTs. These studies have revealed that the GSTs comprise a widespread and highly versatile superfamily which show similarities to non-GST stress-related proteins. Independent classification systems have arisen for groups of organisms such as plants and insects. This review surveys the classification of GSTs in non-mammalian sources, such as bacteria, fungi, plants, insects and helminths, and attempts to relate them to the more mainstream classification system for mammalian enzymes. The implications of this classification with regard to the evolution of GSTs are discussed.
Article
Full-text available
Pseudomonas strain ADP metabolizes the herbicide atrazine via three enzymatic steps, encoded by the genes atzABC, to yield cyanuric acid, a nitrogen source for many bacteria. Here, we show that five geographically distinct atrazine-degrading bacteria contain genes homologous to atzA, -B, and -C. The sequence identities of the atz genes from different atrazine-degrading bacteria were greater than 99% in all pairwise comparisons. This differs from bacterial genes involved in the catabolism of other chlorinated compounds, for which the average sequence identity in pairwise comparisons of the known members of a class ranged from 25 to 56%. Our results indicate that globally distributed atrazine-catabolic genes are highly conserved in diverse genera of bacteria.
Article
Glutathione and soluble thiol content were examined in a broad spectrum of bacteria. Significant soluble thiol was present in all cases. The thiol compound was glutathione in most of the gram-negative bacteria but not in most of the gram-positive bacteria studied. Glutathione was absent in four anerobes and one microaerophile but was present in a blue-green bacterium. The glutathione content of Escherichia coli increased significantly during transition from exponential to stationary phase.
Article
The membrane associated proteins in eicosanoid and glutathione metabolism (MAPEG) superfamily includes structurally related membrane proteins with diverse functions of widespread origin. A total of 136 proteins belonging to the MAPEG superfamily were found in database and genome screenings. The members were found in prokaryotes and eukaryotes, but not in any archaeal organism. Multiple sequence alignments and calculations of evolutionary trees revealed a clear subdivision of the eukaryotic MAPEG members, corresponding to the six families of microsomal glutathione transferases (MGST) 1, 2 and 3, leukotriene C4 synthase (LTC4), 5-lipoxygenase activating protein (FLAP), and prostaglandin E synthase. Prokaryotes contain at least two distinct potential ancestral subfamilies, of which one is unique, whereas the other most closely resembles enzymes that belong to the MGST2/FLAP/LTC4 synthase families. The insect members are most similar to MGST1/prostaglandin E synthase. With the new data available, we observe that fish enzymes are present in all six families, showing an early origin for MAPEG family differentiation. Thus, the evolutionary origins and relationships of the MAPEG superfamily can be defined, including distinct sequence patterns characteristic for each of the subfamilies. We have further investigated and functionally characterized representative gene products from Escherichia coli, Synechocystis sp., Arabidopsis thaliana and Drosophila melanogaster, and the fish liver enzyme, purified from pike (Esox lucius). Protein overexpression and enzyme activity analysis demonstrated that all proteins catalyzed the conjugation of 1-chloro-2,4-dinitrobenzene with reduced glutathione. The E. coli protein displayed glutathione transferase activity of 0.11 micromol.min(-1).mg(-1) in the membrane fraction from bacteria overexpressing the protein. Partial purification of the Synechocystis sp. protein yielded an enzyme of the expected molecular mass and an N-terminal amino acid sequence that was at least 50% pure, with a specific activity towards 1-chloro-2,4-dinitrobenzene of 11 micromol.min(-1).mg(-1). Yeast microsomes expressing the Arabidopsis enzyme showed an activity of 0.02 micromol.min(-1).mg(-1), whereas the Drosophila enzyme expressed in E. coli was highly active at 3.6 micromol.min(-1).mg(-1). The purified pike enzyme is the most active MGST described so far with a specific activity of 285 micromol.min(-1).mg(-1). Drosophila and pike enzymes also displayed glutathione peroxidase activity towards cumene hydroperoxide (0.4 and 2.2 micromol.min(-1).mg(-1), respectively). Glutathione transferase activity can thus be regarded as a common denominator for a majority of MAPEG members throughout the kingdoms of life whereas glutathione peroxidase activity occurs in representatives from the MGST1, 2 and 3 and PGES subfamilies.
Article
The residue Cys10 in E. coli glutathione transferase (GST) is apparently equivalent in primary structure to the catalytic Ser residue of the class theta GST and is located near the thiol group of GSH bound to the enzyme. The mutation of Cys10 to Ala, however, increased the specific activity toward 1-chloro-2,4-dinitrobenzene at pH 6.5. This mutation increased both the kcat and Km values for GSH and affected the pH-activity profile the enzyme. The side chain of Cys10 is thought to be important for construction of the GSH-binding site and partly for lowering the pKa of the GSH thiol, but not to be essential for the catalytic activity.
Article
Discharge of DCE and VC to an aerobic surface water system simultaneously represents a significant environmental concern and, potentially, a non-engineered opportunity for efficient contaminant bioremediation. The potential for bioremediation, however, depends on the ability of the stream-bed microbial community to efficiently and completely degrade DCE and VC over a range of contaminant concentrations. The purposes of the studies reported here were to assess the potential for aerobic DCE and VC mineralization by stream-bed microorganisms and to evaluate the effects of DCE and VC concentrations on the apparent rates of aerobic mineralization. Bed-sediment microorganisms indigenous to a creek, where DCE-contaminated groundwater continuously discharges, demonstrated rapid mineralization of DCE and VC under aerobic conditions. Over 8 days, the recovery of [1,2-14C]DCE radioactivity as 14CO2 ranged from 17% to 100%, and the recovery of [1,2-14C]VC radioactivity as 14CO2 ranged from 45% to 100%. Rates of DCE and VC mineralization increased significantly with increasing contaminant concentration, and the response of apparent mineralization rates to changes in DCE and VC concentrations was adequately described by Michaelis−Menten kinetics.
Article
A novel superfamily designated MAPEG (Membrane Associated Proteins in Eicosanoid and Glutathione metabolism), including members of widespread origin with diversified biological functions is defined according to enzymatic activities, sequence motifs, and structural properties. Two of the members are crucial for leukotriene biosynthesis, and three are cytoprotective exhibiting glutathione S-transferase and peroxidase activities. Expression of the most recently recognized member is strongly induced by p53, and may therefore play a role in apoptosis or cancer development. In spite of the different biological functions, all six proteins demonstrate common structural characteristics typical of membrane proteins. In addition, homologues are identified in plants, fungi, and bacteria, demonstrating this superfamily to be generally occurring.