ArticlePDF AvailableLiterature Review

Genetics and Biochemistry of Phenol Degradation by Pseudomonas sp. CF600

Authors:

Abstract and Figures

Pseudomonas sp. strain CF600 is an efficient degrader of phenol and methylsubstituted phenols. These compounds are degraded by the set of enzymes encoded by the plasmid located dmpoperon. The sequences of all the fifteen structural genes required to encode the nine enzymes of the catabolic pathway have been determined and the corresponding proteins have been purified. In this review the interplay between the genetic analysis and biochemical characterisation of the catabolic pathway is emphasised. The first step in the pathway, the conversion of phenol to catechol, is catalysed by a novel multicomponent phenol hydroxylase. Here we summarise similarities of this enzyme with other multicomponent oxygenases, particularly methane monooxygenase (EC 1.14.13.25). The other enzymes encoded by the operon are those of the well-known meta-cleavage pathway for catechol, and include the recently discovered meta-pathway enzyme aldehyde dehydrogenase (acylating) (EC 1.2.1.10). The known properties of these meta-pathway enzymes, and isofunctional enzymes from other aromatic degraders, are summarised. Analysis of the sequences of the pathway proteins, many of which are unique to the meta-pathway, suggests new approaches to the study of these generally little-characterised enzymes. Furthermore, biochemical studies of some of these enzymes suggest that physical associations between meta-pathway enzymes play an important role. In addition to the pathway enzymes, the specific regulator of phenol catabolism, DmpR, and its relationship to the XylR regulator of toluene and xylene catabolism is discussed.
Content may be subject to copyright.
Biodegradation
5: 219-, 1994. 219
(g) 1994
Kluwer Academic Publishers. Printed in the Netherlands.
Genetics and biochemistry of phenol degradation by
Pseudomonas
sp. CF600
Justin Powlowski 1 & Victoria Shingler 2
1 Department of Chemistry and Biochemistry, Concordia University, 1455 de Maisonneuve Boulevard West,
Montreal, Quebec, Canada H3G 1MS; 2 Department of Cell and Molecular Biology, Umed University, S-901 87
Umed, Sweden
Received 20 September 1993; accepted 31 January t994
Key words:
Phenols, oxygenases,
meta-cleavage
pathway, dehydrogenases,
Pseudomonas
Abstract
Pseudomonas
sp. strain CF600 is an efficient degrader of phenol and methylsubstituted phenols. These compounds
are degraded by the set of enzymes encoded by the ptasmid located
dmpoperon.
The sequences of all the fifteen
structural genes required to encode the nine enzymes of the catabolic pathway have been determined and the
corresponding proteins have been purified. In this review the interplay between the genetic analysis and biochemical
characterisation of the catabolic pathway is emphasised. The first step in the pathway, the conversion of phenol
to catechol, is catalysed by a novel multicomponent phenol hydroxylase. Here we summarise similarities of
this enzyme with other nmlticomponent oxygenases, particularly methane monooxygenase (EC 1.14.13.25). The
other enzymes encoded by the operon are those of the well-known
meta-cleavage
pathway for catechol,
and
include the recently discovered
meta-pathway
enzyme aldehyde dehydrogenase (acylating) (EC t.2.1.I0). The
known properties of these
meta-pathway
enzymes, and isofunctional enzymes from other aromatic degraders, are
summarised. Analysis of the sequences of the pathway proteins, many of which are unique to the
meta-pathway,
suggests new approaches to the study of these generally little-characterised enzymes. Furthermore, biochemical
studies of some of these enzymes suggest that physical associations between
recta-pathway
enzymes play an
important role. In addition to the pathway enzymes, the specific regulator of phenol catabolism, DmpR, and its
relationship to the XylR regulator of toluene and xylene catabolism is discussed.
Introduction
Utilisation of phenol as a source of microbial nourish-
ment has perhaps been overshadowed by awareness of
its generally unhealthy effects on humans. Although
the use of phenol as an antiseptic for treating wounds
and wound dressings was of unquestionable value in
helping to establish the importance of sterile tech-
niques in medicine, and probably helped save the lives
of many patients, the toxicity of the compound was
evident from, amongst other things, its contemporane-
ous use as a popular orally-administered suicidal agent
(Lister 1867; McGrew 1985). While the use of phenol
as an antiseptic dwindled by the turn of the century,
its application as a general disinfectant continued, as
did development of new uses for phenol and its deriva-
tives. Today approximately 1700 million kilograms of
phenol are synthesised annually in the United States
alone (Anonymous 1993), and countless quantities of
phenol and its derivatives are produced as industrial
by-products.
Considering the large scale on which this highly
toxic compound and related chemicals are produced,
its degradability is of obvious interest. Persistence of
phenol itself in the environment is not a particular prob-
lem, with both physico-chemical and biological agents
contributing to its degradation (Howard et al. 1991).
However, as with any metabolic process, biodegrad-
ability can be affected by many variables, including,
for example, the presence of alternative carbon sources
or substituents on the phenol ring. While such limita-
tions may be overcome by genetic manipulation of
the catabolic pathway involved, dissimilatory mecha-
nisms and their regulation must first be characterized
220
at the molecular level in individual microbial species.
Since phenol is a relatively simple, unadorned, aromat-
ic compound for which much information on biodegra-
dation is already available, phenol-degrading bacteria
provide good model systems for these kinds of stud-
ies. Microbial phenol degradation is unlikely to be a
recently-acquired trait. While the use of phenol on a
large scale by humans is relatively recent, microbes
have long been exposed to biologically generated phe-
nols and methyl-substituted phenols produced, for
example, by enterobacteria from tyrosine (Kumagai
et al. 1970; Spoelstra 1977) and by the break-down
of plant materials. The isolation of phenol-degrading
bacteria was achieved as early as 1932 (Happold &
Key 1932), and numerous others have been unearthed
since then. While many of the most-studied phenol
degraders are pseudomonads, other bacterial genera,
such as Bacillus, as well as yeast, are also represent-
ed (Gurujeyalakshmi & Oriel 1989; Neujahr & Gaal
1973).
Work during the nineteen sixties and seventies with
phenol-degrading Pseudomona putida U was respon-
sible for much of the progress in understanding the
biochemistry of the meta-cleavage pathway (reviewed
in Bayly & Barbour 1984). In this pathway, ring cleav-
age adjacent to two aromatic hydroxyls (Fig. 1) is
the distinguishing feature, with the subsequent series
of reactions dictated by the nature of the ring-fission
product. This contrasts with the ortho-cleavage path-
way, used by some phenol degraders, in which cate-
chol (1,2-dihydroxybenzene) produced from phenol is
cleaved between the catechol hydroxyls, with an entire-
ly different set of enzymes required for metabolism
of the distinctive ring-cleavage product (reviewed by
Ornston & Yeh 1982). One consequence of the enzy-
mologies of the two pathways is that while the ortho-
cleavage pathways of pseudomonads cannot tolerate
methyl-substituents, all three mono-methyl substitut-
ed phenols can be degraded via the meta-cleavage
pathway. Ortho-cleavage degraders can, however, cir-
cumvent this problem by first oxidizing the methyl
group to a carboxyl, and then using a separate set of
enzymes for the resulting hydroxybenzoic acid (see
Dagley 1986).
The main emphasis in this review is work done
on the dmp system which encodes the phenol degra-
dation pathway of Pseudomonas sp. strain CF600.
In addition to phenol, this organism metabolizes all
monomethylphenols, as well as 3,4-dimethylphenol,
via the dmp-encoded hydroxylation and meta-cleavage
pathway enzymes. This review highlights the interplay
between the analysis of the genes of this and other
pathways, and the biochemical studies of the enzymes
involved.
Genetic location and organisation of the dmp
system
Like many other pathways for the catabolism of aro-
matic compounds, the phenol/dimethylphenol (dmp)
pathway of Pseudomonas sp. strain CF600 is plasmid
encoded. The plasmid, designated pVI150, belongs to
the P-2 incompatibility group (Shingler et al. t989).
Like two other IncP-2 catabolic plasmids, OCT and
CAM (Fennewald et al. 1978; Rheinwald et al. t973),
pVI150 is very large, and exceeds 200 kb in size. Due to
the large size of pVI 150 and attendant difficulties in its
purification, initial isolation of the pathway-encoding
DNA necessitated utilization of methods more often
associated with isolation of chromosomally-encoded
genes (Shingler et al. 1989; Bartilson et al. 1990).
The results of genetic dissection of the 15 kb DNA
region encoding the dmp system, and the locations
of the sixteen genes involved, are summarised in Fig.
1. This figure also illustrates the biochemical route,
via hydroxylation and a meta-cleavage pathway, for
the catabolism of the pathway substrates. The fifteen
genes encoding the enzymes of the pathway (Table
1) are clustered in the dmp operon. Our attention ini-
tially focused on the DNA encoding the first enzyme
of the pathway, phenol hydroxylase, that catalyses
the conversion of phenol to catechol. DNA encod-
ing this enzyme activity was distinguished from that
encoding the remaining pathway enzymes by express-
ing different parts of the dmp operon in a strain of
Pseudomonas that possesses an ortho-cleavage path-
way capable of further catabolism of catechol, but
not methylcatechols. Subsequent testing of the growth
range of the resulting strains functionally defined the
DNA tbr the phenol hydroxylase (growth on phenol
only) and the entire pathway (growth on phenol and
methylphenols).
A combination of nucleotide sequencing and
polypeptide analysis identified six genes, dmp-
KLMNOP, and their corresponding products, P0, P1,
P2, P3, 174 and P5, within the phenol hydroxylase-
coding region. Deletions within each gene in turn, and
expression of each individual open reading frame, cou-
pled with polypeptide analysis, confirmed the assign-
ment of the polypeptide products to individual genes
(Nordlund et al. 1990a). As discussed below, the find-
N
I
0
4
Bg P B Bg P P N P E N
I I. I I II. I
I. I.
I
i 2 ~ i ~ 6 ~ h 9 10 fl 12 is f,
| l I u II I I II I
R K L M N 0 P Q B C D E F (3 H
[ Phenol Hydroxylase
II
M~m -Pathway Enzymes
221
P
15 kb
,Q
I
I
dmp(K)LMNOP
OH
©
dmp(Q)B dmpD dmpE dmpG drapF
NADH, H +. q ~H /CHO H I"120 ~H 2 C~OH HzO CH 3 COOH (~,-I 3 NAD +. CoASH Ve~I3
k >
.o o --x--"
,usa2 - ,L
0 ~ NADH, H 0 ~ SCoA
, ~ "OH HCOOH v
"0
+ o
N~r~ n" ~ ampc
o ~ "COOH
~/cOOHcooH dmpI CO0~ooH
OH ~
~(0
Fig. 1.
Schematic representation of the phenol/dimethylphenol pathway encoded by a 15 kb region of the pVI150 catabolic plasmid of
Pseudomonas
sp. strain CF600. The arrows indicate the divergent transcription of the
dmpR
regulatory gene and the
dmpKLMNOPQBCDE-
FGHI-operon.
Enzyme functions are as shown in Table 1. Both the hydrolytic
(dmpD-encoded)
and the 4-oxalocrotonate
(DmpCIH-encoded)
branches of the
meta-cleavage
pathway are shown. Restriction enzyme recognition site are:
B, BamHI; Bg, BgllI; E, EcoRI; N, NotI; P,
PvulI.
Table 1.
Summary of the genes and gene products of the
dmp
operon.
Gene Amino acid Molecular mass (kDa) Product Function Reference
residues predicted/estimated
dmpK
92 10.6/12.5 P0
dmpL
331 38.2/34.0 P1
dmpM
90 10.5/10.0 P2
dmpN
517 60.5/58.0 P3
dmpO
119 13.2/13.0 P4
dmpP
353 38.5/39.0 P5
dmpQ
112 12.2/12.0 DmpQ
dmpB
307 35.2/32.0 C230
dmpC
486 51.7/50.0 HMSD
dmpD
283 31.0/30.0 HMSH
dmpE
261 27.9/28.0 OEH
dmpF
312 32.7/35.0 ADA
dmpG
345 37.5/39.0 HOA
dmpH
264 28.4/28.5 4OD
dmpl
63 7.1/6.7 401
Unknown
Phenol hydroxylase component
Phenol hydroxylase component
Phenol hydroxylase component
Phenol hydroxylase component
Phenol hydroxylase component
Ferredoxin-like protein
Catechol 2,3-dioxygenase, EC 1.13.11.2
2-Hydroxymueonic semialdehyde,
dehydrogenase, EC 1.2.1 .-
2-Hydroxymuconie semialdehyde hydrolase
2-Oxopent-4-dienoate hydratase, EC 4.2.1.80
Aldehyde dehydrogenase (acylating), EC 1.2.1.10
4-Hydroxy-2-oxovalerate aldolase, EC 4.1.3.-
4-Oxalocrotonate decarboxylase, EC 4.1.1 .-
4-Oxalocrotonate isomemse, EC 5.3.2.-
Nordlund et al. 1990
Nordlund et al. 1990
Nordlundet al. 1990
Nordlund et al. 1990
Nordlund et al. 1990
Nordlnnd et al. 1990
Shingler et al. 1992
Bartilson & Shingler 1989
Nordlund & Shingler 1990
Nordlund & Shingler 1990
Shingler et al. 1992
Shingler et al. 1992
Shingler et al. 1992
Shingler et al. 1992
Shingler et al. 1992
ing that the phenol hydroxylase encoded by pV1150 is
multicomponent was initially surprising and prompted
further studies of this unusual enzyme.
The remaining nine genes of the operon,
dmpQBCDEFGHI,
encode the
meta-cleavage
pathway
enzymes for conversion of catechol to pyruvate and
acetyl-CoA (Table 1, Fig. 2). Assignment of a func-
tion to each of these genes was achieved using a com-
bination of expression of open reading frames, both
alone and in combination, nucleotide sequencing, and
222
sion of growth substrates to catechols appear to be clus-
tered in a single operon together with the meta-pathway
genes. However, the bph and tod operons only code for
an abridged version of the meta-cleavage pathway and
do not include genes for the 4-oxalocrotonate branch
(see Table 2), which is not necessary for the metabo-
lites of these two pathways. Comparisons between all
of these pathways and the dmp-encoded meta-cleavage
pathway will be included. However, before discussion
of the meta-cleavage pathway, the dmp-encoded phe-
nol hydroxylase that feeds it will be considered.
Initial hydroxylation of phenol
Fig.2. Proteinsencodedbythedmpoperon.2to 5 ~gofprotein were
analysed on a gradient ( 10 to 20%) SDS polyacrylamide gel. Lane 1,
DmpK (P0); 2, DmpLNO (P1 -P3-P4); 3, DmpM (P2); 4, DmpP (P5);
5, DmpQ; 6, DmpB (C230); 7, DmpC (HMSD); 8, DmpD (HMSH);
9, DmpEH (OEH/4OD); 10, DmpFG (ADA/HOA); I 1, Dmpl (4OI):
Molecular mass standards are given in kDa on the right-hand side.
All polypeptides were purified from phenol-grown
Pseudomonas
sp. strain CF600 except DmpK (lane 1), DmpQ (lane5) and DmpC
(lane7), which were purified from separate E. coli strains express-
hag the corresponding genes. Enzyme abbreviations and estimated
molecular masses are given in Table 1.
biochemical techniques (Bartilson & Shingler 1989;
Nordlund & Shingler 1990; Shingler et al. 1992).
Since many pathways for the catabolism of aro-
matic compounds include meta-cleavage pathway
enzymes, the dmp genes and proteins have counterparts
in other systems. The most extensively studied of these
are those from pseudomonads harbouring the pWW0-
encoded toluene/xylene xyl pathway, the NAH7- and
pWW60-1-eneoded naphthalene/salicylate nal (sal)
pathways, and the phenol degradation pathway of P.
putida U. All these systems have both branches of
the meta-cleavage pathway as depicted in Fig. 1 (see
Assinder & Williams 1990; Williams & Sayer 1994 for
recent reviews). Using gene-specific probing, this has
also been shown to be true for the meta-pathway genes
of the Pseudomonas sp. strain IC biphenyl degradation
pathway (Carrington et al. 1994). Recently, nucleotide
sequences have been completed for two chromosoma-
lly encoded meta-cleavage pathways involved in the
catabolism of (chloro-)biphenyl and toluene respec-
tively: the Pseudomonas sp. strain KKS 102 bph system
(Kikuchi et al. I993), and the P. putida F1 rod systern
(Lau et al. 1993). In these two systems, like that of
Pseudomonas sp. strain CF600, the genes for conver-
Although some of the first well-studied bacterial
aromatic degraders included phenol-degrading pseu-
domonads that hydroxylate the ring as aprelude to ring-
cleavage, little biochemical information about phe-
nol hydroxylases from these bacteria was published.
Work that required meagurements of phenol hydrox-
ylase activity relied on oxygen-uptake experiments
using whole cells. In fact, the first reported purification
of a hydroxylase specific to a microbial phenol degra-
dation pathway was for phenol hydroxylase from the
yeast Trichosporon cutaneum (Neujahr & Gaal 1973).
This enzyme is a flavoprotein hydroxylase with prop-
erties similar to those of many bacterial flavoprotein
hydroxylases used for catabolism of such compounds
asp-hydroxybenzoate and melilotate (reviewed by Bal-
lou 1982) (Fig. 3).
All of these flavoprotein hydroxylases mono-
oxygenate the aromatic ring at a position ortho (or
in some cases para) to a pre-existing hydroxyl group.
The role of the flavin in these enzymes is the activa-
tion of molecular oxygen, which is normally unreactive
with organic compounds, Activation initially involves
reaction of 2-dectron reduced flavin with 02to form a
flavin 4a-hydroperoxide, which is an electrophilic, rel-
atively weak, oxygenating agent (Entsch et al. 1976).
Oxygenation by this enzyme-bound species is aided
by delocalization of electrons into the ring from the
hydroxyl group of the substrate. Conversion of pheno-
lies to dihydroxylated ring-fission substrates by pro-
teins containing flavin as the sole prosthetic group is
thus a well-documented occurrence in microbial aro-
matic metabolism.
This contrasts with oxygenases for compounds,
such as benzene, toluene, or phthalate, that are less
susceptible to oxygenation by virtue of the lack of an
electron-donating substituent on the ring. For these
compounds, a more potent enzyme-generated oxy-
223
Table 2.
Compilation of sequence homologies between the
recta-cleavage
pathway proteins of the
dmp a-, xyl b-, tod c- and
bph a-systems.
Ferredoxin-like proteins DmpQ
2,3-dioxygenases, EC 1. t3.11 .- DmpB
Dehydrogenases (HMSD), EC 12.I.- DmpC
Hydrolases (HMSH) DmpD
Hydratases (OEH), EC 4.2,I.80 DmpE
CoA-dependent dehydrogenases (ADA), EC 1,2.1.10 DmpF
Aldolases (HOA), EC 4.1.3.- DmpG
Decarboxylases (4OD), EC 4.1.1.- DmpH
Isomerases (4OI), EC 5.3.2.- DmpI
XylT (64) -
XylE (84) TodE (22) BphC (22)
XylG (84) -
XylF (75) TodF (64) BphD (32)
XylJ (89) TodG (41) BphE (43)
XylQ (90) TodI (75) BphG (78)
XytK (87) TodH (78) BphF (81)
XylI (89) -
XylH (78) -
TodE/BphC (50)
The percent identity of the deduced amino acid sequences when compared with those of the
drop
system are shown in
parenthesis.
a The phenol/dimethylphenol pVl 150-encoded pathway of
Pseudomonas
sp. CF600: references as in Table 1.
b The toluene/xylene pWW0-encoded pathway of
Pseudomonasputida
mt-2: Nakai et al. (1983), Harayama et al. (1991),
Harayama & Rekik (1993), Horn et al. (1991), Chen et al. (1992).
c The chromosomally-encoded toluene pathway of
Pseudomonasputida
F1: Zylstra & Gibson (1989), Menn et al. (1991),
Lau et al. (1993) and Lau (pers. comm.).
a The chromosomally-encoded (chloro-)biphenyl pathway of
Pseudomonas
sp. KKS102: Kimbara et al. (l 989), Kikuchi et
al. (1993).
Homologies not determined in the above references were analysed using the GCG Wisconsin sequence analysis software
package .2
MONOOXYGENASES DIOXYGENASES
NADPH 2
phenol
NADP,,.~ ~t, j/~k~l~
hydroxylase
(I. cutaneum)
t1~-11
NADH2
H methane
H~~H
H~O OH
NADH2~
NAP"
~V
4mothoxy n to NADH2
p eno, NAo' C
hydroxylase
(Pseudomonas CF600)
phthalate
dioxygenase
benzoate
dioxygenase
toluene
(or biphenyl)
dioxygenase
naphthalene
dioxygenase
Fig. 3.
Composition of representative hydroxylating mono- and di-oxygenases involved in microbial metabolism. All are
Pseudomonas
enzymes except phenol hydroxylase (T.
cutaneum),
and methane monooxygenase
(M. capsulatus
(Bath),
M. trichosporium).
Subscripts denote
ferredoxin (Fd) or Reiske (R)- type 2Fe-2S eentres. Original references may be found in: Mason & Cammack (1992), Haryama et al. (1992),
and Neujahr & Gaal (1973).
224
genating agent is required. This is commonly thought
to be generated by complexation of 02 with iron in
some form at the active site of the enzyme (Fig. 3),
although in most cases shown here the nature of the
activated oxygen species is poorly defined. One or
more electron-transferring proteins are associated with
the oxygenating component of each enzyme in order to
shunt electrons from NAD(P)H for complete reduction
of 02 (Fig. 3). While these enzyme systems dioxy-
genate, some examples of multicomponent monooxy-
genases used for unactivated non-aromatic compounds
are also shown in Fig. 3.
In the light of the above, it was quite surprising to
find that Pseudomonas sp. strain CF600 elaborates a
multicomponent phenol hydroxylase that has little in
common with single-component flavoprotein hydrox-
ylases. As will be discussed below, while flavopro-
tein phenol hydroxylases are now known to be used by
some other phenol-degrading species of Pseudomonas,
the multicomponent variety is not restricted to Pseu-
domonas sp. strain CF600.
Phenol hydroxylase: A new aromatic oxygenase
As described earlier, the dmpKLMNOP gene products,
P0 to P5, were all found to be essential for growth of the
catechol-degrading test strain on phenol. Furthermore,
a series of strains harbouring plasmids lacking expres-
sion of each of the genes in turn confirmed that all six
of these genes are required (Nordlund et al. 1990a).
However, these results did not rule out the possibility
that one or more of these gene products is involved in
something other than hydroxylase activity, for example
phenol transport.
At the time the nucleotide sequences of these genes
were determined, database searches revealed similari-
ty only for the amino terminal of P5 (38.5 kDa) with
other proteins. These are plant-type ferredoxins, which
are small (11 kDa) electron-transfer proteins contain-
ing an iron-sulfur centre that is liganded by four cys-
teine residues with characteristically well-conserved
spacing. Purification of P5 revealed the presence of
flavin adenine dinucleotide (FAD) in addition to the
expected [2Fe-2S] centre, which was confirmed by
visible absorbance spectroscopy to be of the ferredox-
in type (Powlowski & Shingler 1990). The purified
protein catalyzes the NAD(P)H-dependent reduction
of cytochrome c, a property that is shared with other
proteins containing these prosthetic groups (reviewed
in Mason & Cammack 1992). The role of the fiavin in
these reductases is to accept electrons from the obligate
2-electron donor, NAD(P)H, and presumably to then
pass them one at a time to the oxygenase active site on
another polypeptide via the [2Fe-2S] centre. In some
multicomponent oxygenases the ravin and iron-sulfur
centre prosthetic groups may be present on separate
polypeptides (Fig. 3).
Since this work was published, sequences for a
number of these electron-transferring proteins have
appeared, and up-to-date comparisons including P5
can be found in another review in this volume
(Williams & Sayers 1994). Possible evolutionary rela-
tionships between ferredoxin-NADP + reductases and
P5, as well as other oxygenase reductases, are dis-
cussed by Andrews et al. 1992.
Further characterization of phenol hydroxylase,
and the roles of the five other polypeptides, P0-P4,
required the development of an in vitro assay and purifi-
cation of the individual components. In vitro activity of
the enzyme was found to be dependent on the addition
of NADH, or, less effectively, NADPH, to an assay
mixture, in addition to Fe +2 (with Fe +3 stimulating
a lower level of activity). 1 The polypeptides required
for hydroxylase activity were identified by assaying
crude extracts from strains in which genes for each
polypeptide in turn had been deleted. These experi-
ments revealed that of the six polypeptides demon-
strated to be required for growth on phenol by a
catechol-degrading test strain, only P0 was dispens-
able for hydroxylase activity ~owlowski & Shingler
t990). The function of PO, whose primary structure
does not resemble that of any protein currently in the
data bases, z is not yet known but roles in phenol trans-
port or hydroxylase regulation are two possibilities.
Purification of P1-P4 (J Powlowski & V Shingler,
unpubl.) resolves these polypeptides into two com-
ponents, one consisting of P2 alone, and the other a
complex of P1-P3-P4 (see Fig. 2). The purified pro-
teins are active in the in vitro assay when supplemented
with the reductase (P5) (Powlowski & Shingler 1990,
unpubl.). Visible absorbance spectroscopy showed that
none of the purified P1-P4 components appear to con-
tain any of the prosthetic groups commonly associated
with microbial aromatic mono- and di-oxygenases (see
Fig. 3). However, Fe +2 is still required in the assay,
raising the possibility that one of the components had
lost an intact iron centre: since S -2 was not added to
the assay, reconstitution of an iron-sulfur centre is not a
likely explanation. The presence of a Reiske-type iron-
sulfur centre, common in numerous oxygenase com-
ponents (Fig. 3), is also unlikely, based on sequence
comparisons with other Reiske-centre containing pro-
teins (Powlowski & Shingler 1990).
On the basis of these studies a general similarity
between the phenol hydroxylase from this organism
and methane monooxygenase was noted (Powlowski
& Shingler 1990). Thus, both enzymes are made up of
five polypeptides of similar sizes: one is a reductase
with FAD and a ferredoxin-type [2Fe-2S] centre; three
co-purify, and in the case of methane monooxygenase
have been demonstrated to comprise the oxygenase
itself; and one is a low molecular-weight polypeptide
that can be resolved from the others and which possess-
es no chromophoric prosthetic group (P2 in phenol
hydroxylase, component B in methane monooxyge-
nase) (see Fig. 3). The function of the latter component
in methane monooxygenase is somewhat controversial,
but appears to be involved in regulation of the oxyge-
nase activity (Green & Dalton 1985; Froland et al.
1992). The sequence similarities between components
of these two oxygenases include 28% identity between
the reductases of the two systems (Stainthorpe et al.
1989, 1990; Nordlund et al. 1990a). The next high-
est sequence similarity is found between the P2 and
MmoB components of the two systems, which share
27% identity if two large gaps are included to optimise
alignment. However, little overall sequence identity
is evident between the other polypeptides of phenol
hydroxylase and methane monooxygenase, even after
inclusion of multiple gaps.
Among the properties not yet known about phenol
hydroxylase is the identity of the oxygen-activating
prosthetic group. One possible candidate, despite the
lack of overall sequence identity, is the binuclear iron
centre like that of methane monooxygenase (Fox et
al. 1988; Ericson et al. 1988; Rosenzweig et al.
1993). While variants of this centre axe found in a
number of different enzymes, that found in the e~-
subunit of the A-component of methane monooxyge-
nase (MmoX) is similar to the binuclear iron centre of
ribonucleotide reductase, to which methane monooxy-
genase appears to be very distantly related (Nordlund
et al. I992). The crystal structure of ribonucleotide
reductase revealed that the binuclear iron centre is lig-
anded by two histidine, one aspartate, and three glu-
tamate residues (Nordlund et al. 1990b), while the
crystal structure of methane monooxygenase showed
Iigation by two histidine and four glutamate residues
(Rosenzweig et al. 1993). The spacing of these ligands
is conserved in methane monooxygenase and ribonu-
cleotide reductase, as well as in the P3 potypeptide of
phenol hydroxylase (Fig. 4). Despite the lack of over-
225
all sequence identity between ribonucleotide reductase
and methane monooxygenase, the active site struc-
tures of the two enzymes are quite similar (Nordlund
et al. 1992; Rosenzweig et al. 1993). Therefore the
tack of strong sequence identity around the putative
binuclear iron centre ligands of P3 is not particular-
ly worrisome. It is interesting to note that Thr-213 of
methane monooxygenase, thought to play an important
role in some dioxygen-activating centers is conserved
as Thr-204 in P3 (Fig. 4; Rosenzweig et al. 1993, and
references therein).
Several polypeptides of phenol hydroxylase also
share low-level sequence homology with polypeptides
of toluene-4-monooxygenase from Pseudomonas men-
docina KR1. The sequence similarities between these
two proteins have been noted previously and include
homologies with the P1, P2, P3 and P5 polypeptides
of phenol hydroxylase (Yen et al. 1991; Yen & Karl
1992). It is interesting to note that putative binuclear
iron centre ligands are also found in TmoA of toluene
monooxygenase, which shares 35% overall sequence
identity with P3 (Fig. 4). Although the presence of a
binuclear iron centre in toluene and phenol hydroxy-
lases must be confirmed by biochemical studies of the
purified proteins, it appears that oxygenases contain-
ing this centre are not limited to methane monooxyge-
nase.
Phenol hydroxylases from other bacterial strains
The oxygenation of electron-rich phenol is a consid-
erably easier task than the oxygenation of toluene
or methane. Considering that a single FAD-binding
polypeptide is so commonly used by microbes for
hydroxylation of phenolics, it is at the moment puz-
zling that Pseudomonas sp. strain CF600 elaborates
such a complex phenol hydroxylase. Indeed, two other
species of Pseudomonas, P. pickettii PKO1 and Pseu-
domonas EST1001, have clearly been shown to utilize
phenol hydroxylases that are single-component flavo-
proteins related to other aromatic flavoprotein hydrox-
ylases (Kukor & Olsen 1992; Nurk et al. 1991). Further
study of the multicomponent hydroxylase might hint
at what advantages are conferred to compensate for
the energetic expense of synthesizing six polypeptides
rather than one.
These considerations prompted us to ask whether
the multicomponent phenol hydroxylase is an orphan
or whether other phenol degrading bacteria also use
related multicomponent phenol hydroxylases. Reports
in the literature indicated that not all bacterial phenol
226
EcRNR
DmpN(P3}
MmoX
TmoA
P~EKHIFISNL~QTL~QGRSPNVALLPLIS
DARYVNAL_KLFLTAVSPL [E[YQAFQGFSRVGRQFSG
HPKWNE TMKVVSNFLEVG I E [ YNAIAATGMLWD SAQA
DPGWISTL_KSNYGAIAVG E~ ZA~_--VTGEGRNARFSKA
I PELE TWVE TWAFS TI SRSYTH I II~d~IVN- 132
-AGARVACQMQAI DIE[ LR IH[ VQTQVHAMSIIYNK- 155
-AEQKNGYLAQVLD_-]E] I~ [H I THQCAYVNYYFAK-160
-PGNRNMATFGMMD EL~ LR [HI GQLQLFFPNEYC_K-150
EcRNR
DmpN(P3)
MraoX
TmoA
LRELKKKLYLCLMSVNAL
]E I
AI RFYVSFACS FAFAERELME GNAKI IRLIARD_ LTGTQHMLNLLRS-255
DARTAGPFE FLTAVSFSF [El YVLTNLLFVPFMSGAAYNGDMATVTFGFS_AQSD MTLGLEVIKFMLE-250
L~
GFI SGDAVECSLNLQLVG I E [ ACFTNPLIVAVTEWAAANGDE I TPTVFLS iErD ~ S~ ~a MANGYQTVVS IAN-259
I I TGRDAI SVAIMLTFSF LE j TGFT-NMQ FLGLAAD~EAG-Dy T FAN L I SS-~Q T5 _-- AQQGG PALQLL I E - 247
Fig.4.
Conservation ofligands for the binucleariron centre ofdifferent enzymes. The consensus sequenceand extent ofhelices ofribonucleofide
reductase are taken from Nordlund et al. (1990b). E.
coti
ribonucleotide reductase (EcRNR, Carlson et aL 1984), DmpN (P3, Nordlund et al.
1990a), MmoX (Stainthorpe et al. 1990; Rosenzweig et al. 1993), TmoA (Yen et al. 1991). Ligating residues are boxed, other residues present
in at least three out of the four sequences shown are underlined.
hydroxylases are simple flavoproteins (Gurujeyalaksh-
mi & Oriel 1989), and that some multicomponent oxy-
genases for compounds like toluene are capable of turn-
ing over phenols (Spain & Gibson 1988). Considering
the surprising complexity of the phenol hydroxylase
revealed by our studies, it seemed unlikely that similar
enzymes would readily have been isolated, especially
if a single-component enzyme was expected. In order
to address this question, DNA from a collection of
phenol-degrading bacteria was probed for the presence
of each of the phenol hydroxylase genes.
Gene-probing experiments were performed on
eleven phenol-utilizing bacteria in addition to
Pseu-
domonas
sp. strain CF600. One of these strains was R
putida
U, the archetypal phenol-degrader whose phe-
no1 hydroxylase has not been characterised. The other
ten were marine strains collected off the Norwegian
coast. Southern hybridization analysis demonstrated
that five of the marine isolates as well as
R putida
U all possess DNA highly homologous to each of the
components of the
dmp-encoded
phenol hydroxylase.
Furthermore, all of the strains that tested positive were
also found to posses DNA highly homologous to all
nine
meta-cleavage
pathway genes of the
dmp-operon,
and to the specific regulator of phenol catabolism of
this strain (Nordlund et al. 1993).
The nucleotide sequence of the chromosomally-
encoded phenol hydroxylase region from another
phenol-degrading pseudomonad,
P. putida
P35X, has
also recently been determined (Ng et al. 1993). This
strain, like
Pseudomonas
sp. strain CF600, utilizes
a multicomponent phenol hydroxylase and a
meta-
cleavage pathway. The deduced amino acid sequences
from the six open reading frames of the phenol hydrox-
ylase region share between 89% and 99.7% identity
with those of P0 to P5 (LC Ng, pers. comm.).
It is interesting to note that all of these organ-
isms catabolize phenol using a
meta-cleavage
path-
way, while marine isolates that degrade phenol via
an
ortho-cleavage
pathway lack DNA homologous to
the multicomponent phenol hydroxylase. However,
this arrangement does not always hold, as a single-
component flavoprotein phenol hydroxylase has been
shown to be associated with the
meta-cleavage
pathway
of
P. pickettii
PKO1 (Kukor & Olsen 1991). Neverthe-
less, it is now clear that the multicomponent phenol
hydroxylase is not unique to
Pseudomonas
sp. strain
CF600. In all likelihood related enzymes will turn out
to make up a rather significant fraction of the hydrox-
ylases used in aromatic metabolism.
The meta-cleavage pathway for catechol
The biochemical route of the
meta-cleavage
pathway
following phenol hydroxylation is illustrated in Fig. 1,
and the function of the
drop-encoded
enzymes involved
are summarised in Table 1. With the exception of iso-
functional genes homologous with
dmpB and dmpQ,
no other
meta-cleavage
enzyme gene sequences had
been published prior to work with the
dmp
system.
Furthermore, only the amino acid sequence of DmpC
was similar enough to other sequences in the databas-
es to deduce a function with any confidence. Hence
assignment of the functions of individual genes relied
almost exclusively on correlating enzymatic activi-
ty with expression of individual genes. Although the
majority of the enzyme activities of the
meta-cleavage
pathway had been discovered by the mid-seventies, this
type of analysis of the
dmp
operon led to the discovery
of a new meta-cleavage pathway enzyme, aldehyde
dehydrogenase (acylating) (ADA, see Fig. 1, Table
1). In the following sections we will review the indi-
vidual enzymes of the pathway, and where appro-
priate, or where evidence is lacking for the
dmp-
encoded enzymes, isofunctional enzymes of other
meta-cleavage
operons. In particular, reference will
be made to the four
meta-cleavage
pathway systems
shown in Table 2, for which the sequences of all
meta-
cleavage pathway genes are known.
Catechol-2,3-dioxygenase: Cleavage of the aromatic
ring
This enzyme catalyzes the critical ring-opening step of
the
mere-cleavage
pathway (Fig. 1), and contains non-
heine Fe +2 at the active site (reviewed by Yamamo-
to & Ishimura 1991). The
dmpB-encoded
enzyme
shares 83-87% sequence identity with catechol 2,3-
dioxygenases from other
meta-cleavage
pathways of
different
Pseudomonas
species (Bartilson & Shingler
1989). Much lower levels of sequence similarity are
shared with other meta-cleavage dioxygenases (Table
2) that form a different evolutionary subgroup. Evo-
lutionary relationships, and up-to-date comparisons of
different ring cleavage 2,3-dioxygenases are reviewed
in this volume by Williams & Sayers (1994). Despite
the fact that catechol-2,3-dioxygenase was the first
meta-cleavage
pathway enzyme to be isolated and
crystallized, no structure or detailed mechanism has
yet been elucidated for any catechol-2,3-dioxygenase.
However, some progress has been made in manipu-
lating the substrate specificities of the xylE-encoded
catechol 2,3-dioxygenase using genetically-selected
mutants and chimeric proteins (Wasserfallen et al.
t991; Williams et al. 1990).
DmpQ: A protein with homology to plant-type
ferredoxin
The
dmpQ
gene, located between the genes encod-
ing phenol hydroxylase and catechol-2,3-dioxygenase,
encodes a polypeptide about the size of, and with up
to 41% sequence identity with, plant-type ferredoxins
(Shingler et al. i992). Homologous proteins, encod-
ed in analogous locations of the xyl and
nah
operons,
227
share 64% and 52% identity with DmpQ (Harayama
& Rekik 1993; Harayama etal. 1991; You et al. 1991).
Studies with strains in which
dmpQ
has been deleted
from the operon indicate that expression of this protein
is required to allow strains to grow at the expense of 4-
methylphenol or 3,4-dimethylphenol, but not phenol,
2-methylphenol or 3-methylphenol (V Shingler & J
Powlowski unpubl.).
This growth pattern is reminiscent of that recent-
ly reported for mutants of the xyl homologue (xylT),
namely lack of growth on 4-methyl substituted sub-
strates. The xyl homologue has been shown, using
in
vivo
studies, to be involved in reactivation of catechol-
2,3-dioxygenase (Polissi & Harayama 1993). This
enzyme is particularly sensitive to inactivation dur-
ing turnover with 4-substituted catechols, probably by
inadvertent oxidation of the active-site ferrous iron.
Reactivation is postulated to involve transfer of elec-
trons from an as-yet unidentified donor through XylT
to the oxidized iron. Considering the high degree
of similarity between the
dmp
and xyl operons, and
the qualitatively similar growth patterns of
dmpQ and
xylT
mutants, it is likely that DmpQ performs the
same function as XylT. It is interesting to note, that
while the sequence identity between the catechol 2,3-
dioxygenases in the two strains is 84%, XylT and
DmpQ share only 64% identity. It is possible that these
ferredoxins have evolved to accommodate interaction
with different electron donors in the two strains. Hence
the greater divergence of XylT and DmpQ might reflect
these accommodations rather than a response to struc-
tural differences between the corresponding catechol-
2,3-dioxygenases with which they presumably inter-
act.
The presence of the ferredoxin-like protein in the
xyl-, dmp-
and
nah-encoded
pathways supports the
idea that inclusion of these redox proteins represents
a bacterial strategy to expand the substrate speci-
ficity of the
meta-cleavage
pathway to include com-
pounds chanelled through the otherwise suicidal 4-
methylcatechol (Polissi & Harayama 1993). In this
respect the observation that neither the
tod-
nor the
bph-
abridged
recta-pathway
operon appears to encode a
similar protein is illuminating (Lau et al. 1993; Kikuchi
eta|. 1993). In both of these enzyme systems, no com-
pounds that would be metabolised
viapara-substituted
intermediates have been shown to be substrates of the
respective pathways.
228
2-Hydroxymuconic semialdehyde dehydtvgenase
(HMSD) and hydrolase (HMSH): gatekeepers at the
pathway branch point
These enzymes both use ring cleavage products as
substrates, which are 2-hydroxymuconic semialdehyde
from catechol, 5-methyl-2-hydroxymuconic semialde-
hyde from 4-methylcatechol, and 2-hydroxy-6-oxo-
2,4-heptadienoate from 3-methylcatechol. Since the
ring-cleavage product of 3-methylcatechot is a ketone,
rather than an aldehyde, it cannot be further oxidized by
the dehydrogenase, and must therefore be metabolised
via the hydrolytic route (Fig. 1). This was originally
shown using hydrolase-defective strains of
R putida U
that failed to grow at the expense of phenols that are
channelled via the
meta-cleavage
pathway through 3-
methylcatechol (Bayly & Wigmore 1973). Despite the
potential for use of either branch for the ring-cleavage
products of catechol or 4-methylcatechol, these com-
pounds were shown to be preferred substrates for the
dehydrogenase, rather than the hydrolase of the
meta-
cleavage pathway of
P. putida
U (Sala-Trepat et al.
1972). Conclusions from these in vitro results were
reinforced by the observation that mutants of
P. puti-
da
U defective only in the 4-oxalocrotonate branch
failed to grow at the expense of phenols channelled
through catechol or 4-methylcatechol (Bayly & Wig-
more 1973).
These conclusions are also apparently valid for the
dmp-encoded meta-cleavage
pathway. Thus, a dele-
tion within the HMS-dehydrogenase gene
(dmpC),
or
either of the genes for the other two enzymes of the
4-oxalocrotonate branch
(dmpl
and
dmpH),
resulted
in strains that grew on 2-methyl, 3-methyl and 3,4-
dimethylphenols, but not on phenol or 4-methylphenol.
On the other hand, strains deleted within
dmpD,
the
gene encoding the HMS-hydrolase, grew on phenol
and 4-methylphenol, but not at the expense of the oth-
er substrates (V Shingler & J Powlowski unpubl.).
Despite the structural similarity of the substrates
of both enzymes, the lack of sequence homolo-
gy between HMS-hydrolase and HMS-dehydrogenase
indicates different evolutionary origins. The amino
acid sequence of HMS-dehydrogenase shows approxi-
mately 40% identity with various eukaryotic aldehyde
dehydrogenases (Nordlund & Shingler, 1990), and
84% identity with xyl-encoded HMS-dehydrogenase
(Horn et at. 1991). Speculations about the possi-
ble evolutionary significance of this similarity have
been presented for the xylG-encoded enzyme (Horn
et al. 1991), and can equally well be applied to
the dmpC-encoded
counterpart. Sequence comparison
searches for HMS-hydrolase, on the other hand, only
readily identify isofunctional hydrolases including the
xyl and
tod-encoded
enzymes, as well as hydrolases
for phenyl-substituted 2-hydroxymuconic semialde-
hyde, which is an intermediate in microbial biphenyl
metabolism (Table 2). In the latter case, the sequence
identity is considerably lower than that for the oth-
er HMS-hydrolases, possibly reflecting differences in
substituent bulkiness on the substrate.
While information about structure-function rela-
tionships for any HMS-hydrolase or HMS-dehydro-
genase is non-existent, the sequence comparisons sug-
gest avenues for further experimentation. Consider first
the striking similarity between HMS-dehydrogenase
and eukaryotic aldehyde dehydrogenases. While
chemical modification studies have implicated a num-
ber of residues in enzyme activity, including Cys-
302 and Glu-268 of horse aldehyde dehydrogenase
(conserved as Cys-288 and Glu-254 in
dmp-encoded
HMS-dehydrogenase), assignment of definite roles for
these residues has been difficult (reviewed in Hempel
& Jrrnvall 1989). By analogy with glyceraldehyde-3-
phosphate dehydrogenase, the mechanism of aldehyde
dehydrogenase has often been assumed to involve for-
mation of a thiohemiacetal enzyme-substrate interme-
diate that transfers a hydride equivalent to NAD +, and
is then hydrolyzed from the enzyme. However, com-
pelling evidence has been presented that an active-site
serine residue of sheep-liver aldehyde dehydrogenase
is invoh, ed in formation of an acyl-enzyme intermedi-
ate (Loomes et at. 1990). It is interesting that the serine
residue implicated in these studies is conserved in alde-
hyde dehydrogenase isozymes (Johansson et al. 1988),
while it is not conserved in the HMS-dehydrogenase
sequence. HMS-dehydrogenase thus provides a new
variation with which to test roles proposed for the var-
ious conserved aldehyde dehydrogenase residues, and
to probe the involvement of an active-site serine residue
in catalysis.
The participation of an active site serine residue
in HMS-hydrolase catalysis has also been suggested
on the basis of low sequence identity (20% overall)
of the isofunctional xy/-encoded enzyme with atropine
esterase, a serine hydrolase (Horn et al. 1991). This
observation can be extended to note that Set-107 of
the
dmp-encoded
HMS-hydrolase lies within an active-
site lipase consensus sequence GXSXG (reviewed by
Derewenda & Sharp 1993). HMS-hydrolase shares
24% overall identity with a lipase from
Moraxella
(Feller et al. 1991). The serine residue within the lipase
229
consensus sequence is coupled in lipases to Asp (or
Glu) and His, much like the archetypal catalytic triad
of serine proteases (see Derewenda & Sharp 1993).
On the basis of these comparisons it is conceivable
that the mechanism of HMS-hydrolase, which cleaves
a carbon-carbon bond, is similar to the mechanisms of
esterases, in which a carbon-oxygen bond is cleaved. It
is interesting to note that the lipase consensus sequence
with a Cys replacing Ser is a motif conserved in the
ortho-cleavage pathway enzyme dienelactone hydro-
lase (Pathak & Ollis 1990).
4-Oxalocrotonate isomerase (401): a small but
essential participant in the HMS-dehydrogenase
initiated branch
4-oxalocrotonate isomerase (401) catalyzes the iso-
merization of 4-oxalocrotonate, the product of HMS-
dehydrogenase, to 2-keto-3-hexenedioate (Fig. 1).
Studies using mutants of
P. putida U lacking 4OI activ-
ity showed that this enzyme is necessary to support
growth on phenols channelled through catechol or 4-
methylcatechot, which are metabolized via this branch
(Wigmore et al. 1974). As indicated earlier, this is also
the case for the dmp-encoded pathway, since a dele-
tion within the 4OI-encoding gene (dmpl) of the oper-
on prevented the resulting strain from growing at the
expense of phenol and 4-methylphenol. These results
demonstrate the indispensability of 4OI, which is sig-
nificant since the non-enzymatic isomerization reac-
tion is rather fast and could have been sufficient to
support growth in the absence of a specific enzyme
(Sala-Trepat & Evans 1971).
The sequence of 4OI shares 78% identity with
the isofunctional xylH-encoded enzyme, but no sim-
ilarity is found with other protein sequences. 4-
Oxalocrotonate isomerase encoded by the xyl operon
has been isolated and partially characterized (Chen et
al. 1992). Mechanistic studies suggest that the isomer-
ization of
2-oxo-4-trans-hexenedioate proceeds via the
enol form of the substrate shown in Fig. 1, and in
this respect at least, 4OI resembles 3-oxo-AS-steroid
isomerase, as well as other isomerases (Bayly & Bar-
bour 1984; Whitman et al. 1991). It will be interest-
ing to compare the properties of these two isofunc-
tional 4Ois, as 22% of the amino acid sequence is
not identical, in order to see what conclusions can
be drawn about structure-function relationships. X-ray
structures of both the x3'l-encoded enzyme (Davenport
& Whitman 1993) and the drop-encoded enzyme (D.
Roper unpubl.) are currently being sought.
4-Oxalocrotonate decarboxylase (,COD) and
2-oxopent-4-dienoate hydratase (OEH): A complex of
related polypeptides
4OD catalyzes the final step in the 4-oxalocrotonate
branch, while OEH is involved in the subsequent
reaction, which occurs after the 4-oxalocrotonate and
hydrolytic branches merge (Fig. 1). The first attempt-
ed purification of OEH, from phenol-grown P. putida
U, was reported by Collinsworth et al. (1972). The
enzyme proved difficult to purify, and could not be
resolved from some persistent contaminants. One of
the contaminants was undoubtedly 4OD, as years later
it was shown that xyl operon-encoded 4OD and OEH
were tightly enough associated with each other to co-
purify. As the unstable product of 4OD is the substrate
for OEH, it has been proposed that the physical associa-
tion between the two enzymes has arisen to channel the
unstable intermediate (Harayama et al. 1989). Howev-
er, as these workers pointed out, HMS-hydrolase of
the hydrolytic branch generates the same intermedi-
ate, and this enzyme is not tightly associated with the
hydratase (OEH) in the x3,l-encoded system. A loose
association that still allows metabolite channelling has
not been ruled out.
Sequencing of the genes for OEH and 4OD from the
dmpoperon suggested a possible evolutionary expla-
nation for the association of these two polypeptides
(Shingler et al. 1992). As is the case for the xyl-
encoded homologues, the dmp-encoded enzymes are
also tightly associated with each other, and co-purify
(J. Powlowski unpubl., but see Fig. 2). When opti-
mally aligned, their deduced amino acid sequences
show 37% identity, indicating common ancestry. It is
therefore possible that these enzymes evolved from a
protein that was originally multimeric, and this may be
why they are so tightly associated with each other. The
importance of channelling, which could result from
such an arrangement, remains to be demonstrated. If
metabolite channelling is indeed important it may have
played a role in maintaining selective pressure on the
multimeric form of the enzymes.
It seems rather remarkable that an enzyme that cat-
alyzes carbon-carbon bond cleavage is related to one
that involves hydration of a double bond. However,
similar electron shifts can reasonably be postulated for
these two reactions (Fig. 5A, Dagley 1975), so evo-
lution of the two enzymes from a common ancestor
would perhaps not be too surprising. In both cases,
the keto group of the substrate could act as an elec-
tron sink, especially when complexed with a metal
230
A
DmpH DrapE
coon ~3 .coon
COOH CH3 COOH
O= HO/~ ~ "
B
DmpG Ni_fV
COOH
~3 coon no.I o
COOH
+ .COOH
H=c I ~co A
Fig. 5.
Comparisons of reactions catalyzed by DmpE (OEH) and
DmpH (4OD) (A), and DmpG (HOA) and NifV (B). While the
mechanisms of these enzymes are as yet not known, these reasonable
possibilities are illustrated for comparative purposes (see text).
ion, a requirement for which has been demonstrat-
ed for both the decarboxylase (4OD, Sala-Trepat &
Evans 1971; Harayama et al. 1989) and the hydratase
(OEH, Collinsworth et al. 1972; Harayama et al. 1989).
However, it is not currently known whether the iso-
mer of the hydratase substrate shown in Fig. 5A is
catalytically competent. 2-Keto-4-pentenoate, which
is the form shown in Fig. 1, is in rapid equilibri-
um with the enol form, 2-hydroxy-2,4-pentadienoate
(Marcotte & Walsh 1978), which is competent for
turnover by OEH (Harayama et al. 1989). A solution
of these isomers decays to trans-2-keto-3-pentenoate
(Marcotte & Walsh 1978) which is probably the form
not metabolised by the enzyme (Harayama et al. 1989).
The involvement of cis-2-keto-3-pentenoate (the iso-
mer shown in Fig. 5A), as a substrate or transient inter-
mediate in the OEH catalyzed reaction, has not been
specifically addressed.
It is worth emphasizing that from comparisons of
the sequences of all meta-pathway proteins, dmpE
(hydratase, OEH) and dmpH (decarboxylase, 4OD)
represent the only example of possible evolution by
gene duplication. None of the other protein sequences
show significant similarity with each other. Hence,
gene duplication and divergence were not foremost
in evolution of this pathway (Shingler et al. 1992). In
this respect the comparatively low homology between
the hydratase (OEH)-encoding genes of the dmp-
encoded meta-pathway arid those of the abridged tod-
and bph-encoded meta-pathways is notable (Table 2).
Metabolism of the substrates of the tod and bph-
systems does not require the 4-oxalocrotonate branch
of the pathway, and in neither case do the respec-
tive operons include genes for this branch (Lau et al.
1993; Kikuchi et al. 1993, pers. comm.). Therefore,
in these abridged meta-pathways the hydratase coun-
terparts, which use exactly the same substrate as the
dmp-encoded enzyme, would have evolved under dif-
ferent conditions i.e., in the absence of a physically-
associated decarboxylase partner.
4-Hydroxy-2-ketovalerate aldolase (HOA) and
aldehyde dehydrogenase (acylating) (ADA): A second
example of physical association and a new
meta-cleavage pathway enzyme
Close physical association has also been conclu-
sively demonstrated for the last two enzymes of
the dmp-encoded pathway, 4-hydroxy-2-ketovalerate
aldolase (HOA) and aldehyde dehydrogenase (acylat-
ing) (ADA), which generate the end-products pyruvate
and acetyl-CoA, respectively. Like the two preceed-
ing enzymes, HOA and ADA purify to homogeneity
(Powlowski et al. 1993, Fig. 2), but unlike them do not
share common ancestry. Close association in this case
may be required to channel the aldehyde produced by
the aldolase either to reduce toxicity or to ensure that
it is metabolized by ADA, and not some less efficient
route (see below). A consequence of this association
appears to be that the activity of HOA is regulated
by NAD +, and to a lesser extent NADH, which are
used not by the aldolase itself, but by its partner, ADA
(Powlowski et al. 1993). The mechanism by which this
occurs has not yet been elucidated, but conceivably
involves allosteric interactions. This phenomenon, and
the possibility of channelling, may be significant for
modulation of pathway activity at the enzyme level.
Aldolases in general can be divided into two class-
es: those that require a metal ion, and those that rely
on an active-site lysine, for interaction with the keto-
group of the substrate to provide an efficient electron
sink (Walsh 1979). It has long been known that the
meta-cleavage
aldolase, HOA, requires a metal ion
for activity. While that from P.
putida
U is apparent-
ly stimulated by Mg +2 (Dagley & Gibson 1965), the
drop-encoded
HOA is stimulated by Mn +2, with no
increase in activity observed using Mg +2 (Powlowski
et al. 1993). However, the enzyme is still active in the
absence of added metal ions, and no attempt has yet
been made to correlate metal content of the enzyme
with activity.
The
drop-encoded
HOA shares low sequence
homology with enzymes that catalyze reactions that are
mechanistically the reverse of the aldol cleavage reac-
tion catalyzed by HOA (Shingler et al. 1992, and refer-
ences therein). One example is NifV, which is thought
to encode homocitrate synthase that catalyzes Claisen
condensation of acetyl-CoA and o~-ketoglutarate (Fig.
5B).
For years the meta-cleavage pathway has been
depicted as ending at pyruvate and acetaldehyde (or
propionaldehyde in the case of 4-methyl-substituted
growth substrates). Work with the
drop
operon has
uncovered the existence of a new
rneta-cleavage
oper-
on encoded enzyme that converts the aldehyde to the
acyl-CoA derivative using NAD + and coenzyme A
(ADA; Shingler et al. 1992; Powlowski et al. 1993).
Metabolism in other organisms involves oxidation of
acetaldehyde to acetate, followed by ATP-dependent
conversion to the CoA ester (reviewed in Nunn 1987).
The ATP-independent reaction catalyzed by ADA thus
represents an energetically efficient mechanism for
catabolism of the short-chain aldehyde formed by the
pathway. However,
Pseudomonas
CF600 grown at the
expense of phenol also has low levels of enzymes
involved in an ATP-dependent pathway, so some
metabolism of acetaldehyde or propionaldehyde by
these enzymes is possible. It is difficult to test direct-
ly whether or not the
drop-enzyme
is absolutely nec-
essary, as the close coupling of ADA activity with
HOA activity means that mutations or deletions within
the ADA-gene also affect HOA activity second-hand
(Shingler et al. 1992).
Apart from recently-discovered isofuncfional
enzymes (Table 2), the sequence of aldehyde dehy-
drogenase (acylating) (ADA) is not similar to any oth-
er proteins so far entered into the databases. Howev-
er, the amino-terminal of the protein may encompass
the NAD+-binding site, as it contains a characteris-
tic ADP-binding fingerprint sequence (Wierenga et al.
1986, Shingler et al. 1992).
Sequence information is available for at least
one other bacterial CoA-dependent dehydrogenase,
231
namely methylmalonate semialdehyde dehydrogenase
(Steele et al. 1992), and its rat counterpart (Kedishvili
et al. 1992). It is interesting that these sequences show
higher similarity to the CoA-independent eukaryot-
ic aldehyde dehydrogenase, than to the
dmp-encoded
CoA-dependent dehydrogenase, ADA. This suggests
at least two separate evolutionary origins for the
CoA-dependent dehydrogenases, possibly involving
one ancestral NAD +-binding protein and another that
bound CoA. It is perhaps even more intriguing that the
CoA-independent HMS-dehydrogenase from the
dmp-
encoded pathway is also related to eukaryotic alde-
hyde dehydrogenases but not to the CoA-dependent
dehydrogenase (ADA). When the two
dmp-encoded
dehydrogenases are directly compared they show <
20% identity even with insertion of sixteen gaps to
optimise the alignment. Thus, while these two dehy-
drogenases of the
meta-cleavage
pathway might have
been predicted to have evolved from a common ances-
tor, sequence comparisons indicate they have not. It
will be interesting to compare the sequences of other
biochemically-characterized bacterial CoA-dependent
dehydrogenases, e.g.,
Clostridium
acetaldehyde dehy-
drogenase (acylating) (Burton & Stadman 1953), when
they become available.
Transcriptional regulation of the dmp operon
Transcription of the
dmp
operon is tightly regulat-
ed by the divergently transcribed
dmpR
gene prod-
uct (see Fig. 1). As with other regulators of aromatic
catabolism, the activity of the regulator is itself modu-
lated by the presence of aromatic compounds to allow
expression of the specialized catabolic enzymes only
when the substrates of the pathway are present (Shin-
gler et al. 1993). In this capacity the aromatic com-
pound serves as an effector molecule. For a given com-
pound to support growth as the sole carbon and energy
source, it must therefore serve as both an effector for
the regulator and as a substrate for the enzymes of the
pathway. Hence, the effector recognition specificity of
the regulator, in addition to specificities of the catabol-
ic enzymes, is intimately involved in determining the
range of compounds that can be degraded by a metabol-
ic pathway. This has been most elegantly exploited
in manipulation of the pWW0-encoded XylS regula-
tor of benzoate metabolism. Studies of this regulator
have shown that new substrates for thepWW0-encoded
pathway could be selected by mutational expansion of
the effector specificity range of XylS (Abril et al. 1989;
Ramos et al. 1986, 1990).
232
DmpR, like the pWW0-encoded XylR regulator of
toluene and xylene catabolism, belongs to the exten-
sive NtrC-family of transcriptional activators (Shin-
gler et al. 1993). Members of this family respond
to diverse environmental signals to regulate genes
involved in a variety of physiological processes. These
transcriptional activators act by binding to enhancer-
like elements, and regulating transcription from a dis-
tinct set of promoters recognised by ~r54-holoenzyme
RNA polymerase (reviewed in North et al. 1993). As
with eukaryotic enhancer-binding proteins, members
of the bacterial NtrC family are composed of distinct
functional domains. The central and carboxy-terminal
domains of these activators are conserved among all
members of the family and are believed to be involved
in interaction with the 0 .54 RNA polymerase and in
DNA binding, respectively. In many members of the
family the amino-terminal A domains are the sites of
transfer of the signals that are received via sensory
proteins. However, the A domain of DmpR shares
64% homology with the equivalent domain of XylR,
but does not share homology with other members of
the NtrC family or other proteins in the data bases
(Shingler et al. 1993). The unique A-domain homol-
ogy shared by DmpR and XylR, in conjunction with
the signal receptor function of the A-domains in other
members of the family, suggested that this region is
involved in the activation of DmpR and XylR by their
respective aromatic effectors. Studies of the response
of hybrid DmpR/XylR regulators, in which all or parts
of the A-domain of DmpR were exchanged with that of
XylR, have now clearly shown that the distinct effec-
tot recognition specificities of the two regulators spans
and completely resides within the amino terminal 234
residues (Shingler & Moore 1994).
Both DmpR and XylR have broad effector speci-
ficity, but respond differentially depending on the posi-
tion and nature of the substituent(s) on the aromatic
ring of the effector (Abril et al. 1989; Shingler &
Moore 1994). In the case of DmpR the response to
phenolic compounds substituted in the
para-position
is
relatively poor, suggesting that the inefficient growth
observed on these compounds might be due to low
expression of the catabolic enzymes. This is apparent-
ly the case, since a
Pseudomonas
strain carrying the
dmp
system on an RSF1010-based ptasmid of 16-20
copies/celt has a generation time of 66 minutes in min-
imal media containing 4-methylphenol as the carbon
source. This compares favourably with the 132 minute
generation time of this strain harbouring the 1-2 copy
number pVI 150 wild-type plasmid (H. Pavel & V. Shin-
gler unpubl.). However, the improved 4-methytphenol
degrader was unable to grow on phenol, a phenomenon
associated with hyperproduction of the phenol hydrox-
ylase in response to the efficient effector property of
phenol. Although not specifically tested, growth inhi-
bition in this case may be due to interference with
aromatic amino acids, a possibility suggested by the
observation that expression of phenol hydroxylase in
E. coli
results in accumulation of a brown mixture of
pigments produced from aromatic amino acids in the
media &owlowski & Shingler 1990). The detrimen-
tal effects of wholesale over-expression in this case
suggests that more subtle modifications, such as effec-
tor specificity mutants, are likely to be more success-
ful in construction of strains with all-round improved
degradative properties.
Most aromatic catabolic regulators so far analysed
are members of one of three families of bacterial tran-
scriptional regulators: the LysR family e.g., NahR,
CatM, CatR, TcbR, TfdS and ClcR (see Coco et al.
1993, and references therein), the AraC family e.g.,
XylS (Inouye et al. 1986; Ramos et al. 1990), or the
NtrC family e.g., XylR and DmpR (Inouye et al. 1988;
Shingler et al. 1993). For a long time XylR was the
sole example of an NtrC-like regulator. However, as
described earlier, gene-probing experiments using a
dmpR
A-domain gene probe identified highly homolo-
gous DNA associated with the phenol catabolic genes
of
P. putida
U and five marine isolates (Nordlund et
al. 1993). Similarly, a regulatory region with high
homology to
xylR
has been located upstream of the
phenol catabolic genes of P.
putida
P35X (Ng et al.
1993). Therefore, the NtrC-like transcriptional activa-
tors promise to feature more prominently in the future
with respect to regulation of aromatic catabolism.
Concluding remarks
Current knowledge of
meta-cleavage
pathway enzy-
mology, and to some extent chemistry, lags behind
genetic characterization of the pathway. However,
analysis of the genes of the complete pathway, and
comparisons of the amino acid sequences of related
proteins, have identified potentially important residues
for many of these enzymes. Moreover, cloning and
over-expression of these genes can help provide large
quantities of the proteins for enzymological and struc-
tural studies. In the absence of structural data, the infor-
mation gained from genetic analysis provides a rational
basis to begin testing the importance of residues iden-
tiffed as, for example, potentially ligating prosthetic
groups, interacting with co-factors, or being involved
in catalysis. One caveat is that since a number of the
enzymes of the
meta-cleavage
pathway appear to be
unrelated to other proteins, and others are associated
with each other, it is particularly important to initially
characterize the enzymes from the parent strain before
proceeding to large-scale over-expression systems.
A fuller understanding of
meta-cleavage
pathway
enzymology also depends on better characterization of
the lower pathway chemistry. At the very least, bet-
ter methods for preparation of some of the enzyme
substrates are required (e.g., Powlowski et al. 1993).
In addition, since many of these substrates can exist
as keto-enol tautomers and/or
cis-trans
isomers, they
must be very carefully characterized. In this respect
the recent work of Whitman and colleagues with 4-
oxalocrotonate isomerase is exemplary (Whitman et
al. 199t; Chen et al. 1992).
One of the particularly interesting properties of
some of the
meta-cleavage
pathway enzymes, shown in
this and other work, is the formation of tight multien-
zyme complexes. In the
dmp
system tight associations
between the related polypeptides DmpH and DmpE
(4OD and OEH), and the unrelated DmpF and DmpG
polypeptides (ADA and HOA) have been identified
(see Fig. 2). In both cases the enzyme pairs catalyse
sequential steps of the pathway and the member of each
pair that makes the substrate for its partner is appar-
ently dependent on the presence of that partner for
activity (Shingler et al. 1992; Harayama et al. 1989).
In the case of the DmpF/DmpG pair, cross-regulation
of enzyme activity is possible (Powlowski et al. 1993).
These properties suggest that metabolic channelling
may be important for pathway efficiency, thus provid-
ing additional evolutionary impetus for clustering of
the genes involved into an operon. Co-purification of
enzyme activities through multiple steps of purifica-
tion only identifies tight associations and it may well
be that other enzymes of the pathway are also associat-
ed, albeit loosely. The existence of these associations,
and the so-far unproven idea of metabolite channelling,
certainly merit further attention.
The application of microbial metabolic activities to
detoxiffcation and clean-up has stimulated much inter-
est in construction of strains with improved degrada-
tive efficiency or expanded catabolic capacity. While
changes in substrate specificities of enzymes or effector
recognition by regulators have been successful, little
has been reported on genetic manipulation of strains
for improved efficiency. One possible strategy is simple
233
over-expression of the enzymes involved. The obser-
vation of comptexing between
meta-cteavage
pathway
enzymes has obvious practical implications for this
type of work. In addition, the cross-regulation evident
between the DmpF/DmpG pair highlights the fact that
very little is really known about (potential) regulation
at the enzyme level in this pathway. This will undoubt-
edly be remedied as progress on pathway enzymology
is made. Finally, the toxic effects of over-expression of
some of the enzymes (e.g., phenol hydroxylase) sug-
gest further hindrances to increasing strain efficiency
by brute force over-expression.
It is when manipulations for practical applications
are considered that the importance of a comprehen-
sive knowledge of pathway genetics, enzymology, and
chemistry is realized. The current boom in knowledge
of the genetics of the
meta-cleavage
pathways, and
their associated oxygenases, contributes enormously
to bringing this goal closer for at least one type of
model system.
Acknowledgements
We wish to thank our many colleagues, particularly P
Williams, K Kikuchi, PCK Lau, CL Poh, LC Ng, L
Sahlman, D Ahmad and M Gullberg for useful com-
ments, discussion and for providing information prior
to publication. Work in our laboratories was supported
by the National Swedish Research Councils for Nat-
ural Science and Engineering Science, The Natural
Science and Engineering Research Council of Canada,
and Concordia University.
Notes
1. In the original reference (Powlowski & Shingler
1990), the metal ion and NADH concentrations in
the
in vitro
assay was given in mlvI. This was a
typographical error, and the concentrations should
have read 100#M and 280#M respectively.
2. Computer searches of the EMBL/GenBank
(release 35) were made using the TFastA program
of the GCG Wisconsin sequence analysis software
package. One to one protein sequence alignments
were performed using the BestFit program.
234
References
Abril M-A, Michan C, Timmis KN & Ramos JL (1989) Regulator
and enzyme specificities of the TOL plasmid-encoded upper path-
way for degradation of aromatic hydrocarbons and expansion of
the substrate range of the pathway. J. Bacteriol. 171:6782-6790
Andrews SC, Shipley D, Keen JN, Findley JBC, Harrison PM
& Guest JR (1992) The haemoglobin-like protein (HMP) of
Escherichia coli
has ferrisiderophore reductase activity and its
C-terminal domain shares homology with ferredoxin NADP +
reductases. FEBS Lett. 302:247-252
Anonymous (1993) Facts and figures for the chemical industry.
Chem. Engineer. News 71:38-82
Assinder S & Williams PA (1990) The TOL plasmids: determinants
of the catabolism of toluene and xylenes. Adv. Microbial Physiol.
31:1-39
Ballou DP (1982) Flavoprotein monooxygenases. In: Massey V &
Williams CH (Eds) Flavins and Flavoproteins (pp 301-310). Else-
vier Science Publishing, Amsterdam
Bartilson M, Nordlund I & Shingler V (1990)Location and organi-
zation of the dimethylphenol catabolising genes of
Pseudomonas
CF600. Mol. Gen. Genet. 220:294-300
Bartilson M & Shingler V (1989) Nucleotide sequence and expres-
sion of the catechol 2,3-dioxygenase-encoding gene of phenol-
catabolizing
Pseudomonas
CF600. Gene 85:233-238
Bayly RC & Barbour MG (1984) The degradation of aromatic com-
pounds. In: Gibson DT (Ed) Microbial Degradation of Organic
Compounds (pp 253-294). Marcel Dekker, New York
Bayly RC & Wigmore GJ (1973) Metabolism of phenol and cresols
by mutants of
Pseudomonasputida.
J. Bacteriol. 113:1112-1126
Burton RM & Stadman ER (1953) The oxidation of acetaldehyde to
acetyl coenzyme A. J, Biol. Chem. 202:873-890
Carlson J, Fuchs JA & Messing J (1984) Primary structure of the
Escherichia coIi
ribonucleoside diphosphate reductase opemn.
Proc. Natl. Acad. Sci. U.S.A. 81:4294-4297
Carrington B, Lowe A & Williams PA (1994) The lower path-
way operon for benzoate catabolism in biphenyl-utilising
Pseu-
domonas
sp. strain IC and the nucleotide sequence of the
bphE
gene for catechol 2,3-dioxygenase. Microbiology 140:499-508
Chen LH, Kenyon GL, Curtin F, Harayama S, Bembenek ME,
Hajipour G & Whitman CP (1992) 4-Oxalocrotonate tau-
tomerase, an enzyme composed of 62 amino acid residues per
monomer. J. Biol. Chem. 267:17716-17721
Coco WM, Rothmel RK, Henikoff S & Chakrabarty AM (1993)
Nucleotide sequence and initial functional analysis of the
clcR
gene encoding a LysR Family activator of the
clcABD
chlorocat-
echol operon in
Pseudomonasputida.
J. Bacteriol. 175:417-427
Collinsworth WL, Chapman PJ & Dagley S (1972) Stereospecific
enzymes in the degradation of aromatic compounds by
Pseu-
domonas putida.
J. Bacteriol. 113:922-931
Dagley S (1975) A biochemical approach to some problems of envi-
ronmental pollution. Essays Bio.chem, 11:81-138
Dagley S (1986) Biochemistry of aromatic hydrocarbondegradation
in pseudomonads. In: Sokatch JR (Ed) The Bacteria, Vol 10 (pp
527-556). Academic Press, New York
Dagley S & Gibson DT (1965) The bacterial degradation of catechol.
Biochemistry 95:466-474
Davenport RC & Whitman CP (1993) Preliminary analysis of crys-
tals of 4-oxalocrotonate tautomerase, an enzyme composed of
unusually small monomers. J. Mol. Biol. 231:509-512
Derewenda ZS & Sharp AM (1993) News from the interface: the
molecular structure of triacylglyceride lipases. Trends Biochem.
Sci. 18:20-25
Entsch B, Ballou DP & Massey V (1976) Flavin-oxygen derivatives
involved in hydroxylation by p-hydroxybenzoate hydroxylase. J.
Biol. Chem. 251:2550-2563
Ericson A, Hedman B, Hodgson KO, Green J, Dalton H, Bentsen
JG, Beers RH & Lippard SJ (1988) Structural characterization
by EXAFS spectroscopy of the binuclear iron centre in protein
A of methane monooxygenase from
Methylococcus capsulatus
(Bath). J. Amer. Chem. Soc. 110:2330-2332
Feller G, Thiry M & Gerday C (1991) Nucleotide sequence of the
lipase gene
lip3
from the antarctic psychotroph
Moraxella
TA 144.
Biochim. Biophys. Acta 1088:323-324
Fennewald M, Prevatt W, Meyer R & Shapiro J (1978) Isolation of
IncP-2 DNA fromPseudomonasaeruginosa.Plasmid
1:t64-173
Fox BG, Surerus KK, Mt~nck E & Lipscomb JD (1988) Evidence for
a #-oxo-bridged binuclear iron cluster in the hydroxylase com-
ponent of methane monooxygenase. J. Biol. Chem. 263:10553-
10556
Froland WA, Andersson KK, Lee S-K & Lipscomb JD (1992)
Methane monooxygenase component B and reductase alter the
regioselectivity of the hydroxylase-catalyzed reaction. J, Biol.
Chem. 267:17588-17597
Green J & Dalton H (1985)Protein B of soluble methane monooxy-
genase from
Methylococcus capsulatus
(Bath). J. Biol. Chem.
260:15795-15801
Gurujeyalakshmi G & Oriel P (1989) Isolation of phenol-degrading
Bacillus stearothermophilus.
Appl. Environ. Microbiol. 55:500-
5O2
Happold FC & Key A (1932) The bacterial purification of gas-works
liquors. The action of liquors on the bacterial flora of sewage. J.
Hygiene 32:573-580
Harayama S, Kok M & Neidle EL (1992) Functional and evolution-
ary relationships amoung divers oxygenases. Annu. Rev. Micro-
biol. 46:565-601
Harayama S, Polissi A & Rekik M (1991) Divergent evolution of
chloroplast-type ferredoxins. FEBS Lett. 285:85-88
Harayama S & Rekik M (1993) Comparison of the nucleotide
sequence of the
recta-cleavage
pathway genes of TOL plas-
mid pWW0 from
Pseudomonasputida
with other
meta-cteavage
genes suggests that both single and multiple nucleotide substitu-
tions contribute to enzyme evolution. Mol. Gen. Genet. 239:81-
89
Harayama S, Rekik M, Ngai K-1 & Ornston LN (1989) Physical-
ly associated enzymes produce and metabolize 2-hydroxy-2,4-
dienoate, a chemically unstable intermediate formed in catechol
metabolism via
meta-cleavage
in
Pseudomonasputida.
J. Baete-
riol. 171:6251-6258
Hempel J & JBrnvall H (1989) Aldehyde dehydrogenases-structure.
In: Crow KE & Batt RD (Eds) Human Metabolism of Alcohol,
Vol II, Regulation, Enzymology and Metabolites of Ethanol (pp
77-88). CRC Press, Boca Raton, FL
Horn JA, Harayama S & Timmis KN (1991) DNA sequence deter-
mination of the TOL plasmid
xyIGFH
genes
Pseudomonasputi-
da:
implications for the evolution of aromatic catabolism. Mol.
Microbiol. 5:2459-2474
Howard PH, Boethling RS, Jarvis WE Meylan WM & Michalenko
EIvI (1991) Handbook of Environmental Degradation Rates.
Lewis Publishers Inc, Chelsea, MI
Inouye S, Nakazawa A & Nakazawa T (1986) Nncleotide sequence
of the regulatory gene
xylS
on the
Pseudomonas putida TOL
plasmid and identification of the protein product. Gene 44:235-
242
Inouye S, Nakazawa A & Nakazawa T (1988) Nucleotide sequence
of the regulatory gene
xyIR
of the TOL plasmid from
Pseu-
domonasputida.
Gene 66:301-306
Johansson J, Von Bahr-Lindstr6m J, Jeck R, Woenckhaus C &
J6mvall H (1988) Mitochondrial aldehyde dehydrogenase from
horse liver: correlation of the same species variants for both the
cytosolic and mitochondrial forms of an enzyme. Enr. J. Biochem.
172:527-533
Kedishvili NY, Popov KM, Rougraff PM, Zhao Y, Crab DW &
Harris RA (1992) CoA dependent methylmalonate-semialdehyde
dehydrogenase, a unique member of the aldehyde dehydrogenase
superfamily. J. Biol. Chem. 267:19724-19729
Kikuchi K, Yasukochi Y, Nagata Y, Koana T, Fukuda M, Yano K
& Takagi M (1993) Analysis of gene structure for biphenyl and
polychlorinated biphenys (PCBs) degradation in Pseudomonas
sp. KKS102 (Abstract). The Fourth International Symposium on
Pseudomonas: Biotechnology and Molecular Biology, Vancou-
ver, August 1993
Kimbara K, Hashimoto T, Fukuda M, Koana T, Takagi M, Oishi M
& Yano K (1989) Cloning and sequencing of two tandem genes
involved in degradation o f 2,3-dihydroxybiphenolto benzoic acid
in the polychlorinated biphenyl-degrading soil bacterium Pseu-
domonas sp. strain KKSI02. J. Bacteriol. 171:2740-2747
Kukor JJ & Olsen RH (1991) Genetic organization and regulation of a
meta cleavage pathway for catechols produced from catabolism o f
toluene, benzene, phenol, and cresols by Pseudomonas pickettii
PKO1. J. Bacteriol. 173:4587-4594
Kukor JJ & Olsen RH (1992) Complete nucleotide sequence of thuD,
the gene encoding phenol/cresol hydroxylase from Pseudomonas.
pickettii PKO1, and functional analysis of the encoded enzymes.
J. Bacteriol. 174:6518-6526
Kumagai H, Yamada H, Matsui H, Ohkishi H & Ogata H (1970)
Tyrosine phenol lyase I. Purification, crystalization, and proper-
ties. J. Biol. Chem. 245:1767-1777
Lau PCK, Wang Y, Rawlings M, Labb6 D, Bergeron H, Brousseau
R & Gibson DT (1993) Regulation of toluene degradation in
Pseudomonas putida F1 by an unusual sensor-regulator pro-
tein (Abstract) The Fourth International Symposium on Pseu-
domonas: Biotechnology and Molecular Biology, Vancouver,
August 1993
Lister J (1867) On a new method of treating compound fracture,
abscess, etc.: with observations on the conditions of suppuration.
Lancet (March 16):326-329
Loomes KM, Midwinter GG, Blackwell LF & Buckley PD
(1990) Evidence for reactivity of serine-74 with trans-(N, N-
dimethylamino)cinnamaldehyde during oxidation by the cyto-
plasmic aldehyde dehydrogenase from sheep liver, Biochemistry
29:2070-2075
Marcotte P & Walsh C (1978) Sequence of reactions which follows
enzymatic oxidation of allylglycine. Biochemistry 17:5620-5626
Mason JR & Cammack R (1992) The electron-transport proteins
of hydroxylating bacterial dioxygenases. Annu. Rev. Microbiol.
46:277-305
McGrew RE (1985) Encyclopedia of Medical History (pp 20-24).
MacMillan Reference Books, MacMillan Press, London
Menn F-M, Zylslm GJ & Gibson I3T (1991) Location and sequence
of the todF gene encoding 2-hydroxy-6-oxohepta-2,4-dienoate
hydrolase in Pseudomonasputida F1. Gene 104:91-94
Nakai C, Kagamiyama H & Nozaki M (1983) Complete nucleotide
sequence of the metapyrocatachase gene of the TOL plasmid of
Pseudomonas putida rot-2. J. Biol. Chem. 258:2923-2928
Neujahr HY &Gaal A (1973) Phenol hydroxylase from yeast, Purifi-
cation and properties of the enzyme from Trichosporon cutaneum.
Eur. J. Biochem. 35:386-400
Ng LC, Sze CC & Poh CL (1993) Cloning, nucleotide sequence
and expression of chromosomally-encoded phenol hydroxylase
degradative genes in Pseudomonasputida P35X (Abstract) The
235
Fourth International Symposium on Pseudomonas: Biotechnolo-
gy and Molecular Biology, Vancouver, August 1993
Nordlund I & Shingler V (1990) Nucleotide sequence of the meta-
cleavage pathway enzymes 2-hydroxymuconic semialdehyde
dehydrogenase and 2-hydroxymuconic semialdehyde hydrolase
from Pseudomonas CF600. Biochim. Biophys, Acta 1049:227-
230
Nordlund I, Powlowski J, Hagstrfm A & Shingler V (1993) Conser-
vation of regulatory and structural genes for a multicomponent
phenol hydroxylase within phenol-catabolizing bacteria that uti-
lize a meta-cleavage pathway. J, Gen. Mierobiol, 139:2695-2703
Nordlund I, Powtowski J & Shingler V (1990a) Complete nucleotide
sequence and polypeptide analysis of multicomponent phenol
hydroxylase from Pseudomonas sp. strain CF600. J. Bacteriol.
172:6826-6833
Nordlund E SjOberg B-M & Eklund H (1990b) Three-dimensional
structure of the free radical protein of ribonucleotide reductase.
Nature 345:593-598
Nordlund P, Dalton H & Eklund H (1992) The active site structure of
methane monooxygenase is closely related to the binuclear iron
centre of ribonucleotide reductase. FEBS Lett. 307:257-262
North AK, Klose KE, Stedman KM & Kustu S (1993) Prokayotic
enhancer-binding proteins reflect eukaryote-like modularity: the
puzzle of nitrogen regulatory protein C. J, Bacteriol. 175:4267-
4273
Nurk A, Kasak L & Kivisaar M (1991) Sequence of the gene (pheA)
encoding phenol monooxygenase from Pseudomonas EST100t:
expression in Escherichia coli and Pseudomonas putida. Gene
102:13-18
Nunn WP (1987) Two-carbon compounds and fatty acids as carbon
sources. In: Neidhardt FC, lngraham JL, Low KB, Maganasik
B, Schaechter M & Umbarger HE (Eds) Escherichia Coli and
Salmonella Typhimurium: Cellular and Molecular Biology (pp
285-301). American Society for Microbiology, Washington, D.C.
Omston LN & Yeh W-K (1982) Recurring themes and repeated
sequences in metabolic evolution. In: Chakrabarty AM (Ed)
Biodegradation and Detoxification of Environmental Pollutants
(pp 105-126). CRC Press, BocaRaton, FL
Pathak D & Oltis D (1990) The refined structure of dienelactone
hydrolase at 1.8 ,~. J. Mol. Biol. 214:497-525
Polissi A & Harayama S (1993) In vivo reactivation of catechol
2,3-dioxygenase mediated by a chloroplast-type ferredoxin: a
bacterial strategy to expand the substrate specificity of aromatic
degradativepathways. EMBO L 12:3339-3347
Powlowski J, Sahtman L & Shingler V (1993) Purification and
properties of the physically associated meta-cleavage pathway
enzymes 4-hydroxy-2-ketovalerate aldolase and aldehyde dehy-
drogenase (acylating) from Pseudomonas sp. strain CF600. J.
Bacteriol. 175:377-385
Powlowski J & Shingler V (1990) In vitro analysis of polypeptide
requirements of multicomponent phenol hydroxylase from Pseu-
domonas sp. strain CF600. J. Bacteriol. 172:6834--6840
Ramos JL, Stolz, A Reineke W & Timmis KN (1986) Altered effec-
for specificities in regulators of gene expression: TOL plasmid
x3'lS mutants and their use to engineer expansion of the range
of aromatics degraded by bacteria. Proc. Natk Acad. Sci. U.S.A.
83:8467-8471
Ramos JL, Michan C, Rojo E DwyerD & Timmis KN (1990)Signal-
regulator interactions. Genetic analysis of the effector binding
site of xylS, the benzoate-activated positive regulator of Pseu-
domonas TOL plasmid meta-cleavage pathway operon. J. Mol.
Biol. 211:373-382
236
Rheinwald JG, Chakrabarty AM & Gunsalus IC (1973) A trans-
missible plasmid controlling camphor oxidation in Pseudomonas
putida. Proc. Natl. Acad. Sci. U.S.A. 70:885-889
Rosenzweig AC, Fredrick CA, Lippard SJ & Nordlund P (t993)
Crystal structure of a bacterial non-haem iron hydroxylase that
catalyses the biological oxidation of methane. Nature 366:537-
543
Sala-Trepat JM, Murray K & Williams PA (1972) The metabolic
divergence in the meta cleavage of catechols by Pseudomonas
putida. Eur. J. Biochem. 28:347-356
Sala-Trepat JM & Evans WC (1971) The meta-cleavage of cate-
chol by Azotobacter species: 4-oxalocrotonate pathway. Eur. J.
Bioehem. 20:400-413
Shingler V, Bartilson M & Moore T (1993) Cloning and nucleotide
sequence of the gene encoding the positive regulator (DmpR) of
the phenol catabolic pathway encoded by pVI150 and identifica-
tion of DmpR as a member of the NtrC family of transcriptional
activators. J. Bacteriol. 175:1596-1604
Shingler V, Franklin FCH, Tsuda M, Holroyd D & Bagdasarian M
(1989) Molecular analysis of a plasmid-encoded phenol hydrox-
ylase from Pseudomonas CF600. J. Gen. Microbiol, 135:1083-
1092
Shingler V & Moore T (1994) Sensing of aromatic compounds by
the transcriptional activator
ofphenol-catabolisingPseudomonas
sp. strain CF600. J. Bacteriol. 176:1555-1560
Shingler V, Powlowski J & Marklund U (1992) Nucleotide
sequence and functional analysis of the complete phenoI/3,4-
dimethylphenol catabolic pathway of Pseudomonas sp. strain
CF600. J. Bacteriol. 174:711-724
Spain JC & Gibson DT (1988) Oxidation of substituted phenols by
Pseudomonasputida F1 and Pseudomonas sp. strain JS6. Appl.
Environ. Microbiol. 54:1399-1404
Spoelstra SF (1977) Degradation of tyrosine in anaerobically stored
piggery wastes and pig faeces, Appl. Environ. Microbiol. 36:631-
638
Stainthorpe AC, Lees V, Salmond GPC, Dalton H & Murrell JC
(1990) The methane monooxygenase gene cluster of Methyto-
coccus capsulatus (Bath). Gene 91:27-34
Stainthorpe AC, Murrell JC, Salmond GPC, Dalton H & Lees
V (1989) Molecular analysis of methane monooxygenase from
Methylococcus capsulatus (Bath). Arch. MicrobioL 152:154-159
Steele MI, Lorenz D, Hatter K, Park A & Sokatch (1992) Character-
ization of the mmsAB operon of Pseudomonas aeruginosa PAO
encoding methylmalonate-semialdehyde dehydrogenase and 3-
hydroxyisobutyrate dehydrogenase. J. Biol. Chem. 267:13585-
13592
Walsh C (1979) Enzymatic Reaction Mechanisms (pp 745-759).
Freeman and Company, San Francisco
Wasserfallen A, Rekik M & Harayama S (1991) A Pseu-
domonas strain able to degrade m-toluate in the presence of 3-
chlorocatechol. Bio/Technol. 9:296-298
Whitman CP, Aird BA, Gillespie WR & Stolowich NJ ( 1991 ) Chem-
ical and enzymatic ketonization of 2-hydroxymuconate, a conju-
gated enol. J. Amer. Chem. Soc. 113:3154-3162
Wierenga RK, Terpstra P & Hol WGJ (1986) Prediction of the occur-
rence of the ADP-binding/3-c~-fl fold in proteins using an amino
acid sequence finger print. J. Mot. Biol. I87:101-107
Wigmore G J, Bayly RC & Berardino D (1974) Pseudomonasputida
mutants defective in metabolism of the products of meta-fisson
of cateehol and its methyl analogues. J. Bacteriol. 120:31-37
Williams PA, Assinder SJ & Shaw LE (1990) Construction of hybrid
xylE genes between the two duplicate homologous genes from
TOL plasmid pWW53: comparison of the kinetic properties of
the gene product. J. Gen. Mierobiol. 136:1583-1589
Williams PA & Sayers JR (1994) The evolution of pathways for
aromatic hydrocarbon oxidation in Pseudomonas. This Volume
Yamamoto S & tshimura Y (1991) Dioxygenases and monooxyge-
nases. In: Kuby J (Ed) A Study of Enzymes (pp 315-344). CRC
Press, Boca Raton, FL
Yen K-M & Karl MR (1992) Identification of a new gene, tmoF, in
the
Pseudomonas mendocina KR 1 gene cluster encoding toluene-
4-monooxygena,se. J. Bacteriol. 174:7253-7261
Yen K-M, Karl MR, Blatt LM, Simon MJ, Winter RB, Fansset PR,
Lu HS, Harcourt AA & Chen KK (1991) Cloning and characteri-
zation of a Pseudomonas mendocina KR 1 gene cluster encoding
toluene-4-monooxygenase. J. Bacteriol. 173:5315-5327
You l-S, Ghosal D & Gunsahs IC (I 991) Nucleotide sequence analy-
sis of the Pseudomonasputida PpG7 salicylate hydroxylase gene
(nahG) and its 3t-flanking region. Biochemistry 30:1635-I641
Zylstra GJ & Gibson DT (1989) Toluene degradation by Pseu-
domonas Ft: nueleotide sequence of the todC1C2BADE genes
and their expression in Escherichia coli. J. Biol. Chem.
264:14940-14946
... The former was sourced from the Pseudomonas sp. strain CF600 (Powlowski and Shingler 1994), while the latter was from Rhodococcus erythropolis CCM2595 (Zídková et al. 2013). As shown in Table 2, for each query, at least one hit was found. ...
... Interestingly, there was a consistent distribution of dmp gene-harboring microorganisms from terrestrial environments, such as Pseudomonas sp. (Powlowski and Shingler 1994;Sazinsky et al. 2006) and Acinetobacter sp. (Doukyu et al. 2003), which also predominantly belong to Gammaproteobacteria, specifically within the orders Pseudomonadales and Burkholderiales. ...
Article
Full-text available
Phenolic compounds, as well as other aromatic compounds, have been reported to be abundant in hadal trenches. Although high-throughput sequencing studies have hinted at the potential of hadal microbes to degrade these compounds, direct microbiological, genetic and biochemical evidence under in situ pressures remain absent. Here, a microbial consortium and a pure culture of Pseudomonas, newly isolated from Mariana Trench sediments, efficiently degraded phenol under pressures up to 70 and 60 MPa, respectively, with concomitant increase in biomass. By analyzing a high-pressure (70 MPa) culture metatranscriptome, not only was the entire range of metabolic processes under high pressure generated, but also genes encoding complete phenol degradation via ortho- and meta-cleavage pathways were revealed. The isolate of Pseudomonas also contained genes encoding the complete degradation pathway. Six transcribed genes (dmpKLMNOPsed) were functionally identified to encode a multicomponent hydroxylase catalyzing the hydroxylation of phenol and its methylated derivatives by heterogeneous expression. In addition, key catabolic genes identified in the metatranscriptome of the high-pressure cultures and genomes of bacterial isolates were found to be all widely distributed in 22 published hadal microbial metagenomes. At microbiological, genetic, bioinformatics, and biochemical levels, this study found that microorganisms widely found in hadal trenches were able to effectively drive phenolic compound degradation under high hydrostatic pressures. This information will bridge a knowledge gap concerning the microbial aromatics degradation within hadal trenches. Supplementary Information The online version contains supplementary material available at 10.1007/s42995-024-00224-2.
... Biological treatment has emerged as an effective approach that has attracted significant interest due to its gentle nature, costeffectiveness, efficacy, and environmental friendliness (Kobayashi et al. 2012;Panigrahy et al. 2022). Over the past few decades, various microorganisms, such as Pseudomonas, Candida, Rhodococcus, Bacillus, Acinetobacter and Oceanobacillus, have been shown to degrade phenol (Banerjee and Ghoshal 2010;He et al. 2022;Lee et al. 2020;Long et al. 2015;Powlowski and Shingler 1994;Xu et al. 2003). Many studies have focused on the microbial metabolism of phenol, and some studies have attempted to use microorganisms for the bioremediation of phenol-contaminated soils. ...
Article
Full-text available
Phenols are highly toxic chemicals that are extensively used in industry and produce large amounts of emissions. Notably, phenols released into the soil are highly persistent, causing long-term harm to human health and the environment. In this study, a gram-positive, aerobic, and rod-shaped bacterial strain, Z13T, with efficient phenol degradation ability, was isolated from the soil of sugarcane fields. Based on the physiological properties and genomic features, strain Z13T is considered as a novel species of the genus Rhodococcus, for which the name Rhodococcus sacchari sp. nov. is proposed. The type strain is Z13T (= CCTCC AB 2022327T = JCM 35797T). This strain can use phenol as its sole carbon source. Z13T was able to completely degrade 1200 mg/L phenol within 20 h; the maximum specific growth rate was μmax = 0.93174 h⁻¹, and the maximum specific degradation rate was qmax = 0.47405 h⁻¹. Based on whole-genome sequencing and Kyoto Encyclopedia of Genes and Genomes (KEGG) analysis, strain Z13T contains a series of phenol degradation genes, including dmpP, CatA, dmpB, pcaG, and pcaH, and can metabolize aromatic compounds. Moreover, the potential of strain Z13T for soil remediation was investigated by introducing Z13T into simulated phenol-contaminated soil, and the soil microbial diversity was analyzed. The results showed that 100% of the phenol in the soil was removed within 7.5 d. Furthermore, microbial diversity analysis revealed an increase in the relative species richness of Oceanobacillus, Chungangia, and Bacillus. Graphical abstract
... In particular, Novosphingobium sp. and Pseudomonas sp. were abundant, showing an average contribution of over 10% to the bacterial community composition, suggesting their importance in leaf carbon cycling in lentic freshwaters. Pseudomonas sp. is commonly found during the early-stage decomposition of leaves (Purahong et al., 2016;Tl askal et al., 2016), and it can utilize various polymers as its carbon source, for example, carbohydrates (Udaondo et al., 2018), phenols (Powlowski & Shingler, 1994), and even highly recalcitrant plastic polymers (Wilkes & Aristilde, 2017). The relative abundance of Gammaproteobacteria decreased in all treatments during the succession, as found also by other studies (Newman et al., 2015;Zhao et al., 2017). ...
Article
Full-text available
Terrestrial organic matter subsidizes aquatic food webs and plays an important role in carbon cycling in lake ecosystems, where it is decomposed mainly by microbes. However, the contribution of terrestrial and aquatic microbiomes on terrestrial carbon cycling and their effects on the biochemical fate of carbon has remained understudied. Therefore, we explored the microbial carbon utilization of three chemically differing leaf species in lake water in microcosms and quantified the biochemical endpoints of leaf carbon in CO2, CH4, and microbial biomass. Additionally, we identified microbial taxa responsible for leaf carbon recycling and studied the role of epiphytic and endophytic leaf microbiomes in microbial community succession in lake water. Microbially utilized leaf carbon was mainly respired (82.7 ± 1.4%), whereas a small proportion (17.1 ± 1.4%) was assimilated into biomass. Carbon from nitrogen‐rich alder leaves was taken up at the fastest rate, whereas birch leaf addition produced the highest concentrations of CH4, suggesting that leaf chemistry affects the decomposition rate and biochemical fate of carbon. In particular, terrestrial bacteria shaped the succession of aquatic bacterial communities. The addition of leaves resulted in the equal contribution of epiphytic and endophytic bacteria in the lake water, whereas epiphytic fungi dominated the fungal community structure. Our results suggest that terrestrial bacteria originating from terrestrial leaves influence the microbiome succession in lake ecosystems and play a key role in linking terrestrial carbon to an aquatic food web and determining the quality of carbon emissions that are released into the atmosphere.
... Other enzymes in the cleavage pathway of catechol include aldehyde dehydrogenase, which uses the ring cleavage products from catechol, 4-methyl-2-hydroxymuconic semialdehyde, and 3-methylcatechol as substrates. These products include 2-hydroxymuconic semialdehyde and 2-hydroxy-6-oxo2,4-heptadienoate, which have been shown to be preferred substrates for the dehydrogenase 28 . Fig. 3B shows a promoter region in the Ph2 DNA sequence. ...
Article
Full-text available
The objective of this study was to identify genes associated with the biodegradation of phenol by Acinetobacter sp. strain DF4 through the use of differential display (DD) methodology. The bacteria were grown in YEPG medium, and total RNA was extracted and analyzed using labeled primers to detect gene expression differences. Three distinctively expressed cDNA bands (ph1, ph2, and ph3) were identified, cloned, and sequenced. DNA analysis involved searching for open reading frames (ORFs), verifying results with the NCBI database, predicting promoter regions, and constructing phylogenetic trees using bioinformatics tools. The ph1 gene displayed a 97% identity with the AraC transcriptional regulator, suggesting its potential role in regulating the ortho-catabolic pathway of phenol. The ph2 gene showed a 98% identity with aspartate semialdehyde dehydrogenase, which is involved in phenol degradation. The ph3 gene had a 93% identity with acetyltransferase. Essential transcription factors, such as TATA, GTGTGT, CACA, and CTTTT, were detected, and the three genes promoter regions were predicted. This study successfully identified functional genes involved in the metabolism of cyclic chemicals, particularly phenol, using the DD technique. These findings provide insights into the biodegradation pathways of phenol by Acinetobacter sp. Strain DF4 and may contribute to the development of more efficient bioremediation strategies for phenol-contaminated environments.
... Of the two enzymes assayed in this study, aldehyde oxidase is reported to metabolise catechol into 2-hydroxymuconate semialdehyde and then to cis,cis-2-hydroxy muconic acid, via meta cleavage [21]. The end metabolite of cis,cis-2-hydroxy muconic acid is acetaldehyde or pyruvate, the latter two being glycolytic intermediate metabolite to tricarboxylic acid cycle for energy production which contribute to cell replication and growth [22]. The other enzyme oxygen oxidoreductase is reported to metabolise CBZ into carbamazepine-2,3-epoxide [23], which is hydroxylated to produce non-harmful derivatives [24]. ...
... A common process for metabolizing aromatic substances like phenol is the dihydroxylation of the aromatic ring to form derivative of catechol, followed by ortho-or meta-oxidation. The enzymes catechol 2,3-dioxygenase, and catechol 1,2-dioxygenase, convert catechol derivatives to -ketoadipate or semialdehyde either the ortho-cleavage route or the meta pathway (12). ...
Article
Full-text available
Biological treatments for olive mill waste cleanup using bacterial strains have proven to be effective and environmentally friendly. Three phenol-degrading bacterial strains identified as Lysinibacillus macroides, Lysinibacillus boronitolerans, and Brevundimonas olei were isolated from olive mill wastewater (OMWW) have high efficacy in degrading polyphenols. These strains exhibited both enzymes, catechol 1,2-dioxygenase and catechol 2,3-dioxygenase, responsible for degradation of phenolic compounds to varying degrees. Lysinibacillus macroides, rather than Lysinibacillus boronitolerans or Brevundimonas olei, have the highest catechol-dioxygenase enzyme activities. Various factors were optimized during the OMWW fermentation process to maximize the phenol degradation process. Maximum degradation rates were reached after 35 days at 20% dilution and pH ranged from 6 to 7 with starch and yeast reaching 1 g/l when added as carbon and nitrogen sources. In greenhouse experiments, the final fermentation product was evaluated as an organic biofertilizer. Bioassay data showed that the application of fermented OMWW significantly increases barley height, fresh and dry weight, nitrogen, potassium, and phosphorus contents compared to non-fermented barley, regardless of the type of bacteria used. The results of this study demonstrate that biologically fermented OMWW is proved to be economically beneficial to be used as an organic biofertilizer.
... Aldehyde oxidase has been reported to metabolise catechol (carbamazepine central metabolite) into acetaldehyde or pyruvate, via meta cleavage [20]. Despite both substrates can enter the glycolytic and tricarboxylic acid cycle to produce energy that contributes to cell replication and growth [21], it is possible that the mixed culture favours other metabolites to support growth. A study on the growth of similar species Rodococcus equi demonstrated that this isolate prefers to metabolise organic acids such as acetate, lactate, succinate, malate, and fumarate instead of pyruvate [22], the latter being the product of oxidoreductase concentration and carbamazepine degradation and bacterial growth suggests that this enzyme was expressed as a strategy to obtain carbon from carbamazepine to increase bacterial biomass. ...
Article
Full-text available
A mixed culture of Leifsonia shinshuensis and Rhodococcus zopfii was assessed on its ability to degrade carbamazepine (CBZ). The mixed culture was found to metabolise CBZ by utilizing two enzymes, oxygen oxidoreductase and aldehyde oxidase. Correlation analysis showed that the degraded CBZ was strongly correlated (R 2 ≈0.9) with bacterial culture growth. However, only the concentrations of oxygen oxidoreductase is strongly correlated (R 2 ≈0.9) with both degradation of CBZ and bacterial growth, whereas the aldehyde oxidase concentration was less strongly correlated (R 2 ≈0.8) with both degradation of CBZ and bacterial growth. This suggests that the mixed culture adopted two different strategies simultaneously to metabolise CBZ to support growth and reduce toxicity.
Article
Full-text available
The coir retting industries in Southern regions of Cherthala provides a combative environment for the multiplication of specialized microbes with high phenol biodegradation potential. The present study aimed at isolation and characterization of phenol degrading bacteria from coir retting contaminated soils. After a series of screening enrichment techniques in MSPM at 37ºC, two bacteria were found to have the highest efficiency of pronounced growth in phenol stress conditions. Also, these bacteria have shown to be able to remove 22.5μM of phenol concentration that was measured by modified Mordaque et al, 1999 protocol at 24, 48 and 72 hours of incubation with a regular interval of 24 hours . The strong phenol degrading bacterial isolates were identified and the sequences were deposited in genbank as Lysini bacillus sp strain GAKVM (MT476862.1) and Paenibacillus sp strain. GA2KVM (MT509865.1). In the concentrations of 22.5μM of phenol, MT476862.1was able to degrade 74.44% and MT509865.1were able to degrade 74.77% within 72 hours of incubation at 37ºC.
Article
This research reports on the efficiency of Rhodococcus zopfii and Leifsonia shinshuensis to degrade approximately 99.98 ± 0.20% carbamazepine (CBZ) within 5 days. The degradation of CBZ was strongly correlated (R2 ≥ 0.9) to bacterial growth in L. shinshuensis, but only moderate (R2 ≈ 0.6) in R. zopfii. Both oxygen oxidoreductase and aldehyde oxidase concentrations were positively and strongly correlated (R2 ≈ 0.9) to the degradation rate of CBZ in both cultures. However, when it comes to bacterial growth (CFU mL−1/h), only the culture of L. shinshuensis showed a strong positive correlation (R2 ≈ 0.9) to both enzyme concentrations, whereas the culture of R. zopfii only shows a moderate correlation (R2 ≈ 0.7). The degradation of CBZ by R. zopfii did not directly translate into higher bacterial growth. This demonstrated the possibility of different CBZ degradation strategies was being adopted by these two different bacteria. Hence, this piece of information is important to elucidate the mechanisms of carbamazepine degradation and sustainable treatment.
Article
Full-text available
An understanding of the mechanism of biological methane oxidation has been hampered by the lack of purified proteins. We describe here a purification protocol for the previously uncharacterized protein B of the soluble methane monooxygenase from the obligate methanotroph Methylococcus capsulatus (Bath). Soluble methane monooxygenase is a multicomponent enzyme consisting of a hydroxylase component, protein A, a reductase component, protein C, and protein B. All three proteins are required for monooxygenase activity. Protein B proves to be a low molecular weight (16,000) single subunit protein devoid of prosthetic groups. The protein is a powerful regulator of soluble methane monooxygenase activity, possessing the capacity to convert the enzyme from an oxidase to an oxygenase. Proteins A and C together catalyze the reduction of molecular oxygen to water, a reaction prevented by protein B. The uncoupling of soluble methane monooxygenase in this manner displays a number of novel features. First, the product of the uncoupled reaction is water, and second, the uncoupling is independent of substrate. Free hydrogen peroxide is not an intermediate in the reduction of oxygen by the incomplete methane monooxygenase enzyme complex. Finally, electron transfer can occur between protein C and protein A in the absence of protein B and protein B prevents the steady-state transfer of electrons in the absence of an oxidizable substrate, such as methane. It is demonstrated that oxygen reduction occurs at the active site of the hydroxylase component, protein A. A unifying mechanism, describing the interaction of the three proteins of soluble methane monooxygenase, is proposed.
Chapter
Full-text available
Article
Full-text available
The nucleotide sequence of the todC1C2BADE genes which encode the first three enzymes in the catabolism of toluene by Pseudomonas putida F1 was determined. The genes encode the three components of the toluene dioxygenase enzyme system: reductaseTOL (todA), ferredoxinTOL (todB), and the two subunits of the terminal dioxygenase (todC1C2); (+)-cis-(1S, 2R)-dihydroxy-3-methylcyclohexa-3,5-diene dehydrogenase (todD); and 3-methylcatechol 2,3-dioxygenase (todE). Knowledge of the nucleotide sequence of the tod genes was used to construct clones of Escherichia coli JM109 that overproduce toluene dioxygenase (JM109(pDT-601]; toluene dioxygenase and (+)-cis-(1S, 2R)-dihydroxy-3-methylcyclohexa-3,5-diene dehydrogenase (JM109(pDTG602]; and toluene dioxygenase, (+)-cis-(1S, 2R)-dihydroxy-3-methylcyclohexa-3,5-diene dehydrogenase, and 3-methylcatechol 2,3-dioxygenase (JM109(pDTG603]. The overexpression of the tod-C1C2BADE gene products was detected by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The three E. coli JM109 strains harboring the plasmids pDTG601, pDTG602, and pDTG603, after induction with isopropyl-β-D-thiogalactopyranoside, oxidized toluene to (+)-cis-(1S, 2R)-dihydroxy-3-methylcyclohexa-3,5-diene, 3-methylcatechol, and 2-hydroxy-6-oxo-2,4-heptadienoate, respectively. The tod-C1C2BAD genes show significant homology to the reported nucleotide sequence for benzene dioxygenase and cis-1,2-dihydroxycyclohexa-3,5-diene dehydrogenase from P. putida 136R-3 (Irie, S., Doi, S., Yorifuji, T., Takagi, M., and Yano, K. (1987) J. Bacteriol. 169, 5174–5179). In addition, significant homology was observed between the nucleotide sequences for the todDE genes and the sequences reported for cis-1,2-dihydroxy-6-phenylcyclohexa-3,5-diene dehydrogenase and 2,3-dihydroxybiphenyl-1,2-dioxygenase from Pseudomonas pseudoalcaligenes KF707 (Furukawa, K., Arimura, N., and Miyazaki, T. (1987) J. Bacteriol. 169, 427–429).
Article
The TOL plasmid upper pathway operon encodes enzymes involved in the catabolism of aromatic hydrocarbons such as toluene and xylenes. The regulator of the gene pathway, the XylR protein, exhibits a very broad effector specificity, being able to recognize as effectors not only pathway substrates but also a wide variety of mono- and disubstituted methyl-, ethyl-, and chlorotoluenes, benzyl alcohols, and p-chlorobenzaldehyde. Benzyl alcohol dehydrogenase and benzaldehyde dehydrogenase, two upper pathway enzymes, exhibit very broad substrate specificities and transform unsubstituted substrates and m- and p-methyl-, m- and p-ethyl-, and m- and p-chloro-substituted benzyl alcohols and benzaldehydes, respectively, at a high rate. In contrast, toluene oxidase only oxidizes toluene, m- and p-xylene, m-ethyltoluene, and 1,2,4-trimethylbenzene [corrected], also at a high rate. A biological test showed that toluene oxidase attacks m- and p-chlorotoluene, albeit at a low rate. No evidence for the transformation of p-ethyltoluene by toluene oxidase has been found. Hence, toluene oxidase acts as the bottleneck step for the catabolism of p-ethyl- and m- and p-chlorotoluene through the TOL upper pathway. A mutant toluene oxidase able to transform p-ethyltoluene was isolated, and a mutant strain capable of fully degrading p-ethyltoluene was constructed with a modified TOL plasmid meta-cleavage pathway able to mineralize p-ethylbenzoate. By transfer of a TOL plasmid into Pseudomonas sp. strain B13, a clone able to slowly degrade m-chlorotoluene was also obtained.
Article
Catechol was metabolized through 2‐hydroxymuconic semialdehyde by cell‐free extracts of benzoate‐grown Azotobacter Strains. Some properties of catechol 2,3 oxygenase preparations from Azotobacter vinelandii 206 are described. Two different enzymatic activities able to attack 2‐hydroxymuconic semialdehyde have been found in crude extracts from benzoate‐grown cells; one catalyses a hydrolytic release of formate from the semialdehyde and the other a dehydrogenation of this compound to 4‐oxalocrotonate. However, the low, non‐inducible levels of 2‐hydroxymuconic semialdehyde hydrolase activity appear negligible for metabolic purposes and the semialdehyde seems to be dissimilated almost exclusively via 4‐oxalocrotonate, by the action of a NAD ⁺ ‐dependent dehydrogenase, in Azotobacter strains. A tautomerase activity responsible for the interconversion of the enol and keto forms of 4‐oxalocrotonic acid was found in extracts from benzoate‐grown cells. 4‐Oxalocrotonate was stoicheiometrically converted to CO 2 and 4‐hydroxy‐2‐oxovalerate by a partially purified extract, with the transient formation of a compound that appears to be 2‐oxopent‐4‐enoic acid. The 4‐oxalocrotonate decarboxylase activity was stimulated by Mg ²⁺ or Mn ²⁺ ions and was inhibited by EDTA. Cell‐free extracts from Azotobacter strains converted synthetic 4‐hydroxy‐2‐oxovalerate to acetaldehyde and pyruvate. A reaction sequence, termed the 4‐oxalocrotonate pathway, for the dissimiation of catechol to acetaldehyde and pyruvate by Azotobacter species is presented. All the enzymes operative in this pathway were inducible, except the 4‐hydroxy‐2‐oxovalerate aldolase. The findings described here are discussed in connection with the two previously reported meta cleavage pathways for the oxidation of catechol in Pseudomonas strains.