Characterization of Cupriavidus Metallidurans CYP116B1 - A Thiocarbamate Herbicide Oxygenating P450-Phthalate Dioxygenase Reductase Fusion Protein

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Characterization of Cupriavidus metallidurans CYP116B1

A thiocarbamate herbicide oxygenating P450phthalate


dioxygenase reductase fusion protein
Ashley J. Warman1, Jacob W. Robinson2, Dominika Luciakova2, Andrew D. Lawrence3,
Ker R. Marshall1, Martin J. Warren3, Myles R. Cheesman4, Stephen E. J. Rigby2,
Andrew W. Munro2 and Kirsty J. McLean2
1 Department of Biochemistry, University of Leicester, Leicester, UK
2 Faculty of Life Sciences, Manchester Interdisciplinary Biocentre, University of Manchester, Manchester, UK
3 Department of Biosciences, University of Kent, Canterbury, Kent, UK
4 School of Chemical Sciences, University of East Anglia, Norwich, UK

Keywords The novel cytochrome P450 redox partner fusion enzyme CYP116B1 from
CYP116B1 cytochrome P450 (116B1) from Cupriavidus metallidurans was expressed in and puried from Escherichia coli.
Cupriavidus metallidurans; fusion enzyme;
Isolated CYP116B1 exhibited a characteristic Fe(II)CO complex with Soret
phthalate dioxygenase reductase;
maximum at 449 nm. EPR and resonance Raman analyses indicated low-
potentiometry
spin, cysteinate-coordinated ferric haem iron at both 10 K and ambient
Correspondence temperature, respectively, for oxidized CYP116B1. The EPR of reduced
Kirsty J. McLean or Andrew W. Munro, CYP116B1 demonstrated stoichiometric binding of a 2Fe-2S cluster in the
Faculty of Life Sciences, Manchester reductase domain. FMN binding in the reductase domain was conrmed
Interdisciplinary Biocentre, University of by avin uorescence studies. Steady-state reduction of cytochrome c and
Manchester, 131 Princess Street,
ferricyanide were supported by both NADPH NADH, with NADPH used
Manchester M1 7DN, UK
more efciently (Km[NADPH] = 0.9 0.5 lM and Km[NADH] = 399.1
Fax: 0044 161 306 8918
Tel: 0044 161 306 4194 2715 52.1 lM). Stopped-ow studies of NAD(P)H-dependent electron transfer
E-mail: kirsty.mclean@manchester.ac.uk to the reductase conrmed the preference for NADPH. The reduction
or potential of the P450 haem iron was -301 7 mV, with retention of haem
Tel: 0044 161 306 5151 thiolate ligation in the ferrous enzyme. Redox potentials for the 2Fe-2S
Fax: 0044 161 306 8918 and FMN cofactors were more positive than that of the haem iron. Multi-
E-mail: andrew.munro@manchester.ac.uk
angle laser light scattering demonstrated CYP116B1 to be monomeric.
Type I (substrate-like) binding of selected unsaturated fatty acids (myristo-
(Received 14 December 2011, revised 10
February 2012, accepted 17 February 2012) leic, palmitoleic and arachidonic acids) was shown, but these substrates
were not oxidized by CYP116B1. However, CYP116B1 catalysed hydroxyl-
doi:10.1111/j.1742-4658.2012.08543.x ation (on propyl chains) of the herbicides S-ethyl dipropylthiocarbamate
(EPTC) and S-propyl dipropylthiocarbamate (vernolate), and the subse-
quent N-dealkylation of vernolate. CYP116B1 thus has similar thiocarba-
mate-oxidizing catalytic properties to Rhodoccocus erythropolis CYP116A1,
a P450 involved in the oxidative degradation of EPTC.

Abbreviations
CO, carbon monoxide; CPR, cytochrome P450 reductase; CYP101A1, Pseudomonas putida camphor hydroxylase P450cam or CYP102A1;
CYP116B1, cytochrome P450 (116B1) from Cupriavidus metallidurans CH34; ddH2O, distilled, deionized water; EPTC, S-ethyl
dipropylthiocarbamate; Fe-S, iron-sulfur; HS, high-spin; ImC10, imidazolyl decanoic acid; ImC11, imidazolyl undecanoic acid; ImC12,
imidazolyl dodecanoic acid; IPTG, isopropyl thio-b-D-galactoside; LC, liquid chromatography; MALLS, multi-angle laser light scattering; NHE,
normal hydrogen electrode; NO, nitric oxide; P450, cytochrome P450 or CYP; P450 BM3, cytochrome P450 BM3 from Bacillus megaterium
or CYP102A1; PDO, phthalate dioxygenase; PDOR, Burkholderia cepacia phthalate dioxygenase reductase; SEC, size-exclusion
chromatography; vernolate, S-propyl dipropylthiocarbamate.

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1675
Biochemical analysis of CYP116B1 A. J. Warman et al.

Introduction
The cytochromes P450 (P450s) are a superfamily of putidaredoxin [9,10]. The second system (class 2) is
cysteine thiolate-coordinated haem b-binding enzymes exemplied by the mammalian hepatic P450 enzymes
widespread in nature [1]. The vast majority of P450s (that are pivotal to drug and xenobiotic metabolism)
catalyse the reductive scission of molecular oxygen and involves membrane-bound P450s and a diavin
bound to the haem iron, resulting in monooxygenation reductase [cytochrome P450 reductase (CPR)] that
of a substrate molecule bound close to the haem [2]. A sources its electrons from NADPH [11]. However,
highly reactive iron-oxo species (compound I) is con- more recent studies of diverse P450 systems, particu-
sidered to be the major oxidant in P450-mediated sub- larly from microorganisms, have demonstrated that the
strate oxidations. Only recently has the rst compelling repertoire of P450 redox partners is much broader
evidence been presented for the formation and the than was rst considered [12] and that certain P450s
reactivity of compound I. Using the P450 CYP119 have disposed altogether with redox partner proteins
from the thermophilic bacterium Sulfolobus acidocalda- and interact directly with NADH or hydrogen perox-
rius, Rittle and Green demonstrated (using EPR, ide to facilitate catalysis [13,14]. Also, other P450s
Mossbauer and UV-visible absorption spectroscopy) exist as natural fusions to both CPR-like and other
characteristic spectra near-identical to those observed forms of redox partners [12,15].
for compound I species in peroxidase hemoproteins, The rst major class of P450redox partner fusion
and an apparent second-order rate constant for oxida- enzymes characterized were the CYP102A family,
tion of unactivated hydrocarbons of 1.1 107 m)1s)1 headed by the well characterized Bacillus megaterium
for the CYP119 compound I species [3]. However, P450 BM3 (BM3, CYP102A1) [16]. This is a cytoplas-
other iron-oxo species in the P450 catalytic cycle are mic enzyme in which a fatty acid hydroxylase P450
also potential substrate oxidants in selected P450 trans- (N terminal) is fused to an NADPH-dependent CPR
formations. For instance, the ferric hydroperoxo spe- [17]. P450 BM3 has the highest reported rate of sub-
cies (compound 0) that precedes compound I is also a strate oxidation for any P450 oxygenase
potential oxidant (e.g. for epoxidation across carbon ( 17 000 min)1 with arachidonate), facilitated by the
carbon double bonds, or for sulfoxidation), albeit with efcient electron transport system operating both
conicting evidence for its general catalytic relevance within its CPR domain and between the CPR and the
emerging from experimental and computational studies P450 [18,19]. Members of the CYP102A family are
[46]. There is a broad spectrum of substrate specicity widespread in Bacillus and in other bacteria, and a
seen across the P450 enzyme superfamily, and a wide similar type of P450CPR fatty acid hydroxylase
array of oxidative transformations are catalysed (e.g. fusion is also found in lower eukaryotes (e.g. Fusari-
hydroxylation, epoxidation, demethylation, S- and um oxysporum P450foxy, CYP505A1) [20,21]. P450
N-oxidation and C-C bond cleavage) [1,2,6]. Various BM3 has been the subject of intensive studies aimed at
P450 enzymes can also catalyse reductive chemistry, in understanding its structure and mechanism, with recent
addition to isomerizations and dehydrations [e.g. studies demonstrating that the dimeric form of the
2,7,8]. enzyme is catalytically functional in fatty acid oxida-
P450-mediated substrate oxygenation requires the tion, with electron transfer between monomers proba-
delivery of two consecutive electrons to a substrate- bly supporting its catalytic function [2224]. Moreover,
bound, ferric enzyme. These events achieve reduction as a consequence of its catalytic efciency and the
of the iron, rst to the ferrous form (which binds availability of good crystal structure data, P450 BM3
dioxygen) and then to the ferric peroxy state [1,2]. Suc- has been extensively engineered in order to explore its
cessive protonations of the latter intermediate produce structurefunction relationships and to enable oxida-
rst compound 0 and then (with water loss) the tran- tion of novel substrates such as short-chain fatty acids,
sient and highly reactive compound I [3]. Electrons are alkanes and steroids [2528].
typically derived from NAD(P)H and are delivered to A distinctive type of P450redox partner fusion sys-
the P450 by iron-sulfur and or avoproteins. Histori- tem was identied from genome analysis of pathogenic
cally, two major classes of P450 redox partners were Burkholderia species, the heavy metal-tolerant bacte-
thought to exist. The rst (class 1) is typied by the rium Cupriavidus metallidurans CH34 and Rhodococcus
Pseudomonas putida camphor hydroxylase P450cam sp. NCIMB 9784 [29]. In this system, P450s of unde-
(CYP101A1), to which electrons are shuttled from ned structure and substrate selectivity (classied in
NADH through the FAD-binding putidaredoxin the CYP116B P450 family) are fused to a reductase
reductase and the ironsulfur (2Fe-2S) cluster-binding module with an amino-acid sequence resembling that

1676 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

of Burkholderia cepacia phthalate dioxygenase reduc- In this manuscript we present the rst report of the
tase (PDOR), an enzyme that provides electrons to cloning, expression and detailed kinetic, spectroscopic,
phthalate dioxygenase (PDO) to enable this enzyme to hydrodynamic and thermodynamic analysis of the co-
oxygenate phthalate in a pathway for its degradation factor-replete C. metallidurans P450PDOR fusion
and exploitation as a carbon source for growth [30,31]. enzyme (CYP116B1). The enzyme exhibits characteris-
As with P450 BM3, the P450 module in the novel tic haem Soret spectral properties for the P450 class,
CYP116PDOR fusion enzymes is at the N terminus but displays some notable differences to features
of the fusion protein (Fig. 1). The Rhodococcus enzyme observed for CYP116B2. In this work, light scattering
from this family (CYP116B2) was expressed and puri- studies showed CYP116B1 to be monomeric and to
ed, and shown to exhibit P450-like spectral character- bind polyunsaturated lipids with an induced shift in
istics. Genetic dissection of the enzyme enabled haem iron spin-state equilibrium. We showed that
characterization of FMN and 2Fe-2S centres in the CYP116B1 catalyses NADPH-dependent oxygenation
expressed PDOR module [32,33]. CYP116B2 was also of thiocarbamate herbicides, an activity consistent with
shown to have weak activity with a prototypical uo- its evolutionary relationship with a distinct (non-part-
rescent substrate for the P450 superfamily (7-ethoxy- ner fused) P450 enzyme (CYP116A1, also known as
coumarin). However, no physiologically relevant ThcB) from Rhodococcus erythropolis NI86 21 that
substrate was identied [32]. oxidatively degrades the thiocarbamate herbicide
S-ethyl dipropylthiocarbamate (EPTC) [34].

A Results

Expression, purification and UV-visible


absorption features of CYP116B1
The gene encoding the C. metallidurans CH34
B CYP116B1 P450 (gene 4932, UniProt Q1LDI2) was
cloned by PCR from bacterial genomic DNA and
expressed in Escherichia coli (Origami (DE3)) using a
pET11a construct. CYP116B1 was puried using a
combination of ion-exchange and hydrophobic afnity
chromatography, producing a single major band on
SDS PAGE (Fig. 2A, inset), consistent with the pre-
dicted 86.4-kDa molecular mass of the P450PDOR
Fig. 1. Schematic showing domain organization and electron flow fusion protein. The UV-visible absorption properties
in CYP116B and CYP102A (P450 BM3) enzymes. (A) The genetic
of the pure CYP116B1 in its oxidized, sodium dithio-
arrangement of the individual domains (indicated by cofactor con-
tent: haem [P450] in the P450 domain, and FMN and 2Fe-2S [Fe-S]
nite-reduced and reduced carbon monoxide (CO)-
in the reductase domain) in CYP116B1 is shown at the left. The bound forms are illustrated in Fig. 2A. For oxidized
diagram at the right shows the direction of electron flow from CYP116B1, the spectral features of a haem cofactor
NADPH through FMN, iron-sulfur and onto the P450 haem within a are clearly evident. The main Soret band has its
CYP116B1 monomer. (B) A similar diagram to (A) of genetic absorption maximum at 418 nm, with longer-wave-
arrangement for P450 BM3 (CYP102A1) is shown, with FMN and length alpha beta bands positioned at 566 and 532 nm,
FAD cofactors in the reductase domain. The diagram on the right
respectively. A pyridine haemochromagen assay to
illustrates electron flow within the P450 BM3 dimer. Electron trans-
fer occurs from NADPH to FAD and then to FMN in monomer 1.
determine the haem concentration of CYP116B1 was
Thereafter, electrons are transferred from FMN of monomer 1 to performed in triplicate and a CYP116B1-specic
the haem in monomer 2 [22]. In the electron-transfer diagrams, the extinction coefcient was determined to be e418 =
flow of electrons from NADPH is indicated by large blue arrows. 121 mm)1cm)1 at the Soret peak of the oxidized
The haem domains are represented by pink circles with the haem enzyme. This value includes the spectral contributions
shown in red; the iron-sulfur domain is in a coral circle with the from iron-sulfur and avin centres in the cofactor-
iron-sulfur cluster in red and yellow spacefill; and the FMN and
replete CYP116B1, and was used to determine the
FAD domains are in orange and yellow circles, respectively, with
the isoalloxazine ring portions shown in the centre. CYP116B1 is
CYP116B1 concentration in all further experiments.
represented as a monomer, while one monomer of the P450 BM3 On aerobic reduction with sodium dithionite, the
dimer is shown coloured as described above, while the second Soret band decreased in intensity and was slightly
monomer is shown in greyscale. blue-shifted to 417 nm. The less intense haem alpha

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1677
Biochemical analysis of CYP116B1 A. J. Warman et al.

A rous, thiolate-coordinated state) [35,36]. The less-exten-


sive Soret shift for CYP116B1 is a consequence of
only partial reduction of the haem iron by dithionite
under aerobic conditions, which, in turn, is a result of
the rapid reaction of its ferrous haem iron with dioxy-
gen and its subsequent re-oxidation with production of
superoxide. Anaerobic titration of CYP116B1 with
dithionite led to complete reduction of the haem iron
and a more extensive Soret shift (see the section
Examination of the thermodynamic properties of
CYP116B1). However, the simultaneous drop in
absorption intensity in the region beyond  425 nm
suggests that the major cause of the apparent Soret
absorption decrease is actually the reduction of, and
B concomitant changes in the spectral contributions of,
the other cofactors bound to the protein. The reduc-
tase (PDOR) domain (amino acids  445780 of the
780-residue CYP116B1) is expected to bind both avin
(FMN) and ironsulfur cofactors, by homology with
both the B. cepacia PDOR and the reductase domain
of Rhodococcus sp. NCIMB9784 CYP116B2 [33].
Reductive bleaching of FMN, and 2Fe-2S cofactor
absorption by dithionite thus underlies the decreased
Soret intensity and the absorption bleaching observed
beyond  425 nm.
In view of the differing coenzyme selectivity of the
B. cepacia PDOR (which favours NADH) and the
Rhodococcus CYP116B2 (which favours NADPH)
[30,32], we also analyzed spectral changes on aerobic
Fig. 2. UV-visible absorption spectroscopy of CYP116B1. (A) mixing of oxidized CYP116B1 with NADH and
UV-visible absorption spectra for CYP116B1 (2.9 lM) showing the NADPH (50 lm). Interestingly, spectral changes simi-
purified oxidized enzyme (thick solid line) with Soret maximum at lar to those observed with dithionite were seen using
418 nm, the dithionite-reduced enzyme (dashed line) and the redu-
NADPH as reductant, but NADH only effected par-
ced CO-bound adduct (thin solid line) showing the Fe(II)CO peak at
tial reduction of the FMN ironsulfur centres (data
449 nm with a minor peak at 420 nm from the (protonated) cyste-
ine thiol-bound form. Inset: SDS PAGE gel (10%) of CYP116B1 not shown). In view of our subsequent work (vide
showing the purified enzyme at 86.4 kDa at three different concen- infra), which shows that both NADPH and NADH
trations of 1, 0.5 and 0.1 lg (left to right), sizes of marker bands support CYP116B1 catalytic activity, it is likely that
indicated. (B) Complexes of CYP116B1 with small-molecule inhibi- this reects faster enzyme reoxidation (and or slower
tors. Spectra shown are for oxidized CYP116B1 (2.3 lM, thick solid reduction) for NADH-reduced CYP116B1 compared
line); NO-bound CYP116B1 (thin solid line), showing a Soret inten-
with the NADPH-reduced form. This conclusion is
sity decrease and shift to 433 nm with Q-bands at 574 and
consistent with stopped-ow studies reported later,
545 nm; cyanide (CN)-bound (50 mM KCN) CYP116B1 (dotted line),
showing a Soret intensity decrease and shift to 430 nm; and imid- under Steady-state and stopped-ow kinetic studies on
azole (2 mM)-bound CYP116B1 (dashed line), exhibiting a Soret shift CYP116B1.
to 424 nm with a small decrease in intensity. Inset: CYP116B1- Addition of CO to the dithionite-reduced
imidazole binding curve generated from difference spectral data (as CYP116B1 resulted in a shift of the Soret peak to
described in the Materials and methods), yielding a Kd value of 449 nm, with a much smaller feature at  424 nm.
646 13 lM.
The former band is the characteristic signature for
the cysteine thiolate-ligated P450 haems in their
Fe(II)CO complexes. The minor feature at  424 nm
and beta bands remained at  566 and 532 nm, respec- probably reects a small proportion of the (proton-
tively. The extent of Soret absorbance shift was less ated) cysteine thiol-coordinated P420 form. Proton-
than that observed for other P450s (e.g. the P450 BM3 ation of the proximal ligand can arise as a
Soret shifted to  409 nm on full reduction to the fer- consequence of either haem reduction or the binding

1678 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

of CO to ferrous haem iron, or both [37,38]. Other were no signicant optical changes observed with the
spectroscopic studies indicate that the cysteine thio- saturated fatty acids tested (lauric acid [C12], myris-
late-coordinated state is predominant in the oxidized tic acid [C14], pentadecanoic acid [C15] and palmitic
enzyme (see Characterization of CYP116B1 haem acid [C16]), small type-I spectral shifts (typically
and ironsulfur clusters by EPR). In the Fe(II)CO 23 nm, spin-state change of  410%) were seen with
complex of CYP116B1, there is a single absorption a number of monounsaturated fatty acids (myristo-
Q-band peak in the visible region at  549 nm, indic- leic acid [C14], palmitoleic acid [C16] and oleic acid
ative of complete conversion of the haem iron to the [C18]), and with the polyunsaturated C20 arachidonic
ferrous state (Fig. 2A). acid. The low solubility of oleic acid in aqueous buf-
To further dene the spectral features of CYP116B1, fer did not allow for afnity to be determined by
we generated ligand complexes using the small-mole- optical titration. However, this was feasible for the
cule inhibitors imidazole, cyanide and nitric oxide three other lipids, and data for induced optical
(NO) (Fig. 2B). All three molecules induced red (type change versus fatty acid concentration were tted
II) shifts in the Soret band, which is typical for haem using a hyperbolic function to generate Kd values of
iron-coordinating P450 inhibitors. The imidazole com- 4.8 1.7 lm (myristoleic acid), 33.5 6.3 lm (pal-
plex had its Soret band shifted to 423.5 nm with a mitoleic acid) and 5.6 1.3 lm (arachidonic acid).
small decrease in intensity. An optical binding titration Data are shown for myristoleic acid in Fig. 3A.
with imidazole revealed a hyperbolic dependence of Given the successful demonstration of binding of
induced absorption change on ligand concentration fatty acid molecules to CYP116B1, we also investi-
and a Kd of 646 13 lm (Fig. 2B, inset). The Soret gated whether fatty acid-linked imidazole inhibitors
shift from  418 nm to 423.5 nm on complexation of (imidazolyl decanoic acid [ImC10], imidazolyl
imidazole and triazole inhibitors is typical of behav- undecanoic acid [ImC11] and imidazolyl dodecanoic
iour seen in other P450s (e.g. 38,39). However, it was acid [ImC12]) could act as potent inhibitors of P450.
reported that puried Rhodococcus CYP116B2 has its In previous studies we demonstrated tight binding of
Soret maximum at  424 nm in the ligand-free state ImC10, ImC11 and ImC12 to the P450 BM3 fatty
[32]. Given that this enzyme was puried by imidazole acid oxygenase [43]. These molecules also bound
elution from nickel resin, it is probably the case that tightly to CYP116B1 and induced type-II spectral
the CYP116B2 form puried was the imidazole com- changes (Soret shifts to 424 nm), consistent with
plex. The CYP116B1 complex with (sodium) cyanide haem iron coordination. The Kd values obtained
resulted in a Soret shift to 431 nm and a substantial from tting optical-titration data using a quadratic
decrease in peak intensity (as also observed, for exam- equation for tight-binding ligands were 3.5 0.4 lm
ple, in the Mycobacterium tuberculosis sterol demethy- (ImC10), 0.59 0.12 lm (ImC11) and 0.52 0.11
lase CYP51B1) [38]. Optical titration indicated a Kd lm (ImC12) [44]. Data for the binding of ImC12 to
value of 9.7 0.6 mm for cyanide. Finally, binding of CYP116B1 are shown in Fig. 3B. These Kd values
the gaseous ligand NO led to a Soret shift to 433 nm, compare favourably to those determined for binding
with a distinctive sharpening of features in the visible of ImC10, ImC11 and ImC12 to P450 BM3 (0.9, 7.5
region (Q-bands at 574 and 545 nm), and a spectrum and 1.35 lm, respectively), and with data for binding
comparable to the NO-bound form of P450 BM3 to a rat CYP4A1CPR fatty acid hydroxylase fusion
(Fig. 2B) [40]. protein (12.4, 0.37 and 0.35 lm, respectively) [43,45].
The binding of P450 substrate molecules often In view of the close relationship (52% amino-acid
induces spectral shifts in the Soret region. Blue shifts sequence identity) of the CYP116B1 haem domain
(towards  390 nm) are observed when substrates with the R. erythropolis CYP116A1 (also known as
cause displacement of an aqua ligand bound weakly ThcB, a non-fused P450 shown to catalyse the N-deal-
to the sixth coordination position on the haem iron, kylation of EPTC [34]), we also analysed the binding
as seen, for example, for camphor binding to of EPTC and the related thiocarbamate herbicide
P450cam (CYP101A1) and for fatty acid binding to S-propyl dipropylthiocarbamate (vernolate) to
P450 BM3 [41,42]. To identify potential substrate CYP116B1. Very small type-I haem perturbations were
molecules for CYP116B1, we undertook extensive observed, but these were not large enough to enable
binding trials with a diverse range of prototypical accurate determination of Kd from optical binding.
P450 substrates, including fatty acids (e.g. lauric acid However, we were able to show that CYP116B1 catal-
and myristic acid), steroids and other cyclic mole- yses the oxidation of both vernolate and EPTC (see
cules (e.g. testosterone, hydrocortisone, 7-ethoxy- Identication of CYP116B1-catalysed reaction prod-
coumarin and 6-deoxyerythronolide B). While there ucts of thiocarbamate herbicides).

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1679
Biochemical analysis of CYP116B1 A. J. Warman et al.

A size-exclusion chromatography (SEC) coupled to


multi-angle laser light scattering (MALLS). In contrast
to P450 BM3, CYP116B1 was found to be predomi-
nantly monomeric in solution, eluting from a size-
exclusion column in a single band with an apparent
average molecular mass of 86.1 1.7 kDa, very close
to that predicted for the intact enzyme based on its
protein sequence (86.4 kDa) (Fig. 4). Steady-state
kinetic assays (see Steady-state and stopped-ow
kinetic studies on CYP116B1) revealed no discernible
CYP116B1 concentration dependence on the specic
rate of vernolate-induced NADPH oxidase activity.
This is consistent with a single (monomeric) state of
CYP116B1 and contrasts with the decreased specic
activity of P450 BM3 at low enzyme concentrations
B that occurs as a consequence of dissociation of the
dimer into inactive monomers [22]. It thus appears
likely that CYP116B1 (and probably also CYP116B2
and related CYP116B family enzymes) operate by
intramolecular electron transfer within a monomer, as
opposed to intermolecular electron transfer observed
for P450 BM3 (and the eukaryotic l-arginine oxidase
nitric oxide synthase avocytochromes) [22,46].

Characterization of CYP116B1 haem and


ironsulfur clusters by EPR
EPR spectra for the oxidized and sodium dithionite-
reduced forms of pure CYP116B1 were collected as
Fig. 3. Substrate- and inhibitor-type binding in CYP116B1. (A) Bind- described in the Materials and methods. For oxidized
ing curve for myristoleic acid-induced spectral change in CYP116B1 CYP116B1, a rhombic spectrum consistent with a fer-
(2.0 lM). Data were fitted using a quadratic function [44], generat- ric cysteinate-coordinated haem iron was observed,
ing a Kd value of 4.8 1.7 lM for myristoleic acid. Inset: difference
spectra showing the progressive substrate-like Soret absorbance
shifts to shorter wavelength (blue shifts) upon addition of myristo-
leic acid at 7.96, 15.92, 23.88, 47.76 and 79.60 lM. (B) Binding
curve for ImC12-induced spectral change in CYP116B1. Data were
fitted as described for (A), producing a Kd value of 0.52 0.11 lM.
Inset: UV-visible absorbance spectra for CYP116B1 in ligand-free
form (2.1 lM, solid line) and with ImC12 at a near-saturating con-
centration (5.13 lM, dashed line). There is a shift in the Soret peak
maximum from 418 to 424 nm on binding ImC12.

Multi-angle laser light scattering analysis of


CYP116B1
The P450 the BM3 P450-CPR fusion enzyme was Fig. 4. SEC-MALLS analysis of CYP116B1. The MALLS profile is
shown to be catalytically functional in its dimeric state, shown for a pure CYP116B1 sample resolved by SEC on a Super-
dex 200 gel-filtration column before passing directly through
with electron transfer between the FMN of monomer
MALLS and refractive index (RI) detectors (the RI trace is shown
1 and the haem of monomer 2 probably supporting its as a solid line). The number average molar mass (Mn) across the
fatty acid hydroxylase activity (Fig. 1) [22,24]. In order single peak resolved (filled blue circles) was determined as
to analyze the aggregation state of the P450PDOR 86.1 1.7 kDa. These data are consistent with CYP116B1 being a
fusion enzyme (CYP116B1) in solution, we performed monomeric enzyme.

1680 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

with g-values of 2.42 (gz), 2.25 (gy) and 1.93 (gx) 3750 Gauss. The reduced CYP116B1 spectrum has a
(Fig. 5). These g-values lie in the range typical for rhombic shape and g-values of gz = 2.06, gy = 1.97
other low-spin P450 enzymes [e.g. 38,47,48], and are and gx = 1.88 that fall within the typical ranges
consistent with an essentially completely low-spin P450 reported by More et al. [52], clearly identifying the pres-
enzyme. There were no signicant signals in the region ence of a reduced 2Fe-2S cluster. There is no EPR signal
associated with the high-spin P450 haem iron (e.g. that could be clearly assigned to a avin semiquinone in
g = 8.18 7.85, 3.44 3.97 and 1.66 1.78 for P450 the reduced CYP116B1. The EPR spectrum for the
BM3 P450cam in substrate-bound forms [48,49]), sug- 2Fe-2S cofactor-reconstituted PDOR (2Fe-2S- and
gesting that ferric CYP116B1 is hexacoordinated with FMN-binding) domain of CYP116B2 from Rhodococ-
a water ligand trans to the cysteine thiolate. Similar cus sp. NCIMB 9784 produces an identical set of
sets of g-values were obtained for the low-spin forms g-values [33]. The CYP116B1 2Fe-2S EPR spectrum is
of other bacterial P450s. These include P450 BM3 and also highly similar to that for the homologous B. cepa-
P450 BioI from Bacillus subtilis (CYP107H1), whose cia PDOR (g = 2.04, 1.95 and 1.90), and to that of the
EPR spectra yield g-values of 2.42, 2.26 and 1.92, and Arabidopsis photosynthetic ferredoxin Fd2 (g = 2.05,
2.41, 2.25 and 1.92, respectively [47,48]. 1.95, 1.89) [53,54]. Integration of the EPR signals for
For oxidized CYP116B1, there were no apparent the oxidized haem and for the reduced ironsulfur
EPR signals that could be ascribed to either ironsulfur cofactors in CYP116B1 indicate that the 2Fe-2S centre
or avin centres that are likely to be present in the is present at  98% of the concentration of the haem,
PDOR domain of the enzyme. Oxidized avins are EPR thus suggesting near-complete incorporation of the
silent, and the 2Fe-2S cluster in its oxidized form (i.e. iron-sulfur cluster in CYP116B1.
[2Fe-2S]2+) should contain two ferric iron atoms, in To identify the avin bound to CYP116B1, we iso-
which the magnetic moments of the unpaired electrons lated the bound cofactor from a puried enzyme sam-
are spin-coupled, giving S = 0. However, reduced 2Fe- ple, as described in the Materials and methods.
2S clusters are EPR active in the [2Fe-2S]+ form, with Thereafter, avin uorescence was determined at both
one ferric and one ferrous iron atom [e.g. 50,51]. An pH 7.7 and pH 2.0, as described previously [55], and
EPR spectrum was collected for the dithionite-reduced compared with FAD and FMN standards treated in
form of CYP116B1 and is shown in Fig. 5. The spec- the same way. Decreased avin uorescence was
trum is more compact than that for the oxidized enzyme observed for both the CYP116B1 cofactor extract and
with the major features appearing between 3250 and for FMN standards on acidication, while enhanced
uorescence was seen for FAD (owing to hydrolysis of
its pyrophosphoryl bond and release of ADP and the
more highly uorescent FMN). FMN content was esti-
mated at 55% by comparison with the quantity of
bound haem, consistent with the previous report of
variable FMN content in CYP116B2 [33]. The inclu-
sion of free FMN (1 mm) in buffers used for
CYP116B1 purication enabled increased FMN con-
tent in the puried enzyme (typically to > 80%).

Steady-state and stopped-flow kinetic studies on


CYP116B1
Analysis of the ability of CYP116B1 to reduce electron
acceptor molecules (potassium ferricyanide, cyto-
Fig. 5. X-Band EPR spectra for CYP116B1. Upper spectrum: oxi- chrome c) in a NAD(P)H-dependent manner and to
dized CYP116B1 (184 lM) showing a rhombic signal with g-values perform P450 substrate-dependent oxidation of
typical for a low-spin, six-coordinate ferric haem iron at gz = 2.42, NAD(P)H was examined in steady-state reactions
gy = 2.25 and gx = 1.93. Lower spectrum: CYP116B1 (184 lM), (Table 1). There was a strong preference for NADPH
reduced anaerobically by dithionite, showing almost complete loss
over NADH, with a difference of  440-fold in Km
of the signal from the ferric haem species (as haem is reduced to
the EPR-silent ferrous state) and the emergence of bands at
values for NADPH (0.9 lm) and NADH (399.1 lm)
gz = 2.06, gy = 1.97 and gx = 1.88, typical of a plant-like 2Fe-2S determined in cytochrome c-reduction studies. The kcat
cluster in the EPR active [2Fe-2S]+ form. No significant signal is values for both ferricyanide and cytochrome c reduc-
observed for a flavin semiquinone. tion were also three-fold greater with NADPH as the

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1681
Biochemical analysis of CYP116B1 A. J. Warman et al.

Table 1. Steady-state kinetic properties of CYP116B1. Km and kcat Despite favourable haem optical changes induced by
values for CYP116B1 were determined from steady-state kinetic myristoleic, palmitoleic, oleic and arachidonic acids,
assays with NADH and NADPH as reductants, and using the elec-
none of these molecules stimulated CYP116B1-depen-
tron acceptors potassium ferricyanide (FeCN) and cytochrome c,
and the P450 substrates vernolate and EPTC. Km values shown are
dent oxidation of either NADH NADPH to any signif-
for NAD(P)H in the experiments using cytochrome c and FeCN, icant levels. By contrast, substantial stimulation of
and for vernolate EPTC in the experiments using the thiocarba- oxidation of both coenzymes was observed with EPTC
mates. In experiments using the thiocarbamate substrates, NADH and vernolate as substrates. A hyperbolic dependence
(1.5 mM) and NADPH (200 lM) were used at near-saturating con- of coenzyme oxidation rate on EPTC concentration
centrations. All assays were performed on a dual-beam spectropho- was obtained using both NADH and NADPH, and
tometer (Jasco) in buffer A at 25 C using CYP116B1 at a final
data were tted using the Michaelis function to pro-
concentration of 75100 nM
duce values of kcat = 30.7 min)1 and Km EPTC =
Reductant 50.3 lm (with NADPH as the donor) and kcat = 17.0 -
min)1 and Km EPTC = 110.4 lm (with NADH as the
Acceptor or substrate NADPH NADH
donor). For vernolate, the comparable values were
Cytochrome c kcat = 57.3 min)1 and Km vernolate = 210.2 lm (using
Km (lM) 0.9 0.5 399.1 52.1 NADPH as the donor) and kcat = 34.0 min)1 and
kcat (min)1) 151.1 8.0 57.1 2.7 Km vernolate = 270.4 lm (using NADH as the donor)
(Table 1). Thus, EPTC (a substrate that is dealkylated
FeCN
Km (lM) 3.0 1.0 102.1 12.2
in whole cells of R. erythropolis that express
kcat (min)1) 833.1 17.8 286.8 7.5 CYP116A1) clearly stimulates NAD(P)H oxidation in
CYP116B1, and the related dithiocarbamate vernolate
Vernolate enhances NAD(P)H oxidation to an even greater
Km (lM) 210.2 30.7 270.4 58.8 extent. In the absence of vernolate, NADPH-dependent
kcat (min)1) 57.3 1.9 34.0 2.2 reduction of the CYP116B1 haem iron is slow in a CO-
saturated solution ( 0.2 min)1) and results mainly in
EPTC
Km (lM) 50.3 9.6 110.4 10.0
the P420 form of the enzyme. In the presence of satu-
kcat (min)1) 30.7 0.5 17.0 0.4 rating vernolate, there is some stimulation of the rate
of haem iron reduction, but data analysis is compli-
cated by a greater tendency of the enzyme to aggregate
with development of turbidity in the solution. Despite
the progressive aggregation of vernolate-bound
coenzyme than with NADH (e.g. 833.1 min)1 versus CYP116B1 in this experiment, it is notable that the pre-
286.8 min)1 for ferricyanide), resulting in a 99-fold dif- dominant species is now the P450 form, indicating sub-
ference in catalytic efciency for ferricyanide reduction strate-dependent stabilization of the thiolate ligand, as
(measured as the ratio of kcat Km[NADPH] to kcat Km[- observed previously in other systems [37,38,56]. Given
NADH]) and a 1174-fold difference for cytochrome c the limited solubility of vernolate, its marginal effect on
reduction in favour of NADPH (Table 1). CYP116B2 haem iron spin-state equilibrium and its inuence on
also demonstrated preference for NADPH aggregation of CYP116B1, it is difcult to establish the
(Km = 3.0 lm) over NADH (102.1 lm) in steady-state extent to which electron transfer to the CYP116B1
ferricyanide-reduction assays [32]. In contrast, the haem iron is thermodynamically gated in the presence
B. cepacia PDOR avoprotein has strong preference of this substrate and a kinetic gate to haem reduction
for NADH (Km  10 lm) over NADPH, and NADPH in CYP116B1 cannot be ruled out [57]. Consistent with
did not support measurable activity with either cyto- the monomeric form of CYP116B1 being the catalyti-
chrome c or the partner protein PDO, pointing to an cally relevant species, specic rates of vernolate-depen-
important evolutionary divergence between the stand- dent NADPH oxidation, determined across a range of
alone PDOR enzyme and the fused PDOR domains in CYP116B1 concentrations (10500 nm), showed no sig-
the CYP116B enzymes. nicant variation. In contrast, similar studies with P450
In view of the preceding data indicating substrate- BM3 and the substrate lauric acid showed diminished
like (type I) Soret spectral shifts induced on binding specic activity at low enzyme concentration as a result
both unsaturated fatty acids and (to a lesser extent) the of disassociation of the catalytically active dimer [22].
thiocarbamate herbicides EPTC and vernolate, we To further characterize the reaction of CYP116B1
analysed steady-state oxidation of NADPH by with NAD(P)H, stopped-ow absorption spectroscopy
CYP116B1 in the presence of these potential substrates. was used to analyse the kinetics of reduction of

1682 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

CYP116B1 by NADH and NADPH. Absorption tran- A


sients were collected at 465 nm (i.e. near the expected
absorption maximum for the oxidized FMN cofactor
[33]) after mixing CYP116B1 with different concentra-
tions of NAD(P)H. Reaction transients were biphasic
in all cases and were tted accurately using a double
exponential function. Approximately 80% (NADH)
and 90% (NADPH) of the overall absorption change
observed on coenzyme-dependent reduction of the
FMN 2Fe-2S centres occurred in the fast phase of the
reaction transients. The CYP116B1 haem was not
reduced by NAD(P)H to any extent in its substrate-
free form. Figure 6A shows a plot of the individual
rate constants for the fast (kobs1) and slow (kobs2)
phases of the transients against NADH concentration,
while Fig. 6B shows the same data set with NADPH B
as reductant. For NADH-dependent reduction, kobs1
shows a hyperbolic dependence on the NADH concen-
tration, and data were tted to yield a limiting rate
constant (klim) of 21.9 0.2 s)1 and an apparent
NADH Kd of 21.7 0.8 lm. The kobs2 values for
NADH are much slower (0.47 0.03 s)1) and show
no dependence on the NADH concentration (Fig. 6A).
For NADPH-dependent reduction of CYP116B1, the
reaction rate constants are higher, but there is no
dependence of either kobs1 (72 6 s)1) or kobs2
(5.5 3.0 s)1) on the NADPH concentration
(Fig. 6B).

Examination of the thermodynamic properties of Fig. 6. Stopped-flow kinetic analysis of NADH- and NADPH-depen-
CYP116B1 dent CYP116B1 reduction. Stopped-flow absorption kinetic studies
of CYP1161B1 (3.3 lM final concentration) were performed by
To analyse the redox properties of CYP116B1, a monitoring the reduction of FMN 2Fe-2S centres at 465 nm (the
spectroelectrochemical redox titration was performed, wavelength of maximal overall change between oxidized and
using methods described previously [58,59]. For the reduced CYP116B1) using both NADH (A) and NADPH (B) reducing
intact CYP116B1, the overlapping spectral contribu- coenzymes at a range of concentrations up to 500 lM. Open circles
tions of the individual haem, FMN and 2Fe-2S cofac- represent individual rate constants derived from the fast phase
tors resulted in difculty in deconvoluting potentials (kobs1) of double-exponential fits to the DA465 versus time tran-
sients, while open squares represent kobs2 rate constants for the
for the three centres in the multicofactor enzyme.
slow phase of the reaction transients. There is no apparent depen-
However, it is clear that the reduction process occurs dence of either kobs1 (72 6 s)1) or kobs2 (5.5 3.0 s)1) on
in two phases. In the rst phase (black spectra in [NADPH]. With NADH as reductant, there is a hyperbolic depen-
Fig. 7A, completed in the range of 0 to )200 mV ver- dence of kobs1 on [NADH], giving a limiting rate constant (klim) of
sus the normal hydrogen electrode [NHE]), spectral 21.9 0.2 s)1. However, there is no apparent dependence of kobs2
bleaching of the contributions of the ironsulfur and on [NADH] (0.47 0.03 s)1).
FMN cofactors occurs at  400600 nm (as seen for
the dithionite-reduced spectrum in Fig. 2A). There is positive than that for the oxidized semiquinone cou-
no obvious development of a long wavelength signal ple (E1). With the assumption that absorption
(at  600 nm) during this reductive phase (Fig. 7A), changes at 485 nm (in the potential range where there
suggesting that there is negligible formation of the is negligible haem contribution) reect primarily the
FMN neutral semiquinone, as also observed in the transition from FMN quinone to hydroquinone (E12)
dithionite-reduced CYP116B1 EPR spectrum (Fig. 5). (with some absorption change also resulting from
This may indicate that the reduction potential for the the reduction of the 2Fe-2S cluster), the midpoint
FMN semiquinone hydroquinone couple (E2) is more potential for the two-electron FMN reduction was

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1683
Biochemical analysis of CYP116B1 A. J. Warman et al.

estimated as )122 9 mV (Fig. 7B). Data tted at Fig. 7. Spectroelectrochemical titration of CYP116B1. (A) UV-visible
409 nm (an isosbestic point for the haem iron Fe(III) spectra recorded during a redox titration of CYP116B1 (6.3 lM).
The arrows show directions of absorbance changes occurring at dif-
to Fe(II) transition) produced a similar value
ferent wavelengths during the reductive phase of the titration.
()134 8 mV) for the combined reduction of the Absorbance spectra demonstrating the reduction of the FMN and
FMN 2Fe-2S centres. Although differing proportions 2Fe-2S clusters at more positive potentials are shown as black
of absorption contribution probably come from the lines, while spectra showing mainly reduction of the CYP116B1
FMN and ironsulfur cluster at the two wavelengths, haem iron from ferric to ferrous (at more negative potentials) are
the midpoint potential values are within error and it shown as red lines. (B) A plot of absorbance data at 485 nm versus
applied potential in the range from  +25 to )200 mV (versus
NHE). Data are fitted using the Nernst function to define a midpoint
A
potential of )122 9 mV for the reduction of both the FMN and
2Fe-2S centres in CYP116B1. (C) A plot of absorbance data at
418 nm versus applied potential in the range from  +25 to
)450 mV (versus NHE). The plot shows two absorbance transi-
tions, and data were fitted using a two-electron Nernst function to
define midpoint potentials of 160 9 mV (encompassing both
FMN and 2Fe-2S centres) and )301 7 mV (for the CYP116B1
haem iron Fe(III) Fe(II) couple).

is clear that complete reduction of both cofactors


occurs in this potential range.
In the second stage of CYP116B1 reduction that
occurs in the range from )200 to )500 mV versus
NHE, the haem iron is converted from the ferric to
B the ferrous state. Haem iron reduction occurs with a
shift in the Soret peak from 418 to 410 nm (red spectra
in Fig. 7A), consistent with the retention of a cystei-
nate-ligated CYP116B1 haem iron and similar to the
spectral maximum observed for the reduced form of
P450 BM3 [58]. Data tted at 393 nm, where there are
negligible absorbance changes from the early part of
the reductive (FMN and iron-sulfur) titration, show a
single transition with a midpoint potential of
)291 9 mV (data not shown). Absorption versus
potential data plots at the oxidized Soret maximum
(418 nm) show a clear two-stage transition (Fig. 7C).
The decrease in A418 intensity in the rst stage of the
titration is attributed to the reduction of the FMN
and iron-sulfur cofactors. This accounts for approxi-
C mately half of the overall absorption change and is
indicative of high levels of incorporation of these
cofactors in CYP116B1. The midpoint potentials
for the two transitions were determined using a two-
electron Nernst function [58], yielding values of (a)
)160 9 mV to encompass the FMN E12 and iron
sulfur cluster [2Fe-2S]2+ [2Fe-2S]+ couples and (b)
)301 7 mV for the haem iron Fe(III) Fe(II) couple.
These midpoint potential values are consistent with the
data ts at other wavelengths. Variations in the rela-
tive optical contributions of the FMN and 2Fe-2S
cofactors may explain the small differences in redox-
potential estimates for their combined reduction at
different analysis wavelengths.

1684 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

almost identical mass spectrum may also be indicative


Identification of CYP116B1-catalysed reaction
of hydroxylation occurring at different positions on
products of thiocarbamate herbicides
the N-propyl chain, although we cannot dene their
Following the fatty acid thiocarbamate binding and exact position(s) from these data. The mass spectrum
steady-state kinetic studies reported above, we under- of the third small species observed at 37.7 min in the
took experiments to establish whether CYP116B1 chromatogram (indicated by an arrow in Fig. 8A) gave
catalysed the oxidation of unsaturated fatty acids or of a distinct single peak with an m z of 162 (Fig. 8E).
the thiocarbamates vernolate and EPTC. No formation This m z ratio is consistent with dealkylation (loss of a
of oxidized products was observed for NAD(P)H-driven propyl group) from the parent vernolate. This is most
turnover of CYP116B1 with monounsaturated fatty likely to be an N-dealkylation reaction given the differ-
acids. However, preliminary studies indicated that prod- ent m z ratio to that observed for the loss of the S-pro-
ucts were formed from both vernolate and EPTC, with pyl group upon fragmentation. However, this
superior levels of product formation seen using dealkylated product is a minor species compared with
NADPH as the reducing coenzyme (consistent with the the hydroxylated form(s) of vernolate, as seen from the
steady-state kinetic data in Table 1). A detailed charac- difference in sizes of the respective peaks in the HPLC
terization of the products from the NADPH-dependent chromatogram. Similar hydroxylated products on the
turnover of CYP116B1 with these thiocarbamates was N-propyl chains are also observed for the thiocarbamate
thus undertaken by liquid chromatography (LC)-MS, as EPTC, which differs by a CH2 group in the S-ethyl
described in the Materials and methods. chain (S-propyl in vernolate) (Fig. S3), although these
HPLC was used to analyse products from are found in lower amounts compared with the verno-
CYP116B1-catalysed oxidation of vernolate and late products. There is no evidence of CYP116B1-medi-
revealed major features in the chromatogram at 34 ated N-dealkylation with EPTC, although we cannot
39 min after sample loading, in addition to the remain- rule out small amounts of such a product. Given the
ing parent molecule that eluted between 50 and 52 min greater proportion of hydroxylation and N-dealkylation
(Fig. 8A). The new features consisted of two major seen with vernolate over EPTC, it appears likely that
product peaks, with a further small peak distinguish- CYP116B1 binds more favourably to the S-propyl
able between these major peaks. These product fea- group in vernolate than to the S-ethyl group in EPTC
tures are absent from vernolate-only or enzyme-only to enable oxidative catalysis.
controls. The shorter retention time of these peaks
indicates that they arise from compounds more polar
Discussion
in nature than the parent vernolate. The mass spectra
of the major bands at 37.1 and 38.4 min in the chro- We report here the rst biochemical characterization
matogram are nearly identical, with two major peaks of CYP116B1, a representative of a family of cyto-
observed with m z values at 220 and 144 (Fig. 8B). chrome P450PDOR fusion enzymes with biotechno-
The larger peak, at m z 220, corresponds to the mass logical potential. Like P450 BM3 and other related
(M+H) of vernolate with a single hydroxylation. Fur- P450P450 reductase fusion enzymes, this makes
ther analysis of this peak performed by collision- CYP116B1 a catalytically self-sufcient P450 enzyme,
induced disassociation (MS2) identied a single MS2 requiring only substrates [NAD(P)H coenzyme and
fragment with m z ratio 144. Analysis of the vernolate thiocarbamate herbicide] for turnover (2,12,32).
structure revealed this MS2 fragment to be a hydroxyl- The C. metallidurans CYP116B1 was expressed in
ated form of vernolate lacking the S-propyl group and E. coli and puried to homogeneity using a combina-
probably arising from cleavage of the S-C bond tion of anion-exchange and hydrophobic afnity
(Fig. 8C). MS2 analysis of the secondary HPLC- chromatography. Heterologous expression of cofactor-
derived peak at m z 144 revealed a single feature with binding proteins in E. coli can result in incomplete
an m z ratio of 116, which again is a hydroxylated incorporation of non-covalently bound cofactors.
product of a vernolate fragment lacking both the Indeed, in the case of the Rhodococcus sp. CYP116B2
S-propyl group and the carbonyl group (Fig. 8D). This enzyme, deciencies in both FMN and 2Fe-2S cofac-
fragmentation pattern that shows loss of the S-propyl tors were noted [33]. In our study of the C. metallidu-
group is consistent with that observed in the vernolate- rans CYP116B1, EPR quantication revealed
only controls (Fig. S2A,B) and conrms the position(s) essentially stoichiometric incorporation of the 2Fe-2S
of CYP116B1-catalysed hydroxylation to be on one of cluster and the P450 haem (Fig. 5). The FMN cofactor
the N-propyl chains. The fact that we observed two was more weakly bound, but inclusion of FMN in
major peaks in the HPLC chromatogram with an purication buffers enabled FMN incorporation to

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1685
Biochemical analysis of CYP116B1 A. J. Warman et al.

B C

D E

Fig. 8. Analysis of products of CYP116B1-catalysed vernolate oxidation by liquid chromatography (LC)-MS. (A) HPLC chromatogram showing
elution profiles of vernolate substrate (upper blue line) and CYP116B1-catalysed oxidation products of vernolate (lower black line). Vernolate
elutes as a single peak between 50 and 52 min. Two additional peaks are observed following turnover with CYP116B1. These elute between
34 and 39 min and were determined to be hydroxylated forms of vernolate. A further sharp peak of lower intensity is also identified within
this region, eluting at 37.7 min and indicated by the arrow. This feature is assigned to an N-dealkylated product of vernolate. (B) The mass
spectrum of the chromatogram peaks at 37.1 and 38.4 min (both found to contain almost identical masses) showing two major species with
m z values of 220 and 144, and a smaller m z feature at 162. The m z 220 species results from hydroxylation of vernolate
(m z = 204 + 16). The structure of a potential hydroxylated product of vernolate is shown as an inset. (C) MS2 analysis of the m z 220 spe-
cies with a resultant m z at 144. This 144 m z species is assigned to a hydroxylated fragment of vernolate, with the S-propyl group
removed. A representative structure of the hydroxylated vernolate fragment structure is shown as an inset. (D) MS2 analysis of the m z 144
species with a resultant m z at 116. This is assigned to a hydroxylated fragment of vernolate, with the S-propyl and adjacent carbonyl
groups removed. These data are further confirmatory that the hydroxylation occurs on one of the N-propyl chains, but do not define an exact
position. A representative structure of this hydroxylated fragment is shown as an inset. (E) Mass spectrum of the minor peak at 37.7 min in
the CYP116B1 + vernolate chromatogram (A), showing a predominant m z peak at 162. This is consistent with dealkylation (loss of a pro-
pyl group) from vernolate. While the dealkylation could theoretically occur with removal of either an N- (shown inset) or an S-propyl chain,
the likely product is the N-dealkylated form (as seen for the reaction of Rhodococcus erythropolis ThcB with EPTC), occurring by CYP116B1-
catalysed C-N bond cleavage following successive hydroxylations at adjacent carbons on the N-propyl chain [34].

> 80%. UV-visible, EPR and resonance Raman spec- is instead indicative of an enzyme in which imidazole
troscopy (see the Supporting Information) conrmed (used to elute the enzyme from nickel afnity resin)
that that enzyme was isolated in a predominantly has displaced water as the sixth ligand to produce an
low-spin ferric state, with cysteinate and water as the inhibited form of the enzyme [32]. Previous studies of
fth and sixth (axial) ligands to the haem iron intact CYP116B2 and its PDOR domain indicated a
(Figs 2A and 5, Fig. S1). The CYP116B1 Soret maxi- low content of the 2Fe-2S cluster (in the range of
mum at 418 nm is typical of such coordination and 436%), necessitating ironsulfur cofactor reconstitu-
spin-state in P450s, whereas the previous report of the tion of the enzyme under anaerobic conditions [33].
related CYP116B2 enzyme with Soret peak at 424 nm Given the essentially stoichiometric incorporation of

1686 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

the 2Fe-2S cluster reported here for CYP116B1 (puri- from C14 to C20, producing substrate-like type I spec-
ed without using a nickel afnity step), it appears tral changes with accumulation of HS haem iron (e.g.
possible that the nickel Sepharose resin used for puri- Fig. 3A). However, no evidence for their oxidation by
cation of CYP116B2 resulted in stripping of iron from CYP116B1 could be obtained either through measure-
the 2Fe-2S centre in this enzyme. ments of fatty acid-dependent stimulation of
There is no EPR signal attributable to bound avin in CYP116B1 NAD(P)H oxidation or isolation of oxi-
its EPR active (semiquinone) form in the dithionite- dized fatty acid products. In contrast, the thiocarba-
reduced form of CYP116B1 (Fig. 5), and no obvious mate herbicides vernolate and EPTC, despite showing
optical signals typical of a avosemiquinone are rather small P450 type I (HS) spectral changes, not
observed for either dithionite- or NADPH-reduced only stimulated CYP116B1-dependent NAD(P)H oxi-
CYP116B1 (Fig. 2A). We thus infer that the reduction dation (Table 1), but were also shown to be substrates
potential for the CYP116 FMN oxidized semiquinone for CYP116B1, and to be hydroxylated (EPTC and
couple (E1) is more negative than that for the semiqui- vernolate) and N-dealkylated (vernolate) in an
none hydroquinone couple (E2), meaning that little NADPH-dependent manner. The modest type I shifts
FMN semiquinone accumulates during avin reduction. induced by these herbicides possibly indicates that they
Bound avin is converted from quinone to the two-elec- can occupy different positions in the CYP116B1
tron hydroquinone form on reduction with dithionite, active-site pocket, with transient occupancy of the HS
and thus remains EPR silent. Redox potentiometry of position probably linked to oxidative catalysis by the
CYP116B1 (Fig. 7) demonstrates clearly that the reduc- P450. Steady-state kinetic analysis indicates a prefer-
tion of the non-haem centres in CYP116B1 takes place ence for vernolate over EPTC in CYP116B1, and this
at a more positive potential than does the reduction of is conrmed by the greater amounts of hydroxylated
the P450 haem iron. Bleaching of absorption occurs in products obtained with vernolate, as well as there
the avin (and ironsulfur) region between  420 and being clear evidence for N-dealkylation only in the
520 nm before there is any notable shift in the position case of vernolate. Vernolate differs from EPTC in the
of the haem Soret band. At more negative potentials, length of its S-alkyl side chain (S-propyl for vernolate
the Soret shifts from 418 to  410 nm, consistent with and S-ethyl for EPTC). The vernolate S-propyl chain
retention of cysteine thiolate as the haem iron fth may confer greater specicity for CYP116B1 than does
ligand in the reduced (ferrous) state of CYP116B1 the respective S-ethyl chain in EPTC. Although
[37,38]. The CYP116B1 haem iron potential CYP116B1 catalyses N- rather than S-dealkylation of
()301 7 mV versus NHE from data analysis at vernolate, it is conceivable that the S-propyl chain por-
418 nm) is more positive than that previously reported tion of the thiocarbamate makes stabilizing interac-
for the isolated haem (P450) domain of the related tions in the active site that are important for
CYP116B2 enzyme ()380 4 mV), but is also more orientating one of the N-propyl chains of vernolate
negative than the reduction potential of the 2Fe-2S and towards the CYP116B1 haem to enable its oxidation
FMN centres in CYP116B1, whose reduction is com- and subsequent N-dealkylation. At this stage we can-
pleted by  )200 mV. This suggests that binding of not entirely rule out the absence of CYP116B1-medi-
substrate to the CYP116B1 P450 active site may make ated S-dealkylation, although products are not
the haem iron potential more positive (through displace- detected under the experimental conditions used. Other
ment of the sixth water ligand to the ferric iron and N-dealkylation reactions are known in P450 chemistry,
concomitant shift in haem iron spin-state towards high- and such oxidations are common in human enzymes.
spin), as is the case for several other P450s [e.g. 41,58]. For instance, several such reactions are catalysed by
The limited solubility of vernolate EPTC, their mar- the major human drug-metabolizing enzymes CYP2D6
ginal effects on CYP116B1 haem spin-state equilibrium and CYP3A4, including N-dealkylation of the opioid
and their promotion of CYP116B1 aggregation prevents addiction drug buprenorphine [e.g. 60,61]. In the case
us from establishing whether the binding of these dith- of the vernolate N-dealkylation reaction catalysed by
iocarbamates is associated with a large elevation in CYP116B1, there are clear functional similarities
haem iron potential. This makes it difcult to establish between CYP116B1 and the non-partner fused P450
the extent to which substrate-dependent haem iron spin- ThcB in R. erythropolis NI86 21 [34]. ThcB
state modulation is important for control of enzyme (CYP116A1 in the P450 superfamily) was shown to
catalysis in CYP116B1, or whether an alternative mode oxidatively dealkylate EPTC. This is probably the rst
of regulation may be involved [e.g. 57,58]. of a series of reactions enabling thiocarbamate catabo-
In other work reported here, CYP116B1 was shown lism in the Rhodococcus strain. CYP116A1 and the
to bind to unsaturated fatty acids with chain lengths haem domain of CYP116B1 show 52% amino-acid

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1687
Biochemical analysis of CYP116B1 A. J. Warman et al.

sequence identity and are clearly orthologues. In of Birmingham, UK. To generate the megaplasmid DNA
R. erythropolis, genes encoding a 2Fe-2S ferredoxin template for Rmet_4932 cloning, single colonies of C. me-
(rhodocoxin) and a ferredoxin reductase (rhodocoxin tallidurans were resuspended in 1 mL of distilled, deionized
reductase) are located immediately downstream of the water (ddH2O), and 50 lL of the sample was heated at
P450, forming a classical bacterial class I P450 redox 99 C for 9 min in a PHC-2 thermal cycler (Techne, Bur-
system [2,34]. The divergent evolution of the CYP116B lington, NJ, USA). The PCR reaction to amplify the
family has clearly involved the genetic fusion of the CYP116B1 gene involved 30 cycles of (a) denaturation at
CYP116-type P450 with a distinctive redox partner 95 C for 45 s, (b) annealing at 59 C for 1 min and (c)
polymerization at 72 C for 5 min, followed by a nal poly-
system, presumably as a result of advantages gained in
merization step of 72 C for 10 min. The forward and
catalytic efciency. While it is uncertain whether thioc-
reverse primers used for this reaction were based upon the
arbamates such as EPTC and vernolate are physiologi-
published genome sequence of C. metallidurans (http://gen-
cally relevant substrates of CYP116B1, it is quite
ome.ornl.gov/microbial/rmet/), and were as follows:
possibly the case that C. metallidurans exploits this
CYP116B1F, GGACTAATCTCGCTGGACTGATTAATG
P450 as part of its armoury of enzymes for metabolism CCGC; and CYP116B1R, CGCTCAGCATCCGTGATG
of xenobiotics. C. metallidurans is a bacterium able to CCGTGCTGAG (Nucleotides in bold are sites for restric-
thrive in stressful pollutant heavy metal-contaminated tion enzymes AseI and BlpI, respectively, which were incor-
and nutrient-limited environments, and oxidases of porated to facilitate the subcloning of CYP116B1.
diverse substrate selectivity are probably important Underlined nucleotides indicate the initiation (ATG) and
components of its survival strategy [62]. termination (TGA) codons.) The PCR ( 2.4 kbp) product
In conclusion, we present the rst report of the bio- was puried using a Min-Elute gel-purication kit (Qiagen,
chemical, spectroscopic, thermodynamic and catalytic Crawley, UK) using the manufacturers protocol. The PCR
properties of C. metallidurans CYP116B1, a novel, cat- fragment was then A-tailed and cloned into the plasmid
alytically self-sufcient P450PDOR fusion enzyme vector pGEM-T (Promega, Southampton, UK). CYP116B1
with oxidative activity towards thiocarbamate herbi- clones were veried by blue white screening following the
cides. From a biotechnological viewpoint, there is manufacturers instructions, and subsequently by DNA
growing interest in the exploitation of such single-com- sequencing of the entire genes (PNACL, University of
ponent P450redox partner fusion enzymes. This Leicester). The CYP116B1 gene was transferred from the
comes both from the perspective of using protein engi- pGEM-T cloning vector into the expression plasmid,
neering on high-activity fusion enzymes, such as P450 pET11a (Merck, Nottingham, UK). CYP116B1 was excised
BM3, to enable industrially useful oxidation chemistry from the pGEM-T clone by restriction digestion with AseI
[e.g. 2528], and also with respect to mimicking the and BlpI and the resultant gene fragment (2423 bp) was
resolved by electrophoresis through a 1% agarose gel and
efciency of the naturally fused P450 enzymes by gen-
puried as above. The pET11a vector was digested using
erating articial fusions of exogenous P450s to redox
BlpI and NdeI (the latter enzyme produces ends cohesive
partner systems. In the case of P450 BM3, fusions of
with those generated by AseI) and similarly puried. Resul-
exogenous P450s to the BM3 reductase have met with
tant colonies from ligation reactions were identied by
limited success, probably as a consequence of its
diagnostic restriction enzyme digestions and veried by
dimeric nature and the difculty in orientating reduc- DNA sequencing. Validated CYP116B1pET11a constructs
tase and P450 domains to achieve efcient intermolec- were used for CYP116B1 production.
ular electron transfer. However, CYP116B1, and Expression trials were performed with CYP116B1
probably the other CYP116B family members, are pET11a under a variety of conditions (varying growth
monomeric enzymes, and thus fusions of exogenous temperature, culture time, growth medium and isopropyl
P450s to their PDOR modules could provide a more thio-b-d-galactoside (IPTG) concentration) and in various
robust route to the generation of novel self-sufcient DE3 lysogen E. coli strains (Rosetta, Origami and
P450 oxidases for biotechnological applications [63]. HMS174, all from Merck). Optimal expression of
CYP116B1 was obtained in the Origami(DE3) CYP116B1
pET11a transformant with growth at 37 C in Luria
Bertani (LB) medium until the culture reached an optical
Materials and methods
density at 600 nm (D600) of 0.6. Thereafter, the culture tem-
perature was reduced to 18 C and growth continued until
Cloning, expression and purification of CYP116B1 an D600 of 0.8 was reached. At this point, IPTG was added
The CYP116B1 gene (Rmet_4932) was cloned from geno- to a nal concentration of 100 lm and growth was contin-
mic DNA prepared from a culture of C. metallidurans ued overnight (typically 1216 h). For protein preparation,
CH34 supplied by Professor Nigel Brown of the University a total culture volume of  6 L was used. Cells were

1688 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

harvested by centrifugation (30 min at 10 000 g in a Beck- potassium phosphate (pH 8.0), containing 200 mm KCl, on
man JLA 8.1000 rotor). Cell pellets were resuspended in a Dionex BioLC HPLC at 0.71 mLmin)1. Protein was
buffer A (50 mm Tris HCl, pH 7.2, containing 1 mm passed directly through a Wyatt EOS 18-angle laser pho-
EDTA) and then re-pelleted as before. Cells were resus- tometer coupled to a Wyatt Optilab rEX refractive index
pended in the same buffer ( 50 mL) and kept on ice. A detector. The molecular weight moments and concentra-
CompleteTM protease inhibitor tablet (Roche, West Sussex, tions of the resulting peaks were analysed using the Astra
UK) was added at this stage, along with the protease inhib- v.5.3.2 software (Wyatt Technology, Haverhill, UK). Sam-
itors phenylmethanesulfonyl uoride (1 mm) and benzami- ples were passed through the column twice to ensure that
dine hydrochloride (1 mm). These protease inhibitors were no aggregating species were included in the mass calcula-
added to all buffers from this stage in the purication. The tions.
cells were disrupted by passage (twice) through a French
Press (ThermoFisher, Loughborough, UK) at 950 psi, fol-
lowed by ultrasonication using a Bandelin Sonopuls EPR
GM2600 sonicator (6 20-s bursts at 50% full power, with EPR spectra were recorded at 10 K on a Bruker ER-300D
5 min cooling time on ice allowed between bursts). The series electromagnet and microwave source, interfaced to a
lysate was centrifuged at 46 000 g for 30 min at 4 C and Bruker EMX control unit and tted with a dual microwave
then the supernatant was transferred to a DEAE column cavity from the same supplier (model ER-4116DM). Spectra
pre-equilibrated in buffer A. CYP116B1-containing frac- were recorded at 2 mW microwave power at a modulation
tions were eluted using a gradient (500 mL) of buffer A to amplitude of 10 G. Spectra were recorded for both oxidized
buffer A + 500 mm KCl. CYP116B1 eluted at  350 mm and sodium dithionite-reduced forms of CYP116B1.
KCl. Fractions were pooled and transferred to a phenyl
Sepharose column pre-equilibrated with buffer B (buffer A
containing 0.75 m ammonium sulphate solution). The col- Determination of CYP116B1 flavin content
umn was washed extensively in buffer B and CYP116B1
was eluted using a gradient of buffer B to buffer A Flavin uorescence was recorded with sample excitation at
(500 mL). Fractions were pooled, concentrated by ultral- 450 nm, and emission data were collected at 500600 nm
tration (using Centriprep concentrators; Millipore, using a Cary Eclipse uorescence spectrometer in a 2 ml
Watford, UK) and dialysed extensively against buffer A to quartz uorescence cuvette of 1 cm pathlength. To remove
desalt the protein. The protein solution was then trans- avin cofactor from CYP116B1, enzyme samples (150 nm
ferred to a Q-Sepharose column and CYP116B1 was eluted 3 lm) were heated to 95 C for 3 min and aggregated protein
as described for the DEAE step above. CYP116B1 was was separated by centrifugation at 14 000 g for 10 min at
pure at this stage and was concentrated by ultraltration 4 C. The supernatant was retained and analysed. Fluores-
(as already described) to a nal concentration of  500 lm, cence data were collected for the protein in 100 mm potas-
and then dialysed into buffer A, containing 50% (v v) ster- sium phosphate (pH 7.7) and following acidication with
ile glycerol, before storage at )80 C. perchloric acid at pH 2.0, according to the method of Knight
and Hardy [66]. Fluorescence data at the different pH values
were compared with those for standard FAD and FMN solu-
UV-visible spectroscopic analysis and cofactor tions for identication and quantication of the CYP116B1-
identification bound avin, as described in our earlier review [55].
UV-visible spectra of CYP116B1 were collected using a
Cary UV-50 scanning spectrophotometer (Varian) with a Substrate- and ligand-binding studies
1 cm-pathlength quartz cuvette. Spectra were recorded for
oxidized, sodium dithionite-reduced and various substrate- Analysis of interactions of CYP116B1 with various ligands
(fatty acid and thiocarbamate) and ligand- (CO, NO, and candidate substrates was performed by spectral titration
cyanide, imidazole and substituted imidazole) bound forms, in buffer A, monitoring ligand-induced absorption changes
as described previously [35,38]. Reduction of CYP116B1 across the spectral region from 300 to 750 nm. Difference
and binding of CO was performed as described previously spectra were created at each point in the titrations by sub-
[64]. The haem concentration of CYP116B1 was determined traction of the spectrum for the ligand-free enzyme from
using the pyridine haemochromagen method [65]. that of the relevant ligand-bound form. Titrations were con-
tinued until no further spectral change could be discerned.
Maximal induced absorption change data were computed as
SEC-MALLS studies of CYP116B1 the difference in absorption at the peak and trough in the
MALLS analysis of CYP116B1 was coupled to SEC and difference spectra, and these values were plotted versus the
performed on a Supredex-200 24 30 gel-ltration column relevant ligand concentration applied. To determine Kd val-
(GE Healthcare, Chalfont St Giles, UK) run in 50 mm ues, these data were tted to either a standard hyperbolic

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1689
Biochemical analysis of CYP116B1 A. J. Warman et al.

function or to a quadratic function (for very tight-binding and single-wavelength data were recorded for up to 0.5 s.
ligands), as described previously [39]. Titrations were per- Data were tted accurately using a double exponential
formed at 25 C, with ligands added either in buffer A or function and the applied photophysics software.
(for fatty acids and thiocarbamate herbicides) in ethanol or
dimethylsulfoxide. The total volume of ligand addition was
< 1% of the nal volume in the assay. Volume changes as
Thermodynamic characterization
a result of ligand additions were accounted for by spectral A spectroelectrochemical redox titration of CYP116B1 was
correction before data analysis. performed, as described previously [58,59], with CYP116B1
at  8 lm. Spectral data were tted using Nernst functions
at wavelengths reecting transitions from the oxidized to
Steady-state and stopped-flow analysis of
the reduced forms of the bound cofactors, in order to
CYP116B1
derive midpoint reduction potentials for the CYP116B1
Steady-state kinetic analysis was carried out with haem iron Fe3+ Fe2+ transition and to establish the poten-
CYP116B1 in buffer A at 25 C on a dual-beam spectro- tial range in which the 2Fe-2S and FMN cofactors were
photometer (Jasco). Assays for reduction of electron accep- reduced.
tor substrates were performed with cytochrome c (horse
heart) and potassium ferricyanide. Reactions typically used
Product analysis
75100 nm CYP116B1, and the Km values for cytochrome c
(29.9 1.7 lm) and ferricyanide (103.6 20.5 lm) were In order to identify the products of vernolate and EPTC
established in preliminary assays using NADPH at a satu- oxidation by CYP116B1, turnover reactions were per-
rating concentration (200 lm). Thereafter, the kcat and Km formed according to the methods reported by Lawson et al.
values for NADPH and NADH were established in assays [47]. The 10-ml reactions contained CYP116B1 at a nal
in which the coenzyme concentration was varied for both concentration of 6 lm, 1 mm NADPH, 20 mm glucose and
NADH (02 mm) and NADPH (0200 lm), and the elec- 8.25 units of glucose dehydrogenase (an NADPH regenera-
tron acceptor concentration was maintained at a constant tion system), and 0.33 mm of either vernolate or EPTC.
near-saturating concentration (300 lm for cytochrome c The reaction constituents were all solubilized in buffer A,
and 1 mm for ferricyanide). Rates were determined at with the exception of vernolate and EPTC, which were pre-
550 nm for cytochrome c using De550 = 22640 m)1cm)1 pared in ethanol. The reactions were performed in 30-mL
[18] and at 420 nm for ferricyanide using De420 = glass vials that had been wrapped in foil to exclude light
1020 m)1 cm)1. Similar assays for P450 substrate-dependent and washed with methanol, dichloromethane and ddH2O to
NAD(P)H oxidation (De340 = 6210 m)1 cm)1) were carried remove potential contaminants. Control reactions lacking
out using various fatty acids (myristoleic acid, palmitoleic either CYP116B1 or substrate were also performed. Reac-
acid and arachidonic acid, typically 0500 lm) and using tion mixtures were incubated at 25 C and stirred continu-
the thiocarbamate herbicides EPTC and vernolate, at con- ously for 16 h. Reactions were terminated by the addition
centrations up to 2 mm, using previously described methods of 0.5 mL of 1.0 m HCl. Reaction products were extracted
[48]. Vernolate-dependent NADPH oxidation was also by adding 3 mL of HPLC-grade dichloromethane to each
measured across a range of CYP116B1 concentrations reaction. Organic phases were retained and aqueous phases
(10500 nm) in order to evaluate the effects on the specic were subjected to a second extraction with dichloromethane
catalytic rate. CO-trapping experiments (to examine the to maximize product recovery. Organic phases from each
inuence of dithiocarbamate substrate on the reduction of reaction were pooled and the residual aqueous phase was
CYP116B1 haem) were performed anaerobically in the removed by the addition of anhydrous magnesium sulphate.
absence and presence of vernolate, and in saturated CO Samples were ltered to remove any particulates before
buffer with 3 lm CYP116B1. Electron transfer was initiated being evaporated to dryness under vacuum. The resultant
by the addition of 10-fold excess NAD(P)H, and spectral pellets were resuspended in a minimal volume of HPLC-
data were recorded to monitor the formation of (reduced- grade methanol ( 50 lL) for LC-MS analysis. LC-MS
CO) P450-adducts. was performed using an Agilent 1100 series HPLC-MSD
Stopped-ow studies of CYP116B1 were carried out (Agilent Technologies, Waldbronn, Germany). Samples
under the same conditions as the steady-state assays. were resolved using a Vydac C8 column (21 150 mm) run
Assays were performed using both NADH and NADPH as at a ow rate of 1 mLmin)1 and maintained at 30 C. Fol-
reductants and at 465 nm (the wavelength of maximal over- lowing sample loading, an isocratic gradient of the mobile
all change between oxidized and PDOR domain-reduced phase (66% solvent A [100% HPLC-grade acetonitrile] and
CYP116B1). Data were recorded using an SX18MV 34% solvent B [0.2% acetic acid in ddH2O]) was run for
UV-visible stopped-ow instrument (Applied Photophysics, 28 min. Subsequently, the concentration of solvent B was
Leatherhead, UK). Enzyme (6.6 lm) was mixed with raised to 100% and the column was run for an additional
NAD(P)H (up to 500 lm) in the stopped-ow instrument 32 min. Detection of reaction products from the resolved

1690 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

samples was performed using the ion trap mass spectrome- 11 Wang M, Roberts DL, Paschke R, Shea TM, Masters
ter in the positive electrospray mode. BS & Kim JJ (1997) Three dimensional structure of
NADPH-cytochrome P450 reductase: prototype for
FMN- and FAD-containing enzymes. Proc Natl Acad
Acknowledgements Sci USA 94, 84118416.
The authors thank the Biotechnology and Biological 12 Munro AW, Girvan HM & McLean KJ (2007) Cyto-
Sciences Research Council UK (BBSRC) for funding chrome P450-redox partner fusion enzymes. Biochim
(PhD studentship to Ashley Warman). The authors Biophys Acta 1770, 345359.
also thank Dr Harriet Seward (University of Leicester) 13 Daiber A, Shoun H & Ullrich V (2005) Nitric oxide
for assistance with EPR data collection, Dr Tom reductase (P450nor) from Fusarium oxysporum. J Inorg
Biochem 99, 185193.
Jowitt and Mrs Marj Howard for assistance with
14 Shoji O, Fujishiro T, Nakajima H, Kim M, Nagano S,
MALLS collection and analysis and Mrs Marina
Shiro Y & Watanabe Y (2007) Hydrogen peroxide
Golovanova (University of Manchester) for excellent
dependent monooxygenations by tricking the substrate
technical support.
recognition of cytochrome P450BSbeta. Angew Chem
Int Ed 46, 36563659.
References 15 Hannemann F, Bichet A, Ewen KM & Bernhardt R
(2007) Cytochrome P450 systems biological variations
1 Denisov IG, Makris TM, Sligar SG & Schlichtling I of electron transport chains. Biochim Biophys Acta
(2005) Structure and chemistry of cytochrome P450. 1770, 330344.
Chem Rev 105, 22532277. 16 Warman AJ, Roitel O, Neeli R, Girvan HM, Seward HE,
2 Munro AW, Girvan HM & McLean KJ (2007) Varia- Murray SA, McLean KJ, Joyce MG, Toogood H, Holt
tions on a (t)heme novel mechanisms, redox partners RA et al. (2005) Flavocytochrome P450 BM3: an update
and catalytic functions in the cytochrome P450 super- on structure and mechanism of a biotechnologically
family. Nat Prod Rep 24, 585609. important enzyme. Biochem Soc Trans 33, 747753.
3 Rittle J & Green MT (2010) Cytochrome P450 com- 17 Narhi LO & Fulco AJ (1987) Identication and charac-
pound I: capture, characterization, and C-H bond acti- terization of two functional domains in cytochrome
vation kinetics. Science 330, 933937. P-450BM-3, a catalytically self-sufcient monooxygen-
4 Cryle MJ & De Voss JJ (2006) Is the ferric hydroperoxy ase induced by barbiturates in Bacillus megaterium.
species responsible for sulfur oxidation in cytochrome J Biol Chem 262, 66836690.
P450s? Angew Chem Int Ed 45, 82218223. 18 Noble MA, Miles CS, Chapman SK, Lysek DA, Mac-
5 Shaik S, de Visser SP & Kumar D (2004) One oxidant, Kay AC, Reid GA, Hanzlik RP & Munro AW (1999)
many pathways : a theoretical perspective of monooxy- Roles of key active-site residues in avocytochrome
genation mechanisms by cytochrome P450 enzymes. P450 BM3. Biochem J 339, 371379.
JBiol Inorg Chem 9, 661668. 19 Munro AW, Daff S, Coggins JR, Lindsay JG & Chap-
6 Guengerich FP (2001) Common and uncommon cyto- man SK (1996) Probing electron transfer in avocyto-
chrome P450 reactions related to metabolism and chem- chrome P450 BM3 and its component domains. Eur J
ical toxicity. Chem Res Toxicol 14, 611650. Biochem 239, 403409.
7 Rylott EL, Jackson RG, Sabbadin F, Seth-Smith HM, 20 Gustafsson MC, Roitel O, Marshall KR, Noble MA,
Edwards J, Chong CS, Strand SE, Grogan G & Bruce Chapman SK, Pessegueiro A, Fulco AJ, Cheesman
NC (2011) The explosive-degrading cytochrome P450 MR, von Wachenfeldt C & Munro AW (2004) Expres-
XplA: biochemistry, structural features and prospects for sion, purication and characterization of Bacillus subtil-
bioremediation. Biochim Biophys Acta 1814, 230236. is cytochromes P450 CYP102A2 and CYP102A3:
8 Song W-C, Baertschis SW, Boeglin WE, Harris TM & avocytochrome homologues of P450 BM3 from Bacil-
Brash AR (1993) Formation of epoxyalcohols by a lus megaterium. Biochemistry 43, 54745487.
puried allene oxide synthase. Implications for the 21 Kitazume T, Takaya N, Nakayama N & Shoun H
mechanism of allene oxide synthase. J Biol Chem 268, (2000) Fusarium oxysporum fatty-acid subterminal
62936298. hydroxylase (CYP505) is a membrane-bound eukaryotic
9 Poulos TL, Finzel BC & Howard AJ (1987) High-reso- counterpart of Bacillus megaterium cytochrome P450
lution crystal structure of cytochrome P450cam. J Mol BM3. J Biol Chem 275, 3973439740.
Biol 195, 687700. 22 Neeli R, Girvan HM, Lawrence A, Warren MJ, Leys
10 Sevrioukova IF, Poulos TL & Churbanova IY (2010) D, Scrutton NS & Munro AW (2005) The dimeric form
Crystal structure of the putidaredoxin reductase-puti- of avocytochrome P450 BM3 is catalytically functional
daredoxin electron transfer complex. J Biol Chem 285, as a fatty acid hydroxylase. FEBS Lett 579, 55825588.
1361613620.

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1691
Biochemical analysis of CYP116B1 A. J. Warman et al.

23 Kitazume T, Haines DC, Estabrook RW, Chen B & substitution mutants at the haem site of P450 BM3.
Peterson JA (2007) Obligatory intermolecular electron Biochem J 427, 455466.
transfer from FAD to FMN in dimeric P450 BM-3. 36 Ost TW, Miles CS, Munro AW, Murdoch J, Red GA
Biochemistry 46, 1189211901. & Chapman SK (2001) Phenylalanine 393 exerts ther-
24 Girvan HM, Dunford AJ, Neeli R, Ekanem IS, Wal- modynamic control over the heme of avocytochrome
tham TN, Joyce MG, Leys D, Curtis RA, Williams P, P450 BM3. Biochemistry 40, 1342113429.
Fisher K et al. (2011) Flavocytochrome P450 BM3 37 Perera R, Sono M, Sigman JA, Pster TD, Lu Y &
mutant W1046A is a NADH-dependent fatty acid Dawson JH (2003) Neutral thiol as a proximal ligand
hydroxylase: implications for the mechanism of electron to ferrous heme iron: implications for heme proteins
transfer in the P450 BM3 dimer. Arch Biochem Biophys that lose cysteine thiolate ligation on reduction. Proc
507, 7585. Natl Acad Sci USA 100, 36413646.
25 Ost TW, Miles CS, Murdoch J, Cheung Y, Reid GA, 38 McLean KJ, Warman AJ, Seward HE, Marshall KR,
Chapman SK & Munro AW (2000) Rational re-design Girvan HM, Cheesman MR, Waterman MR & Munro
of the substrate binding site of avocytochrome P450 AW (2006) Biophysical characterization of the sterol
BM3. FEBS Lett 486, 173177. demethylase P450 from Mycobacterium tuberculosis, its
26 Peters MW, Meinhold P, Glieder A & Arnold FH cognate ferredoxin and their interactions. Biochemistry
(2005) Regio- and enantioselective alkane hydroxylation 45, 84278443.
with engineered cytochromes P450 BM-3. J Am Chem 39 McLean KJ, Cheesman MR, Rivers SL, Richmond A,
Soc 125, 1344213450. Leys D, Chapman SK, Reid GA, Price NC, Kelly SM,
27 Kille S, Zilly FE, Acevedo JP & Reetz MT (2011) Clarkson J et al. (2002) Expression and purication of
Regio- and stereoselectivity of P450-catalysed hydroxyl- the cytochrome P450 CYP121 from Mycobacterium
ation of steroids controlled by laboratory evolution. tuberculosis. J Inorg Biochem 91, 527541.
Nat Chem 14, 738743. 40 Quaroni LG, Seward HE, McLean KJ, Girvan HM, Ost
28 Whitehouse CJ, Bell SG & Wong LL (2012) P450 BM3 TW, Noble MA, Kelly SM, Price NC, Cheesman MR,
(CYP102A1): connecting the dots. Chem Soc Rev 41, Smith WE et al. (2004) Interaction of nitric oxide with
12181260. cytochrome P450 BM3. Biochemistry 43, 1641616431.
29 De Mot R & Parret AH (2002) A novel class of self-suf- 41 Sligar SG & Gunsalus IC (1976) A thermodynamic
cient cytochrome P450 monooxygenases in prokary- mode of regulation: modulation of redox equilbria in
otes. Trends Microbiol 10, 502508. camphor monoxygenase. Proc Natl Acad Sci USA 73,
30 Gassner GT, Ludwig ML, Gatti DL, Correll CC & Bal- 10781082.
lou DP (1995) Structure and mechanism of the iron-sul- 42 Miles JS, Munro AW, Rospendowski BN, Smith WE,
fur avoprotein phthalate dioxygenase reductase. McKnight JE & Thomson AJ (1992) Domains of the
FASEB J 9, 14111418. catalytically self-sufcient cytochrome P-450 BM-3.
31 Batie CJ & Ballou DP (1990) Phthalate dioxygenase. Genetic construction, overexpression, purication and
Methods Enzymol 188, 6170. spectroscopic characterization. Biochem J 288, 503509.
32 Roberts GA, Celik A, Hunter DJ, Ost TW, White JH, 43 Noble MA, Quaroni L, Chumanov GD, Turner KJ,
Chapman SK, Turner NJ & Flitsch SL (2003) A self- Chapman SK, Hanzlik RP & Munro AW (1998)
sufcient cytochrome P450 with a primary structural Imidazolyl carboxylic acids as mechanistic probes of
organization that includes a avin domain and a avocytochrome P-450 BM3. Biochemistry 37, 15799
[2Fe-2S] redox center. J Biol Chem 278, 4891448920. 15807.
33 Hunter DJ, Roberts GA, Ost TW, White JH, Muller S, 44 Morrison JF (1969) Kinetics of the reversible inhibition
Turner NJ, Flitsch SL & Chapman SK (2005) Analysis of enzyme-catalysed reactions by tight-binding inhibi-
of the domain properties of the novel cytochrome P450 tors. Biochim Biophys Acta 185, 269286.
RhF. FEBS Lett 579, 22152220. 45 Alterman MA, Chaurasia CS, Lu P, Hardwick JP &
34 Nagy I, Schoofs G, Compornolle F, Proost P, Vander- Hanzlik RP (1995) Fatty acid discrimination and omega
leyden J & de Mot R (1995) Degradation of the thioc- hydroxylation by cytochrome P450 4A1 and a cyto-
arbamate herbicide EPTC (S-ethyl chrome P450 4A1 NADPH-P450 reductase fusion pro-
dipropylcarbamothioate) and biosafening by Rhodococ- tein. Arch Biochem Biophys 320, 289296.
cus sp. Strain NI86 21 involve an inducible cytochrome 46 Siddhanta U, Presta A, Fan B, Wolan D, Rousseau DL
P-450 system and aldehyde dehydrogenase. J Bacteriol & Stuehr DJ (1998) Domain swapping in inducible
177, 676687. nitric-oxide synthase. Electron transfer occurs between
35 Girvan HM, Levy CW, Williams P, Fisher K, avin and heme groups on adjacent subunits in the
Cheesman MR, Rigby SE & Munro AW (2010) Gluta- dimer. J Biol Chem 273, 1895018958.
mate-haem ester bond formation is disfavoured in avo- 47 Lawson RJ, Leys D, Sutcliffe MJ, Kemp CA, Chees-
cytochrome P450 BM3: characterization of glutamate man MR, Smith SJ, Clarkson J, Smith WE, Haq I,

1692 FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS
A. J. Warman et al. Biochemical analysis of CYP116B1

Perkins JB et al. (2004) Thermodynamic and biophysi- tion reactions and qualitative metabolite predictions
cal characterisation of cytochrome P450 BioI from using a combined protein and pharmacophore model
Bacillus subtilis. Biochemistry 43, 1241012426. for CYP2D6. J Med Chem 42, 40624070.
48 Girvan HM, Marshall KR, Lawson RJ, Leys D, Joyce 61 Kobayashi K, Yamamoto T, Chiba K, Tani M, Shi-
MG, Clarkson J, Smith WE, Cheesman MR & Munro mada N, Ishizaki T & Kuroiwa Y (1998) Human bupr-
AW (2004) Flavocytochrome P450 BM3 mutant A264E enorphine N-dealkylation is catalyzed by cytochrome
undergoes substrate- dependent formation of a novel P450 3A4. Drug Metab Dispos 26, 818821.
heme iron ligand set. J Biol Chem 279, 2327423286. 62 Janssen PJ, Van-Houdt R, Moors H, Monsieurs P,
49 Lipscomb JD (1980) Electron paramagnetic resonance Morin N, Michaux A, Benotmane MA, Leys N,
detectable states of cytochrome P-450cam. Biochemistry Vallaeys T, Lapidus A et al. (2010) The complete
19, 35903599. genome sequence of Cupriavidus metallidurans CH34,
50 Evans EH, Rush JD, Johnson CE & Evans MCW a master survivalist in harsh and anthropogenic
(1979) Mossbauer spectra of Photosystem-I reaction environments. PLoS One 5, e10433.
centres from the blue-green alga Chlorogloea fritschii. 63 Sabbadin F, Hyde R, Robin A, Hilgarth E-A, Delenn
Biochem J 182, 861865. M, Flitsch S, Turner N, Grogan G & Bruce NC (2010)
51 Sellers VM, Wang K, Johnson MK & Dailey HA LICRED: a versatile drop-in vector for rapid genera-
(1998) Evidence that the fourth ligand to the [2Fe-2S] tion of redox-self-sufcient cytochrome P450s. Chem-
cluster in animal ferrochelatase is a cysteine. J Biol BioChem 11, 987994.
Chem 273, 2231122316. 64 Omura T & Sato R (1964) The carbon monoxide-bind-
52 More C, Belle V, Asso M, Fournel A, Roger G, Guigliar- ing pigment of liver microsomes. I. Evidence for its
elli B & Bertrand P (1999) EPR spectroscopy: a powerful hemoprotein nature. J Biol Chem 239, 23702378.
technique for the structural and functional investigation 65 Berry EA & Trumpower BL (1987) Simultaneous deter-
of metalloproteins. Biospectroscopy 5(S5), S3S18. mination of hemes a, b, and c from pyridine hemo-
53 Batie CJ, Lahaie E & Ballou DP (1987) Purication chrome spectra. Anal Biochem 161, 115.
and characterization of phthalate oxygenase and phtha- 66 Knight E Jr & Hardy RF (1966) Isolation and charac-
late oxygenase reductase from Pseudomonas cepacia. teristics of avodoxin from nitrogen-xing Clostridium
J Biol Chem 262, 15101518. pasteurianum. J Biol Chem 241, 27522756.
54 Voss I, Goss Y, Murozuka E, Altmann B, McLean KJ,
Rigby SE, Munro AW, Scheibe R, Hase T & Hanke
GT (2011) FdC1, a novel ferredoxin protein capable of
Supporting information
alternative electron partitioning, increases in conditions The following supplementary material is available:
of acceptor limitation at photosystem I. J Biol Chem Fig. S1. Analysis of the CYP116B1 heme cofactor
286, 5059. properties by resonance Raman spectroscopy
55 Munro AW & Noble MA (1999) Fluorescence analysis Fig. S2. Vernolate and EPTC LC-MS controls
of avoproteins. Methods Mol Biol 131, 2548. Fig. S3. Product analysis of CYP116B1 EPTC turn-
56 Driscoll MD, McLean KJ, Levy C, Mast N, Pikuleva
over by LC-MS
IA, Late P, Rigby SE, Leys D & Munro AW (2010)
Fig. S4. Amino acid alignment of the structurally
Structural and biochemical characterization of Myco-
resolved Burkholderia cepacia phthalate dioxygenase
bacterium tuberculosis CYP142: evidence for multiple
reductase (B. cepacia PDOR) and the reductase domain
cholesterol 27-hydroxylase activities in a human patho-
of Cupriavidus metallidurans CYP116B1 (CYP116B1-
gen. J Biol Chem 285, 3827038282.
57 Honeychurch MJ, Hill AO & Wong LL (1999) The ther-
RED). Amino acid alignment of the reductase domain
modynamics and kinetics of electron transfer in the cyto- of CYP116B1 (RED) with the structurally resolved
chrome P450cam enzyme system. FEBS Lett 451, 351353. Burkholderia cepacia phthalate dioxygenase reductase
58 Daff SN, Chapman SK, Turner KL, Holt RA, Govind- (B. cepacia PDOR) and key amino acids involved in
araj S, Poulos TL & Munro AW (1997) Redox control FMN binding and pyridine cofactor binding specicity.
of the catalytic cycle of avocytochrome P450 BM3. This supplementary material can be found in the
Biochemistry 36, 1381613823. online version of this article.
59 Dutton PL (1978) Redox potentiometry: determination Please note: As a service to our authors and readers,
of midpoint potentials of oxidation-reduction compo- this journal provides supporting information supplied
nents of biological electron-transfer systems. Methods by the authors. Such materials are peer-reviewed and
Enzymol 54, 411435. may be reorganized for online delivery, but are not
60 de Groot MJ, Ackland MJ, Horne V, Alex AA & Jones copy-edited or typeset. Technical support issues arising
BC (1999) A novel approach to predicting P450 medi- from supporting information (other than missing les)
ated drug metabolism. CYP2D6 catalyzed N-dealkyla- should be addressed to the authors.

FEBS Journal 279 (2012) 16751693 2012 The Authors Journal compilation 2012 FEBS 1693

You might also like