theoretical study of the catalytic mechanism of catechol

14
ORIGINAL PAPER Theoretical study of the catalytic mechanism of catechol oxidase Mireia Gu ¨ell Per E. M. Siegbahn Received: 20 June 2007 / Accepted: 16 August 2007 / Published online: 20 September 2007 Ó SBIC 2007 Abstract The mechanism for the oxidation of catechol by catechol oxidase has been studied using B3LYP hybrid density functional theory. On the basis of the X-ray structure of the enzyme, the molecular system investigated includes the first-shell protein ligands of the two metal centers as well as the second-shell ligand Cys92. The cycle starts out with the oxidized, open-shell singlet complex with oxidation states Cu 2 (II,II) with a l-g 2 :g 2 bridging peroxide, as suggested experimentally, which is obtained from the oxidation of Cu 2 (I,I) by dioxygen. The substrate of each half-reaction is a catechol molecule approaching the dicopper complex: the first half-reaction involves Cu(I) oxidation by peroxide and the second one Cu(II) reduction. The quantitative potential energy profile of the reaction is discussed in connection with experimental data. Since no protons leave or enter the active site during the catalytic cycle, no external base is required. Unlike the previous density functional theory study, the dicopper complex has a charge of +2. Keywords Catechol oxidase Á Copper enzymes Á O 2 cleavage Á Hybrid density functional theory Introduction Proteins containing copper ions at their active sites are usually involved as redox catalysts in a wide range of biological processes. Type-3 active-site copper proteins contain a dicopper core in which both copper ions are surrounded by three nitrogen donor atoms from histidine residues [1, 2]. A characteristic feature of the proteins with this active site is their ability to reversibly bind dioxygen at ambient conditions. The Cu(II) ions in the oxy state of these proteins are strongly antiferromagnetically coupled, leading to electron paramagnetic resonance (EPR) silent behavior. This class is represented by three proteins: hemocyanin, catechol oxidase and tyrosinase. Proteins with type-3 copper centers can serve either as oxygenase/oxidase enzymes or as dioxygen transport pro- teins [2]. An example of an oxygen carrier is hemocyanin. Hemocyanins can be divided into two classes depending on their biological source: the arthropodan and the molluscan hemocyanins [35]. Catechol oxidase, which is also known as o-diphenol oxidase, catalyzes exclusively the oxidation of catechols (i.e., o-diphenols) to the corresponding o-quinones (called catecholase activity) (Fig. 1)[6]. The resulting highly reactive quinones autopolymerize to form brown poly- phenolic catechol melanins, a process thought to protect damaged plants from pathogens or insects [7]. The rate for catechol conversion in sweet potatoes has been measured to be 2.3 · 10 3 /s [8], corresponding to a rate-limiting free- energy barrier of around 13 kcal/mol. In contrast to catechol oxidase, the strongly related tryrosinase shows additional monooxygenase activity (Fig. 1). This so-called cresolase activity enables the enzyme to accept also monophenols (like tyrosine and cresol). Catechol oxidases are found in plant tissues and in M. Gu ¨ell (&) Institut de Quı ´mica Computacional, Universitat de Girona, Campus de Montilivi, 17071 Girona, Spain e-mail: [email protected] P. E. M. Siegbahn Department of Biochemistry and Biophysics, Stockholm University, 106 91 Stockholm, Sweden 123 J Biol Inorg Chem (2007) 12:1251–1264 DOI 10.1007/s00775-007-0293-z

Upload: others

Post on 12-Jul-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Theoretical study of the catalytic mechanism of catechol

ORIGINAL PAPER

Theoretical study of the catalytic mechanism of catechol oxidase

Mireia Guell Æ Per E. M. Siegbahn

Received: 20 June 2007 / Accepted: 16 August 2007 / Published online: 20 September 2007

� SBIC 2007

Abstract The mechanism for the oxidation of catechol

by catechol oxidase has been studied using B3LYP hybrid

density functional theory. On the basis of the X-ray

structure of the enzyme, the molecular system investigated

includes the first-shell protein ligands of the two metal

centers as well as the second-shell ligand Cys92. The cycle

starts out with the oxidized, open-shell singlet complex

with oxidation states Cu2(II,II) with a l-g2:g2 bridging

peroxide, as suggested experimentally, which is obtained

from the oxidation of Cu2(I,I) by dioxygen. The substrate

of each half-reaction is a catechol molecule approaching

the dicopper complex: the first half-reaction involves Cu(I)

oxidation by peroxide and the second one Cu(II) reduction.

The quantitative potential energy profile of the reaction is

discussed in connection with experimental data. Since no

protons leave or enter the active site during the catalytic

cycle, no external base is required. Unlike the previous

density functional theory study, the dicopper complex has a

charge of +2.

Keywords Catechol oxidase � Copper enzymes �O2 cleavage � Hybrid density functional theory

Introduction

Proteins containing copper ions at their active sites are

usually involved as redox catalysts in a wide range of

biological processes. Type-3 active-site copper proteins

contain a dicopper core in which both copper ions are

surrounded by three nitrogen donor atoms from histidine

residues [1, 2]. A characteristic feature of the proteins with

this active site is their ability to reversibly bind dioxygen at

ambient conditions. The Cu(II) ions in the oxy state of

these proteins are strongly antiferromagnetically coupled,

leading to electron paramagnetic resonance (EPR) silent

behavior. This class is represented by three proteins:

hemocyanin, catechol oxidase and tyrosinase.

Proteins with type-3 copper centers can serve either as

oxygenase/oxidase enzymes or as dioxygen transport pro-

teins [2]. An example of an oxygen carrier is hemocyanin.

Hemocyanins can be divided into two classes depending on

their biological source: the arthropodan and the molluscan

hemocyanins [3–5].

Catechol oxidase, which is also known as o-diphenol

oxidase, catalyzes exclusively the oxidation of catechols

(i.e., o-diphenols) to the corresponding o-quinones (called

catecholase activity) (Fig. 1) [6]. The resulting highly

reactive quinones autopolymerize to form brown poly-

phenolic catechol melanins, a process thought to protect

damaged plants from pathogens or insects [7]. The rate for

catechol conversion in sweet potatoes has been measured

to be 2.3 · 103/s [8], corresponding to a rate-limiting free-

energy barrier of around 13 kcal/mol.

In contrast to catechol oxidase, the strongly related

tryrosinase shows additional monooxygenase activity

(Fig. 1). This so-called cresolase activity enables the

enzyme to accept also monophenols (like tyrosine and

cresol). Catechol oxidases are found in plant tissues and in

M. Guell (&)

Institut de Quımica Computacional,

Universitat de Girona,

Campus de Montilivi,

17071 Girona, Spain

e-mail: [email protected]

P. E. M. Siegbahn

Department of Biochemistry and Biophysics,

Stockholm University,

106 91 Stockholm, Sweden

123

J Biol Inorg Chem (2007) 12:1251–1264

DOI 10.1007/s00775-007-0293-z

Page 2: Theoretical study of the catalytic mechanism of catechol

some insects and crustaceans, whereas tryrosinases can be

isolated from a broader variety of plants, fungi, bacteria,

mammalians, crustaceans and insects [2]. The differentia-

tion between catechol oxidase and tryrosinase is not

rigorous, as some catechol oxidases also show weak

monooxygenase activity. However, these catechol oxidases

often do not accept tyrosine as a substrate [9, 10].

Klabunde et al. [11] isolated the first crystal structures of

the catechol oxidase from Ipomoea batatas (sweet potato)

in three catalytic states: the native met [Cu(II)Cu(II)] state,

the reduced deoxy [Cu(I)Cu(I)] form, and in the complex

with the inhibitor phenylthiourea. This enzyme was found

to be monomeric and ellipsoidal in shape (Fig. 2). Its

secondary structure is primarily a-helical with the core of

the enzyme formed by a four-helix bundle. The helical

bundle accommodates the catalytic dinuclear copper cen-

ter, where each of the two copper ions is coordinated by

three histidine residues. One of the key features of the

catechol oxidase active site is an unusual thioether bridge

between Cys92 and His109, one of the ligands of CuA

(Fig. 3). Apart from the geometrical constraints added to

the CuA site, no function of the chemistry performed by the

enzyme has been ascribed to this covalent bridge.

In the native met state, the two copper ions are 2.9 A

apart. In addition to six histidine residues, a bridging sol-

vent molecule, most likely hydroxide anion, was refined in

close proximity to the two metal centers (CuA–O 1.9 A,

CuB–O 1.8 A), completing the coordination sphere of the

copper ions to a trigonal pyramid. EPR data reveal a strong

antiferromagnetic coupling between the copper ions, in line

with a solvent molecule bridging two metal centers, as

found in the crystal structure.

Upon reduction of the Cu(II) ions to the +1 oxidation

state, the distance between them increases to 4.4 A, while

the histidine residues move only slightly, and no significant

change was observed for other residues of the protein [11].

On the basis of the residual electron density maps, a water

molecule was positioned at a distance of 2.2 A from the

CuA atom. Thus, the coordination sphere around the CuA

ion is a distorted trigonal pyramid, with three nitrogen

atoms from the histidine residues forming a basal plane,

while the coordination sphere around the CuB ion is best

described as square planar with one missing coordination

site.

When phenylthiourea binds to catechol oxidase, it

replaces the hydroxo bridge, present in the met form. The

sulfur atom of phenylthiourea is coordinated to both Cu(II)

centers, increasing the distance between them to 4.2 A. The

amide nitrogen interacts weakly with the CuB center (Cu–N

distance of 2.6 A), completing its square-pyramidal

geometry.

The oxy form of catechol oxidase can be obtained by

treating the met form of the enzyme with hydrogen per-

oxide [12].

As previously mentioned, catechol oxidase catalyzes the

oxidation of catechols to the respective quinones through a

four-electron reduction of dioxygen to water. Krebs and

coworkers proposed a mechanism for the catalytic process,

based on biochemical and spectroscopic data [2, 12, 13], as

well as structural data [11], which is depicted in Fig. 4

[14]. The catalytic cycle begins with the met form of cat-

echol oxidase, which is the resting form of the enzyme. The

dicopper(II) center of the met form reacts with 1 equivalent

of catechol, leading to the formation of quinone and to the

reduced deoxy dicopper(I) state. This step is supported by

the observation that stoichiometric amounts of the quinone

product form immediately after the addition of catechol,

even in the absence of dioxygen [11, 14]. On the basis of

the structure of catechol oxidase with the bound inhibitor

phenylthiourea, the monodentate binding of the substrate to

the CuB center has been proposed. Afterwards, dioxygen

binds to the dicopper(I) active site, replacing the solvent

molecule bonded to CuA in the reduced enzyme. UV–vis

spectroscopy and Raman data suggested that dioxygen

OH OH

OH

O

O

1/2 O2 1/2 O2 H2O

cresolase activity

(tyrosinase)

catecholase activity

(tyrosinase andcatechol oxidase)

Fig. 1 Reaction pathway of the

oxygenation and oxidation

catalyzed by tyrosinase and

catechol oxidase

Fig. 2 Tertiary structure of the oxidized catechol oxidase (PDB code

1BT3) [11]. Cu(II) ions are given as yellow spheres and important

active-site residues are shown. The picture was generated using the

VMD 1.8.5 molecular visualization program. See Fig. 3 for a more

detailed picture of the active site

1252 J Biol Inorg Chem (2007) 12:1251–1264

123

Page 3: Theoretical study of the catalytic mechanism of catechol

binds in a bridging side-on l-g2:g2 binding mode with a

copper–copper separation of 3.8 A, as determined by

extended X-ray absorption fine structure spectroscopy for

the oxy state [12]. The observed binding mode of phenyl-

thiourea and the modeled catechol-binding mode suggest

that simultaneous binding of catechol and dioxygen is

possible. In this model, CuB is six-coordinated with a

tetragonal planar coordination by His240, His244 and the

dioxygen molecule in the basal plane. The CuA site retains

the tetragonal pyramidal geometry with dioxygen, His88

and His118 in the equatorial positions, His109 in an axial

position and a vacant sixth coordination site. In this pro-

posed ternary catechol oxidase–O22––catechol complex, two

electrons can be transferred from the substrate to the

Fig. 3 Active site of the

oxidized catechol oxidase (PDB

code 1BT3) [11]

CuIAHis

His

His

BCuI His

His

His

H2O + H+

OH

OH

CuIIHis

His

His O

OCuII His

His

His

O

O

H+

OH

OH

O

O

H2O +

OH2

O2 +

O

HO

CuIIAHis

His

His

OH

BCuII His

His

His

O

HO

2H+

CuIIAHis

His

His

OH

BCuII His

His

His

met state

deoxy state

oxy state

A B

Fig. 4 Catalytic cycle of

catechol oxidase as proposed

by Klabunde et al. [11]

J Biol Inorg Chem (2007) 12:1251–1264 1253

123

Page 4: Theoretical study of the catalytic mechanism of catechol

peroxide, followed by cleavage of the O–O bond, loss of

water and the formation of the quinone product, together

with the restoration of the met state, completing the cata-

lytic cycle.

A very similar catalytic mechanism had been proposed

earlier by Solomon et al. [2] for the catecholase activity of

the structurally related type-3 protein tyrosinase (Fig. 5).

The cycle starts from the oxy and met states. A diphenol

substrate binds to the met state (for example), followed by

the oxidation of the substrate to the first quinone and the

formation of the reduced state of the enzyme. Binding of

dioxygen leads to the oxy state, which is subsequently

attacked by the second diphenol molecule. Oxidation to the

second quinone forms the met state again and closes the

catalytic cycle.

The main difference between the two mechanistic pro-

posals involves the binding mode of the substrate to the

dicopper(II) core: whereas a monodentate asymmetric

coordination of the substrate was proposed by Klabunde

et al. [11, 14], a simultaneous coordination of the substrate

to both copper centers in a bidentate bridging fashion was

suggested by Solomon et al. [2].

Also of interest in the present context is the study by

Granata et al. [15] of the activity of tyrosinase. They

concluded that the interpretation of the diphenolase reac-

tion is complicated by the fact that cleavage of the peroxide

bond involves coupled electron and proton transfers from

the substrate [16–18]. For this reason, it is difficult to

predict the sequence of events, since it is likely that protein

residues at the active site, as well as ionizable groups on

the substrate, participate as proton storage/delivery devi-

ces. Compared with the monophenolase reaction, proton

transfers from the diphenol are greatly facilitated by its

stronger acidity. We can thus formulate the evolution of the

ternary complex according to the two alternative routes

represented in Fig. 6, depending on whether proton transfer

from bound catecholate occurs before or after O–O bond

cleavage.

Energetic considerations suggest that the two pathways

should be essentially equivalent, since proton transfers

involve a small energy barrier and heterolytic cleavages of

peroxo or hydroperoxo O–O bonds are expected to be

similar in energy [18, 19].

A rather different mechanism of the catalytic cycle has

also been proposed [18]. That mechanism was built on the

growing number of theoretical [21] and experimental [22,

23] studies suggesting that the active site of an enzyme

should not change its charge during the catalytic cycle. In

the mechanism, proposed by Klabunde et al. [11, 14], the

charge of the active site changes from +1 to +3. This in turn

implies the availability of several external nearby bases,

which could store protons, released during the cycle.

However, the X-ray crystal structure does not reveal the

presence of such candidates in the region of the active site.

Another fact of importance in this context is that for the

similar enzyme tyrosinase, it is required for the conserva-

tion of charge (without external base) that the dicopper

complex has a charge of +1. From the structural similarity

between tyrosinase and catechol oxidase, the charge of the

dicopper complex was therefore chosen as +1 also for

CuIIHis

His

His O

O

CuII His

His

His

CuIHis

His

His

CuI His

His

His

OH

OH

CuIIHis

His

His O

OCuII His

His

His

O

O

OH

OH

O

O

O O CuIIHis

His

His

OH

CuII His

His

His

O O

CuIIHis

His

His

OH

CuII His

His

His 2H+

3H+

2H+

H2O +

+ H2O

H+

oxy state deoxy stateO2

met state

Fig. 5 Catalytic cycle for the

oxidation of o-diphenols to o-

quinones by tyrosinase proposed

by Solomon and coworkers. [2]

1254 J Biol Inorg Chem (2007) 12:1251–1264

123

Page 5: Theoretical study of the catalytic mechanism of catechol

catechol oxidase. In that proposal, the catalytic cycle thus

starts with a bridging hydroxide ligand between two Cu(I)

ions (Fig. 7). In the first stage, catechol binds to the deoxy

form, transferring the proton to the bridging hydroxide,

with the subsequent generation of a bridging water mole-

cule between the metal centers. Afterwards, dioxygen

displaces the water molecule, binding as a superoxide

radical anion and resulting in the formation of the mixed-

valence dicopper(II,I) species (step a). The superoxide

subsequently abstracts a hydrogen atom (a proton and an

electron) from the bound substrate (step b). To release the

quinone molecule, an electron is then transferred from the

semiquinone radical to the Cu(II) ion, leading to the res-

toration of the dicopper(I) state (step c). The next step

involves the cleavage of the O–O bond, which is accom-

panied by the transfer of two protons from the substrate and

two electrons (from one of the Cu(I) ions and the substrate)

to the peroxide moiety (step d). Altogether this leads to a

product which can be best described as a Cu(II)Cu(I)

species with a semiquinone radical anion. The second

electron transfer from the semiquinone radical to the Cu(II)

center leads to the restoration of the initial hydroxo-bridged

dicopper(I) form (step e). However, it should be noted that

at the present moment this mechanism lacks support from

experimental findings. In particular, the existence of a

bridging l-1,1-superoxide radical anion [18] has never

been reported in the literature for any copper species.

On the other hand, bioinspired synthetic catalysts con-

stitute a valuable tool to explore the reaction mechanisms

that enzymes use to perform their chemistry. The recent

experimental structure determination of catechol oxidase

[11] has encouraged an extensive investigation on model

compounds of this enzyme [25]. The approaches used to

study the mechanism for catecholase activity of Cu(II)

complexes can be divided into four major groups: sub-

strate-binding studies [17, 26–31], structure–activity

relationship studies [31–36], kinetic studies on the catalytic

reaction [26, 28, 33, 37–39] and studies of stochiometric

oxidation of catechol substrates by peroxo–dicopper and

oxo–dicopper complexes [26, 27, 30, 38, 40, 41]. We are

presently working on the study of the catalytic mechanism

of biomimetic complexes of catechol oxidase and we will

return to this point in another paper.

In the present study, it was decided to leave the analogy

to tyrosinase and instead start with a dicopper(II) complex

with charge +2, to investigate if a reasonable reaction

pathway could still be found without adding or releasing

protons from the active complex.

Computational details

All the calculations were done using the B3LYP [42–45]

hybrid density functional. Open shell systems were treated

using unrestricted density functional theory. Geometry

optimizations were performed using a standard valence

LACVP basis set as implemented in the Jaguar 5.5 program

[46]. For the first- and second-row elements, LACVP

implies a 6-31G double-n basis set. For the copper atoms,

LACVP uses a nonrelativistic effective core potential [47],

where the valence part is essentially of double-n quality.

Local minima were optimized using the Jaguar 5.5 program

[46]. The optimizations of the transition states and the

frequency calculations were performed using the Gaussian

03 program [48]. The zero-point vibrational energies were

included in our theoretical results, although no thermal

effects were considered because the use of nonoptimized

frozen coordinates (see ‘‘Chemical models’’) does not give

reliable entropy effects. Accurate single-point energies

were obtained using the cc-pvtz(-F) basis set. For copper

atoms, an effective core potential was used. The sur-

rounding protein was treated with a self-consistent reaction

field method, using a Poisson–Boltzmann solver [49, 50] as

implemented in Jaguar 5.5. The dielectric constant of the

CuO

OCu

O OH

CuO

OCu

O OH

CuO

OCu

O

CuO

OCu

O OH

CuO

OCu

O

CuO

OCu

O OH

CuO

OCu

O OH2H+

H2O

2H+

H2O

(a)

(b)H

H

Fig. 6 The two possible ways

in which the ternary complex

can evolve according to Battaini

and coworkers. [20]

J Biol Inorg Chem (2007) 12:1251–1264 1255

123

Page 6: Theoretical study of the catalytic mechanism of catechol

homogeneous dielectric medium was set equal to 4.0, in

line with previous modeling of enzymes [51]. The probe

radius was set to 2.50 A. No geometry optimizations

including the dielectric continuum were made since the

calculated dielectric effects were found to be quite small.

The accuracy of the B3LYP functional has been tested in

the extended G3 benchmark set [52], which consists of

enthalpies of formation, ionization potentials, electron

affinities and proton affinities for molecules containing

first- and second-row atoms. The B3LYP functional gives

an average error of 4.3 kcal/mol [52] for 376 different

entries. Owing to the lack of accurate experimental data for

transition metals, there are only a few benchmark tests.

They indicate that for normal metal–ligand bond strengths

the errors are in the range 3–5 kcal/mol [52, 53]. Different

aspects of modeling enzyme active sites have been

reviewed [51, 54, 55].

Results and discussion

Chemical models

The present modeling of the catechol oxidase active site is

based on the X-ray structure of the oxidized form from

sweet potato (PDB code 1BT3; Figs. 2, 3) [11]. In view

of the mechanistic proposals (see ‘‘Introduction’’), the

coppers and the first-shell histidine ligands are the key

chemical species needed to be accurately modeled. Since a

covalent link is present between one of these histidines and

Cys92, this cysteine is also included in the model. The

structural histidines were modeled by imidazoles and the

cysteine was modeled by SCH3. This type of modeling has

been shown to be appropriate on the basis of previous

density functional theory studies of various enzymes [21].

To reproduce the protein strain and a realistic positioning

of the different chemical units, specific restrictions on some

nuclear coordinates were applied. This is important in

modeling fragments not directly bound to the metal [56,

57].

The present description of the mechanism of catechol

oxidase will start from the oxidized, open-shell singlet

complex with oxidation states Cu2(II,II), where the OH was

replaced by a l-g2:g2 bridging peroxide in order to obtain

the intermediate l-g2:g2 bridging peroxide suggested

experimentally [58]. This structure is obtained owing to the

oxidation of Cu2(I,I) by dioxygen to Cu2(II,II). The total

charge for the complex is +2. With this choice of charge

state, the protons needed at later stages in the cycle are

available on the metal complex. This is important since

there is no structural evidence for any external proton

donors or acceptors in the neighborhood of the copper

complex, and it is general experience that active sites

deeply buried in enzymes should not change their charge

CuI CuI

His

His

His

His

His

His

O

O

HH

O

OH

CuII CuI

His

His

His

His

His

HisO

OH

O

CuII CuI

His

His

His

His

His

HisO

O

OH

CuI CuI

His

His

His

His

His

His

O

OH

H

OH

CuI CuII

His

His

His

His

His

His

O

OH

H

OH

OH

OH

OH

OH

O

O

step astep e

step d step b

step c

H2O +

H2O

O2

O

O

OO

O

Fig. 7 The mechanism of the

catalytic cycle of catechol

oxidase proposed previously

based on DFT calculations [24]

1256 J Biol Inorg Chem (2007) 12:1251–1264

123

Page 7: Theoretical study of the catalytic mechanism of catechol

states [21–23]. The triplet spin state was also considered,

but the triplet structures that we obtained were always at

least 2.0 kcal/mol higher in energy than the singlet ones.

The first half-reaction

As mentioned already, in the present proposal the pathway

starts from a structure where there is a l-g2:g2 bridging

peroxide, as suggested experimentally (Fig. 8). When the

catechol is added to the active site, two hydrogen bonds are

formed between the substrate and the bridging peroxide

(OA–H and OB–H distances are 1.71 and 1.68 A,

respectively) stabilizing the system by 1.6 kcal/mol.

Moreover, the CuB–OB distance increases from 2.05 to

2.38 A.

In the first step of the mechanism the peroxide abstracts

a proton from the catechol. The fully optimized transition

state (TS12) is shown in Fig. 9. At the end, the spin pop-

ulation is –1.06 on the substrate and the charge is –0.5,

showing that one electron has also been transferred. The

net result is that a hydrogen atom (a proton and an electron)

is transferred from the substrate. The spin population on

CuB has changed from –0.57 to 0.18, indicating a change of

the oxidation states of the dimer to Cu2(II,I). This reaction

step is exothermic by –0.5 kcal/mol.

Fig. 8 Fully optimized starting

point (structure 0) for the

catalytic cycle. Atoms marked

with an asterisk were kept fixed,

from the X-ray structure in the

geometry optimization

Fig. 9 Fully optimized

transition state for the first

hydrogen-atom transfer (TS12).

Atoms marked with an asteriskwere kept fixed, from the X-ray

structure in the geometry

optimization. Distances are in

angstroms

J Biol Inorg Chem (2007) 12:1251–1264 1257

123

Page 8: Theoretical study of the catalytic mechanism of catechol

A comment should be made at this point, since there is

so far no explicit experimental evidence for radicals in

catechol oxidase. However, radicals are strongly impli-

cated in the formation of the covalent link between Cys92

and His109. In other enzymes where such covalent links

are present, radical chemistry in the catalytic mechanism

has also been suggested [59]. For example, tyrosyl radicals

are suggested to be important in both galactose oxidase

[60] and cytocrome c oxidase [61].

The next step is the most complicated and demanding

one in the mechanism, since it involves the cleavage of the

O–O bond (Fig. 10).

At the transition state the O–O bond has increased from

1.53 to 1.81 A. The spin-density population on CuB has

changed from 0.18 to –0.62 and that for OA has increased

from 0.19 to 0.70. This means that there has been an

electron transfer from CuB to OA. The barrier obtained for

this step is 12.1 kcal/mol, which is in reasonable agreement

with the experimental rate (13 kcal/mol) [8]. The fully

optimized transition state is shown in Fig. 11. After the

O–O bond cleavage, the CuB–OB distance has decreased

from 2.91 to 2.07 A.

In the last step of the first half-reaction, a proton transfer

is accompanied by an electron transfer (the spin-density

population on the substrate goes from –1.02 to 0.00 and

that on OA goes from 0.70 to –0.01), which means that a

hydrogen atom is again transferred. Consequently, the first

molecule of quinone is obtained and another catechol can

enter into the catalytic cycle. The fully optimized transition

state for this step (TS34) is shown in Fig. 12.

According to the mechanisms suggested by Solomon

et al. [2] and Eicken et al. [14] the first molecule of water is

obtained after the first half-reaction, apart from the first

quinone molecule. In the present mechanism, there are

instead two OH ligands coordinated to both copper atoms

at this stage. The two molecules of water are going to be

obtained during the second half-reaction.

Some geometrical parameters of the structures that

appear in the first half-reaction as well as their relative

energies can be found in Table 1. The most relevant spin

populations for these structures can be found in Table 2.

According to the results obtained, in the first half-reac-

tion one catechol molecule is oxidized to one quinone and

peroxide is reduced to hydroxides (Fig. 13). Although the

oxidation state of the dicopper core is the same at the start

and at the end of this half-reaction, CuB has a very

important role in it, since it makes the reduction of per-

oxide possible.

With the departure of the quinone, the second catechol

substrate can bind to the dimer. The quinone–catechol

exchange is calculated to be endothermic by 1.4 kcal/mol.

The changes in entropy and the changes in hydrogen

bonding to external residues are assumed to be small,

CuIIHis

His

His O

OCuI His

His

His

OOH

HCuIIHis

His

His O

OCuII His

His

His

OOH

H

32

Fig. 10 The second step of the mechanism suggested for catechol

oxidase

Fig. 11 Fully optimized

transition state for the O–O

bond cleavage (TS23). Atoms

marked with an asterisk were

kept fixed, from the X-ray

structure in the geometry

optimization. Distances

are in angstroms

1258 J Biol Inorg Chem (2007) 12:1251–1264

123

Page 9: Theoretical study of the catalytic mechanism of catechol

which is probably a reasonable approximation. Modeling

this process more accurately is quite difficult and beyond

present interest.

The second half-reaction

When the second catechol enters the system, the direction

of the OH groups changes in order for hydrogen bonds to

be formed with the substrate (Fig. 14).

In the first step, one of the oxygen atoms of the substrate

binds to CuB and one proton is transferred from the cate-

chol to OB, leading to the formation of the first molecule of

water (Fig. 15). The CuB–OB distance increases from 2.05

to 3.16 A. Since there is no spin on the substrate and the

oxidation state of the copper atoms does not change, no

electrons are transferred in this step. This part of the

mechanism is in agreement with experimental suggestions

that propose an attack on CuB [62].

In the final step of the second half-reaction a hydrogen

atom (a proton and an electron) is transferred from the

substrate and the second molecule of water is obtained. The

fully optimized transition state for this step (TS67) is

shown in Fig. 16.

The spin of CuA changes from –0.64 to 0.00 but for CuB

it remains 0.60. The spin on the substrate increases from

0.00 to –0.60. This means that in order to release the

second quinone product, an electron has to be transferred

from the quinone radical anion to CuB(II), leading to the

reduced Cu2(I,I) dimer.

Some geometrical parameters of the structures that

appear in the second half-reaction as well as their relative

energies can be found in Table 3. The most relevant spin

populations for these structures can be found in Table 4.

According to the results obtained, in the second half-

reaction one catechol molecule is oxidized to one quinone,

the dicopper core is reduced from Cu2(II,II) to Cu2(I,I) and

two molecules of water are obtained from two hydroxides

Fig. 12 Fully optimized

transition state for the second

hydrogen-atom transfer (TS34).

Atoms marked with an asteriskwere kept fixed, from the X-ray

structure in the geometry

optimization. Distances

are in angstroms

Table 1 Comparison of

geometrical parameters for the

structures of the first half-

reaction of the catalytic cycle

of catechol oxidase. Relative

energies are also reported

Structure Distance (A) DE (kcal/mol)

CuA–OA CuA–OB CuB–OA CuB–OB OA–OB

0 2.05 2.04 1.99 2.05 1.51

1 2.10 2.02 1.97 2.38 1.54 0.0

TS12 2.10 2.04 2.13 2.98 1.53 4.9

2 2.11 2.15 2.06 2.91 1.53 –0.5

TS23 2.07 2.00 1.94 3.03 1.81 11.6

3 2.04 2.00 1.93 2.07 2.20 –6.9

TS34 2.04 1.97 1.95 2.16 2.18 –6.0

4 2.03 1.98 1.94 2.08 2.60 –41.0

J Biol Inorg Chem (2007) 12:1251–1264 1259

123

Page 10: Theoretical study of the catalytic mechanism of catechol

(Fig. 17). With one new oxygen molecule and a catechol

substrate the cycle can start again.

Rather than attempting to calculate the energy of the

rather complicated step where the product quinone and

water go out and a dioxygen molecule and a new substrate

catechol come in, this energy can be estimated form the

overall reaction (Fig. 1). The calculated exothermicity for

this reaction in the gas phase is 20 kcal/mol. To this value,

the energies of placing the reactants and products in a

suitable surrounding must be added. Assuming that the

hydrogen-bonding energies and entropies for the catechol

and the quinone are approximately the same, the energy

contribution from these molecules will be close to zero.

The remaining contributions are then on the reactant

side from one dioxygen and on the product side from

two water molecules. The entropy of dioxygen is estimated

to be 12 kcal/mol and the binding energies of the two

water molecules are estimated to be 14 kcal/mol each. This

will make the total reaction exergonic by 36 (20 –

12 + 14 + 14) kcal/mol.

Conclusions

The catalytic cycle of catechol oxidase has been studied

using methods and models similar to those used before for

many other metalloenzymes. On the basis of experience

from modeling other enzymes [21], the catalytic cycle was

constructed in a way where no proton enters or leaves the

active-site region, thus keeping the charge constant at the

active site. The catalytic cycle suggested is shown sche-

matically in Fig. 18 and the energetics are shown in

Fig. 19. The cycle starts with the oxidation of Cu2(I,I)

by dioxygen to Cu2(II,II), forming a l-g2:g2 bridging

OH

OH

O

O

+ O22- + 2OH -

Fig. 13 The first half-reaction

Fig. 14 Fully optimized

minimum for the starting point

of the second half-reaction.

Atoms marked with an asteriskwere kept fixed, from the X-ray

structure in the geometry

optimization. Distances are in A

Table 2 Comparison of spin populations for the structures of the first

half-reaction of the catalytic cycle of catechol oxidase

Structure Spin density

CuA CuB OA OB Substrate

0 0.50 –0.50 –0.01 0.03

1 0.56 –0.57 –0.02 0.00 0.00

TS12 0.55 –0.31 0.15 0.14 –0.58

2 0.45 0.18 0.19 0.08 –1.06

TS23 0.60 –0.33 0.21 0.44 –1.02

3 0.59 –0.62 0.70 0.36 –1.02

TS34 0.62 –0.61 0.62 0.42 –0.97

4 0.62 –0.61 –0.01 0.01 0.00

1260 J Biol Inorg Chem (2007) 12:1251–1264

123

Page 11: Theoretical study of the catalytic mechanism of catechol

peroxide. The charge is thus +2. The peroxide abstracts a

hydrogen atom from the first catechol substrate. Subse-

quently, there is cleavage of the O–O bond, followed by

transfer of a second proton, resulting in the first molecule

of quinone. When the second catechol enters the system,

there is transfer of a proton and the first molecule of water

is obtained. When the last hydrogen is transferred, one of

the coppers is reduced and the second water molecule is

obtained. With one new oxygen molecule and a catechol

substrate, the cycle can start again.

In this new mechanism the most critical step is the

peroxide O–O bond cleavage, which has a barrier that is in

Fig. 16 Fully optimized

transition state for the last

hydrogen transfer (TS67).

Atoms marked with an asteriskwere kept fixed, from the X-ray

structure in the geometry

optimization. Distances are in A

Table 3 Comparison of

geometrical parameters for the

structures of the second half-

reaction of the catalytic cycle

of catechol oxidase. Relative

energies are also reported

Structure Distance (A) DE (kcal/mol)

CuA–OA CuA–OB CuB–OA CuB–OB OA–OB

5 2.04 1.97 1.95 2.05 2.47 –39.6

TS56 2.05 1.96 1.92 2.47 2.59 –37.2

6 2.08 1.98 2.03 3.16 2.71 –38.4

TS67 3.03 2.00 2.02 3.44 3.02 –31.1

7 3.89 2.06 2.04 3.27 3.27

Table 4 Comparison of spin

populations for the structures

of the second half-reaction of

the catalytic cycle of catechol

oxidase

Structure Spin density

CuA CuB OA OB Substrate

5 –0.63 0.65 –0.02 0.08 0.00

TS56 –0.64 0.64 –0.03 –0.09 0.00

6 –0.64 0.59 0.03 –0.06 0.08

TS67 –0.50 0.60 0.05 –0.04 –0.12

7 0.00 0.61 0.05 –0.05 –0.60

CuIIHis

His

His O

O

CuII His

His

His

H

H

OOH

H

CuIHis

His

His O

OH2

CuII His

His

His

H

OOH

5 6

Fig. 15 The first step for the second half-reaction of the suggested

mechanism for catechol oxidase

J Biol Inorg Chem (2007) 12:1251–1264 1261

123

Page 12: Theoretical study of the catalytic mechanism of catechol

reasonable agreement with the experimental rate. In some

steps there is a monodentate coordination of the substrate

to the dicopper core, which is in line with the proposal

by Klabunde et al. [11, 14]. Owing to the structural

similarities between catechol oxidase and tyrosinase,

whose crystallographic structure has recently been reported

[63], the understanding of the activity of the former could

shed some light on how the latter works.

CuIIHis

His

His O

OCuII His

His

His

OOH

H

CuIIHis

His

His O

OCuI His

His

His

OOH

H

CuIIHis

His

His O

OCuII His

His

His

OOH

H

CuIIHis

His

His O

OCuII His

His

His

OO

H

H

CuIIHis

His

His O

O

CuII His

His

His

H

H

OOH

H

CuIIHis

His

His O

OH2

CuII His

His

His

H

OOH

CuIHis

His

His H2O

OH2

CuII His

His

His

OO

O

O

OH

OH

O

O

OH

OH

+ 2H2O

O2 +

1

5 4

3

2

6

7

step astep g

step f step b

step e step c

step d

Fig. 18 Suggested catalytic

cycle for catechol oxidase

OH

OH

O

O

+ 2OH - + 2Cu2+ + 2H2O + 2Cu+

Fig. 17 The second

half-reaction

1 TS343TS232 7TS676TS565

Reaction Coordinate

0.0

-38.4-37.2

-39.6-41.0

-6.0-6.9

11.6

-0.5

4.9

-42.9

-31.1

Quinone

Catechol

-36.0Catechol

O2

Quinone2 H2O

1

)lom/lac k( seigren

E evitaleR

TS12 4

Fig. 19 Potential energy profile

obtained for the catalytic

cycle of catechol oxidase

1262 J Biol Inorg Chem (2007) 12:1251–1264

123

Page 13: Theoretical study of the catalytic mechanism of catechol

Acknowledgment M.G. thanks the MEC for research grants and

J.M. Luis for valuable discussions. We thank the reviewers for helpful

comments.

References

1. Solomon EI, Baldwin MJ, Lowery MD (1992) Chem Rev

92:521–542

2. Solomon EI, Sundaram UM, Machonkin TE (1996) Chem Rev

96:2563–2605

3. Cuff ME, Miller KI, van Holde KE, Hendrickson WA (1998)

J Mol Biol 278:855–870

4. Gaykema WPJ, Hol WGJ, Vereijken JM, Soeter NM, Bak HJ,

Beintema JJ (1984) Nature 309:23–29

5. Magnus KA, Tonthat H, Carpenter JE (1994) Chem Rev 94:727–

735

6. Cary JW, Lax AR, Flurkey WH (1992) Plant Mol Biol 20:245–

253

7. Deverall BJ (1961) Nature 189:311–315

8. Baruah P, Swain T (1959) J Sci Food Agric 10:125–129

9. Mayer AM, Harel E (1979) Phytochemistry 18:193–215

10. Walker JRL, Ferrar PH (1998) Biotechnol Genet Eng Rev

15:457–498

11. Klabunde T, Eicken C, Sacchettini JC, Krebs B (1998) Nat Struct

Biol 5:1084–1090

12. Eicken C, Zippel F, Buldt-Karentzopoulos K, Krebs B (1998)

FEBS Lett 436:293–299

13. Wilcox DE, Porras AG, Hwang YT, Lerch K, Winkler ME,

Solomon EI (1985) J Am Chem Soc 107:4015–4027

14. Eicken C, Krebs B, Sacchettini JC (1999) Curr Opin Struct Biol

9:677–683

15. Granata A, Monzani E, Bubacco L, Casella L (2006) Chem Eur J

12:2504–2514

16. Ros JR, Rodriguezlopez JN, Garciacanovas F (1994) Biochim

Biophys Acta Protein Struct Mol Enzymol 1204:33–42

17. Granata A, Monzani E, Casella L (2004) J Biol Inorg Chem

9:903–913

18. Siegbahn PEM (2004) J Biol Inorg Chem 9:577–590

19. Siegbahn PEM (2003) J Biol Inorg Chem 8:567–576

20. Battaini G, Granata A, Monzani E, Gullotti M, Casella L (2006)

Adv Inorg Chem Bioinorg Stud 58:185–233

21. Siegbahn PEM (2003) Q Rev Biophys 36:91–145

22. Lee SY, Lipscomb JD (1999) Biochemistry 38:4423–4432

23. Orville AM, Lipscomb JD (1997) Biochemistry 36:14044–14055

24. Bassan A, Borowski T, Siegbahn PEM (2004) Dalton Trans

20:3153–3162

25. Koval IA, Gamez P, Belle C, Selmeczi K, Reedijk J (2006) Chem

Soc Rev 35:814–840

26. Ackermann J, Meyer F, Kaifer E, Pritzkow H (2002) Chem Eur J

8:247–258

27. Berreau LM, Mahapatra S, Halfen JA, Houser RP, Young VG,

Tolman WB (1999) Angew Chem Int Ed Engl 38:207–210

28. Wegner R, Gottschaldt M, Gorls H, Jager EG, Klemm D (2001)

Chem Eur J 7:2143–2157

29. Torelli S, Belle C, Hamman S, Pierre JL, Saint-Aman E (2002)

Inorg Chem 41:3983–3989

30. Koval IA, Belle C, Selmeczi K, Philouze C, Saint-Aman E,

Schuitema AM, Gamez P, Pierre JL, Reedijk J (2005) J Biol

Inorg Chem 10:739–750

31. Than R, Feldmann AA, Krebs B (1999) Coord Chem Rev

182:211–241

32. Kao CH, Wei HH, Liu YH, Lee GH, Wang Y, Lee CJ (2001) J

Inorg Biochem 84:171–178

33. Torelli S, Belle C, Gautier-Luneau I, Pierre JL, Saint-Aman E,

Latour JM, Le Pape L, Luneau D (2000) Inorg Chem 39:3526–3536

34. Belle C, Beguin C, Gautier-Luneau I, Hamman S, Philouze C,

Pierre JL, Thomas F, Torelli S (2002) Inorg Chem 41:479–491

35. Merkel M, Moller N, Piacenza M, Grimme S, Rompel A, Krebs B

(2005) Chem Eur J 11:1201–1209

36. Koval IA, Sehmeczi K, Belle C, Philouze C, Saint-Aman E,

Gautier-Luneau I, Schuitema AM, van Vliet M, Gamez P, Rou-

beau O, Luken M, Krebs B, Lutz M, Spek AL, Pierre JL, Reedijk

J (2006) Chem Eur J 12:6138–6150

37. Monzani E, Battaini G, Perotti A, Casella L, Gullotti M, Santa-

gostini L, Nardin G, Randaccio L, Geremia S, Zanello P,

Opromolla G (1999) Inorg Chem 38:5359–5369

38. Selmeczi K, Reglier M, Giorgi M, Speier G (2003) Coord Chem

Rev 245:191–201

39. Born K, Comba P, Daubinet A, Fuchs A, Wadepohl H (2007)

J Biol Inorg Chem 12:36–48

40. Santagostini L, Gullotti M, Monzani E, Casella L, Dillinger R,

Tuczek F (2000) Chem Eur J 6:519–522

41. Mahadevan V, DuBois JL, Hedman B, Hodgson KO, Stack TDP

(1999) J Am Chem Soc 121:5583–5584

42. Becke AD (1988) Phys Rev A 38:3098–3100

43. Becke AD (1993) J Chem Phys 98:5648–5652

44. Becke AD (1993) J Chem Phys 98:1372–1377

45. Lee CT, Yang WT, Parr RG (1988) Phys Rev B 37:785–789

46. Schrodinger LLC (2003) MacroModel 8.5. Portland, OR

47. Hay PJ, Wadt WR (1985) J Chem Phys 82:299–310

48. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA,

Cheeseman JR, Montgomery JA Jr, Vreven T, Kudin KN, Burant

JC, Millam JM, Iyengar SS, Tomasi J, Barone V, Mennucci B,

Cossi M, Scalmani G, Rega N, Petersson GA, Nakatsuji H, Hada

M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nak-

ajima T, Honda Y, Kitao O, Nakai H, Klene M, Li X, Knox JE,

Hratchian HP, Cross JB, Bakken V, Adamo C, Jaramillo J,

Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R,

Pomelli C, Ochterski JW, Ayala PY, Morokuma K, Voth GA,

Salvador P, Dannenberg JJ, Zakrzewski G, Dapprich S, Daniels

AD, Strain MC, Farkas O, Malick DK, Rabuck AD, Raghava-

chari K, Foresman JB, Ortiz JV, Cui Q, Baboul AG, Clifford S,

Cioslowski J, Stefanov BB, Liu G, Liashenko A, Piskorz P,

Komaromi I, Martin RL, Fox DJ, Keith T, Al-Laham MA, Peng

CY, Nanayakkara A, Challacombe M, Gill PMW, Johnson B,

Chen W, Wong MW, Gonzalez C, Pople JA (2003) Gaussian 03,

revision B.05. Gaussian, Pittsburgh

49. Tannor DJ, Marten B, Murphy R, Friesner RA, Sitkoff D,

Nicholls A, Ringnalda M, Goddard WA, Honig B (1994) J Am

Chem Soc 116:11875–11882

50. Marten B, Kim K, Cortis C, Friesner RA, Murphy RB, Ringnalda

MN, Sitkoff D, Honig B (1996) J Phys Chem 100:11775–11788

51. Blomberg MRA, Siegbahn PEM, Babcock GT (1998) J Am

Chem Soc 120:8812–8824

52. Curtiss LA, Raghavachari K, Redfern PC, Pople JA (2000) J

Chem Phys 112:7374–7383

53. Siegbahn PEM, Blomberg MRA (1999) Annu Rev Phys Chem

50:221–249

54. Siegbahn PEM, Blomberg MRA (2000) Chem Rev 100:421–437

55. Blomberg MRA, Siegbahn PEM (2001) J Phys Chem B

105:9375–9386

56. Pelmenschikov V, Cho KB, Siegbahn PEM (2004) J Comput

Chem 25:311–321

57. Pelmenschikov V, Siegbahn PEM (2003) J Biol Inorg Chem

8:653–662

58. Rompel A, Fischer H, Meiwes D, Buldt-Karentzopoulos K,

Dillinger R, Tuczek F, Witzel H, Krebs B (1999) J Biol Inorg

Chem 4:56–63

J Biol Inorg Chem (2007) 12:1251–1264 1263

123

Page 14: Theoretical study of the catalytic mechanism of catechol

59. Fontecave M, Ollagnier-de-Choudens S, Mulliez E (2003) Chem

Rev 103:2149–2166

60. Whittaker MM, Whittaker JW (1988) J Biol Chem 263:6074–

6080

61. Proshlyakov DA, Pressler MA, Babcock GT (1998) Proc Natl

Acad Sci USA 95:8020–8025

62. Decker H, Dillinger R, Tuczek F (2000) Angew Chem Int Ed

Engl 39:1591–1595

63. Matoba Y, Kumagai T, Yamamoto A, Yoshitsu H, Sugiyama M

(2006) J Biol Chem 281:8981–8990

1264 J Biol Inorg Chem (2007) 12:1251–1264

123