Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Jul 1.
Published in final edited form as: Arch Biochem Biophys. 2015 May 14;0:24–34. doi: 10.1016/j.abb.2015.04.007

Mechanistic Studies of the Tyrosinase-Catalyzed Oxidative Cyclocondensation of 2-Aminophenol to 2-Aminophenoxazin-3-one

Courtney Washington 1, Jamere Maxwell 1, Joenathan Stevenson 1, Gregory Malone 1, Edward W Lowe Jr 2, Qiang Zhang 1, Guangdi Wang 1, Neil R McIntyre 1,*
PMCID: PMC4456232  NIHMSID: NIHMS690905  PMID: 25982123

Abstract

Tyrosinase(EC 1.14.18.1) catalyzes the monophenolase and diphenolase reaction associated with vertebrate pigmentation and fruit/vegetable browning. Tyrosinase is an oxygen-dependent, dicopper enzyme that has three states: Emet, Eoxy, and Edeoxy. The diphenolase activity can be carried out by both the met and the oxy states of the enzyme while neither mono- nor diphenolase activity results from the deoxy state. In this study, the oxidative cyclocondensation of 2-aminophenol (OAP) to the corresponding 2-aminophenoxazin-3-one (APX) by mushroom tyrosinase was investigated. Using a combination of various steady- and pre-steady state methodologies, we have investigated the kinetic and chemical mechanism of this reaction. The kcat for OAP is 75±2 sec−1, KMOAP=1.8±0.2mM, KMO2=25±4μM with substrates binding in a steady-state preferred fashion. Stopped flow and global analysis support a model where OAP preferentially binds to the oxy form over the met (k7 ≫ k1). For the met form, His269 and His61 are the proposed bases, while the oxy form uses the copper-peroxide and His61 for the sequential deprotonation of anilinic and phenolic hydrogens. Solvent KIEs show proton transfer to be increasingly rate limiting for kcat/KMOAP as [O2] −>0 μM (1.38±0.06) decreasing to 0.83±0.03 as [O2]−>∞ reflecting a partially rate limiting μ-OH bond cleavage (Emet) and formation (Eoxy) following protonation in the transition state. The coupling and cyclization reactions of o-quinone imine and OAP pass through a phenyliminocyclohexadione intermediate to APX, forming at a rate of 6.91±0.03 μM−1s−1 and 2.59E-2±5.31E-4 s−1. Differences in reactivity attributed to the anilinic moiety of OAP with o-diphenols are discussed.

Keywords: Isotope effect, tyrosinase, 2-aminophenol, mechanism, cyclocondensation, kinetics

Graphical abstract

graphic file with name nihms690905u1.jpg

Introduction

Tyrosinase (E.C. 1.14.18.1) is a member of the type-3 copper protein family catalyzing the monophenolase and diphenolase activity on phenolic/anilinic and catechol, 2-aminophenol and 2-aminoaniline substrates, respectively [16]. Representative members of the type-3 copper protein family include: catechol oxidase (CO), hemocyanin (Hc) and tyrosinase (Ty) though members have diverse functions. Tyrosinase is associated with pigmentation catalyzing the first step in melanin biosynthesis – a pigment ubiquitously present in all phyla [6]. The structure of mushroom (Agaricus bisporus) tyrosinase (abTy) is a tetrameric protein with 2 heavy (H, 392 aa, ~43 kDa) and 2 light (L, 150 aa, ~14 kDa) chains (2.30 Å resolution, 2Y9W) [7, 8]. As a derivative of the ppo3 gene, the fungal enzyme is structurally homologous to other tyrosinases, only larger with 110 more residues present as loops connecting conserved secondary structural elements. H-subunits of abTy contain 13 α-helices, 8 mostly short β-strands and many loops, similar to known type-3 protein structures including: tyrosinase structures from Streptomyces castaneoglobisporus [9] and Bacillus megaterium [10, 11] as well as Octopus dolfeini Hc [12] and I. batatas catechol oxidase [13]. Within the H-subunit, the dicopper binding domain is comprised of four antiparallel helices (α3, α4, and α10, α11) positioned perpendicularly with CuA is coordinated by H61 (α3), H94 (α4) and H85 (loop connecting α3 to α4) and CuB by H259, H263 (α10) and H296 (α11). The structure of the L-subunit in abTy is best described as a β-trefoil fold with 12 antiparallel β-strands assembled in a cylindrical barrel of six 2-stranded sheets [7]. The L-subunit appears to play a structural role in tetramer formation positioned 25 Å away from the active site as H-subunit turnover number is unaffected by its absence.

The di-copper domain in tyrosinase is classified by three distinct geometric and electronic architectures with alternate function toward substrate oxidation: Edeoxy, Emet, and Eoxy in addition to an inactive mixed valent copper form [14]. The oxy-state (Eoxy) of tyrosinase has molecular oxygen bound in a side-on μ−η22 geometry linked to the solvent-bridged met-state (μ-OH) and the reduced deoxy-state through the reversible exchange of the ligand as oxygen (Edeoxy) or peroxide (Emet) [15]. The mechanism for tyrosinase-catalyzed phenol (monophenolase) and o-diphenol (diphenolase) oxidation has been investigated using steady-state, transient phase and kinetic isotope effects [1622]. The monophenolase (monooxygenase) activity is catalyzed by tyrosinase in the Eoxy form while the o-diphenolase (oxidase) activity results from both Emet and Eoxy forms of the enzyme [19]. Several tyrosinase studies demonstrate that the Eoxy form is stable and subsequently reduced to Emet with the concomitant oxidation of o-diphenol to the corresponding o-quinone [2325]. Evaluation of the abTy o-diphenolase reaction mechanism demonstrate tight O2 binding in addition to an increased second-order rate constant for the o-diphenol substrate to preferentially bind with the Emet state over the Eoxy form of the enzyme [20]. The monophenolase activity of tyrosinase utilizes an electrophilic aromatic substitution (EAS) mechanism for ortho-hydroxylation.

Interestingly, tyrosinase and type-3 model complexes lack an observable kinetic isotope effect (KIE) with ortho-deuterated monophenol substrates. This suggests that the ortho-deprotonation step in this mechanism is either extremely slow or concerted with the oxygen insertion step. The observation of a normal KIE (>1) was revealed upon deuteration of the phenolic hydrogen for both phenol and o-diphenol substrates [16, 17, 26, 27]. With proton inventory solvent KIE (sKIE) studies, the mechanism for monophenolase and diphenolase tyrosinase activity support a partially-rate limiting proton transfer in both Emet and Eoxy forms attributed to nucleophilic attack by the corresponding phenolate ion to copper [16, 17].

Structure-function relationships illustrate drastically altered sKIE values based on the para substituent of the substrate. Therein, the fractionation factors (ϕ) decrease as the observed sKIE for kcat increase with each mode of abTy-dependent oxidation. For monophenolase activity, the lowest fractionation factor values and highest sKIE were observed for substrates with an ionic para substituent. For diphenolase activity, the trend is less clear coupling higher fractionation factors with lower sKIE to those reported for the monophenolase reaction. These finding are consistent with a general base catalytic mechanism important to deprotonate substrate phenolic hydrogens that contribute to the stabilization of slightly altered transition state geometries from either the Eoxy or Emet reaction coordinates [2834]. The extent of transition state stabilization through altered protonic interactions support the presence of a far different microenvironment for the diphenolase reaction coordinate compared to that observed for monophenolase activity.

The precedence for 2-aminophenol oxidase activity has been previously observed in several copper containing enzymes (laccase [36], tyrosinase [1] and various homologues such as: griF [37], phenoxazinone synthase [38]) as well as biomimetic complexes ranging from copper through manganese [39]. The oxidation of 2-aminophenols by tyrosinase has been previously studied showing a mechanism similar to catechol compounds [2]. For o-aminophenol oxidation, the turnover number is decreased relative to the o-diphenol substrates though KM values does not appear to be dramatically affected by the structural difference [1]. The resulting o-quinone imine (Q) undergoes a sequential coupling and cyclization towards the corresponding planar 2–3H-aminophenoxazin-3-one (APX) structure passing through several proposed intermediates (Scheme 1) [35, 40]. Our current study of 2-aminophenol oxidation by abTy was performed to increase the present body of data contrasting the oxidase mechanism for 2-aminophenol with o-diphenol. The role of the phenylamine moiety was studied using a mixture of steady-state and transient phase kinetics, sKIEs, global analysis and molecular modeling studies to characterize multiple aspects of the enzymatic and non-enzymatic reaction coordinate for 2-aminophenol oxidation. Therein, we present experimental evidence supporting a distinct preference for 2-aminophenol oxidation by the Eoxy form of abTy (over the more stable Emet form) to produce the o-quinone imine product which passes through a spectroscopically observed transient phenyliminocyclodione intermediate prior to APX product formation.

Scheme 1.

Scheme 1

Reaction stoichiometry for the tyrosinase catalyzed oxidation of 2-aminophenol to o-quinone imine followed by non-enzymatic coupling with 2-aminophenol passing through several proposed intermediate structures to give 2-amino-3H-phenoxazin-3-one (product). Please note, estimates for the λmax for intermediates were based on previous work by Barry et al [35].

Materials and Methods

Reagents

Mushroom tyrosinase (3300 U/mg), 2-aminophenol, L-ascorbic acid, D2O, mono-, di- and tri-sodium phosphate, DMSO, DMSO-d6 and purchased from Sigma-Aldrich Co. (St. Louis, MO). HPLC-grade solvents (acetonitrile, methanol, and water) were purchased from Fisher Scientific (Fair Lawn, NJ). Analytical standards were purchased from Sigma-Aldrich Co. (St. Louis, MO). All other experimental reagents were purchased from commercial sources at highest purity grade available and used without additional modification. Commercial enzyme was purified following the procedure of Duckworth and Coleman [41] or the IMAC method [42]. Protein concentration was determined by the bicinchoninic acid assay [43] with bovine serum albumin as standard. Protein purity was assessed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis.

Oxygen Electrode

Reactions at 37.0 ± 0.1 °C were initiated by the addition of ~0.10–0.50 μM mushroom tyrosinase (4–5 μL) into 2.0 mL of 100 mM sodium phosphate (pH 6.4). Initial velocities were measured at varying concentrations of 2-aminophenol and oxygen. Initial rates were measured by following the tyrosinase-dependent consumption of O2 using a Yellow Springs Instrument Model 5300 oxygen monitor interfaced with a personal computer using a Dataq Instruments analogue/digital converter (model DI-158RS) interfaced to Microsoft Excel through the Windaq module, modified from McIntyre et al. [44]. Using a Maxtec Low flow oxygen blender, in situ [O2] was changed by varying the percent mixtures of oxygen and argon standardized to air-saturated distilled water at 37.0 ± 0.1 °C. Turnover numbers were normalized to controls performed at 8.99 mM 2-aminophenol to account for differences in specific activity between different lots of enzyme. Background O2 consumption rates were first determined without enzyme and were subtracted from the rate obtained upon tyrosinase addition. Initial rates were determined from the consumption of dissolved oxygen and converted to o-quinone imine production or 2-aminophenol consumption using a 2:1 substrate or product to oxygen ratio (equation 1).

vo=12d[O2]dt=2d[OAP]dt=2d[Q]dt (1)

Initial rate data was then fit to the bi-substrate Michaelis-Menten equation in SigmaPlot 11.2 using the Enzyme Kinetics module (version 1.3)

rate=kcat[OAP][O2]KIO2KMOAP+KMO2[OAP]+KMOAP[O2]+[OAP][O2] (2)

where kcat is the turnover number (maximal number of OAP and O2 molecules converted to o-quinone imine per tyrosinase active site per second), KMOAP and KMO2 are the Michaelis constants for 2-aminophenol (OAP) and molecular oxygen (O2), respectively, and KIO2 is the limiting value of KMO2 when the [OAP] approaches zero. Due to the high degree of error for the KMO2 parameter using equation 2, a replot of kcatapp versus [O2] was fit to equation 3 to measure the Michaelis constant for [O2] under pseudo-first order conditions and fit in Kaleidagraph™.

rate=kcatapp×[O2]KMO2+[O2] (3)

Molecular Modeling

The initial coordinates of the Agaricus bisporus (mushroom) tyrosinase at 2.30 Å resolution were obtained from the Protein Data Bank (http://www.rcsb.org/pdb/2Y9X). The structure was prepared for simulation using the protein preparation module of Schrodinger Drug Discovery Suite 2011-2. Water molecules were removed and copper valence was assigned. In the copper binding domains of the active-site, the μ−η22 oxygen bridged structure (Eoxy) and η2-hydroxyl bridged structures (Emet) were created through the manual insertion of the oxygen species followed by minimization. The Coulson charges for the ligands were generated and Glide XP was used to generate the initial poses. The partial atomic charges utilizing electrostatic potential fitting of the protein-bound ligand were calculated using Q-site then re-docked with these updated charges using Glide XP.

Solvent Kinetic Isotope Effects

The dependence of D2O on the catalytic efficiency of 2-aminophenol oxidation as a function of oxygen tension was performed in a non-competitive fashion following tyrosinase-dependent oxygen consumption. A 10 mM stock of deuterated 2-aminophenol was prepared with 10% DMSO-d6 in 50 mM sodium phosphate buffer (pH 6.4) dissolved in D2O, the reaction buffer was 100 mM sodium phosphate buffer (pH 6.4) dissolved in D2O. Reactions were initiated at 37.0 ± 0.1 °C with 5 μL of enzyme (0.5 nM), six 2-aminophenol concentrations from 0.90 to 8.99 mM and five oxygen concentrations ranging from 65–850 μM were used, respectively. The kcat/KMOAP data was determined using standard Michaelis-Menten kinetics (equation 4) and plotted against oxygen tension fitting the data to equation 5. The dependence of pH on the kcat/KMOAP term was performed using the 2-aminophenol concentration range 0.90 to 8.99 mM under the same conditions listed above over a phosphate buffer pH range of 5.8 to 8.5. The individual parameters of the kcat/KMOAP term (kcat and KM) were determined using equation 4 and plotted against pH and fit by a linear regression function.

rate=kcatx[OAP]KM+[OAP] (4)

The  D(kcat/KMOAP) data versus [O2] was fit to the following exponential curve in Kaleidagraph™.

 D(kcatKMOAP)observed=[ D(kcatKMOAP)[O2]0μM D(kcatKMOAP)[O2]]eax[O2]+ D(kcatKMOAP)[O2] (5)

Where

α=Dk1KDO2k7+k9+ D(kcatKMOAP)[O2]0μM+1 (6)

Oxy-Tyrosinase

The addition of L-ascorbic acid and hydroxylamine to tyrosinase assays prevent substrate oxidation and reduce the Emet form of the enzyme to Eoxy [45]. Suppression of the UV-vis spectroscopic signal of oxidized 2-aminopehnol was followed using an OLIS refurbished Hewlett-Packard 8452 diode array spectrophotometer. The relative rate of 2-aminophenoxazin-3-one (APX) formation at 432 nm was measured versus L-ascorbic acid concentrations intended to perform the in situ reduction of the o-quinone imine product ranging from 0–6 molar equivalents (Figure S.1.).

The stability of the Eoxy form of abTy was examined by mixing the Edeoxy form with an oxygenated buffer in the absence of the 2-aminophenol substrate. Therein, 80 μM of the reduced enzyme (Edeoxy) was prepared anaerobically (5 vacuum-argon gas cycles) with 6-equivalents of hydroxylamine and rapidly mixed with an oxygen saturated buffer. The signal for Eoxy was followed at 345 nm (ε345 = 18 mM−1cm−1) with a time course collected at 2 millisecond intervals [24, 25]. Consumption of the Eoxy signal in the presence of the 2-aminophenol substrate used 20 μM Eoxy was prepared from the reduced enzyme and oxygen gas (vide supra) then rapidly mixed with a 1 mM 2-aminophenol solution containing 6 molar equivalents of L-ascorbic acid [20].

All rapid-mix studies were performed using a Kintek Mini-Mixer (KinTek Corp., Austin, TX) at 37±0.1 °C with temperature maintained with a Neslab RTE-110 water circulator. Data was collected and processed using an apparatus designed in-house with supplies purchased from Ocean Optics (Ocean Optics, St. Petersburg, Fl.). Specifically, the signal was collected with the SpectraSuite interface using a USB 2000+XR spectrophotometer and DH-2000 light source linked to the observation cell through Series 74 collimating lenses (200–2000 nm) and solarization resistant fiber optic cables. All curve fitting and kinetic simulation of this data was analyzed with Kaleidagraph™ and Kintek Explorer Professional (KinTek Corp., Austin, TX)

Non-Enzymatic formation of 2-Aminophenoxazin-3-one

100 μM tyrosinase was mixed with 1 mM 2-aminophenol under ambient oxygen saturation at 37±0.1 °C. The proposed intermediate was observed at 380 nm and the 2-aminophenoxazin-3-one (APX) product was followed at 432 nm (ε432= 23.4 mM−1cm−1). 3-Dimensional analysis of absorbance, time and wavelength were exported from SpectraSuite (Ocean Optics, St. Petersburg, Fl.) into the Specfit module of Kintek Explorer Professional (KinTek Corp., Austin, TX) where single value decomposition (SVD) analysis was performed. The spectra was reconstructed from the SVD fits then subtracted from the experimental data to ensure the SVD analysis was statistically relevant. Kinetic constants were constrained by the experimental parameters and calculated using FitSpace to measure the values and show a contour plot.

Kinetic Simulation

Kintek Explorer Professional (KinTek Corp., Austin, TX) with the SpecFit module was used for global analysis through direct numerical integration to simulate experimental data [4648]. The mechanism proposed in Scheme 2 was simulated in two different parts: catalyzed and non-catalyzed reactions. Each respective path was modeled using the steady- and transient phase data to constrain the minimal kinetic mechanism for the reaction. For the tyrosinase-catalyzed oxidation of 2-aminopehnol to the corresponding o-quinone imine, a global analysis approach was utilized. Therein, mixtures of pre- and steady-state kinetics were used to evaluate the catalyzed minimal kinetic model (Scheme 2). The three states of the enzyme (Edeoxy, Emet and Eoxy) were fit as observables within this model and the experimental data for Eoxy was added. The experimental Eoxy data was compared to models showing k1k7 and k1k7 to distinguish the relative ability of Emet and Eoxy to oxidize 2-aminophenol to o-quinone imine, respectively. For the non-enzymatic coupling and cyclization (non-catalyzed) reaction of the o-quinone imine (Q) to 2-aminophenoxazin-3-one (APX) the time resolved spectra were recalculated using a single value decomposition (SVD) analysis calculated from a matrix-representation. This was the ideal method over global analysis because there were far more wavelengths than species. Using singular values, the statistical relevance of the reconstructed spectra (relative contribution and weights) was compared with experimental values and evaluated through residual analysis. For the non-catalyzed reaction, an irreversible two-step kinetic model was used for the formation of APX through a single intermediate. This intermediate represents the accumulation of a transient species associated with early stage coupling steps in the reaction of Q with OAP required to form APX. FitSpace was used to calculate a contour plot for rate constants associated accumulation and decay of this spectroscopically detected intermediate.

Scheme 2.

Scheme 2

Minimal kinetic mechanism of the tyrosinase dependent catalytic oxidation of 2-aminophenol (S) to o-quinone imine (Q) and its non-enzymatic cyclo-addition to 2-amino-3H-phenoxazin-3-one (APX) passing through an intermediate (I).

HPLC-MS/MS Analysis

Lyophilized samples were prepared from 8.99 mM (18 micromoles) of 2-aminophenol oxidized with 750 picomoles mushroom tyrosinase in 100 mM sodium phosphate (200 micromoles) buffer (pH 6.4). The reaction was incubated at 37 °C for about 6 hours. Prior to HPLC analysis, samples were precipitated by centrifugation at room temperature, and the solvent was evaporated with a stream of nitrogen at 37 °C. The residual solution was applied to 6-mL SUPELCO C18 solid-phase extraction (SPE) columns pretreated with water and methanol. The columns were washed with HPLC-grade water followed by elution with methanol. The effluents were again concentrated by a nitrogen stream at 37°C to near dryness before reconstitution with methanol to 1 mL volume for HPLC injection. Samples were analyzed on a linear ion trap tandem mass spectrometer (LTQ) interfaced with a Surveyor HPLC system (Thermo-Fisher Scientific, West Palm Beach, FL). Chromatographic separations were achieved by using a Synergi 4u Hydro-RP 80A column (150 × 2.0 mm, 4 μm particle size; Phenomenex, Torrance, CA). Samples were eluted from the analytical column with mobile phase A (water with 0.1% formic acid) and mobile phase B (acetonitrile with 0.1% formic acid) at a flow-rate of 0.2 mL/min over 22 min using a linear gradient of 0–100% B for 13 min; holding at 100% B for 5 min before returning to 0% B.

Results

Product Characterization

Tandem LC-MS/MS analysis demonstrates two retention times associated with substrate and product structure (Figure S.2). The first retention time of 3.41 minutes corresponds to a high relative intensity peak (M+1) at 110.09 m/z for 2-aminophenol (MW 109.13), tandem MS/MS analysis of this peak gave a fragmentation pattern with peaks at 93.17, 92.13 m/z corresponding to the hydroxide loss (C6H6N+ and C6H5N+) and an 82.16 m/z representing phenol rearrangement resulting in an M-28 (CO removal) fragment (C5H6N). The retention time at 11.60 minutes shows the highest relative intensity peak (M+1) at 213.06 m/z for 2-amino-3H-phenoxazin-3-one (MW 212.06). Tandem MS/MS analysis of the 213.06 m/z peak gives two peaks at 185.03 and 186.11 m/z corresponding to loss of methylamine (−CNH, 27 m/z and −CNH2, 28 m/z).

Steady and Transient State Kinetics

The steady-state kinetics for the oxidation of 2-aminophenol to the corresponding o-quinone imine (Q) was fit to the bi-substrate Michaelis-Menten equation (2). Shown in figure 1, the kcat is 75±2 sec−1 with a of 1.8±0.3 mM with a large error on the KMO2 and KDO2 values. Replotting the rate of 2-aminophenol oxidation as a function of oxygen concentration under pseudo-first order conditions provides a KMO2 value of 25±4 μM (Figure S.3.). Shown in figure S.2A, the double reciprocal plot of 1/rate versus 1/[2-aminophenol] intersects on the abscissa while the 1/rate versus 1/[Oxygen] intersects to the left with a positive y-coordinate, represented as y =[ETotal]−1 ((k6kcat)/(k5kcat)) = 0.097 sec [20]. A replot of the 1/rate versus 1/[2-aminophenol] slope determined for each oxygen concentration (plotted as 1/[O2]) crosses the ordinate at a value greater than 0 (Figure S.4.).

Figure 1.

Figure 1

Steady-state kinetics for the tyrosinase dependent oxidation of 2-aminophenol as a function of oxygen tension. Rates were fit with the bi-substrate Michaelis-Menten equation (equation 2).

The binding of O2 to chemically reduced enzyme prepared anaerobically (Edeoxy) shows the formation of Eoxy observed at 345 nm to be stoichiometric to the initial concentration of the enzyme, activating approximately 92%. Following equilibration, the rate of decay for Eoxy was observed to be slow, 6.50 × 10−4 μM sec−1 (Figure S.5.). The Eoxy form of tyrosinase was rapidly mixed with 2-aminophenol with L-ascorbic acid for the in situ reduction of o-quinone imine (Q). Under saturating oxygen conditions, the consumption of the Eoxy signal corresponding to 13 ±2 μM enzyme followed a single exponential decay function corresponding to 3.1±5.6E-2 μM−1 sec−1 (Figure 2).

Figure 2.

Figure 2

Stopped-flow analysis of Eoxy consumption in the presence of 2-aminophenol. Pre-steady state kinetics were measured at 345 nm upon mixing oxy-tyrosinase (Eoxy) with 1 mM 2-aminophenol with data fit to a single exponential function.

The kinetic complexity of the tyrosinase-dependent oxidation of 2-aminophenol to o-quinone imine (Q) make direct analysis difficult (see appendix for distribution equations). Therefore, elementary rate constants for the minimal kinetic mechanism presented in scheme 2 were determined through a mixture of steady-state and transient phase analysis. Coefficients within the bi-substrate Michaelis-Menten equation are defined according to Scheme 2 as:

Kinetic simulation using global fitting analysis of the proposed minimal kinetic mechanism (scheme 2) is shown in tables 1 and 2. Using Emet, Edeoxy, and Eoxy as ‘observables’ to validate our kinetic model against the experimentally determined rate for Eoxy consumption (k7) to analyze whether 2-aminophenol preferentially reacts with Emet or Eoxy (table 2). Rate constants in our model assumed that product release (Q) was equal to kcat representing with k3 and k9 while the rate constants k1, k5, k6, k7 were fixed to the values derived in table 1 while k2 and k8 were each fixed to 0.1 sec−1. The values for k7 were determined through global fitting analysis of the mechanism to be 0.804 ± 0.041 μM−1sec−1. According to figure 3 (top), k1k7 displayed strong correlation between the experimental and theoretical values for Eoxy throughout the time course. Validation of this model was performed by reversing the values of k1 and k7 (k7k1) and comparing the experimental values for Eoxy with simulated values where dramatic shift in the theoretical Eoxy away the experimental value was observed (figure 3, bottom).

Table 1.

Derivation and determination of rate constants from for the 2-aminophenol oxidase activity from steady-state kinetics according to scheme 2.

kcat = k3 k9/(k3 + k9)
KI,O2KMOAP=k3k6k9/[k5k7(k3+k9)]=kcatKDO2/k7
KMO2=k3k9/[k5(k3+k9)]=kcat/k5
KMOAP=[k3k9(k1+k7)]=[kcat(k1+k7)/k1k7]
kcatKMOAP=k5=3.00±0.0048μM1s1
kcatKMOAP=k1×k7k1×k7
If k7k1 the 2nd order rate constant for OAP binding to Emet become
kcatKMOAP=k1=0.0421±0.0048μM1s1
KDO2=k6k5
k6=KDO2×k5=97±73s1

Table 2.

Steady-state kinetic parameters and rate constants corresponding to the minimal kinetic mechanism representing the best global fit analysis of steady and pre-steady state kinetics of 2-aminophenol (S) oxidation to o-quinone imine (Q) and single value decomposition (SVD) analysis of the cyclization of Q and S to form 2-amino-3H-phenoxazin-3-one (APX). Please note that k2 and k8 were fixed to 0.1 sec−1 and that k11 and k13 were determined using the FitSpace module of Kintek Explorer Professional.

Steady State Kinetic Parameters
kcat 75±2 s−1
KM,OAP 1.8±0.2 mM
KM,O2 25±4 μM
KD,O2 33 ±24 μM

Rate Constants
k1 42.1 ± 4.8 mM−1s−1
k2 0.1* s−1
k3 150 s−1
k5 3,000 mM−1s−1
k6 97 s−1
k7 804 ± 41 mM−1s−1
k8 0.1* s−1
k9 150 s−1
k11 6,910 ± 31 mM−1s−1
k13 2.54E-2 ± 5.31E-4 s−1

Figure 3.

Figure 3

Model validation correlating simulated data representing the distribution of tyrosinase enzyme states with experimentally derived data for a time course. Each simulation was calculated used kinetic data fit to the global model with k7 ≪ k1 (top) or k7 ≫ k1 (bottom) where the calculated values were switched and compared directly with Eoxy values determined experimentally.

The mechanism for the non-enzymatic coupling of the o-quinone imine (Q) product with 2-aminophenol was predicted to pass through several intermediates leading to the APX product (Scheme 1). Figure 4 shows time resolved spectra for a transient peak at 380 nm (I) followed by a peak at 432 nm (APX). SVD de-convolution of observable species and spectra reconstruction demonstrate that the residual of the experimental and simulated spectra are well matched (Figure S.6.). Analysis of rate constants for the reaction model with global analysis show k11 and k13 to be 6.91 ± 0.03 μM−1sec−1 and 2.59 E-2 ± 5.31E-4E-6 sec−1 (Table 2) with contour diagram generated though FitSpace showing the data to be well constrained within the model (Figure S.7.). By correlating the absorbance of APX (ε432 = 23.4 mM−1cm−1) with the observed intermediate, a ε380 = 7.7 mM−1cm−1 was determined and the extracted data was fit to a time course that reflected concentration (Figure S.8.).

Figure 4.

Figure 4

Time resolved spectra for 2-amino-3H-phenoxazinone (APX) formation from 2-aminophenol and mushroom tyrosinase.

The dependence of sKIE on oxygen tension was studied for kcat/KMOAP (Figure 5). The results clearly demonstrate that proton transfer is increasingly rate limiting to the catalytic efficiency for the tyrosinase-dependent oxidation of OAP as [O2] → 0 mM, extrapolating to a value of D2O(k/K) = 1.38±0.06. As [O2] →∞, the reaction displays an inverse solvent kinetic isotope effect, 0.83±0.03. Pertinent to the sKIE studies, pH versus kcat/KMOAP showed near constant value of ~4 mM−1sec−1 over a broad range (5.7–8.5) (Figure S.8.).

Figure 5.

Figure 5

Solvent kinetic isotope effects on the catalytic efficiency (kcat/KMOAP) as a function of oxygen tension and fit to equation 5. The inset displays the dependence of (kcat/KMOAP) with oxygen concentration.

Molecular Modeling

Docking orientation in the Emet form shows both 2-aminophenol and catechol to interact with the imidazole side chain of His263 in the CuB binding domain. The distance between Nε−His263 and the anilinic hydrogen of 2-aminophenol was 2.26 Å and 3.71 Å for the corresponding phenolic hydrogen of catechol (figure 6). For Eoxy the distance of the 2-aminophenol N-H distance to the peroxide was 2.42 Å and 2.34 Å for the OH group of catechol (Figure 7). For both Emet and Eoxy active site architectures, both substrates share nearly superimposable benzene ring stacking with the His296 imidazole moiety and hydrogen bonding with Asn260.

Figure 6.

Figure 6

Binding orientation of 2-aminophenol in the active site of the Eoxy form of mushroom tyrosinase. The active site contains copper coordinating histidines (tube), a bridging peroxide for CuA and CuB which hydrogen bonds the amino moiety of 2-aminophenol(OAP). For clarity, catechol was not included but (like Emet) shares an indistinguishable binding mode with 2-aminophenol. All other residues that extend beyond the active site are represented by their secondary structural elements.

Figure 7.

Figure 7

Homologous binding mode of 2-aminophenol and catechol (‘Substrates’) in the active site of the Emet form of mushroom tyrosinase. Please note, the Emet active site contains copper coordinating histidines (tube) with bridging hydroxide (ball and stick) ion for CuA and CuB. Please note the hydrogen bonding between the hydroxyl moiety of each substrate with Asn260. All other residues that extend beyond the active side are represented by their secondary structural elements.

Discussion

Oxygen binds with high affinity to mushroom tyrosinase [49]. Binding of O2 to abTy is best characterized as a two-electron exchange stabilization of triplet O2 [5052]. In the Eoxy form, the O2 is reversibly activated through a peroxide dicopper(II) complex that is highly conserved in type-3 proteins, including Hc and Ty. Binding of O2 to the dicuprous Edeoxy state results in filling the π* HOMO (highest occupied molecular orbital) through two electron reduction to peroxide (O2−2) to overcome the spin-forbidden triplet state oxygen (3O2) containing two unpaired electrons situated in the doubly degenerate π* orbital. In the EOXY form, the side-on μ−η22 geometry is responsible for direct orbital overlap (coordination) necessary to produce the intense ligand to metal charge transfer (LMCT) band and is the most favorable orbital overlap for dioxygen reduction to peroxide (Eoxy) [53]. The resulting peroxide character of the LMCT is associated with O2−2 π*σ → Cu(II) (dx2y2) interactions responsible for significant peroxide character in the LUMO. These π*σ interactions allow strong σ donation into dx2y2 of Cu allowing a π*σ superexchange pathway to stabilize the singlet state of the Cu2/O2 complex and subsequently weaken the O-O bond resembling a bis-μ-oxo complex [54].

The oxidation of 2-aminophenol to o-quinone imine (Q) by abTy was explored using steady-state kinetic analysis (Figure 1). The similarity in KMO2 and KDO2 values (25±4 and 33±24 μM) using the bi-substrate Michaelis-Menten equation (2) suggests an equilibrium ordered mechanism for substrate binding, though the replot of the slopes in the 1/rate versus 1/[OAP] plot versus 1/[O2] clearly show the line does not pass through the origin, supporting that the difference in KMO2 and KDO2 values is relevant (Figures S.3. and S.4.) [55]. With KMO2<KDO2, there is synergistic binding of substrates suggesting 2-aminophenol binds tighter to the enzyme in the presence of oxygen to form Eoxy. The positive y-coordinate within the x, y intersection in the 1/rate versus 1/[oxygen] (figure S.2A.) indicates that the rate of O2 release is faster than that of Q release (k6 > kcat) which is further demonstrated through derivation of this rate constant in the table 1 (k6 = 97 and kcat = 75 sec−1). Therefore, the binding regime for substrates to the Edeoxy form is best described as a steady-state ordered mechanism with high error in the KDO2 measurement attributed to the tighter binding of the oxygen to the type-3 catalytic center compared with 2-aminophenol [49]. Transient state kinetics following the Eoxy form of abTy show the signal for the normally stable Eoxy active site architecture to be consumed in single exponential fashion in the presence of 2-aminophenol (Figures S.5. and 2). The rate for this process (k7) was determined to be over 15-fold faster than the corresponding rate for the Emet (k1) form of abTy derived from kcat/KMOAP (Table 2). The observation that k7k1 supports a model of preferred 2-aminophenol binding to Eoxy over Emet which is in stark contrast to work pursued with o-diphenol substrates [20]. The rate for k7 became an experimental constraint for the global fitting analysis of the rate constants to validate our proposed model of 2-aminophenol oxidation where k7k1 (Table 2). Using global analysis, the best correlation of the experimentally determined values for k7 was observed for the k7k1 simulation supporting our model that 2-aminophenol preferentially binds to the Eoxy rather than Emet form of abTy (Figures 3).

An independent measure of the increased reactivity of the o-aminophenol substrates for Eoxy over Emet was further investigated using solvent KIE (sKIE). Previous sKIE values measured at ambient oxygen concentrations show normal sKIE values and an accompanying decrease in fractionation factors from reactant to product state (ϕR ~1 −> ϕP < 1) for both mono- and di-phenolase activity that results from a single protonic interaction (n=1) [16, 17]. The behavior of the same exchangeable position in the abTy reaction coordinate reveal distinct differences from the reactant and product (and presumably the transition) states distinct to the reactions of both Emet and Eoxy states. Our present sKIE study sought to identify the distribution and relative rate limitation of bridging protonic interactions within the transition state structures for both Emet and Eoxy forms of the enzyme. To study this effect relative to the abTy enzyme forms present in 2–aminophenol oxidation (Emet and Eoxy), oxygen tension was varied with data represented as kcat/KMOAP (Figure 5). We chose catalytic efficiency (kcat/KMOAP) over kcat or KM based on the derivations present in table 1. Specifically, this term involves both the binding and chemical steps of 2-aminophenol oxidation and is well suited to discriminate proton transfer events between with Emet and Eoxy forms of abTy. This experimental design requires that the pH-rate profile be independent of the catalytic efficiency. The constant pH-rate profile demonstrates that kinetic data collected in either protiated or deuterated solvent are isolated to the di-coppers involved in catalysis and cannot be attributed to matrix perturbations associated with pH versus pD effects (Figure S.9.). Our results show a dramatic change in the magnitude of the sKIE as a function of oxygen concentration supporting a model in which the magnitude of the sKIE values are a direct transition state probe of μ-OH formation and consumption in the Emet state. Relative to the transition from Eoxy to Emet, the extent of μ-OH bond formation is represented as the inverse sKIE (D(kcat/KMOAP)<1) region of the curve at high [O2]. The increase toward the normal sKIE values (  D(kcat/KMOAP)>1 as [O2] decreases) demonstrate the partial rate limitation of μ-OH bond cleavage as Emet is converted to Edeoxy following protonation in the transition state. Compared with 2-aminophenol oxidase sKIE values, the EAS mechanism for phenol hydroxylation by Eoxy display the highest sKIE values and support an altered binding mode compared with o-diphenol substrates [5658]. Therefore, the transition from Eoxy to Emet is rate-limiting for phenol hydroxylation while it serves as a thermodynamic drive for the 2-aminophenol oxidase reaction.

Based on sKIE results from this study and others, a mechanism was postulated to address the unexpected preference of 2-aminophenol for Eoxy form of abTy. In Scheme 3, pose (A) has the 2-aminophenol substrate orienting itself with the anilinic group near His296. Structure (C) follows anilinic group de-protonation by His296. Based on crystal structure data for abTy and Bacillus megaterium tyrosinase there is precedence for numerous stable conformations and copper mobility present in type-3 dicopper architectures [7, 9, 59]. Therein, multiple metal ion binding modes for this structure were described through a cooperative model describing high copper ion plasticity controlling reactivity and substrate selectivity of the active site coppers is thought to limit CO to diphenolase activity and Hc to reversible oxygen binding. In support of copper ion plasticity and sampling between different enzyme forms, the reversible oxygen binding with Hc was altered through the addition of a denaturant (SDS) that moved blocking residues opening the active site to the catalytic oxidation of the diphenolic substrate dopamine while CO has copper coordinating histidine residues positioned in a conformationally labile loop regions [60, 61]. Based on our molecular modeling studies, the proposed binding orientations for 2-aminophenol and catechol are essentially superimposable with respect to benzene ring positioning near H296 and proximity to the CuA – CuB binding domains for both Emet and Eoxy forms of abTy (Figures 6 and 7). The similarity in spatial orientation providing a π−π binding interaction with His296 and the benzene ring proposes that any difference in reactivity is solely due to differences in the functional group substitution of aniline for a phenolic moiety within the o-diphenol substrate scaffold.

Scheme 3.

Scheme 3

Proposed mechanism for the tyrosinase dependent oxidation of 2-aminophenol to o-quinone imine.

Beginning with the Emet form of the enzyme, we propose a mechanism in which the μ-phenylamine of 2-aminophenol coordinates the dicopper domain resulting in a conformational change in the tertiary structure. Similar to reversible O2 binding in the type-3 proteins, the μ-phenylamine structure would maintain the appropriate superexchange pathway while increasing the hydroxide ion character of the solvent derived μ-OH allowing the capture of a phenolic proton by His61 allowing reduction of the copper domains to Edeoxy with concomitant oxidation to the o-quinone imine (Q) product (Complex C). The resulting Edeoxy form binds molecular oxygen then 2-aminophenol where the bases in this form are the μ−η22 dicopper peroxide for the anilinic proton and His61 for the phenolic proton followed by coordination to the copper centers. The resulting 2e, 2H+ redox process yields the second o-quinone imine and water with His61 deprotonated to give the characteristic μ-OH of Emet (Complexes E and F).

From a combination of kinetic and molecular modeling coupled with solvent kinetic isotope effects, it appears that the Emet form has a greater degree of conformational sampling allowing stable quasi-productive binding modes to occur. The μ-phenylamine (complex B) is one postulated intermediate that could cause the reduction of the μ-OH of Emet to become slow or rate-limiting to the global catalytic model (Scheme 2). Additional support for conformational mobility in the abTy active site is found by correlating secondary structural elements of the protein with ligand and binding pocket surfaces (Figures 8 and 9). With respect to 2-aminophenol binding modes in both Emet and Eoxy, the anilinic moiety is buried in the surface accessible active-site. With a similar interaction with His296 for each enzyme form, the resulting loop extending from the α11 helix defines a significant portion of the substrate binding face. Following proton transfer, subsequent electronic changes in the active site would be expected to sufficiently affect the loop conformation allowing the phenolic group to position near the CuA domain for deprotonation with H61.

Figure 8.

Figure 8

Representation of the ligand and active site binding surface for 2-aminophenol and the Emet form of mushroom tyrosinase. Please note, 2-aminophenol is shown in blue, the binding pocket is covered with a grey with the loop (teal) extending from the α11-helix (green).

Figure 9.

Figure 9

Binding orientation of 2-aminophenol in the CuB domain of Eoxy form of mushroom tyrosinase. Please note, 2-aminophenol is shown in blue, the binding pocket is covered with a grey showing the loop (teal) extending from the α11-helix (green).

Following release of the o-quinone imine product, tandem coupling and cyclization of o-quinone imine (Q) with 2-aminophenol (OAP) to forms the isolated product 2-amino-3H-phenoxazin-3-one (APX) (Figure S.2. and 4). Previous work on melanin carried out by Riley [62] and dopamine o-quinone cyclization [63] explored intramolecular cyclization for compounds biologically relevant to melanogenesis. Based on proposed intermediates and projected λmax values, the spectroscopic signals observed at 380 and 432 nm corresponding to intermediate 2 and product in the nucleophilic intermolecular conjugate addition pathway for APX formation (Scheme 1). Using SVD, the rate of coupling is much faster than the corresponding cyclization (6.91 ± 0.3μM−1sec−1 and 2.59 ± 5.31E-4 sec−1) passing through a phenyliminocyclohexadione intermediate (Table 2). Using the sequential model proposed in scheme 1, the extracted data suggest the short lived o-quinone imine (Q) enzymatic product undergoes a Michael-type addition with 2-aminophenol resulting in an observed phenyliminocyclohexadione intermediate that cyclizes to the APX product (Figure 4).

In conclusion, this study shows the oxidative cyclocondensation of 2-aminohenol to undergo an abTy dependent oxidation to its corresponding o-quinone imine releasing a water molecule from both Emet and Eoxy followed by several conjugate addition intermediates leading to the eventual 2-amino-3H-phenoxazin-3-one (APX) product (Scheme 3). Within this reaction, the abTy-dependent oxidation of 2-aminophenol was preferentially carried out by the Eoxy rather than Emet forms of the enzyme. We suggest the additional stabilization of the Emet form by the anilinic group of 2-aminophenol relative to o-diphenol substrates as an explanation for the difference in reactivity (Scheme 3). The sKIE studies support this model as the rate limitation for proton transfer is greatest with the Emet form and lowest (inverse) with Eoxy. This suggests that μ-OH formation (Eoxy to Emet) is favorable compared the rate limiting transition of Emet to Edeoxy likely caused by the formation of a quasi-stable conformation of the Emet containing a μ-phenylamine bridge between CuA and CuB that mimics the bis-μ-oxo dicopper geometry observed in the Eoxy form of abTy. Overall, the present study characterized distinct mechanistic differences in the conventional mechanism between o-diphenolase and 2-aminophenol oxidase with abTy and followed the non-enzymatic cyclization through a previously undetected intermediate leading to the APX product.

Supplementary Material

supplement

Highlights.

  • Tyrosinase preferentially oxidizes 2-aminophenol (OAP) from the oxy-state

  • sKIE dependence on [O2] reflect rate limitation of μ-OH formation/consumption

  • Transient phenyliminocyclohexadione cyclooxidation intermediate detected

Acknowledgments

Neil R. McIntyre would like to acknowledge support from NIH-1SC3GM112558-01, NIH-RCMI grant 2G12MD007595-06, Louisiana Cancer Research Consortium (LCRC), NSF-Louis Stokes Alliance for Minority Participation (JM), Center for Undergraduate Research (CUR)(JS), Louisiana Board of Regents Research Commercialization and Education Enhancement Program (RCEEP)(GM), Department of Chemistry Work/Study. Edward W. Lowe, Jr. acknowledges NSF support through the CI-TraCS fellowship (OCI-1122919). Guangdi Wang would like to acknowledge support from Department of Defense grant W81XWH-04-1-0557, Office of Naval Research Grant N00014-99-1-0763, and LCRC. We would also like to acknowledge the insightful recommendations of Dr. Larry Byers (Tulane) during the preparation of this manuscript.

Appendix

Distribution Equations

EMETET=k2k6k8+k3k5k7k9[O2][S]+k2k5k7k9[O2][S]+k2k7k9[S]+k2k6k9/Δ
EMETSET=k1k6k8[S]+k6k8+k1k5k7k9[O2][S]2+k1k7k9[O2][S]+k1k6k9[S]/Δ
EDEOXYET=k1k3k6k8[S]+k3k6k8+k2k6k8+k1k3k7k9[S]2+k1k3k6k9[S]/Δ
EOXYETk1k3k5k8[O2][S]+k3k5k8[O2]+k2k5k8[O2]+k2k8+k1k3k5k9[O2][S]/Δ
EOXYSETk1k3k5k7[O2][S]2+k3k5k7[O2][S]+k2k5k7[O2][S]+k2k7[S]+k2k6/Δ

Please note the ET is total enzyme and represents the sum of all distribution equations for each enzyme form.

Rate of product formation

dQdt=k3[EMETS]+k9[EOXYS]
dQdt=k3k1k6k8[S]+k6k8+k1k5k7k9[O2][S]2+k1k7k9[O2][S]+k1k6k9[S]+k9k1k3k5k7[O2][S]2+k3k5k7[O2][S]+k2k5k7[O2][S]+k2k7[S]+k2k6k2k6k8+k3k5k7k9[O2][S]+k2k5k7k9[O2][S]+k2k7k9[S]+k2k6k9+k1k6k8[S]+k6k8+k1k5k7k9[O2][S]2+k1k7k9[O2][S]+k1k6k9[S]+k1k3k6k8[S]+k3k6k8+k2k6k8+k1k3k7k9[S ]2+k1k3k6k9[S]+k1k3k5k8[O2][S]+k3k5k8[O2]+k2k5k8[O2]+k2k8+k1k3k5k9[O2][S]+k1k3k5k7[O2][S]2+k3k5k7[O2][S]+k2k5k7[O2][S]+k2k7[S]+k2k6

If k3k2 AND k9k8

dQdt=k1k5k7k9[O2][S]2+k1k7k9[O2][S]+k1k6k9[S]+k9k1k3k5k7[O2][S]2+k3k5k7[O2][S]k3k5k7k9[O2][S]+k1k5k7k9[O2][S]2+k1k7k9[O2][S]+k1k6k9[S]+k1k3k7k9[S]2+k1k3k6k9[S]+k1k3k5k9[O2][S]+k1k3k5k7[O2][S]2+k3k5k7[O2][S]

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Toussaint O, Lerch K. Catalytic oxidation of 2-aminophenols and ortho hydroxylation of aromatic amines by tyrosinase. Biochemistry. 1987;26:8567–8571. doi: 10.1021/bi00400a011. [DOI] [PubMed] [Google Scholar]
  • 2.Munoz-Munoz JL, Garcia-Molina F, Garcia-Ruiz PA, Varon R, Tudela J, Rodriguez-Lopez JN, Garcia-Canovas F. Catalytic oxidation of o-aminophenols and aromatic amines by mushroom tyrosinase. Biochimica et biophysica acta. 2011;1814:1974–1983. doi: 10.1016/j.bbapap.2011.07.015. [DOI] [PubMed] [Google Scholar]
  • 3.Wright CI, Mason HS. The oxidation of catechol by tyrosinase. J Biol Chem. 1946;165:45–53. [PubMed] [Google Scholar]
  • 4.Mason HS. The chemistry of melanin; mechanism of the oxidation of dihydroxyphenylalanine by tyrosinase. J Biol Chem. 1948;172:83–99. [PubMed] [Google Scholar]
  • 5.Mason HS. The chemistry of melanin; mechanism of the oxidation of catechol by tyrosinase. J Biol Chem. 1949;181:803–812. [PubMed] [Google Scholar]
  • 6.Sanchez-Ferrer A, Rodriguez-Lopez JN, Garcia-Canovas F, Garcia-Carmona F. Tyrosinase: a comprehensive review of its mechanism. Biochimica et biophysica acta. 1995;1247:1–11. doi: 10.1016/0167-4838(94)00204-t. [DOI] [PubMed] [Google Scholar]
  • 7.Ismaya WT, Rozeboom HJ, Weijn A, Mes JJ, Fusetti F, Wichers HJ, Dijkstra BW. Crystal structure of Agaricus bisporus mushroom tyrosinase: identity of the tetramer subunits and interaction with tropolone. Biochemistry. 2011;50:5477–5486. doi: 10.1021/bi200395t. [DOI] [PubMed] [Google Scholar]
  • 8.Ismaya WT, Rozeboom HJ, Schurink M, Boeriu CG, Wichers H, Dijkstra BW. Crystallization and preliminary X-ray crystallographic analysis of tyrosinase from the mushroom Agaricus bisporus. Acta Crystallogr Sect F Struct Biol Cryst Commun. 2011;67:575–578. doi: 10.1107/S174430911100738X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Matoba Y, Kumagai T, Yamamoto A, Yoshitsu H, Sugiyama M. Crystallographic evidence that the dinuclear copper center of tyrosinase is flexible during catalysis. J Biol Chem. 2006;281:8981–8990. doi: 10.1074/jbc.M509785200. [DOI] [PubMed] [Google Scholar]
  • 10.Sendovski M, Kanteev M, Shuster Ben-Yosef V, Adir N, Fishman A. Crystallization and preliminary X-ray crystallographic analysis of a bacterial tyrosinase from Bacillus megaterium. Acta Crystallogr Sect F Struct Biol Cryst Commun. 66:1101–1103. doi: 10.1107/S1744309110031520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Sendovski M, Kanteev M, Ben-Yosef VS, Adir N, Fishman A. First structures of an active bacterial tyrosinase reveal copper plasticity. Journal of molecular biology. 405:227–237. doi: 10.1016/j.jmb.2010.10.048. [DOI] [PubMed] [Google Scholar]
  • 12.Cuff ME, Miller KI, van Holde KE, Hendrickson WA. Crystal structure of a functional unit from Octopus hemocyanin. Journal of molecular biology. 1998;278:855–870. doi: 10.1006/jmbi.1998.1647. [DOI] [PubMed] [Google Scholar]
  • 13.Klabunde T, Eicken C, Sacchettini JC, Krebs B. Crystal structure of a plant catechol oxidase containing a dicopper center. Nature structural biology. 1998;5:1084–1090. doi: 10.1038/4193. [DOI] [PubMed] [Google Scholar]
  • 14.Ramsden CA, Riley PA. Tyrosinase: the four oxidation states of the active site and their relevance to enzymatic activation, oxidation and inactivation. Bioorganic & medicinal chemistry. 2014;22:2388–2395. doi: 10.1016/j.bmc.2014.02.048. [DOI] [PubMed] [Google Scholar]
  • 15.Chen P, Solomon EI. O2 activation by binuclear Cu sites: noncoupled versus exchange coupled reaction mechanisms. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:13105–13110. doi: 10.1073/pnas.0402114101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Fenoll LG, Penalver MJ, Rodriguez-Lopez JN, Garcia-Ruiz PA, Garcia-Canovas F, Tudela J. Deuterium isotope effect on the oxidation of monophenols and o-diphenols by tyrosinase. Biochem J. 2004;380:643–650. doi: 10.1042/BJ20040136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Penalver MJ, Rodriguez-Lopez JN, Garcia-Ruiz PA, Garcia-Canovas F, Tudela J. Solvent deuterium isotope effect on the oxidation of o-diphenols by tyrosinase. Biochimica et biophysica acta. 2003;1650:128–135. doi: 10.1016/s1570-9639(03)00208-5. [DOI] [PubMed] [Google Scholar]
  • 18.Penalver MJ, Hiner AN, Rodriguez-Lopez JN, Garcia-Canovas F, Tudela J. Mechanistic implications of variable stoichiometries of oxygen consumption during tyrosinase catalyzed oxidation of monophenols and o-diphenols. Biochimica et biophysica acta. 2002;1597:140–148. doi: 10.1016/s0167-4838(02)00264-9. [DOI] [PubMed] [Google Scholar]
  • 19.Fenoll LG, Rodriguez-Lopez JN, Garcia-Sevilla F, Garcia-Ruiz PA, Varon R, Garcia-Canovas F, Tudela J. Analysis and interpretation of the action mechanism of mushroom tyrosinase on monophenols and diphenols generating highly unstable o-quinones. Biochimica et biophysica acta. 2001;1548:1–22. doi: 10.1016/s0167-4838(01)00207-2. [DOI] [PubMed] [Google Scholar]
  • 20.Rodriguez-Lopez JN, Fenoll LG, Garcia-Ruiz PA, Varon R, Tudela J, Thorneley RN, Garcia-Canovas F. Stopped-flow and steady-state study of the diphenolase activity of mushroom tyrosinase. Biochemistry. 2000;39:10497–10506. doi: 10.1021/bi000539+. [DOI] [PubMed] [Google Scholar]
  • 21.Munoz-Munoz JL, Garcia-Molina Mdel M, Garcia-Molina F, Garcia-Ruiz PA, Garcia-Sevilla F, Rodriguez-Lopez JN, Garcia-Canovas F. Deuterium isotope effect on the suicide inactivation of tyrosinase in its action on o-diphenols. IUBMB life. 2013;65:793–799. doi: 10.1002/iub.1191. [DOI] [PubMed] [Google Scholar]
  • 22.Munoz-Munoz JL, Garcia-Molina F, Varon R, Tudela J, Garcia-Canovas F, Rodriguez-Lopez JN. Kinetic cooperativity of tyrosinase. A general mechanism. Acta biochimica Polonica. 2011;58:303–311. [PubMed] [Google Scholar]
  • 23.Jackman MP, Huber M, Hajnal A, Lerch K. Stabilization of the oxy form of tyrosinase by a single conservative amino acid substitution. Biochem J. 1992;282(Pt 3):915–918. doi: 10.1042/bj2820915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Jolley RL, Jr, Evans LH, Makino N, Mason HS. Oxytyrosinase. J Biol Chem. 1974;249:335–345. [PubMed] [Google Scholar]
  • 25.Makino N, Mason HS. Reactivity of oxytyrosinase toward substrates. J Biol Chem. 1973;248:5731–5735. [PubMed] [Google Scholar]
  • 26.Osako T, Ohkubo K, Taki M, Tachi Y, Fukuzumi S, Itoh S. Oxidation mechanism of phenols by dicopper-dioxygen (Cu(2)/O(2)) complexes. J Am Chem Soc. 2003;125:11027–11033. doi: 10.1021/ja029380+. [DOI] [PubMed] [Google Scholar]
  • 27.Matsumoto T, Furutachi H, Kobino M, Tomii M, Nagatomo S, Tosha T, Osako T, Fujinami S, Itoh S, Kitagawa T, Suzuki M. Intramolecular arene hydroxylation versus intermolecular olefin epoxidation by (mu-eta2:eta2-peroxo)dicopper(II) complex supported by dinucleating ligand. J Am Chem Soc. 2006;128:3874–3875. doi: 10.1021/ja058117g. [DOI] [PubMed] [Google Scholar]
  • 28.Quinn DM, Sutton LD. Theoretical basis and mechanistic utility of solvent isotope effects. In: Cook PF, editor. Enzyme Mechanisms from Isotope Effects. CRC Press; Boca Raton, Fl: 1991. pp. 73–126. [Google Scholar]
  • 29.Kresge AJ, More O’Ferrall RA, Powell MF. Solvent isotope effects, fractionation factors and mechanisms of proton transfer reactions. In: Buncel E, Lee CC, editors. Isotopes in Organic Chemistry. Vol. 7. Elsevier; Amsterdam: 1987. pp. 177–273. [Google Scholar]
  • 30.Alvarez FJ, Schowen RL. Mechanistic deductions from solvent isotope effects. In: Buncel E, Lee CC, editors. Isotopes in Organic Chemistry. Elsevier; Amsterdam: 1987. pp. 1–60. [Google Scholar]
  • 31.Venkatasubban KS, Schowen RL. The proton inventory technique. Crit Rev Biochem. 1984;17:1–44. doi: 10.3109/10409238409110268. [DOI] [PubMed] [Google Scholar]
  • 32.Schowen KBJ, Schowen RL. Solvent isotope effects on enzyme systems. Methods in enzymology. 1982;87:551–606. [PubMed] [Google Scholar]
  • 33.Schowen KBJ. Solvent hydrogen isotope effects. In: Gandour RD, Schowen RL, editors. Transition State of Biochemical Processes. Plenum; New York: 1978. pp. 225–283. [Google Scholar]
  • 34.Schowen RL. Solven isotope effects on enzyme reactions. In: Cleland WW, O’Leary MH, Northrop DB, editors. Isotope Effects on Enzyme-Catalyzed Reactions. University Park Press; Baltimore, MD: 1977. pp. 64–99. [Google Scholar]
  • 35.Barry CE, 3rd, Nayar PG, Begley TP. Phenoxazinone synthase: mechanism for the formation of the phenoxazinone chromophore of actinomycin. Biochemistry. 1989;28:6323–6333. doi: 10.1021/bi00441a026. [DOI] [PubMed] [Google Scholar]
  • 36.Bruyneel F, Payen O, Rescigno A, Tinant B, Marchand-Brynaert J. Laccase-mediated synthesis of novel substituted phenoxazine chromophores featuring tuneable water solubility. Chemistry. 2009;15:8283–8295. doi: 10.1002/chem.200900681. [DOI] [PubMed] [Google Scholar]
  • 37.Suzuki H, Furusho Y, Higashi T, Ohnishi Y, Horinouchi S. A novel o-aminophenol oxidase responsible for formation of the phenoxazinone chromophore of grixazone. J Biol Chem. 2006;281:824–833. doi: 10.1074/jbc.M505806200. [DOI] [PubMed] [Google Scholar]
  • 38.Smith AW, Camara-Artigas A, Wang M, Allen JP, Francisco WA. Structure of phenoxazinone synthase from Streptomyces antibioticus reveals a new type 2 copper center. Biochemistry. 2006;45:4378–4387. doi: 10.1021/bi0525526. [DOI] [PubMed] [Google Scholar]
  • 39.Kaizer J, Barath G, Csonka R, Speier G, Korecz L, Rockenbauer A, Parkanyi L. Catechol oxidase and phenoxazinone synthase activity of a manganese(II) isoindoline complex. Journal of inorganic biochemistry. 2008;102:773–780. doi: 10.1016/j.jinorgbio.2007.11.014. [DOI] [PubMed] [Google Scholar]
  • 40.Freeman JC, Nayar PG, Begley TP, Villafranca JJ. Stoichiometry and spectroscopic identity of copper centers in phenoxazinone synthase: a new addition to the blue copper oxidase family. Biochemistry. 1993;32:4826–4830. doi: 10.1021/bi00069a018. [DOI] [PubMed] [Google Scholar]
  • 41.Duckworth HW, Coleman JE. Physicochemical and kinetic properties of mushroom tyrosinase. J Biol Chem. 1970;245:1613–1625. [PubMed] [Google Scholar]
  • 42.Powell A, Siu N, Inlow JK, Flurkey WH. Immobilized Metal Affinity Chromatography (IMAC) of Mushroom Tyrosinase. Journal of Chemistry. 2007;1:1–13. [Google Scholar]
  • 43.Smith PK, Krohn RI, Hermanson GT, Mallia AK, Gartner FH, Provenzano MD, Fujimoto EK, Goeke NM, Olson BJ, Klenk DC. Measurement of protein using bicinchoninic acid. Anal Biochem. 1985;150:76–85. doi: 10.1016/0003-2697(85)90442-7. [DOI] [PubMed] [Google Scholar]
  • 44.McIntyre NR, Lowe EW, Jr, Chew GH, Owen TC, Merkler DJ. Thiorphan, tiopronin, and related analogs as substrates and inhibitors of peptidylglycine alpha-amidating monooxygenase (PAM) FEBS Lett. 2006;580:521–532. doi: 10.1016/j.febslet.2005.12.058. [DOI] [PubMed] [Google Scholar]
  • 45.Goodwin DC, Yamazaki I, Aust SD, Grover TA. Determination of rate constants for rapid peroxidase reactions. Analytical biochemistry. 1995;231:333–338. doi: 10.1006/abio.1995.0059. [DOI] [PubMed] [Google Scholar]
  • 46.Johnson KA. Fitting enzyme kinetic data with KinTek Global Kinetic Explorer. Methods in enzymology. 2009;467:601–626. doi: 10.1016/S0076-6879(09)67023-3. [DOI] [PubMed] [Google Scholar]
  • 47.Johnson KA, Simpson ZB, Blom T. FitSpace explorer: an algorithm to evaluate multidimensional parameter space in fitting kinetic data. Analytical biochemistry. 2009;387:30–41. doi: 10.1016/j.ab.2008.12.025. [DOI] [PubMed] [Google Scholar]
  • 48.Johnson KA, Simpson ZB, Blom T. Global kinetic explorer: a new computer program for dynamic simulation and fitting of kinetic data. Analytical biochemistry. 2009;387:20–29. doi: 10.1016/j.ab.2008.12.024. [DOI] [PubMed] [Google Scholar]
  • 49.Rodriguez-Lopez JN, Ros JR, Varon R, Garcia-Canovas F. Oxygen Michaelis constants for tyrosinase. Biochem J. 1993;293(Pt 3):859–866. doi: 10.1042/bj2930859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Solomon EI, Ginsbach JW, Heppner DE, Kieber-Emmons MT, Kjaergaard CH, Smeets PJ, Tian L, Woertink JS. Copper dioxygen (bio)inorganic chemistry. Faraday discussions. 2011;148:11–39. doi: 10.1039/c005500j. discussion 97–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Solomon EI, Heppner DE, Johnston EM, Ginsbach JW, Cirera J, Qayyum M, Kieber-Emmons MT, Kjaergaard CH, Hadt RG, Tian L. Copper active sites in biology. Chemical reviews. 2014;114:3659–3853. doi: 10.1021/cr400327t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Shearer J, Zhang CX, Zakharov LN, Rheingold AL, Karlin KD. Substrate oxidation by copper-dioxygen adducts: mechanistic considerations. J Am Chem Soc. 2005;127:5469–5483. doi: 10.1021/ja045191a. [DOI] [PubMed] [Google Scholar]
  • 53.Holm RH, Kennepohl P, Solomon EI. Structural and Functional Aspects of Metal Sites in Biology. Chemical reviews. 1996;96:2239–2314. doi: 10.1021/cr9500390. [DOI] [PubMed] [Google Scholar]
  • 54.Kieber-Emmons MT, Ginsbach JW, Wick PK, Lucas HR, Helton ME, Lucchese B, Suzuki M, Zuberbuhler AD, Karlin KD, Solomon EI. Observation of a Cu(II)(2) (mu-1,2-peroxo)/Cu(III)(2) (mu-oxo)(2) equilibrium and its implications for copper-dioxygen reactivity. Angew Chem Int Ed Engl. 2014;53:4935–4939. doi: 10.1002/anie.201402166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Cleland WW. Statistical analysis of enzyme kinetic data. Methods in enzymology. 1979;63:103–138. doi: 10.1016/0076-6879(79)63008-2. [DOI] [PubMed] [Google Scholar]
  • 56.Molina FG, Munoz JL, Varon R, Lopez JN, Canovas FG, Tudela J. An approximate analytical solution to the lag period of monophenolase activity of tyrosinase. Int J Biochem Cell Biol. 2007;39:238–252. doi: 10.1016/j.biocel.2006.08.007. [DOI] [PubMed] [Google Scholar]
  • 57.Naish-Byfield S, Riley PA. Tyrosinase autoactivation and the problem of the lag period. Pigment Cell Res. 1998;11:127–133. doi: 10.1111/j.1600-0749.1998.tb00722.x. [DOI] [PubMed] [Google Scholar]
  • 58.Olivares C, Solano F. New insights into the active site structure and catalytic mechanism of tyrosinase and its related proteins. Pigment cell & melanoma research. 2009;22:750–760. doi: 10.1111/j.1755-148X.2009.00636.x. [DOI] [PubMed] [Google Scholar]
  • 59.Sendovski M, Kanteev M, Ben-Yosef VS, Adir N, Fishman A. First structures of an active bacterial tyrosinase reveal copper plasticity. Journal of molecular biology. 2011;405:227–237. doi: 10.1016/j.jmb.2010.10.048. [DOI] [PubMed] [Google Scholar]
  • 60.Cong Y, Zhang Q, Woolford D, Schweikardt T, Khant H, Dougherty M, Ludtke SJ, Chiu W, Decker H. Structural mechanism of SDS-induced enzyme activity of scorpion hemocyanin revealed by electron cryomicroscopy. Structure. 2009;17:749–758. doi: 10.1016/j.str.2009.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Baird S, Kelly SM, Price NC, Jaenicke E, Meesters C, Nillius D, Decker H, Nairn J. Hemocyanin conformational changes associated with SDS-induced phenol oxidase activation. Biochimica et biophysica acta. 2007;1774:1380–1394. doi: 10.1016/j.bbapap.2007.08.019. [DOI] [PubMed] [Google Scholar]
  • 62.Land EJ, Ramsden CA, Riley PA. An MO study of regioslelctive amine addition to ortho-quinones relevant to mealnogenesis. Tetrahedron Lett. 2006;62:4884–4891. [Google Scholar]
  • 63.Borovansky J, Edge R, Land EJ, Navaratnam S, Pavel S, Ramsden CA, Riley PA, Smit NP. Mechanistic studies of melanogenesis: the influence of N-substitution on dopamine quinone cyclization. Pigment Cell Res. 2006;19:170–178. doi: 10.1111/j.1600-0749.2006.00295.x. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

supplement

RESOURCES