Introduction

The catalytic oxidation of ortho-diphenols to quinones is the only known activity of catechol oxidase [1] and one of the activities of tyrosinase, which is also able to oxidize monophenols to quinones [2]. Dioxygen reduction to water is coupled to the oxidation of two catechol molecules in these processes. Both types of enzyme contain a dinuclear, type 3 Cu center in which each metal ion is coordinated to three histidine ligands. Important details of the mechanism by which catechol oxidase and tyrosinase oxidize catechols are completely unknown. In particular, the mode of substrate binding and product release from the enzyme, and the mode of dioxygen activation and cleavage at the dinuclear copper center, are still under debate [1, 3, 4]. Catecholase activity has been recently associated also with bovine brain Cu-S100B, a small neural protein which binds two Cu ions with type 2 spectroscopic features [5]. However, in this case the ratio between quinone produced and dioxygen consumed is 1:1, and the latter is reduced to hydrogen peroxide. Several model Cu complexes exhibit catecholase activity [6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18], but thorough mechanistic studies are scarce [11, 12, 13, 14, 15, 16], and in particular only in a few cases has the mode of reduction of dioxygen, to hydrogen peroxide [12, 13, 14] or water [15, 16], been established. The dinuclear copper(II) complex derived from the ligand L55 (α,α′-bis{bis[1-(1′-methyl-2′-benzimidazolyl)methyl]amino}-m-xylene), [Cu2L55]4+, appears to exhibit the highest activity in the catalytic oxidation of catechols among the available Cu systems [15]. An important feature of the reaction mediated by [Cu2L55]4+ is the possibility to separate the two steps of the catecholase catalytic cycle corresponding to oxidation of the two molecules of catechol (Scheme 1). Herein we report mechanistic details on these two steps, which provide new insight into the catecholase activity of type 3 Cu enzymes.

structure 1
Scheme 1
scheme 1

Simplified mechanism of the catechol oxidase catalytic cycle

Materials and methods

The complexes [Cu2L55(H2O)2][ClO4]4 and [Cu2L55(CH3CN)2][ClO4]2 were prepared as described previously [19]. The catalytic oxidation of 3,5-di-tert-butylcatechol (DTBCH2) by [Cu2L55]4+ and O2 was studied at various temperatures in a mixed solvent of 75% methanol/glycerol (7/1 v/v) and 25% (v/v) aqueous 50 mM Hepes buffer at pH 6.8, saturated with atmospheric oxygen. The reactions were followed through the development of the 3,5-di-tert-butylquinone (DTBQ) band at 412 nm (ε 1550 M−1 cm−1 in the above medium). In each experiment, the curve of quinone absorption versus time was found to be biphasic: a fast initial increase of absorbance, corresponding to the amount of copper catalyst present, was followed by a slower increase due to the catalytic reaction (Fig. 1).

Fig. 1
figure 1

Biphasic trace of absorbance versus time curves in typical kinetic experiments of DTBCH2 oxidation in the mixed solvent of 75% methanol/glycerol (7/1) and 25% (v/v) Hepes buffer (50 mM) at 20±0.1 °C; mixing of the reagents was performed with the stopped-flow apparatus: (a) 0.050 mM [Cu2(L55)]4+; (b) 0.025 mM [Cu2(L55)]4+; in both cases, [DTBCH2] was 4 mM

The first step of DTBCH2 oxidation was studied with an Applied Photophysics model RS-1000 stopped-flow instrument, with 1 ms dead time and 1-cm pathlength cell, coupled with a HP 8452A diode array spectrophotometer. The solution of [Cu2L55]4+ (10 μM) in the aqueous/organic solvent was mixed with solutions of variable concentrations of DTBCH2 (0.2–4 mM) in the same solvent. Upon mixing, the final concentrations of the complex and DTBCH2 was reduced to one half of the original concentrations. The rate dependence on [DTBCH2] was studied in the range from 14 to 33 °C; experiments at lower temperature were hampered by tarnish problems of the cuvette and at higher temperatures the reactions were too fast to give reliable rate data. Temperature was controlled at ±0.1 °C with a thermostat. All the experiments were repeated at least four times. The rates in ΔA/s units were calculated from the slope of the trace at 412 nm in the first 2–3 s after mixing of the reagents; the conversion from ΔA/s to s−1 units was made using the quinone extinction coefficient and the complex concentration. The kinetic parameter k1 and the thermodynamic dissociation constant KD1 were obtained by fitting of the rate versus [DTBCH2] plots. The above procedure was repeated, changing the pH of the aqueous buffer added to the organic solvent from 6.0 to 8.5. The experiments were performed at constant DTBCH2 (2 mM) and catalyst (6 μM) concentrations, at 20±0.1 °C. For determining the exact pH of the mixed solvent, the glass electrode was calibrated through the addition of measured quantities of standard perchloric acid solution to a 75% methanol/glycerol (7/1 v/v) and 25% (v/v) water solution, at the desired temperature, according to Gran’s method [20]. Blank experiments were carried out in the absence of copper catalyst, but no significant catechol autoxidation was detected in the time range investigated.

The second step of the reaction was studied using a diode array spectrophotometer, in a magnetically stirred and thermostatted (at ±0.1 °C) 1-cm pathlength cell. The rate dependence of [DTBCH2] was determined by maintaining the concentration of [Cu2L55]4+ at 2.5 μM and varying the substrate concentration from 0.05 to 5 mM. The temperature dependence of the rate was studied in the 0–27 °C range. All measurements were performed in duplicate. In all cases the reactions showed a biphasic behavior. Since the first step was completed in the mixing time of the solutions, reliable data could be obtained only for the second step, taking the data in the time interval of 5–20 s, where the absorbance variation was linear. At long reaction times, some inactivation of the catalyst was noted. The conversion from ΔA/s to s−1 units was made using the quinone extinction coefficient and the complex concentration, and then dividing the value obtained by 2; the latter operation is required because in each catalytic cycle two quinone molecules are produced (see below). The kinetic parameter k2 and the dissociation constant KD2 were obtained by fitting of the rate versus [DTBCH2] plots.

The above procedure was repeated, changing the pH of the aqueous buffer added to the organic solvent from 6.0 to 8.5. The experiments were performed at constant DTBCH2 (2 mM) and catalyst (6 μM) concentrations, at 20±0.1 °C.

The reaction rates for the two steps of the reaction were also determined with an alternative method. The absorbance at 412 nm versus time profile was fitted to the following equation (see Appendix for details):

$$ {\text{Abs}} = \varepsilon _{{\text{P}}} {\text{ $ \times $ }}\frac{{k_{{{\text{obs1}}}} \times {\left[ {{\text{Cu}}_{{\text{2}}} {\text{L55}}} \right]}}} {{k_{{{\text{obs1}}}} + k_{{{\text{obs2}}}} }} \times {\left\{ {{\text{2}} \times k_{{{\text{obs1}}}} \times t - \frac{{k_{{{\text{obs1}}}} - k_{{{\text{obs2}}}} }} {{k_{{{\text{obs1}}}} + k_{{{\text{obs2}}}} }} \times {\left[ {{\text{1}} - {\text{exp}}{\left[ { - {\left( {k_{{{\text{obs1}}}} + k_{{{\text{obs2}}}} } \right)} \times t} \right]}} \right]}} \right\}} $$
(1)

where P represents the quinone, εP its molar extinction coefficient at the reading wavelength, and kobs1 and kobs2 the observed kinetic constants for the first and second steps, respectively, so that the rate of each step can be obtained by directly multiplying the observed constant with the [Cu2L55]4+ concentration. This method gave reliable data only using the stopped-flow mixing apparatus.

The experiments at variable O2 concentration were also performed using the stopped-flow apparatus; the concentrations of DTBCH2 and [Cu2L55]4+ were maintained fixed at 4 mM and 2.5 μM, respectively, in the cuvette in the usual solvent, at 20±0.1 °C. Solutions of both the substrate and catalyst containing the desired O2 concentration were prepared by mixing appropriate volumes of a deoxygenated solution and an O2-saturated (1 atm) solution of the compounds. Deoxygenation and O2 saturation of the solutions were performed by bubbling for 20 min pre-equilibrated (with the same solvent) streams of argon or dioxygen, respectively, at constant temperature.

The determination of dissolved dioxygen in the solvent mixture employed for the kinetic measurements was made at different temperatures, three times at each temperature, after equilibration with air, by Winkler’s method [21]. The experimental dioxygen concentration ranged from 0.57 mM at 12 °C to 0.50 mM at 27 °C.

The low-temperature experiments were performed using a custom-designed immersible fiber-optic quartz probe (5 mm pathlength, Hellma) fitted to a Schlenk vessel with gas and vacuum inlets and connected to the diode array spectrophotometer. The catechol complex of [Cu2L55]4+ was obtained by adding to an anaerobic solution of the complex (0.2 mM) in degassed methanol at −90 °C a chilled, degassed methanolic solution of DTBCH2 (1.8 mM final concentration). The solution was kept at low temperature for 3 h, recording spectra from time to time, and then it was slowly warmed to room temperature. A similar experiment was carried out using [Cu2L55]2+ and DTBCH2, but no appreciable spectral changes were detected. The catechol complex of [Cu2L55]2+ could be instead characterized by NMR, by recording anaerobic spectra of the complex in CD3OD at low temperature in the absence and in the presence of 10 molar equiv DTBCH2. The experiments attempting to characterize the ternary complex of [Cu2L55]2+-O2-DTBC were carried out by adding to a degassed methanolic solution of [Cu2L55(CH3CN)2]2+ at −90 °C, degassed methanolic solutions of DTBCH2 (to final concentrations up to 3.2 mM), and then adding chilled pure dioxygen (1 atm) to the cold solution and recording optical spectra.

Results and discussion

The CuI and CuII complexes of the dinucleating ligand L55 constitute one of the few biomimetic systems exhibiting phenol monooxygenase activity [22] and catecholase activity [15]. Both the CuI and CuII complexes have been recently structurally characterized [23]. The previous study of the catalytic oxidation of DTBCH2 by [Cu2L55]4+ was carried out at pH 5.1 [15]. The mixed solvent of 75% methanol/glycerol (7/1 v/v) and 25% (v/v) aqueous 50 mM Hepes buffer allows good solubility of the reagents, thus extending the investigation of the catecholase reaction over a range of temperatures and pH. Also in this aqueous/organic solvent, the catecholase reaction by [Cu2L55]4+ proceeds with a biphasic rate of DTBQ development (Fig. 1). The initial phase is faster and involves the stoichiometric oxidation of one molecule of catechol per molecule of [Cu2L55]4+; this is followed by a second phase which involves the true catalytic process. At longer reaction times, some catalyst degradation occurs; the complex performs at least 70 cycles, at room temperature, before significant inactivation. This process is probably related to the presence of a large amount of water (25% v/v) in the solvent; in fact, when the reaction was performed in a 30/1 (v/v) methanol/buffer solution, no significant degradation was observed, even at longer reaction times [15].

We used two different methods for the analysis of the data. In the first one the two steps were analyzed separately; the initial rate was used to obtain the observed rate constant for the stoichiometric process, while the turnover rate in the catalytic phase was used for the second step. In the second method, the rate constants were obtained by fitting the development of the quinone absorbance with time to Eq. 1. In spite of the differences in the two methods of analysis, the results obtained were identical.

The observed rate constants for both the reaction steps were found to exhibit hyperbolic dependence on DTBCH2 concentration (Fig. 2). In addition, the rate of the first step was unaffected by changing the dioxygen concentration of the reaction solution (data not shown). In contrast, the rate of the second step depends linearly on [O2] (Fig. 3), independent of the catechol concentration. The nonzero value for the rate in the absence of dioxygen observed in the plot of Fig. 3 is probably due to unavoidable leaks in the transfer of the deoxygenated and partially oxygenated solutions. These results can be rationalized on the basis of the following mechanism:

Fig. 2A, B
figure 2

Rate dependence of the oxidation of DTBCH2 catalyzed by [Cu2L55)]4+ on catechol concentration in the mixed solvent of 75% methanol/glycerol (7/1) and 25% (v/v) Hepes buffer (50 mM, pH 6.8) at different temperatures. A First phase: mixing of the reagents was performed with the stopped-flow apparatus; 5 μM [Cu2L55)]4+; fitting of the curve was performed with Eq. 2. B Second phase: 2.5 μM [Cu2L55)]4+; fitting of the curve was performed with Eq. 3

Fig. 3
figure 3

Rate dependence for the second step of the oxidation of DTBCH2 (4 mM) catalyzed by [CuL55]4+ (2.5 μM) on dioxygen concentration in the mixed solvent of 75% methanol/glycerol (7/1) and 25% (v/v) Hepes buffer (50 mM) at pH 6.8 at 20±0.1 °C

  1. 1.

    First Step:

$$ {\text{[Cu}}^{{\text{II}}} {\text{ - OH - Cu}}^{{\text{II}}} {\text{]}}^{{\text{3 + }}} {\text{ + DTBCH}}_{\text{2}} {\text{ }} \rightleftarrows {\text{ [Cu}}^{{\text{II}}} _{\text{2}} {\text{ - DTBC]}}^{{\text{2 + }}} {\text{ + H}}_{\text{2}} {\text{O + H}}^{\text{ + }} $$
(Ia)
$$ {\text{[Cu}}^{{\text{II}}} _{\text{2}} {\text{ - DTBC]}}^{{\text{2 + }}} {\text{ }} \to {\text{ [Cu}}^{\text{I}} _{\text{2}} {\text{]}}^{{\text{2 + }}} {\text{ + DTBQ}} $$
(Ib)
  1. 2.

    Second step:

$$ {\text{[Cu}}^{\text{I}} _{\text{2}} {\text{]}}^{{\text{2 + }}} {\text{ + DTBCH}}_{\text{2}} {\text{ }} \rightleftarrows {\text{ [Cu}}^{\text{I}} _{\text{2}} {\text{ - DTBC] + 2H}}^{\text{ + }} $$
(IIa)
$$ {\text{[Cu}}^{\text{I}} _{\text{2}} {\text{ - DTBC] + O}}_{\text{2}} {\text{ }} \rightleftarrows {\text{ [Cu}}_{\text{2}} {\text{ - O}}_{\text{2}} {\text{ - DTBC] }} $$
(IIb)
$$ {\text{[Cu}}_{{\text{2}}} {\text{/O}}_{{\text{2}}} {\text{/DTBC]}}\xrightarrow{{\,\,\,}}\,\,\,\xrightarrow{{{\text{3H}}^{{\text{ + }}} }}{\text{ [Cu}}^{{{\text{II}}}} {\text{ - OH - Cu}}^{{{\text{II}}}} {\text{]}}^{{{\text{3 + }}}} {\text{ + DTBQ + H}}_{{\text{2}}} {\text{O}} $$
(IIc)

The first phase of the reaction is connected to reactions Ia and Ib and consists of binding catechol to the [Cu2L55]4+ complex, followed by electron transfer to give the quinone. The rate is given by the equation (see Appendix):

$$ {\text{rate}} = \frac{{k_{\text{1}} \times {\text{[DTBCH}}_{\text{2}} {\text{]}} \times {\text{[complex]}}}} {{K_{{\text{D1}}} + {\text{[DTBCH}}_{\text{2}} {\text{]}}}} $$
(2)

where KD1 represents the dissociation constant for the [CuII2/DTBC] adduct, i.e. reverse of reaction Ia, and k1 the rate constant for reaction Ib.

In reaction Ia, the active form of the [Cu2L55]4+ catalyst is represented as a monohydroxo species. In fact, the [Cu2L55]4+ complex undergoes pH-dependent dissociation equilibria from a moderately coupled bis(aqua)-dicopper(II) form to a strongly magnetically coupled bis(μ-hydroxo)-dicopper(II) form [23]:

$$ {\text{[Cu}}_{\text{2}} {\text{L55(H}}_{\text{2}} {\text{O)}}_{\text{2}} {\text{]}}^{{\text{4 + }}} {\text{ }} \rightleftarrows {\text{ [Cu}}_{\text{2}} {\text{L55(H}}_{\text{2}} {\text{O)(OH)]}}^{{\text{3 + }}} {\text{ + H}}^{\text{ + }} $$
(III)
$$ {\text{[Cu}}_{\text{2}} {\text{L55(H}}_{\text{2}} {\text{O)(OH)]}}^{{\text{3 + }}} {\text{ }} \rightleftarrows {\text{ [Cu}}_{\text{2}} {\text{L55(OH)}}_{\text{2}} {\text{]}}^{{\text{2 + }}} {\text{ + H}}^{\text{ + }} $$
(IV)

with the intermediate monohydroxo species appreciably present between pH 6 and 8, but in lower than 20% abundance. The pKa values for dissociation of bound water in reactions III and IV were determined previously as 6.93 and 6.25, respectively (in acetonitrile/water, 4/1 v/v) [23]. The bis(μ-dimethoxo)-dicopper(II) complex derived from [Cu2L55(OH)2]2+ by water–methanol interchange has been characterized by X-ray crystallography [23]. When the rate of the first step of the catecholase reaction was studied as a function of the pH of the medium, it was found that the rate was maximal around pH 7, but it dropped dramatically above pH≈7.5 (Fig. 4). Taking into account that deprotonation of DTBCH2 is favored at higher pH, the rate profile parallels the distribution curve of the monohydroxo species [Cu2L55(H2O)(OH)]3+ [23], while the bis(μ-hydroxo) species [Cu2L55(OH)2]2+, which is dominant above pH≈6.5, is catalytically inactive. This is probably due to the difficulty of the catechol to attack the tightly dibridged Cu2(OH)2 core. For the same reason, an analogous monohydroxo-dicopper(II) intermediate was found to be critical for hydrogen peroxide binding and activation in the biomimetic dicopper(II) complex with the closely related ligand L66 [24]. It is interesting to note that a monohydroxo-bridged dicopper(II) center is also contained in the structure of oxidized catechol oxidase [25] and is assumed to be present in met-tyrosinase [26]. The bridging hydroxo group may be essential to remove the second proton of the catechol residue and favor the binding of the substrate in a bridging mode (Scheme 2).

Fig. 4
figure 4

Rate dependence for the first (open squares) and second (solid circles) steps of the oxidation of DTBCH2 (2 mM) catalyzed by [Cu2L55)]4+ (6 μM) on the pH in the mixed solvent of 75% methanol/glycerol (7/1) and 25% (v/v) Hepes buffer (50 mM) at 20±0.1 °C

Scheme 2
scheme 2

The catechol binding process to the dicopper(II) center

To characterize the catechol–[Cu2L55]4+ interaction in the first step, a methanolic solution of [Cu2L55]4+ was reacted anaerobically with excess DTBCH2 at −90 °C. The dominant UV/Vis ligand-to-metal charge transfer (LMCT) features of the fraction of bis(μ-hydroxo) species [19] present in solution (λmax 320 and 350 nm) were slowly replaced by weaker absorptions at 345 and 440 nm, attributable to LMCT bands of the catecholate complex (Fig. 5). The resulting spectrum is quite similar to that of the complex formed by [Cu2L55]4+ with the dianion of methyl 3,4-dihydroxybenzoate that, due to the electron-withdrawing substituent on the catechol ring, is stable and could be isolated and characterized previously [22]. In the present case, warming the solution of the [CuII2-DTBC] complex to room temperature results in quantitative catechol oxidation and CuII reduction. The multicomponent structure of the LMCT band suggests that catecholate binds as a bridging ligand to the dicopper(II) complex, as this is a common feature observed in the LMCT spectra of dicopper(II) complexes containing other bridging ligands like azide [19, 27] or even peroxide [28, 29]. As a matter of fact, chelated catecholate mononuclear CuII complexes exhibit a single LMCT band [18, 23]. The η21 bridging mode proposed for catecholate binding in Scheme 2 is based on the results of a recent spectroscopic investigation of nitrocatecholate-copper(II) complexes [30].

Fig. 5
figure 5

Electronic spectra recorded anaerobically in methanol solution at −90 °C of: (a) [Cu2(L55)]4+ (0.2 mM) and (b) its complex with DTBCH2 (1.8 mM). Spectrum (c) shows the stoichiometric formation of DTBQ after warming the solution to room temperature. When the solution was exposed to air, quantitative conversion of DTBCH2 into DTBQ occurred

The second phase of the reaction involves the global turnover, reactions IaIIc, where the rate-determining step is reaction IIc. Catechol binding to the CuI2 species in reaction IIa occurs as a pre-equilibrium before dioxygen binding, and accounts for the observed substrate saturation behavior. This pre-equilibrium is required because [Cu2L55]2+ does not form a stable dioxygen complex even at low temperature, due to the poor coordination of the ligand L55 tertiary amino groups to the CuI ions, as shown by the crystal structure of the complex [23]. Charge donation from bound catechol to the CuI centers partially compensates this effect, facilitating the formation of the dioxygen adduct. We have tried to characterize the binding of catechol to [Cu2L55(CH3CN)2]2+ by spectroscopic methods in anaerobic conditions. Catecholate complex formation does not bear useful near-UV features but can be followed by NMR (Fig. 6). Since at the concentrations required for the NMR experiment the presence of base causes precipitation of the complex in the presence of the catechol, an excess of free DTBCH2 had to be used to force the binding process. To reduce the effect of exchange phenomena, due to the kinetic lability of Cu(I), the spectra were recorded at low temperature. The proton signals of the benzimidazole, xylyl, and methylene groups of the ligand are slightly shifted and both the bound CH3CN molecules of the starting complex are clearly displaced upon formation of the catechol-[Cu2L55]2+ complex. Since the signals of the L55 residues bound to both CuI atoms are affected and the spectrum of the CuI2-catechol complex maintains the symmetric pattern present in that of the starting [Cu2L55(CH3CN)2]2+ complex, catechol appears to bind to both CuI atoms. Although we cannot discriminate the mode of binding of the catechol, we believe that in order to keep the two CuI ions in close proximity to favor the dioxygen binding step (reaction IIb), DTBC must act as a symmetric, likely η11 bridging, ligand in this case (Scheme 3).

Fig. 6
figure 6

NMR spectra of [Cu2L55(CH3CN)2]2+ (lower trace) and the complex in the presence of 10 molar equiv DTBCH2 (upper trace) in CD3OD at −40 °C under anaerobic conditions. Labeling refers to the structure of L55. Signals due to excess DTBCH2 are not shown. The chemical shift of methyl protons j at 3.77 ppm for [Cu2L55]2+ is little affected by DTBC binding. The spectra are basically unchanged, but decrease in resolution at still lower temperature

Scheme 3
scheme 3

The catechol binding process to the dicopper(I) center

Attempts to characterize the ternary complex of [Cu2L55]2+-O2-DTBC by UV/Vis or NMR techniques at low temperature were unsuccessful, even in the presence of large excess DTBCH2. Thus, the ternary complex does not accumulate in solution during turnover. The observed data could be explained considering a slow oxygenation of the complex [CuI2-DTBC], reaction IIb, followed by faster catechol oxidation. This hypothesis agrees with the kinetic behavior but can be discarded on the basis of the analysis of the activation parameters of the reaction (see below). A second possibility to explain the lack of accumulation of the ternary complex in the second step considers that oxygenation is faster than electron transfer, but that the equilibrium IIb is shifted to the left, so that only a minor fraction of the complex is present in solution as a ternary complex. This implies that quinone formation according to reaction IIc is preceded by two successive equilibria, that is coordination of DTBCH2 (reaction IIa) and then O2 (reaction IIb) to [Cu2L55]2+. The rate of the process then becomes (see Appendix):

$$ {\text{rate}} = \frac{{k_{{\text{IIc}}} \times K_{{\text{O2}}} \times {\text{[O}}_{\text{2}} {\text{]}} \times {\text{[DTBCH}}_{\text{2}} {\text{]}} \times {\text{[complex]}}}} {{K_{{\text{D2}}} + {\text{[DTBCH}}_{\text{2}} {\text{]}} + K_{{\text{O2}}} \times {\text{[O}}_{\text{2}} {\text{]}} \times {\text{[DTBCH}}_{\text{2}} {\text{]}}}} \approx \frac{{k_{\text{2}} \times {\text{[DTBCH}}_{\text{2}} {\text{]}} \times {\text{[complex]}}}} {{K_{{\text{D2}}} + {\text{[DTBCH}}_{\text{2}} {\text{]}}}} $$
(3)

where kIIc represents the rate constant for reaction IIc, KO2 the dioxygen affinity constant in the ternary complex, and k2=kIIc×KO2×[O2]. Equation 3 agrees with the observation that the rate of the second step depends both on substrate and dioxygen concentrations.

When the rate of the second step was analyzed as a function of the pH of the medium (Fig. 4), it was found that the maximum activity occurs above pH 7. The reaction rates drops dramatically at acidic pH. Probably this is due to protonation of the ligand, with concomitant release of copper(I), or protonation of the ternary complex of [Cu2L55]2+-O2-DTBC, with concomitant release of hydrogen peroxide; the last option is consistent with the observed catalyst degradation at long reaction times.

Fitting the rate data obtained at different temperatures with Eq. 2 for the first step, and Eq. 3 for the second step, allowed the determination of k1, KD1, k2 and KD2 (data at 20 °C in Table 1). According to the proposed mechanism, KD1 and KD2 refer to the catechol dissociation constant for the CuII2 and CuI2 forms of the complex, respectively. The temperature dependence of KD1−1 and KD2−1, through the van’t Hoff equation (Fig. 7), gives the ΔH° and ΔS° values for reactions Ia and IIa, respectively (Table 1). Surprisingly, ΔH° is less favorable for the first step, involving CuII2, with respect to the second step, involving CuI2. This is due to the fact that, in the first step, binding of the catechol to [Cu2L55(H2O)(OH)]3+ occurs with displacement of the hydroxo group (reaction Ia). On the other hand, this OH group keeps the CuII centers of the dinuclear complex in close proximity and this favors, from an entropy point of view (ΔS° is less negative), the coordination of DTBCH2 as a bridge between the two metal ions. The large negative ΔS° term in the case of [Cu2L55]2+ is probably related to the fact that the two ligand arms binding the CuI centers are largely free to move and catechol binding to both metal ions strongly reduces the mobility of the molecule. Overall, the binding is stronger in the first step than in the second step (KD1−1>KD2−1), due to the less unfavorable entropy term.

Table 1 Kinetic and thermodynamic parameters for the two steps of the catalytic catecholase reaction
Fig. 7
figure 7

van’t Hoff plots for the two steps of the catalytic oxidation of DTBCH2 by [Cu2L55)]4+ in a mixed solvent of 75% methanol/glycerol (7/1) and 25% Hepes buffer (50 mM) at pH 6.8: open squares, first step (KDi−1=KD1−1); solid circles, second step (KDi−1=KD2−1)

The activation parameters for reactions Ib, IIb and IIc were obtained through the temperature dependence of k1 and k2, respectively (Fig. 8, Table 1). For step 1, k1 measures the rate of electron transfer of reaction Ib, which is followed by a presumably fast quinone release. The ΔH value is moderately large, its magnitude being determined by the difference in redox potential of the two partners and by the overlap between the orbitals involved. Interestingly, ΔS has a negative value, in spite that reaction Ib occurs with an increase in the number of species.

Fig. 8
figure 8

Eyring plots for the two steps of the catalytic oxidation of DTBCH2 by [Cu2L55)]4+ in a mixed solvent of 75% methanol/glycerol (7/1) and 25% Hepes buffer (50 mM) at pH 6.8: solid circles, first step (k i =k1); open squares, second step (k i =k2/[O2])

For the analysis of the rate of the second step, we note that k2 is the combination of an electron transfer rate constant, a binding constant and [O2] (k2=kIIC×KO2×[O2]). Therefore, for the determination of the ΔH and ΔS parameters for the second step it was necessary to determine the value of [O2] at each temperature and apply the Eyring equation to k2/[O2] (Fig. 8). Since k2/[O2]=kIIC×KO2, the observed activation parameters are combinations of kinetic and thermodynamic parameters:

$$ \Delta {\text{H}}^{ \ne } = \Delta {\text{H}}^{ \ne } _{{{\text{ET}}}} + \Delta {\text{H $ ^\circ $ }}_{{{\text{oxyg}}}} {\text{ }} $$
(4)
$$ \Delta {\text{S}}^{ \ne } = \Delta {\text{S}}^{ \ne } _{{{\text{ET}}}} + \Delta {\text{S $ ^\circ $ }}_{{{\text{oxyg}}}} {\text{ }} $$
(5)

where ΔHET and ΔSET refer to the electron transfer process from bound catechol to the Cu2O2 core in reaction IIc, and ΔH°oxyg and ΔS°oxyg refer to the oxygenation equilibrium of reaction IIb. The range of ΔH°oxyg values for a variety of dinuclear CuI complexes related to [Cu2L55]2+ extends from −28 to −74 kJ mol−1 [31]. We expect that, in the present case, ΔH°oxyg will be less favorable, i.e. shifted toward the positive range, as [Cu2L55]2+ does not form a stable dioxygen complex. In any case, the ΔH value for the second step of the catalytic cycle is larger than that for the first step, making reactions IIb and IIc the rate-determining steps for the catechol oxidation. The large ΔH for reaction IIc is probably essentially determined by the cleavage of the O–O bond in the transition state, in agreement with theoretical calculations [4]. In addition, as opposed to the first step, ΔS is now large and positive, in spite of the fact that ΔS°oxyg is expected to be large and negative, with values ranging from −100 to −250 J K−1 mol−1 [31]. For reaction IIb, ΔS°oxyg probably approaches the more negative end of this range, due to sizeable steric crowding in the ternary complex of [Cu2L55]2+-O2-DTBC. As a consequence, ΔSETS−ΔS°oxyg must have a large positive value, probably because passing from the ternary complex to the transition state the partial cleavage of the O–O and Cu–O(catechol) bonds relieves the steric crowding.

As stated above, the lack of accumulation of the ternary complex in the second step could in principle be explained also by a slow oxygenation of the complex [CuI2-DTBC] (reaction IIb), followed by faster catechol oxidation. The rate equation describing this process would be analogous to Eq. 3, with k2=koxyg×[O2] (see Appendix), and in that case ΔS would refer to the oxygenation process of reaction IIb. However, the observation of a positive ΔS contrasts this possibility, since it would imply a negative activation entropy. In fact, oxygenation studies performed on dinuclear CuI complexes systematically show large negative values of ΔS, ranging from −66 to −167 J K−1 mol−1 [31].

In summary, the present investigation analyzes the catalytic catechol oxidase activity by the biomimetic dinuclear copper complex derived from the ligand L55. The elementary steps of the catalytic cycle have been separated, allowing, for the first time, the determination of the kinetic and thermodynamic parameters controlling the two phases of the process. The data obtained on this model system can be useful for the interpretation of the intimate mechanisms of the catecholase reactions promoted by type 3 copper enzymes such as catechol oxidase and tyrosinase, which are still completely unclear.