Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Marcus-type driving force correlations reveal the mechanism of proton-coupled electron transfer for phenols and [Ru(bpy)3]3+ in water at low pH

Janne Soetbeer , Prateek Dongare * and Leif Hammarström *
Department of Chemistry – Ångström Laboratory, Uppsala University, Box 523, SE-751 20, Uppsala, Sweden. E-mail: leif.hammarstrom@kemi.uu.se; prateek.dongare@kemi.uu.se

Received 7th February 2016 , Accepted 1st April 2016

First published on 1st April 2016


Abstract

Proton-coupled electron transfer (PCET) from tyrosine and other phenol derivatives in water is an important elementary reaction in chemistry and biology. We examined PCET between a series of phenol derivatives and photogenerated [Ru(bpy)3]3+ in low pH (≤4) water using the laser flash-quench technique. From an analysis of the kinetic data using a Marcus-type free energy relationship, we propose that our model system follows a stepwise electron transfer-proton transfer (ETPT) pathway with a pH independent rate constant at low pH in water. This is in contrast to the concerted or proton-first (PTET) mechanisms that often dominate at higher pH and/or with buffers as primary proton acceptors. The stepwise mechanism remains competitive despite a significant change in the pKa and redox potential of the phenols which leads to a span of rate constants from 1 × 105 to 2 × 109 M−1 s−1. These results support our previous studies which revealed separate mechanistic regions for PCET reactions and also assigned phenol oxidation by [Ru(bpy)3]3+ at low pH to a stepwise PCET mechanism.


Introduction

Proton-coupled electron transfer (PCET) reactions are ubiquitous in chemistry and biology. PCET is a term given to a wide array of reactions where proton transfer accompanies electron transfer.1–5 Various mechanistic regimes such as sequential electron–proton transfer (ETPT or PTET), concerted electron–proton transfer (CEP) and hydrogen atom transfer (HAT) reactions fall under the umbrella of PCET reactions.5,6 In many cases the step-wise and concerted mechanisms compete with each other, as has been found in model compounds as well as complex systems (polypeptides, enzymes) and therefore the ability to distinguish between the mechanisms is of prime importance. However, the intermediate species of the step-wise reactions are usually very short-lived and cannot be directly detected. In the majority of cases, thermodynamic-kinetic arguments are used to judge whether a step-wise mechanism is possible, i.e. if observed rate constants or activation energies are consistent with expected differences in E0 or pKa between the reagents. Also, high H/D kinetic isotope effects (KIEs) are used to support assignment to a CEP reaction, but the value is often moderate, KIE ≈ 2.7–13 More rarely, rate constants vs. driving force correlations are used to support a mechanistic assignment. In one example, direct measurements of all the limiting ET, PT and CEP rate constants were made for the PCET oxidation of [W(Cp)(CO)3H] to conclude that CEP is the operating mechanism when both the oxidants and the bases are weak.14

Tyrosine and other phenol derivatives constitute the most studied class of compounds for PCET reactions, in synthetic systems as well as in proteins. The role of H2O in PCET reactions of phenolic compounds is topical with reference to natural photosynthetic membranes and other protein systems. Therefore the mechanism of PCET in water at different pH ranges remains an active area of intense investigation and debate.10,12,15–29,48 However, in aqueous systems it is often even more difficult to distinguish between stepwise and concerted reactions. Differences in reactant pKa and E0 values are often less extreme than in organic solvents, and observed H/D KIEs may be affected by solvent KIEs of the H2O/D2O exchange.

A prototypical PCET reaction is the oxidation of phenol by [Ru(bpy)3]3+ derivatives with water as the primary proton acceptor. In neutral and alkaline solution the mechanism has been assigned to a CEP reaction, or a PTET reaction involving OH.10,15,17,30 In some literature reports the pH-independent PCET mechanism at low pH (2–4), when OH was not the primary acceptor, was assigned to a CEP reaction based on the kinetic isotope effect KIE = 2.15,17 Irebo et al.10 instead assigned it to an ETPT reaction, as a pathway with different kinetic characteristics (e.g. KIE = 3.0) that dominated at neutral pH was already assigned to a CEP reaction; however, they did not exclude a CEP mechanism also at low pH. It should be noted that neither of these two mechanisms could be excluded based on thermodynamic limits, as discussed above for organic solvents. Also, while KIEs around 2 are often taken as evidence for a CEP reaction, even pure ET reactions may show a significant solvent KIE when H2O and D2O are compared.10,31–33

In order to investigate the PCET mechanism for this important case study reaction at low pH, we employed a series of phenols with varying redox potentials and pKa values (Chart 1 and Table S3), which were oxidized by laser-flash generated [Ru(bpy)3]3+. We investigated the correlation of experimental rates of phenol oxidation and the ones predicted for the CEP and stepwise mechanisms. Similar correlations were previously established by Mayer et al.34 to test PCET mechanisms in phenolic systems with organic bases.3,35–37 Reports have appeared where oxidation of phenol derivatives in organic solvents by [Ru(2,2′-bipyrazine)3]2+ has been studied.38,39 However, this is the first example of this method for deducing a PCET mechanism where water is acting as the proton acceptor.


image file: c6sc00597g-c1.tif
Chart 1 Structural formulae of [Ru(bpy)3]2+ and phenol derivatives. Phenols 8–12 were included only in the electrochemical analysis (see ESI).

Results and discussion

The PCET reaction was studied using the ‘flash-quench’ method. The reaction was initiated by irradiation of [Ru(bpy)3]2+ at 460 nm using a 10 ns laser flash in the presence of an electron acceptor, [Co(NH3)5Cl]Cl2 or methylviologen (MV2+), which resulted in generation of a photooxidized [Ru(bpy)3]3+ species on a 100 ns time scale. The [Ru(bpy)3]3+ then oxidized the phenol on a much longer time scale, leading to a recovery of the [Ru(bpy)3]2+ bleach at around 450 nm and phenoxyl radical absorption at around 410 nm (Fig. 1).
image file: c6sc00597g-f1.tif
Fig. 1 Representative kinetic traces at 450 and 410 nm for the oxidation of (a) phenol 1 (4-MeO) and (b) phenol 7 (2,6-F). The remaining traces are shown in the ESI (Fig. S1).

The resulting transient absorption trace was subjected to a single exponential fit to extract the pseudo-first order rate constants for oxidation of phenol (kobs). To nullify the effect of recombination between MV+ and [Ru(bpy)3]3+, [Co(NH3)5Cl]3+ was used as a sacrificial acceptor instead of MV2+, which ensured that even the slowest PCET rates could be observed. To avoid interference from irreversible phenol degradation, phenol was added in great excess of [Ru(bpy)3]2+ and fresh, deoxygenated solutions which were protected from ambient light were used. Separate experiments were performed for phenols with faster rates (i.e.1–3) using the reversible acceptor MV2+ to ensure that the PCET rates are independent of the choice of external oxidant. At 450 nm and pH 2 the rate constants of 1.94 × 109 M−1 s−1 and 3.69 × 105 M−1 s−1 were obtained for phenols 1 and 3, respectively. This rules out any interference from by-products of the cobalt complex as rate constants of 1.9 × 109 M−1 s−1 and 3.8 × 105 M−1 s−1, respectively, were obtained when [Co(NH3)5Cl]Cl2 was used as an external electron acceptor.

The rate of phenol oxidation depends upon the pH of the medium due to the presence of a pH dependent, very reactive phenolate ion (PhO) species. However, in the low pH range of 0–2 (pH = 0 in the case of 3F and 2,6-F) the rate of phenol oxidation is entirely due to the protonated (PhOH) species, as seen by a pH-independent reaction rate. Phenols 8–12 did not show pH independent rates even at pH = 0 and were excluded from the correlation in Fig. 2. The PCET mechanistic studies were performed in neat water as a solvent to ensure that water was the only proton acceptor in the system. The pH was adjusted using HCl or H2SO4 and NaOH. At this low pH the buffer capacity of water is sufficient for convenient pH stability.


image file: c6sc00597g-f2.tif
Fig. 2 Correlation plots for measured against predicted values of the second order PCET rate constant ln(krel), relative to that for phenol 3, for (a) reversible ETPT (r2 = 0.91) (b) irreversible ETPT (r2 = 0.86) and (c) CEP (r2 = 0.71) reaction mechanisms (r2 values relate to the solid line). The broken line shows a linear fit to the data points and the solid line is drawn along the diagonal representing an ideal correlation of slope = 1. The numbering of phenols follows Chart 1. The vertical error bars represent the 95% confidence interval (Table S1). The horizontal error bars reflect the variation in literature values for E0red and pKa (Table S3).

The mechanism can be deduced by analysis and comparison of the observed variation in the PCET rate constant for the series of phenols with that predicted for the respective mechanism of ETPT and CEP (Fig. 2). As is customary for ET and PCET reactions with an homologous series of reactants, we assume that the stability constant of the precursor and successor complexes is invariant,34,40 and that the variation in reorganization energy (λ) between the phenols is small; the aqueous λ value of PhOH–PhOH+˙ for 1 and 3 has indeed been reported to be equal within 5%.41 In the following paragraphs we derive the predicted variations before we comment on the results of Fig. 2.

At pH = 0–2, phenol deprotonation will occur by water (H2O) and not by OH (eqn (1)). Because of the large ΔpKa between the conjugate acid H3O+ (pKa = 0) and phenol (pKa = 10), PTET where water is a proton acceptor is too slow to be consistent with the observed rate constant (kPT = 1011−pKa = 10 s−1).42–44 Moreover, the phenols with lower pKa values give slower kobs, in contradiction with a PTET mechanism. Therefore, the PTET mechanism can be ruled out of this discussion.

 
image file: c6sc00597g-u1.tif(1)

In the case of ETPT (eqn (2)), the rate constant is not pH dependent, as pKa < −1 for all the PhOH+˙ studied and deprotonation to H2O dominates, but rather it depends upon the substituents on the phenol. This mechanism can be studied in terms of two separate stepwise mechanistic regimes, namely, irreversible ETPT and reversible ETPT.

 
image file: c6sc00597g-u2.tif(2)

Reversible ETPT

In the case where reversible ET is followed by PT (k−ETkPT), the overall rate constants can be derived from pre-equilibrium kinetics, which yields; kobs = kET/k−ET·kPT. Here kET/k−ET is decreased by a factor of 10 for each 59 meV increase in ΔG0ET. Deprotonation of an Eigen acid in water follows kPT ≈ 1011−pKa s−1.8 This results in the following rate constant expression for reversible ETPT (pKa refers here to the PhOH+˙ species):
 
image file: c6sc00597g-t1.tif(3a)
 
image file: c6sc00597g-t2.tif(3b)
where kETPTrev(rel.) = kETPTrev/kETPTrev (phenol 3), ΔΔG and ΔpKa is the value relative to phenol 3, A is a pre-exponential factor and the pKa values can be obtained from the literature (Table S2). Eqn (4) was employed to obtain the driving force for the ET step, ΔG0ET with an assumption that the coulombic interaction between the involved species is negligible.
 
image file: c6sc00597g-t3.tif(4)
here, z is the number of electrons transferred, E0RuIII/II = 1.26 V vs. NHE45 and image file: c6sc00597g-t4.tif is the pH-independent potential of the substituted phenol obtained from the literature or experimentally determined in this work (Table S3).

Irreversible ETPT

Following eqn (2), the oxidation of phenol leads to a significant drop in its pKa to around −2. As a result the PhOH+˙ species is rapidly deprotonated which could make electron transfer the rate limiting step (k−ETkPT in addition to kETkPT). This results in kobs = kET. The rate of irreversible ETPT can thus be determined by a standard Marcus-type rate expression (eqn. (5)). The value relative to that for phenol 3 was calculated using eqn (5b);46 this assumes that |ΔG0CEP| ≪ λCEP, which is reasonable as ΔG0ET varies from −0.34 to 0.00 eV in the series of phenols 1–7.
 
image file: c6sc00597g-t5.tif(5a)
 
image file: c6sc00597g-t6.tif(5b)
Here A is the pre-exponential factor and λET is the reorganizational energy for the electron transfer step, which is assumed to be constant for the series of phenols. Eqn (5b) gives the predicted dependence of ln[thin space (1/6-em)]kETPTirr (=ln[thin space (1/6-em)]kET) on the driving force for the series of phenols.

Concerted mechanism (CEP)

In this mechanism, electron and proton transfer occurs in a single kinetic step (eqn (6)).
 
image file: c6sc00597g-u3.tif(6)

The driving force for CEP with water as an acceptor is equal to the sum of the driving forces for oxidation to PhOH+˙ (eqn (4)) and its subsequent deprotonation (upper path in Scheme 1):

 
image file: c6sc00597g-t7.tif(7)
with pKa(H3O(aq)+) = 0. Analogous to ETPTirr, the rate constant for CEP can be determined by using a Marcus-type rate expression (eqn (8)), assuming |ΔG0CEP| ≪ λCEP in eqn (8b).46,47 We note that a value of λCEP = 0.45 eV for the reaction between [Ru(bpy)3]3+ and 3 in water has been suggested;8 however, the use of that value to predict relative values of kCEP resulted in a poor correlation with experimental data, with a curvature that clearly suggests that the value of λCEP should be significantly larger than 0.45 eV, see Fig. S4.
 
image file: c6sc00597g-t8.tif(8a)
 
image file: c6sc00597g-t9.tif(8b)


image file: c6sc00597g-s1.tif
Scheme 1 The mechanism of phenol oxidation via ETPT, PTET and CEP pathways.

Fig. 2 shows the correlation of experimental vs. calculated rate constants relative to the value for the case of unsubstituted PhOH; hence PhOH is distinctly represented in the centre of all plots. The relative rate constants for the three mechanisms were calculated according to eqn (3a), (3b), (5a), (5b), (8a) and (8b), using reduction potentials and pKa values of phenol derivatives (Table S3). The horizontal error bars on the data points are associated with the uncertainty in the literature values for E0 and pKa of the studied phenols.48–50 The operative mechanism under the present conditions should be revealed by a good correlation between experimental and predicted values. We point out that the plots for irreversible ET and CEP (Fig. 2b and c, respectively) are identical to their respective Marcus plots of ln[thin space (1/6-em)]k vs. ΔG0, with a linear transformation of the x-axis according to eqn (5b) and (8b), respectively (1 ln-unit = 50 meV driving force), and the solid line shows the slope of 1/(50 meV) predicted by eqn (5b) and (8b).

A simple inspection of the three correlation plots (Fig. 2) indicates that the best agreement between experimental and predicted relative rate constants is in the case of ETPT, while for CEP the correlation is comparatively poor. More precisely, a linear regression analysis reveals that in the case of reversible ETPT (Fig. 2a) the data points obey linearity with respect to the unsubstituted phenol (point 3) with a slope of ≈1. This is illustrated by the nearly perfect overlap between the linear fit (dashed line) and ideal correlation (solid line). The irreversible ETPT mechanism (Fig. 2b) shows the second best correlation with a slope of 1.2 and some deviation of the data points from the reference diagonal line. On the other hand, the slope of 1.6 for the CEP mechanism (Fig. 2c) shows a strong deviation from the typical Marcusian rate dependence on the CEP driving force. Phenols with both the highest and lowest observed rate constants deviate substantially from the predicted values in this case.

From the above comparison of the correlations in Fig. 2, we draw the conclusion that the PCET reaction for this series of phenols with [Ru(bpy)3]3+ in water follows a step-wise, ETPT mechanism. Regarding which of the two kinetic limits is most likely, we note that reversible ETPT (Fig. 2a) shows a slightly better agreement with predictions than the irreversible ETPT (Fig. 2b). On the other hand, this would require that the reverse ET in the solvent cage is faster than deprotonation of the PhOH+˙ intermediate, which occurs with τ = 0.1–1 ps (kPT ≈ 1011−pKa s−1).8 While this cannot be excluded, it is questionable that this would hold for the entire series of phenols with different potentials and pKa values. It is of course possible that the mechanism gradually changes between ETPTrev and ETPTirrev within the series. In the absence of clearer proof, we leave it as an open question whether the step-wise ETPT mechanism follows an irreversible or a reversible pathway.

Conclusions

A central conclusion that can be drawn at this juncture is that the oxidation of the series of phenols by [Ru(bpy)3]3+ shows that a stepwise ETPT mechanism is most likely to occur at low pH with water as a proton acceptor. This conclusion is based on rate correlations of phenols yielding rate constants that vary from 1 × 105 to 2 × 109 M−1 s−1. The results are in line with our previous assignment for intramolecular PCET in a Ru–Tyrosine complex, where at low pH a stepwise ETPT mechanism was proposed when water is the proton acceptor.10 In contrast the Marcus-type analysis suggests that the operating mechanism under the conditions investigated is not a concerted reaction (CEP), as was suggested before, and as a consequence the reorganization energy value of 0.45 eV reported is not correct.8 This impacts our understanding of the competition between concerted and step-wise PCET mechanisms of tyrosine and other phenols in water. The mechanistic investigation using a Marcus-type relationship offers a successful tool to discern among these PCET mechanisms. This study provides insight from a model system into the mechanisms of PCET for the working of various biological processes where phenoxyl radicals play a pivotal role.

Acknowledgements

J. S. would like to thank Dr Starla D. Glover for help with electrochemistry measurements. This work was supported by the Swedish Research Council, the Swedish Energy Agency, and the Knut and Alice Wallenberg Foundation.

References

  1. J. L. Dempsey, J. R. Winkler and H. B. Gray, Chem. Rev., 2010, 110, 7024–7039 CrossRef CAS PubMed.
  2. S. Hammes-Schiffer and A. A. Stuchebrukhov, Chem. Rev., 2010, 110, 6939–6960 CrossRef CAS PubMed.
  3. J. J. Warren, T. A. Tronic and J. M. Mayer, Chem. Rev., 2010, 110, 6961–7001 CrossRef CAS PubMed.
  4. J. Stubbe, D. G. Nocera, C. S. Yee and M. C. Y. Chang, Chem. Rev., 2003, 103, 2167–2202 CrossRef CAS PubMed.
  5. D. R. Weinberg, C. J. Gagliardi, J. F. Hull, C. F. Murphy, C. A. Kent, B. C. Westlake, A. Paul, D. H. Ess, D. G. McCafferty and T. J. Meyer, Chem. Rev., 2012, 112, 4016–4093 CrossRef CAS PubMed.
  6. C. J. Gagliardi, B. C. Westlake, C. A. Kent, J. J. Paul, J. M. Papanikolas and T. J. Meyer, Coord. Chem. Rev., 2010, 254, 2459–2471 CrossRef CAS.
  7. L. Hammarstrom and S. Styring, Energy Environ. Sci., 2011, 4, 2379–2388 CAS.
  8. J. Bonin, C. Costentin, C. Louault, M. Robert and J.-M. Savéant, J. Am. Chem. Soc., 2011, 133, 6668–6674 CrossRef CAS PubMed.
  9. C. J. Fecenko, T. J. Meyer and H. H. Thorp, J. Am. Chem. Soc., 2006, 128, 11020–11021 CrossRef CAS PubMed.
  10. T. Irebo, M.-T. Zhang, T. F. Markle, A. M. Scott and L. Hammarström, J. Am. Chem. Soc., 2012, 134, 16247–16254 CrossRef CAS PubMed.
  11. D. J. Stewart, M. K. Brennaman, S. E. Bettis, L. Wang, R. A. Binstead, J. M. Papanikolas and T. J. Meyer, J. Phys. Chem. Lett., 2011, 2, 1844–1848 CrossRef CAS.
  12. C. Costentin, M. Robert and J.-M. Saveant, Phys. Chem. Chem. Phys., 2010, 12, 11179–11190 RSC.
  13. R. Balasubramanian, G. Blondin, J. C. Canales, C. Costentin, J.-M. Latour, M. Robert and J.-M. Savéant, J. Am. Chem. Soc., 2012, 134, 1906–1909 CrossRef CAS PubMed.
  14. M. Bourrez, R. Steinmetz, S. Ott, F. Gloaguen and L. Hammarström, Nat. Chem., 2015, 7, 140–145 CrossRef CAS PubMed.
  15. J. Bonin, C. Costentin, C. Louault, M. Robert, M. Routier and J.-M. Savéant, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 3367–3372 CrossRef CAS PubMed.
  16. J. Bonin, C. Costentin, M. Robert, M. Routier and J.-M. Savéant, J. Am. Chem. Soc., 2013, 135, 14359–14366 CrossRef CAS PubMed.
  17. C. Costentin, V. Hajj, C. Louault, M. Robert and J.-M. Savéant, J. Am. Chem. Soc., 2011, 133, 19160–19167 CrossRef CAS PubMed.
  18. C. Costentin, C. Louault, M. Robert and J.-M. Savéant, J. Am. Chem. Soc., 2008, 130, 15817–15819 CrossRef CAS PubMed.
  19. M.-T. Zhang and L. Hammarström, J. Am. Chem. Soc., 2011, 133, 8806–8809 CrossRef CAS PubMed.
  20. M.-T. Zhang, T. Irebo, O. Johansson and L. Hammarström, J. Am. Chem. Soc., 2011, 133, 13224–13227 CrossRef CAS PubMed.
  21. M.-T. Zhang, J. Nilsson and L. Hammarstrom, Energy Environ. Sci., 2012, 5, 7732–7736 CAS.
  22. C. J. Fecenko, H. H. Thorp and T. J. Meyer, J. Am. Chem. Soc., 2007, 129, 15098–15099 CrossRef CAS PubMed.
  23. M. R. Seyedsayamdost, C. S. Yee, S. Y. Reece, D. G. Nocera and J. Stubbe, J. Am. Chem. Soc., 2006, 128, 1562–1568 CrossRef CAS PubMed.
  24. H. Ishikita, A. V. Soudackov and S. Hammes-Schiffer, J. Am. Chem. Soc., 2007, 129, 11146–11152 CrossRef CAS PubMed.
  25. C. Costentin, M. Robert, J.-M. Savéant and A.-L. Teillout, ChemPhysChem, 2009, 10, 191–198 CrossRef CAS PubMed.
  26. N. Song and D. M. Stanbury, Inorg. Chem., 2008, 47, 11458–11460 CrossRef CAS PubMed.
  27. N. Song and D. M. Stanbury, Inorg. Chem., 2011, 50, 12762–12773 CrossRef CAS PubMed.
  28. N. Song and D. M. Stanbury, Inorg. Chem., 2012, 51, 4909–4911 CrossRef CAS PubMed.
  29. P. Dongare, S. Maji and L. Hammarström, J. Am. Chem. Soc., 2016, 138, 2194–2199 CrossRef CAS PubMed.
  30. C. Costentin, C. Louault, M. Robert and J.-M. Savéant, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 18143–18148 CrossRef CAS PubMed.
  31. L. R. Khundkar, J. W. Perry, J. E. Hanson and P. B. Dervan, J. Am. Chem. Soc., 1994, 116, 9700–9709 CrossRef CAS.
  32. L. P. McMahon, W. J. Colucci, M. L. McLaughlin and M. D. Barkley, J. Am. Chem. Soc., 1992, 114, 8442–8448 CrossRef CAS.
  33. A. Masuda and Y. Kaizu, Inorg. Chem., 1998, 37, 3371–3375 CrossRef CAS.
  34. J. M. Mayer, J. Phys. Chem. Lett., 2011, 2, 1481–1489 CrossRef CAS PubMed.
  35. T. F. Markle, I. J. Rhile, A. G. DiPasquale and J. M. Mayer, Proc. Natl. Acad. Sci. U. S. A., 2008, 105, 8185–8190 CrossRef CAS PubMed.
  36. J. P. Roth, J. C. Yoder, T.-J. Won and J. M. Mayer, Science, 2001, 294, 2524–2526 CrossRef CAS PubMed.
  37. J. N. Schrauben, M. Cattaneo, T. C. Day, A. L. Tenderholt and J. M. Mayer, J. Am. Chem. Soc., 2012, 134, 16635–16645 CrossRef CAS PubMed.
  38. J. J. Concepcion, M. K. Brennaman, J. R. Deyton, N. V. Lebedeva, M. D. E. Forbes, J. M. Papanikolas and T. J. Meyer, J. Am. Chem. Soc., 2007, 129, 6968–6969 CrossRef CAS PubMed.
  39. O. S. Wenger, Acc. Chem. Res., 2013, 46, 1517–1526 CrossRef CAS PubMed.
  40. N. Sutin, in Prog. Inorg. Chem., John Wiley & Sons, Inc., 2007,  DOI:10.1002/9780470166314, ch. 9, pp. 441–498.
  41. J. J. Guerard, P. R. Tentscher, M. Seijo and J. Samuel Arey, Phys. Chem. Chem. Phys., 2015, 17, 14811–14826 RSC.
  42. L. Hammarström and S. Styring, Philos. Trans. R. Soc., B, 2008, 363, 1283–1291 CrossRef PubMed.
  43. The deprotonation rate constant kPT is given by kPT = k−PTKa, where Ka is the acid dissociation constant (=1 × 10−10 M for phenol) and kPT is the second-order rate constant for protonation of the corresponding base, which is not larger than diffusion controlled, 1 × 1011 M−1 s−1 for proton in water. This puts the limit for deprotonation at k−PT ≤ 10 s−1.
  44. M. Gutman, in Methods Biochem. Anal., John Wiley & Sons, Inc., 2006, ch. 1, pp. 1–103 DOI:10.1002/9780470110515.
  45. A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser and A. von Zelewsky, Coord. Chem. Rev., 1988, 84, 85–277 CrossRef CAS.
  46. R. A. Marcus and N. Sutin, Biochim. Biophys. Acta, 1985, 811, 265–322 CrossRef CAS.
  47. S. Hammes-Schiffer, Acc. Chem. Res., 2001, 34, 273–281 CrossRef CAS PubMed.
  48. A. Harriman, J. Phys. Chem., 1987, 91, 6102–6104 CrossRef CAS.
  49. C. Li and M. Z. Hoffman, J. Phys. Chem. B, 1999, 103, 6653–6656 CrossRef CAS.
  50. J. Lind, X. Shen, T. E. Eriksen and G. Merenyi, J. Am. Chem. Soc., 1990, 112, 479–482 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Experimental details, phenol reduction potentials and pKa values, and example kinetic and electrochemical data. See DOI: 10.1039/c6sc00597g

This journal is © The Royal Society of Chemistry 2016