1: %\documentclass[preprint,prb,aps,nobibnotes]{revtex4}%endfloats
2: \documentclass[prb,aps,nobibnotes,twocolumn]{revtex4}%endfloats
3: \usepackage{graphicx}%
4: \usepackage{dcolumn}
5: \usepackage{amsmath}
6: \usepackage{bm}
7: \usepackage{dcolumn}
8: \usepackage{longtable}
9: \usepackage{pifont,color}
10: %\usepackage[dvips]{color}
11: \voffset=1.5cm
12: %\hoffset=0.8cm
13:
14: \bibliographystyle{achemsol}
15:
16: \begin{document}
17:
18: \title{\centering\Large\bf What are the benefits of bound (protonation) states
19: for the electron-transfer kinetics? }
20: \author{Dmitry V.\ Matyushov}
21: \affiliation{Center for Biological Physics,
22: Arizona State University, PO Box 871604, Tempe, AZ 85287-1604}
23: \date{\today}
24: \begin{abstract}
25: We describe a model of electron transfer reactions affected by local
26: binding to the donor or acceptor sites of a particle in equilibrium
27: with the solution. The statistics of fluctuations of the
28: donor-acceptor energy gap caused by binding/unbinding events are
29: non-Gaussian, and the resulting free energy surfaces of electron
30: transfer are non-parabolic. The band-width of the charge-transfer
31: optical transition is predicted to pass through a maximum as a
32: function of the concentration of binding particles in the solution.
33: The model is used to rationalize recent observations of
34: pH-dependence of electron transfer rates involving changes in the
35: protonation state of the donor-acceptor complex.
36: \end{abstract}
37: \preprint{Submitted to J.\ Phys.\ Chem.\ B}
38: \maketitle
39:
40: \section{Introduction}
41: \label{sec:1}
42: Many redox reaction in chemistry and biology involve bound states
43: which are weaker than common chemical bonds, but stronger than
44: intermolecular interactions in bulk molecular liquids. A prominent
45: example of such association is binding of water molecules to solutes
46: via hydrogen bonds. Since the strength of such bonds typically varies
47: between different electronic states of the solute, water association
48: can be recorded by optical
49: solvatochromism.\cite{Reichardt:94,DMjpcb:97,SolvPolarity:04,Agmon:05}
50: Another example is the one of proton equilibria involved in biological
51: electron transfer (ET) reactions responsible for photosynthesis and
52: respiration.\cite{McEvoy:06} In Photosystem II, the primary donor
53: P$_{680}^+$ is oxidized by tyrosine which changes its pK$_a$-value
54: from 10 to $-2$ upon oxidation. As a result, it cannot hold onto its
55: phenolic proton in aqueous solution releasing it in what is argued to
56: be a concerted electron-proton transfer mechanism.\cite{Cukier:04}
57: Another related example is photoacidity when photoexcited
58: intramolecular charge transfer lowers pK$_a$ for the release of a
59: proton to the solvent.\cite{Agmon:05} In view of the wide spread of
60: such reactions, often referred to as proton-coupled
61: ET,\cite{Cukier:96,Soudackov:00,Hammes-Schiffer:01,Cukier:04} in
62: biological systems in general and enzymetic reactions in
63: particular,\cite{Ferguson-Miller:96,Kirby:97} one wonders if their
64: occurrence is just a coincidence caused by ubiquity of acid-base
65: equilibria in proteins, or, alternatively, protonation/deprotonation
66: transitions are used by nature to fine-tune the ET energetics. This
67: question, also used for the paper title, is the subject of this study.
68:
69: It has been argued that coupling of ET to the change in protonation
70: alters the reaction free energy, often turning uphill reactions along
71: the ET coordinate into downhill reactions along a combined
72: electron-proton reaction path.\cite{Mayer:04} In order to make this
73: argument more qualitative, at least one more parameter, the
74: reorganization energy, needs to be
75: specified.\cite{Marcus:93,Borgis:96} The picture of crossing parabolas
76: used to calculate barriers of charge-transfer
77: reactions\cite{Marcus:65} involves in fact three components: the
78: reaction free energy $\Delta G$ representing the vertical shift of the
79: minima, the curvature $(2\lambda)^{-1}$ given in terms of the reorganization
80: energy $\lambda$, and the horizontal shift of the parabolas' minima
81: responsible for the Stokes shift, $\Delta X=X_{01}-X_{02}$ (Figure
82: \ref{fig:1}).
83:
84: \begin{figure}
85: \centering
86: \includegraphics*[width=6.5cm]{fig1}
87: \caption{Diagram of the free energy surfaces of ET obtained from
88: three parameters, the reaction free energy $\Delta G$, the free energy
89: curvature $(2\lambda)^{-1}$, and the Stokes shift (separation between
90: the minima $\Delta X$). The reaction coordinate $X$ is the energy
91: separation between the donor and acceptor electronic states. The
92: activated state is given by the resonance of the donor and
93: acceptor levels, $X=0$, while $X_{01}$ and $X_{02}$ indicate the
94: positions of the minima. }
95: \label{fig:1}
96: \end{figure}
97:
98: The reaction coordinate $X$, used to monitor the progress of a
99: charge-transfer reaction, is the instantaneous energy gap between the
100: donor and acceptor electronic energy levels.\cite{King:90} The
101: vertical axis in Figure \ref{fig:1} refers to \textit{free energies}
102: $G_i(X)$ as functions of \textit{energy} $X$. They cross at zero
103: energy gap $X=0$ allowing electronic tunneling ($i=1,2$ refer to
104: reactants and products, respectively). In fact, the Marcus-Hush
105: description of ET\cite{Marcus:89} imposes one more
106: additional constraint requiring the Stokes shift to be identically
107: equal to $2\lambda$,
108: %
109: \begin{equation}
110: \label{eq:0}
111: \Delta X = 2 \lambda = \beta \langle (\delta X)^2 \rangle
112: \end{equation}
113: where $\beta = 1/(k_{\text{B}}T)$. Therefore, only two parameters, $\Delta G$
114: and $\lambda$, are required to determine the activation barrier. The strong
115: link between the Stokes shift and the reorganization energy is a
116: consequence of the Gaussian statistics of $X$. Once the statistics
117: become non-Gaussian, the free energy surfaces $G_i(X)$ are
118: non-parabolic and eq \ref{eq:0} does not hold.\cite{DMacc:07} Here, we
119: propose a simple model for a charge-transfer reactions triggering the
120: change in the state of binding of some particle dissolved in the
121: solvent. In this development, we are less concerned with the details
122: of a particular binding mechanisms, but more interested in addressing
123: a more general question of whether the involvement of bound states can
124: potentially provide a new mechanism of tuning the energetics of ET not
125: incorporated in the picture of crossing parabolas. We find that the
126: statistics of fluctuations of the donor-acceptor energy gap, induced
127: by local binding/unbinding events, are non-Gaussian thus violating the
128: link between the Stokes shift and reorganization energy given by eq
129: \ref{eq:0}. The main consequence of this result is more flexibility,
130: compared to the standard picture, in tuning the energetics of ET.
131:
132: \section{Model}
133: \label{sec:2}
134: We will consider a somewhat simplified situation in which the bound
135: state is fully thermalized on the time-scale of ET i.e.\ on the time
136: $\tau_{\text{ET}}=k_{\text{ET}}^{-1}$ required for the system to climb
137: the top of the potential barrier from the equilibrium bottom of the
138: free energy surface ($k_{\text{ET}}$ is the ET rate constant). This
139: assumption implies that populations of bound and unbound states follow
140: Boltzmann statistics along the ET reaction coordinate, a situation
141: similar to the treatment of intramolecular vibrations in Sumi-Marcus
142: model.\cite{Sumi:86} This condition can be achieved when the rate of
143: release of the binding particle B, $k_b$, is much higher that the rate
144: of ET
145: %
146: \begin{equation}
147: \label{eq:1}
148: k_{\text{ET}} \ll k_b
149: \end{equation}
150:
151: When this requirement is satisfied, one can use statistical mechanics
152: to construct a one-dimensional free-energy surface for the ET reaction
153: in which the exchange between localized and delocalized states of the
154: binding particle produce fluctuations of the donor-acceptor energy
155: gap, just as any other solvent mode interacting with the transferred
156: electron.\cite{Stuchebrukhov:03}
157:
158:
159: \begin{figure}
160: \centering
161: \includegraphics*[width=7cm]{fig2}
162: \caption{Diagram of particle B exchanging between the bound and
163: dissolved states with the rate constant $k_b$. The equilibrium
164: population of the bound state assumes that the unbinding rate is
165: much faster than the rate of ET ($k_{\text{ET}}\ll k_b$); $\Delta
166: \epsilon=\epsilon_2-\epsilon_1$ denotes the change in the binding energy caused by the
167: electronic transition. }
168: \label{fig:2}
169: \end{figure}
170:
171: The release of one particle from the bound localized state requires
172: the Gibbs energy $g_i$ balancing the binding energy $\epsilon_i$ with an
173: entropy gain from moving the particle into the bulk. The binding
174: energy $\epsilon_i$ is generally different in the initial, D--A ($i=1$), and
175: final, D$^+$--A$^-$ ($i=2$), configurations of the donor-acceptor
176: complex (Fig.\ \ref{fig:2}). The difference in free energies needed to
177: release particle B to the solution is equal to the difference in
178: binding energies, $\Delta g = \Delta\epsilon = \epsilon_2-\epsilon_1$. Therefore, this quantity can
179: be estimated from the difference of the equilibrium dissociation
180: constants pK$_a^{(i)}$:
181: %
182: \begin{equation}
183: \label{eq:2}
184: \Delta \epsilon = k_{\text{B}}T \ln 10 \Delta \mathrm{pK}_a
185: \end{equation}
186:
187: The requirement of non-adiabaticity of ET, which we implicitly assume
188: here, imposes an additional constraint on the rate of bound particle
189: release. The change of the donor and acceptor energy levels caused by
190: the motion of the bound particle should not break the Landau-Zener
191: non-adiabaticity condition close to donor-acceptor energy resonance:
192: %
193: \begin{equation}
194: \label{eq:3}
195: \frac{2\pi}{\hbar} |V_{DA}|^2/ \dot\epsilon \ll 1
196: \end{equation}
197: Here, $V_{DA}$ is the donor-acceptor electronic coupling. Assuming the
198: rate of binding energy change as $\dot \epsilon \simeq k_b \Delta \epsilon $ and $V_{DA}$ of
199: the order of 1 cm$^{-1}$, eq \ref{eq:3} requires the rate of release
200: of B to be faster than 1 ns$^{-1}$. This threshold rate is comparable
201: to the rate of proton release from bound protonation states of
202: proteins,\cite{Gutman:97} while the time of water exchange between the
203: protein surface and the bulk is faster, in the range of tens of
204: picoseconds.\cite{Halle:04}
205:
206: Once particle B is released from its bound state, it becomes a part of
207: the bulk, which is the aqueous solution or a solvent mixture for water
208: binding or the ionic atmosphere for ionic (proton) association. The
209: interaction of the bulk with electronic states of the donor and
210: acceptor is relatively well understood. For water binding,
211: fluctuations of the dipolar polarization produce the thermal noise.
212: For electrolytes, the Debye-H{\"u}ckel electric field both shifts the
213: donor-acceptor energy gap and creates its fluctuations thus producing
214: a corresponding reorganization energy.\cite{German:92} This ionic
215: reorganization energy $\simeq \kappa^2 e^2 R / \epsilon_s$ scales quadratically with
216: the inverse Debye-H{\"u}ckel length $\kappa$ and linearly with a solute
217: dimension $R$. Since it is also inversely proportional to the static
218: dielectric constant of the solvent $\epsilon_s$, the effect of fluctuations
219: of the ionic atmosphere in polar solvents is commonly small relative
220: to the reorganization energy of polarization fluctuations. We will
221: therefore assume that once particle B is released from its bound
222: state, it does not interact any more with the transferred electron and
223: becomes a part of the many-particle electrostatic potential.
224:
225: This approximation allows us to apply the tight-binding approximation
226: and to use the following expression for the Hamiltonian of the
227: donor-acceptor system in a polar solvent
228: %
229: \begin{equation}
230: \label{eq:4}
231: \begin{split}
232: H_i & = E_{0i} - \left(\mathbf{E}_i^{u} |u\rangle\langle u| + \mathbf{E}_i^{b} |b\rangle\langle b|\right)*\mathbf{P} \\
233: & - \epsilon_i |b\rangle \langle b| + \frac{1}{2} \mathbf{P}*\chi^{-1}*\mathbf{P}
234: \end{split}
235: \end{equation}
236: Here, $E_{0i}$ incorporates the vacuum energy and the free energy of
237: solvation by the electronic degrees of freedom of the solvent and
238: $\mathbf{E}_i^{u,b}$ denotes the vacuum electric field of the
239: donor-acceptor complex in two ET states ($i=1,2$) and two binding
240: states (u,b). Also, $|b\rangle \langle b|$ and $|u\rangle \langle u|$ in eq \ref{eq:4}
241: describe the population operators for the bound (b) and unbound (u)
242: states of particle B with the binding energy $\epsilon_i$. One can expect
243: only a small change in the electric field of the donor-acceptor
244: complex for weak binding of a neutral molecule, $\mathbf{E}_i^b \simeq
245: \mathbf{E}_i^u$. However, the dependence of the electric field on the
246: binding state needs to be incorporated for equilibria of charged
247: particles.
248:
249: In application to protonation/deprotonation equilibria, the
250: Hamiltonian considered here (eq \ref{eq:4}) is different from the ones
251: used by Cukier\cite{Cukier:96} and Soudackov and
252: Hammes-Schiffer.\cite{Soudackov:00} In their case, the proton was
253: considered to move between two localized states within the
254: donor-acceptor complex, while in the problem considered here the
255: proton is released to the bulk and loses any connection to electronic
256: transitions within the donor-acceptor pair except for the influence of
257: the Debye-H{\"u}ckel field it becomes a part of. The different physics
258: of the problem considered here demands the different Hamiltonian.
259:
260: The electric field interacts with the (nuclear) dipolar polarization
261: of the solvent $\mathbf{P}$, and the asterisk in eq \ref{eq:4} implies
262: both the tensor contraction and the volume integration. The
263: statistics of $\mathbf{P}$ are Gaussian with the response function
264: $\mathbf{\chi}$ such that the solvent reorganization energy is
265: %
266: \begin{equation}
267: \label{eq:5}
268: \lambda_{u,b} = \frac{1}{2}\mathbf{\Delta E}_{u,b}*\chi*\mathbf{\Delta E}_{u,b}, \quad
269: \Delta \mathbf{E}_{u,b}= \mathbf{E}_2^{u,b} -\mathbf{E}_1^{u,b}
270: \end{equation}
271: Although $\lambda_{u,b}$ in eq \ref{eq:5} are formally the reorganization
272: energies of the dipolar polarization field of the solvent, we will
273: attach a more general meaning to them as the reorganization energies
274: of the nuclear degrees of freedom of the bulk. The corresponding
275: formal definition is easy to obtain from eq \ref{eq:4} by replacing
276: the Gaussian polarization noise with a sum of statistically
277: independent Gaussian fields.
278:
279: The free energies of ET are obtained from the Hamiltonian in eq
280: \ref{eq:4} by projecting all possible nuclear motions on the reaction
281: coordinate $X=\Delta H = H_2 - H_1$:
282: %
283: \begin{equation}
284: \label{eq:6}
285: e^{-\beta G_i(X)} = \int \mathcal{D} \mathbf{P}
286: \mathrm{Tr}\left[\delta(X - \Delta H) e^{-\beta H_i}\right]
287: \end{equation}
288: Here, $\mathcal{D} \mathbf{P}$ implies functional integration over the
289: field $\mathbf{P}$ and ``Tr'' refers to the sum over bound, $|b\rangle $, and unbound,
290: $|u\rangle$, states of particle B with statistical weights $f_{b,u}$:
291: %
292: \begin{equation}
293: \label{eq:7}
294: \mathrm{Tr}[\dots ] = \sum_{a=b,u} \langle a|f_a \dots |a\rangle .
295: \end{equation}
296: The ratio of the statistical weights gives the entropy of releasing
297: particle B to the bulk, $k_{\text{B}}\ln(f_u/f_b)$.
298:
299:
300: Taking the integral and trace in eq \ref{eq:6} results in the following
301: equation for the free energy surfaces of ET
302: %
303: \begin{equation}
304: \label{eq:8}
305: \begin{split}
306: G_i(X)& = G_i^u + \frac{(X-\Delta E_i^u)^2}{4\lambda_u} \\
307: & - \beta^{-1}\ln\left(1 + 10^{\mathrm{pK}_a^{(i)}-\mathrm{pB}}f_i(X)\right)
308: \end{split}
309: \end{equation}
310: In eq \ref{eq:8}, $G_i^u$ is the free energy of the donor-acceptor
311: complex in the unbound state. The solvation free energy entering
312: $G_i^u$ is thus of electrostatic origin and does not include the free
313: energy of binding particle B. Correspondingly, the vertical energy
314: gap $\Delta E_i^u$ in the second summand is given as $\lambda_u + \Delta G_u$ for the
315: forward transition $1\to 2$ and as $\lambda_u - \Delta G_u$ for the backward
316: transition $2\to 1$. Binding of particle B is expressed by the term
317: under the logarithm where function $f_i(X)$ is the ratio of Boltzmann
318: factors for activation through bound and unbound states:
319: %
320: \begin{equation}
321: \label{eq:8-1}
322: f_i(X) = \exp\left[-\beta \frac{(X+\Delta \epsilon - \Delta E_i^b)^2}{4\lambda_b} + \beta \frac{(X- \Delta E_i^u)^2}{4\lambda_u} \right]
323: \end{equation}
324: Here, $\Delta E_i^b$ are the vertical ET gaps in the bound state of
325: particle B which are given as $\lambda_b + \Delta G_b$ ($i=1$) and $\lambda_b - \Delta G_b$
326: ($i=2$).
327:
328:
329: In writing eq \ref{eq:8} we have also represented the Boltzmann factor
330: for the release of particle B in terms of pK$_a^{(i)}$ and pB$= -
331: \log[B]$ as follows
332: %
333: \begin{equation}
334: \label{eq:9}
335: e^{-\beta g_i - \beta \Delta G_i }= 10^{\mathrm{pB} - \mathrm{pK}_a^{(i)}}
336: \end{equation}
337: In this equation $g_i=\epsilon_i - k_{\text{B}}T\ln(f_u/f_b)$ is the free
338: energy of releasing the bound particle from the molecular fragment in
339: the donor-acceptor complex (e.g.\ tyrosine) and $\Delta G_i = G_i^u -
340: G_i^b$ is the change in the solvation energy of the entire
341: donor-acceptor complex caused by unbinding event. The equilibrium
342: constants $\mathrm{pK}_a^{(i)}$ thus reflects the equilibrium of the
343: entire complex and can be modified compared to the equilibrium
344: constants of the molecular fragment alone.
345:
346:
347: The free energies $G_i(X)$ can be used to calculate the first and
348: second moments of the reaction coordinate $X$, e.g.
349: %
350: \begin{equation}
351: \label{eq:10}
352: \langle X\rangle_i = Q_i^{-1} \int X e^{-\beta G_i(X)} dX
353: \end{equation}
354: where
355: %
356: \begin{equation}
357: \label{eq:11}
358: Q_i = \int e^{-\beta G_i(X)} dX
359: \end{equation}
360: One gets for the average
361: \begin{equation}
362: \label{eq:12}
363: \langle X\rangle_i = (1 - n_i) \Delta E_i^u + n_i (\Delta E_i^b- \Delta\epsilon)
364: \end{equation}
365: and for the variance
366: %
367: \begin{equation}
368: \label{eq:13}
369: \begin{split}
370: \langle (\delta X)^2 \rangle_i &= 2k_{\text{B}}T\left[\lambda_b n_i + \lambda_u(1-n_i)\right]\\
371: & +n_i(1-n_i)\left[\Delta E_i^u - \Delta E_i^b + \Delta \epsilon\right]^2
372: \end{split}
373: \end{equation}
374: In eqs \ref{eq:12} and \ref{eq:13},
375: %
376: \begin{equation}
377: \label{eq:14}
378: n_i = \left[1 + 10^{\mathrm{pB}-\mathrm{pK}_a^{(i)}} \right]^{-1}
379: \end{equation}
380: is the equilibrium population of the bound site.
381:
382: \begin{figure}
383: \centering
384: \includegraphics*[width=6.5cm]{fig3}
385: \caption{Free energy surfaces of ET (eq \ref{eq:14-1}) at different
386: values of pB indicated in the plot assuming pK$_a^{(1)}=10$ and
387: pK$_a^{(2)}=-2$ (equilibrium constants of tyrosine). The final
388: charge-transfer state ($i=2$) is always deprotonated while the
389: initial charge-transfer state ($i=1$) is protonated at
390: pB$<$pK$_a^{(1)}$ and is deprotonated at pB$>$pK$_a^{(1)}$. The
391: Stokes shift for the protonated state 1 is $2\lambda_s - \Delta\epsilon = 2.71$ eV
392: and is equal to $2\lambda_s=2$ eV for the deprotonated state 1. The
393: solvent bulk component of the reaction Gibbs energy reaction is $\Delta
394: G = -0.1$ eV. }
395: \label{fig:3}
396: \end{figure}
397:
398:
399: Equations \ref{eq:6}--\ref{eq:14} provide a general description of ET
400: when binding affects both the statistics of the energy gap
401: fluctuations and the electronic density responsible for the electric
402: field. In order to reduce the number of independent parameters, we
403: first neglect the second effect assuming that the electric field does
404: not change with binding, i.e.\ $\lambda_u=\lambda_b=\lambda_s$, $\Delta E_i^u=\Delta E_i^b=\Delta E_i$,
405: and $G_i^u=G_i^b=G_i$. This situation most closely corresponds to
406: binding of a neutral molecule (water) while the case of a charged
407: particle (protonation) is postponed to the next section.
408:
409: The free energy surfaces of ET can then be written as follows:
410: %
411: \begin{equation}
412: \label{eq:14-1}
413: \begin{split}
414: &G_i(X) = G_i + \frac{(X-\Delta E_i+\Delta \epsilon)^2}{4\lambda_s} \\
415: & - \beta^{-1} \ln\left(10^{\mathrm{pK}_a^{(i)} - \mathrm{pB}} + \exp\left[\frac{\beta\Delta\epsilon}{2\lambda_s}(X-\Delta E_i +\Delta\epsilon/2)\right]\right)
416: \end{split}
417: \end{equation}
418: where $G_i$ refers to the free energy of the donor acceptor complex
419: with particle B released to the solution.
420:
421: Figure \ref{fig:3} illustrates the change of the free energy surfaces
422: with pB according to eq \ref{eq:14-1} (pK$_a^{(i)}$ values of
423: tyrosine). For the reaction $1\to 2$, the system needs to climb the
424: activation barrier from the bottom of the free energy well $G_1(X)$ to
425: the crossing point at $X=0$. The bottom of the potential well rises
426: with increasing pB at pB$<$pK$_a^{(1)}$ thus resulting in a smaller
427: activation barrier. The reaction rate depends on pB. The barrier
428: height stops changing once pB reaches pK$_a^{(1)}$, and the reaction
429: rate is insensitive to pB at pB$>$pK$_a^{(1)}$. Notice that the rate
430: of the reverse transition $2\to 1$ remains unchanged in the whole range
431: of pB values. This invariance makes the effect of pB on the ET rate
432: distinct from the effect of the driving force which alters the
433: activation barriers for both the forward and backward reactions.
434:
435: When binding does not affect the electric field of the donor-acceptor
436: complex the equations for the first and second cumulants (eqs
437: \ref{eq:12} and \ref{eq:13}) simplify to
438: %
439: \begin{equation}
440: \label{eq:15}
441: \langle X\rangle_i = \Delta E_i - \Delta\epsilon n_i
442: \end{equation}
443: and
444: %
445: \begin{equation}
446: \label{eq:16}
447: \langle (\delta X)^2 \rangle_i = 2k_{\text{B}}T \lambda_s + \Delta\epsilon^2 n_i(1-n_i)
448: \end{equation}
449: Several important observations follow from eqs \ref{eq:15} and
450: \ref{eq:16}. First, the binding/unbinding fluctuations break the
451: connection (eq \ref{eq:0}) between the reorganization energy from the
452: free energy curvature
453: %
454: \begin{equation}
455: \label{eq:17}
456: \lambda_i = (\beta/2) \langle (\delta X)^2 \rangle_i
457: \end{equation}
458: and the Stokes shift
459: %
460: \begin{equation}
461: \label{eq:18}
462: \Delta X = 2\lambda_s + \Delta\epsilon \Delta n,\quad \Delta n = n_2 - n_1
463: \end{equation}
464: This effect is the consequence of the local character of the unbinding
465: events contrasting with the quasi-macroscopic nature of the bulk
466: fluctuations resulting in eq \ref{eq:0}. Second, the reorganization
467: energy in eqs \ref{eq:16} and \ref{eq:17} arising from binding
468: fluctuations depends on the state of the donor-acceptor complex, i.e.\
469: $\lambda_1 \neq \lambda_2$. This condition implies that the free energy surfaces
470: $G_i(X)$ must be non-parabolic\cite{DMacc:07} to comply with the
471: energy conservation,\cite{King:90} $G_2(X)=G_1(X)+X$. The effect of thermalized
472: bound states on ET cannot therefore be accounted for within the
473: Marcus-Hush formalism. We also note that the variance of the reaction
474: coordinate in eq \ref{eq:14} does not follow the
475: fluctuation-dissipation theorem\cite{Landau5} requiring $ \langle (\delta X)^2\rangle_i
476: \propto T$ and instead has a more complex temperature dependence arising
477: from the equilibrium dissociation constant entering the equilibrium
478: population in eqs \ref{eq:13} and \ref{eq:16}.
479:
480:
481: \begin{figure}
482: \includegraphics*[width=6.5cm]{fig4}
483: \caption{Stokes shift, $\Delta X$, and the reaction coordinate variance,
484: $\langle (\delta X)^2\rangle_1$, vs pB for a bound state with the dissociation
485: constants of tyrosine (pK$_a^{(1)}=10$ and pK$_a^{(2)}=-2$) and
486: the solvent reorganization energy $\lambda_s=1$ eV. }
487: \label{fig:4}
488: \end{figure}
489:
490: The dependence of both the energy gap variance and the Stokes shift on
491: pB are shown in Figure \ref{fig:4}. The equality of $\beta \langle (\delta X)^2\rangle_1$
492: and $\Delta X$, expected from eqs \ref{eq:0}, is seen when both the initial
493: and final ET states are in the same binding state. They are shifted by
494: the binding energy $\Delta \epsilon$ when the the binding states are different in
495: the two ET states (left corner in Figure \ref{fig:4}). The most
496: interesting region is when pB$\simeq$pK$_a$ and the term proportional to
497: $n_i(1-n_i)$ in eqs \ref{eq:13} and \ref{eq:16} can potentially
498: dominate the energy gap variance. Our model thus makes a simple,
499: experimentally testable prediction that the band-width of a
500: charge-transfer optical transition (absorption for binding to the
501: donor) is expected to pass through a maximum as a function of pB.
502:
503:
504:
505: \section{Protonation affecting electron transfer}
506: \label{sec:3}
507: Measurements of oxidation rates of primary pair P$_{680}^+$ by
508: tyrosine in Photosystem II have produced intriguing
509: results.\cite{Ahlbrink:98} The reaction rate increases with increasing
510: pH until it levels off at pH approximately equal to pK$_a$ of phenolic
511: proton. Similar results were reported by Hammarstr{\"o}m and co-workers
512: for intramolecular ET from tyrosine to Ru(III) covalently connected in
513: a donor-acceptor complex.\cite{Sjodin:00,Lomoth:06} In order to
514: explain the observed pH-dependence, they used the Marcus equation for
515: the activation free energy in which a pH-dependent redox potential was
516: substituted. The pH-dependence of the rate arises in their analysis
517: from the dependence of the tyrosine reduction potential on the
518: concentration of protons in the solution.
519:
520: This practice,\cite{Sjodin:00,Carra:03} which had not been anticipated
521: in the original Marcus formulation,\cite{Marcus:65} was criticized by
522: Krishtalik\cite{Krishtalik:03} and, more recently, by Sav{\'e}ant and
523: co-workers.\cite{Costenin:07} Krishtalik argued that the Gibbs free
524: energy of ET reactions cannot possibly depend on the ideal mixing
525: entropy of protons in the bulk, which is the origin of the
526: pH-dependent term in the Nernst equation for the redox
527: potential.\cite{Bockris:70} We can only add to this, absolutely
528: correct, argument that using a pH-dependent $\Delta G$ in the Marcus
529: equation clearly violates the Franck-Condon principle. The vertical
530: \textit{energy} gap $\Delta G + \lambda $ does not involve entropy change since
531: the nuclei do not move on the time-scale of electronic transitions.
532: When the entropic pH term appears in $\Delta G$ and is not compensated by a
533: corresponding term in $\lambda$, the unphysical entropy of protons mixing
534: appears in the vertical transition energy. The present model allows us
535: to account for the pH-dependence of the ET rate without introducing
536: unphysical approximations.
537:
538: \begin{figure}
539: \centering
540: \includegraphics*[width=6.5cm]{fig5}
541: \caption{Charge shift rates from tyrosine to Ru(III) from ref
542: \onlinecite{Sjodin:00} (points) and the fit of the data using eq
543: \ref{eq:14-1} (solid lines). The lower curve is obtained using
544: $\lambda_{u}=2.3$ eV, $\lambda_{b}=1.8$ eV, and the non-ergodicity multiplier
545: $\alpha=0.27$ ($\Delta G_u=-0.54$ eV and $\Delta G_b=0.2$ eV are taken from ref
546: \onlinecite{Carra:03}). The upper curve was obtained with the
547: same activation parameters by increasing the rate preexponent (see
548: text for discussion). The vertical arrow indicates the pK$_a^{(1)}$
549: value for tyrosine at which the ET rate becomes independent of pH.
550: }
551: \label{fig:5}
552: \end{figure}
553:
554: For protonation affecting ET, B$=$H and pB$=$pH. This problem is,
555: however, potentially more complex than binding/unbinding of neutral
556: molecules. The process of deprotonation proceeds by formation of a
557: geminate pair (on the time-scale of picoseconds for photoexcited
558: states\cite{Agmon:05}) followed by a slower diffusion of the released
559: proton in the Coulomb potential of the deprotonated negative
560: charge.\cite{Agmon:05,Perez:07} This slow process may imply that the
561: unbound proton will not be able to sample all possible states
562: available to a particle in an ideal solution on the time-scale of ET,
563: $\tau_{\text{ET}}$. A full account of such effects\cite{DMacc:07}
564: presently appears hard to achieve. We will therefore introduce an
565: empirical non-ergodicity multiplier $\alpha$ to account for the effects of
566: insufficient sampling. The term pK$_a^{(i)}-$pH in eq \ref{eq:8} is
567: replaced with $\alpha(\mathrm{pK}_a^{(i)}-\mathrm{pH})$, where $\alpha$ is a
568: non-ergodicity coefficient between zero and unity. With this
569: correction, the model qualitatively reproduces the dependence of the
570: ET rate on pH observed by Sj\"odin \textit{et al}.\cite{Sjodin:00} for
571: the oxidation of tyrosine by Ru(III).
572:
573: Sj\"odin \textit{et al}.\cite{Sjodin:00} have monitored the recovery of
574: Ru(II) from Ru(III) produced by flash photolysis and accelerated by
575: electron transfer from tyrosine covalently linked to the ruthenium
576: complex. Their observed rates, monitoring the arrival of electrons,
577: mathematically correspond to projecting the complex, potentially
578: multi-coordinate,\cite{Agmon:83,Sumi:86,Soudackov:00} dynamics of the
579: system onto one single ET reaction coordinate, which is exactly the
580: scenario considered here. Since proton is a charged particle, we need
581: the full formulation given by eq \ref{eq:8} to analyze the
582: experimental rates. The reaction Gibbs energies and reorganization
583: energies of ET will potentially be different for the bound and unbound
584: ET pathways and indeed the reaction free energies are $\Delta G_u = - 0.54$
585: eV and $\Delta G_b=0.2$ eV for the unbound and bound proton states,
586: respectively.\cite{Carra:03} The unknown parameters are the rate
587: preexponent, reorganization energies $\lambda_{u,b}$, and the non-ergodicity
588: multiplier $\alpha$. This latter parameter is expected to be small because
589: of the slow rate of proton dissociation from tyrosine\cite{Sjodin:00}
590: and thus a lower extent of modulation of the donor-acceptor gap by
591: unbinding events than it would be possible for full thermalization.
592:
593: \begin{figure}
594: \centering
595: \includegraphics*[width=6.5cm]{fig6}
596: \caption{Stokes shift, $\Delta X$, and the variance, $\langle(\delta
597: X)^2 \rangle_1$, calculated from eqs \ref{eq:12} and \ref{eq:13} using
598: the reorganization parameters obtained by fitting the experimental
599: data in Figure \ref{fig:5}. }
600: \label{fig:6}
601: \end{figure}
602:
603:
604: In order to see if the model can qualitatively account for the
605: experimental observations, we have used the rate preexponent,
606: reorganization energies $\lambda_{u,b}$, and $\alpha$ as free parameters to fit
607: the data from ref \onlinecite{Sjodin:00}. The result is shown in
608: Figure \ref{fig:5}. The fitting curve ($\lambda_u=2.3$ eV, $\lambda_b=1.8$ eV,
609: and $\alpha=0.26$) of $\log_{10}(k_{\text{ET}})$ vs pH rises linearly with
610: the slope $\alpha$ when pH is below pK$_a^{(1)}$ and levels off after
611: reaching this value. The reorganization energies $\lambda_{u,b}$ here
612: include reorganization of classical intramolecular vibrations in
613: addition to solvent reorganization. The parameters extracted from the
614: fit of the rates result in a dramatic violation of eq \ref{eq:0} as is
615: shown in Figure \ref{fig:6}.
616:
617:
618: Fitting the experimental jump in the rate at
619: $\mathrm{pH}=\mathrm{pK}_a^{(1)}$ requires a higher value of the rate
620: preexponent as is shown by the fragment of the curve obtained by using
621: the same parameters as in the low-pH fit, but varying the preexponent.
622: The increase of the rate at $\mathrm{pH}=\mathrm{pK}_a^{(1)}$ has been
623: addressed by Carra \textit{et al}.\cite{Carra:03} The low-pH portion
624: of the curve corresponds to electronic transition accompanied by
625: simultaneous proton release (proton-coupled
626: ET\cite{Hammes-Schiffer:01}). The nonadiabatic matrix element in the
627: rate preexponent then includes small Franck-Condon overlap between
628: proton bound and free states.\cite{Cukier:96} This overlap disappears
629: in the flat portion of the plot at $\mathrm{pH}>\mathrm{pK}_a^{(1)}$
630: when tyrosine is deprotonated in both ET states. Only electronic
631: coupling enters the rate preexponent in this regime resulting in a
632: higher rate. A more quantitative analysis would require extensive
633: calculations. Since the model presented here does not aim at a
634: detailed quantitative analysis of proton-coupled ET, we limit our
635: discussion to qualitative observations only.
636:
637:
638:
639:
640: \section{Conclusions}
641: \label{sec:4}
642: %
643: Traditional theories of ET in polar liquids emphasize activation of
644: electronic transitions by long-range, quasi-macroscopic solvent
645: fluctuations.\cite{Marcus:65} Local binding to the donor and acceptor
646: sites provides an additional modulation of the donor-acceptor energy
647: gap. This study addressed the question of whether this additional
648: thermal noise can be accounted for within the standard framework of
649: Gaussian fluctuations of the energy gap, that is by adjusting the
650: magnitudes of the driving force and reorganization energy.
651:
652: We have found that the statistics of binding events are non-Gaussian,
653: and the resultant free energy surfaces cannot be reduced to crossing
654: parabolas. The model predicts a regime of a significant dependence of
655: the activation barrier on the concentration of the binding particles
656: in the solution. When the concentration pB is close to the binding
657: equilibrium constant pK$_a$, the variance of the energy gap passes
658: through a maximum which can be observed by spectroscopy of
659: charge-transfer bands. Despite these new features, the main effect of
660: binding is in shifting the free energy gap of ET, as was suggested
661: previously,\cite{Mayer:04} leaving the reorganization energy mostly
662: unaffected and within the realm of standard models. Finally, the
663: model helps to rationalize some recent observations of the dependence
664: of rates of intramolecular ET, involving deprotonation of reactants,
665: on the pH of the solution.
666:
667: \begin{acknowledgments}
668: This work was supported by the NSF (CHE-0616646).
669: \end{acknowledgments}
670:
671:
672: %\bibliographystyle{apsrev}
673: % \bibliography{/home/dmitry/p/bib/chem_abbr,/home/dmitry/p/bib/photosynthNew,/home/dmitry/p/bib/glass,/home/dmitry/p/bib/et,/home/dmitry/p/bib/dm,/home/dmitry/p/bib/pt,/home/dmitry/p/bib/solvation,/home/dmitry/p/bib/heteroet,/home/dmitry/p/bib/bioet,/home/dmitry/p/bib/protein}
674:
675: \providecommand{\refin}[1]{\\ \textbf{Referenced in:} #1}
676: \begin{thebibliography}{10}
677:
678: \bibitem{Reichardt:94}
679: Reichardt,~C. \textit{Chem.\ Rev.} \textbf{1994,} \textsl{94,} 2319-2358.
680:
681: \bibitem{DMjpcb:97}
682: Matyushov,~D.~V.;\ \ Schmid,~R.;\ \ Ladanyi,~B.~M. \textit{J. Phys. Chem. B}
683: \textbf{1997,} \textsl{101,} 1035.
684:
685: \bibitem{SolvPolarity:04}
686: Katritzky,~A.~R.;\ \ Fara,~D.~C.;\ \ Yang,~H.;\ \ T{\"a}mm,~K. \textit{Chem.
687: Rev.} \textbf{2004,} \textsl{104,} 175.
688:
689: \bibitem{Agmon:05}
690: Agmon,~N. \textit{J. Phys. Chem. A} \textbf{2005,} \textsl{109,} 13.
691:
692: \bibitem{McEvoy:06}
693: McEvoy,~J.~P.;\ \ Brudvig,~G.~W. \textit{Chem. Rev.} \textbf{2006,}
694: \textsl{106,} 4455.
695:
696: \bibitem{Cukier:04}
697: Cukier,~R.~I. \textit{Biochim. Biophys. Acta} \textbf{2004,} \textsl{1655,}
698: 37-44.
699:
700: \bibitem{Cukier:96}
701: Cukier,~R. \textit{J.\ Phys.\ Chem.} \textbf{1996,} \textsl{100,} 15428.
702:
703: \bibitem{Soudackov:00}
704: Soudackov,~A.;\ \ Hammes-Schiffer,~S. \textit{J. Chem. Phys.} \textbf{2000,}
705: \textsl{113,} 2385.
706:
707: \bibitem{Hammes-Schiffer:01}
708: Hammes-Schiffer,~S. \textit{Acc. Chem. Res.} \textbf{2001,} \textsl{34,} 273.
709:
710: \bibitem{Ferguson-Miller:96}
711: Ferguson-Miller,~S.;\ \ Babcock,~G. \textit{Chem. Rev.} \textbf{1996,}
712: \textsl{96,} 2889.
713:
714: \bibitem{Kirby:97}
715: Kirby,~A. \textit{Acc. Chem. Res.} \textbf{1997,} \textsl{30,} 290-296.
716:
717: \bibitem{Mayer:04}
718: Mayer,~J.~M. \textit{Annu. Rev. Phys. Chem.} \textbf{2004,} \textsl{55,} 363.
719:
720: \bibitem{Marcus:93}
721: Marcus,~R.~A. \textit{Rev. Mod. Phys.} \textbf{1993,} \textsl{65,} 599.
722:
723: \bibitem{Borgis:96}
724: Borgis,~D.;\ \ Hynes,~J. \textit{J.\ Phys.\ Chem.} \textbf{1996,} \textsl{100,}
725: 1118.
726:
727: \bibitem{Marcus:65}
728: Marcus,~R.~A. \textit{J. Chem. Phys.} \textbf{1965,} \textsl{43,} 679.
729:
730: \bibitem{King:90}
731: King,~G.;\ \ Warshel,~A. \textit{J. Chem. Phys.} \textbf{1990,} \textsl{93,}
732: 8682.
733:
734: \bibitem{Marcus:89}
735: Marcus,~R.~A. \textit{J. Phys. Chem.} \textbf{1989,} \textsl{93,} 3078.
736:
737: \bibitem{DMacc:07}
738: Matyushov,~D.~V. \textit{Acc. Chem. Res.} \textbf{2007,} \textsl{40,} 294.
739:
740: \bibitem{Sumi:86}
741: Sumi,~H.;\ \ Marcus,~R.~A. \textit{J. Chem. Phys.} \textbf{1986,} \textsl{84,}
742: 4894.
743:
744: \bibitem{Stuchebrukhov:03}
745: Stuchebrukhov,~A.~A. \textit{J. Theor. Comp. Chem.} \textbf{2003,} \textsl{2,}
746: 91.
747:
748: \bibitem{Gutman:97}
749: Gutman,~M.;\ \ Nachliel,~E. \textit{Annu. Rev. Phys. Chem.} \textbf{1997,}
750: \textsl{48,} 329.
751:
752: \bibitem{Halle:04}
753: Halle,~B. \textit{Phil. Trans. R. Soc. Lond.} \textbf{2004,} \textsl{359,}
754: 1207.
755:
756: \bibitem{German:92}
757: German,~E.~D.;\ \ Kuznetsov,~A.~M. \textit{Electrokhimiya} \textbf{1992,}
758: \textsl{28,} 294.
759:
760: \bibitem{Landau5}
761: Landau,~L.~D.;\ \ Lifshits,~E.~M. \textit{Statistical Physics;} Pergamon Press:
762: New York, 1980.
763:
764: \bibitem{Ahlbrink:98}
765: Ahlbrink,~R.;\ \ Haumann,~M.;\ \ Cherepanov,~D.;\ \ B{\"o}gershausen,~O.;\ \
766: Mulkidjanian,~A.;\ \ Junge,~W. \textit{Biochemistry} \textbf{1998,}
767: \textsl{37,} 1131.
768:
769: \bibitem{Sjodin:00}
770: Sj{\"o}din,~M.;\ \ Styring,~S.;\ \ {\AA}kermark,~B.;\ \ Sun,~L.;\ \
771: Hammarstr{\"o}m,~L. \textit{J. Am. Chem. Soc.} \textbf{2000,} \textsl{122,}
772: 3932.
773:
774: \bibitem{Lomoth:06}
775: Lomoth,~R.;\ \ Magnuson,~A.;\ \ Sj{\"o}din,~M.;\ \ Huang,~P.;\ \ Styring,~S.;\
776: \ Hammarstr{\"o}m,~L. \textit{Photosynth. Res.} \textbf{2006,} \textsl{87,}
777: 25.
778:
779: \bibitem{Carra:03}
780: Carra,~C.;\ \ Iordanova,~N.;\ \ Hammes-Schiffer,~S. \textit{J. Am. Chem. Soc.}
781: \textbf{2003,} \textsl{125,} 10429.
782:
783: \bibitem{Krishtalik:03}
784: Krishtalik,~L.~I. \textit{Biochim. Biophys. Acta} \textbf{2003,} \textsl{1604,}
785: 13.
786:
787: \bibitem{Costenin:07}
788: Costenin,~C.;\ \ Robert,~M.;\ \ Sav{{\'e}}ant,~J.-M. \textit{J. Am. Chem. Soc.}
789: \textbf{2007,} \textsl{129,} 5870.
790:
791: \bibitem{Bockris:70}
792: Bockris,~J.~O. \textit{Modern Electrochemistry;} McDonald: London, 1970.
793:
794: \bibitem{Perez:07}
795: P{\'e}rez-Lustres,~J.~L.;\ \ Rodriguez-Prieto,~F.;\ \ Mosquera,~M.;\ \
796: Senyushkina,~T.~A.;\ \ Ernsting,~N.~P.;\ \ Kovalenko,~S.~A. \textit{J. Am.
797: Chem. Soc.} \textbf{2007,} \textsl{129,} 19.
798:
799: \bibitem{Agmon:83}
800: Agmon,~N.;\ \ Hopfield,~J.~J. \textit{J. Chem. Phys.} \textbf{1983,}
801: \textsl{78,} 6947.
802:
803: \end{thebibliography}
804:
805:
806:
807:
808: \end{document}
809:
810: \newpage
811:
812: \begin{center}
813: \textbf{FIGURE CAPTIONS}
814: \end{center}
815:
816: \begin{description}
817: \item[FIGURE 1] Diagram of the free energy surfaces of ET obtained from
818: three parameters, the reaction free energy $\Delta G$, the free energy
819: curvature $(2\lambda)^{-1}$, and the Stokes shift (separation between
820: the minima $\Delta X$). The reaction coordinate $X$ is the energy
821: separation between the donor and acceptor electronic states. The
822: activated state is given by the resonance of the donor and
823: acceptor levels, $X=0$, while $X_{01}$ and $X_{02}$ indicate the
824: positions of the minima.
825: \item[FIGURE 2] Diagram of particle B exchanging between the bound and
826: dissolved states with the rate constant $k_b$. The equilibrium
827: population of the bound state assumes that the unbinding rate is
828: much faster than the rate of ET ($k_{\text{ET}}\ll k_b$); $\Delta
829: \epsilon=\epsilon_2-\epsilon_1$ denotes the change in the binding energy caused by the
830: electronic transition.
831: \item[FIGURE 3] Free energy surfaces of ET (eq \ref{eq:14-1}) at different
832: values of pB indicated in the plot assuming pK$_a^{(1)}=10$ and
833: pK$_a^{(2)}=-2$ (equilibrium constants of tyrosine). The final
834: charge-transfer state ($i=2$) is always deprotonated while the
835: initial charge-transfer state ($i=1$) is protonated at
836: pB$<$pK$_a^{(1)}$ and is deprotonated at pB$>$pK$_a^{(1)}$. The
837: Stokes shift for the protonated state 1 is $2\lambda_s - \Delta\epsilon = 2.71$ eV
838: and is equal to $2\lambda_s=2$ eV for the deprotonated state 1. The
839: solvent bulk component of the reaction Gibbs energy reaction is $\Delta
840: G = -0.1$ eV.
841: \item[FIGURE 4] Stokes shift, $\Delta X$, and the reaction coordinate variance,
842: $\langle (\delta X)^2\rangle_1$, vs pB for a bound state with the dissociation
843: constants of tyrosine (pK$_a^{(1)}=10$ and pK$_a^{(2)}=-2$) and
844: the solvent reorganization energy $\lambda_s=1$ eV.
845: \item[FIGURE 5] Charge shift rates from tyrosine to Ru(III) from ref
846: \onlinecite{Sjodin:00} (points) and the fit of the data using eq
847: \ref{eq:14-1} (solid lines). The lower curve is obtained using
848: $\lambda_{u}=2.3$ eV, $\lambda_{b}=1.8$ eV, and the non-ergodicity multiplier
849: $\alpha=0.27$ ($\Delta G_u=-0.54$ eV and $\Delta G_b=0.2$ eV are taken from ref
850: \onlinecite{Carra:03}). The upper curve was obtained with the
851: same activation parameters by increasing the rate preexponent (see
852: text for discussion). The vertical arrow indicates the pK$_a^{(1)}$
853: value for tyrosine at which the ET rate becomes independent of pH.
854: \item[FIGURE 6] Stokes shift, $\Delta X$, and the variance, $\langle(\delta
855: X)^2 \rangle_1$, calculated from eqs \ref{eq:12} and \ref{eq:13} using
856: the reorganization parameters obtained by fitting the experimental
857: data in Figure \ref{fig:5}.
858: \end{description}
859:
860: \newpage
861: \includegraphics*[width=9cm]{fig1}
862: \vfill
863: \textbf{Figure 1. Matyushov}
864: \newpage
865: \includegraphics*[width=9cm]{fig2}
866: \vfill
867: \textbf{Figure 2. Matyushov}
868: \newpage
869: \includegraphics*[width=9cm]{fig3}
870: \vfill
871: \textbf{Figure 3. Matyushov}
872:
873: \newpage
874: \includegraphics*[width=9cm]{fig4}
875: \vfill
876: \textbf{Figure 4. Matyushov}
877:
878: \newpage
879: \includegraphics*[width=9cm]{fig5}
880: \vfill
881: \textbf{Figure 5. Matyushov}
882:
883: \newpage
884: \includegraphics*[width=9cm]{fig6}
885: \vfill
886: \textbf{Figure 6. Matyushov}
887:
888: \end{document}
889: