Next Article in Journal
Thiourea Derivative Metal Complexes: Spectroscopic, Anti-Microbial Evaluation, ADMET, Toxicity, and Molecular Docking Studies
Previous Article in Journal
Hybrids Composed of an Fe-Containing Wells–Dawson Polyoxometalate and Carbon Nanomaterials as Promising Electrocatalysts for the Oxygen Reduction Reaction
Previous Article in Special Issue
Structure–Function Relationship within Cu-Peptoid Electrocatalysts for Water Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Applicability of Transition State Theory to the (Proton-Coupled) Electron Transfer in Photosynthetic Water Oxidation with Emphasis on the Entropy of Activation

Department of Physics, Freie Universität Berlin, 14195 Berlin, Germany
*
Authors to whom correspondence should be addressed.
Inorganics 2023, 11(10), 389; https://doi.org/10.3390/inorganics11100389
Submission received: 9 August 2023 / Revised: 23 September 2023 / Accepted: 25 September 2023 / Published: 30 September 2023

Abstract

:
Recent advancements in the study of the protein complex photosystem II have clarified the sequence of events leading to the formation of oxygen during the S3 → S4 → S0 transition, wherein the inorganic Mn4Ca(µ-O)6(OHx)4 cluster finishes photo-catalyzing the water splitting reaction (Greife et al., Nature 2023, 617, 623–628; Bhowmick et al., Nature 2023, 617, 629–636). During this final step, a tyrosine radical (TyrZ), stable for a couple of milliseconds, oxidizes a cluster-bound oxygen while the hydrogen bonding patterns of nearby waters shift a proton away. A treatment of this redox reaction within the context of accepted transition state theories predicts rate constants that are significantly higher than experimentally recovered values (1012 s−1 versus 103 s−1). In an effort to understand this disparity, temperature-dependent experiments have revealed large entropic contributions to the rates with only a moderate enthalpy of activation. We suggest that the entropic source may be related to the observed proton rearrangements, and further possible near isoenergetic variations in the nearby extended H-bonding network delaying the realization of an ‘ideal’ transition state. In the following, we explore this relation in the context of Eyring’s transition state theory and Marcus’ electron transfer theory and evaluate their compatibility with the experimental evidence.

Graphical Abstract

1. Introduction

In oxygenic photosynthesis, water is oxidized at a Mn4Ca(µ-O)6(OHx)4 complex (denoted here as the Mn4Ca-cluster) bound to the proteins of photosystem II (PSII) [1,2]. Light drives water oxidation at the Mn4Ca-cluster indirectly [3,4], with a primary charge separation occurring at the electron donor side of PSII resulting in the formation of a delocalized chlorophyll cation radical (P680+). This is followed by the oxidation of a specific tyrosine residue, denoted as TyrZ or YZ. The subsequent electron transfer (ET) steps from the Mn4Ca cluster to TyrZox are closely related to the water oxidation chemistry coupled with proton movements over a large area on various length scales, as can be seen in Figure 1. Herein, we will discuss, in a broad sense, whether the Mn4Ca cluster–TyrZox ET can be described favorably by transition state (TS) theory in the framework of the Eyring–Polanyi equation [5,6,7], or whether elements of ET theories, e.g., the well-known semi-classical Marcus theory [8,9], are applicable or even essential for a correct description. We note that the following Results and Discussion section corresponds in parts to Sections SII.8 and SII.9 of the Supplementary Material of Greife et al. [10]. Our discussion specifically addresses the water oxidation reaction (or oxygen evolution reaction) in PSII, but it may, in parts, also be of relevance for the redox reactions of other proteins, molecules, and redox-active electrodes in electrocatalysis.

2. Results and Discussion

2.1. TS Theory Applied to Oxygen-Evolution Step of PSII

In Eyring’s TS theory, the reaction rate (kTST) is governed by the Eyring–Polanyi equation [5,6] involving the entropy (ΔS) and enthalpy (ΔH) of activation:
k T S T = k B T h κ e Δ S k B e Δ H k B T
with kB, h, and T denoting the Boltzmann constant, the Planck constant, and the absolute temperature in Kelvin. The κ parameter describes the transmission probability, i.e., the probability that the reaction will occur once the TS is reached; typically, κ-values of 0.5 or 1.0 are assumed to be reasonable [5,6]. In Equation (1), the activation enthalpy, ΔH, describes the temperature dependence of the rate constant and is closely related to the activation energy, Ea, as determined by conventional Arrhenius analysis (ΔHEaκBT0, where the difference between ΔH and Ea stems from the weak temperature dependence of k B T h in Equation (1), as can be seen in ref. [12]). The activation entropy, ΔS, is usually a negative number that corresponds to a temperature-independent reduction in the kTST.
In research on PSII water oxidation, Equation (1) has been applied to the rate constant of the oxygen-evolution transition [10,13,14], which is initiated by the third flash of saturating light applied to dark-adapted PSII [15]. The oxygen evolution transition is referred to as the S3 → S0 transition and involves an ET from the Mn4Ca-cluster to TyrZox. This ET step precedes the O–O bond formation and O2-release. It is generally believed that the rate-determining step of the overall reaction is the Mn4Ca-cluster—TyrZox ET step, which we suggested to be coupled to the movement of four protons [10], as can be seen in Figure 2.
Based on an experimental determination of the rate constants for the O2-formation at temperatures ranging from −5 °C to 40 °C, we calculated (by application of Equation (1)) the enthalpy and entropy of activation of the rate-determining (proton-coupled) ET step for PSII of the thermophilic cyanobacterium Thermosynechococcus elongatus (T. elongatus), as can be seen in Extended Data Figure 5 and elsewhere in Greife et al. [10]. We obtained values for the activation enthalpy of 310 meV (7.15 kcal/mol) and the activation entropy of 284 meV (6.55 kcal/mol, value of −T0ΔS for T0 = 293 K), for κ = 1 [10]. A variation of the κ-value from 1.0 to 0.5 would leave the activation enthalpy unchanged and only marginally affects the numerical value of the activation entropy.
For the PSII of T. elongatus and a closely related thermophilic cyanobacterium, Thermosynechococcus vulcanus (T. vulcanus), detailed crystallographic data are available, not only for the dark-stable state, but also for intermediates of the reaction cycle [11,16,17,18,19,20]. Due to the availability of atomic-resolution structures, computational studies on the mechanism of PSII water oxidation have mostly been based on coordinates obtained for PSII of T. elongatus or T. vulcanus [21,22,23,24,25,26,27,28]. Therefore, herein we focus on the enthalpy and entropy of activation experimentally determined for the O2-formation step of T. elongatus.
The application of Equation (1) to experimental data is straightforward and has resulted in values for the enthalpy and entropy of activation for the rate-determining ET transfer step in the oxygen evolution transition. The question arises as to how the application of TS theory relates to ET theories.

2.2. Relation to ET Theory

In the framework of the Eyring–Polanyi equation, essentially instantaneous ET would have to occur whenever the appropriate nuclear coordinates are reached, corresponding to a κ-value near unity in Equation (1). In terms of non-adiabatic ET theory, however, the relocation of electron density (and thus spin density) does not always occur once a suitable nuclear geometry has been reached, but with a finite probability that is determined by the electron-tunneling distance. A tunneling probability well below the frequency factor in Equation (1) might mimic an entropic contribution to the activation energy. Therefore, the possibility of whether a non-adiabaticity of the ET step could significantly contribute to the experimentally determined entropy of activation is explored in the following.
Following Moser et al. [29], we estimate the tunneling distance as the molecular edge-to-edge distance between donor and acceptor moiety, where, herein, the acceptor moiety is the tyrosine radical and the donor is the Mn4Ca cluster. The ‘edges’ or rather the extension of the relevant orbital systems are not obvious and cannot reliably be identified without using computational approaches. We calculated several potentially relevant interatomic (internuclear) distances, as can be seen in Figure 3 and Table 1. An internuclear distance close to 5.7 Å is obtained between TyrZ-O and O6. This distance could be especially relevant because, in the rate-limiting ET step of the S3 → S0 transition, the formation of the O6 radical is facilitated by ET to the oxidized TyrZ, with a high spin density expected on the phenolic oxygen. This corresponds to a van der Waals tunneling distance of only about 2.5 Å (estimate obtained by subtracting a van der Waals radius of 1.6 Å [30] for each oxygen atom from the internuclear (TyrZ-)O–O6 distance). Other potentially relevant internuclear distances in Table 1 are even shorter, predicting a particularly high tunneling probability. According to Moser et al., the tunneling distance of 2.5 Å corresponds to a free-energy optimized rate constant for electron tunneling on the order of 1012 s−1, or a time constant of 1 ps. The frequency that describes nuclear movements in the TS regime is about six times greater (kBT/h in Equation (1), about 6 × 1012 s−1 at 20 °C) suggesting that for oxygen-evolution steps in PSII, a minor slow-down resulting from a limiting ET probability cannot be excluded. Staying in the framework of TS theory, such a slowdown could be described by a reduced transmission probability (reduced κ-value), which would result in only a small reduction in the activation entropy that we calculated from experimental data using the Eyring–Polanyi equation (Equation (1)).
The above considerations involve several approximations and estimates. Presumably, the tunneling probability exceeds the above estimate of 1012 s−1 because the ‘medium’ between the donor and acceptor moieties comprises atom groups (specifically W3 and W4) that could significantly lower the tunneling barrier. In addition, multiple parallel tunneling paths could also increase the overall tunneling probability. Thus, it is well conceivable that the adiabatic limit, where electron tunneling probabilities become irrelevant, is reached.
In summary, uncertainties remain regarding the exact numerical value of the activation entropy, because of possible limitations due to electron tunneling. However, it is likely that the limiting case of adiabatic ET is almost, or even fully reached. If the adiabatic ET limit were not reached, the activation entropy would be moderately lower than the value formally determined by the application of Equation (1) to experimental data. However, this does not affect the central conclusion that there is a pronounced entropic slowdown of the oxygen-evolution transition.
Based on minimum energy path (MEP) calculations, we proposed that the ET from the Mn4Ca cluster to TyrZox is coupled to the movement of four specific protons [10]. However, the used computational approach cannot clarify whether concerted ET and proton transfer (PT) is involved, that is, the simultaneous tunneling of the electron and one or more protons at the TS. This question is of relevance when assessing the significance of activation entropy formally calculated by the application of Equation (1) to experimental data, because, in the case of a truly concerted electron–proton transfer (CEPT), the proton tunneling probability could influence the reaction rate and mimic an entropic slowdown. However, the low value of the experimentally observed H/D kinetic isotope effect (about 1.2 [31,32]) clearly disfavors a limitation of the reaction rate by the proton-tunneling probability.
Further evidence supporting the entropic slowdown hypothesis comes from experimental indications for entropy–enthalpy compensation in the total free energy of activation of the oxygen-evolution step. In PSII of higher plants (spinach) at room temperature, the rate constant is very similar to that of the thermophilic cyanobacterium (T. elongatus). However, the Arrhenius activation energy, and consequently, the activation enthalpy, are significantly lower for the higher-plant PSII [14]. For the cyanobacterium Synechocystis sp. PCC 6803, again another combination of action enthalpy and entropy has been reported [13]. We propose that the compensation of the decreased activation enthalpy by an increased activation entropy explains the identical rate constant values for differing organisms at similar temperatures in the physiological temperature range. This entropy–enthalpy compensation in the activation energy of the oxygen-evolution step is still insufficiently understood and requires further investigation (work in progress). To this end, the experimental determination of the entropy and enthalpy of activation for various organisms as well as for PSII with genetically modified H-bond networks could address both (i) a possible physiological role in the genetic adaptation of various organisms to different environmental temperatures; and (ii) the elucidation of the molecular determinants (in terms of differing amino acid environments) of the experimentally observable shifts between the enthalpy and entropy of activation, complemented by computational investigation. In general, entropy–enthalpy compensation is well known, albeit somewhat controversial, for the free energy of activation relating to various catalytic processes in biological and non-biological systems [33,34,35,36,37,38,39]. Irrespectively of the explanation of entropy–enthalpy compensation at the molecular level, finding entropy–enthalpy compensation in PSII oxygen evolution supports the existence of a major entropic contribution to the total free energy of activation, and it disfavors alternative explanations like a dominating role of electron or proton tunneling probabilities.

2.3. Origin of Activation Entropy and Its Relation to Marcus Theory

In PSII water oxidation, the environment of the two key protagonists, TyrZ and the Mn4Ca cluster, is characterized by an extended network of H-bonded water molecules and amino-acid residues. See Figure 4 for a two-dimensional mapping of the H-bond network (HBN) within 20 Å of the Mn4Ca cluster. There are over 40 resolved waters to be found, grouped into several different water channels. Some of these groups are likely involved in proton and water transport and have been shown to be variable during the S3 → S0 transition [10,11]. Similarly, it has been shown that a significant amount (>50.000) of quasi-isoenergetic protonation ‘microstates’ exist in bacterial photosynthetic reaction center photocycle intermediates [40]. We propose that multiple conformations, that is, multiple H-bond patterns with multiple orientations of water molecules and residues, can explain the experimentally found activation entropy in PSII. The weak interactions of this internal network could facilitate the observed compensation in a similar manner to what has been proposed in other systems with surface bound solvation water interactions [33,41,42,43]. The interplay between evolutionary optimized order (stabilization of favorable conformations) and thermally driven dynamics needs to be considered when discussing the functional relevance of the extended H-bonded protein–water network. Following Greife et al. [10] (Supplementary Information, section SII.9), we point out that:
(1) It is expected and has been verified by extensive molecular dynamics simulations on PSII that the protein internal H-bond networks (HBN) are highly dynamic, especially regarding water positions and H-bonding directionality, with a multitude of roughly isoenergetic conformations reached within nanoseconds at room temperature (see, e.g., ref. [44]). Even though these HBN dynamics may, in part, be conducive regarding PT reactions, they are largely ‘inevitable’ thermodynamic fluctuations.
(2) The arrangement of the water molecules in the extended H-bonded protein–water network surrounding the Mn4CaO5 cluster and the TyrZ is well resolved in the crystallographic structures determined at both cryogenic and room temperature. This implies that the HBN dynamics evolve around the mean positions of the individual nuclei that are detected by protein crystallography. The HBN and the related mean-value atom coordinates are likely evolutionarily optimized for efficient (fast) water oxidation.
In light of (1) and (2), the qualitative explanation for the energy of activation in Greife et al. [10] appears plausible: ‘Although well-defined coordinates of individual oxygen atoms are resolved in crystal structures, the presence of an HBN that is at the same time and in every detail perfectly arranged for the here discussed proton-coupled ET, will still be a rare event. The limited probability to reach this perfect conformation of all atoms of the HBN explains the entropic contributions to the activation energy.’
We emphasize that, even if the optimal HBN conformation perfectly matches the atomic coordinates resolved by protein crystallography, it will still be rare that all atoms are simultaneously at their evolutionarily optimized positions at physiological temperatures.
Figure 4. Two-dimensional representation of the resolved extended H-bonded water (blue circles) and amino acid (red hexagons) network within 20 Å of the Mn4Ca cluster. Distances less than 2.7 Å are indicated with a solid line while distances less than 3.3 Å are dashed (distances based of the structure PDB 8EZ5 [11]). A black line marks an association with a sidechain while a grey line indicates a backbone connection. The colored borders of water molecules represent participation in water channels possibly used for water and proton transport [20,28,45,46]. Only oxygen and nitrogen atoms were included in the analysis (H-bond distances measured between O/N donor and O/N acceptor). While the first-sphere ligands of the Mn4Ca cluster were included in the analysis, connections within are not shown for simplicity. Numbering is based on the structure PDB 8EZ5 and Supplementary Table 1 of ref. [11]. Waters not present in the table are denoted by their chain and residue ID number.
Figure 4. Two-dimensional representation of the resolved extended H-bonded water (blue circles) and amino acid (red hexagons) network within 20 Å of the Mn4Ca cluster. Distances less than 2.7 Å are indicated with a solid line while distances less than 3.3 Å are dashed (distances based of the structure PDB 8EZ5 [11]). A black line marks an association with a sidechain while a grey line indicates a backbone connection. The colored borders of water molecules represent participation in water channels possibly used for water and proton transport [20,28,45,46]. Only oxygen and nitrogen atoms were included in the analysis (H-bond distances measured between O/N donor and O/N acceptor). While the first-sphere ligands of the Mn4Ca cluster were included in the analysis, connections within are not shown for simplicity. Numbering is based on the structure PDB 8EZ5 and Supplementary Table 1 of ref. [11]. Waters not present in the table are denoted by their chain and residue ID number.
Inorganics 11 00389 g004

3. Conclusions

Marcus and related theories on ET or CEPT reactions are powerful tools because they can predict rate constants based on a small number of clearly defined parameters, which are often experimentally accessible. The extended H-bonded protein–water network in PSII might be seen as a special case of the solvent environment that in Marcus theory is described involving a quadratic energy dependence with the outer-sphere reorganization energy being the (only) decisive parameter [9]. However, such an approximative treatment does not consider individual solvent molecules. Instead, Marcus theory assumes a homogeneous solvent environment which responds to charges at the redox factors by a linear dielectric response; the total energy of charge solvation is decisive for the contribution of the outer-sphere reorganization to the activation energy.
In contrast to the considerations on the ET in PSII of the Results and Discussion section, in Marcus theory, there is no special conformation of the solvent environment foreseen that uniquely supports the ET step; it is the total electrostatic solvation energy that matters, realized by a multitude of conformations of the solvent environment. Therefore, the underlying approximations of Marcus theory render its applicability to redox reactions in a protein environment with multiple conformations of a specific, evolutionary optimized protein–water network problematic, inter alia because it cannot easily capture the entropic contributions to the free energy of activation discussed herein. Whether the predictive power of ET theory can be fruitfully used regarding the contribution of inner-sphere processes of contributing redox factors (in PSII and other redox proteins), remains to be answered. In conclusion, the common assumption (in ET theories) of nuclear movements in approximately quadratic potentials needs to be scrutinized for redox factors surrounded by a fluctuating protein-internal HBN.
In summary, our considerations support the appropriateness of discussing the rate-limiting proton-coupled ET step in photosynthetic water oxidation in terms of Eyring’s TS theory. We point out the important but apparently poorly understood role of activation entropy and hope that future work can shed more light on the interplay between order and statistical dynamics in PSII water oxidation, biological redox processes in general, homogeneous and heterogeneous catalysis.

Author Contributions

The main text was written, edited, and reviewed by H.D. and P.G., with the initial conceptualization by H.D. Data analysis and visualization was performed by P.G. Supervision, project administration and funding acquisition was performed by H.D. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) provided to the collaborative research center on Protonation Dynamics in Protein Function (SFB 1078, project A4/Dau, DFG project number 221545957) and under Germany’s Excellence Strategy EXC 2008: UniSysCat (DFG project number 390540038).

Data Availability Statement

The structural data used herein are publicly available from the protein data bank (PDB).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Junge, W. Oxygenic photosynthesis: History, status and perspective. Q. Rev. Biophys. 2019, 52, e1. [Google Scholar] [CrossRef] [PubMed]
  2. Cox, N.; Pantazis, D.A.; Lubitz, W. Current Understanding of the Mechanism of Water Oxidation in Photosystem II and Its Relation to XFEL Data. Annu. Rev. Biochem. 2020, 89, 795–820. [Google Scholar] [CrossRef] [PubMed]
  3. Dau, H.; Haumann, M. Eight steps preceding O-O bond formation in oxygenic photosynthesis--a basic reaction cycle of the Photosystem II manganese complex. Biochim. Biophys. Acta 2007, 1767, 472–483. [Google Scholar] [CrossRef] [PubMed]
  4. Tran, R.; Kern, J.; Hattne, J.; Koroidov, S.; Hellmich, J.; Alonso-Mori, R.; Sauter, N.K.; Bergmann, U.; Messinger, J.; Zouni, A.; et al. The Mn4Ca photosynthetic water-oxidation catalyst studied by simultaneous X-ray spectroscopy and crystallography using an X-ray free-electron laser. Philos. Trans. R. Soc. B 2014, 369, 20130324. [Google Scholar] [CrossRef] [PubMed]
  5. Eyring, H. The activated complex in chemical reactions. J. Chem. Phys. 1935, 3, 107–115. [Google Scholar] [CrossRef]
  6. Evans, M.G.; Polanyi, M. Some applications of the transition state method to the calculation of reaction velocities, especially in solution. J. Chem. Soc. Faraday Trans. 1935, 31, 875–894. [Google Scholar] [CrossRef]
  7. Mortimer, R.G.; Eyring, H. Elementary transition state theory of the Soret and Dufour effects. Proc. Natl. Acad. Sci. USA 1980, 77, 1728–1731. [Google Scholar] [CrossRef]
  8. Marcus, R.A. On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. I. J. Chem. Phys. 1956, 24, 966–978. [Google Scholar] [CrossRef]
  9. Marcus, R.A.; Sutin, N. Electron transfers in chemistry and biology. Biochim. Biophys. Acta 1985, 811, 265–322. [Google Scholar] [CrossRef]
  10. Greife, P.; Schönborn, M.; Capone, M.; Assunção, R.; Narzi, D.; Guidoni, L.; Dau, H. The electron–proton bottleneck of photosynthetic oxygen evolution. Nature 2023, 617, 623–628. [Google Scholar] [CrossRef]
  11. Bhowmick, A.; Hussein, R.; Bogacz, I.; Simon, P.S.; Ibrahim, M.; Chatterjee, R.; Doyle, M.D.; Cheah, M.H.; Fransson, T.; Chernev, P.; et al. Structural evidence for intermediates during O2 formation in photosystem II. Nature 2023, 617, 629–636. [Google Scholar] [CrossRef] [PubMed]
  12. Guo, Y.; Sekharan, S.; Liu, J.; Batista, V.S.; Tully, J.C.; Yan, E.C.Y. Unusual kinetics of thermal decay of dim-light photoreceptors in vertebrate vision. Proc. Natl. Acad. Sci. USA 2014, 111, 10438–10443. [Google Scholar] [CrossRef] [PubMed]
  13. Bao, H.; Burnap, R.L. Structural rearrangements preceding dioxygen formation by the water oxidation complex of photosystem II. Proc. Natl. Acad. Sci. USA 2015, 112, E6139–E6147. [Google Scholar] [CrossRef] [PubMed]
  14. Assunção, R.; Zaharieva, I.; Dau, H. Ammonia as a substrate-water analogue in photosynthetic water oxidation: Influence on activation barrier of the O2-formation step. Biochim. Biophys. Acta 2019, 1860, 533–540. [Google Scholar] [CrossRef]
  15. Kok, B.; Forbush, B.; McGloin, M. Cooperation of charges in photosynthetic O2 evolution - I. A linear four-step mechanism. Photochem. Photobiol. 1970, 11, 457–475. [Google Scholar] [CrossRef]
  16. Kern, J.; Alonso-Mori, R.; Tran, R.; Hattne, J.; Gildea, R.J.; Echols, N.; Glöckner, C.; Hellmich, J.; Laksmono, H.; Sierra, R.G.; et al. Simultaneous Femtosecond X-ray Spectroscopy and Diffraction of Photosystem II at Room Temperature. Science 2013, 340, 491–495. [Google Scholar] [CrossRef]
  17. Suga, M.; Akita, F.; Sugahara, M.; Kubo, M.; Nakajima, Y.; Nakane, T.; Yamashita, K.; Umena, Y.; Nakabayashi, M.; Yamane, T.; et al. Light-induced structural changes and the site of O=O bond formation in PSII caught by XFEL. Nature 2017, 543, 131–135. [Google Scholar] [CrossRef]
  18. Kern, J.; Chatterjee, R.; Young, I.D.; Fuller, F.D.; Lassalle, L.; Ibrahim, M.; Gul, S.; Fransson, T.; Brewster, A.S.; Alonso-Mori, R.; et al. Structures of the intermediates of Kok’s photosynthetic water oxidation clock. Nature 2018, 563, 421–425. [Google Scholar] [CrossRef]
  19. Kato, K.; Miyazaki, N.; Hamaguchi, T.; Nakajima, Y.; Akita, F.; Yonekura, K.; Shen, J.R. High-resolution cryo-EM structure of photosystem II reveals damage from high-dose electron beams. Commun. Biol. 2021, 4, 382. [Google Scholar] [CrossRef]
  20. Hussein, R.; Ibrahim, M.; Bhowmick, A.; Simon, P.S.; Chatterjee, R.; Lassalle, L.; Doyle, M.; Bogacz, I.; Kim, I.S.; Cheah, M.H.; et al. Structural dynamics in the water and proton channels of photosystem II during the S2 to S3 transition. Nat. Commun. 2021, 12, 6531. [Google Scholar] [CrossRef]
  21. Siegbahn, P.E. Mechanisms for proton release during water oxidation in the S2 to S3 and S3 to S4 transitions in photosystem II. Phys. Chem. Chem. Phys. 2012, 14, 4849–4856. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, J.; Askerka, M.; Brudvig, G.W.; Batista, V.S. Crystallographic Data Support the Carousel Mechanism of Water Supply to the Oxygen-Evolving Complex of Photosystem II. ACS Energy Lett. 2017, 2, 2299–2306. [Google Scholar] [CrossRef] [PubMed]
  23. Narzi, D.; Capone, M.; Bovi, D.; Guidoni, L. Evolution from S3 to S4 state of the oxygen evolving complex in Photosystem II monitored by QM/MM dynamics. Chem. Eur. J. 2018, 24, 10820–10828. [Google Scholar] [CrossRef] [PubMed]
  24. Shoji, M.; Isobe, H.; Shigeta, Y.; Nakajima, T.; Yamaguchi, K. Concerted Mechanism of Water Insertion and O2 Release during the S4 to S0 Transition of the Oxygen-Evolving Complex in Photosystem II. J. Phys. Chem. B 2018, 122, 6491–6502. [Google Scholar] [CrossRef] [PubMed]
  25. Capone, M.; Guidoni, L.; Narzi, D. Structural and dynamical characterization of the S4 state of the Kok-Joliot’s cycle by means of QM/MM Molecular Dynamics Simulations. Chem. Phys. Lett. 2020, 742, 137111. [Google Scholar] [CrossRef]
  26. Isobe, H.; Shoji, M.; Suzuki, T.; Shen, J.-R.; Yamaguchi, K. Exploring reaction pathways for the structural rearrangements of the Mn cluster induced by water binding in the S3 state of the oxygen evolving complex of photosystem II. J. Photochem. Photobiol. A Chem. 2021, 405, 112905. [Google Scholar] [CrossRef]
  27. Allgöwer, F.; Gamiz-Hernandez, A.P.; Rutherford, A.W.; Kaila, V.R.I. Molecular Principles of Redox-Coupled Protonation Dynamics in Photosystem II. J. Am. Chem. Soc. 2022, 144, 7171–7180. [Google Scholar] [CrossRef]
  28. Sirohiwal, A.; Pantazis, D.A. Functional Water Networks in Fully Hydrated Photosystem II. J. Am. Chem. Soc. 2022, 144, 22035–22050. [Google Scholar] [CrossRef]
  29. Moser, C.C.; Keske, J.M.; Warncke, K.; Farid, R.S.; Dutton, P.L. Nature of biological electron transfer. Nature 1992, 355, 796–802. [Google Scholar] [CrossRef]
  30. Bondi, A. van der Waals Volumes and Radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  31. Gerencser, L.; Dau, H. Water oxidation by Photosystem II: H2O−D2O Exchange and the influence of pH support formation of an intermediate by removal of a proton before dioxygen creation. Biochemistry 2010, 49, 10098–10106. [Google Scholar] [CrossRef]
  32. Klauss, A.; Haumann, M.; Dau, H. Alternating electron and proton transfer steps in photosynthetic water oxidation. Proc. Natl. Acad. Sci. USA 2012, 109, 16035–16040. [Google Scholar] [CrossRef] [PubMed]
  33. Borsarelli, C.D.; Braslavsky, S.E. Enthalpy, Volume, and Entropy Changes Associated with the Electron Transfer Reaction between the 3MLCT State of Ru(Bpy)32+ and Methyl Viologen Cation in Aqueous Solutions. J. Phys. Chem. A 1999, 103, 1719–1727. [Google Scholar] [CrossRef]
  34. Losi, A.; Wegener, A.A.; Engelhard, M.; Braslavsky, S.E. Enthalpy−Entropy Compensation in a Photocycle: The K-to-L Transition in Sensory Rhodopsin II from Natronobacterium pharaonis. J. Am. Chem. Soc. 2001, 123, 1766–1767. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, L.; Guo, Q.-X. Isokinetic Relationship, Isoequilibrium Relationship, and Enthalpy−Entropy Compensation. Chem. Rev. 2001, 101, 673–696. [Google Scholar] [CrossRef]
  36. Ferrante, A.; Gorski, J. Enthalpy-entropy compensation and cooperativity as thermodynamic epiphenomena of structural flexibility in ligand-receptor interactions. J. Mol. Biol. 2012, 417, 454–467. [Google Scholar] [CrossRef]
  37. Chodera, J.D.; Mobley, D.L. Entropy-enthalpy compensation: Role and ramifications in biomolecular ligand recognition and design. Annu. Rev. Biophys. 2013, 42, 121–142. [Google Scholar] [CrossRef]
  38. Åqvist, J.; Kazemi, M.; Isaksen, G.V.; Brandsdal, B.O. Entropy and Enzyme Catalysis. Acc. Chem. Res. 2017, 50, 199–207. [Google Scholar] [CrossRef]
  39. Cornish-Bowden, A. Enthalpy–entropy compensation and the isokinetic temperature in enzyme catalysis. J. Biosci. 2017, 42, 665–670. [Google Scholar] [CrossRef]
  40. Khaniya, U.; Mao, J.; Wei, R.J.; Gunner, M.R. Characterizing Protein Protonation Microstates Using Monte Carlo Sampling. J. Phys. Chem. B 2022, 126, 2476–2485. [Google Scholar] [CrossRef]
  41. Lumry, R.; Rajender, S. Enthalpy-entropy compensation phenomena in water solutions of proteins and small molecules: A ubiquitous property of water. Biopolymers 1970, 9, 1125–1227. [Google Scholar] [CrossRef] [PubMed]
  42. Dunitz, J.D. Win some, lose some: Enthalpy-entropy compensation in weak intermolecular interactions. Chem. Biol. 1995, 2, 709–712. [Google Scholar] [CrossRef] [PubMed]
  43. Breiten, B.; Lockett, M.R.; Sherman, W.; Fujita, S.; Al-Sayah, M.; Lange, H.; Bowers, C.M.; Heroux, A.; Krilov, G.; Whitesides, G.M. Water Networks Contribute to Enthalpy/Entropy Compensation in Protein–Ligand Binding. J. Am. Chem. Soc. 2013, 135, 15579–15584. [Google Scholar] [CrossRef]
  44. Guerra, F.; Siemers, M.; Mielack, C.; Bondar, A.N. Dynamics of Long-Distance Hydrogen-Bond Networks in Photosystem II. J. Phys. Chem. B 2018, 122, 4625–4641. [Google Scholar] [CrossRef] [PubMed]
  45. Bondar, A.-N.; Dau, H. Extended protein/water H-bond networks in photosynthetic water oxidation. Biochim. Biophys. Acta 2012, 1817, 1177–1190. [Google Scholar] [CrossRef]
  46. Sakashita, N.; Ishikita, H.; Saito, K. Rigidly hydrogen-bonded water molecules facilitate proton transfer in photosystem II. Phys. Chem. Chem. Phys. 2020, 22, 15831–15841. [Google Scholar] [CrossRef]
Figure 1. H-bonded protein–water network surrounding the Mn4Ca cluster of PSII (S3-state structure, PDB 8EZ5 of T. elongatus [11], rendered in Pymol). The water molecules ligated to Mn1 and Ca (W1, W2, W3, and W4) are part of an extended H-bonding network reaching from the redox-active TyrZ towards the Glu65/Glu312 ‘proton gate’. H-bond distances of less than or equal to 3.4 Å are indicated by dashed yellow lines. The depicted amino-acid sidechains are either part of the PsbA (D1) or PsbD (D2) subunits; TyrZ denotes Tyr161 of the PsbA subunit.
Figure 1. H-bonded protein–water network surrounding the Mn4Ca cluster of PSII (S3-state structure, PDB 8EZ5 of T. elongatus [11], rendered in Pymol). The water molecules ligated to Mn1 and Ca (W1, W2, W3, and W4) are part of an extended H-bonding network reaching from the redox-active TyrZ towards the Glu65/Glu312 ‘proton gate’. H-bond distances of less than or equal to 3.4 Å are indicated by dashed yellow lines. The depicted amino-acid sidechains are either part of the PsbA (D1) or PsbD (D2) subunits; TyrZ denotes Tyr161 of the PsbA subunit.
Inorganics 11 00389 g001
Figure 2. Scheme of the electron transfer between O6 and TyrZ with accompanying proton rearrangements as reported in Greife et al. [10]. Alongside the ET, resulting in the oxidation of O6, three protons move within an H-bond (from H-bond donor to H-bond acceptor) shifting the protonation by about 7 Å from O6 to Asp61. This proton-coupled electron transfer deprotonates and oxidizes O6, thereby allowing for the subsequent peroxide formation between O5 and O6. The ET step is further coupled to the movement of a fourth proton from His190 to TyrZ (not shown).
Figure 2. Scheme of the electron transfer between O6 and TyrZ with accompanying proton rearrangements as reported in Greife et al. [10]. Alongside the ET, resulting in the oxidation of O6, three protons move within an H-bond (from H-bond donor to H-bond acceptor) shifting the protonation by about 7 Å from O6 to Asp61. This proton-coupled electron transfer deprotonates and oxidizes O6, thereby allowing for the subsequent peroxide formation between O5 and O6. The ET step is further coupled to the movement of a fourth proton from His190 to TyrZ (not shown).
Inorganics 11 00389 g002
Figure 3. Shortest distances between the sidechain of Tyr161 (TyrZ) and the first-sphere ligand atoms (all oxygen atoms) of the Mn and Ca atoms of the Mn4Ca cluster in the S3 state (PDB 8EZ5 [11], rendered in PyMol). For seven distances, connecting lines are indicated starting at four oxygen atoms, which are W4, W3, O6, and one carboxylate oxygen of the ligating Asp170. The seven corresponding distances and further distances are indicated in Table 1 with comparative distances for the structure PDB 6DHO (2nd flash, predominantly S3 state) [18].
Figure 3. Shortest distances between the sidechain of Tyr161 (TyrZ) and the first-sphere ligand atoms (all oxygen atoms) of the Mn and Ca atoms of the Mn4Ca cluster in the S3 state (PDB 8EZ5 [11], rendered in PyMol). For seven distances, connecting lines are indicated starting at four oxygen atoms, which are W4, W3, O6, and one carboxylate oxygen of the ligating Asp170. The seven corresponding distances and further distances are indicated in Table 1 with comparative distances for the structure PDB 6DHO (2nd flash, predominantly S3 state) [18].
Inorganics 11 00389 g003
Table 1. Distances between atoms of the Mn4Ca cluster and the nearest atoms of TyrZ for two S3-state coordinates sets (8EZ5 of ref. [11], 6DHO of ref. [18]).
Table 1. Distances between atoms of the Mn4Ca cluster and the nearest atoms of TyrZ for two S3-state coordinates sets (8EZ5 of ref. [11], 6DHO of ref. [18]).
AtomDistance to YZO (Å)
8EZ56DHO
W42.72.9
3.6 (to CZ ring)3.7
W33.94
4.1 (to CE2 ring)4.2
Ca4.64.7
(CA1-Asp170) OD24.9 (to CE2 ring)5.1
5.35.4
(CA1-Ala344) O5.65.7
O65.75.7
(Mn4-Asp170) OD15.8 (to CE2 ring)5.8
O16.16.2
(Mn1-Glu189) OE26.26.1
Mn17.07.1
Mn47.47.4
Mn27.77.7
Mn38.08.1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dau, H.; Greife, P. Applicability of Transition State Theory to the (Proton-Coupled) Electron Transfer in Photosynthetic Water Oxidation with Emphasis on the Entropy of Activation. Inorganics 2023, 11, 389. https://doi.org/10.3390/inorganics11100389

AMA Style

Dau H, Greife P. Applicability of Transition State Theory to the (Proton-Coupled) Electron Transfer in Photosynthetic Water Oxidation with Emphasis on the Entropy of Activation. Inorganics. 2023; 11(10):389. https://doi.org/10.3390/inorganics11100389

Chicago/Turabian Style

Dau, Holger, and Paul Greife. 2023. "Applicability of Transition State Theory to the (Proton-Coupled) Electron Transfer in Photosynthetic Water Oxidation with Emphasis on the Entropy of Activation" Inorganics 11, no. 10: 389. https://doi.org/10.3390/inorganics11100389

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop