Next Article in Journal
Preliminary Study on the Restoration of the Phospholipid Profile in Serum from Patients with COVID-19 by Treatment with Vitamin E
Previous Article in Journal
Isolation of a Virulent Clostridium perfringens Strain from Elaphurus davidianus and Characterization by Whole-Genome Sequence Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Removal of External Proteins PsbO, PsbP and PsbQ on Flash-Induced Molecular Oxygen Evolution and Its Biphasicity in Tobacco PSII

Faculty of Physics and Applied Computer Science, AGH University of Krakow, al. Mickiewicza 30, 30-059 Krakow, Poland
*
Author to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2024, 46(7), 7187-7218; https://doi.org/10.3390/cimb46070428
Submission received: 2 June 2024 / Revised: 30 June 2024 / Accepted: 2 July 2024 / Published: 8 July 2024
(This article belongs to the Topic Metalloproteins and Metalloenzymes)

Abstract

:
The oxygen evolution within photosystem II (PSII) is one of the most enigmatic processes occurring in nature. It is suggested that external proteins surrounding the oxygen-evolving complex (OEC) not only stabilize it and provide an appropriate ionic environment but also create water channels, which could be involved in triggering the ingress of water and the removal of O2 and protons outside the system. To investigate the influence of these proteins on the rate of oxygen release and the efficiency of OEC function, we developed a measurement protocol for the direct measurement of the kinetics of oxygen release from PSII using a Joliot-type electrode. PSII-enriched tobacco thylakoids were used in the experiments. The results revealed the existence of slow and fast modes of oxygen evolution. This observation is model-independent and requires no specific assumptions about the initial distribution of the OEC states. The gradual removal of exogenous proteins resulted in a slowdown of the rapid phase (~ms) of O2 release and its gradual disappearance while the slow phase (~tens of ms) accelerated. The role of external proteins in regulating the biphasicity and efficiency of oxygen release is discussed based on observed phenomena and current knowledge.

Graphical Abstract

1. Introduction

The appearance of the first photosynthetic organisms capable of extracting electrons and protons from water determined the direction of the evolution of living organisms on Earth. This is the result of a process called oxygenic photosynthesis, which is responsible for increasing and maintaining the amount of O2 in the Earth’s atmosphere. Only organisms possessing photosystem II (PSII), including cyanobacteria, algae, and higher plants, are able to oxidize water 2H2O → 4e + 4H++ O2↑. PSII is the only photosystem with a high enough redox potential for water oxidation [1,2]. It has a water-oxidizing enzyme known as an oxygen-evolving complex (OEC), which can cyclically accumulate four positive charges, bind water, release electrons and protons, and support the formation of O=O bonds and the release of O2 molecules [3,4,5]. The PSII complex and the components involved in the linear electron transport process are shown in Figure 1a. PSII is a light-driven plastoquione oxidoreductase.
A major step forward in understanding the functioning of the OEC was Joliot’s observation of periodic oscillations of oxygen evolution in dark-adapted chloroplasts under the influence of short saturating flashes [3] and the linear four-step model of the oxidation cycle proposed by Kok et al. [4]. The cycling transient states of the OEC are called Si(i = 0,1,2,3,4) states and are assigned to a particular arrangement and oxidation state of the Mn4CaO5 complex (subscript indicates the number of accumulated charges, see Figure 1c, Equations (S1)–(S3) in the Supplementary Materials). When sufficient oxidizing power is accumulated, water molecules are split, an O=O bond is formed, and O2 is released.
The detected high degree of conformational variability of the individual Si states may suggest an adaptive ability of the Mn4CaO5 complex to maintain an optimal level of OEC activity under certain external and internal conditions. Two redox isomers of the S2 state with distinctive EPR signals were detected [6,7,8,9]. It was shown that the S2 state may exist in a low spin (LS) state characterized by a multiline signal with g ≈ 2 and in a high spin (HS) state with g ≈ 4. These two states are isoenergetic and can be converted into each other (Figure 1b). The transition between open (denoted A, LS) and closed (denoted B, HS) Mn4CaO5 structures is associated with changes in the coupling between the Mn ions that form the manganese cluster, resulting in an Mn1(III)Mn4(IV) ↔ Mn1(IV)Mn4(III) electron exchange and O5 ligand displacement [10]. Recently, even two stable transient S2 closed conformers were suggested [11]. The heterogeneity observed in S2 may already be the result of two different intermediate states S1YZ (where YZ is a tyrosyl radical) [12], for which two isomeric S1 states can coexist as a consequence of a quasi-reversible structure change induced by proton migration [13]. Two S1 conformations, closed and open, resembling S2 states, have been proposed [14]. Due to the orientational Jahn–Teller isomerism, they could convert to each other (open with LS = 1 and closed with HS = 3), and this would determine which Mn(III), Mn1, or Mn4 is oxidized [15,16,17,18]. Two S0 isomers (Mn(III, IV, III, III) with relatively similar energies analogous to the open and closed states of S2 were also considered. This time, a singly protonated O5 was assumed in both structures and assigned as a slow H2O-exchanging substrate [19,20], but the open form of S0 was considered preferable [21]. In the subsequent steps of cyclic water oxidation, the S3 state, activated by two flashes, also shows isomerism. This is probably closely correlated with the two different conformations of the S2 state. In the S3 state, all manganese ions remain oxidized Mn(IV), but depending on whether Mn4(IV) is five- or six-coordinated, the Mn4CaO5 complex adopts different spin and conformational states, respectively, as either an LS state (S = 3), with a stable closed cubane form, or an HS state (S = 6), with a stable open cubane form [17,22,23,24]. Thus, the S1, S2, and S3 states can adopt at least two potentially stable conformations, while it is proposed that the S0 state exists rather than an open cubane [21,25].
Figure 1. (a) Simplified scheme of photosystem II (PSII) for higher plants localized in the inner membranes of chloroplasts, called thylakoids. For clarity, small proteins are omitted except for cytochrome b559 [26,27]. Two polypeptides, D1 (PsbA) and D2 (PsbD), form the core of PSII. Two inner antenna subunits, CP43 and CP47, are attached to it. The OEC on the PSII donor side, containing four Mn ions, one Ca ion, and five oxygen atoms forming the Mn4CaO5 cluster, is protected in higher plants by three outer proteins of about 17 kDa (PsbQ), 23 kDa (PsbP), and 33 kDa (PsbO) [28]. The thick orange lightning shows the excitation of the chlorophyll pair at the reaction center, P680, and the black arrows show the direction of linear electron flow within PSII. (b) The currently accepted open (left column) and closed (right column) conformations of the S2 and S3 states and their possible transformations upon attaching a water molecule are shown. The water molecules W1 and W2 (here it is deprotonated) are ligands to Mn4, and W3 and W4 are ligands to Ca. The ligand O5 is nominally closest to the entrance of the broad channel (also called Cl1 channel), the ligand O4 to the entrance of the narrow channel (also called O4 channel), and the ligand O1 to the entrance of the large channel (also called O1 channel) [29,30,31,32,33,34,35]. The additional water molecule bound to Ca ion during the S2 → S3 transition described in [36], which can be a precursor of the O6 atom, is not included here. (c) Scheme of the 5S-state model explicitly considering the S4 state and its longer-lived isomer S4, which are associated with fast (d—the probability parameter) and slow ((1 − d)—the probability parameter) O2 release, respectively [37,38]. The transition probability between Si → Si+1 states is (1 − αi), where αi is the miss parameter.
Figure 1. (a) Simplified scheme of photosystem II (PSII) for higher plants localized in the inner membranes of chloroplasts, called thylakoids. For clarity, small proteins are omitted except for cytochrome b559 [26,27]. Two polypeptides, D1 (PsbA) and D2 (PsbD), form the core of PSII. Two inner antenna subunits, CP43 and CP47, are attached to it. The OEC on the PSII donor side, containing four Mn ions, one Ca ion, and five oxygen atoms forming the Mn4CaO5 cluster, is protected in higher plants by three outer proteins of about 17 kDa (PsbQ), 23 kDa (PsbP), and 33 kDa (PsbO) [28]. The thick orange lightning shows the excitation of the chlorophyll pair at the reaction center, P680, and the black arrows show the direction of linear electron flow within PSII. (b) The currently accepted open (left column) and closed (right column) conformations of the S2 and S3 states and their possible transformations upon attaching a water molecule are shown. The water molecules W1 and W2 (here it is deprotonated) are ligands to Mn4, and W3 and W4 are ligands to Ca. The ligand O5 is nominally closest to the entrance of the broad channel (also called Cl1 channel), the ligand O4 to the entrance of the narrow channel (also called O4 channel), and the ligand O1 to the entrance of the large channel (also called O1 channel) [29,30,31,32,33,34,35]. The additional water molecule bound to Ca ion during the S2 → S3 transition described in [36], which can be a precursor of the O6 atom, is not included here. (c) Scheme of the 5S-state model explicitly considering the S4 state and its longer-lived isomer S4, which are associated with fast (d—the probability parameter) and slow ((1 − d)—the probability parameter) O2 release, respectively [37,38]. The transition probability between Si → Si+1 states is (1 − αi), where αi is the miss parameter.
Cimb 46 00428 g001
To unravel the mystery of the mechanism of O-O bond formation and oxygen release, it seems crucial to know the last of the metastable states of the Kok cycle, i.e., the S3 state. Therefore, knowing the individual stages of S2S3 transition is as important as understanding the possible intermediate S4 states during the flash that triggers S3 → (S4) → S0 transition [39,40,41,42,43]. It is now known that any electron uptake from Mn4CaO5 is preceded by the oxidation of TyrZ (Yz) and its deprotonation [36,44,45] and that the accumulation of positive charges on the manganese cluster during the transitions between SiSi+1 states requires prior proton release [36,46,47,48,49,50,51,52,53]. The only exception is the transition from S0S1, when the transfer of electrons to YZ precedes deprotonation [52,53]. There is consensus that during S0S1 and S1S2 transitions, oxidation is concentrated on Mn [54,55], but there is no such consensus for S2S3 transition, where some groups favor an interpretation of the data as indicating Mn-centered oxidation [23,56,57,58] and others suggest ligand-centered oxidation [59,60,61,62,63]. Furthermore, opinions remain divided on proton release during S1 → S2 transition. Since no proton release into the bulk has been observed, it is generally assumed that no proton release occurs during this transition [36,43,44,45]. However, during this transition, both theory and experiments suggest that the OEC is deprotonated and H+ is taken up in the water clusters, thus preventing the release of protons to the bulk [48,49,52,64,65,66,67].
There are several experimental approaches, combined with computational ones, that attempt to explain how the different conformational forms of the Si states, in particular S2 and S3, can affect the mechanism of oxygen release and its efficiency. At present, two pathways of O-O bond formation involving Mn4CaO5 are the most considered for reconstructing the Kok cycle. One of them assumes an oxo-oxyl radical coupling mechanism, where the S4 intermediate state may involve the electrophilic Mn(V)-oxo or Mn(IV)-oxyl radical [46,68,69,70,71]. The other one involves a mechanism of nucleophilic attack by water [29,72,73,74]. A completely different mechanism has been proposed for O-O bond formation within the MnVII dioxo site on Mn4 [75]. Furthermore, different scenarios for the pathway leading to the formation of the O-O bond have been proposed, depending on the assumption of the conformation of the subsequent Si states and the stage at which the second exchangeable substrate water is bound to the Mn4CaO5 cluster. For example, one of the models considers additional H2O binding during S2S3 transition and only open structures of Si states [42], while the other assumes that structural changes in the Mn4CaO5 complex (open ↔ close) for the Si states facilitate the coordination of substrate water binding, proton release, and O-O formation [43,75,76,77,78].
The coupling of PSII conformational changes, which can lead to the reorganization of the immediate environment of the OEC, regulating substrate water access, proton release, and ultimately dioxygen formation, and thus affecting its enzyme activity, has been suggested many times; for example, see [78,79,80,81]. In addition to the protein network, the stability/variability of the hydrogen bonds between the water molecules, as well as between the water molecules and the amino acid residues, plays a vital role in this process (for review, see [81]). The presence of more water molecules in the vicinity of the manganese complex, which may be directly or indirectly involved in the oxygen evolution process, was predicted via EPR, mass spectroscopy, or FTIR measurements, among others [50,82,83], and further confirmed in structural studies of the OEC [26,27,84].
Due to the cyclic nature of the light-driving force-dependent operation of the Mn4CaO5 cluster and its location and protection by external proteins on the lumenal side of the thylakoids, it has been suggested that the pathways for the entry and exit of substrates (water molecules) and products (protons and O2) may be essential for the efficient functioning of PSII [85,86,87,88]. To date, based on theoretical studies, three water channels have been identified that have counterparts in cyanobacteria, algae, and higher plants [89,90]. Two of them, large and broad, are branched and, using the nomenclature based on the entry point of each channel into the OEC region, are also known as the O1 and Cl1 channels, respectively. The third one is a single narrow channel, also known as the O4 channel. In cyanobacteria, the O4 channel is formed by residues D1, D2, CP43, CP47, PsbO (a 33 kDa external protein), and PsbU (an external protein subunit). It connects O4 to the lumen with the participation of protonated D1-D61. But on the other side of D1-D61, there is the Cl1 channel that is at the interface of D2 and PsbO subunits [27,91,92]. The Cl1 channel with its branching arms is indicated as an H-channel rather than a water delivery path [92,93,94,95], although the latter function cannot be ruled out either [27,32]. The O1 water chain, a branched network, is formed in cyanobacteria by the same protein subunits as the Cl1 channel, with one difference: instead of PsbO, the PsbV subunit is involved. This channel is thought to remove O2 and/or deliver water to the OEC [91,96,97,98]. On the other hand, the sometimes recognized so-called back channel appears to be inaccessible to water but permeable to O2 [99,100]. Hydrophobic channels could be an effective way to remove O2 [91]. Among the many water channels reported in cyanobacteria or algae, the organization of amino acids of the Cl1 channel starting at Mn4, W1, W2, and W3 is the most evolutionary conserved. The recognized proton gate residues D1-E65/D2-E312/D1-R334/D1-N335 associated with the proton network rearrangement along this channel were found in all photosynthetic species [89]. The O4 channel is also very conservative at its entrance to the manganese complex near O4 and W1, including D1-D61, but shows a different orientation at its end near the surface. The O1 (large) channel, like the O4 channel, shows a high degree of conservation in the vicinity of the Mn4CaO5 complex, reaching Ca, O1, and W4, but at the same time, has different orientations at the end of the path close to the bulk, as well as different degrees of branching. For more details, see: [89,90,101]. In plants, however, fewer PSII subunits form the channels mentioned above. The subunits D1, CP43, PsbP (23 kDa external protein), and PsbQ (17 kDa external protein) are involved in forming the O1 water chain, D1, D2, CP43, CP47, and PsbP are involved in forming the O4 water chain, and only three subunits D1, D2 and PsbO (33 kDa external protein) are involved in forming the Cl1 water chain [90,101]. The small radius of the O4 channel (~1.4 Å) in all species suggests that water molecules are also arranged as a single chain in higher plants as well [90]. A rigid water chain with particularly strong hydrogen bonds between the initial water molecules near the Mn4CaO5 cluster and with its O4 atom indicates that the narrow channel can transport protons via the Grotthuss mechanism since the activation energy of proton transfer is lowest when all water molecules are strongly bound in the H-bond network [53,102,103]. Proton release during the S0S1 [42,53,103,104] and S2S3 transition via this channel was proposed [71]. However, the involvement of the O4 channel in the removal of protons during S2S3 transition was considered to be rather unlikely [42,89], and the Cl1 channel has been proposed as a proton release pathway during this transition [105,106,107,108], with D1-E65 (branch Cl1A) serving as the gate for proton transport to minimize the back reaction [40,89]. Recently, this has been confirmed experimentally through the use of pump–probe time-resolved femtosecond crystallography (TR-SFX) [36]. On the other hand, the O4 channel is suggested to be responsible for supplying water to the OEC during S4S0 transition [77]. Compared with the binding of ammonia to Mn4CaO5, water with similar structural and electrical properties was suggested to be supplied to Mn4 also through the O4 channel during S2S3 transition [100,109,110]. Especially in spinach, it is expected that the O4 channel is able to deliver water efficiently to the manganese complex. This is because the entrance to the channel is wider due to the presence of Ala in D1 at position 87 than in cyanobacteria, where Asn is found [111,112]. The competing hypothesis is that during the transition from S2 to S3, a water molecule is introduced into Mn4CaO5 through the O1 channel on the Ca side [40,42,44,89,109,113], and it could be a precursor of the O6 oxygen occurring in the manganese cluster [36]. This channel has been shown to exhibit the highest water exchange rate [89]. It has also been proposed that in cyanobacteria, the O1 channel may supply water, and both O1 and Cl1 channels release protons in S2S3 and S3 → (S4) → S0 transitions. But during S0S1 transition, the O4 channel has been suggested to be the proton exit pathway, which was found to be disconnected in the S2 state and restored only in the S0 state [42]. Recently, an interesting observation was made. Namely, glycerol (commonly used as a cryoprotectant and stabilizer of isolated PSII) incorporated into the O4 channel in cyanobacteria at a distance > 10 Å from the OEC affects the LS stabilization of the S2 state of the Mn4CaO5 complex, which adopts the ‘open’ conformation (Figure 1b) as a result of disruption to the hydrogen bond network involving D1-D61 when it remains protonated. In the absence of glycerol (D1–61 becomes deprotonated), both states of S2, i.e., LS and HS (open and closed conformations), are virtually isoenergetic [114]. An analogous effect of regulation of the Mn4CaO5 complex by allosteric interactions may also occur in higher plants, as may be indicated by the usual occurrence in their case of both spin states of the S2 state and the disappearance of the HS signal with an increase in the concentration of glycerol [115]. A diagram of the distribution of water channels identified in a higher plant (spinach), with an indication of the location of conserved amino acids, is shown in Figure 2.
The search for water, proton, and O2 transport pathways has been and continues to be carried out through the use of in silico experiments with various computational methods, including QM (quantum mechanics)/MM (molecular mechanics), MD (molecular dynamics), and CE (continuum electrostatics)/MC (Monte Carlo studies) using available PSII structures, even the PSII thylakoid membrane model [116]. Various research groups have pointed out similar patterns of water channels in cyanobacteria, but their purpose cannot always be clearly determined. Some counterparts have been found in higher plants, but even minor differences from cyanobacteria may be relevant to their distinct control of protons and O2 output during water oxidation or water delivery to the OEC. Identifying these mechanisms is a challenge. Learning about them is the key to achieving a complete picture of how OEC and PSII function as a whole. In general, water diffusion tends to require water-filled channels. Each water channel can potentially evacuate O2, but its diffusion can also occur through hydrophobic pathways. It has been suggested that lipid clusters within PSII, due to their predominantly hydrophobic nature, may serve as an oxygen drain and mediate efficient, PSII-safe, and rapid release of O2 [94,117]. A highly conserved small hydrophobic pathway in cyanobacteria and algae (prokaryotes and eukaryotes) has been identified at the beginning of the O1 channel and has been suggested to be responsible for facilitating O2 release from the OEC [117,118].
As can be seen, the factors regulating the process of water evolution by PSII are extremely complex. To understand them, it is necessary to consider not only the conformation of the OEC and its immediate environment but, most likely, the entire PSII, which is a dynamic system [36,79].
In this study, we focused on investigating the effects of the external proteins PsbO (~33 kDa), PsbP (~23 kDa), and PsbQ (~17 kDa) on the process of oxygen evolution in a higher plant, tobacco. Although the role of these proteins is not yet fully understood, it is known that they play a protective role in stabilizing the binding site of the Mn4CaO5 complex, ensuring ionic balance in its environment, mainly preventing the loss of calcium and chlorine ions [119,120,121,122,123,124]. In addition, as outlined above, current knowledge suggests that they may be involved in regulating the access of water molecules to the OEC and the removal of O2 from it [65,89,90].
In this work, we utilize a fast polarographic measurement technique that enables direct measurement of the kinetics of photosynthetic oxygen evolution without relying on models or assumptions about the initial conditions of the systems under investigation. The primary objectives of this approach are as follows: (i) to examine whether O2 evolution is a heterogeneous process, as postulated by previous analyses of the oscillatory pattern of O2 release during water oxidation in PSII influenced by short saturating flashes [37,38,125] and (ii) to investigate the impact of external proteins on the kinetics and predicted biphasicity of O2 yield in order to gain a better understanding of the factors that govern the efficiency and dynamics of oxygen evolution.
By addressing these questions, the study aims to contribute to the understanding of the complex mechanisms involved in photosynthetic oxygen evolution and the role of external proteins in this process. Understanding the various factors that contribute to the exceptional efficiency of the OEC in water splitting has significant implications beyond the field of photosynthesis research. This knowledge is also important in the design of high-efficiency fuel cells.

2. Materials and Methods

The procedures used for the isolation of PSII-enriched thylakoids, PSII BBY, and subsequent removal of extrinsic proteins were carried out according to [126,127] with some modifications [128]. PSII BBY was isolated from freshly harvested laboratory-grown tobacco, Nicotiana tabacum var. John William’s Broadleaf (JWB, from Prof. G.H.Schmid’s seed collection) under control conditions (25 °C; humidity, 65%; 14 h day/10 h night; white light—500 µE; red light—20 µE; infrared light; collected leaves from plants 3 months old) in PSI FytoScope growth chamber (PSI Photon System Instruments, Drasov, Czech Republic). The culture was carried out in summer. In the original procedure, two consecutive Triton washes were used to obtain a sample with high PSII enrichment. We wanted to maintain a sufficiently large natural pool of plastoquinone in the sample to avoid the need to add external acceptors. For this reason, the second Triton wash series was not carried out. A Triton X-100-to-chlorophyll (Chl) ratio 20:1 was used. The sample was washed four times with a large volume of Hepes II buffer to remove any residual Triton, which could lead to further degradation of the sample during storage. The washing procedure was continued until the supernatant was completely transparent. The PSII membrane (2 mg Chl/mL) was suspended in Hepes II buffer (15 mM NaCl, 5 mM MgCl2, 20 mM Hepes, and 400 mM sucrose; pH 6.5). The two extrinsic proteins, PsbQ (~17 kDa) and PsbP (~23 kDa) were removed from PS II BBY via incubation in a buffer containing 1.5 M NaCl, 400 mM sucrose, 40 mM Hepes, and 5 mM MgCl2 (pH 6.5) for 30 min at 0 °C (kept on ice). To remove three extrinsic proteins, PsbQ, PsbP, and PsbO (~33 kDa) instead of NaCl 1.5 M MgCl2 was used. The samples contained 1 mg Chl/mL. After treatment with the high concentrations of sodium and magnesium salts and two washings in Hepes II, the PSII membranes were resuspended in the same medium. Our experience [128], as well as that of other groups [for example, [127]], shows that 1.5 M MgCl2 and CaCl2 are similarly efficient in removing the three extrinsic proteins from PSII, but CaCl2 induces a much greater inhibition of oxygen evolution than MgCl2. Although both the CaCl2 and MgCl2 washes do not remove Mn from the OEC, the oxygen evolution in the samples treated with high concentrations of CaCl2 shows more than twofold greater decrease in the activity of PSII in the oxygen evolution than in the case of MgCl2. As shown by the data presented in [127,128], the use of CaCl2 would make it virtually impossible to perform a similar analysis of oxygen evolution. Since the samples showed the expected activity, the active site had to contain bound Mn ions. There was no need to add additional Ca2+ ions to the buffer as we were not using chelators. Furthermore, these would induce additional OEC modifications, which we wanted to avoid.
All of the prepared samples were divided into small portions and frozen in liquid nitrogen and stored at −80 °C until measurement (no longer than three months).
Samples lacking the two outer proteins PsbP and PsbQ and the three outer proteins PsbO, PsbP, and PsbQ are referred to in the paper as PSII BBY—P,Q and PSII BBY—O,P,Q, respectively.
To confirm protein removal, denaturing SDS-PAGE was conducted using a vertical polyacrylamide gel system in a Mini-Protean Tetra Cell apparatus (BioRad, Hercules, CA, USA) according to the standard protocol [127], as shown in Figure S6 in the Supplementary Materials.

Fast Polarography Experiments

Amperometric measurements of oxygen evolution in PSII BBY untreated and PSII BBY—P,Q and PSII BBY—O,P,Q were performed under short saturating flashes. We did not use any external acceptors because our experience and other groups (see, e.g., [39]) show that different external acceptors modify the O2 yield pattern in different ways. We also did not activate the sample with a single flash. This also affects the distribution of the Si states and other initial conditions of the samples. The assumption that only the S1 state is occupied is not necessarily true. Its stabilization also depends on other factors affecting the donor and acceptor sides of PSII in the dark (see Section 4). Furthermore, one had to assume that the transition from S0 to S1 must occur with probability 1, i.e., there are no misses, i.e., α0 = 0. Secondly, the sample exposed to a saturating flash of light changes the system, and even assuming the validity of the assumptions made, this certainly cannot be applied to the BBY PSII—O,P,Q sample, which means that the results obtained could not be compared for all samples.
A three-electrode system (a Joliot-type electrode, descrbed in [129,130]) was used. The polarization voltage was set at −680 mV. A xenon lamp (X-strobe. Perkin Elmer, Salem, MA, USA) emitted flashes of ~3 μs half-width. According to the manufacturer, the spectral bandwidth of the flash lamp is 250–1100+ nm, with infrared radiation accounting for 16% of the light output for wavelengths between 800 nm and 1100 nm in the total spectrum of the xenon lamp. The lamp is equipped with an optical fibre with a 13 mm diameter. It is placed at a distance not exceeding 1 cm from the electrode surface. The approximate light output from the fibre optic light guide is about 150 mJ per flash per cm2 (~2 J input energy per flash). With a realistic assumption of 70% luminous flux at the surface of our electrode and an average conversion factor for daylight, one may estimate that this gives about 500 µmole/cm2 per flash (about 98% accumulated in the peak of the flash). We verified that changes in the distance of the optical fibre from the surface within 1 cm while maintaining the other measurement parameters had no effect on the observed oscillations of the O2 yield under these short flashes. The average O2 evolution pattern, determined from three independent measurements, was maintained for each of these distances. This means that there was no change in the distribution of the initial Si states and the miss parameters. Thus, under our experimental conditions, we may assume that the light is saturating. The surface of our electrode on which the sample is spread is about 1.13 cm2 (its external diameter is about 1.2 cm). The sample is evenly distributed on the electrode surface, forming a very thin layer after sedimentation on a special Millipore filter of the same size as the electrode. Thus, under these experimental conditions, we may assume that the light is saturating. The volume of buffer (Hepes I) above the sample was about 600 µL.
The experimental protocol shown in Figure 3 was designed to directly observe the heterogeneity of oxygen evolution in PSII. Two independent experiments were performed in which (i) the intervals between flashes 1 and 2 were varied (as shown in Figure 3—upper scheme), and the remaining 14 flashes were delivered at 300 m intervals, or (ii) the intervals between flashes 2 and 3 were varied (as shown in Figure 3—lower scheme), and the remaining 14 flashes were delivered at 300 m intervals. To illustrate the course of the experiment, Figure 4 shows example raw data for PSII BBYcontrol for Δt2–3 = 120 m. It is the simplest and most natural approach to monitor kinetics of oxygen evolution under the third flash (Y3) as a function of Δt1–2 or Δt2–3. The idea was to check if the observed kinetics for these two protocols were different or the same. The first scenario would imply a bifurcation of the oxygen evolution process between these transitions, while the second one that linearity is maintained between the transitions, and that the classic Kok’s model can be applied.
Each measurement for a given fixed period Δt1–2 or Δt2–3 (varied from 5 m to 500 m) was performed on a freshly applied sample, previously incubated in the dark, as described above. In a control protocol typically used as a standard, all flashes were 300 m apart. The final volume of samples containing about 30 µg Chl suspended in a Hepes I buffer (15 mM NaCl, 5 mM MgCl2, 20 mM Hepes, pH 6.5) was 500 μL. Thus, the concentration of chlorophyll in the polarographically measured samples was 60 µg/mL. After 7 min of dark incubation on ice (2 min) and sedimentation on the filter at room temperature (5 min), the samples were transferred to the electrode in dim light conditions and were incubated for another 10 min in the dark. This time was sufficient for the electrode signal to stabilize at a constant level. A thawed portion of the sample was used within approximately 1.5 h. Once the sample was thawed, it was stored on ice in the dark. The activity and stability of the sample were monitored for a standard period of 300 m between all flashes at the beginning, in the middle, and at the end of the measurement series. For a given series of measurements, variable periods of Δt1–2 or Δt2–3 were randomly selected. The amplitude of oxygen evolution Y3 was normalized to the sum of all amplitudes (Ys) in each measurement, and then these values were normalized to the mean value of the ratio Y3(300 m)/Ys (300 m) obtained for the standard protocol (i.e., Δt1–2 or Δt2–3 of 300 m) in a given series of measurements. Finally, all of the normalized results from a given series of measurements were scaled to the average of all individual measurements for a given Δt1–2 or Δt2–3 (normalized as described above) for a range of variable time periods from about 200 m to 500 m in order to show all of the data in one figure, as shown in the Figure in Section 3.2.
Thus, the approach to obtain information on the kinetics of oxygen evolution based on the variability of the periods between the initial flashes is different from the protocols used to measure the lifetimes of the Si states and the method of their determination, which depends on the model adopted [131,132]. The variable period in these experiments usually ranges from 0.5 s to a few hundred seconds and concerns the dark adaptation between a short series of pre-flashes and the actual sequence of a dozen flashes.
The proposed method of measuring oxygen evolution time using fast polarography has significant advantages over the shape analysis of the electrode response signal. Analysis of the polarographic signal is complex and does not provide direct information on the biphasic nature of O2 evolution in PSII. Furthermore, as shown in [133], the signal recorded on the electrode is strongly dependent on the contact of the sample with the electrode or the thickness of the sample, apart from the voltage applied to the electrode. However, the porosity of the sample and/or its external potential, which may change due to sample modification (e.g., protein removal), will also affect the shape of the signal.
Typical shapes of the polarographic signals obtained for the PSII BBY control, PSII BBY—P,Q and PSII BBY—O,P,Q under the same experimental conditions are shown in Figure S5 in the Supplementary Materials.

3. Results

3.1. Analyzing Oxygen Evolution Patterns

In the standard procedure for the measurement of oxygen release under short, saturating flashes using a fast three-electrode system, the time interval between the flashes is 300 ms, which provides optimal experimental conditions. Experimental and theoretical data of the flash-induced oxygen evolution in native and two or three extrinsic-protein-depleted PSII BBY are shown in Figure 5 (example raw data are presented in Figure S1 in the Supplementary Materials). No external acceptors were used in this study to avoid altering the functioning of the active PSII complexes [39]. Therefore, the total oscillatory oxygen release signal slightly decreases with an increasing number of flashes, as one may expect [37,39,132]. This is due to the partial depletion of a natural acceptor (plastoquinone PQ-9), which reduces the number of active photosystems. Typically, the mobile PQ pool is estimated to be 5–10 molecules per PSII RC in thylakoids and does not change significantly in isolated membranes due to the lipophilicity of quinone but is not uniformly distributed [134,135,136]. In our case, this is determined by the C parameter, which is 1 for a fully functional system with a sufficient abundance of the acceptor, which means that the number of active centers is maintained (i.e.,   i = 0 4 S i = 1 , where Si is a fraction of the OEC in a specific oxidized state i, Markov chain). We assumed that the decay of the signal follows a geometric progression with the ratio C, which is slightly less than one (see Equation (S4) in Supplementary Materials). The quenching of the total signal is stronger in PSII BBY—P,Q than in the control sample (Figure 5) because the removal of two external proteins (PsbP and PsbQ) reduced the available plastoquinone pool. In this case, the C parameter has the lowest value (Table 1). However, the activity of PSII BBY—P,Q was only reduced by about 10% compared to intact PSII BBY (Figures S2 and S5a in the Supplementary Materials). In contrast, after the removal of three external proteins (PsbO, PsbP, and PsbQ), when the number of active centers is significantly reduced by about 60% (see Figures S2 and S5a in the Supplementary Materials), the effect of signal quenching due to PQ deficiency is much less pronounced (Table 1 and Figure 5).
The failure rate α of the trapping centers (called misses) leads to a redistribution of the Si states and, consequently, to a damping of the oscillations of the O2 release. The original Kok model assumed equal misses for light-driven transitions SiSi+1 and additionally doubled effective excitation in a fraction γ of the centers, which are in the S0 and S1 states (called double hits and also equal). However, it has been shown that when this homogeneous model is used, there are significant discrepancies between the theoretical and experimental O2 yield patterns [137,138]. Double hits were introduced to explain the often-significant oxygen release that occurs at the second flash. In addition, double hits lead to an increase in oscillation damping at higher flash numbers. This further reduces the observed differences between experimental and theoretical data but still does not allow for the satisfactory reproduction of the observed O2 evolution oscillations. In the Supplementary Materials, Figure S3 shows fits using the Kok model with equal misses or equal misses and double hits to our experimental data of the samples studied here, i.e., PSII BBY control, PSII BBY—P,Q, and PSII BBY—O,P,Q. The fitted parameters are given in Table S1. Various approaches to the modification of the Kok model for a better representation of the experimental data are mentioned in the Supplementary Materials. However, it has been demonstrated that the progressive damping of the oscillations is mainly due to misses, and the heterogeneous model with different αi misses, omitting double hits, gives a better quantitative agreement with the experimental data obtained for other systems [37,139,140,141,142,143]. Unequal misses for SiSi+1 transitions for i = 0, 1, 2, 3, 4 were also experimentally proven via EPR measurements [144]. Furthermore, it was found that extending the Kok model to a 5S-state model by explicitly including the S4 state and introducing biphasicity in the O2 evolution by introducing a bifurcation of the S3 → (S4/S4’)S0 transition (Figure 1c) resulted in an excellent reproduction of the experimentally observed patterns of O2 release [37].
Figure 5. Flash-induced oxygen yield pattern in intact PSII BBY (A) and PSII BBY depleted of the extrinsic proteins via NaCl (B) and MgCl2 (C) washing. Samples were suspended in the Hepes I buffer pH 6.5. The flashes were 300 m apart. Amplitudes are always normalized to the amplitude of O2-evolution under the third flash. The experimental data are the mean values obtained from 3 independent measurements in each case. The error bars correspond to the maximum error.
Figure 5. Flash-induced oxygen yield pattern in intact PSII BBY (A) and PSII BBY depleted of the extrinsic proteins via NaCl (B) and MgCl2 (C) washing. Samples were suspended in the Hepes I buffer pH 6.5. The flashes were 300 m apart. Amplitudes are always normalized to the amplitude of O2-evolution under the third flash. The experimental data are the mean values obtained from 3 independent measurements in each case. The error bars correspond to the maximum error.
Cimb 46 00428 g005
The distinction between the short-lived and the metastable S4 state allowed us to estimate the contribution of PSII involved in the fast (d) and slow (1 − d) phases of oxygen evolution; for example, see the scheme in Figure 1c and Section 4. Due to its simplicity, the proposed model allows all parameters αi, parameter d, and the initial distribution of Si states, to be fitted to the experimental data without any assumptions. This results in higher accuracy and precision than the classical Kok models (see Figure S3 and Table S1 in the Supplementary Materials). Note that S4 = 0, and S3 = 1 − S0S1S2 leaves eight free parameters to be fitted. However, the initial conditions are sufficiently well defined. The first four amplitudes are a good approximation of the initial Si states. Consequently, in our model, the minimum of the fit quality as a function of the parameters used is stable. The sensitivity of the heterogeneous model to αi parameters is presented and discussed in the Supplementary Materials (Figure S4).
The results of the theoretical evaluation of the experimental data for the control sample, PSII BBY—P,Q and PSII BBY—O,P,Q obtained using the 5S-state heterogeneous model are shown in Table 1. It is worth noting that the same trends in changes in initial Si state occupancy and misses are also obtained for the control sample and samples lacking two or three external proteins using the Kok model with equal misses and equal misses and double hits (see the Supplementary Materials, Table S1, Figure S3). As can be seen, these homogenous models do not allow for the satisfactory reconstruction of the oxygen evolution oscillations, especially in the case of the control sample and PSII BBY—P,Q. The quality of the fit compared to the extended 5S-state model is about 6–13 times worse in the first case and about 2–5 times worse in the second case. Due to the low signal intensity and increased oscillation damping observed for PSII BBY—O,P,Q, the fit quality parameter is only about 1.5 times worse than with the extended 5S-state model. The use of the heterogeneous 5S state model allows for a more detailed analysis of the causes of the decrease in the probability of transitions between the SiSi+1 states.

3.2. Kinetics of the Fast and Slow O2 Release Pathways

To verify the heterogeneity (biphasicity) of the oxygen release process in direct measurement, we used a protocol for measuring oxygen release using a fast electrode type, as presented in this paper (Figure 3 and Figure 4). Two protocol variations were used, one with different periods between the first and second flashes (Δt1–2) and the other with varying periods between the second and third flashes (Δt2–3).
Figure 6 shows the dependence of the changes in the oxygen evolution after the third flash as a function of the flash interval when Δt1–2 (red symbols) and when Δt2–3 (black symbols) changed. Note that for a given sample, whether the intervals between the first and second flashes or between the second and third flashes were altered, the dependence of the changes in oxygen evolution after the third flash on the flash interval is the same. This is in accordance with Kok’s linear sequence of transitions between the Si states.
Figure 6. The dependence of O2 release under the third flash on the time interval between the first and second flashes, Δt1–2, (red symbols) and the second and third flashes, Δt2–3 (black symbols) for control PSII BBY (a), PSII BBY—P,Q (b), and PSII BBY—O,P,Q (c). Each experimental point is a mean of at least 3 independent measurements. Error bars represent the root mean square error. The theoretical curve shows the fit obtained using a biexponential function (Equation (1)).
Figure 6. The dependence of O2 release under the third flash on the time interval between the first and second flashes, Δt1–2, (red symbols) and the second and third flashes, Δt2–3 (black symbols) for control PSII BBY (a), PSII BBY—P,Q (b), and PSII BBY—O,P,Q (c). Each experimental point is a mean of at least 3 independent measurements. Error bars represent the root mean square error. The theoretical curve shows the fit obtained using a biexponential function (Equation (1)).
Cimb 46 00428 g006
The experimental data were fitted using a dual relaxation function, defined as the sum of two exponential relaxations (solid lines in Figure 6a–c):
Y 3 = A f a s t ( 1 exp ( t τ f a s t ) ) + A s l o w ( 1 exp ( t τ s l o w ) ,  
where Afast (Aslow) and τfast (τslow) are the amplitudes and time constants characterizing the fast (slow) O2 release, respectively. For the normalized data, Afast and Aslow denote the contributions of the individual phases. The parameters Afast, Aslow, τfast, and τslow obtained from the data for control PSII BBY and PSII depleted of two or three external proteins are listed in Table 2.
In the case of the control sample and PSII BBY—P,Q, two components were necessary to obtain a satisfactory fit to the experimental data, whereas in the case of PSII BBY—O,P,Q, one component was sufficient. We checked that the behavior of the experimental data obtained for samples depleted of extrinsic proteins could not be satisfactorily reproduced by assuming two phases with time constants as for the control sample (Figure S7 and Discussion in the Supplementary Materials). The contribution of the fast component shows excellent agreement with the values and direction of change in the parameter d, corresponding to the fast phase of O2 release, obtained from the analysis of the O2 yield pattern using the 5S-state model.
Applying this experimental approach, it was also possible to determine the time constants of the fast and slow pathways of oxygen release in the samples. For the control sample, the time constant of oxygen release in the slow (~44 m) and fast (~4 m) phases differs by an order of magnitude. In the case of PSII BBY—P,Q, the difference is only about 3.5 times. This is because the fast phase is slowed down by about 2 m, while the slow phase is twice as fast. The removal of all three proteins resulted in a further acceleration of the slow phase by a factor of about 1.7 (Table 2).

4. Discussion

4.1. Analyzing Oxygen Evolution Patterns

In the control sample, we observed the highest initial occupancy of the S1 state (~87%) and comparable occupancies of the S0 and S2 states at about 5% and 8%, respectively. The deletion of the PsbQ and PsbP proteins did not affect the initial distribution of the Si states. It is known that during prolonged darkness, the OEC is mainly in the S1 state [145]. This is due to the oxidation of the S0 state by an electron carrier, TyrD+ (tyrosine Tyr160 of peptide D2, YD in Figure 1) [7,32,61,93,146,147,148]. Moreover, it has been proposed that S1 is stable in the dark because the oxidation of Mn3(III) to Mn3(IV) forces the deprotonation of a μ-hydroxo group at the O4 position in the Mn4CaO5 cluster, and the proton is transferred along the O4 water channel up to ~13.5 Å from O4 [53]. So, the S1 state does not return to the ground state of S0 in the usual experimental time of a few minutes. Consequently, the occupancy of the S1 state in darkness is dominant, and therefore, the first maximum is observed under the third flash. A low occupancy of the S2 state, i.e., a low O2 yield under the second flash, is observed in a number of samples. Even shortening the duration of the flash at mid-maximum intensity from µs to ns did not result in the disappearance of oxygen release under the influence of the second flash [149,150]. For a long time, it was thought that the S2 state could not be stable after long dark incubation and that only the S0 and S1 states would be stable in the dark. Moreover, reduced TyrD, accumulated in the dark due to its involvement in the oxidation of the S0-state, S0TyrDoxdS1TyrDred, has also been shown to be able to reduce S2 and S3 states [151,152]. However, the efficiency of electron transfer from the reduced TyrD to the S2/S3 states depends on the protonation in the vicinity of the TyrD and the OEC. It was observed that the oxidized TyrD present in the S1 centers showed high stability, although its slow reduction was detected in the dark, most likely via S1 TyrDoxdS2 TyrDred transition [147]. This would explain why, in the absence of electron donation from the acceptor side, a small population of the S2 state is always present in samples with long dark adaptation. This is our case. The higher states are expected to be unstable [37,153,154].
The additional removal of the PsbO protein significantly affected the initial distribution of the Si states. On the one hand, this decreased the stability of the S1 state in the dark and, on the other hand, increased the stability of the S2 and S3 states. This observation is consistent with the fact that PsbO is an already-known MSP protein (i.e., manganese complex stabilizing protein). Most probably, it is responsible for stabilizing not only the Mn4CaO5 complex but also the surroundings of the cluster, including TyrZ and TyrD and the entire hydrogen network in their vicinity [155]. The effect of increased stability of the higher states caused by the PSII depletion of PsbO has also been observed by other groups [156,157] but has not been discussed. We suspect that these changes are related to increased uncontrolled water molecule access to the OEC and modification of the surrounding hydrogen network.
In PSII BBY—P,Q, a significant decrease in transition efficiency between the Si states was observed (Table 1 and Table S1). There was an increase in the parameter αt (total miss: α t = i = 0 3 α i ) mainly due to an increase in α0 and α1, while the parameter α2 remained unchanged. The value of parameter α3 also increased, indicating that about 88% of the S3 states efficiently transitioned to S4 states. The probability of this transition in the control sample was nearly 100% under our experimental conditions. At the same time, the contribution of the fast mode to the O2 yield decreased (parameter d decreased almost threefold compared to the control sample). The removal of all three external proteins did not affect the transition probability between SiSi+1 states compared to PSII BBY—P,Q. However, the removal of the additional PsbO protein affected the initial distribution of the Si states and further reduced the d parameter to a value of about 0.02, indicating almost complete inactivation of the fast O2 release channel.
It is noteworthy that in the case of the control sample, the misses came almost exclusively from α2, which maintained its value in samples lacking outer proteins of two or three, giving the highest contribution to αt. The lowest efficiency of the transition S2S3 is consistent with the unique character of this transition, confirmed via various measurements sensitive to the reorganization and charge transfer changes during this step of the cyclic transformation of the OEC [44,71,78,89]. EXAFS experiments implied significant structural changes during S2 to S3 transition, observing the Mn-Mn and Mn-Ca distances [48,158]. The significant value of the α2 miss parameter may also reflect the structural changes observed around TyrD under two-flash illumination. This is due to the partial oxidation of TyrD resulting from inefficient electron donation from the Mn4CaO5 complex, as suggested in [71]. In addition, FTIR measurements showed that the binding of one of the substrate water molecules to the Mn4CaO5 complex occurs during S2S3 transition and the other during the S3 → (S4)→ S0 transition [36,50,159]. In addition, it has been suggested that releasing a proton during S2S3 transition could be a rate-limiting step in this transition [108]. Mn4CaO5 binds four water molecules. Two water molecules, called W1 and W2, are ligated to Mn4, and two others, W3 and W4, to Ca (Figure 1b). During S2 to S3 transition, additional bridging oxygen (O6) ligated to Mn1 near O5 was observed [42,44,71,78]. The sixth oxygen bound to Mn1, presumably as an OH group, may originate from substrate water directly bound to Mn1 earlier in S1S2 transition [48] or from a water molecule already present in the S1 state, proposed to be a Ca (W3) [113,160,161] or Mn4 (W1) ligand [110,162,163]. In the latter case, when an additional water molecule is bound to the Ca or Mn4 ion (which requires Mn4(IV) in the S2 state) during S2S3 transition, one of the internal water molecules (W1 or W3) moves to a new position, i.e., O6 [76,164,165]. However, recently, a completely new scenario has been proposed based on TR-SFX experiments, which can be used to follow PSII structural dynamics in the ns to ms timescale [36]. The O6 ligand may have its origin in a water molecule from the outer sphere of the water lattice. It binds to the calcium ion in less than a microsecond during S2S3 transition. However, this is preceded by a significant reorganization of the Mn4CaO5 cluster and the surrounding proton network, including water molecules.
It is currently not possible to determine whether the heterogeneity in the structures of the Si states, described in the Introduction, indicates multiple pathways for O-O bond formation, or rather a single pathway within a given reaction center that is optimized during the cycle, leading to O2 release [25,43,48,76,77,80,166,167]. In the latter case, the different conformations of the Si would be transition states that occur in a single chain of events in the Kok cycle. For example, the topology of a high-spin closed cubane in the S2 state (formed, e.g., when H+ is released from the S2Y state) was proposed to be essential for the transition to the HS S3 (closed cubane) state, which changes to the low spin LS S3 (open cubane) state on the addition of H2O [35,168]. It is often suggested that the emergence of a high-spin closed cubane topology in the S2 state is of mechanistic importance for the subsequent catalytic steps [45,163]. However, the specific steps leading to the formation of O-O and the release of O2 may be much more complicated or different from those mentioned above. For example, the observed HS S2 may not result from conformational changes in the Mn4CaO5 complex but from the protonation of O4 in the open cubane form [169]. Proton isomerism of the S2 state has been independently suggested in [161]. Proton isomerism of high- and low-spin S3 states has also been proposed [80,170].
Furthermore, the question arises as to the causal relationship between the changes in the manganese complex itself and the protein network of its proximal and distal environment. This point is crucial because of the need to synchronize the delivery of water, the reception of protons, and the release of O2 with the cycling of the Mn4CaO5 complex. Obtaining a complete picture of the process of OEC functioning is hampered because of, among other things, the insufficient resolution of PSII structures, the blurriness of the images obtained from them, the difficulty of capturing the various stages of the Kok cycle (mixing of Si states), the influence of the measurement conditions on the stability of the sample and the oxidation states of the manganese cluster or the preparation method on the functioning of PSII, the degree of its hydration, and simplified theoretical models applied [36,77,171,172,173,174,175,176,177,178].
Based on the existing knowledge, one may suggest the possible causes of the changes in the αi parameters due to the extraction of the two outer proteins, PsbP and PsbQ. As a first approximation, the removal of these two external proteins should primarily be attributed to proton channel dysfunction. This assumes that the S0S1, S1S2, S2S3 and S3 → (S4) → S0 transitions primarily require efficient proton extraction to minimize ‘back reactions’, mainly related to electron flow. This will, of course, be accompanied by changes in the coupling efficiency within the Mn4CaO5 complex, as well as changes in its interaction with the immediate protein–water environment. They will certainly be reflected in the conformation and stability of the subsequent Si states resulting from these transitions. Referring to the proposed contribution of PsbP in stabilizing O1B and O4 channel outlets and PsbQ in stabilizing the O1A channel outlet, recognizing in higher plants [90,101] (see Figure 2), the observed increase in miss parameters very well reflects the changes in the activity of these channels. Thus, the lack of regulation of proton uptake at the output of the O4 water channel, which has been indicated to be the main, if not the only, proton transfer channel for the S0S1 transition [42,53,61,104], translates into a high increase in α0. The dysfunction of both the O4 channel and the two branches of the O1 channel leads to an increase in α1 and α3 but to a much lesser extent. For S1S2 and S3S0 transitions, α1 and α3 are approximately 2 and more than 4 times lower than α0, respectively. This may indicate a smaller contribution of the O4 channel to deprotonation processes during these transitions, especially in terms of S3S0 transition, and a larger contribution of the other channels. It has been suggested that the O4 channel only opens during S0S1 transition [42]. It is also likely that the smaller changes in O1 channel function induced by the deletion of the PsbQ and PsbP proteins are due to the specificity of this channel, which is wider than the others. It contains highly mobile water molecules along almost its entire length and may function primarily as a water supply channel [89]. This has recently been experimentally demonstrated for water binding during S2S3 transition [36]. Since the extraction of all three outer proteins, i.e., PsbQ, PsbP, and PsbO, did not further alter the parameters α0, α1, and α3, it can be expected that mainly the Cl1A branch (Figure 2) is responsible for the reception of H+ during S1S2 and S3S0 transitions. It is also the most evolutionarily conserved water channel, also in higher plants [179,180,181,182]. It contains the very important conserved amino acid sequences E65/E312/R334, which are suggested to be gates that regulate the release of protons into the lumen [183,184] as well as being possibly involved in the uptake of water (see Figure 2) [100,185]. Because no proton release to the bulk is observed during S1S2 transition [51,186], it is proposed that, in this case, the proton is temporarily stored in the form of hydronium in the hydrogen bond network in the vicinity of Ca [64,172] or in the Cl1 channel to which it can be transferred via D61 [182]. For example, Asp 170 has been suggested to be involved [69] in this process, but other deprotonation pathways of Mn4CaO5 cannot be excluded. Indeed, the formation of a cationic water cluster was observed, which indicates transient storage of the proton in the S2 state in the form of the nH2O(H3O)+ cluster, where n = 5 [65]. It should be noted here that the proton network of the Cl1 channel, which connects to the O4 channel in the vicinity of O4 of the manganese cluster, extends through a network of hydrogen bonds involving unbound and bound water molecules on the Mn4CaO5 cluster and neighboring amino acids up to TyrZ [27,42,91]. On the opposite side of the manganese complex, close to Ca, water molecules can penetrate from the O1 channel [42] (Figure 2).
The highest value of the miss parameter assigned to α2, regardless of the presence of external proteins, is related to the specific nature of S2S3 transition. As mentioned above and shown in recent experiments [36,40,41], this transition requires large rearrangements of the Mn4CaO5 cluster associated with substrate water binding. Water availability and binding it to a specific site for the required conformation of the complex may be the main reason for the low efficiency of this transition [185]. In particular, the transport of a water molecule from the water network of the O1 channel to the Mn4CaO5 cluster, which requires a coordinated reorganization of the hydrogen network beyond the first or even the second coordination sphere of the manganese complex and results in the attachment of an additional water molecule to the calcium ion, as proposed in [36], could be a critical step in the oxygen evolution process. Thus, mere deprotonation of Ca-bound H2O involving the Cl1 channel, as proposed in [36], would not be the least efficient step for the S2S3 transition [43,44,109,187,188,189]. The independence of S2S3 transition from the presence of PsbQ, PsbP, and PsbO may also suggest that Cl1A may act as a water transport channel during this transition. This is consistent with previous predictions [184]. On the other hand, if access to substrate water is not a bottleneck for this transition, then perhaps the assumption of a mechanism (‘carousel’ or ‘pivot’) implying that the O6 position is occupied by one of the internal waters already present in the manganese cluster is still possible. Then, the high stabilization of the S2 transient state before the electron flow to TyrZ and/or the instability of the S3 state would explain the low efficiency of this step of the Kok cycle. The much higher probability of the S3 → (S4) → S0 transition than in the previous step suggests that, in this case, the problem is not water access but deprotonation and probably O-O bond formation, the initial stage of which begins with the formation of the S3 state [36,40,41], i.e., during S2S3. This could mean that during the formation of the S0 state, the water molecule is already attached in the immediate vicinity of the Mn4CaO5 cluster.
The 5S-state model employed in this study has provided further insights into the effects of external proteins on the heterogeneity of the oxygen evolution process. Specifically, the decrease in the d parameter observed in samples without external proteins suggests a diminished contribution of the fast oxygen release pathway. Remarkably, the removal of PsbQ and PsbP resulted in a nearly threefold reduction in the contribution of the fast pathway to O2 evolution. Furthermore, in PSII samples lacking all three external proteins (PsbQ, PsbP, and PsbO), the oxygen evolution process occurred almost exclusively through the slow pathway, represented by the S4’ state in the model.

4.2. Kinetics of the Fast and Slow O2 Release Pathways

Oxygen release times obtained using the fast polarographic method are typically in the range of 0.8 ms to 4 ms, depending on the measurement technique and conditions, type of the sample, and how the O2 release times are determined [190,191,192,193,194,195,196]. The lower limit of oxygen release, determined experimentally using EPR oximetry measurements and obtained via direct analysis of the signal after the third flash from the fast electrode, is approximately 500 µs [133,197]. This limitation also results from the electron transition time between QA and QB, which varies from about 0.2 ms to 0.8 ms. It depends on the redox state of QB and the number of positive charges accumulated in the OEC [198,199,200,201,202]. It should be noted that regardless of the state of the acceptor side, at the start of OEC operation after the dark adaptation period, in order to achieve O2 release in the four-cycle action of each reaction center, the reduced and protonated quinone at the QB site must be replaced at least once by the oxidized quinone from the available external PQ pool. If the external pool of quinones was sufficient and the observed specific PQ diffusion pathway between the membrane and the QB-binding pocket would allow for the attachment of quinones [84,203], the acceptor side of PSII is not a bottleneck in the kinetics of the oxygen release process. An increase in misses can be expected if one takes into account the cyclic flow around PSII postulated by various research groups. Cytochrome b559 plays an important role in this flow, among other things by causing the reduction of P680+ through a network of Car and Chl cofactors (for review see: [204,205,206]). Cytochrome b559, reduced by the PSII acceptor site, has also been shown to be able to reduce TyrD and TyrZ as well as Si(i=2,3) states [207,208]. Recently, it has been postulated that the microstates (QAQB)2− can reduce the Y+S3 and Y+S2 states, and this is coupled to the deprotonation of the PQ bound at the QC site [209,210]. As a result, the efficiency of the transitions between the S2S3 and S3S4 states can be reduced due to YS2 ↔ Y+S1 and YS3 ↔ Y+S2. This should lead to an increased damping of the oscillations of O2 evolution. It could be hypothesized that the partial relaxation of the S2 and S3 states to lower states, affecting backward recombination involving oxidized P680 and/or the cyclic electron flow around PSII, are responsible for the biphasic effect of oxygen evolution that we observed. Back reactions are certainly responsible for the increase in misses. However, if the effects we observed were related to this mechanism, there would be a problem explaining why the slow phase of oxygen evolution accelerates as the miss values increase. In fact, the opposite effect would be expected. The effect of cyclic electron transfer around PSII on the observed two-phase oxygen evolution will certainly be investigated.
The approximately 10-fold slower time constant of the O2 signal observed here is comparable in magnitude to the oxygen release times observed for thylakoids isolated from Synechocystis sp. PCC6803 mutants: D1-V185N/T, D1-N181A/S and D1-D61N/A. In these mutants, the time after which the oxygen release signal reached a maximum varied between 20 and 33 ms at pH 6.5 [211,212,213]. It is important to note that these times are about 20 to 30 times longer than those obtained for the wild form of Synechocystis sp. PCC6803. Mutations of the hydrophobic D1-V185 and hydrophilic D1-N181 are located near water molecules situated between TyrZ and D1-D61 and disrupt the extensive hydrogen bonding network between water molecules in the Cl1 water chain [95]. D1-N181 interacts with the chloride ion via hydrogen bonding and is one of the closest residues for water W2 bound to Mn4 [27,213]. The residue D1-D61 is directly hydrogen-bonded to the W1 ligand of Mn4 and is thought to be involved in proton removal from the OEC to the lumen and/or water entry [51,91,100,114,214]. The D1-V185 residue, being about 3.7 Å from the Mn4CaO5 complex, faces the Cl1 channel, which interacts with TyrZ, D1-Asp170, D1-D61, and D2-K317. Therefore, together with the surrounding water cluster, it may be involved in tuning the efficient relaxation processes of H-bond networks and/or proteins in the vicinity of the OEC. In this way, it may influence the efficiency of electron transfer to TyrZ+ [212].
Both the theoretical approach to the analysis of the oxygen release sequence under the influence of short flashes, which takes into account the bifurcation of the path of the O2 release and the experimental approach, based on a time-dependent measurement of the increase in O2 yield under the influence of the third flash, allow us to conclude that the observed heterogeneity is due to the functioning of the OEC. The consistency of the observed changes in the proportions of fast and slow O2 release pathways for the different samples, control and PSII with external proteins removed allows us to relate the obtained times to the fast and slow release of the oxygen molecule and thus to the kinetics of the processes occurring on the donor side of PSII. The reduction in the contribution of the fast phase to oxygen release in the case of PSII BBY—P,Q indicates a destabilization of the proton network of water inside the O4 and O1 channels due to the destruction of their outputs towards the lumen. This indicates not only an impairment of proton transport through these channels but also a reduced control of water access to the Mn4CaO5 complex. The impairment of the proton uptake through these channels is also manifested by the hindered transitions between the S0S1, S1S2, and S3S0. The additional extraction of the PsbO protein had no further effect on the efficiency of the transitions between the SiSi+1 states. Still, it did result in the disappearance of the fast oxygen release phase. This means the transition between the S3 and S0 states in PSII BBY—O,P,Q occurs via the metastable S4 state. In this case, the only functional channel is Cl1A, as mentioned above, although some modifications of its functionality cannot be excluded. The PsbO protein has a region containing a hydrogen-bonding network near Cl-1 ion and a conserved residue R262 near the lumenal side, which can interact with the Mn4CaO5 cluster through a D1-D61-mediated hydrogen-bonding network extending to the PSII core proteins [215]. So, the organizational modification of D1 and D2, as shown on the donor side of PSII due to the extraction of external proteins, particularly PsbO, is possible in several ways. The absence of this protein destabilizes the Mn4CaO5 complex, affecting not only the organization of its protein ligands but also the entire proton network, to which water molecules and amino acids from the first and further coordination spheres contribute. The critical role of the hydrogen bonding network on the lumenal side of PSII, not only in the immediate vicinity of the OEC, was discussed in [92]. Furthermore, donor CP43 conformational changes also need to be considered [36]. The observed slowing of the fast phase in the sample lacking the PsbP and PsbQ proteins compared to the control is not surprising, as one would expect that a disruption of the coordinated deprotonation with water binding would lead to a slowing of the O2 evolution. Disabling some channels and undoubtedly reducing their efficiency in both transport and gating contribute significantly to this effect. The structural change in the Mn4CaO5 complex must also be considered. In contrast, the acceleration of oxygen release via the slow pathway may seem surprising. However, assuming that water is mainly supplied to the Mn4CaO5 complex through the O1 channel, the lack of gating of the water flow as a result of modification of the channel is the simplest explanation for the observed effect. Such a mechanism to control water access to the O1 channel with D1-E329 has been proposed for cyanobacteria [98]. The above hypothesis is highly plausible if one assumes that water binding to the manganese complex is an equally important mechanism regulating OEC function in addition to proton extraction. This thesis is supported by the further significant acceleration of oxygen release in PSII BBY—O,P,Q when the Mn4CaO5 complex has open access to the aqueous environment, and the aqueous network could provide an effective channel for the removal of oxygen molecules. On the other hand, if the hydrophobic matrix is responsible for the removal of O2, this would imply that the penetration of the oxygen molecule into the lipid layer is enhanced in PSII BBY—O,P,Q.
However, another completely different scenario can be outlined to explain the observed changes in the contribution and rate of these two O2 release pathways due to depletion of external proteins. EPR measurements showed that both LS and HS S2 states exist in higher plants [215]. This is attributed to the coexistence of two different isomers of the Mn4CaO5 complex (see Section 4 above). The depletion of the two outer proteins, PsbQ and PsbP, led to a significant decrease in the ratio of the HS signal to the LS signal for the S2 state. The depletion of all three outer proteins, i.e., including the PsbO protein, resulted in a complete loss of the HS signal for this state [215]. These results provide evidence for the regulatory role of the external proteins in stabilizing the HS S2 state. Thus, the fast transition pathway from S3 to S0 could be associated with the HS S2 state and the slow pathway, which requires consideration of the metastable S4 state, with the LS S2 state. We do not know how the following stages of the Kok cycle proceed. However, this direct correlation of the two S2 spin states with the respective fractions of the two oxygen release pathways suggests that the bifurcation of the mechanism leading to O-O bond formation and O2 release occurs at an earlier stage than the final transition S3 → (S4/S4’) → S0.
However, the following examples, based on the study of oxygen evolution in the above-mentioned mutants, illustrate how difficult it is to identify the main cause that determines whether PSII will release oxygen at a fast or slow rate.
Although D1-D61A/N and D1-V185N mutants affect OEC functioning differently, both lead to the slowing of oxygen release. It has been, therefore, suggested that in these mutants, the Mn4CaO5 complex adopts a structural rearrangement and/or tautomerism that allows a similar mechanism of O-O bond formation and oxygen release [211]. The D1-V185 residue was shown to be involved in stabilizing the S2 state of the Mn4CaO5 complex in the LS state, and a significant slowing of O2 release (t1/2 ≈ 20 ms) was observed in the D1-V185T mutant, in which the HS S2 state was dominant [212]. However, in an independent experiment, an O2 release rate (t1/2 ≈ 1.5 ms) similar to the wild type was observed for this mutant, albeit with a reduced oxygen production efficiency of approximately 40% [216]. In contrast to the V185T mutant [212], the Mn4CaO5 complex in the V185N, D61A, and D61N mutants, for which slow O2 evolution was observed, gives rise to a multiband S2 state signal that is qualitatively similar to the LS S2 state favoured in wild-type cyanobacteria at pH ~6 [217,218,219].
From the results of experiments performed on these cyanobacteria mutants, it can be concluded that the open (LS) or closed (HS) states of the manganese complex in S2 are not closely related to either the fast or slow oxygen release mechanism. Understanding the reasons for the different oxygen release times observed in the D1-V185T mutant by different research groups may provide valuable information about the underlying mechanism governing the rate of this process. Moreover, for the S2 and S3 states, there are potentially more possible configurations if differences in the hydration of the Mn4CaO5 cluster are also taken into account [33,34]. Nevertheless, it is unlikely that differences in the OEC due to the conformational heterogeneity of the S2 and/or S3 states themselves can explain the observed biphasic oxygen release. Thus, based on the results obtained for the mutants mentioned above, one may suggest that disruption of the hydrogen bond network in the immediate vicinity of the Mn4CaO5 complex slows down the release of oxygen as a result of delayed proton and/or electron transfer. In contrast, studies of the D1-N181A/S mutants have led to the conclusion that proton transfer is not impaired in their case and that the significant slowdown in oxygen yield is due to delayed O-O bond formation [213]. Interestingly, the S2 state was observed in both the LS and HS states in these mutants. The authors suggested that the positions and dynamics of critical water molecules required for efficient O–O bond formation may be perturbed in these mutants. Similarly, in the case of the D1-V185 and D1-D61 mutants, it was proposed that the changes introduced could affect substrate water’s movement and possibly the cluster’s associated isomerization [214,216].
Thus, it is clear that many factors can determine whether the release of oxygen in PSII will be fast or slow. Identifying the sequence of events that make a system fast or slow is currently very difficult, especially as it is so far not even certain when and how the O-O bond is formed [for review, see [31,43,168,220]]. Some believe that the formation of a bond between two oxygen atoms is possible only after the accumulation of four positive charges, i.e., during the transition of the manganese complex to the S4 state [44,74,75,217,221], while others think that it can already happen during the formation of the S3 state, which, like the S2 state, shows heterogeneity [23,31,52,69,71,161,222,223,224,225,226,227,228]. Each of the possible configurations of the Mn4CaO5 complex at the various stages of its reorganization associated with the accumulation of positive charge, the attachment of further water molecules, and the release of protons is closely related to the valence changes in specific Mn ions, and consequently, to the magnetic properties of the entire manganese cluster [for review, see [31]]. Due to the very short lifetime of the S4 state, one can only try to determine the organization of the OEC by modeling. This means that it is not possible to state unequivocally what is the actual stage of O-O formation and O2 release. How the reorganization of the Mn4CaO5 cluster itself may occur during the Kok cycle has been the subject of many hypotheses [for review, see [30,34,35,75,187,229]]. Recent studies of the S3 → (S4) → S0 transition kinetics using microsecond Fourier transform infrared (FTIR) spectroscopy [41] and serial femtosecond X-ray crystallography snapshots [40] have allowed for the approximation of the timescale of subsequent steps in the multistep O2 formation process. At least a two-step deprotonation of Mn4CaO5 was observed in the stage before TyrZ+ reduction and after water binding refilling the vacant site created by O2 release. Both experiments showed that the appearance of the peroxide after about 1.2–2.5 ms is the intermediate and slowest step before the formation of the O2 molecule. Furthermore, in both works, the D1-D61 pathway (Cl1 channel) was identified as responsible for the exit of both protons from Mn4CaO5 and D65/D312 as a regulator of these deprotonations. The mechanism of Mn4CaO5 cluster reorganization and substrate water binding proposed by both groups is based on its open cubic structure. This structure is typically observed in cyanobacteria. In higher plants, high-spin S2 and S3 states are detected, which are often explained by the closed cubic structure of the Mn4CaO5 complex [35]. Consequently, there have also been theoretical proposals for a different pathway leading to O-O bond formation and O2 release than the one proposed, for example, in [41], which involves Ca/Mn3/Mn4 μ3-oxo (O5) and Mn1(IV)-oxyl (Ox), or in [40], where different variants have been suggested, not excluding the possibility of the involvement of the Mn(V) state in the final transition step, i.e., in the S4 state, prior to water binding to the substrate. A ‘nucleophilic oxy-oxo coupling’ mechanism between Mn4(V) = oxo and μ3-oxo (O5) has been proposed for the final S4 state in the case of the closed cubic structure of the Mn4CaO5 complex [230]. However, even here, a second pathway for the open cubane structure in the S4 state leading to an identifiable final S0 state has not been ruled out. Nevertheless, each of the proposed scenarios of O-O bond formation and O2 release requires synchronized receipt of protons (H+) and O2 and water binding to restore the S0 state.

5. Conclusions

In summary, this study emphasizes the critical role of external proteins in promoting the fast oxygen release pathway, which contributes to the biphasicity observed in the overall oxygen evolution process. These findings underscore the importance of external proteins in regulating the efficiency and dynamics of water splitting within the OEC. Moreover, it is evident that the conformation of the Si states of the water oxidizing enzyme is not the primary factor influencing the rate of oxygen release in the studied systems. Independent experimental and theoretical evidence from other research groups further supports the notion that the O2 evolution process must be considered in the context of the mutual influence of structural changes in the Mn4CaO5 complex, its protonation states, bound water molecules, and the functioning of water channels. Numerous studies have demonstrated the impact of external proteins on the stabilization and activity of the Mn4CaO5 complex [65,122,231,232,233,234,235,236]. However, their role in PSII heterogeneity has not been explored until now. Furthermore, it is essential to recognize that the entire protein–lipid–dye matrix forming the PSII system significantly influences the oxygen release process from water. The overall reorganization capacity of the PSII system plays a critical role in this process, including the local dynamics of the hydrogen network near the OEC and the influence of external proteins, as well as the dynamics on the acceptor side [36,42,44,71]. While changes on the acceptor side may potentially account for some biphasicity in oxygen release, this explanation becomes less plausible in PSII samples lacking PsbQ, PsbP, or PsbO proteins. In these samples, it is difficult to attribute any acceleration of the slow oxygen release pathway to conformational changes in the QB binding site. Additionally, no significant changes in transfer probabilities between Si states were observed compared to samples depleted of only PsbQ and PsbP.
This research did not examine the impact of Cl- ions on the heterogeneity of the O2 evolution process, as the systems were operated in an environment with adequate chlorine ion concentration. However, it is possible that the release of Cl- ions could affect the observed results, given their potential influence on the Cl1 channel’s efficiency in accepting protons from the Mn4CaO5 complex and delivering water molecules to it [237,238,239,240].
A key aspect of this work is the use of a method that is independent of any OEC operating model and assumptions regarding the initial states of the acceptor and donor sides of PSII. This allows for a more unbiased investigation into the two-phase nature of oxygen evolution, ensuring that the findings are not influenced by preconceived notions or specific model parameters.
In conclusion, these studies emphasize the pivotal role of extrinsic proteins in controlling the efficiency of the fast pathway for oxygen evolution in photosystem II (PSII).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cimb46070428/s1, Figure S1: Examples of raw data of oxygen evolution under short saturating flashes separated by 300 ms obtained for (a) PSII BBY control, (b) PSII BBY—P,Q and (c) PSII BBY—O,P,Q.; Figure S2. Reduction of active reaction centers due to elution of two (PSII BBY—P,Q) and three (PSII BBY—O,P,Q) external proteins compared to the PSII BBY control sample; Figure S3. Flash-induced oxygen yield pattern in intact PSII BBY (a) and PSII BBY depleted of the extrinsic proteins by NaCl (b) and MgCl2 (c) washing (…) The red and blue symbols represent the theoretical fits of the corresponding experimental data, which were obtained using homogeneous Kok models with equal misses and with equal misses and double hits, respectively; Figure S4. Flash-induced oxygen yield pattern in control PSII BBY (A) and PSII BBY depleted of the extrinsic proteins by NaCl (B) and MgCl2 (C) washing. (…) In all cases, the black squares represent experimental data. The green, light blue, red, and dark blue symbols represent data simulated using the heterogeneous 5S-state extended Kok model for the parameters given in Table 1 (main text) for cyclic changes in the parameters αi when the largest missing parameter is set to α0, α1, α2 and α3, respectively, as indicated in the legends of the following figures: Figure S5. The example polarographic signals of oxygen evolution under the 3rd flash of light for the control PSII BBY (black symbols), PSII BBY—P,Q (magenta symbols) and PSII BBY—O,P,Q (blue symbols): (a) raw data and (b) normalized, smoothed signals shifted to 0 on the time scale; Figure S6. SDS-PAGE electrophoretogram of PSII BBY isolated from tobacco: Contr—control, NaCl—the sample depleted of PsbQ (17 kDa) and PsbP (23 kDa) proteins (PSII BBY—P,Q), MgCl2—the sample depleted of PsbQ (17 kDa), PsbP (23 kDa) and PsbO (33k Da) proteins (PSII BBY—O,P,Q), M—protein ladder; Figure S7: The differences between the experimental points and the theoretical curve when the data presented in Figure 6 (main text) were fitted freely without any assumptions (a) for PSII BBY—P,Q and (b) for PSII BBY—O,P,Q). The difference spectrum between the experimental data and the theoretical curve, assuming the time constants obtained for the PSII BBYcontrol sample (c) for PSII BBY—P,Q and (d) for PSII BBY—O,P,Q). The deviations of the fixed parameter curve from the free fit curve are shown by the blue line in (c) and (d).(…); Table S1. Transition parameters and the initial Si-state distribution estimated according to the standard Kok model assuming equal misses (red filled or open circles in Figure S3) or equal misses and double hits (blue filled circles in Figure S3) for the PSII BBY control sample and PSII BBY depleted of two or three external proteins. Refs. [241,242,243,244,245,246] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, K.B.; methodology, K.B.; validation, K.B. and S.K.; formal analysis, S.K. and K.B.; investigation, S.K.; writing—original draft preparation, K.B. and S.K.; writing—review and editing, K.B. and S.K.; visualization, S.K. and K.B.; supervision, K.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

S.K. has been partly supported by the EU Project POWR.03.02.00-00-I004/16.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Pecoraro, V.L.; Baldwin, M.J.; Caudle, M.T.; Hsieh, W.-Y.; Law, N.A. A proposal for water oxidation in photosystem II. Pure Appl. Chem. 1998, 70, 925–929. [Google Scholar] [CrossRef]
  2. Wood, P.M. The potential diagram for oxygen at pH 7. Biochem. J. 1988, 253, 287–289. [Google Scholar] [CrossRef] [PubMed]
  3. Joliot, P.; Barbieri, G.; Chabaud, R. Un nouveau modele des centres photochimiques du systeme II. Photochem. Photobiol. 1969, 10, 309–329. [Google Scholar] [CrossRef]
  4. 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] [PubMed]
  5. Shevela, D.; Kern, J.F.; Govindjee, G.; Messinger, J. Solar energy conversion by photosystem II: Principles and structures. Photosynth. Res. 2023, 156, 279–307. [Google Scholar] [CrossRef] [PubMed]
  6. Casey, J.L.; Sauer, K. EPR detection of a cryogenically photogenerated intermediate in photosynthetic oxygen evolution. Biochim. Biophys. Acta 1984, 767, 21–28. [Google Scholar] [CrossRef]
  7. Dismukes, G.C.; Siderer, Y. Intermediates of a polynuclear manganese center involved in photosynthetic oxidation of water. Proc. Natl. Acad. Sci. USA 1981, 78, 274–278. [Google Scholar] [CrossRef] [PubMed]
  8. Hansson, Ö.; Andréasson, L.-E. EPR-detectable magnetically interacting manganese ions in the photosynthetic oxygen-evolving system after continuous illumination. Biochim. Biophys. Acta 1982, 679, 261–268. [Google Scholar] [CrossRef]
  9. Zimmermann, J.L.; Rutherford, A.W. EPR studies of the oxygen-evolving enzyme of Photosystem II. Biochim. Biophys. Acta 1984, 767, 160–167. [Google Scholar] [CrossRef]
  10. Pantazis, D.A.; Ames, W.; Cox, N.; Lubitz, W.; Neese, F. Two Interconvertible Structures that Explain the Spectroscopic Properties of the Oxygen-Evolving Complex of Photosystem II in the S2 State. Angew. Chem. Int. Ed. 2012, 51, 9935–9940. [Google Scholar] [CrossRef]
  11. Miyagawa, K.; Kawakami, T.; Suzuki, Y.; Isobe, H.; Shoji, M.; Yamanaka, S.; Okumura, M.; Nakajima, T.; Yamaguchi, K. Relative stability among intermediate structures in S2 state of CaMn4O5 cluster in PSII by using hybrid-DFT and DLPNO-CC methods and evaluation of magnetic interactions between Mn ions. J. Photochem. Photobiol. A 2020, 405, 112923. [Google Scholar] [CrossRef]
  12. Sioros, G.; Koulougliotis, D.; Karapanagos, G.; Petrouleas, V. The S1YZ Metalloradical EPR Signal of Photosystem II Contains Two Distinct Components That Advance Respectively to the Multiline and g = 4.1 Conformations of S2. Biochemistry 2007, 46, 210–217. [Google Scholar] [CrossRef]
  13. Kusunoki, M. S1-state Mn4Ca complex of Photosystem II exists in equilibrium between the two most-stable isomeric substates: XRD and EXAFS evidence. J. Photochem. Photobiol. B 2011, 104, 100–110. [Google Scholar] [CrossRef]
  14. Narzi, D.; Mattioli, G.; Bovi, D.; Guidoni, L. A Spotlight on the Compatibility between XFEL and Ab Initio Structures of the Oxygen Evolving Complex in Photosystem II. Chem. Eur. J. 2017, 23, 6969–6973. [Google Scholar] [CrossRef]
  15. Campbell, K.A.; Peloquin, J.M.; Pham, D.P.; Debus, R.J.; Britt, R.D. Parallel polarization EPR detection of an S1-state “multiline” EPR signal in Photosystem II particles from Synechocystis sp. PCC 6803. J. Am. Chem. Soc. 1998, 120, 447–448. [Google Scholar] [CrossRef]
  16. Dexheimer, S.L.; Klein, M.P. Detection of a paramagnetic intermediate in the S1 state of the photosynthetic oxygen-evolving complex. J. Am. Chem. Soc. 1992, 114, 2821–2826. [Google Scholar] [CrossRef]
  17. Drosou, M.; Zahariou, G.; Pantazis, D.A. Orientational Jahn–Teller Isomerism in the Dark-Stable State of Nature’s Water Oxidase. Angew. Chem. Int. Ed. 2021, 60, 13493–13499. [Google Scholar] [CrossRef]
  18. Yamauchi, T.; Mino, H.; Matsukawa, T.; Kawamori, A.; Ono, T.-A. Parallel polarization electron paramagnetic resonance studies of the S1-state manganese cluster in the photosynthetic oxygen-evolving system. Biochemistry 1997, 36, 7520–7526. [Google Scholar] [CrossRef]
  19. Hillier, W.; Wydrzynski, T. Substrate water interactions within the Photosystem II oxygen evolving complex. Phys. Chem. Chem. Phys. 2004, 6, 4882–4889. [Google Scholar] [CrossRef]
  20. Messinger, J.; Badger, J.; Wydrzynski, T. Detection of one slowly exchanging substrate water molecule in the S3 state of photosystem II. Proc. Natl. Acad. Sci. USA 1995, 92, 3209–3213. [Google Scholar] [CrossRef]
  21. Lohmiller, T.; Krewald, V.; Sedoud, A.; Rutherford, A.W.; Neese, F.; Lubitz, W.; Pantazis, D.A.; Cox, N. The First State in the Catalytic Cycle of the Water-Oxidizing Enzyme: Identification of a Water-Derived μ-Hydroxo Bridge. J. Am. Chem. Soc. 2017, 139, 14412–14424. [Google Scholar] [CrossRef]
  22. Boussac, A.; Rutherford, A.W.; Sugiura, M. Electron transfer pathways from the S2-states to the S3-states either after a Ca2+/Sr2+ or a Cl/I exchange in Photosystem II from Thermosynechococcus elongatus. Biochim. Biophys. Acta 2015, 1847, 576–586. [Google Scholar] [CrossRef]
  23. Cox, N.; Retegan, M.; Neese, F.; Pantazis, D.A.; Boussac, A.; Lubitz, W. Electronic structure of the oxygenevolving complex in photosystem II prior to O-O bond formation. Science 2014, 345, 804–808. [Google Scholar] [CrossRef]
  24. Krewald, V.; Retegan, M.; Neese, F.; Lubitz, W.; Pantazis, D.A.; Cox, N. Spin State as a Marker for the Structural Evolution of Nature’s Water-Splitting Catalyst. Inorg. Chem. 2016, 55, 488–501. [Google Scholar] [CrossRef]
  25. Krewald, V.; Retegan, M.; Cox, N.; Messinger, J.; Lubitz, W.; DeBeer, S.; Neese, F.; Pantazis, D.A. Metal oxidation states in biological water splitting. Chem. Sci. 2015, 6, 1676–1695. [Google Scholar] [CrossRef]
  26. Ferreira, K.N.; Iverson, T.M.; Maghlaoui, K.; Barber, J.; Iwata, S. Architecture of the Photosynthetic Oxygen-Evolving Center. Science 2004, 303, 1831–1838. [Google Scholar] [CrossRef]
  27. Umena, Y.; Kawakami, K.; Shen, J.-R.; Kamiya, N. Crystal structure of oxygen-evolving photosystem II at a resolution of 1.9 Å. Nature 2011, 473, 55–61. [Google Scholar] [CrossRef]
  28. De Las Rivas, J.; Balsera, M.; Barber, J. Evolution of oxygenic photosynthesis: Genome-wide analysis of the OEC extrinsic proteins. Trends Plant Sci. 2004, 9, 18–25. [Google Scholar] [CrossRef]
  29. Guo, Y.; Messinger, J.; Kloo, L.; Sun, L. Reversible Structural Isomerization of Nature’s Water Oxidation Catalyst Prior to O–O Bond Formation. J. Am. Chem. Soc. 2022, 144, 11736–11747. [Google Scholar] [CrossRef]
  30. Krewald, V.; Neese, F.; Pantazis, D.A. Implications of structural heterogeneity for the electronic structure of the final oxygen-evolving intermediate in photosystem II. J. Inorg. Biochem. 2019, 199, 110797. [Google Scholar] [CrossRef]
  31. Pantazis, D.A. The S3 State of the Oxygen-Evolving Complex: Overview of Spectroscopy and XFEL Crystallography with a Critical Evaluation of Early-Onset Models for O–O Bond Formation. Inorganics 2019, 7, 55. [Google Scholar] [CrossRef]
  32. Sakashita, N.; Watanabe, H.C.; Ikeda, T.; Saito, K.; Ishikita, H. Origins of Water Molecules in the Photosystem II Crystal Structure. Biochemistry 2017, 56, 3049–3057. [Google Scholar] [CrossRef] [PubMed]
  33. Shoji, M.; Isobe, H.; Miyagawa, K.; Yamaguchi, K. Possibility of the right-opened Mn-oxo intermediate (R-oxo(4444)) among all nine intermediates in the S3 state of the oxygen-evolving complex of photosystem II revealed by large-scale QM/MM calculations. Chem. Phys. 2019, 518, 81–90. [Google Scholar] [CrossRef]
  34. Yamaguchi, K.; Shoji, M.; Isobe, H.; Kawakami, T.; Miyagawa, K.; Suga, M.; Akita, F.; Shen, J.-R. Geometric, electronic and spin structures of the CaMn4O5 catalyst for water oxidation in oxygen-evolving photosystem II. Interplay between experiments and theoretical computations. Coord. Chem. Rev. 2022, 471, 214742. [Google Scholar] [CrossRef]
  35. Zahariou, G.; Ioannidis, N.; Sanakis, Y.; Pantazis, D.A. Arrested SubstrateBinding Resolves Catalytic Intermediates in Higher-Plant Water Oxidation. Angew. Chem. Int. Ed. 2021, 60, 3156–3162. [Google Scholar] [CrossRef] [PubMed]
  36. Li, H.; Nakajima, Y.; Nango, E.; Owada, S.; Yamada, D.; Hashimoto, K.; Luo, F.; Tanaka, R.; Akita, F.; Kato, K.; et al. Oxygen-evolving photosystem II structures during S1–S2–S3 transitions. Nature 2024, 626, 670–677. [Google Scholar] [CrossRef]
  37. Burda, K.; Schmid, G.H. On the Determination of the 5-State Distribution in the Kok Model. Z. Naturforsch. 1996, 51, 329–341. [Google Scholar] [CrossRef]
  38. Burda, K.; Schmid, G.H. Heterogeneity of the mechanism of water splitting in photosystem II. Biochim. Biophys. Acta 2001, 1506, 47–54. [Google Scholar] [CrossRef] [PubMed]
  39. Ananyev, G.; Roy-Chowdhury, S.; Gates, C.; Fromme, P.; Dismukes, G.C. The Catalytic Cycle of Water Oxidation in Crystallized Photosystem II Complexes: Performance and Requirements for Formation of Intermediates. ACS Catal. 2019, 9, 1396–1407. [Google Scholar] [CrossRef]
  40. 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]
  41. 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]
  42. 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]
  43. Lubitz, W.; Pantazis, D.A.; Cox, N. Water oxidation in oxygenic photosynthesis studied by magnetic resonance techniques. FEBS Lett. 2023, 597, 6–29. [Google Scholar] [CrossRef] [PubMed]
  44. Ibrahim, M.; Fransson, T.; Chatterjee, R.; Cheah, M.H.; Hussein, R.; Lassalle, L.; Sutherlin, K.D.; Young, I.D.; Fuller, F.D.; Gul, S.; et al. Untangling the sequence of events during the S2 → S3 transition in photosystem II and implications for the water oxidation mechanism. Proc. Natl. Acad. Sci. USA 2020, 117, 12624–12635. [Google Scholar] [CrossRef]
  45. Retegan, M.; Cox, N.; Lubitz, W.; Neese, F.; Pantazis, D.A. The first tyrosyl radical intermediate formed in the S2–S3 transition of photosystem II. Phys. Chem. Chem. Phys. 2014, 16, 11901–11910. [Google Scholar] [CrossRef] [PubMed]
  46. Haumann, M.; Liebisch, P.; Müller, C.; Barra, M.; Grabolle, M.; Dau, H. Photosynthetic O2 formation tracked by time-resolved x-ray experiments. Science 2005, 310, 1019–1021. [Google Scholar] [CrossRef]
  47. Lavergne, J.; Junge, W. Proton release during the redox cycle of the water oxidase. Photosynth. Res. 1993, 38, 279–296. [Google Scholar] [CrossRef]
  48. Pushkar, Y.; Ravari, A.K.; Jensen, S.C.; Palenik, M. Early Binding of Substrate Oxygen Is Responsible for a Spectroscopically Distinct S2 State in Photosystem II. J. Phys. Chem. Lett. 2019, 10, 5284–5291. [Google Scholar] [CrossRef] [PubMed]
  49. Rappaport, F.; Lavergne, J. Proton release during successive oxidation steps of the photosynthetic water oxidation process: Stoichiometries and pH dependence. Biochemistry 1991, 30, 10004–10012. [Google Scholar] [CrossRef]
  50. Suzuki, H.; Sugiura, M.; Noguchi, T. Monitoring Water Reactions during the S-State Cycle of the Photosynthetic Water-Oxidizing Center: Detection of the DOD Bending Vibrations by Means of Fourier Transform Infrared Spectroscopy. Biochemistry 2008, 47, 11024–11030. [Google Scholar] [CrossRef]
  51. Yang, K.R.; Lakshmi, K.V.; Brudvig, G.W.; Batista, V.S. Is Deprotonation of the Oxygen-Evolving Complex of Photosystem II during the S1 → S2 Transition Suppressed by Proton Quantum Delocalization? J. Am. Chem. Soc. 2021, 143, 8324–8332. [Google Scholar] [CrossRef] [PubMed]
  52. 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]
  53. Saito, K.; Rutherford, A.W.; Ishikita, H. Energetics of proton release on the first oxidation step in the water-oxidizing enzyme. Nat. Commun. 2015, 6, 8488. [Google Scholar] [CrossRef] [PubMed]
  54. Haumann, M.; Müller, C.; Liebisch, P.; Iuzzolino, L.; Dittmer, J.; Grabolle, M.; Neisius, T.; Meyer-Klaucke, W.; Dau, H. Structural and Oxidation State Changes of the Photosystem II Manganese Complex in Four Transitions of the Water Oxidation Cycle (S0 → S1, S1 → S2, S2 → S3, and S3,4 → S0) Characterized by X-ray Absorption Spectroscopy at 20 K and Room Temperature. Biochemistry 2005, 44, 1894–1908. [Google Scholar] [CrossRef] [PubMed]
  55. Iuzzolino, L.; Dittmer, J.; Dörner, W.; Meyer-Klaucke, W.; Dau, H. X-ray absorption spectroscopy on layered photosystem II membrane particles suggests manganese-centered oxidation of the oxygen-evolving complex for the S0-S1, S1-S2, and S2-S3 transitions of the water oxidation cycle. Biochemistry 1998, 37, 17112–17119. [Google Scholar] [CrossRef] [PubMed]
  56. Dau, H.; Iuzzolino, L.; Dittmer, J. The tetra-manganese complex of photosystem II during its redox cycle—X-ray absorption results and mechanistic implications. Biochim. Biophys. Acta 2001, 1503, 24–39. [Google Scholar] [CrossRef]
  57. Dau, H.; Liebisch, P.; Haumann, M. X-ray absorption spectroscopy to analyze nuclear geometry and electronic structure of biological metal centers--potential and questions examined with special focus on the tetra-nuclear manganese complex of oxygenic photosynthesis. Anal. Bioanal. Chem. 2003, 376, 562–583. [Google Scholar] [CrossRef]
  58. Schuth, N.; Zaharieva, I.; Chernev, P.; Berggren, G.; Anderlund, N.; Styring, S.; Dau, H.; Haumann, M. Kα X-ray Emission Spectroscopy on the Photosynthetic Oxygen-Evolving Complex Supports Manganese Oxidation and Water Binding in the S3 State. Inorg. Chem. 2018, 57, 10424–10430. [Google Scholar] [CrossRef]
  59. Guiles, R.D.; Zimmermann, J.L.; McDermott, A.E.; Yachandra, V.K.; Cole, J.L.; Dexheimer, S.L.; Britt, R.D.; Wieghardt, K.; Bossek, U.; Sauer, K.; et al. The S3 state of photosystem II: Differences between the structure of the manganese complex in the S2 and S3 states determined by x-ray absorption spectroscopy. Biochemistry 1990, 29, 471–485. [Google Scholar] [CrossRef]
  60. MacLachlan, D.J.; Nugent, J.H.A.; Evans, M.C.W. A XANES study of the manganese complex of inhibited PS II membranes indicates manganese redox changes between the modified S1, S2 and S3 states. Biochim. Biophys. Acta 1994, 1185, 103–111. [Google Scholar] [CrossRef]
  61. Mandal, M.; Kawashima, K.; Saito, K.; Ishikita, H. Redox Potential of the Oxygen-Evolving Complex in the Electron Transfer Cascade of Photosystem II. J. Phys. Chem. Lett. 2020, 11, 249–255. [Google Scholar] [CrossRef] [PubMed]
  62. Messinger, J.; Robblee, J.H.; Bergmann, U.; Fernandez, C.; Glatzel, P.; Visser, H.; Cinco, R.M.; McFarlane, K.L.; Bellacchio, E.; Pizarro, S.A.; et al. Absence of Mn-Centered Oxidation in the S2 → S3 Transition: Implications for the Mechanism of Photosynthetic Water Oxidation. J. Am. Chem. Soc. 2001, 123, 7804–7820. [Google Scholar] [CrossRef]
  63. Roelofs, T.A.; Liang, W.; Latimer, M.J.; Cinco, R.M.; Rompel, A.; Andrews, J.C.; Sauer, K.; Yachandra, V.K.; Klein, M.P. Oxidation states of the manganese cluster during the flash-induced S-state cycle of the photosynthetic oxygen-evolving complex. Proc. Natl. Acad. Sci. USA 1996, 93, 3335–3340. [Google Scholar] [CrossRef] [PubMed]
  64. Polander, B.C.; Barry, B.A. Detection of an intermediary, protonated water cluster in photosynthetic oxygen evolution. Proc. Natl. Acad. Sci. USA 2013, 110, 10634–10639. [Google Scholar] [CrossRef] [PubMed]
  65. Barry, B.A.; Brahmachari, U.; Guo, Z. Tracking Reactive Water and Hydrogen-Bonding Networks in Photosynthetic Oxygen Evolution. Acc. Chem. Res. 2017, 5, 1937–1945. [Google Scholar] [CrossRef] [PubMed]
  66. Siegbahn, P.E.M. Water oxidation in photosystem II: Oxygen release, proton release and the effect of chloride. Dalton Trans. 2009, 45, 10063–10068. [Google Scholar] [CrossRef] [PubMed]
  67. Boussac, A.; Sugiura, M.; Sellés, J. Probing the proton release by Photosystem II in the S1 to S2 high-spin transition. Biochim. Biophys. Acta (BBA)—Bioenergetics 2022, 1863, 148546. [Google Scholar] [CrossRef]
  68. Li, X.; Siegbahn, P.E.M. Alternative mechanisms for O2 release and O–O bond formation in the oxygen evolving complex of photosystem II. Phys. Chem. Chem. Phys. 2015, 17, 12168–12174. [Google Scholar] [CrossRef]
  69. Siegbahn, P.E.M. Structures and Energetics for O2 Formation in Photosystem II. Acc. Chem. Res. 2009, 42, 1871–1880. [Google Scholar] [CrossRef]
  70. Siegbahn, P.E.M. Nucleophilic water attack is not a possible mechanism for O–O bond formation in photosystem II. Proc. Natl. Acad. Sci. USA 2017, 114, 4966–4968. [Google Scholar] [CrossRef]
  71. 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] [PubMed]
  72. Barber, J. A mechanism for water splitting and oxygen production in photosynthesis. Nat. Plants 2017, 3, 17041. [Google Scholar] [CrossRef] [PubMed]
  73. Sproviero, E.M.; Gascón, J.A.; McEvoy, J.P.; Brudvig, G.W.; Batista, V.S. Quantum Mechanics/Molecular Mechanics Study of the Catalytic Cycle of Water Splitting in Photosystem II. J. Am. Chem. Soc. 2008, 130, 3428–3442. [Google Scholar] [CrossRef] [PubMed]
  74. Vinyard, D.J.; Khan, S.; Brudvig, G.W. Photosynthetic water oxidation: Binding and activation of substrate waters for O–O bond formation. Faraday Discuss. 2015, 185, 37–50. [Google Scholar] [CrossRef]
  75. Zhang, B.; Sun, L. Why nature chose the Mn4CaO5 cluster as water-splitting catalyst in photosystem II: A new hypothesis for the mechanism of O–O bond formation. Dalton Trans. 2018, 47, 14381–14387. [Google Scholar] [CrossRef] [PubMed]
  76. Cox, N.; Messinger, J. Reflections on substrate water and dioxygen formation. Biochim. Biophys. Acta 2013, 1827, 1020–1030. [Google Scholar] [CrossRef] [PubMed]
  77. Guo, Y.; Li, H.; He, L.-L.; Zhao, D.-X.; Gong, L.-D.; Yang, Z.-Z. Theoretical reflections on the structural polymorphism of the oxygen-evolving complex in the S2 state and the correlations to substrate water exchange and water oxidation mechanism in photosynthesis. Biochim. Biophys. Acta 2017, 1858, 833–846. [Google Scholar] [CrossRef]
  78. Suga, M.; Akita, F.; Yamashita, K.; Nakajima, Y.; Ueno, G.; Li, H.; Yamane, T.; Hirata, K.; Umena, Y.; Yonekura, S.; et al. An oxyl/oxo mechanism for oxygen-oxygen coupling in PSII revealed by an x-ray free-electron laser. Science 2019, 366, 334–338. [Google Scholar] [CrossRef]
  79. Burda, K. Dynamics of electron transfer in photosystem II. Cell Biochem. Biophys. 2007, 47, 271–284. [Google Scholar] [CrossRef]
  80. 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 2021, 405, 112905. [Google Scholar] [CrossRef]
  81. Linke, K.; Ho, F.M. Water in Photosystem II: Structural, functional and mechanistic considerations. Biochim. Biophys. Acta 2014, 1837, 14–32. [Google Scholar] [CrossRef] [PubMed]
  82. Hansson, Ö.; Andréasson, L.-E.; Vänngård, T. Oxygen from water is coordinated to manganese in the S2 state of photosystem II. FEBS Lett. 1986, 195, 151–154. [Google Scholar] [CrossRef]
  83. Burda, K.; Bader, K.P.; Schmid, G.H. An estimation of the size of the water cluster present at the cleavage site of the water splitting enzyme. FEBS Lett. 2001, 491, 81–84. [Google Scholar] [CrossRef] [PubMed]
  84. Loll, B.; Kern, J.; Saenger, W.; Zouni, A.; Biesiadka, J. Towards complete cofactor arrangement in the 3.0 Å resolution structure of photosystem II. Nature 2005, 438, 1040–1044. [Google Scholar] [CrossRef] [PubMed]
  85. Anderson, J.M. Does functional photosystem II complex have an oxygen channel? FEBS Lett. 2001, 488, 1–4. [Google Scholar] [CrossRef] [PubMed]
  86. Anderson, J.M.; Chow, W.S. Structural and functional dynamics of plant photosystem II. Philos. Trans. R Soc. Lond. B Biol. Sci. 2002, 357, 1421–1430; discussion 1469–1470. [Google Scholar] [CrossRef] [PubMed]
  87. Rutherford, A.W. Photosystem II, the water-splitting enzyme. Trends Biochem. Sci. 1989, 14, 227–232. [Google Scholar] [CrossRef] [PubMed]
  88. Wydrzynski, T.; Hillier, W.; Messinger, J. On the functional significance of substrate accessibility in the photosynthetic water oxidation mechanism. Physiol. Plant. 1996, 96, 342–350. [Google Scholar] [CrossRef]
  89. 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]
  90. Sakashita, N.; Watanabe, H.C.; Ikeda, T.; Ishikita, H. Structurally conserved channels in cyanobacterial and plant photosystem II. Photosynth. Res. 2017, 133, 75–85. [Google Scholar] [CrossRef]
  91. Ho, F.M.; Styring, S. Access channels and methanol binding site to the CaMn4 cluster in Photosystem II based on solvent accessibility simulations, with implications for substrate water access. Biochim. Biophys. Acta 2008, 1777, 140–153. [Google Scholar] [CrossRef] [PubMed]
  92. Vogt, L.; Vinyard, D.J.; Khan, S.; Brudvig, G.W. Oxygen-evolving complex of Photosystem II: An analysis of second-shell residues and hydrogen-bonding networks. Curr. Opin. Chem. Biol. 2015, 25, 152–158. [Google Scholar] [CrossRef] [PubMed]
  93. 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] [PubMed]
  94. Guskov, A.; Kern, J.; Gabdulkhakov, A.; Broser, M.; Zouni, A.; Saenger, W. Cyanobacterial photosystem II at 2.9Å resolution and the role of quinones, lipids, channels and chloride. Nat. Struct. Mol. Biol. 2009, 16, 334–342. [Google Scholar] [CrossRef]
  95. Ishikita, H.; Saenger, W.; Loll, B.; Biesiadka, J.; Knapp, E.-W. Energetics of a Possible Proton Exit Pathway for Water Oxidation in Photosystem II. Biochemistry 2006, 45, 2063–2071. [Google Scholar] [CrossRef]
  96. Gabdulkhakov, A.G.; Kljashtorny, V.G.; Dontsova, M.V. Analysis of Molecular Oxygen Exit Pathways in Cyanobacterial Photosystem II: Molecular Dynamics Studies. Kristallografiya 2015, 60, 926–931. [Google Scholar] [CrossRef]
  97. Gabdulkhakov, A.G.; Kljashtorny, V.G.; Dontsova, M.V. Molecular Dynamics Studies of Pathways of Water Movement in Cyanobacterial Photosystem II. Kristallografiya 2015, 60, 91–97. [Google Scholar] [CrossRef]
  98. Vassiliev, S.; Zaraiskaya, T.; Bruce, D. Molecular dynamics simulations reveal highly permeable oxygen exit channels shared with water uptake channels in photosystem II. Biochim. Biophys. Acta 2013, 1827, 1148–1155. [Google Scholar] [CrossRef]
  99. Murray, J.W.; Barber, J. Structural characteristics of channels and pathways in photosystem II including the identification of an oxygen channel. J. Struct. Biol. 2007, 159, 228–237. [Google Scholar] [CrossRef]
  100. Vassiliev, S.; Zaraiskaya, T.; Bruce, D. Exploring the energetics of water permeation in photosystem II by multiple steered molecular dynamics simulations. Biochim. Biophys. Acta 2012, 1817, 1671–1678. [Google Scholar] [CrossRef]
  101. Hussein, R.; Ibrahim, M.; Bhowmick, A.; Simon, P.S.; Bogacz, I.; Doyle, M.D.; Dobbek, H.; Zouni, A.; Messinger, J.; Yachandra, V.K.; et al. Evolutionary diversity of proton and water channels on the oxidizing side of photosystem II and their relevance to function. Photosynth. Res. 2023, 158, 91–107. [Google Scholar] [CrossRef]
  102. Stuchebrukhov, A.A. Mechanisms of proton transfer in proteins: Localized charge transfer versus delocalized soliton transfer. Phys. Rev. E 2009, 79, 031927. [Google Scholar] [CrossRef] [PubMed]
  103. Takaoka, T.; Sakashita, N.; Saito, K.; Ishikita, H. pKa of a Proton-Conducting Water Chain in Photosystem II. J. Phys. Chem. Lett. 2016, 7, 1925–1932. [Google Scholar] [CrossRef]
  104. Shimizu, T.; Sugiura, M.; Noguchi, T. Mechanism of Proton-Coupled Electron Transfer in the S0-to-S1 Transition of Photosynthetic Water Oxidation As Revealed by Time-Resolved Infrared Spectroscopy. J. Phys. Chem. B 2018, 122, 9460–9470. [Google Scholar] [CrossRef]
  105. 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]
  106. Kuroda, H.; Kawashima, K.; Ueda, K.; Ikeda, T.; Saito, K.; Ninomiya, R.; Hida, C.; Takahashi, Y.; Ishikita, H. Proton transfer pathway from the oxygen-evolving complex in photosystem II substantiated by extensive mutagenesis. Biochim. Biophys. Acta Bioenerg. 2021, 1862, 148329. [Google Scholar] [CrossRef] [PubMed]
  107. Okamoto, Y.; Shimada, Y.; Nagao, R.; Noguchi, T. Proton and Water Transfer Pathways in the S2 → S3 Transition of the Water-Oxidizing Complex in Photosystem II: Time-Resolved Infrared Analysis of the Effects of D1-N298A Mutation and NO3 Substitution. J. Photochem. Photobiol. B 2021, 125, 6864–6873. [Google Scholar] [CrossRef]
  108. Takemoto, H.; Sugiura, M.; Noguchi, T. Proton Release Process during the S2-to-S3 Transition of Photosynthetic Water Oxidation As Revealed by the pH Dependence of Kinetics Monitored by Time-Resolved Infrared Spectroscopy. Biochemistry 2019, 58, 4276–4283. [Google Scholar] [CrossRef]
  109. Askerka, M.; Brudvig, G.W.; Batista, V.S. The O2-Evolving Complex of Photosystem II: Recent Insights from Quantum Mechanics/Molecular Mechanics (QM/MM), Extended X-ray Absorption Fine Structure (EXAFS), and Femtosecond X-ray Crystallography Data. Acc. Chem. Res. 2017, 50, 41–48. [Google Scholar] [CrossRef] [PubMed]
  110. Askerka, M.; Wang, J.; Vinyard, D.; Brudvig, G.; Batista, V. S3 State of the O2-Evolving Complex of Photosystem II: Insights from QM/MM, EXAFS, and Femtosecond X-ray Diffraction. Biochemistry 2016, 55, 981–984. [Google Scholar] [CrossRef]
  111. Retegan, M.; Pantazis, D.A. Interaction of methanol with the oxygen-evolving complex: Atomistic models, channel identification, species dependence, and mechanistic implications. Chem. Sci. 2016, 7, 6463–6476. [Google Scholar] [CrossRef]
  112. Retegan, M.; Pantazis, D.A. Differences in the Active Site of Water Oxidation among Photosynthetic Organisms. J. Am. Chem. Soc. 2017, 139, 14340–14343. [Google Scholar] [CrossRef]
  113. Ugur, I.; Rutherford, A.W.; Kaila, V.R.I. Redox-coupled substrate water reorganization in the active site of Photosystem II—The role of calcium in substrate water delivery. Biochim. Biophys. Acta 2016, 1857, 740–748. [Google Scholar] [CrossRef]
  114. Flesher, D.A.; Liu, J.; Wiwczar, J.M.; Reiss, K.; Yang, K.R.; Wang, J.; Askerka, M.; Gisriel, C.J.; Batista, V.S.; Brudvig, G.W. Glycerol binding at the narrow channel of photosystem II stabilizes the low-spin S2 state of the oxygen-evolving complex. Photosynth. Res. 2022, 152, 167–175. [Google Scholar] [CrossRef]
  115. Zimmermann, J.L.; Rutherford, A.W. Electron paramagnetic resonance properties of the S2 state of the oxygen-evolving complex of photosystem II. Biochemistry 1986, 25, 4609–4615. [Google Scholar] [CrossRef]
  116. Ogata, K.; Yuki, T.; Hatakeyama, M.; Uchida, W.; Nakamura, S. All-Atom Molecular Dynamics Simulation of Photosystem II Embedded in Thylakoid Membrane. J. Am. Chem. Soc. 2013, 135, 15670–15673. [Google Scholar] [CrossRef]
  117. Gabdulkhakov, A.; Guskov, A.; Broser, M.; Kern, J.; Müh, F.; Saenger, W.; Zouni, A. Probing the Accessibility of the Mn4Ca Cluster in Photosystem II: Channels Calculation, Noble Gas Derivatization, and Cocrystallization with DMSO. Structure 2009, 17, 1223–1234. [Google Scholar] [CrossRef]
  118. Caspy, I.; Fadeeva, M.; Mazor, Y.; Nelson, N. Structure of Dunaliella Photosystem II reveals conformational flexibility of stacked and unstacked supercomplexes. biorxiv 2021, 1–29. [Google Scholar] [CrossRef]
  119. Bricker, T.M.; Frankel, L.K. The structure and function of the 33 kDa extrinsic protein of Photosystem II: A critical assessment. Photosynth. Res. 1998, 56, 157–173. [Google Scholar] [CrossRef]
  120. Cammarata, K.V.; Cheniae, G.M. Studies on 17, 24 kD Depleted Photosystem II Membranes: I. Evidences for High and Low Affinity Calcium Sites in 17, 24 kD Depleted PSII Membranes from Wheat versus Spinach. Plant Physiol. 1987, 84, 587–595. [Google Scholar] [CrossRef]
  121. Enami, I.; Tomo, T.; Kitamura, M.; Katoh, S. Immobilization of the three extrinsic proteins in spinach oxygen-evolving Photosystem II membranes: Roles of the proteins in stabilization of binding of Mn and Ca2+. Biochim. Biophys. Acta 1994, 1185, 75–80. [Google Scholar] [CrossRef]
  122. Ifuku, K.; Noguchi, T. Structural Coupling of Extrinsic Proteins with the Oxygen-Evolving Center in Photosystem II. Front. Plant Sci. 2016, 7, 84. [Google Scholar] [CrossRef]
  123. Kruk, J.; Burda, K.; Jemioła-Rzemińska, M.; Strzałka, K. The 33 kDa protein of photosystem II is a low-affinity calcium- and lanthanide-binding protein. Biochemistry 2003, 42, 14862–14867. [Google Scholar] [CrossRef]
  124. Vass, I.; Ono, T.; Inoue, Y. Removal of 33 kDa extrinsic protein specifically stabilizes the S2QA− charge pair in photosystem II. FEBS Lett. 1987, 211, 215–220. [Google Scholar] [CrossRef]
  125. Burda, K.; He, P.; Bader, K.P.; Schmid, G.H. Temperature dependence of the O2-oscillation pattern in the filamentous cyanobacterium Oscillatoria chalybea and in Chlorella kessleri. Z. Naturforsch. 1996, 51, 823–832. [Google Scholar] [CrossRef]
  126. Berthold, D.A.; Babcock, G.T.; Yocum, C.F. A highly resolved, oxygen-evolving photosystem II preparation from spinach thylakoid membranes. FEBS Lett. 1981, 134, 231–234. [Google Scholar] [CrossRef]
  127. Ono, T.-A.; Inoue, Y. Mn-preserving extraction of 33-, 24- and 16-kDa proteins from O2-evolving PS II particles by divalent salt-washing. FEBS Lett. 1983, 164, 255–260. [Google Scholar] [CrossRef]
  128. Burda, K.; Strzałka, K.; Schmid, G.H. Europium- and Dysprosium-Ions as Probes for the Study of Calcium Binding Sites in Photosystem II. Z. Naturforsch. 1995, 50, 220–230. [Google Scholar] [CrossRef]
  129. Joliot, P.; Joliot, A. A polarographic method for detection of oxygen production and reduction of Hill reagent by isolated chloroplasts. Biochim. Biophys. Acta 1968, 153, 625–634. [Google Scholar] [CrossRef]
  130. Schmid, G.H.; Thibault, P. Evidence for a Rapid Oxygen-Uptake in Tobacco Chloroplasts. Z. Naturforsch. 1979, 34, 414–418. [Google Scholar] [CrossRef]
  131. Ananyev, G.; Gates, C.; Dismukes, G.C. The Oxygen quantum yield in diverse algae and cyanobacteria is controlled by partitioning of flux between linear and cyclic electron flow within photosystem II. Biochim. Biophys. Acta Bioenerg. 2016, 1857, 1380–1391. [Google Scholar] [CrossRef]
  132. Pham, L.V.; Janna Olmos, J.D.; Chernev, P.; Kargul, J.; Messinger, J. Unequal misses during the flash-induced advancement of photosystem II: Effects of the S state and acceptor side cycles. Photosynth. Res. 2019, 139, 93–106. [Google Scholar] [CrossRef]
  133. Schulder, R.; Burda, K.; Strzałka, K.; Bader, K.P.; Schmid, G.H. Study on the Parameters Affecting Oxygen Release Time Measurements by Amperometry. Z. Naturforsch. 1992, 47, 465–473. [Google Scholar] [CrossRef]
  134. Kirchhoff, H.; Horstmann, S.; Weis, E. Control of the photosynthetic electron transport by PQ diffusion microdomains in thylakoids of higher plants. Biochim. Biophys. Acta Bioenerg. 2000, 1459, 148–168. [Google Scholar] [CrossRef]
  135. Lavergne, J.; Bouchaud, J.-P.; Joliot, P. Plastoquinone compartmentation in chloroplasts. II. Theoretical aspects. Biochim. Biophys. Acta Bioenerg. 1992, 1101, 13–22. [Google Scholar] [CrossRef]
  136. McCauley, S.W.; Melis, A. Quantitation of plastoquinone photoreduction in spinach chloroplasts. Photosynth. Res. 1986, 8, 3–16. [Google Scholar] [CrossRef]
  137. Thibault, P. A new attempt to study the oxygen evolving system of photosynthesis: Determination of transition probabilities of a state i. J. Theor. Biol. 1978, 73, 271–284. [Google Scholar] [CrossRef]
  138. Meunier, P.C. Oxygen evolution by Photosystem II: The contribution of backward transitions to the anomalous behaviour of double-hits revealed by a new analysis method. Photosynth. Res. 1993, 36, 111–118. [Google Scholar] [CrossRef]
  139. Delrieu, M.-J. Simple explanation of the misses in the cooperation of charges in photosynthetic O2 evolution. Photochem. Photobiol. 1974, 20, 441–454. [Google Scholar] [CrossRef]
  140. Delrieu, M.-J. Evidence for Unequal Misses in Oxygen Flash Yield Sequence in Photosynthesis. Z. Naturforsch. 1983, 38, 247–258. [Google Scholar] [CrossRef]
  141. Lavorel, J. Matrix analysis of the oxygen evolving system of photosynthesis. J. Theor. Biol. 1976, 57, 171–185. [Google Scholar] [CrossRef]
  142. Lavorel, J. On the origin of the damping of the O2 yield in sequences of flashes. In Photosynthetic Oxygen Evolution; Metzner, H., Ed.; Academic Press: New York, NY, USA, 1978. [Google Scholar]
  143. Han, G.; Chernev, P.; Styring, S.; Messinger, J.; Mamedov, F. Molecular basis for turnover inefficiencies (misses) during water oxidation in photosystem II. Chem. Sci. 2022, 13, 8667–8678. [Google Scholar] [CrossRef]
  144. Han, G.; Mamedov, F.; Styring, S. Misses during Water Oxidation in Photosystem II Are S State-dependent. J. Biol. Chem. 2012, 287, 13422–13429. [Google Scholar] [CrossRef]
  145. Vermaas, W.F.J.; Renger, G.; Dohnt, G. The reduction of the oxygen-evolving system in chloroplasts by thylakoid components. Biochim. Biophys. Acta 1984, 764, 192–202. [Google Scholar] [CrossRef]
  146. Styring, S.; Rutherford, A.W. In the oxygen-evolving complex of photosystem II the S0 state is oxidized to the S1 state by D+ (signal IIslow). Biochemistry 1987, 26, 2401–2405. [Google Scholar] [CrossRef]
  147. Vass, I.; Styring, S. pH-dependent charge equilibria between tyrosine-D and the S states in photosystem II. Estimation of relative midpoint redox potentials. Biochemistry 1991, 30, 830–839. [Google Scholar] [CrossRef]
  148. Vass, I.; Deák, Z.; Hideg, É. Charge equilibrium between the water-oxidizing complex and the electron donor tyrosine-D in Photosystem II. Biochim. Biophys. Acta Bioenerg. 1990, 1017, 63–69. [Google Scholar] [CrossRef]
  149. Weiss, C.; Sauer, K. Activation kinetics of photosynthetic oxygen evolution under 20–40 nanosecond laser flashes. Photochem. Photobiol. 1969, 11, 495–501. [Google Scholar] [CrossRef]
  150. Weiss, C., Jr.; Solnit, K.T.; von Gutfeld, R.J. Flash activation kinetics and photosynthetic unit size for oxygen evolution using 3-nsec light flashes. Biochim. Biophys. Acta 1971, 253, 298–301. [Google Scholar] [CrossRef]
  151. Vermass, W.F.; Rutherford, A.W.; Hansson, O. Site-directed mutagenesis in photosystem II of the cyanobacterium Synechocystis sp. PCC 6803: Donor D is a tyrosine residue in the D2 protein. Proc. Natl. Acad. Sci. USA 1988, 85, 8477–8481. [Google Scholar] [CrossRef]
  152. Debus, R.J.; Barry, B.A.; Babcock, G.T.; McIntosh, L. Site-directed mutagenesis identifies a tyrosine radical involved in the photosynthetic oxygen-evolving system. Proc. Natl. Acad. Sci. USA 1988, 85, 427–430. [Google Scholar] [CrossRef]
  153. Yano, J.; Yachandra, V. Mn4Ca Cluster in Photosynthesis: Where and How Water is Oxidized to Dioxygen. Chem. Rev. 2014, 114, 4175–4205. [Google Scholar] [CrossRef]
  154. Pokhrel, R.; Service, R.J.; Debus, R.J.; Brudvig, G.W. Mutation of Lysine 317 in the D2 Subunit of Photosystem II Alters Chloride Binding and Proton Transport. Biochemistry 2013, 52, 4758–4773. [Google Scholar] [CrossRef]
  155. Sirohiwal, A.; Neese, F.; Pantazis, D.A. Microsolvation of the Redox-Active Tyrosine-D in Photosystem II: Correlation of Energetics with EPR Spectroscopy and Oxidation-Induced Proton Transfer. J. Am. Chem. Soc. 2019, 141, 3217–3231. [Google Scholar] [CrossRef]
  156. Bricker, T.M.; Roose, J.L.; Fagerlund, R.D.; Frankel, L.K.; Eaton-Rye, J.J. The extrinsic proteins of Photosystem II. Biochim. Biophys. Acta 2012, 1817, 121–142. [Google Scholar] [CrossRef]
  157. Popelkova, H.; Yocum, C.F. PsbO, the manganese-stabilizing protein: Analysis of the structure–function relations that provide insights into its role in photosystem II. J. Photochem. Photobiol. B 2011, 104, 179–190. [Google Scholar] [CrossRef]
  158. Liang, W.; Roelofs, T.A.; Cinco, R.M.; Rompel, A.; Latimer, M.J.; Yu, W.O.; Sauer, K.; Klein, M.P.; Yachandra, V.K. Structural change of the Mn cluster during the S2 → S3 state transition of the oxygen-evolving complex of Photosystem II. Does it reflect the onset of water/substrate oxidation? Determination by Mn X-ray absorption spectroscopy. J Am. Chem. Soc. 2000, 122, 3399–3412. [Google Scholar] [CrossRef]
  159. Noguchi, T.; Sugiura, M. Flash-induced FTIR difference spectra of the water oxidizing complex in moderately hydrated photosystem II core films: Effect of hydration extent on S-state transitions. Biochemistry 2002, 41, 2322–2330. [Google Scholar] [CrossRef]
  160. Kim, C.J.; Debus, R.J. Evidence from FTIR Difference Spectroscopy That a Substrate H2O Molecule for O2 Formation in Photosystem II Is Provided by the Ca Ion of the Catalytic Mn4CaO5 Cluster. Biochemistry 2017, 56, 2558–2570. [Google Scholar] [CrossRef] [PubMed]
  161. Siegbahn, P.E.M. The S2 to S3 transition for water oxidation in PSII (Photosystem II), revisited. Phys. Chem. Chem. Phys. 2018, 20, 22926–22931. [Google Scholar] [CrossRef] [PubMed]
  162. Capone, M.; Bovi, D.; Narzi, D.; Guidoni, L. Reorganization of Substrate Waters between the Closed and Open Cubane Conformers during the S2 to S3 Transition in the Oxygen Evolving Complex. Biochemistry 2015, 54, 6439–6442. [Google Scholar] [CrossRef]
  163. Narzi, D.; Bovi, D.; Guidoni, L. Pathway for Mn-cluster oxidation by tyrosine-Z in the S2 state of photosystem II. Proc. Natl. Acad. Sci. USA 2014, 111, 8723–8728. [Google Scholar] [CrossRef]
  164. Siegbahn, P.E.M. Substrate Water Exchange for the Oxygen Evolving Complex in PSII in the S1, S2, and S3 States. J. Am. Chem. Soc. 2013, 135, 9442–9449. [Google Scholar] [CrossRef]
  165. 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]
  166. Chrysina, M.; Heyno, E.; Kutin, Y.; Reus, M.; Nilsson, H.; Nowaczyk, M.M.; DeBeer, S.; Neese, F.; Messinger, J.; Lubitz, W.; et al. Five-coordinate MnIV intermediate in the activation of nature’s water splitting cofactor. Proc. Natl. Acad. Sci. USA 2019, 116, 16841–16846. [Google Scholar] [CrossRef]
  167. Cox, N.; Pantazis, D.A.; Neese, F.; Lubitz, W. Biological Water Oxidation. Acc. Chem. Res. 2013, 46, 1588–1596. [Google Scholar] [CrossRef]
  168. Orio, M.; Pantazis, D.A. Successes, challenges, and opportunities for quantum chemistry in understanding metalloenzymes for solar fuels research. Chem. Commun. 2021, 57, 33952–33974. [Google Scholar] [CrossRef]
  169. Corry, T.A.; O’Malley, P.J. Proton Isomers Rationalize the High- and Low-Spin Forms of the S2 State Intermediate in the Water-Oxidizing Reaction of Photosystem II. J. Phys. Chem. Lett. 2019, 10, 5226–5230. [Google Scholar] [CrossRef]
  170. Rogers, C.J.; Hardwick, O.; Corry, T.A.; Rummel, F.; Collison, D.; Bowen, A.M.; O’Malley, P.J. Magnetic and Electronic Structural Properties of the S3 State of Nature’s Water Oxidizing Complex: A Combined Study in ELDOR-Detected Nuclear Magnetic Resonance Spectral Simulation and Broken-Symmetry Density Functional Theory. ACS Omega 2022, 7, 41783–41788. [Google Scholar] [CrossRef] [PubMed]
  171. Nass, K.; Foucar, L.; Barends, T.R.; Hartmann, E.; Botha, S.; Shoeman, R.L.; Doak, R.B.; Alonso-Mori, R.; Aquila, A.; Bajt, S.; et al. Indications of radiation damage in ferredoxin microcrystals using high-intensity X-FEL beams. J. Synchrotron. Radiat. 2015, 22, 225–238. [Google Scholar] [CrossRef] [PubMed]
  172. Tanaka, A.; Fukushima, Y.; Kamiya, N. Two different structures of the oxygen-evolving complex in the same polypeptide frameworks of photosystem II. J. Am. Chem. Soc. 2017, 139, 1718–1721. [Google Scholar] [CrossRef] [PubMed]
  173. Amin, M.; Askerka, M.; Batista, V.S.; Brudvig, G.W.; Gunner, M.R. X-ray Free Electron Laser Radiation Damage through the S-State Cycle of the Oxygen-Evolving Complex of Photosystem II. J Phys. Chem. B 2017, 121, 9382–9388. [Google Scholar] [CrossRef] [PubMed]
  174. Li, Y.; Yao, R.; Chen, Y.; Xu, B.; Chen, C.; Zhang, C. Mimicking the Catalytic Center for the Water-Splitting Reaction in Photosystem II. Catalysts 2020, 10, 185. [Google Scholar] [CrossRef]
  175. Nass, K. Radiation damage in protein crystallography at X-ray free-electron lasers. Acta Cryst. 2019, 75, 211–218. [Google Scholar] [CrossRef]
  176. Sirohiwal, A.; Pantazis, D.A. Functional Water Networks in Fully Hydrated Photosystem II. J. Am. Chem. Soc. 2022, 144, 22035–22050. [Google Scholar] [CrossRef] [PubMed]
  177. Suga, M.; Shimada, A.; Akita, F.; Shen, J.-R.; Tosha, T.; Sugimoto, H. Time-resolved studies of metalloproteins using X-ray free electron laser radiation at SACLA. Biochim. Biophys. Acta 2020, 1864, 129466. [Google Scholar] [CrossRef]
  178. Mandal, M.; Saito, K.; Ishikita, H. Release of Electrons and Protons from Substrate Water Molecules at the Oxygen-Evolving Complex in Photosystem II. J. Phys. Soc. Jpn. 2022, 91, 091012. [Google Scholar] [CrossRef]
  179. Haumann, M.; Barra, M.; Loja, P.; Löscher, S.; Krivanek, R.; Grundmeier, A.; Andreasson, L.E.; Dau, H. Bromide does not bind to the Mn4Ca complex in its S1 state in Cl(-)-depleted and Br(-)-reconstituted oxygen-evolving photosystem II: Evidence from X-ray absorption spectroscopy at the Br K-edge. Biochemistry 2006, 45, 13101–13107. [Google Scholar] [CrossRef] [PubMed]
  180. Lindberg, K.; Andréasson, L.E. A one-site, two-state model for the binding of anions in photosystem II. Biochemistry 1996, 35, 14259–14267. [Google Scholar] [CrossRef]
  181. Olesen, K.; Andréasson, L.E. The function of the chloride ion in photosynthetic oxygen evolution. Biochemistry 2003, 42, 2025–2035. [Google Scholar] [CrossRef]
  182. Rivalta, I.; Amin, M.; Luber, S.; Vassiliev, S.; Pokhrel, R.; Umena, Y.; Kawakami, K.; Shen, J.-R.; Kamiya, N.; Bruce, D.; et al. Structural–Functional Role of Chloride in Photosystem II. Biochemistry 2011, 50, 6312–6315. [Google Scholar] [CrossRef] [PubMed]
  183. Service, R.J.; Hillier, W.; Debus, R.J. Evidence from FTIR Difference Spectroscopy of an Extensive Network of Hydrogen Bonds near the Oxygen-Evolving Mn4Ca Cluster of Photosystem II Involving D1-Glu65, D2-Glu312, and D1-Glu329. Biochemistry 2010, 49, 6655–6669. [Google Scholar] [CrossRef] [PubMed]
  184. Shimada, Y.; Sugiyama, A.; Nagao, R.; Noguchi, T. Role of D1-Glu65 in Proton Transfer during Photosynthetic Water Oxidation in Photosystem II. J. Phys. Chem. B 2022, 126, 8202–8213. [Google Scholar] [CrossRef] [PubMed]
  185. de Lichtenberg, C.; Kim, C.J.; Chernev, P.; Debus, R.J.; Messinger, J. The exchange of the fast substrate water in the S2 state of photosystem II is limited by diffusion of bulk water through channels—Implications for the water oxidation mechanism. Chem. Sci. 2021, 12, 12763–12775. [Google Scholar] [CrossRef] [PubMed]
  186. Junge, W.; Haumann, M.; Ahlbrink, R.; Mulkidjanian, A.; Clausen, J. Electrostatics and proton transfer in photosynthetic water oxidation. Philos. Trans. R Soc. Lond. B Biol. Sci. 2002, 357, 1407–1417; discussion 1417–1420. [Google Scholar] [CrossRef] [PubMed]
  187. Retegan, M.; Krewald, V.; Mamedov, F.; Neese, F.; Lubitz, W.; Cox, N.; Pantazis, D.A. A five-coordinate Mn(IV) intermediate in biological water oxidation: Spectroscopic signature and a pivot mechanism for water binding. Chem. Sci. 2016, 7, 72–84. [Google Scholar] [CrossRef] [PubMed]
  188. Bovi, D.; Narzi, D.; Guidoni, L. The S2 State of the Oxygen-Evolving Complex of Photosystem II Explored by QM/MM Dynamics: Spin Surfaces and Metastable States Suggest a Reaction Path Towards the S3 State. Angew. Chem. Int. Ed. 2013, 52, 11744–11749. [Google Scholar] [CrossRef] [PubMed]
  189. Kawashima, K.; Takaoka, T.; Kimura, H.; Saito, K.; Ishikita, H. O2 evolution and recovery of the water-oxidizing enzyme. Nat. Com. 2018, 9, 1247. [Google Scholar] [CrossRef]
  190. Bouges-Bocquet, B. Limiting steps in photosystem II and water decomposition in chlorella and spinach chloroplasts. Biochim. Biophys. Acta 1973, 292, 772–785. [Google Scholar] [CrossRef]
  191. Canaani, O.; Malkin, S.; Mauzerall, D. Pulsed photoacoustic detection of flash-induced oxygen evolution from intact leaves and its oscillations. Proc. Natl. Acad. Sci. USA 1988, 85, 4725–4729. [Google Scholar] [CrossRef]
  192. Etienne, A.L. Étude de l’étape thermique de l’émission photosynthétique d’oxygène par une méthode d’écoulement. Biochim. Biophys. Acta 1968, 153, 895–897. [Google Scholar] [CrossRef] [PubMed]
  193. Jursinic, P.A.; Dennenberg, R.J. Oxygen release time in leaf discs and thylakoids of peas and Photosystem II membrane fragments of spinach. Biochim. Biophys. Acta 1990, 1020, 195–206. [Google Scholar] [CrossRef]
  194. Lavergne, J. Mitochondrial responses to intracellular pulses of photosynthetic oxygen. Proc. Natl. Acad. Sci. USA 1989, 86, 8768–8772. [Google Scholar] [CrossRef] [PubMed]
  195. Mauzerall, D. Determination of Oxygen Emission and Uptake in Leaves by Pulsed, Time Resolved Photoacoustics. Plant Physiol. 1990, 94, 278–283. [Google Scholar] [CrossRef] [PubMed]
  196. Tang, X.S.; Moussavi, M.; Dismukes, G.C. Monitoring oxygen concentration in solution by ESR oximetry using lithium phthalocyanine: Application to photosynthesis. J. Am. Chem. Soc. 1991, 113, 5914–5915. [Google Scholar] [CrossRef]
  197. Strzalka, K.; Walczak, T.; Sarna, T.; Swartz, H.M. Measurement of Time-Resolved Oxygen Concentration Changes in Photosynthetic Systems by Nitroxide-Based EPR Oximetry. Arch. Biochem. Biophys. 1990, 281, 312–318. [Google Scholar] [CrossRef]
  198. de Wijn, R.; van Gorkom, H.J. Kinetics of Electron Transfer from QA to QB in Photosystem II. Biochemistry 2001, 40, 11912–11922. [Google Scholar] [CrossRef]
  199. Vinyard, D.J.; Ananyev, G.M.; Dismukes, G.C. Photosystem II: The Reaction Center of Oxygenic Photosynthesis. Annu. Rev. Biochem. 2013, 82, 577–606. [Google Scholar] [CrossRef] [PubMed]
  200. Bowes, J.; Crofts, A.R.; Arntzen, C.J. Redox Reactions on the Reducing Side of Photosystem II in Chloroplasts with Altered Herbicide Binding Properties. Arch. Biochem. Biophys. 1980, 200, 303–308. [Google Scholar] [CrossRef]
  201. Robinson, H.H.; Crofts, A.R. Kinetics of the oxidation-reduction reactions of the photosystem II quinone acceptor complex, and the pathway for deactivation. FEBS Lett. 1983, 153, 221–226. [Google Scholar] [CrossRef]
  202. Weiss, W.; Renger, G. UV-spectral characterization in Tris-washed chloroplasts of the redox component D1 which functionally connects the reaction center with the water-oxidizing enzyme system Y in photosynthesis. FEBS 1984, 169, 219–223. [Google Scholar] [CrossRef]
  203. Krivanek, R.; Kern, J.; Zouni, A.; Dau, H.; Haumann, M. Spare quinones in the QB cavity of crystallized photosystem II from Thermosynechococcus elongatus. Biochim. Biophys. Acta 2007, 1767, 520–527. [Google Scholar] [CrossRef] [PubMed]
  204. Whitmarsh, J.; Pakrasi, H.B. Form and Function of Cytochrome b-559. In Oxygenic Photosynthesis: The Light Reactions; Ort, D.R., Yocum, C.F., Heichel, I.F., Eds.; Springer: Dordrecht, The Netherlands, 1996; pp. 249–264. [Google Scholar]
  205. Shinopoulos, K.E.; Brudvig, G.W. Cytochrome b559 and cyclic electron transfer within photosystem II. Biochim. Biophys. Acta Bioenerg. 2012, 1817, 66–75. [Google Scholar] [CrossRef] [PubMed]
  206. Müh, F.; Zouni, A. Cytochrome b559 in Photosystem II. In Cytochrome Complexes: Evolution, Structures, Energy Transduction, and Signaling; Cramer, W.A., Kallas, T., Eds.; Springer: Dordrecht, The Netherlands, 2016; pp. 143–175. [Google Scholar]
  207. Falkowski, P.G.; Fujita, Y.; Ley, A.; Mauzerall, D. Evidence for Cyclic Electron Flow around Photosystem II in Chlorella pyrenoidosa. Plant Physiol. 1986, 81, 310–312. [Google Scholar] [CrossRef] [PubMed]
  208. Feyziyev, Y.; Deák, Z.; Styring, S.; Bernát, G. Electron transfer from Cyt b559 and tyrosine-D to the S2 and S3 states of the water oxidizing complex in photosystem II at cryogenic temperatures. J. Bioenerg. Biomembr. 2013, 45, 111–120. [Google Scholar] [CrossRef] [PubMed]
  209. Zournas, A.; Mani, K.; Dismukes, G.C. Cyclic electron flow around photosystem II in silico: How it works and functions in vivo. Photosynth. Res. 2023, 156, 129–145. [Google Scholar] [CrossRef]
  210. Gates, C.; Ananyev, G.; Roy-Chowdhury, S.; Fromme, P.; Dismukes, G.C. Regulation of light energy conversion between linear and cyclic electron flow within photosystem II controlled by the plastoquinone/quinol redox poise. Photosynth. Res. 2023, 156, 113–128. [Google Scholar] [CrossRef] [PubMed]
  211. 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]
  212. Sugiura, M.; Tibiletti, T.; Takachi, I.; Hara, Y.; Kanawaku, S.; Sellés, J.; Boussac, A. Probing the role of Valine 185 of the D1 protein in the Photosystem II oxygen evolution. Biochim. Biophys. Acta Bioenerg. 2018, 1859, 1259–1273. [Google Scholar] [CrossRef]
  213. Pokhrel, R.; Debus, R.J.; Brudvig, G.W. Probing the Effect of Mutations of Asparagine 181 in the D1 Subunit of Photosystem II. Biochemistry 2015, 54, 1663–1672. [Google Scholar] [CrossRef]
  214. Dilbeck, P.L.; Hwang, H.J.; Zaharieva, I.; Gerencser, L.; Dau, H.; Burnap, R.L. The D1-D61N Mutation in Synechocystis sp. PCC 6803 Allows the Observation of pH-Sensitive Intermediates in the Formation and Release of O2 from Photosystem II. Biochemistry 2012, 51, 1079–1091. [Google Scholar] [CrossRef] [PubMed]
  215. Taguchi, S.; Shen, L.; Han, G.; Umena, Y.; Shen, J.-R.; Noguchi, T.; Mino, H. Formation of the High-Spin S2 State Related to the Extrinsic Proteins in the Oxygen Evolving Complex of Photosystem II. J. Phys. Chem. Lett. 2020, 11, 8908–8913. [Google Scholar] [CrossRef] [PubMed]
  216. Dilbeck, P.L.; Bao, H.; Neveu, C.L.; Burnap, R.L. Perturbing the water cavity surrounding the manganese cluster by mutating the residue D1-Valine 185 has a strong effect on the water oxidation mechanism of Photosystem II. Biochemistry 2013, 52, 6824–6833. [Google Scholar] [CrossRef] [PubMed]
  217. de Lichtenberg, C.; Avramov, A.P.; Zhang, M.; Mamedov, F.; Burnap, R.L.; Messinger, J. The D1-V185N mutation alters substrate water exchange by stabilizing alternative structures of the Mn4Ca-cluster in photosystem II. Biochim. Biophys. Acta Bioenerg. 2021, 1862, 148319. [Google Scholar] [CrossRef] [PubMed]
  218. Kaur, D.; Szejgis, W.; Mao, J.; Amin, M.; Reiss, K.M.; Askerka, M.; Cai, X.; Khaniya, U.; Zhang, Y.; Brudvig, G.W.; et al. Relative stability of the S2 isomers of the oxygen evolving complex of photosystem II. Photosynth. Res. 2019, 141, 331–341. [Google Scholar] [CrossRef] [PubMed]
  219. Oyala, P.H.; Stich, T.A.; Debus, R.J.; Britt, R.D. Ammonia Binds to the Dangler Manganese of the Photosystem II Oxygen Evolving Complex. J. Am. Chem. Soc. 2015, 137, 8829–8837. [Google Scholar] [CrossRef] [PubMed]
  220. Lubitz, W.; Chrysina, M.; Cox, N. Water oxidation in photosystem II. Photosynth. Res. 2019, 142, 105–125. [Google Scholar] [CrossRef] [PubMed]
  221. Kusunoki, M. Mono-manganese mechanism of the photosytem II water splitting reaction by a unique Mn4Ca cluster. Biochim. Biophys. Acta 2007, 1767, 484–492. [Google Scholar] [CrossRef] [PubMed]
  222. Corry, T.A.; O’Malley, P.J. Evidence of O-O Bond Formation in the Final Metastable S3 State of Nature’s Water Oxidising Complex Implies a Novel Mechanism of Water Oxidation. J. Phys. Chem. Lett. 2018, 9, 6269–6274. [Google Scholar] [CrossRef]
  223. Mandal, M.; Saito, K.; Ishikita, H. The Nature of the Short Oxygen−Oxygen Distance in the Mn4CaO6 Complex of Photosystem II Crystals. J. Phys. Chem. Lett. 2020, 11, 10262–10268. [Google Scholar] [CrossRef]
  224. Pérez-Navarro, M.; Neese, F.; Lubitz, W.; Pantazis, D.A.; Cox, N. Recent developments in biological water oxidation. Curr. Opin. Chem. Biol. 2016, 31, 113–119. [Google Scholar] [CrossRef] [PubMed]
  225. Pushkar, Y.; Davis, K.M.; Palenik, M.C. Model of the Oxygen Evolving Complex Which is Highly Predisposed to O–O Bond Formation. J. Phys. Chem. Lett. 2018, 9, 3525–3531. [Google Scholar] [CrossRef] [PubMed]
  226. Renger, G. Mechanism of light induced water splitting in Photosystem II of oxygen evolving photosynthetic organisms. Biochim. Biophys. Acta 2012, 1817, 1164–1176. [Google Scholar] [CrossRef] [PubMed]
  227. Shoji, M.; Isobe, H.; Yamaguchi, K. QM/MM Study of the S2 to S3 Transition Reaction in the Oxygen-Evolving Complex of Photosystem II. Chem. Phys. Lett. 2015, 636, 172–179. [Google Scholar] [CrossRef]
  228. Siegbahn, P.E.M. A Structure-Consistent Mechanism for Dioxygen Formation in Photosystem II. Eur. J. Chem. 2008, 14, 8290–8302. [Google Scholar] [CrossRef] [PubMed]
  229. Hatakeyama, M.; Ogata, K.; Fujii, K.; Yachandra, V.K.; Yano, J.; Nakamura, S. Structural changes in the S3 state of the oxygen evolving complex in photosystem II. Chem. Phys. Lett. 2016, 651, 243–250. [Google Scholar] [CrossRef] [PubMed]
  230. Guo, Y.; Messinger, J.; Kloo, L.; Sun, L. Alternative Mechanism for O2 Formation in Natural Photosynthesis via Nucleophilic Oxo–Oxo Coupling. J. Am. Chem. Soc. 2023, 145, 4129–4141. [Google Scholar] [CrossRef] [PubMed]
  231. Burnap, R.L.; Shen, J.R.; Jursinic, P.A.; Inoue, Y.; Sherman, L.A. Oxygen yield and thermoluminescence characteristics of a cyanobacterium lacking the manganese-stabilizing protein of photosystem II. Biochemistry 1992, 31, 7404–7410. [Google Scholar] [CrossRef] [PubMed]
  232. Campbell, K.A.; Gregor, W.; Pham, D.P.; Peloquin, J.M.; Debus, R.J.; Britt, R.D. The 23 and 17 kDa extrinsic proteins of photosystem II modulate the magnetic properties of the S1-state manganese cluster. Biochemistry 1998, 37, 5039–5045. [Google Scholar] [CrossRef] [PubMed]
  233. Enami, I.; Okumura, A.; Nagao, R.; Suzuki, T.; Iwai, M.; Shen, J.R. Structures and functions of the extrinsic proteins of photosystem II from different species. Photosynth. Res. 2008, 98, 349–363. [Google Scholar] [CrossRef]
  234. Tomita, M.; Ifuku, K.; Sato, F.; Noguchi, T. FTIR evidence that the PsbP extrinsic protein induces protein conformational changes around the oxygen-evolving Mn cluster in photosystem II. Biochemistry 2009, 48, 6318–6325. [Google Scholar] [CrossRef] [PubMed]
  235. Nagao, R.; Tomo, T.; Noguchi, T. Effects of extrinsic proteins on the protein conformation of the oxygen-evolving center in cyanobacterial photosystem II as revealed by Fourier transform infrared spectroscopy. Biochemistry 2015, 54, 2022–2031. [Google Scholar] [CrossRef] [PubMed]
  236. Offenbacher, A.R.; Polander, B.C.; Barry, B.A. An intrinsically disordered photosystem II subunit, PsbO, provides a structural template and a sensor of the hydrogen-bonding network in photosynthetic water oxidation. J. Biol. Chem. 2013, 288, 29056–29068. [Google Scholar] [CrossRef] [PubMed]
  237. Boussac, A.; Setif, P.; Rutherford, A.W. Inhibition of tyrosine Z photooxidation after formation of the S3-state in calcium-depleted and chloride-depleted photosystem-II. Biochemistry 1992, 31, 1224–1234. [Google Scholar] [CrossRef] [PubMed]
  238. Saito, K.; Rutherford, A.W.; Ishikita, H. Mechanism of tyrosine D oxidation in Photosystem II. Proc. Natl. Acad. Sci. USA 2013, 110, 7690–7695. [Google Scholar] [CrossRef] [PubMed]
  239. van Vliet, P.; Rutherford, A.W. Properties of the chloride-depleted oxygen-evolving complex of photosystem II studied by electron paramagnetic resonance. Biochemistry 1996, 35, 1829–1839. [Google Scholar] [CrossRef] [PubMed]
  240. Wincencjusz, H.; van Gorkom, H.J.; Yocum, C.F. The photosynthetic oxygen evolving complex requires chloride for its redox state S2 → S3 and S3 → S0 transitions but not for S0 → S1 or S1 → S2 transitions. Biochemistry 1997, 36, 3663–3670. [Google Scholar] [CrossRef]
  241. Packham, N.K.; Hodges, M.; Etienne, A.L.; Briantais, J.M. Changes in the flash-induced oxygen yield pattern by thylakoid membrane phosphorylation. Photosynth. Res. 1988, 15, 221–232. [Google Scholar] [CrossRef] [PubMed]
  242. Shinkarev, V.P. Flash-Induced Oxygen Evolution in Photosynthesis: Simple Solution for the Extended S-State Model that includes Misses, Double-Hits, Inactivation, and Backward-Transitions. Biophys. J. 2005, 88, 412–421. [Google Scholar] [CrossRef]
  243. Isgandarova, S.; Renger, G.; Messinger, J. Functional differences of photosystem II from Synechococcus elongatus and spinach characterized by flash induced oxygen evolution patterns. Biochemistry 2003, 42, 8929–8938. [Google Scholar] [CrossRef]
  244. Vinyard, D.J.; Zachary, C.E.; Ananyev, G.; Dismukes, G.C. Thermodynamically accurate modeling of the catalytic cycle of photosynthetic oxygen evolution: A mathematical solution to asymmetric Markov chains. Biochim. Biophys. Acta 2013, 1827, 861–868. [Google Scholar] [CrossRef] [PubMed]
  245. Messinger, J.; Schroeder, W.P.; Renger, G. Structure-function relations in photosystem II. Effects of temperature and chaotropic agents on the period four oscillation of flash-induced oxygen evolution. Biochemistry 1993, 32, 7658–7668. [Google Scholar] [CrossRef] [PubMed]
  246. Pham, L.V.; Messinger, J. Probing S-state advancements and recombination pathways in photosystem II with a global fit program for flash-induced oxygen evolution pattern. Biochim. Biophys. Acta 2016, 1857, 848–859. [Google Scholar] [CrossRef] [PubMed]
Figure 2. Scheme of water channels identified in a higher plant Spinacia oleracea [101]. The location of conserved amino acids [69,89,90,101,112] are indicated (magenta from the D1 core protein; gray—from the D2 core protein and blue—from the inner light-harvesting complex CP43 protein). Areas of the channels partially stabilized by the outer proteins PsbQ, PsbP, and PsbO are highlighted in light red, blue, and green, respectively. The Mn4CaO5 cluster is also included in the figure. The purple circles represent manganese atoms, the dark red circles oxygen atoms and the light red circles oxygen atoms from bound water. All are numbered according to generally accepted convention. The calcium atom is marked in yellow.
Figure 2. Scheme of water channels identified in a higher plant Spinacia oleracea [101]. The location of conserved amino acids [69,89,90,101,112] are indicated (magenta from the D1 core protein; gray—from the D2 core protein and blue—from the inner light-harvesting complex CP43 protein). Areas of the channels partially stabilized by the outer proteins PsbQ, PsbP, and PsbO are highlighted in light red, blue, and green, respectively. The Mn4CaO5 cluster is also included in the figure. The purple circles represent manganese atoms, the dark red circles oxygen atoms and the light red circles oxygen atoms from bound water. All are numbered according to generally accepted convention. The calcium atom is marked in yellow.
Cimb 46 00428 g002
Figure 3. The experimental protocol for oxygen evolution experiments. The time separation between the 1st and the 2nd (Δt1–2) or the 2nd and the 3rd flash (Δt2–3) varied from 5 m to 500 m.
Figure 3. The experimental protocol for oxygen evolution experiments. The time separation between the 1st and the 2nd (Δt1–2) or the 2nd and the 3rd flash (Δt2–3) varied from 5 m to 500 m.
Cimb 46 00428 g003
Figure 4. Example raw data for PSII BBYcontrol for Δt2–3 = 120 m. The experimental conditions were as described in the text.
Figure 4. Example raw data for PSII BBYcontrol for Δt2–3 = 120 m. The experimental conditions were as described in the text.
Cimb 46 00428 g004
Table 1. Transition parameters and the initial Si-state distribution estimated according to the 5S-state model [37] (see Equation (S3) and the discussion in Section 2) for the PSII BBY control sample and PSII BBY depleted of two or three external proteins. Theoretical data are presented in Figure 5 (red-filled circles).
Table 1. Transition parameters and the initial Si-state distribution estimated according to the 5S-state model [37] (see Equation (S3) and the discussion in Section 2) for the PSII BBY control sample and PSII BBY depleted of two or three external proteins. Theoretical data are presented in Figure 5 (red-filled circles).
Parametersα0α1α2α3dS0S1S2S3Cpfq
PSII BBY control0.0010.0010.7850.0010.850.050.870.080.000.9950.0054
PSII BBY—P,Q0.550.280.780.120.300.040.880.080.000.9670.0165
PSII BBY—O,P,Q0.510.270.790.160.020.2820.5300.1300.0580.9900.0139
Parameters: αi—the failure rate of the trapping centers (called misses); d—the contribution of the fast mode to the O2 yield; Si—the initial distribution of the manganese states in the OEC; C—the parameter describing the fraction of active photosystems (the quenching parameter due to not sufficient amount of quinone acceptors); pfq—the parameter of fit quality described in Supplementary Materials, Equation (S5).
Table 2. Fitted parameters Afast, Aslow, τfast, and τslow related to the amplitudes and time constants characterizing the fast and slow mode of O2 release (using Equation (1)).
Table 2. Fitted parameters Afast, Aslow, τfast, and τslow related to the amplitudes and time constants characterizing the fast and slow mode of O2 release (using Equation (1)).
ParametersAfastτfast [ms]Aslowτslow [ms]
PSII BBY control0.73 ± 0.084.1 ± 1.80.27 ± 0.0844.2 ± 14.7
PSII BBY—P,Q0.26 ± 0.096.2 ± 3.60.74 ± 0.0622.2 ± 4.4
PSII BBY—O,P,Q0-------1.00 ± 0.0313 ± 2
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

Krysiak, S.; Burda, K. The Effect of Removal of External Proteins PsbO, PsbP and PsbQ on Flash-Induced Molecular Oxygen Evolution and Its Biphasicity in Tobacco PSII. Curr. Issues Mol. Biol. 2024, 46, 7187-7218. https://doi.org/10.3390/cimb46070428

AMA Style

Krysiak S, Burda K. The Effect of Removal of External Proteins PsbO, PsbP and PsbQ on Flash-Induced Molecular Oxygen Evolution and Its Biphasicity in Tobacco PSII. Current Issues in Molecular Biology. 2024; 46(7):7187-7218. https://doi.org/10.3390/cimb46070428

Chicago/Turabian Style

Krysiak, Sonia, and Kvetoslava Burda. 2024. "The Effect of Removal of External Proteins PsbO, PsbP and PsbQ on Flash-Induced Molecular Oxygen Evolution and Its Biphasicity in Tobacco PSII" Current Issues in Molecular Biology 46, no. 7: 7187-7218. https://doi.org/10.3390/cimb46070428

Article Metrics

Back to TopTop