Significance
Photosynthesis—the biological process via which solar energy is stored in the form of energy-rich molecules—fuels life on Earth and provides the molecular oxygen we breathe. The crucial starting point for this reaction is the splitting of water, which is carried out by a unique catalyst in Photosystem II. Unraveling the details of this reaction provides the blueprint for how to extract protons and electrons from water using abundant and cheap metal catalysts—a pre-requisite for the sustainable production of green fuels and chemicals. In this study, we identify a key feature of nature’s water-splitting unit, the binding site of one of the two water molecules involved in making O2.
Keywords: photosynthesis, photosystem II, water oxidation mechanism, membrane inlet mass spectrometry (MIMS), electron paramagnetic resonance (EPR)
Abstract
Identifying the two substrate water sites of nature’s water-splitting cofactor (Mn4CaO5 cluster) provides important information toward resolving the mechanism of O-O bond formation in Photosystem II (PSII). To this end, we have performed parallel substrate water exchange experiments in the S1 state of native Ca-PSII and biosynthetically substituted Sr-PSII employing Time-Resolved Membrane Inlet Mass Spectrometry (TR-MIMS) and a Time-Resolved 17O-Electron-electron Double resonance detected NMR (TR-17O-EDNMR) approach. TR-MIMS resolves the kinetics for incorporation of the oxygen-isotope label into the substrate sites after addition of H218O to the medium, while the magnetic resonance technique allows, in principle, the characterization of all exchangeable oxygen ligands of the Mn4CaO5 cofactor after mixing with H217O. This unique combination shows i) that the central oxygen bridge (O5) of Ca-PSII core complexes isolated from Thermosynechococcus vestitus has, within experimental conditions, the same rate of exchange as the slowly exchanging substrate water (WS) in the TR-MIMS experiments and ii) that the exchange rates of O5 and WS are both enhanced by Ca2+→Sr2+ substitution in a similar manner. In the context of previous TR-MIMS results, this shows that only O5 fulfills all criteria for being WS. This strongly restricts options for the mechanism of water oxidation.
Nature’s water splitting catalyst, a tetra-manganese penta-oxygen calcium cofactor (Mn4CaO5; Fig. 1A) is found in a unique membrane protein complex, Photosystem II (PSII) (1–4). The catalytic cycle of the cofactor is comprised of five distinct redox intermediates, the Sn states, where the subscript indicates the number of stored oxidizing equivalents (n = 0 to 4) required to split two water molecules and release molecular oxygen (5) (Fig. 1B). Importantly, each Sn state transition is multi-step, with the cofactor’s oxidation coupled to its deprotonation into the bulk and conformational changes (with the exception of the S1 to S2 transition) (6–15). Sn state progression is driven by the reaction center of PSII, which is a multi-chlorophyll/pheophytin pigment assembly. Light absorption and subsequent charge separation generate an in situ photo-oxidant (P680•+), coupled to the Mn4CaO5 cofactor via an intermediary redox-active tyrosine residue, YZ (Fig. 1C) (1). After four charge separation events, the transiently formed [S4] state decays to the S0 state with the concomitant release of molecular triplet oxygen and rebinding of one substrate water molecule (14, 15). The term substrate water is used irrespective of the protonation state of the bound oxygen species (16–19).
Fig. 1.
(A) The water oxidizing Mn4O5Ca cofactor with selected ligands and water channels that regulate water access and proton egress. Mn ions are displayed in purple and labeled 1 to 4, oxygen bridges in red and numbered O1 to O5, Ca in yellow, and terminal water/hydroxide ligands are named W1-W4 (based on PDB 6W1O). (B) Catalytic (Kok) cycle of the water oxidation reaction (5). Sn states (n = 0,…,4) indicate oxidation state changes of the Mn4CaO5/6 cluster. Light-induced electron and proton removal as well as water binding events are indicated. The moon symbol indicates that the S1 state is the dark-stable state, and the plus sign an extra positive charge due to lack of proton release from S1 to S2. (C) Schematic view of the Mn4CaO6 cluster and the tyrosine Z (YZ) – His190 pair in the S3 state. The additional oxygen bridge between Ca and Mn1, inserted during the S2→S3 transition, is known as Ox or O6 (12, 20). YZ acts as electron relay between the Mn4CaO5/6 cluster and P680•+. (D) Proposed equilibria between conformations of the Mn4CaO5 cluster in its S1 state (21). Mn ions are represented by Roman numbers indicating their oxidation state, while oxygen bridges are depicted in red. The red arrow indicates the direction of the Jan–Teller (JT) axis on the Mn4 ion. Below shows one potential pathway for O5 exchange in the S1 state via water binding to the Mn1 (22). (E) Results of substrate water exchange measurements by membrane-inlet mass spectrometry (MIMS) in the S3 state of PSII core complexes from T. vestitus at m/z 34 (16O18O) and m/z 36 (18O18O). The m/z 34 trace displays a biphasic rise reflecting the rates of the fast (Wf) and slowly (WS) exchanging substrate waters, while the rise of m/z 36 is limited by the exchange of WS (data replotted from ref. 23). (F) 17O EDNMR spectra (black line) obtained at W-band with highly concentrated PSII samples in absence (Top) and after addition (Bottom) of ammonia. The blue line signifies the 14N coupling arising from the D1-His332 ligand of Mn1, see panel A (24). The red dotted line gives the four component fit of the signal (see text). The figure is adapted from ref. 25.
The dark-stable (S1 state) structure of the water splitting cofactor resembles a “distorted chair” with Mn1, Mn2, Mn3, and the Ca2+ ion making up the base of the chair, or the open cubane unit as there is no bond between Mn1 and O5 (12, 26, 27) (Fig. 1 A and D). The outer Mn4 is linked to the open cubane unit via the additional oxygen bridge O4 to Mn3 and by binding to the O5 bridge of the open cubane forming the back of the chair. In the S1 state, the oxidation states of the four Mn ions are III, IV, IV, and III (Mn1 to Mn4) (28–30). Recently, two structural isomers were proposed for the S1 state (21), S1A, and S1B, that are distinguished by the orientation of the Jahn–Teller axis at the outer Mn4(III) ion (Fig. 1D). All Mn ions are six coordinate (octahedral) with the important exception of Mn1, which is five-coordinate (square pyramidal), allowing for binding of one water molecule to Mn in the S2→S3 transition (Fig. 1B) as well as O5 exchange (Fig. 1D) (12, 18, 20, 22, 31–38). The additional water-derived ligand in the S3 state forms a bridge between Mn1 and Ca (Fig. 1C) and is known as Ox or O6 (11, 12).
Two substrate waters are required to form O2 in the S3→S4→S0 transition. The crystal structures of PSII (26) resolve seven candidates at the Mn4CaO5 cluster: i) the waters bound to the Ca2+ ion (W3, and W4) which form the endpoint of the O1 water channel; ii) the water-derived ligands of the outer Mn4 (W1 and W2) and possibly the oxygen bridge O4 which are connected to the bulk via the Cl1 and the O4 channels; and finally iii) the oxygen bridges O5, and, in the S3 state, Ox/O6 (Fig. 1 A and C) (11, 12, 14, 15, 18, 36, 39–41).
The two substrate waters can be directly monitored using time-resolved membrane inlet mass spectrometry (TR-MIMS). In this experiment, dark-adapted PSII samples are first advanced to the desired Si state using single turn-over flashes. Then, H218O is injected and after various incubation times, O2 evolution is induced by a sequence of closely spaced flashes. Finally, the isotopic composition of O2 is determined by isotope ratio mass spectrometry (16, 18). As the rate of induction of the singly (16O18O) and doubly (18O18O) labeled products are significantly different (Fig. 1E), the two bound substrates must be chemically distinct. The two substrates are thus referred to as the slowly exchanging substrate (Water slow – WS) and the fast-exchanging substrate (Water fast – Wf) (16). Rates for WS exchange have been measured for all Sn states and vary for Thermosynechococcus (T.) vestitus BP-1 (previously T. elongatus BP-I) PSII core preparations between 0.4 and 1.1 s−1, whereas rates for Wf are only well-defined for the higher Sn states (S2, S3) and are much faster, see Fig. 1E (18, 23, 38). In the S2 state, the exchange of Wf is about 100 s−1 and was recently shown to be limited by diffusion of bulk and internal water molecules through the channels leading to the Mn4CaO5 cofactor (40). Importantly, in the S3 state, the exchange of Wf is slower (about 40 s−1) and the exchange of both substrates is arrested in the S3YZ• state (23). Thus, in S3 the exchange rate of Wf is no longer diffusion limited and hence it is unlikely to be bound to Ca2+ as terminal ligand as the rate of ligand exchange for the solvated Ca2+ cation is on the order of 108 s−1 and thereby much too fast (42). Chemical modification of the cofactor—substitution of the Ca2+ ion with Sr2+, either by chemical exchange or biosynthetic incorporation—enhances the rate of WS exchange but leaves Wf unchanged, suggesting WS is somehow associated with the Ca2+ site (23, 38, 43). As terminal Ca ligands are excluded, and since an Sn state dependence of the WS exchange rate is observed, it was proposed that WS could represent an oxygen bridge between the Ca2+ ion and a neighboring Mn ion (39, 43). Specifically, the central oxo-bridge was proposed in ref. 39.
For obtaining structural information on exchangeable oxygen ligands a spectroscopic tool is required that is isotope sensitive. Here, in principle, EPR methods are especially suitable owing to their ability to probe the interaction of nuclear spins with the EPR signals of the catalytic site. However, the characterization of oxygen ligands of metallocofactors is challenging owing to the NMR active 17O nucleus having a high nuclear spin (I = 5/2, quadrupolar nucleus), a small nuclear g-value, and a low natural abundance, with only a handful of published studies using conventional pulse EPR methods (44–48). As a consequence, we developed high field (94 GHz, W-band) 17O-Electron-electron Double resonance detected NMR (17O-EDNMR), as a sensitive assay for their detection (18, 49). In the high field regime (W-band, 3.4 T), 17O ligands of transition metal complexes appear about the characteristic (Larmor) frequency split by the electron-nuclear hyperfine coupling. At this magnetic field, the Larmor frequency of 17O [ν(17O) = 19.6 MHz at 3.4 T] is sufficiently different from background nuclei [e. g. ν(14N) = 10.5 MHz at 3.4 T] that they can be unambiguously identified (49). The electron-nuclear hyperfine interaction (signal width) can be used to differentiate between different types of 17O ligands (O2−, OH, OH2) (25, 49–51).
Typical EDNMR spectra are shown in Fig. 1F (49). Two species are readily observed: i) a background doublet centered at ν(14N) assigned to the only nitrogen ligand of the cofactor, His332 (Fig. 1A) (24), and a broader envelope centered at ν(17O) assigned to exchanged oxygen sites of the cofactor (49). In our original study, the profile was fitted as three distinct signals: i) a broad signal (splitting ≈10 MHz; red in Fig. 1F), which describes the width of the envelope (49). This was assigned to a single, exchanged oxygen bridge (O5), based on comparisons to model systems, and by chemical modification of the cofactor (25, 49–51); ii) a narrow matrix signal (≈1 MHz; dark gray), assigned to exchanged water ligands of the cofactor, dominantly W1; and iii) a third signal of intermediate width (≈4 MHz; light gray) that is hidden under the signal envelope but is better resolved in the double quantum region. This was assigned to a terminal hydroxide ligand of Mn4, W2 (49). Treatment of PSII with ammonia confirms this basic description (lower part of Fig. 1F). NH3 displaces W1 leading to the suppression of the 17O matrix signal, with the central peak broadening as it now has a larger contribution from W2 (25).
Importantly, our initial report demonstrated that all terminal ligands of the cofactor (W1-W4) and O5 exchange within 10 to 15 s with bulk solvent but obtaining exchange rates remained impossible (49). Nevertheless, this structural information from EDNMR further supported our original suggestion made based on the kinetic information from TR-MIMS (39) and point to O5 being WS. This was important, since the exchange of an oxygen bridge at such a rate was contested, because in model complexes the rates are much slower owing to the acidity of the μ-oxo bridge motif (51, 52). The reason why O5 likely exchanges rapidly is its conformational flexibility and the neighboring open coordination site at Mn1 that allows extra water to bind and to bring O5, in fully protonated form, to a terminal position for water exchange (Fig. 1D) (22, 38).
While EDNMR and MIMS experiments thus far reported strongly point to O5 being the slowly exchanging substrate, they do not conclusively exclude all other water-derived Mn ligands. As such, in the literature, the data have been interpreted to be still consistent with W2 being the WS (53–58) (for discussion see also ref. 38). To resolve this question, we have developed time-resolved EDNMR (TR-EDNMR) that is similar to TR-MIMS (Fig. 2A) and allows comparing the exchange rates for O5 and WS. Here we demonstrate that these rates are highly similar in Ca-PSII core complexes from T. vestitus, and that Ca2+/Sr2+ substitution significantly enhances the rates of both O5 and WS exchange. Thus, our present data strongly endorse the assignment of the central O5 bridge of the Mn4CaO5 cluster as the slowly exchanging substrate for photosynthetic dioxygen production.
Fig. 2.
(A) Experimental sequences of the water exchange experiments in the S1 state of the Mn4CaO5 cluster employing time-resolved membrane-inlet mass spectrometry (TR-MIMS; Top) and time-resolved ELDOR detected NMR (TR-EDNMR; Bottom). In both cases, isotope-labeled water (H218O resp. H217O) is rapidly mixed with the dark-adapted sample. The degree of exchange of the substrate bound at the Mn4CaO5 cluster with bulk water is then probed after various delay times (Δt). In MIMS, three flashes are given to produce O2, which is then analyzed by isotope ratio mass spectrometry, while in EDNMR the exchange process is stopped by freezing and the incorporation of the label is probed by EDNMR after low-temperature (198 K) illumination into the S2 state. (B) Experimental set-up of the rapid freeze quench system (based on a Biologic QFM 400) for mixing H217O (blue) with PSII (green) to study water exchange kinetics at the Mn4CaO5 cluster in PSII. Specifications: 110 ms deadtime (for PSII), 15 µL sample volume, time resolution determined by delay line (magenta), sample collection on a cold aluminum surface of the collection unit in contact with liquid N2, followed by packing into an EPR tube through the central hole of the collection unit.
Results
Time-resolved H217O Exchange with Freeze-quench.
To determine the rate of O5 exchange we measured the induction of the TR-17O-EDNMR envelope as a function of incubation in labeled water employing a purpose-built micro-rapid freeze quench (μRFQ) apparatus (Fig. 2B), which is described in detail in SI Appendix, Figs. S1–S7 and Text S1), along with calibration data. The system allows the mixing and collection of dark-adapted PSII (S1 state) with 17O-labeled water, with sample volumes as little as 15 μL, minimizing sample waste—an important consideration when using the expensive 17O-labeled water. The ejected sample was frozen on a liquid nitrogen–cooled aluminum surface (funnel), below which EPR capillaries were located. The sample was transferred to the capillary through a hole in the center of the aluminum funnel using a thin rod. It should be noted that the effective concentration of PSII in our μRFQ samples was lower than what we could typically achieve using our standard loading approach, owing to the lower packing efficiency of our capillaries with pre-frozen sample. To combat this, we switched to a higher Q (narrower bandwidth) resonator for all EDNMR measurements, improving overall S/N, but excluding measurements of 17O double quantum transitions. Importantly, we can quantify the packing efficiency of each sample tube on the basis of the S2 EPR multiline signal induced by 200 K illumination. In this way, we can normalize the entire dataset and thereby compensate for possible sample concentration variations across the set of timepoints.
Detection of the oxygen exchange by EDNMR occurred after a 200 K illumination that quantitatively transferred the sample into the S2 state (Fig. 2A), which exhibits the S2 EPR multiline signal. It is important to note that the signal intensities in the EDNMR experiment are dependent on experimental conditions—and so are not strictly quantitative (59, 60). The length and amplitude of the first microwave pulse of the EDNMR sequence, termed the high turning angle pulse (HTA)—has optimal values for each 17O species (18). A short/low amplitude HTA pulse will amplify the species with the largest hyperfine coupling e.g., in this instance O5 and W2, whereas a long/high amplitude pulse will amplify species with smaller hyperfine couplings, e.g., the background 14N signal and in particular the matrix signal (W1). That said, as long as experimental conditions are kept constant, then the relative intensity of one signal relative to another is robust allowing kinetic data to be collected (59).
Fig. 3A shows a set of selected TR-17O-EDNMR traces collected using different mixing times for PSII core complexes with the natural Ca in the water oxidation complex (Ca-PSII) and of one trace of a sample where Ca was exchanged against Sr (Sr-PSII) (see SI Appendix, Fig. S9 for the complete time course of the Sr sample and a replicate dataset for Ca-PSII). The top trace shows the 17O envelope seen in Ca-PSII following 60 s incubation in H217O. This length of time is longer than required for complete exchange as can be seen from the very similar envelopes obtained at 8 s and 4 s. The spectral profile is essentially the same as reported earlier (Fig. 1F). As compared to these earlier studies (49), the 14N and the central 17O peak are somewhat suppressed relative to the O5 bridge signal due to selecting an initial HTA pulse that maximizes the O5 signal.
Fig. 3.
(A) Time-resolved 17O ELDOR-detected NMR (TR-17O-EDNMR) spectra obtained after incubating PSII core complexes from T. vestitus at room temperature (22 °C) and pH 6.5 for various times with H217O. The first seven traces were obtained with Ca-PSII (60 s to 0.5 s) and the bottom trace with Sr-PSII (0.4 s). Additional spectra are shown in SI Appendix, Fig. S9. The integrated intensities of the O5 resonances are indicated (red), and the matrix 17O line (gray). Experimental parameters are listed in Materials and Methods and the SI Appendix, Text S2. (B) Exchange time dependence of the integrated O5 resonance in the 17O EDNMR spectrum (red area in panel A) for Ca-PSII (blue) and Sr-PSII (green). The blue-dashed line is a mono-exponential fit of the Ca-PSII data yielding a rate of 0.9 s−1. Error bars were determined by the integrated fitting error, normalized to the overall integral value in the fitting region. (C) Normalized 36O2 flash yields of Ca-PSII (blue) and Sr-PSII (green) core complexes of T. vestitus induced by three saturating flashes after different exchange times with H218O at 20 °C and pH 6.5. Each dot represents a separate measurement. The dashed lines represent fits of the exchange rates of the slow substrate water (WS). A mono-exponential rise with 0.4 s−1 was found for Ca-PSII, while a bi-exponential fit with 50% unresolved exchange and a rate of 1.9 s−1 for the slower fraction was obtained for the Sr-PSII data.
To probe the exchange rate of O5, a series of TR-17O-EDNMR spectra were collected for Ca-PSII after successively shorter incubation times with H217O, down to 0.5 s. The results show a successive decline of the integral O5 EDNMR signal (red) at incubation times shorter than 4 s, such that the O5 EDNMR signal is strongly diminished at 0.5 s (Fig. 3A). By contrast, the 14N signal, stemming from the D1His332 ligand to Mn1 (24), is observed at approximately the same intensity throughout the time course.
The trace at the bottom of Fig. 3A shows the 17O envelope seen in Sr-PSII following a 0.4 s incubation with H217O. In line with the direct ligation of O5 with Ca/Sr, the O5 signal is slightly narrower (10%) in Sr-PSII as compared to that obtained with Ca-PSII samples, as seen in our previous study (50). Remarkably, the O5 EDNMR signal intensity at 0.4 s is already similar to that in the 60 s trace of Ca-PSII, and to longer incubation times in Sr-PSII (SI Appendix, Fig. S9). This shows that the O5 exchange is significantly faster in Sr-PSII than in Ca-PSII.
For both Ca-PSII and Sr-PSII, the central peak, marked in gray, contains information on the exchange rates of W1 and W2 bound to Mn4 (Fig. 1A). This central line decreases somewhat at shorter incubation times. However, at this point, we cannot accurately disentangle the exchange of these two water ligands, since the central peak has approximately the same shape across the entire time course (see also Ca-PSII dataset2 in SI Appendix, Fig. S9). Reliable analysis is also complicated by the small signal intensity of the central line relative to the O5 signal leading to significant scatter and its variation between datasets. Thus, to characterize these two species (W1 and W2), we need to collect data with much shorter exchange times, and likely switch to 17O-ENDOR (51) to reliably decompose the central peak into two components. Both potential improvements are presently challenging to implement (SI Appendix, S1).
Comparison of O5 and WS Exchange Rates.
Fig. 3B shows the time course for the induction of the O5 EDNMR signal in Ca-PSII as a function of incubation time with H217O water. The intensity of the O5 signal can be estimated from either the maximum of the high field edge of the signal, or by determining the area of the O5 signal by fitting the envelope of the broad rectangular-shaped signal indicated in red in Fig. 3A. For better signal-to-noise we employed the latter method, which slightly differs from our original assignment described above, owing to the subsequently found good match of the shape of the broad O5 signal with that seen for exchanged μ-oxo bridges in Mn model complexes and Mn-catalase (49, 51); for further justification see SI Appendix, S3 and Fig. S8. Both methods yield a similar induction curve. The resulting data are normalized such that the signal intensity at full exchange is one. The time course of the O5 EDNMR signal rise is well-described by an induction rate of approximately 0.9 s−1 (dashed blue line in Fig. 3B; see SI Appendix, Fig. S10 for a replicate experiment yielding a rate of 1.1 s−1). The Sr-PSII data (green symbols) were treated in the same way. They demonstrate that the O5 exchange rate is significantly faster in Sr-PSII than in Ca-PSII – too fast for the determination of an exchange rate with our approach.
For quantitative comparison, the WS exchange rates for Ca-PSII and Sr-PSII were measured by TR-MIMS under nominally the same conditions with regards to pH, buffer composition (SI Appendix, Table S4) and temperature employing the same preparations as used for EDNMR (see Fig. 2A for the employed flash/injection sequence). The results obtained at the mass-to-charge (m/z) 36 signal are displayed in Fig. 3C. The rise kinetic of this signal reflects the exchange of the more tightly bound substrate WS, since for generating 36O2 both substrates at a catalytic site need to exchange (16, 18) (m/z 34 data are shown in SI Appendix, Fig. S11). The data show that in Ca-PSII the WS exchange rate has a comparable value (0.4 s−1) to the O5 exchange (0.9 s−1 and 1.1 s−1) measured by EDNMR in two independent datasets (Fig. 3B and SI Appendix, Figs. S9 and S10). We speculate that the 2.5-fold faster O5 exchange seen in TR-17O-EDNMR is caused by modest sample heating, from room temperature (22 °C) to about 28 to 30 °C (SI Appendix, Text S3.3), owing to friction during the fast mixing of the viscous PSII sample with H217O. In contrast, during TR-MIMS experiments, H218O is injected with much less friction into a dilute PSII suspension thermostated to 20°C; thus, no temperature increase is expected.
As in the case of O5 exchange in Sr-PSII, the exchange of WS also exhibits a significantly faster rate in Sr-PSII than in Ca-PSII. However, using MIMS, the WS exchange in Sr-PSII can be partially resolved in line with the 2.5-fold slower kinetics compared to TR-17O-EDNMR discussed above. Interestingly, the WS exchange in the S1 state of Sr-PSII was biphasic, with approximately 50% of centers exchanging at a rate faster than can be measured, while the rate of the second population was approximately fivefold faster (1.9 s−1) than the one observed for Ca-PSII. While this biphasic behavior in the 36O2 data is unusual, we have previously observed it in the m/z 36 data of substrate exchange experiments in the S2-state of Sr-PSII and attributed this observation to a slow conformational equilibrium between a low- and a high-spin form of the S2 state (38).
Discussion
With the aim to structurally identify the slow substrate water WS of the photosynthetic water oxidation reaction, we performed parallel experiments of substrate water exchange employing TR-MIMS and water-ligand exchange at the Mn4CaO5-cluster using a specially developed TR-17O-EDNMR method. These experiments were performed under highly comparable conditions in the S1 state on parallel samples of Ca-PSII and Sr-PSII core complexes isolated from T. vestitus. This allows a direct kinetic comparison of WS exchange with that of the O5 bridge of the Mn4CaO5 cluster.
Assignment of O5 to Ws.
Our present TR-17O-EDNMR data (Fig. 3) show that O5 exchanged in the S1 state of Ca-PSII with a rate that is very similar to that measured by TR-MIMS for WS exchange. Similarly, the exchange of both O5 and WS was significantly faster in Sr-PSII. Thus, our data unambiguously demonstrate that O5 matches the TR-MIMS characteristics of WS. This strongly supports our previous proposal that O5 is the slowly exchanging substrate in all S states (39, 49).
As discussed previously (38), W2 bound to Mn4 is the only other possible assignment for WS. The present data do, unfortunately, not allow a direct kinetic analysis of the W2 exchange in the S1 state via measuring the double quantum signals, which would allow unobscured observation of W2 exchange (49). This is due to the fact that the experiments had to be performed at lower sample concentrations than our previous experiments in order to allow mixing times with H217O that match the WS exchange rates established by TR-MIMS. Thus, our present data alone do not exclude W2 as WS; nevertheless, in combination with previous TR-MIMS results this remaining alternative can be eliminated. For this, we discuss in the following two scenarios in which W2 has been proposed to act as WS.
In the first scenario, the exchange of WS = W2 would be determined by the isotopic equilibration of both W2 and O5 with bulk water, an arrangement that was proposed to explain the observed S state dependence of WS exchange (54, 61). For the S1 state, it was proposed that O5 exchanges via W2 and thereby markedly slows the observable exchange rate of W2 [therefore the authors formally designated O5 as WS (54, 61)], while in S2 and S3 only W2 can exchange. First, we contest the kinetic viability of this model, since a slow exchange of O5 with W2 cannot markedly slow the much faster exchange of W2 with bulk water. Second, this proposal does not straight forwardly account for the Ca/Sr dependence of WS exchange, since W2 has no direct connection with Ca. Finally, in this (54, 61) and other nucleophilic attack proposals (54, 57, 58, 61) Wf is assigned to W3 that remains bound to Ca in the S3 and S4 states. Thus, these proposals struggle to account for the slowing of Wf exchange during the S2→S3 transition, and cannot explain the arrest of Wf exchange in the S3YZ• state (18, 23).
In a second scenario, WS is assumed to be W2 in the S0 to S2 states where its exchange rates would be modulated via S state–dependent conformational equilibria (38, 62). In contrast to scenario 1, W2 may then rotate during the S2→S3 transition via a pivot/carousel water insertion at Mn4 into the original O5 binding site (63, 64). In such a scenario, O-O bond formation may occur between the original W2, now the central oxo-bridge, and the original O5, which is proposed to be pushed by this water insertion path into the Ox/O6 binding site (Fig. 1C) (63, 64). Therefore, in the S1 state, Wf would be O5, which is in conflict with the present data that yield, in the S1 state, an exchange rate for O5 that is very similar to that measured by TR-MIMS for WS and thereby more than 100-fold slower than that of Wf (SI Appendix, Fig. S11 and refs. 38 and 40).
In conclusion, the combination of TR-17O-EDNMR and TR-MIMS demonstrates that O5 is the only viable candidate for WS.
Biphasic WS exchange in the S1 state of Sr-PSII.
The exchange of WS in Sr-PSII measured by TR-MIMS displayed a biphasic behavior, with an equal distribution between an unresolved fast phase and a resolved slower phase. Compared to Ca-PSII, the WS exchange in the resolved part of Sr-PSII was nearly fivefold faster.
The biphasic exchange of WS in the S1 state of Sr-PSII may indicate the presence of two S1 state structures. Evidence for two S1 structures was reviewed by Pantazis (4). More recently, Drosou et al. (21) proposed that the two S1 state conformations differ in the orientation of the JT axis at the Mn4(III) ion: in the more stable S1A state the JT axis is proposed to be along the coordination axis of the D1-Asp170 and D1-Glu333 ligands, while in the less stable S1B the JT axis is oriented along the W1-Mn4-O5 axis (Fig. 1D). Although this proposal awaits experimental confirmation, we propose that these two S1-state conformations provide an explanation for the two WS exchange rates. Two cases are possible, both assume that the two conformers are close in energy (thus populated about 50% each) and that S1B state facilitates fast O5 exchange. In case 1, the two rates reflect the O5 exchange rate in S1A and S1B, respectively. This assumes that there is a high barrier for the conversion of conformer S1A to S1B. In case 2, which we consider more likely, the faster rate reflects the exchange in the S1B state, while the slower rate reflects the conversion of S1A to S1B. In this scenario, the faster WS exchange in the S1B state can be understood on the basis of a proposal made for O5 exchange in the S1 state (22), in which an extra water molecule needs to bind to the open coordination site of Mn1 (Fig. 1D). This rate-limiting step involves a proton transfer to O5 that will be energetically more favorable in the S1B state where the JT axis is oriented toward O5 as this makes O5 a better proton acceptor. Consequently, the new water molecule can replace the O5 bridge more easily, which makes O5H to a terminal ligand of Mn4—a prerequisite for further protonation and its exchange with bulk water.
The slower, monophasic O5 exchange in Ca-PSII as compared to Sr-PSII may then indicate that both the energy difference between the A and B forms of the S1 state as well as the barrier for conversion of S1A to S1B are slightly larger in Ca-PSII. This would lead to nearly 100% S1A population and the observed 5-fold slower conversion into the still rapidly exchanging S1B conformation.
Consequences for the Mechanism of O-O Bond Formation.
Presently, a number of different mechanisms are discussed for O-O bond formation, which involve W2, W3, O5 and/or Ox/O6 as substrates. Our study strongly favors those mechanisms that assign O5 as the slowly exchanging substrates. In contrast, our present data provides no support, for example, for O2 formation via nucleophilic attack of W2 by W3 (54, 55, 57, 58). Nevertheless, a number of mechanisms remain, since the fast-exchanging substrate water has not been uniquely identified. At present, the water-derived ligands at the W2 and Ox/O6 binding sites of the S3 state both remain acceptable assignments on the basis of substrate water exchange results that indicate that the fast substrate is bound to Mn in the S3 state (discussed in ref. 40). Fig. 4 depicts the prospective S4 or S3YZ• states of five such proposals, drawn from current computational studies (65–71). It is remarkable that despite differences in geometry of the cluster or lack of water binding during the S2-S3 transition, in all cases a near linear Mn4-O O-Mn1 arrangement is found. Proposals A-D have in common that the spins on Mn4 and Mn1 are antiparallel, owing to the coupling within the Mn4CaO6 cofactor, which allows triplet 3O2 formation without the need for a spin conversion (65). In model E a higher spin state (S = 6) is utilized in which Mn4 and Mn1 have parallel spins to form a [O-O]3− intermediate (model F) at the level of the S3YZ• state by reducing Mn4. Note that formation of the peroxide intermediate, [O-O]2−, from F requires flipping the unpaired spins on Mn1 to allow its reduction. Thus, also in this case an antiparallel spin arrangement between Mn1 and Mn4 is required to form a peroxide intermediate (72, 73). We suggest that the options displayed in Fig. 4 capture the key design features needed for efficient Mn-based water oxidation, both in biology and by synthetic catalysts.
Fig. 4.
Comparison of selected mechanisms for O-O bond formation in photosystem II that includes O5 as a substrate. In all cases (A–F), a sketch of the calculated structure of the transient S4 or S3YZ• state is shown from which O-O bond formation is initiated. S4AW states (A, C, E, and F) are open cubane (or “right open”) structures that have taken up one substrate water into the open binding site at Mn1 during the S2→S3 transition. The position of O5 and the identity of the other substrate oxygen relative to the S1 state structure (Fig. 1A) depend on the water insertion mechanism, which remains controversial (for review, see refs. 4 and 74). Thus, no distinction is made between the slow- and fast-exchanging substrate water. For clarity, both substrate oxygens are shown in green. The two closed cubane (or “left open”) structures of the S4 state (B and D) are denoted as S4BW and S4B, respectively. While S4BW has taken up one substrate water during the S2→S3 transition, for S4B it is assumed that water binding does not have to occur up to the S4 state (75). The two oxo-oxyl radical coupling mechanisms (A and B) are denoted as OORC, the concerted bond switching mechanism (C) as CBS, and the nucleophilic oxo-oxo coupling (D) is signified as NOOC. Structures (E) and (F) are intermediates of a not named O-O bond formation mechanism that starts from a higher spin state (S = 6) and is proposed to occur prior to oxidation of the cluster by YZ•, which is achieved by sequential electron donation to two Mn(IV) ions. Note that there is a proton release between E and F and that peroxide formation starting from F requires a flip of the unpaired spins on Mn1 (69, 70). For better comparability and for emphasizing the similarity of the five proposals, the unpaired spins at Mn4 are denoted in all cases as spin down, while those on Mn1 as spin up (except E and F). Similarly, the bent half-arrows indicating the electron movements upon peroxide formation have been adjusted from the original proposals. Substrate oxidation in the OORC mechanisms (A and B) is indicated by a black dot near a green O. A, S4AW–OORC (65, 66); B, S4BW–OORC (ref. 66; based on ref. 39); C, S4AW–CBS (67, 76); D, S4B–NOOC (ref. 68; based on ref. 75); (E and F): [S3AW YZ•]+ and S3AW YZ• (69, 70); based on refs. 71–73.
Summary
In this study, we developed the H217O/H216O exchange TR-EDNMR approach and used it to show that O5 is the slowly exchanging substrate of water oxidation in PSII based on direct kinetic comparison to WS exchange kinetics measured by TR-MIMS under highly similar conditions. Our finding limits eligible proposals for the dioxygen formation/release mechanism to a set of five highly related options, all based on a similar O-O bond formation geometry and spin coupling of the manganese cluster. By adding flash illumination and active temperature control, the new TR-17O-EDNMR technique is expected to allow studies of H217O exchange also in the S2, S3, and S3YZox states that may eventually lead, together with snap-shot crystallography at XFELs (14) and TR-FTIR (15), to a further refinement of the experimental identification of the mechanism of this important enzymatic process.
Materials and Methods
PSII core complex preparations from T. vestitus were obtained as described earlier (74, 77, 78) and stored at −80 °C at a chlorophyll concentration of 2 to 3 mg/mL (pH 6.5; for details, see SI Appendix, Text S2 and section 2.1).
TR-MIMS 18O/16O Exchange Measurements.
PSII samples were thawed on ice, diluted to 0.3 mg/mL Chl and poised at S1 by a pre-flash and dark incubated at room temperature for 1 h. The substrate water exchange measurements in the S1-state of the Kok-cycle were performed at 20 °C essentially as described previously (16, 38). For minimizing injection artifacts, dissolved oxygen in the 18O labeled water (97% 18O) was removed by addition of a mix of glucose, glucose oxidase, and catalase prior to rapid mixing of the labeled water with Ca- or Sr-PSII. Flash and injection timing was controlled with trigger pulses generated by a LabView routine, while O2 and Ar isotopologues were detected with an isotope ratio mass spectrometer (Finnigan Delta plus XP). See SI Appendix, Texts S2 and S3 for further details.
TR-EDNMR Measurements.
PSII samples were thawed, and then concentrated to 4 to 5 mg Chl/mL (EDNMR set 1) or 8 to 9 mg/mL (EDNMR set 2) and loaded into the freeze quench system. Samples were mixed with 17O water (set 1: 77% enriched containing buffer with glycerol; set 2: 86%, no additions); see SI Appendix, Table S4 for the buffer composition during the exchange experiments. After freeze quenching, the frozen samples were packed into the EDNMR tubes. The S2 state was generated by illuminating the PSII samples for 3 s at 198 K (ethanol–dry ice bath) with two 250 W halogen lamps filtered by a 2 cm CuSO4 solution (5% w/v) and two filters each: Schott KG 3 (2 mm) and Schott GG 445 (2 mm). The final light intensity at the sample level was about 0.5 W/cm2.
EDNMR measurements at W-band were performed on these samples using a Bruker ELEXSYS E680 spectrometer at T =4 .8 K. Electron spin echo detected field-swept EPR spectra were measured using the pulse sequence: tp -τ-2 tp -τ- echo, with tp = 20 ns and τ = 600 ns. EDNMR spectra were collected using the pulse sequence: tHTA -T- tp -τ- tp -τ- echo, with tHTA = 6 μs, tp = 80 ns, τ = 600 ns, and T = 1 μs. Simulations of the EPR and EDNMR spectra were performed using the EasySpin package (79). See SI Appendix, Texts S2 and S3 for further details.
Supplementary Material
Appendix 01 (PDF)
Acknowledgments
Financial support by the Max Planck Society, the MANGAN project (03EK3545) funded by the Bundesministeriums für Bildung und Forschung, Vetenskaprådet (Grant 2016-05183 and 2020-03809 to J.M.) and the Australian Research Council (Grant DP200100338 to N.C.) is acknowledged. We thank Marcus Lundberg (Uppsala University) for discussions and Birgit Deckers for preparing initial Figures. Dmitry Shevela (SciGrafik) prepared the final Figures.
Author contributions
W.L., J.M., and N.C. designed research; C.d.L. and L.R. performed research; M.R., E.H., and M.M.N. contributed new reagents/analytic tools; C.d.L., L.R., A.S., J.M., and N.C. analyzed data; and J.M. and N.C. wrote the paper.
Competing interests
The authors declare no competing interest.
Footnotes
This article is a PNAS Direct Submission.
Contributor Information
Wolfgang Lubitz, Email: wolfgang.lubitz@cec.mpg.de.
Johannes Messinger, Email: johannes.messinger@kemi.uu.se.
Nicholas Cox, Email: nick.cox@anu.edu.au.
Data, Materials, and Software Availability
All study data are included in the article and/or SI Appendix.
Supporting Information
References
- 1.Shevela D., Kern J. F., Govindjee G., Messinger J., Solar energy conversion by photosystem II: Principles and structures. Photosynth. Res. 156, 279–307 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Lubitz W., Chrysina M., Cox N., Water oxidation in photosystem II. Photosynth. Res. 142, 105–125 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Yamaguchi K., et al. , 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. 471, 214742 (2022). [Google Scholar]
- 4.Pantazis D. A., Missing pieces in the puzzle of biological water oxidation. ACS Catal. 8, 9477–9507 (2018). [Google Scholar]
- 5.Kok B., Forbush B., McGloin M., Cooperation of charges in photosynthetic O2 evolution. Photochem. Photobiol. 11, 457–476 (1970). [DOI] [PubMed] [Google Scholar]
- 6.Lavergne J., Junge W., Proton release during the redox cycle of the water oxidase. Photosynth. Res. 38, 279–296 (1993). [DOI] [PubMed] [Google Scholar]
- 7.Schlodder E., Witt H. T., Stoichiometry of proton release from the catalytic center in photosynthetic water oxidation. J. Biol. Chem. 274, 30387–30392 (1999). [DOI] [PubMed] [Google Scholar]
- 8.Dau H., Haumann M., The manganese complex of photosystem II in its reaction cycle–Basic framework and possible realization at the atomic level. Coord. Chem. Rev. 252, 273–295 (2008). [Google Scholar]
- 9.Rappaport F., Diner B. A., Primary photochemistry and energetics leading to the oxidation of the (Mn)4Ca cluster and to the evolution of molecular oxygen in Photosystem II. Coord. Chem. Rev. 252, 259–272 (2008). [Google Scholar]
- 10.Klauss A., Haumann M., Dau H., Alternating electron and proton transfer steps in photosynthetic water oxidation. Proc. Natl. Acad. Sci. U.S.A. 109, 16035–16040 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Suga M., et al. , An oxyl/oxo mechanism for oxygen-oxygen coupling in PSII revealed by an x-ray free-electron laser. Science 366, 334–338 (2019). [DOI] [PubMed] [Google Scholar]
- 12.Kern J., et al. , Structures of the intermediates of Kok’s photosynthetic water oxidation clock. Nature 563, 421–425 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Ibrahim M., 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. U.S.A. 117, 12624–12635 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Bhowmick A., et al. , Structural evidence for intermediates during O2 formation in photosystem II. Nature 617, 629–636 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Greife P., et al. , The electron–proton bottleneck of photosynthetic oxygen evolution. Nature 617, 623–628 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Messinger J., Badger M., Wydrzynski T., Detection of one slowly exchanging substrate water molecule in the S3 state of photosystem II. Proc. Natl. Acad. Sci. U.S.A. 92, 3209–3213 (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Hillier W., Messinger J., “Mechanism of photosynthetic oxygen production” in Photosystem II. The Light-Driven Water:Plastoquinone Oxidoredutase, Wydrzynski T., Satoh K., Eds. (Springer, Dordrecht, 2005), vol. 22, pp. 567–608. [Google Scholar]
- 18.Cox N., Messinger J., Reflections on substrate water and dioxygen formation. Biochim. Biophys. Acta 1827, 1020–1030 (2013). [DOI] [PubMed] [Google Scholar]
- 19.Lohmiller T., et al. , The first state in the catalytic cycle of the water-oxidizing enzyme: Identification of a water-derived µ-hydroxo bridge. J. Am. Chem. Soc. 139, 14412–14424 (2017). [DOI] [PubMed] [Google Scholar]
- 20.Suga M., et al. , Light-induced structural changes and the site of O=O bond formation in PSII caught by XFEL. Nature 543, 131–135 (2017). [DOI] [PubMed] [Google Scholar]
- 21.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. Engl. 60, 13493–13499 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.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. 135, 9442–9449 (2013). [DOI] [PubMed] [Google Scholar]
- 23.Nilsson H., Rappaport F., Boussac A., Messinger J., Substrate-water exchange in photosystem II is arrested prior to dioxygen formation. Nat Commun 5, 4305 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Stich T. A., Yeagle G. J., Service R. J., Debus R. J., Britt R. D., Ligation of D1-His332 and D1-Asp170 to the manganese cluster of photosystem II from synechocystis assessed by multifrequency pulse EPR spectroscopy. Biochemistry 50, 7390–7404 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Navarro M. P., et al. , Ammonia binding to the oxygen-evolving complex of photosystem II identifies the solvent-exchangeable oxygen bridge (μ-oxo) of the manganese tetramer. Proc. Natl. Acad. Sci. U.S.A. 110, 15561–15566 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Umena Y., Kawakami K., Shen J. R., Kamiya N., Crystal structure of oxygen-evolving photosystem II at a resolution of 1.9 Å. Nature 473, 55–61 (2011). [DOI] [PubMed] [Google Scholar]
- 27.Suga M., et al. , Native structure of photosystem II at 1.95 Å resolution viewed by femtosecond x-ray pulses. Nature 517, 99–103 (2015). [DOI] [PubMed] [Google Scholar]
- 28.Krewald V., et al. , Metal oxidation states in biological water splitting. Chem. Sci. 6, 1676–1695 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Kulik L. V., Epel B., Lubitz W., Messinger J., Electronic structure of the Mn4OxCa cluster in the S0 and S2 states of the oxygen-evolving complex of photosystem II based on pulse 55Mn-ENDOR and EPR spectroscopy. J. Am. Chem. Soc. 129, 13421–13435 (2007). [DOI] [PubMed] [Google Scholar]
- 30.Cheah M. H., et al. , Assessment of the manganese cluster’s oxidation state via photoactivation of photosystem II microcrystals. Proc. Natl. Acad. Sci. U.S.A. 117, 141–145 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Yano J., et al. , Where water is oxidized to dioxygen: Structure of the photosynthetic Mn4Ca cluster. Science 314, 821–825 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.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 47, 11024–11030 (2008). [DOI] [PubMed] [Google Scholar]
- 33.Siegbahn P. E. M., A structure consistent mechanism for dioxygen formation in photosystem II. Chem. Eur. J. 14, 8290–8302 (2008). [DOI] [PubMed] [Google Scholar]
- 34.Cox N., et al. , Effect of Ca2+/Sr2+ substitution on the electronic structure of the oxygen-evolving complex of photosystem II: A combined multifrequency EPR, 55Mn-ENDOR, and DFT study of the S2 state. J. Am. Chem. Soc. 133, 3635–3648 (2011). [DOI] [PubMed] [Google Scholar]
- 35.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. Engl. 51, 9935–9940 (2012). [DOI] [PubMed] [Google Scholar]
- 36.Cox N., et al. , Electronic structure of the oxygen-evolving complex in photosystem II prior to O-O bond formation. Science 345, 804–808 (2014). [DOI] [PubMed] [Google Scholar]
- 37.Kim C. J., Debus R. J., One of the substrate waters for O2 formation in photosystem II is provided by the water-splitting Mn4CaO5 Cluster’s Ca2+ ion. Biochemistry 58, 3185–3192 (2019). [DOI] [PubMed] [Google Scholar]
- 38.de Lichtenberg C., Messinger J., Substrate water exchange in the S2 state of photosystem II is dependent on the conformation of the Mn4Ca cluster. Phys. Chem. Chem. Phys. 22, 12894–12908 (2020). [DOI] [PubMed] [Google Scholar]
- 39.Messinger J., Evaluation of different mechanistic proposals for water oxidation in photosynthesis on the basis of Mn4OxCa structures for the catalytic site and spectroscopic data. Phys. Chem. Chem. Phys. 6, 4764–4771 (2004). [Google Scholar]
- 40.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. 12, 12763–12775 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Hussein R., et al. , Evolutionary diversity of proton and water channels on the oxidizing side of photosystem II and their relevance to function. Photosynth. Res. 158, 91–107 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Helm L., Merbach A. E., Inorganic and bioinorganic solvent exchange mechanisms. Chem. Rev. 105, 1923–1959 (2005). [DOI] [PubMed] [Google Scholar]
- 43.Hendry G., Wydrzynski T., 18O isotope exchange measurements reveal that calcium is involved in the binding of one substrate-water molecule to the oxygen-evolving complex in photosystem II. Biochemistry 42, 6209–6217 (2003). [DOI] [PubMed] [Google Scholar]
- 44.Thomann H., Bernardo M., Goldfarb D., Kroneck P. M. H., Ullrich V., Evidence for water binding to the Fe center in cytochrome P450cam obtained by 17O electron-spin echo envelope modulation spectroscopy. J. Am. Chem. Soc. 117, 8243–8251 (1995). [Google Scholar]
- 45.Burdi D., et al. , The core structure of X generated in the assembly of the diiron cluster of ribonucleotide reductase: 17O2 and H217O ENDOR. J. Am. Chem. Soc. 120, 12910–12919 (1998). [Google Scholar]
- 46.Carepo M., et al. , 17O ENDOR detection of a solvent-derived Ni-(OHx)-Fe bridge that is lost upon activation of the hydrogenase from Desulfovibrio gigas. J. Am. Chem. Soc. 124, 281–286 (2002). [DOI] [PubMed] [Google Scholar]
- 47.Astashkin A. V., Feng C. J., Raitsimring A. M., Enemark J. H., 17O ESEEM evidence for exchange of the axial oxo ligand in the molybdenum center of the high pH form of sulfite oxidase. J. Am. Chem. Soc. 127, 502–503 (2005). [DOI] [PubMed] [Google Scholar]
- 48.Enemark J. H., Raitsimring A. M., Astashkin A. V., Klein E. L., Implications for the mechanism of sulfite oxidizing enzymes from pulsed EPR spectroscopy and DFT calculations for “difficult’’ nuclei. Faraday Discuss. 148, 249–267 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Rapatskiy L., et al. , Detection of the water-binding sites of the oxygen-evolving complex of photosystem II using W-band 17O electron-electron double resonance-detected NMR spectroscopy. J. Am. Chem. Soc. 134, 16619–16634 (2012). [DOI] [PubMed] [Google Scholar]
- 50.Lohmiller T., et al. , Structure, ligands and substrate coordination of the oxygen-evolving complex of photosystem II in the S2 state: A combined EPR and DFT study. Phys. Chem. Chem. Phys. 16, 11877–11892 (2014). [DOI] [PubMed] [Google Scholar]
- 51.Rapatskiy L., et al. , Characterization of oxygen bridged manganese model complexes using multifrequency 17O-hyperfine EPR spectroscopies and density functional theory. J. Phys. Chem. B 119, 13904–13921 (2015). [DOI] [PubMed] [Google Scholar]
- 52.Tagore R., Chen H. Y., Crabtree R. H., Brudvig G. W., Determination of µ-oxo exchange rates in di-µ-oxo dimanganese complexes by electrospray ionization mass spectrometry. J. Am. Chem. Soc. 128, 9457–9465 (2006). [DOI] [PubMed] [Google Scholar]
- 53.Sproviero E. M., Gascon 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. 130, 3428–3442 (2008). [DOI] [PubMed] [Google Scholar]
- 54.Vinyard D. J., Khan S., Brudvig G. W., Photosynthetic water oxidation: Binding and activation of substrate waters for O-O bond formation. Faraday Discuss. 185, 37–50 (2015). [DOI] [PubMed] [Google Scholar]
- 55.Gao Y., Åkermark T., Liu J. H., Sun L. C., Åkermark B., Nucleophilic attack of hydroxide on a MnV oxo complex: A model of the O-O bond formation in the oxygen evolving complex of photosystem II. J. Am. Chem. Soc. 131, 8726–8727 (2009). [DOI] [PubMed] [Google Scholar]
- 56.Huang H.-L., Brudvig G. W., Kinetic modeling of substrate-water exchange in Photosystem II. BBA Adv. 1, 100014 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Shen J.-R., The structure of photosystem II and the mechanism of water oxidation in photosynthesis. Annu. Rev. Plant Biol. 66, 23–48 (2015). [DOI] [PubMed] [Google Scholar]
- 58.Barber J., A mechanism for water splitting and oxygen production in photosynthesis. Nat. Plants 3, 17041 (2017). [DOI] [PubMed] [Google Scholar]
- 59.Cox N., Nalepa A., Lubitz W., Savitsky A., ELDOR-detected NMR: A general and robust method for electron-nuclear hyperfine spectroscopy? J. Magn. Reson. 280, 63–78 (2017). [DOI] [PubMed] [Google Scholar]
- 60.Schosseler P., Wacker T., Schweiger A., Pulsed ELDOR detected NMR. Chem. Phys. Lett. 224, 319–324 (1994). [Google Scholar]
- 61.Vinyard D. J., Brudvig G. W., Progress toward a molecular mechanism of water oxidation in photosystem II. Annu. Rev. Phys. Chem. 68, 101–116 (2017). [DOI] [PubMed] [Google Scholar]
- 62.Boussac A., et al. , The low spin - high spin equilibrium in the S2-state of the water oxidizing enzyme. Biochim. Biophys. Acta 1859, 342–356 (2018). [DOI] [PubMed] [Google Scholar]
- 63.Retegan M., et al. , A five-coordinate Mn(IV) intermediate in biological water oxidation: Spectroscopic signature and a pivot mechanism for water binding. Chem. Sci. 7, 72–84 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Askerka M., Vinyard D. J., Brudvig G. W., Batista V. S., NH3 Binding to the S2 state of the O2-evolving complex of photosystem II: Analogue to H2O binding during the S2 -> S3 transition. Biochemistry 54, 5783–5786 (2015). [DOI] [PubMed] [Google Scholar]
- 65.Siegbahn P. E. M., O-O bond formation in the S4 state of the oxygen-evolving complex in photosystem II. Chem. Eur. J. 12, 9217–9227 (2006). [DOI] [PubMed] [Google Scholar]
- 66.Li X. C., 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. 17, 12168–12174 (2015). [DOI] [PubMed] [Google Scholar]
- 67.Shoji M., Isobe H., Yamaguchi K., Concerted bond switching mechanism coupled with one-electron transfer for the oxygen-oxygen bond formation in the oxygen-evolving complex of photosystem II. Chem. Phys. Lett. 714, 219–226 (2019). [Google Scholar]
- 68.Guo Y., Messinger J., Kloo L., Sun L. C., Alternative mechanism for O2 formation in natural photosynthesis via nucleophilic oxo-oxo coupling. J. Am. Chem. Soc. 145, 4129–4141 (2023). [DOI] [PubMed] [Google Scholar]
- 69.Rummel F., Malcomson T., Barchenko M., O’Malley P. J., Insights into PSII’s S3YZ• state: An electronic and magnetic analysis. J. Phys. Chem. Lett. 15, 499–506 (2024), 10.1021/acs.jpclett.3c03026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Rummel F., O’Malley P. J., How Nature makes O2: An electronic level mechanism for water oxidation in photosynthesis. J. Phys. Chem. B 126, 8214–8221 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.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. 9, 3525–3531 (2018). [DOI] [PubMed] [Google Scholar]
- 72.Renger G., Renger T., Photosystem II: The machinery of photosynthetic water splitting. Photosynth. Res. 98, 53–80 (2008). [DOI] [PubMed] [Google Scholar]
- 73.Messinger J., Renger G., “Photosynthetic water-splitting” in Primary Processes of Photosynthesis - Part 2: Basic Principles and Apparatus, Renger G., Ed. (The Royal Society of Chemistry, Cambridge, UK, 2008), pp. 291–349. [Google Scholar]
- 74.Chrysina M., et al. , Five-coordinate Mn-IV intermediate in the activation of nature’s water splitting cofactor. Proc. Natl. Acad. Sci. U.S.A. 116, 16841–16846 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.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. 199, 110797 (2019). [DOI] [PubMed] [Google Scholar]
- 76.Yamaguchi K., Miyagawa K., Shoji M., Isobe H., Kawakami T., Elucidation of a multiple S3 intermediates model for water oxidation in the oxygen evolving complex of photosystem II. Calcium-assisted concerted O-O bond formation. Chem. Phys. Lett. 806, 140042 (2022). [Google Scholar]
- 77.Boussac A., et al. , Biosynthetic Ca2+/Sr2+ exchange in the photosystem II oxygen-evolving enzyme of Thermosynechococcus elongatus. J. Biol. Chem. 279, 22809–22819 (2004). [DOI] [PubMed] [Google Scholar]
- 78.Nowaczyk M. M., et al. , Psb27, a cyanobacterial lipoprotein, is involved in the repair cycle of photosystem II. The Plant Cell 18, 3121–3131 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Stoll S., Schweiger A., EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 178, 42–55 (2006). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Appendix 01 (PDF)
Data Availability Statement
All study data are included in the article and/or SI Appendix.




