Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2015 Dec 29;113(3):620–625. doi: 10.1073/pnas.1520211113

Redox potential of the terminal quinone electron acceptor QB in photosystem II reveals the mechanism of electron transfer regulation

Yuki Kato 1,1, Ryo Nagao 1, Takumi Noguchi 1,1
PMCID: PMC4725517  PMID: 26715751

Significance

In photosynthesis, photosystem II (PSII) has a function of abstracting electrons from water using light energy and transferring them to a quinone molecule. In addition to the forward electron transfer in PSII, which is essential in energy conversion, backward electron transfer is important in photoprotection of PSII proteins. Forward and backward electron transfers in PSII are regulated by the redox potential (Em) gap of quinone electron acceptors, QA and QB. However, the regulation mechanism is still unclear because Em of QB has not been determined. We directly measured Em of QB using an electrochemical method in combination with Fourier transform infrared spectroscopy. Our results clearly explain the mechanism of electron transfer regulation in PSII relevant to photoprotection.

Keywords: photosynthesis, spectroelectrochemistry, FTIR

Abstract

Photosystem II (PSII) extracts electrons from water at a Mn4CaO5 cluster using light energy and then transfers them to two plastoquinones, the primary quinone electron acceptor QA and the secondary quinone electron acceptor QB. This forward electron transfer is an essential process in light energy conversion. Meanwhile, backward electron transfer is also significant in photoprotection of PSII proteins. Modulation of the redox potential (Em) gap of QA and QB mainly regulates the forward and backward electron transfers in PSII. However, the full scheme of electron transfer regulation remains unresolved due to the unknown Em value of QB. Here, for the first time (to our knowledge), the Em value of QB reduction was measured directly using spectroelectrochemistry in combination with light-induced Fourier transform infrared difference spectroscopy. The Em(QB/QB) was determined to be approximately +90 mV and was virtually unaffected by depletion of the Mn4CaO5 cluster. This insensitivity of Em(QB/QB), in combination with the known large upshift of Em(QA/QA), explains the mechanism of PSII photoprotection with an impaired Mn4CaO5 cluster, in which a large decrease in the Em gap between QA and QB promotes rapid charge recombination via QA.


In oxygenic photosynthesis in plants and cyanobacteria, photosystem II (PSII) has an important function in light-driven water oxidation, a process that leads to the generation of electrons and protons for CO2 reduction and ATP synthesis, respectively (13). Photosynthetic water oxidation also produces molecular oxygen as a byproduct, which is the source of atmospheric oxygen and sustains virtually all life on Earth. PSII reactions are initiated by light-induced charge separation between a chlorophyll (Chl) dimer (P680) and a pheophytin (Pheo) electron acceptor, leading to the formation of a P680+Pheo radical pair (4, 5). An electron hole on P680+ is transferred to a Mn4CaO5 cluster, the catalytic center of water oxidation, via the redox-active tyrosine, YZ (D1-Tyr161). At the Mn4CaO5 cluster, water oxidation proceeds through a cycle of five intermediates denoted Sn states (n = 0–4) (6, 7). On the electron acceptor side, the electron is transferred from Pheo to the primary quinone electron acceptor QA and then to the secondary quinone electron acceptor QB (8, 9). QA and QB have many similarities: they consist of plastoquinone (PQ), are located symmetrically around a nonheme iron center, and interact with D2 and D1 proteins, respectively, in a similar manner (Fig. 1) (10, 11). However, they play significantly different roles in PSII (8, 9). QA is only singly reduced to transfer an electron to QB, whereas QB accepts one or two electrons. When QB is doubly reduced, the resultant QB2− takes up two protons to form plastoquinol (PQH2), which is then released into thylakoid membranes. Differences between QA and QB could be caused by differences in the molecular interactions of PQ with surrounding proteins in QA and QB pockets, although the detailed mechanism remains to be clarified (12, 13).

Fig. 1.

Fig. 1.

Redox cofactors in PSII and the electron transfer pathway (blue arrows). For the PSII structure, the X-ray crystallographic structure at 1.9-Å resolution (Protein Data Bank ID code 3ARC) (9) was used. The electron acceptor side is expanded, showing the arrangements of QA, QB, and the nonheme iron with their molecular interactions. Nearby carboxylic groups are also shown.

Electron transfer reactions in PSII are highly regulated by the spatial localization of redox components and their redox potentials (Em values). Both forward and backward electron transfers are important; backward electron transfers control charge recombination in PSII, and this serves as photoprotection for PSII proteins (5, 1417). PSII involves specific mechanisms to regulate forward and backward electron transfer reactions in response to environmental changes. For instance, in strong light, some species of cyanobacteria increase the Em of Pheo to facilitate charge recombination. Specifically, they exchange D1 subunits originating from different psbA genes to change the hydrogen bond interactions of Pheo (1620). On the other hand, it was found that impairment of the Mn4CaO5 cluster led to a significant increase in the Em of QA by ∼150 mV (2127). This potential increase was thought to inhibit forward electron transfer to QB to promote direct relaxation of QA without forming triplet-state Chl, a precursor of harmful singlet oxygen (2, 5, 14, 15, 17, 23). In addition, charge recombination of QA with P680+ prevents oxidative damage by high-potential P680+ (5). However, the full mechanism of photoprotection by the regulation of the quinone electron acceptor Em values remains to be resolved, because the Em of QB has not been determined conclusively, and the effect of Mn4CaO5 cluster inactivation on it has not been examined (5).

Although the Em of the single reduction of QB has been estimated to be ∼80 mV higher than that of QA from kinetic and thermodynamic data (2832), so far no reports have measured the Em of QB directly. In contrast, the Em of QA was measured extensively using chemical or electrochemical titrations and determined to be approximately −100 mV for oxygen-evolving PSII (2127). The main reason for this difference is due to the fact that the QA reaction can be monitored readily by fluorescence measurement in that an increase in fluorescence indicates QA formation (8, 3335). However, the fluorescence method cannot be used easily to monitor QB reduction. Although UV-Vis absorption and electron spin resonance have also been used to monitor QA in redox titration (summarized in ref. 22), so far these methods have not been used to monitor the titration of QB, likely because QA and QB give similar signals (3639). Another spectroscopic method that can be used to monitor QA and QB reactions is Fourier transform infrared (FTIR) difference spectroscopy, which detects reaction-induced changes in the molecular vibrations of a cofactor and its environment in proteins (4045). It was previously shown that comparison of FTIR difference spectra upon QA and QB formation showed some characteristic differences in spectral features (46). In particular, bands at 1,721 and 1,745 cm−1, which were assigned to ester C=O vibrations of nearby Pheo molecules affected by the reduction of QA and QB, respectively, were suggested to be good markers for discriminating between QA and QB reactions (46).

In this study, we directly measured the Em of QB in PSII using spectroelectrochemistry and light-induced FTIR difference spectroscopy. The effect of Mn4CaO5 cluster depletion on the Em value was also examined. Spectroelectrochemistry has been used to accurately measure the Em values of cofactors in various redox proteins (47, 48) including redox cofactors in PSII (24, 25, 49, 50). FTIR spectroelectrochemistry, which has the additional merit of being able to provide structural information, has also been used to investigate redox reactions of biomolecules and proteins (48, 5054). This method was recently applied to the nonheme iron center of PSII to examine the effect of Mn depletion on the Em value and obtain structural information around the nonheme iron (50). The results in our study showed that the Em of the first reduction of QB [Em(QB/QB)] was much higher than previously estimated, and the Em of the second reduction [Em(PQH2/QB)] was higher than the first reduction. Furthermore, we showed that Mn depletion hardly affected the Em values of QB, in contrast to the large change in the Em of QA (2127). With these results, the mechanism of photoprotection of PSII when the Mn4CaO5 cluster is inactivated is now clearly explained.

Results

FTIR difference spectra of PSII core complexes from a thermophilic cyanobacterium Thermosynechococcus elongatus upon illumination of a single saturating flash were measured at a series of electrode potentials ranging from +250 to +50 mV (Fig. 2A). At the highest potential of +250 mV, a prominent peak was observed at 1,480 cm−1, which is typical of the CO/CC stretching vibration of a QB semiquinone anion (46), whereas at the lowest potential of +50 mV, a similar peak was found at a slightly lower frequency of 1,478 cm−1, which is typical of QA (55, 56). A significant change was observed in the ester C=O region (1,750–1,700 cm−1): the intensity of a positive peak at 1,745 cm−1 decreased gradually, and a peak at 1,721 cm−1 increased, as the potential was lowered from +250 to +50 mV. The former and latter peaks were shown previously to belong to QB/QB and QA/QA differences, respectively, and were proposed to be good markers for discriminating between QB and QA reactions (46). Each peak was assigned to the 132-ester C=O vibration of the nearby Pheo (PheoD2 and PheoD1 for QB and QA, respectively; Fig. 1), which is electrostatically and/or structurally coupled to QB(A) and hence affected by its photoreduction (46). The frequency difference between 1,745 and 1,721 cm−1 is attributed to a difference in hydrogen bond interactions of the Pheo molecules. The hydrogen bond between the 132-ester C=O of PheoD1 and D1-Tyr126 (the X-ray structure is shown in Fig. S1) provides a relatively low frequency of 1,721 cm−1, whereas the corresponding residue near PheoD2 (D2-Phe125) does not form a hydrogen bond, and hence the C=O frequency shows a higher value at 1,745 cm−1. The observed intensity changes in the 1,745 and 1,721 cm−1 peaks indicated that the spectrum gradually changed from QB/QB to QA/QA differences as the potential was lowered from +250 to +50 mV.

Fig. 2.

Fig. 2.

Flash-induced FTIR difference spectra of the intact (A) and Mn-depleted (B) PSII core complexes from T. elongatus at a series of electrode potentials. A single flash from a Nd:YAG laser (532 nm, ∼7 ns) was applied to the sample equilibrated at each potential at 10 °C, and an FTIR difference spectrum was measured. The region in 1,700–1,600 cm−1, which is expressed as dotted lines, is saturated by strong absorption of the amide I and water HOH bending bands.

Fig. S1.

Fig. S1.

Structure around the quinone electron acceptors and Pheo molecules in the PSII core complexes (Protein Data Bank ID code 3ARC). The blue allows indicate the electron transfer pathway. The 132-ester C=O of PheoD1 forms a hydrogen bond with D1-Tyr126, whereas that of PheoD2 is free from a hydrogen bond.

The FTIR spectra also showed a negative band at 1,401 cm−1 (Fig. 2A), which is typical of the S2/S1 difference and was assigned to the symmetric COO stretching vibrations of carboxylate groups surrounding the Mn4CaO5 cluster (5759). This observation was consistent with an intact Mn4CaO5 cluster in the PSII sample in the spectroelectrochemical cell. It was noted that, although the intensity of this band gradually decreased as the electrode potential was lowered, it was almost fully recovered by applying a high potential of +350 mV to the sample after completion of a series of measurements from +250 to +50 mV. This observation indicated that the Mn4CaO5 cluster was probably reduced to lower S states such as the S0 and S−1 states at lower potentials and reoxidized to the S1 state at higher potentials.

FTIR difference spectra of Mn-depleted PSII core complexes at a series of electrode potentials (Fig. 2B) showed spectral features similar to those of intact PSII (Fig. 2A) except for the absence of a negative band at 1,401 cm−1 from the Mn4CaO5 cluster. The intensity of the peak at 1,745 cm−1 decreased as the potential was lowered, and almost disappeared at +100 mV. In contrast, the intensity of the 1,721 cm−1 peak increased at lower potentials, although it was observed even at the highest potential of +250 mV. The latter observation suggested that electron transfer from QA to QB was partly suppressed in Mn-depleted PSII, probably because of the unoccupied QB site in a fraction of centers. Note that the slightly lower frequency of the strongest CO/CC peak at 1,479 cm−1 at higher potentials (Fig. 2B) compared with the frequency of intact PSII at 1,480 cm−1 at similar potentials (Fig. 2A) can be ascribed to the contribution of the QA/QA signal. Indeed, a pure QB/QB spectrum of Mn-depleted PSII, which was obtained at +250 mV after the relaxation of QA by recording a spectrum 20 s after the flash, showed the CO/CC frequency at 1,480 cm−1 (Fig. S2C).

Fig. S2.

Fig. S2.

(A) S2QB/S1QB (black line) and S2/S1 (green line) difference spectra of the intact PSII core complexes, which were measured at +350 and +470 mV, respectively. (B) A double difference spectrum (S2QB/S1QB-minus-S2/S1) representing a QB/QB difference spectrum of intact PSII. (C) A QB/QB difference spectrum of the Mn-depleted PSII core complexes, which was measured at +250 mV.

The reaction of QB was evaluated by the intensity of the 1,745 cm−1 peak specific to the QB/QB difference. To estimate the intensity of the 1,745 cm−1 peak accurately, the minor contribution of the QA/QA signal at this position (a small differential signal at 1,749/1,742 cm−1) was eliminated by subtraction of the QA/QA difference spectrum obtained at +50 mV (Fig. 2 A and B, bottom spectrum) using a factor determined on the basis of the peak intensity at 1,721 cm−1. The relative intensities of the 1,745 cm−1 peak (α) in the corrected spectra at a series of the potentials (Fig. 3A) are shown in a Nernst plot in semilogarithmic form (Fig. 3B). The plots for both intact and Mn-depleted PSII samples revealed a virtually linear relationship, with slopes of 30 ± 2 and 39 ± 1 mV, respectively. These slopes are close to 28 mV, the theoretical value of a two-electron reaction at 10 °C (measurement temperature), but much lower than 56 mV, the theoretical value of a one-electron reaction. The intercepts, which indicate the apparent redox potentials (Emapp) of the reaction, were estimated to be +155 ± 2 and +132 ± 1 mV for intact and Mn-depleted PSII samples, respectively. Nernst curves assuming ideal two-electron reactions with Em values of +155 and +132 mV (Fig. 3C, dotted lines) largely reproduced the experimental data with slight deviations.

Fig. 3.

Fig. 3.

(A) Expanded view of the 1,745 cm−1 band specific to QB/QB at a series of electrode potentials for the intact (Left) and Mn-depleted (Right) PSII core complexes. The contribution of the QA/QA signal (typically at 1,721 cm−1) was eliminated by subtraction of the QA/QA difference spectrum measured at +50 mV from each spectrum. (B) Semilogarithmic Nernst plots of the redox reactions of QB in the intact (blue circles) and Mn-depleted (red circles) PSII core complexes. α indicates the intensity ratio of the 1,745 cm−1 peak at each electrode potential relative to the intensity at +250 mV, at which QB is fully oxidized. The regression lines (dashed lines) with slopes (represented by triangles) and intercepts are also shown. (C) Fitting of the experimental intensity ratios (α) with theoretical Nernst curves for the intact (blue lines) and Mn-depleted (red lines) PSII samples. Dashed lines reveal the theoretical curves of two-electron reactions with Em of +155 (intact) and +132 (Mn-depleted) mV, whereas solid lines reveal the simulated curves with Em values of the first and second reduction of QB as fitting parameters: Em(QB/QB) = +93 mV and Em(PQH2/QB) = +213 mV for intact PSII, and Em(QB/QB) = +87 mV and Em(PQH2/QB) = +157 mV for Mn-depleted PSII.

The fair fit with a two-electron reaction indicates that the redox potential of the second reduction of QB [Em(PQH2/QB)] was higher than that of the first reduction [Em(QB/QB)], and hence QB was electrochemically doubly reduced to PQH2 without forming singly reduced QB as a major component. However, an ideal two-electron curve is not realized unless Em(QB/QB) is at least 180 mV more negative (60) than Em(PQH2/QB) (SI Text and Fig. S3). The observed spectra in Fig. 2 also support the higher Em(PQH2/QB) than Em(QB/QB), because if QB was reduced mainly to QB at a particular potential, flash illumination would have induced a QB•PQH2/QB•PQ difference spectrum providing signals with opposite intensities to the QB/QB spectrum, e.g., negative peaks at 1,480 and 1,745 cm−1. However, such negative peaks were not observed throughout the potentials in Fig. 2. To estimate Em(QB/QB) and Em(PQH2/QB) values, experimental Nernst plots were simulated using Em(QB/QB) and Em(PQH2/QB) as fitting parameters (for details of the simulation, see SI Text and Fig. S4). In the simulation, it should be noted that the experimental relative intensity of the 1,745 cm−1 peak not only reflects the population of neutral QB but also reflects the contribution of the QB population as a negative intensity. The simulated Nernst curves (Fig. 3C, solid lines) provided the following data: Em(QB/QB) = +93 ± 27 mV and Em(PQH2/QB) = +213 ± 36 mV with a 120-mV gap for intact PSII; Em(QB/QB) = +87 ± 16 mV and Em(PQH2/QB) = +157 ± 36 mV with a 70-mV gap for Mn-depleted PSII (Table 1).

Fig. S3.

Fig. S3.

Simulated Nernst curves for the mole fractions of QB (solid lines) and QB (dashed lines) against electrode potential E at 10 °C for various ΔEm (=Em1Em2) centered at 155 mV. ΔEm = +240 mV (blue lines; Em1 = +275 mV, Em2 = +35 mV), +40 mV (light blue lines; Em1 = +175 mV, Em2 = +135 mV), 0 mV (green lines; Em1 = Em2 = +155 mV); −40 mV (light green lines; Em1 = +135 mV, Em2 = +175 mV), −90 mV (brown lines; Em1 = +110 mV, Em2 = +200 mV), and −180 mV (red lines: Em1 = +65 mV, Em2 = +245 mV). A black line represents an ideal two-electron redox reaction with Em = +155 mV.

Fig. S4.

Fig. S4.

Simulated Nernst curves for the mole fraction difference between QB and QB ([QB]/[QB]0 – [QB]/[QB]0) against electrode potential E at 10 °C for various ΔEm (=Em1Em2). ΔEm = +240 mV (blue lines; Em1 = +275 mV, Em2 = +35 mV), +40 mV (light blue lines; Em1 = +175 mV, Em2 = +135 mV), 0 mV (green lines; Em1 = Em2 = +155 mV); −40 mV (light green lines; Em1 = +135 mV, Em2 = +175 mV), −90 mV (brown lines; Em1 = +110 mV, Em2 = +200 mV), and −180 mV (red lines: Em1 = +65 mV, Em2 = +245 mV).

Table 1.

Redox potentials of QB in the intact and Mn-depleted PSII core complexes from T. elongatus determined by FTIR spectroelectrochemistry

PSII sample Emapp,* mV Em(QB/QB), mV Em(PQH2/QB), mV
Intact PSII +155 +93 ± 27 +213 ± 36
Mn-depleted PSII +132 +87 ± 16 +157 ± 36
*

Apparent Em value obtained from the semilogarithmic Nernst plot in Fig. 3B.

Values estimated by simulation using a theoretical Nernst curve with Em(QB/QB) and Em(PQH2/QB) as fitting parameters.

Discussion

To our knowledge, we are the first to report the successful direct measurement of Em values of QB in PSII core complexes from T. elongatus using the spectroelectrochemical method, which uses a flash-induced FTIR signal at 1,745 cm−1 specific to the QB-to-QB change as a marker (46) (Figs. 2 and 3A). A Nernst plot (in semilogarithmic form) of the relative intensity of the 1,745 cm−1 signal suggested a two-electron reaction with a Emapp of +155 mV (Fig. 3B, blue dashed line), suggesting that the Em(PQH2/QB) of the second reduction of QB was higher than the Em(QB/QB) for the first reduction of QB. Further simulation of the Nernst plot estimated Em(QB/QB) and Em(PQH2/QB) to be +93 ± 27 and +213 ± 36 mV, respectively, in PSII core complexes with an intact Mn4CaO5 cluster at pH 6.5 (Fig. 3C and Fig. S5, blue solid line; Table 1). The higher Em for the second reduction compared with the first reduction, which is opposite to the general tendency of two-electron reactions, was realized by the stabilization of a doubly reduced form by the formation of neutral PQH2 by the uptake of two protons (8, 9, 30). This has been widely seen in redox reactions of various quinone molecules in protic solvents (52, 61).

Fig. S5.

Fig. S5.

Simulated curves as best fits for the experimental relative intensities of the 1,745 cm−1 band in the FTIR spectra of intact (blue) and Mn-depleted (red) PSII samples. The difference of the mole fractions of QB and QB ([QB]/[QB]0 – [QB]/[QB]0; solid lines) was fit to the experimental data (circles), and the curves of [QB]/[QB]0 (dotted lines) and [QB]/[QB]0 (dashed lines) are also presented. The parameters for the best fit were Em1 = +93 mV and Em2 = +213 mV (ΔEm = −120 mV) for intact PSII, and Em1 = +87 mV and Em2 = +157 mV (ΔEm = −70 mV) for Mn-depleted PSII.

Taking into account the reported Em(QA/QA) value of approximately −100 mV in intact PSII (2127), the Em(QB/QB) of approximately +90 mV indicates that the ΔEm between QA/QA and QB/QB is ∼190 mV (Fig. 4). This value is much larger than the ΔEm value of ∼80 mV between QA and QB previously estimated from kinetic and thermodynamic measurements using fluorescence detection (2630). However, these measurements were performed using the herbicide 3-(3,4-dichlorophyenyl)-1,1-dimethylurea (DCMU) to accumulate QA as a reference state, without taking into consideration the influence of DCMU binding to the QB site on Em(QA/QA). It was later demonstrated that the binding of DCMU induced a positive shift of Em(QA/QA) by ∼50 mV (62). Thus, the corrected value of ΔEm obtained from thermodynamic measurements should be ∼130 mV, which is in better agreement with, albeit still lower than, the value obtained by direct measurement in this study.

Fig. 4.

Fig. 4.

(A) Diagram of the redox potentials of the electron transfer components in PSII. (B) The effect of Mn depletion on the redox potentials of single reduction of QA and QB. Solid and dashed black arrows indicate forward and backward electron transfer, respectively. The redox potential levels of intact and Mn-depleted PSII are expressed by blue and red bars, respectively, in B. aValues from refs. 1925. bValues from refs. 4 and 46.

We further demonstrated that depletion of the Mn4CaO5 cluster induced only a small negative shift of Emapp by 23 mV (Fig. 3B). Em(QB/QB) and Em(PQH2/QB) values estimated by the simulation of the Nernst plot were +87 ± 27 and +157 ± 36 mV, respectively (Fig. 3C, Fig. S5, and Table 1), which indicate negative shifts of Em by 6 and 56 mV, respectively. Taking into account the relatively large error in this simulation, this result suggests that Em(QB/QB) and Em(PQH2/QB) are virtually unaffected or slightly downshifted by Mn depletion. This is in sharp contrast to the large upshift of Em(QA/QA) by ∼150 mV upon inactivation of the Mn4CaO5 cluster, which was reported previously (2127). As a result of the insensitivity of Em(QB/QB) to Mn depletion, ΔEm between QA/QA and QB/QB largely decreased from ∼190 to ∼40 mV by destruction of the Mn4CaO5 cluster (Fig. 4B).

Based on these data, electron transfer regulation and hence the mechanism of PSII photoprotection are now clear (Fig. 4B). When the Mn4CaO5 cluster is impaired, Em(QA/QA) is significantly upshifted whereas Em(QB/QB) is virtually unaffected, resulting in a large ΔEm decrease. This ΔEm decrease promotes backward electron transfer from QB to QA, facilitating fast relaxation via QA by a direct charge recombination with P680+, which prevents the accumulation of P680+, a strong oxidant that damages PSII (5). The direct recombination of QA with P680+ is further promoted by an increase in the ΔEm between PheoD1/PheoD1 and QA/QA, which inhibits charge recombination via PheoD1 to generate harmful singlet oxygen via the Chl triplet state (Fig. 4A) (5, 14, 15, 17, 23). Thus, the regulation of Em values of QA and QB electron acceptors upon impairment of the Mn4CaO5 cluster or during the process of its photoactivation protects PSII against both oxidative damage by P680+ and damage by singlet oxygen.

Key elements of the mechanism of PSII photoprotection are significantly different effects of Mn4CaO5 cluster inactivation on the Em values of QA and QB, i.e., largely upshifted Em and unaffected or only slightly downshifted Em, respectively. QA and QB are ∼40 Å away from the Mn4CaO5 cluster, and hence the molecular mechanism of the long-range interaction between the Mn4CaO5 cluster and the iron–quinone complex remains poorly understood (5, 14, 21). A recent FTIR electrochemistry study (50) showed that Em(Fe2+/Fe3+) of the nonheme iron was shifted only by +18 mV upon Mn depletion, which was much smaller than the Em shift of QA. The analysis of the Fe2+/Fe3+ FTIR difference spectra showed that Mn depletion caused some structural perturbations around the QB and nonheme iron sites. One change was the increase in the pKa of a nearby Glu residue upon Mn depletion; D1-Glu244 interacting with the nonheme iron through the bicarbonate ligand was proposed to be a tentative candidate, although the possible involvement of other Glu residues located on the stromal side (Fig. 1) was not excluded. In addition, it was suggested that the hydrogen bond of one C=O of QB with D1-His215, which is a ligand to the nonheme iron, was strengthened, whereas the hydrogen bonds of other C=O of QB with the backbone NH and D1-Ser264 were weakened. However, the QB/QB FTIR difference spectra measured for intact and Mn-depleted PSII samples (Fig. S2) exhibited a prominent CO/CC band of QB at the same frequency of 1,480 cm−1, suggesting that at least the hydrogen bond interaction of the QB was not altered significantly, which is consistent with the virtually unaffected Em(QB/QB). In addition, the FTIR study showed that the hydrogen bond interaction of QA was also unaffected by Mn depletion (56). Therefore, it is possible that the pKa changes of protonatable groups including five Glu residues on the stromal side (Fig. 1) and their asymmetric distribution may have affected the Em values of QA and QB differently; the pKa shifts additively affected the Em of QA, whereas they cancelled each other out, resulting in the unchanged Em of QB. Elucidation of the route of the long-range interaction from the Mn4CaO5 cluster to the stromal side and the detailed mechanism of the effects on Em values of QA and QB, which is the essence of the photoprotection of PSII, will require further investigation. In such investigations, FTIR spectroelectrochemistry combined with site-directed mutagenesis of protonatable amino acids on the stromal side will be fruitful.

Methods

The oxygen-evolving PSII core complexes were purified from cells of the T. elongatus 47-H strain (63), in which a (His)6-tag was genetically attached to the carboxyl terminus of the CP47 subunit, using Ni2+-affinity column chromatography as described previously (64). The O2 evolution activity of PSII core complexes was 2,500–2,800 μmol O2⋅(mg Chl)−1⋅h−1. To deplete the Mn4CaO5 cluster, the PSII complexes were treated with 10 mM NH2OH for 30 min at room temperature in the dark (65), followed by washing with Mes buffer, pH 6.5 (Buffer A: 40 mM Mes-NaOH, 5 mM NaCl, 5 mM CaCl2, 0.06% n-dodecyl β-d-maltoside) containing 10% (wt/vol) polyethylene glycol 6000 (PEG6000) by centrifugation. After washing, the sample was suspended in Buffer A. The PSII sample (0.2 mg Chl/mL) was suspended in an electrolytic solution containing 40 mM NaHCO3, 200 mM KCl, 1 M betaine, 10% (wt/vol) PEG6000, and the mixture of redox mediators in Buffer A. The redox mediators were 200 μM 1-methoxy-5-methylphanazinium methosulfate (Em = +63 mV), 200 μM N,N,N′,N′-tetramethyl-p-phenylenediamine (Em = +260 mV), and 4 mM potassium ferricyanide (Em = +430 mV). For measurements of standard S2QB/S1QB and S2/S1 difference spectra, 20 mM potassium ferricyanide was used as a sole redox mediator. The PSII sample in the electrolytic solution was centrifuged at 170,000 × g for 15 min, and the resulting pellet was loaded onto an optically transparent thin-layer electrode (OTTLE) cell for FTIR measurements as reported previously (50).

In the OTTLE cell, a gold mesh (60% transparent, 6-μm thickness; Precision Eforming), which was chemically modified with 4,4′-dithiodipyridine, was used as a working electrode, while a Pt black wire and a Ag/AgCl/3M KCl electrode (Cypress Systems; 66-EE009) were used as counter and reference electrodes, respectively. The sample pellet was loaded onto the gold mesh placed on a CaF2 plate in the cell and was sandwiched with another CaF2 plate. The assembled cell was set in a copper holder, and the sample temperature was adjusted to 10 °C by circulating cold water through the holder.

Flash-induced FTIR spectra were measured using a Bruker IFS-66/S spectrophotometer equipped with an MCT detector (D313-L/3) at 4 cm−1 resolution. A Q-switched Nd:YAG laser (Quanta-Ray INDI-40-10; 532 nm, ∼7-ns full width at half-maximum, ∼7 mJ⋅pulse−1⋅cm−2) was used for flash illumination. The electrode potential of the OTTLE cell was controlled using a potentiostat (Toho Technical Research; model 2020). The electrode potential was expressed against the standard hydrogen electrode (SHE). After a series of spectroelectrochemical measurements, an accurate electrode potential of the Ag/AgCl/3M KCl reference electrode (+208 mV vs. SHE) was evaluated using a standard Ag/AgCl/saturated KCl electrode (+199 mV vs. SHE), and electrode potentials were corrected. To estimate the Em values of the QB redox reactions, single-beam spectra with 20 scans (10-s accumulation) were recorded before and after single-flash illumination on the PSII sample poised at different electrode potentials, and difference spectra upon illumination were calculated. At each electrode potential, the sample was stabilized for 60 min before measurement to equilibrate the redox reaction. To obtain the standard spectra of S2QB/S1QB and S2/S1 differences with high S/N ratios, 20-scan measurements before and after flash illumination on intact PSII at electrode potentials of +350 and +470 mV, respectively, were repeated 24 times with a dark interval of 30 min, and the spectra were averaged. In the case of the S2/S1 difference spectrum, a 4-s delay was inserted after flash illumination to reduce the contamination of the QB/QB signals. For the high-quality QB/QB difference spectrum of Mn-depleted PSII, 20-scan measurements before and 10 s after a single flash at +250 mV were repeated 24 times with a dark interval of 1 h, and the spectra were averaged. The 20-s delay was inserted to reduce the contribution of QA, which relaxed in several seconds.

SI Text: Thermodynamic Simulation of Two-Electron Redox Reactions

We consider that the reduction of QB proceeds by two one-electron processes:

QB+eQBEm1, [S1]
QB+e+2H+PQH2Em2, [S2]

where Em1 and Em2 are the redox potentials of the first and second reduction, respectively [corresponding to Em(QB/QB) and Em(PQH2/QB), respectively, in the main text]. Note that the second reduction is coupled with the uptake of two protons forming neutral plastoquinol. In equilibrium, the electrode potential E is given by the Nernst equations as functions of the concentrations of QB, QB, PQH2, and H+.

E=Em1+RTFln[QB][QB], [S3]
E=Em2+RTFln[QB][H+]2[QH2], [S4]

where R is the gas constant, T is the absolute temperature, and F is the Faraday constant. The initial concentration of QB, [QB]0, is expressed as follows:

[QB]0=[QB]+[QB]+[PQH2]. [S5]

From Eqs. S3, S4, and S5, the mole fraction of QB and QB, [QB]/[QB]0 and [QB]/[QB]0, respectively, are expressed as follows:

[QB][QB]0=11+exp{FRT(Em1E)}+exp{FRT(Em1+Em22E)4.605pH}, [S6]
[QB][QB]0=11+exp{FRT(EEm1)}+exp{FRT(Em2E)4.605pH}. [S7]

From Eqs. S6 and S7, the mole fractions of QB and QB are drawn against the electrode potential E in Fig. S3, assuming different Em gaps (ΔEm) between Em2 and Em1Em = Em1Em2) centered at +155 mV. When Em1 and Em2 are well separated with Em1 > Em2, the mole fractions of QB and QB show the Nernst curves of one-electron redox processes (e.g., blue solid and dashed lines, respectively, for the case of ΔEm = 240 mV). When Em1 and Em2 get closer (e.g., ΔEm = 40 mV; light blue lines), the slope of the curve of the QB fraction becomes sharper and the QB fraction is significantly reduced on the low potential side of Em1. Even if Em1 and Em2 are the same, the slope of the simulated curve of the QB fraction (green solid line) is still lower than the ideal Nernst curve for a two-electron reaction (black solid line). When the order of Em1 and Em2 is “inverted” (i.e., Em1 < Em2) and they are separated by >180 mV (ΔEm less than −180 mV), the simulated curve of the QB fraction becomes virtually the same as the ideal two-electron reaction and the QB fraction disappears (e.g., ΔEm = −180 mV; red solid and dashed lines).

The intensity of the positive peak at 1,745 cm−1 in the flash-induced FTIR spectra (Fig. 3A) reflects both the increase of QB by single reduction of QB and the decrease of QB by reduction of electrochemically produced QB. Thus, the intensity ratio (α) reflects the difference between the mole fractions of QB and QB, [QB]/[QB]0 – [QB]/[QB]0. The curves of ([QB]/[QB]0 – [QB]/[QB]0) against the potential E in various ΔEm cases are presented in Fig. S4. When Em2 is lower than Em1 or when the two Em values are close enough even with higher Em2 than Em1, a negative intensity appears on the low potential side of the curve. The absence of such a negative intensity in the experimental plots (Fig. 3C and Fig. S5) indicates that Em2 is higher than Em1 with a relatively large gap. The simulation curves obtained by fitting experimental plots of intact and Mn-depleted PSII samples are depicted in Fig. S5 together with the calculated mole fractions of QB and QB.

Acknowledgments

This study was supported by JSPS KAKENHI [25410009 (to Y.K.), 24000018, 24107003, and 25291033 (to T.N.), and 26840091 (to R.N.)] and by Grant for Basic Science Research Projects from The Sumitomo Foundation (to Y.K.).

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1520211113/-/DCSupplemental.

References

  • 1.Barber J. Photosynthetic energy conversion: Natural and artificial. Chem Soc Rev. 2009;38(1):185–196. doi: 10.1039/b802262n. [DOI] [PubMed] [Google Scholar]
  • 2.Renger G. Photosynthetic water splitting: Apparatus and mechanism. In: Eaton-Rye JJ, Tripathy BC, Sharkey TD, editors. Photosynthesis: Plastid Biology, Energy Conversion and Carbon Assimilation. Springer; Dordrecht, The Netherlands: 2012. pp. 359–414. [Google Scholar]
  • 3.Messinger J, Noguchi T, Yano J. Photosynthetic O2 evolution. In: Wydrzynski T, Hillier W, editors. Molecular Solar Fuels. Royal Society of Chemistry; Cambridge, UK: 2011. pp. 163–207. [Google Scholar]
  • 4.Rappaport F, Diner BA. 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. 2008;252(3-4):259–272. [Google Scholar]
  • 5.Cardona T, Sedoud A, Cox N, Rutherford AW. Charge separation in photosystem II: A comparative and evolutionary overview. Biochim Biophys Acta. 2012;1817(1):26–43. doi: 10.1016/j.bbabio.2011.07.012. [DOI] [PubMed] [Google Scholar]
  • 6.Joliot P, Barbieri G, Chabaud R. Model of the System II photochemical centers. Photochem Photobiol. 1969;10(5):309–329. [Google Scholar]
  • 7.Kok B, Forbush B, McGloin M. Cooperation of charges in photosynthetic O2 evolution—I. A linear four step mechanism. Photochem Photobiol. 1970;11(6):457–475. doi: 10.1111/j.1751-1097.1970.tb06017.x. [DOI] [PubMed] [Google Scholar]
  • 8.Petrouleas V, Crofts AR. The quinone iron acceptor complex. In: Wydrzynski T, Satoh K, editors. Photosystem II: The Light-Driven Water:Plastoquinone Oxidoreductase. Springer; Dordrecht, The Netherlands: 2005. pp. 177–206. [Google Scholar]
  • 9.Müh F, Glöckner C, Hellmich J, Zouni A. Light-induced quinone reduction in photosystem II. Biochim Biophys Acta. 2012;1817(1):44–65. doi: 10.1016/j.bbabio.2011.05.021. [DOI] [PubMed] [Google Scholar]
  • 10.Guskov A, et al. Cyanobacterial photosystem II at 2.9-Å resolution and the role of quinones, lipids, channels and chloride. Nat Struct Mol Biol. 2009;16(3):334–342. doi: 10.1038/nsmb.1559. [DOI] [PubMed] [Google Scholar]
  • 11.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(7345):55–60. doi: 10.1038/nature09913. [DOI] [PubMed] [Google Scholar]
  • 12.Saito K, Rutherford AW, Ishikita H. Mechanism of proton-coupled quinone reduction in photosystem II. Proc Natl Acad Sci USA. 2013;110(3):954–959. doi: 10.1073/pnas.1212957110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ashizawa R, Noguchi T. Effects of hydrogen bonding interactions on the redox potential and molecular vibrations of plastoquinone as studied using density functional theory calculations. Phys Chem Chem Phys. 2014;16(24):11864–11876. doi: 10.1039/c3cp54742f. [DOI] [PubMed] [Google Scholar]
  • 14.Rutherford AW, Osyczka A, Rappaport F. Back-reactions, short-circuits, leaks and other energy wasteful reactions in biological electron transfer: Redox tuning to survive life in O2. FEBS Lett. 2012;586(5):603–616. doi: 10.1016/j.febslet.2011.12.039. [DOI] [PubMed] [Google Scholar]
  • 15.Krieger-Liszkay A, Fufezan C, Trebst A. Singlet oxygen production in photosystem II and related protection mechanism. Photosynth Res. 2008;98(1-3):551–564. doi: 10.1007/s11120-008-9349-3. [DOI] [PubMed] [Google Scholar]
  • 16.Vass I, Cser K. Janus-faced charge recombinations in photosystem II photoinhibition. Trends Plant Sci. 2009;14(4):200–205. doi: 10.1016/j.tplants.2009.01.009. [DOI] [PubMed] [Google Scholar]
  • 17.Vass I. Role of charge recombination processes in photodamage and photoprotection of the photosystem II complex. Physiol Plant. 2011;142(1):6–16. doi: 10.1111/j.1399-3054.2011.01454.x. [DOI] [PubMed] [Google Scholar]
  • 18.Kós PB, Deák Z, Cheregi O, Vass I. Differential regulation of psbA and psbD gene expression, and the role of the different D1 protein copies in the cyanobacterium Thermosynechococcus elongatus BP-1. Biochim Biophys Acta. 2008;1777(1):74–83. doi: 10.1016/j.bbabio.2007.10.015. [DOI] [PubMed] [Google Scholar]
  • 19.Merry SA, et al. Modulation of quantum yield of primary radical pair formation in photosystem II by site-directed mutagenesis affecting radical cations and anions. Biochemistry. 1998;37(50):17439–17447. doi: 10.1021/bi980502d. [DOI] [PubMed] [Google Scholar]
  • 20.Shibuya Y, et al. Hydrogen bond interactions of the pheophytin electron acceptor and its radical anion in photosystem II as revealed by Fourier transform infrared difference spectroscopy. Biochemistry. 2010;49(3):493–501. doi: 10.1021/bi9018829. [DOI] [PubMed] [Google Scholar]
  • 21.Krieger A, Weis E. Energy-dependent of chlorophyll-a-fluorescence: The involvement of proton-calcium exchange at photosystem 2. Photosynthetica. 1992;27(1-2):89–98. [Google Scholar]
  • 22.Krieger A, Rutherford AW, Johnson GN. On the determination of redox midpoint potential of the primary quinone electron acceptor, QA, in photosystem II. Biochim Biophys Acta. 1995;1229(2):193–201. [Google Scholar]
  • 23.Johnson GN, Rutherford AW, Krieger A. A change in the midpoint potential of the quinone QA in photosystem II associated with photoactivation of oxygen evolution. Biochim Biophys Acta. 1995;1229(2):202–207. [Google Scholar]
  • 24.Shibamoto T, Kato Y, Sugiura M, Watanabe T. Redox potential of the primary plastoquinone electron acceptor QA in photosystem II from Thermosynechococcus elongatus determined by spectroelectrochemistry. Biochemistry. 2009;48(45):10682–10684. doi: 10.1021/bi901691j. [DOI] [PubMed] [Google Scholar]
  • 25.Shibamoto T, et al. Species-dependence of the redox potential of the primary quinone electron acceptor QA in photosystem II verified by spectroelectrochemistry. FEBS Lett. 2010;584(8):1526–1530. doi: 10.1016/j.febslet.2010.03.002. [DOI] [PubMed] [Google Scholar]
  • 26.Ido K, et al. High and low potential forms of the QA quinone electron acceptor in photosystem II of Thermosynechococcus elongatus and spinach. J Photochem Photobiol B. 2011;104(1-2):154–157. doi: 10.1016/j.jphotobiol.2011.02.010. [DOI] [PubMed] [Google Scholar]
  • 27.Allakhverdiev SI, et al. Redox potentials of primary electron acceptor quinone molecule (QA)− and conserved energetics of photosystem II in cyanobacteria with chlorophyll a and chlorophyll d. Proc Natl Acad Sci USA. 2011;108(19):8054–8058. doi: 10.1073/pnas.1100173108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Diner BA. Dependence of the deactivation reactions of photosystem II on the redox state of plastoquinone pool A varied under anaerobic conditions; equilibria on the acceptor side of photosystem II. Biochim Biophys Acta. 1977;460(2):247–258. doi: 10.1016/0005-2728(77)90211-0. [DOI] [PubMed] [Google Scholar]
  • 29.Robinson HH, Crofts AR. Kinetics of the oxidation-reduction reactions of the photosystem II quinone acceptor complex, and the pathway for deactivation. FEBS Lett. 1983;153(1):221–226. [Google Scholar]
  • 30.Crofts AR, Wraight CA. The electrochemical domain of photosynthesis. Biochim Biophys Acta. 1983;726(3):149–185. [Google Scholar]
  • 31.Minagawa J, Narusaka Y, Inoue Y, Satoh K. Electron transfer between QA and QB in photosystem II is thermodynamically perturbed in phototolerant mutants of Synechocystis sp. PCC 6803. Biochemistry. 1999;38(2):770–775. doi: 10.1021/bi981217x. [DOI] [PubMed] [Google Scholar]
  • 32.Rose S, et al. D1-arginine257 mutants (R257E, K, and Q) of Chlamydomonas reinhardtii have a lowered QB redox potential: Analysis of thermoluminescence and fluorescence measurements. Photosynth Res. 2008;98(1-3):449–468. doi: 10.1007/s11120-008-9351-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Papageorgiou GC, Govindjee Photosystem II fluorescence: Slow changes—scaling from the past. J Photochem Photobiol B. 2011;104(1-2):258–270. doi: 10.1016/j.jphotobiol.2011.03.008. [DOI] [PubMed] [Google Scholar]
  • 34.Cao J, Govindjee Chlorophyll a fluorescence transient as an indicator of active and inactive photosystem II in thylakoid membranes. Biochim Biophys Acta. 1990;1015(2):180–188. doi: 10.1016/0005-2728(90)90018-y. [DOI] [PubMed] [Google Scholar]
  • 35.Lazár D. Chlorophyll a fluorescence induction. Biochim Biophys Acta. 1999;1412(1):1–28. doi: 10.1016/s0005-2728(99)00047-x. [DOI] [PubMed] [Google Scholar]
  • 36.van Gorkom HJ. Identification of the reduced primary electron acceptor of photosystem II as a bound semiquinone anion. Biochim Biophys Acta. 1974;347(3):439–442. doi: 10.1016/0005-2728(74)90081-4. [DOI] [PubMed] [Google Scholar]
  • 37.Pulles MPJ, Van Gorkom HJ, Willemsen JG. Absorbance changes due to the charge-accumulating species in system 2 of photosynthesis. Biochim Biophys Acta. 1976;449(3):536–540. doi: 10.1016/0005-2728(76)90162-6. [DOI] [PubMed] [Google Scholar]
  • 38.Fufezan C, Zhang C, Krieger-Liszkay A, Rutherford AW. Secondary quinone in photosystem II of Thermosynechococcus elongatus: Semiquinone-iron EPR signals and temperature dependence of electron transfer. Biochemistry. 2005;44(38):12780–12789. doi: 10.1021/bi051000k. [DOI] [PubMed] [Google Scholar]
  • 39.Sedoud A, et al. Semiquinone-iron complex of photosystem II: EPR signals assigned to the low-field edge of the ground state doublet of QA•−Fe2+ and QB•−Fe2+ Biochemistry. 2011;50(27):6012–6021. doi: 10.1021/bi200313p. [DOI] [PubMed] [Google Scholar]
  • 40.Mäntele W. Reaction-induced infrared difference spectroscopy for the study of protein function and reaction mechanisms. Trends Biochem Sci. 1993;18(6):197–202. doi: 10.1016/0968-0004(93)90186-q. [DOI] [PubMed] [Google Scholar]
  • 41.Noguchi T, Berthomieu C. Molecular analysis by vibrational spectroscopy. In: Wydrzynski T, Satoh K, editors. Photosystem II: The Light-Driven Water:Plastoquinone Oxidoreductase. Springer; Dordrecht, The Netherlands: 2005. pp. 367–387. [Google Scholar]
  • 42.Berthomieu C, Hienerwadel R. Fourier transform infrared (FTIR) spectroscopy. Photosynth Res. 2009;101(2-3):157–170. doi: 10.1007/s11120-009-9439-x. [DOI] [PubMed] [Google Scholar]
  • 43.Chu H-A. Fourier transform infrared difference spectroscopy for studying the molecular mechanism of photosynthetic water oxidation. Front Plant Sci. 2013;4:146. doi: 10.3389/fpls.2013.00146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Debus RJ. FTIR studies of metal ligands, networks of hydrogen bonds, and water molecules near the active site Mn₄CaO₅ cluster in photosystem II. Biochim Biophys Acta. 2015;1847(1):19–34. doi: 10.1016/j.bbabio.2014.07.007. [DOI] [PubMed] [Google Scholar]
  • 45.Noguchi T. Fourier transform infrared difference and time-resolved infrared detection of the electron and proton transfer dynamics in photosynthetic water oxidation. Biochim Biophys Acta. 2015;1847(1):35–45. doi: 10.1016/j.bbabio.2014.06.009. [DOI] [PubMed] [Google Scholar]
  • 46.Suzuki H, Nagasaka MA, Sugiura M, Noguchi T. Fourier transform infrared spectrum of the secondary quinone electron acceptor QB in photosystem II. Biochemistry. 2005;44(34):11323–11328. doi: 10.1021/bi051237g. [DOI] [PubMed] [Google Scholar]
  • 47.Dutton PL. Redox potentiometry: Determination of midpoint potentials of oxidation-reduction components of biological electron-transfer systems. Methods Enzymol. 1978;54:411–435. doi: 10.1016/s0076-6879(78)54026-3. [DOI] [PubMed] [Google Scholar]
  • 48.Melin F, Hellwig P. Recent advances in the electrochemistry and spectroelectrochemistry of membrane proteins. Biol Chem. 2013;394(5):593–609. doi: 10.1515/hsz-2012-0344. [DOI] [PubMed] [Google Scholar]
  • 49.Kato Y, Sugiura M, Oda A, Watanabe T. Spectroelectrochemical determination of the redox potential of pheophytin a, the primary electron acceptor in photosystem II. Proc Natl Acad Sci USA. 2009;106(41):17365–17370. doi: 10.1073/pnas.0905388106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Kato Y, Noguchi T. Long-range interaction between the Mn4CaO5 cluster and the non-heme iron center in photosystem II as revealed by FTIR spectroelectrochemistry. Biochemistry. 2014;53(30):4914–4923. doi: 10.1021/bi500549b. [DOI] [PubMed] [Google Scholar]
  • 51.Mäntele WG, Wollenweber AM, Nabedryk E, Breton J. Infrared spectroelectrochemistry of bacteriochlorophylls and bacteriopheophytins: Implications for the binding of the pigments in the reaction center from photosynthetic bacteria. Proc Natl Acad Sci USA. 1988;85(22):8468–8472. doi: 10.1073/pnas.85.22.8468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Bauscher M, Nabedryk E, Bagley K, Breton J, Mäntele W. Investigation of models for photosynthetic electron acceptors: Infrared spectroelectrochemistry of ubiquinone and its anions. FEBS Lett. 1990;261(1):191–195. [Google Scholar]
  • 53.Gamage RSKA, Umapathy S, McQuillan AJ. OTTLE cell study of the UV-visible and FTIR spectroelectrochemistry of the radical anion and dianion of 1,4-benzoquinone in DMSO solutions. J Electroanal Chem. 1990;284(1):229–235. [Google Scholar]
  • 54.Best SP. Spectroelectrochemistry of hydrogenase enzymes and related compounds. Coord Chem Rev. 2005;249(15-16):1536–1554. [Google Scholar]
  • 55.Berthomieu C, Nabedryk E, Mäntele W, Breton J. Characterization by FTIR spectroscopy of the photoreduction of the primary quinone acceptor QA in photosystem II. FEBS Lett. 1990;269(2):363–367. doi: 10.1016/0014-5793(90)81194-s. [DOI] [PubMed] [Google Scholar]
  • 56.Takano A, Takahashi R, Suzuki H, Noguchi T. Herbicide effect on the hydrogen-bonding interaction of the primary quinone electron acceptor QA in photosystem II as studied by Fourier transform infrared spectroscopy. Photosynth Res. 2008;98(1-3):159–167. doi: 10.1007/s11120-008-9302-5. [DOI] [PubMed] [Google Scholar]
  • 57.Noguchi T, Ono T, Inoue Y. Direct detection of a carboxylate bridge between Mn and Ca2+ in the photosynthetic oxygen-evolving center by means of Fourier transform infrared spectroscopy. Biochim Biophys Acta. 1995;1228(2-3):189–200. [Google Scholar]
  • 58.Noguchi T, Sugiura M. Analysis of flash-induced FTIR difference spectra of the S-state cycle in the photosynthetic water-oxidizing complex by uniform 15N and 13C isotope labeling. Biochemistry. 2003;42(20):6035–6042. doi: 10.1021/bi0341612. [DOI] [PubMed] [Google Scholar]
  • 59.Kimura Y, Mizusawa N, Ishii A, Yamanari T, Ono TA. Changes of low-frequency vibrational modes induced by universal 15N- and 13C-isotope labeling in S2/S1 FTIR difference spectrum of oxygen-evolving complex. Biochemistry. 2003;42(45):13170–13177. doi: 10.1021/bi035420q. [DOI] [PubMed] [Google Scholar]
  • 60.Bard AJ, Faulkner LR. Electrochemical Methods. 2nd Ed Wiley; New York: 2001. [Google Scholar]
  • 61.Rich PR, Bendall DS. The kinetics and thermodynamics of the reduction of cytochrome c by substituted p-benzoquinols in solution. Biochim Biophys Acta. 1980;592(3):506–518. doi: 10.1016/0005-2728(80)90095-x. [DOI] [PubMed] [Google Scholar]
  • 62.Krieger-Liszkay A, Rutherford AW. Influence of herbicide binding on the redox potential of the quinone acceptor in photosystem II: Relevance to photodamage and phytotoxicity. Biochemistry. 1998;37(50):17339–17344. doi: 10.1021/bi9822628. [DOI] [PubMed] [Google Scholar]
  • 63.Iwai M, et al. The PsbK subunit is required for the stable assembly and stability of other small subunits in the PSII complex in the thermophilic cyanobacterium Thermosynechococcus elongatus BP-1. Plant Cell Physiol. 2010;51(4):554–560. doi: 10.1093/pcp/pcq020. [DOI] [PubMed] [Google Scholar]
  • 64.Nakamura S, Nagao R, Takahashi R, Noguchi T. Fourier transform infrared detection of a polarizable proton trapped between photooxidized tyrosine YZ and a coupled histidine in photosystem II: Relevance to the proton transfer mechanism of water oxidation. Biochemistry. 2014;53(19):3131–3144. doi: 10.1021/bi500237y. [DOI] [PubMed] [Google Scholar]
  • 65.Sugiura M, Inoue Y. Highly purified thermo-stable oxygen-evolving photosystem II core complex from the thermophilic cyanobacterium Synechococcus elongatus having His-tagged CP43. Plant Cell Physiol. 1999;40(12):1219–1231. doi: 10.1093/oxfordjournals.pcp.a029510. [DOI] [PubMed] [Google Scholar]

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES