Skip to main content
iScience logoLink to iScience
. 2021 Jun 11;24(7):102719. doi: 10.1016/j.isci.2021.102719

A dimeric chlorophyll electron acceptor differentiates type I from type II photosynthetic reaction centers

Michael Gorka 1, Philip Charles 2, Vidmantas Kalendra 2, Amgalanbaatar Baldansuren 2, KV Lakshmi 2,4,, John H Golbeck 1,3,∗∗
PMCID: PMC8267441  PMID: 34278250

Summary

This research addresses one of the most compelling issues in the field of photosynthesis, namely, the role of the accessory chlorophyll molecules in primary charge separation. Using a combination of empirical and computational methods, we demonstrate that the primary acceptor of photosystem (PS) I is a dimer of accessory and secondary chlorophyll molecules, Chl2A and Chl3A, with an asymmetric electron charge density distribution. The incorporation of highly coupled donors and acceptors in PS I allows for extensive delocalization that prolongs the lifetime of the charge-separated state, providing for high quantum efficiency. The discovery of this motif has widespread implications ranging from the evolution of naturally occurring reaction centers to the development of a new generation of highly efficient artificial photosynthetic systems.

Video abstract

Download video file (41.5MB, mp4)

Subject areas: chemistry, inorganic chemistry, molecular inorganic chemistry, theoretical photochemistry

Graphical abstract

graphic file with name fx1.jpg

Highlights

  • Electron-nuclear hyperfine couplings indicate that A0·– is a chlorophyll dimer

  • DFT calculations reveal electron delocalization over both Chl2 and Chl3 in A0·–

  • Photosystem I utilizes Chl dimers as both the primary donor (P700) and acceptor (A0)

  • A0 is predicted to be a dimer in type I RCs and a monomer in type II RCs


Chemistry; Inorganic chemistry; Molecular inorganic chemistry; Theoretical photochemistry

Introduction

Nature powers present day Earth with oxygenic photosynthesis. In this highly evolved process, solar energy is used to convert carbon dioxide and water into carbohydrates, releasing dioxygen into the atmosphere as a byproduct. Chlorophyll (Chl) molecules are central to oxygenic photosynthesis as they participate in both energy and electron transfer (ET) (Rabinowitch and Govindjee, 1965; Thornber, 1975; Katz et al., 1978; Anderson, 1982). While there are a wide variety of Chls in nature that absorb light in different regions of the visible and near-IR spectrum, the arrangement and function of Chls are conserved in reaction center (RC) proteins across different species of photosynthetic organisms (Olson and Pierson, 1987; Meyer, 1994; Orf et al., 2018; Chen et al., 2020). The type I photosynthetic RC, photosystem (PS) I (Figure 1A), in cyanobacteria consists of approx. 96 peripheral Chl a molecules (Jordan et al., 2001). Additionally, the heterodimeric core of PS I, which comprised the PsaA and PsaB polypeptides, contains three Chl a molecules (Chl1A/1B, Chl2A/2B and Chl3A/3B) in the pseudo-symmetric A and B branch of cofactors (Figure 1B) (Fromme et al., 2001; Jordan et al., 2001). ET in PS I is initiated when the antenna Chl absorb light and transfer the excitation energy with high quantum efficiency to two electronically coupled Chl1A and Chl1B molecules known as the primary donor, P700 (Şener et al., 2011). The photoexcited P700 state leads to rapid ET, resulting in the formation of the first detectable charge-separated state, P700·+A0A·– or P700·+A0B·–, respectively (Wasielewski et al., 1987; Shelaev et al., 2010; Chauvet et al., 2012; Milanovsky et al., 2014). Both branches are active in PS I; however, the A-branch is preferred over the B-branch by a factor of ∼2 in cyanobacteria (Dashdorj et al., 2005). The initial charge separation is stabilized by forward ET to a low-potential phylloquinone, A1A or A1B (Deligiannakis et al., 1998; Ishikita and Knapp, 2003; van der Est, 2006; Srinivasan and Golbeck, 2009), with ET along the A- and B-branches converging at the [4Fe–4S] cluster, FX. Subsequently, the electron is transferred serially through the terminal [4Fe–4S] clusters, FA and FB, in the PsaC polypeptide (Golbeck, 2006) (Figure 1B) which results is the generation of reducing equivalents that are stored as NADPH for use in the carbon fixation reactions.

Figure 1.

Figure 1

Crystal structure of photosystem I and electron transfer cofactors

The (A) polypeptide subunits and (B) electron transfer cofactors as observed in the 2.5 Å resolution X-ray crystal structure of cyanobacterial PS I (Jordan et al., 2001). The core subunits, PsaA, PsaB, and PsaC are displayed as salmon, purple, and green ribbons, respectively. Electron transfer cofactors are denoted with the commonly accepted spectroscopic labels, as well as crystallographic labels, for the six Chl cofactors.

While the kinetics and thermodynamics of each ET step beyond A0 are well characterized (Kurashov et al., 2018; Ptushenko et al., 2008; Srinivasan and Golbeck, 2009), less is known about the mechanism of charge separation and the electronic structure of the A0A/0B·– state. Early research on PS I led to differing models of the reduced primary acceptor, A0A/0B·–. Picosecond optical spectroscopic studies revealed shifts in the absorbance features at 695 nm that suggested the primary acceptor was Chl a dimer (Shuvalov et al., 1979a, 1979b, 1979c). Subsequently, analysis of EPR line widths and electron nuclear double resonance (ENDOR) data concluded that the line width of the reduced primary acceptor was too large for a dimeric complex, instead suggesting the acceptor was better described as monomeric Chl a molecule (Fajer et al., 1980; Forman et al., 1981). The high-resolution X-ray crystal structure has established that there are three distinct pairs of Chl a molecules, Chl1A/1B, Chl2A/2B and Chl3A/3B, in the central core of cyanobacterial PS I (Fromme et al., 2001; Jordan et al., 2001). Chl1A/1B are in close proximity, providing confirmation that P700 functions as a heterodimer of Chl a/Chl a’ molecules (Figures 4C and 4D). Interestingly, Chl2A/3A and Chl2B/3B are oriented in a similar fashion to P700 (Figures 4A and 4B), and while they have slightly longer inter-cofactor distances, their structure could allow for the possibility of a dimeric complex.

Figure 4.

Figure 4

Structural comparison of the A0 and P700 dimers in PS I

Relative orientation and distances of the (A and B) Chl3A and Chl2A cofactors in the A0 acceptor and (C and D) Chl1A/1B in the primary donor, P700, of cyanobacterial PS I (Jordan et al., 2001).

It is important to note that while the spatial orientation of the Chl1A/1B and Chl2A/2B/Chl3A/3B molecules are similar, there are important differences in their respective ligand environments. The Mg2+ ion of Chl1A/1B is axially coordinated by a nitrogen atom of a histidine residue (His680PsaA and His660PsaB), whereas Chl2A/2B and Chl3A/3B are coordinated to a water ligand and soft base sulfur ligand from a methionine residue (Met688PsaA and Met668PsaB), respectively (Jordan et al., 2001). The latter is not predicted by the normative rules of inorganic chemistry (Pearson, 1963; Ayers, 2005), as the central Mg2+ ion of Chl a is a relatively hard acid that is usually coordinated to hard nitrogen (and sometimes oxygen) ligands. In contrast, a sulfur ligand is a highly polarizable soft Lewis base. This suggests that the axial coordination of the central Mg2+ ion could be instrumental in achieving a low midpoint potential that is necessary for establishing extremely rapid charge separation within ∼100 fs of photoexcitation (Shelaev et al., 2010). It is therefore not surprising that replacement of the axial Met ligand with a harder base (His) displays profound effects on the rates and efficiency of ET in both branches (Sun et al., 2014). Additional mutations to this ligand (Xu et al., 2003a, 2003b; Cohen et al., 2004; Dashdorj et al., 2005; Savitsky et al., 2009) and to the hydrogen (H-)bonded tyrosine residue (Li et al., 2006) have been shown to alter ET, but they have not fully elucidated the nature of the A0·− state.

P700 functions as the primary electron donor (Brettel and Leibl, 2001), but to date, the role of the Chl2A/2B cofactor, which is located within a center-to-center distance of ∼8 Å of Chl3A/3B, remains unclear (Figure 1B) (Jordan et al., 2001), and several compelling models have been put forward in recent years. Ultra-fast transient absorption spectroscopy on genetic variants of PS I isolated from Chlamydomonas reinhardtii containing changes in the protein environment surrounding the Chl2A/2B and Chl3A/3B cofactors have suggested that charge separation may be initiated within this pair, after which the hole is rapidly re-reduced by P700 (Müller et al., 2003, 2010; Holzwarth et al., 2006). Other studies, however, have argued that the redox potentials of Chl2A/2B and Chl3A/3B would make initial charge separation from this pair thermodynamically unfavorable (Ptushenko et al., 2008). Instead, it has been suggested that the short distance and redox potential difference between Chl2A/2B and Chl3A/3B could allow them to function as a pseudo-dimer, asymmetrically sharing electron spin density in a ratio of approximately Chl2A/2B:Chl3A/3B::1:7 in the initial charge separated state (Ptushenko et al., 2008). Yet others suggest that these six Chl a molecules that comprise the core of PS I are electronically coupled (Gibasiewicz et al., 2003; Yin et al., 2007; Cherepanov et al., 2020).

Cherepanov et al. (Cherepanov et al., 2020) recently proposed that the primary donor Chl, Chl1A/1B, and the nearest Chl, Chl2A/2B, are combined in a unified exciplex, (Chl2AChl1AChl1BChl2B). Using femtosecond pump-probe spectroscopy, they demonstrated that the excitation of the exciplex led to the formation of an excited state, (Chl2AChl1AChl1BChl2B), that was mixed with two charge-transfer states P700·+Chl2A·− and P700·+Chl2B·−. The negative charge was predominantly localized on Chl2A/2B and subsequently, dielectric polarization of the protein further localized the charge on Chl3A/3B. It was suggested that the electron density was subsequently redistributed between the A- and B-branch that largely generated the A1A·− state. However, empirical evidence has been insufficient to determine the precise role of Chl2A/2B and its interaction with Chl3A/3B in the charge-separated state, P700·+A0·–.

Continuous-wave (CW) and pulsed electron paramagnetic resonance (EPR) spectroscopy have been employed to study the molecular structure and function of the Chl cofactors (Commoner et al., 1956; Käβ et al., 1995, 1996; Käβ and Lubitz, 1996). These methods have made it possible to probe the unpaired electron spin that is formed upon photo-oxidation or reduction of Chl in photosynthetic RCs. In fact, one of the earliest applications of EPR spectroscopy by Commoner et al. detected the steady-state photoaccumulation of the primary donor cation, P700·+, of PS I (Commoner et al., 1956). Subsequently, the signal was attributed to a strongly electronically coupled Chl a dimer (Chl1A and Chl1B in Figure 1B) (Norris et al., 1971; Davis et al., 1993; Mac et al., 1998; Käss et al., 2001), given that the line width was narrower than that of a monomeric Chl a (Norris et al., 1971). It was shown that the relationship, ΔHN = 1/√N⋅ΔHM (where ΔHN and ΔHM are the line width of the multimeric Chl cation with spin delocalization, ChlN·+, and monomeric Chl·+ cation, respectively), accounted for the delocalization of the unpaired electron spin over the dimeric Chl1A/1B cofactors of P700·+ (Norris et al., 1971). More recently, pulsed EPR methods, such as, transient EPR, ENDOR, and electron spin echo envelope modulation (ESEEM) spectroscopy along with computational modeling have provided a wealth of information on the rates of ET and the electronic structure of the cofactors of PS I (Mac et al., 1996, 1998; Käss et al., 2001; Deligiannakis and Rutherford, 2001; Srinivasan et al., 2009, 2011a, 2011b; Poluektov et al., 2005; Norris et al., 1974) and other RCs (Deligiannakis et al., 1999; Lubitz et al., 2002; Britt et al., 2004; Schnegg et al., 2007).

While CW and pulsed EPR spectroscopy of the Chl a·– anion in vitro have provided initial models for the A0 acceptor of PS I (Hoff et al., 1982; Rigby et al., 2003), the precise electronic structure of A0·– has yet to be determined in vivo. In certain cases, in vitro studies can serve as adequate models for in vivo Chl molecules, but in the case of A0, these lack the smart matrix effects of the protein environment that influence its function. The 2.5 Å resolution X-ray crystal structure of PS I (Jordan et al., 2001) sets the stage for the application of state-of-the-art experiments and theory to interrogate the molecular and electronic structure of A0 within the protein. In this work, we utilize a combination of two-dimensional (2D) hyperfine sublevel correlation (HYSCORE) spectroscopy and density functional theory (DFT) to probe the electronic structure of the primary acceptor, A0·–. The paramagnetic A0·– state of PS I is uniquely poised to provide information on the electron-nuclear hyperfine interactions with constituent 1H and 14N atoms. HYSCORE spectroscopy is ideally suited for this study as it provides enhanced resolution from the detection of the hyperfine signals in 2D space, which alleviates the problem of spectral overcrowding from the signals of multiple 1H and 14N atoms present in the A0·– state (Harmer et al., 2009). Additionally, the nuclear transitions are correlated in the two dimensions, which greatly simplifies the analysis and interpretation of the spectra. We have previously demonstrated the application of 2D 1H and 14N HYSCORE to determine the strength and orientation of H-bonds of the quinone cofactors of PS I and PS II that provided insight on the role of H-bonds in the tuning of the midpoint potential of the quinone acceptors (Chatterjee et al., 2011; Srinivasan et al., 2011a).

In conjunction with 2D 14N and 1H HYSCORE measurements of the A0·– state, we performed advanced DFT calculations on an extensive computational model of PS I that includes the Chl1A/1B, Chl2A/2B, Chl3A, and A1A cofactors with all of the associated amino acid and water ligands. This allowed us to model the effects of the unusual axial coordination of the Chl, H-bonding interactions and proximity effects from the cofactors in the A-branch of PS I. This combination of experiment and theory has led to the first direct determination of electron delocalization over the Chl2A and Chl3A dimer in the A0·– state, which was previously believed to be a monomeric Chl electron acceptor. The uncovering of a dimeric Chl motif in the A0·– state has widespread implications ranging from the evolution of RCs in natural photosynthesis to the development of a new generation of bio-inspired artificial photosynthetic systems.

Results

Two-dimensional HYSCORE spectroscopy reveals electron spin delocalization in the A0·– state

Cryogenic illumination of PS I in which forward ET to A1 is inhibited and the FeS clusters are reduced leads to the photoaccumulation of the A0·– state. While the field-sweep electron-spin-echo signal indicates the presence of an unpaired electron spin, S, of ½ (Figure S1), the weak electron-nuclear hyperfine interactions of the electron spin with the surrounding magnetic nuclei are not resolved in the X-band A0·– spectrum. In comparison, the 2D HYSCORE spectrum clearly displays two sets of distinct cross-peaks centered at 3–4 MHz and 14.7 MHz that arise from the hyperfine interactions of the unpaired electron spin with the proximal nitrogen-14 (14N, nuclear spin of I = 1) and hydrogen (1H, I = ½) atoms of the A0·– state, respectively (Figure S2). The 14N and 1H cross-peaks are separated in frequency space due to a large difference in their respective gyromagnetic ratios (γ) (19.3 × 106 and 267.5 × 106 rad·s−1·T−1 for 14N and 1H, respectively). For the purpose of this study, we will focus on the 14N hyperfine interactions of the A0·– state (see supplemental information for a description of the 1H hyperfine interactions (Figure S3)).

The cross-peaks in the experimental and simulated 2D 14N HYSCORE spectrum of the A0·– state in Figures 2A–2D and S4 arise from hyperfine interactions with three 14N atoms, labeled NA – NC. Typically, the spectral features from 14N atoms that are strongly hyperfine-coupled to the unpaired electron spin (A > 2νN, where A is the hyperfine coupling constant) lead to cross-peaks in the (–,+) quadrant of the spectrum, while those from the weakly coupled 14N atoms (A < 2νN) are observed in the (+,+) quadrant (Höfer et al., 1986). Thus, the pair of overlapping cross-peaks at (−6.2, 2.8) MHz that are well-pronounced, intense, and symmetric about the anti-diagonal in the (–,+) quadrant of the spectrum in Figure 2A are identified as double quantum correlations of 14N atoms, NA and NB, that are strongly coupled to the unpaired electron spin of A0·–. The separation between the cross-peaks is nearly four times the corresponding 14N Zeeman frequency (νN of 1.07 MHz). In contrast, there is a single pair of cross-peaks at a frequency of (2.85, 3.7) MHz in the (+,+) quadrant of the spectrum in Figure 2B that is symmetric about the diagonal. These cross-peaks arise from the weakly hyperfine-coupled 14N atom(s), NC. In each case, the small shift along the diagonal from multiples of the 14N Zeeman frequency, νN, is due to the nuclear quadrupolar interaction of the nitrogen atoms.

Figure 2.

Figure 2

Experimental and simulated 2D 14N HYSCORE spectra of the A0·– state

(A and C) Experimental and (B and D) simulated 2D 14N HYSCORE spectra of the A0·– state of PS I. The cross-peaks in the (–,+) and (+,+) quadrant that are centered at multiples of the 14N Zeeman frequency (νN of 1.07 MHz) arise from the hyperfine interactions of the unpaired electron spin with three nitrogen-14 atoms (nuclear spin, I = 1), NA – NC. Figures S1, S2, and S4 display the field-sweep electron-spin-echo spectrum, complete HYSCORE spectrum and a detailed simulation of the (–,+) quadrant of the HYSCORE spectrum of the A0·– state, respectively. The complete set of simulation parameters for the spectra in part (A) and (B) and the calculated hyperfine and quadrupolar couplings are presented in Tables S1 and S4, respectively.

The cross-peaks for NA – NC facilitate the determination of the quadrupolar and hyperfine parameters through numerical simulations of the experimental spectra (Tables 1 and S1). Each set of cross-peaks is uniquely defined by the isotropic hyperfine coupling constant or Fermi contact term, Aiso, anisotropic dipolar hyperfine interaction, T, quadrupolar coupling constant, K, and asymmetry parameter, η. The isotropic hyperfine coupling constant or Fermi contact term, Aiso, provides a measure of the electron spin density at the nucleus, while T reflects the dipole-dipole interaction between the electron and nuclear spin. Additionally, for 14N (nuclear spin I of 1) the quadrupolar coupling and asymmetry parameter, K and η, are valuable probes of the chemical nature and electric field gradient of a quadrupolar nucleus. The quadrupolar coupling constant, K, for all of the hyperfine-coupled nitrogen atoms (NA – NC) is identical, indicating a similar chemical bonding environment (Tables 1 and S1). Moreover, the K value of 0.69 ± 0.02 MHz reveals that NA – NC are pyrrole nitrogen atoms of a Chl ring (Deligiannakis et al., 2000; García-Rubio et al., 2003; van Doorslaer, 2017; Charles et al., 2020). Interestingly, the value of the isotropic hyperfine constant, Aiso, for NA – NC ranges from 2.8–0.52 MHz, indicating that the spin density of the unpaired electron is distributed on at least three of the constituent nitrogen atoms in the A0·– state, albeit to different extents. Additionally, the Aiso of NA – NC for A0·– is smaller than the hyperfine couplings that were previously observed for the nitrogen atoms of both the monomeric Chl a·+ cation (Deligiannakis and Rutherford, 2001) and Chl a· anion (Hoff et al., 1982) in vitro.

Table 1.

Experimentally determined hyperfine and quadrupolar parameters of14N atoms in the A0·– state

Nitrogen Aiso [MHz] K [MHz] ƞ Ring nitrogen Aiso [MHz]a Aiso [MHz]b Aiso [MHz]c
NA 2.8 ± 0.2 0.69 ± 0.02 0.45 ± 0.2 N1 (Chl3A) 5.50 5.2 5.16
NB 2.73 ± 0.2 0.69 ± 0.02 0.45 ± 0.2 N2 (Chl3A) 4.44 8.7 6.42
NC 0.52 ± 0.2 0.69 ± 0.02 0.45 ± 0.2 N2-3 (Chl2A) (−)2.03 1.7 <2
ND (−)0.55 0.8 <2

Experimental 14N isotropic hyperfine (Aiso), quadrupolar (K) and asymmetry (η) parameters of the A0·– state of PS I obtained from the 2D 14N HYSCORE spectroscopy measurements. Also shown are comparisons of Aiso values obtained from isolated monomeric Chl a·– and Pheo·– anions in vitro, as well as Pheo·– measured in PS II. The complete set of simulation parameters for the experimental spectra and calculated hyperfine and quadrupolar couplings are presented in Tables S1 and S4, respectively.

a

Monomeric Chl a·– anion in vitro(Hoff et al., 1982).

b

ESEEM measurements of Pheo·– in PS II(Deligiannakis and Rutherford, 1997).

c

14N-ENDOR measurements of monomeric Pheo·–in vitro(Lendzian et al., 1982).

Electron-nuclear hyperfine parameters of the A0· state are similar to the P700·+ state

For a direct comparison with a paramagnetic Chl species in vivo, we measured the CW (Figure S6) and 2D 14N HYSCORE spectrum of the primary donor cation, P700·+, of PS I (Figures S7 and S8). We observed multiple cross-peaks in the (–,+) and (+,+) quadrant of the spectrum that arise from strongly and weakly hyperfine-coupled 14N atoms, respectively. The hyperfine and quadrupolar coupling parameters from numerical simulations (Figures S7B, S7C, S8B, and S8C) of the experimental spectra (Figure S7A and S8A) are presented in Table S2. There are at least four distinct 14N atoms, NA – ND, (and likely more than four nitrogen atoms if we consider the possibility of overlapping cross-peaks in the (+,+) quadrant that arise from weak hyperfine couplings of 0.4–0.5 MHz) that are interacting with the unpaired electron spin (S = ½) on the primary donor cation, P700·+. Similar to the A0·– state, the K value of 0.69 ± 0.2 MHz indicates that NA – ND are pyrrole nitrogen atoms in P700·+ (Deligiannakis et al., 2000; García-Rubio et al., 2003; van Doorslaer, 2017; Charles et al., 2020). The most pertinent parameter, Aiso, ranges from 1.4–2.8 MHz, indicating that the electron spin is distributed on at least four nitrogen atoms in P700·+ (Deligiannakis and Rutherford, 2001). Once again, the Aiso values of P700·+ are slightly smaller than a monomeric Chl a·+ (Deligiannakis and Rutherford, 2001). This is consistent with previous findings that the primary donor, P700, comprised a strongly electronically coupled Chl a dimer (Norris et al., 1971; Webber and Lubitz, 2001), where the unpaired electron spin density distribution is asymmetric over the two Chl a molecules (3:1 or 4:1 ratio). The similar magnitude of the Aiso values of A0·– and P700·+ indicates that the unpaired electron spin density is also delocalized over two or more Chl a rings in the A0·– state of PS I. Additionally, while the Aiso values range from 1.4–2.8 MHz for the nitrogen atoms of P700·+, the corresponding Aiso values of 0.52–2.80 MHz in the A0·– state suggests a difference in the electron spin density distribution on the pyrrole nitrogen atoms, likely due to a change in the relative distance and orientation of the Chl molecules comprising A0. While electron spin delocalization is apparent in the A0·– state, the 2D 14N HYSCORE measurements do not identify the Chl molecules involved.

DFT confirms electron spin delocalization in the A0·– state

We performed DFT calculations to further understand the electron spin delocalization in the A0·– state. Typically, quantum mechanics/molecular modeling (QM/MM) methods have been used for computational studies of cofactors in redox proteins that include intensive quantum mechanical calculations on a small group of selected atoms, while classical treatment is used for the rest of the system (Kamerlin et al., 2009; Sproviero et al., 2006). The strong electronic coupling between the Chl rings in the primary donor Chl, Chl1A/1B, and the close proximity of Chl1A/1B, Chl2A/2B, Chl3A, and A1A would likely complicate the use of hybrid methods. However, recent studies by Pantazis et al. (Sirohiwal et al., 2020) on highly coupled Chl and pheophytin molecules in photosystem II (PS II) suggests that future refinement of this system using QM/MM methods is possible. In the present study, we used an all-quantum mechanical approach to model the electronic structure of the cofactors as a whole (Charles et al., 2020). We derived the computational model from the X-ray crystal structure of cyanobacterial PS I (Jordan et al., 2001). Initially, we selected a simple monomeric Chl a·– model, Chl3A·–, as a starting point for the DFT calculations. The electron spin density distribution across the isolated Chl3A/3B·– anion (Figure S9) is in agreement with previously published literature on monomeric Chl a·– in vitro (Parusel and Grimme, 2000; O'Malley, 2000a; Huber, 2001; Sinnecker et al., 2000). Interestingly, while the electron spin density has previously been observed to be uniformly distributed across the pyrrole nitrogen atoms in a Chl a·+ cation (O'Malley, 2000a; Sinnecker et al., 2002, O’Malley and Collins, 2001), we (and others) have observed that the distribution is asymmetric in a Chl a·– anion (O'Malley, 2000a, O'Malley, 2000b; Sinnecker et al., 2002).

To determine the effects of protein-cofactor interactions, we systematically expanded the computational model to include the Met668PsaA ligand of Chl3A, a neighboring hydrogen-bonded residue, Tyr696PsaA, the A1A phylloquinone, the ‘accessory’ Chl, Chl2A/2B, and the primary donor Chls, Chl1A/1B. We discovered that, with the exception of A1A, each additional cofactor that was included in the expanded computational models had a significant impact on the electron density distribution. When compared to the Chl a·– anion, including the axial ligand, Met668PsaA, and H-bonded residue, Tyr696PsaA of Chl3A·– saw a shift in the electron spin density distribution across the pyrrole nitrogens, whereby three nitrogens had a slight increase in electron density at the expense of the fourth nitrogen atom (Figure S10). Upon including the ‘accessory Chl’, Chl2A, with its axial water ligand, and the A1A cofactor (Figure 3A) or by considering a ‘full chain model’ with A1A, Chl3A, Chl2A/2B, primary donor Chls, Chl1A/1B and their respective ligands (Figure S11), we observed asymmetric delocalization of electron spin density across both Chl3A and Chl2A in the singly occupied molecular orbital (SOMO) of the A0·– state. Interestingly, the only cofactor that did not demonstrate a discernable impact on charge delocalization in the A0·– state was the phylloquinone acceptor, A1A. To facilitate a direct comparison of the spin delocalization, we also performed DFT calculations on the P700·+ state of PS I using a computational model that contained Chl1A and Chl1B and their respective axial ligands (Jordan et al., 2001). Similar to previous studies, we observed that there is a distribution of the electron spin density on the Chl1A/1B molecules in the SOMO of the primary donor cation, P700·+ (Figure 3B) (Plato et al., 2003), which is similar to the distribution in the A0·– state (Figure 3A) albeit to different extents. This indicates that both the primary donor and acceptor, P700·+ and A0·–, comprised a dimeric Chl motif with extensive electron spin delocalization. Comparison of Figures 3A and 3B indicates that the electron spin density is uniformly distributed across the pyrrole nitrogen atoms on each Chl ring in P700·+ while it is asymmetric in the A0·– state. This is similar to earlier observations on Chl a·+ and Chl a·– in vitro.

Figure 3.

Figure 3

Electron spin density distribution in the singly occupied molecular orbital (SOMO) of the A0·– and P700·+ state

Electron spin density distribution in the singly occupied molecular orbital (SOMO) of the (A) A0·– state in a computational model containing the A1A, Chl3A and Chl2A cofactors as well as the Met688PsaA and Tyr696PsaA residues that are axially ligated and hydrogen bonded to Chl3A cofactor, respectively, and the axial water ligand of the Chl2A cofactor and (B) the P700·+ state in a computational model containing the Chl1A and Chl1B cofactors and their respective axial ligands. The positive and negative spin densities are rendered in purple and yellow, respectively. The calculated hyperfine coupling constants and relative electron spin populations that are obtained from the DFT calculations of the computational model of the A0·– state (part (A)) are shown in Figures S12 and S13, respectively. For comparison, the electron-spin density distribution in the SOMO of an isolated Chl3A· anion, Chl3A· anion with the Met688PsaA and Tyr696PsaA residues and “full-chain” model with the A1A, Chl3A, Chl2A/2B, Chl1A/1B cofactors and the Met688PsaA and Tyr696PsaA residues is shown in Figures S9–S11, respectively.

The electron spin density distribution in the SOMO that was observed in the DFT calculations (Figure 3A) corroborates the delocalization of the unpaired electron spin on at least three nitrogen atoms that were observed in the 2D 14N HYSCORE spectroscopy measurements of the A0·– state (Tables 1 and S1). Moreover, a comparison of the experimental 14N hyperfine coupling constants with the couplings that were calculated by DFT methods indicates that the hyperfine-coupled NA and NB nitrogen atoms in the 2D HYSCORE measurements correspond to the nitrogen atoms at the N1 and N2 of Chl3A and NC matches the N2 and N3 atoms of Chl2A in the A0·– state (Figure S12 and Table S4). The DFT calculations predict hyperfine couplings of 1.3–1.4 MHz with the N3 and N1 atom on Chl3A and Chl2A, respectively. HYSCORE simulations using the calculated hyperfine and quadrupolar parameters that were obtained for the N3 and N1 atom on Chl3A and Chl2A suggest that the experimental cross-peaks should appear at frequencies close to those that arise from NC. However, hyperfine signals from these nitrogen atoms were not detected in the 2D 14N HYSCORE experiments. It is possible that the cross-peaks with hyperfine couplings of 1.3–1.4 MHz are overlapping and may be diminished in intensity due to cross-suppression effects from the hyperfine coupled N1 and N2 atoms of Chl3A (Stoll et al., 2005; Taguchi et al., 2017). The resolution of the hyperfine signals of the N3 and N1 atoms on Chl3A and Chl2A will be the focus of future studies. In principle, it should also be possible to resolve these nitrogens by using PS I isotopically labeled with 15N, as was done with the primary donor in Rhodobacter sphaeroides (Lendzian et al., 1992, 1993). This would be a useful experiment to perform with the understanding, however, that such experiments would come at the loss of information on the quadrupolar coupling (Jarvis et al., 2019). The quantitative prediction of magnetic parameters using DFT methods that are described in this section neglect long-range interactions, such as those of the nearby Chl of the primary donor, P700, which may influence the magnitude of the coupling constants. However, the agreement between the calculated and experimentally measured hyperfine and quadrupolar coupling constants (Figure S12 and Table S4) suggests that such effects may be limited in this system.

Discussion

The most revealing features of the 2D HYSCORE spectroscopic measurements of the A0·– state are the isotropic hyperfine coupling constants, Aiso, that are related to the distribution of electron spin density on the constituent magnetically-active nitrogen atoms (Pandey and Chandra, 1979). The Aiso values of 0.52–2.80 MHz for NA – NC indicate that the electron spin density is distributed on at least three distinct nitrogen atoms of A0·–. Moreover, these Aiso values are less than those of a monomeric Chl a·+ cation (3.18–0.75) (Deligiannakis and Rutherford, 2001) or Chl a·– anion (5.50–0.55) (Hoff et al., 1982) in vitro. The Aiso values are, however, similar in magnitude to the 14N and 15N hyperfine coupling constants (when scaled by the respective gyromagnetic ratios) for the strongly electronically coupled primary donor cation, P700·+ and P865·+, of PS I (Table S2) and the bacterial reaction center (bRC), respectively (Mac et al., 1998; Rigby et al., 2003; Webber and Lubitz, 2001; Käβ et al., 1995; Käβ and Lubitz, 1996; Kaess et al., 1995; Deligiannakis and Rutherford, 2001). These observations provide direct evidence that the unpaired electron spin in the A0·– state is delocalized over two (or more) Chl a molecules in PS I. Additionally, DFT calculations on an extended cofactor model in the A-branch of PS I unambiguously demonstrate that the electron spin density in the A0·– state is delocalized on the Chl2A and Chl3A molecules.

An important factor to consider in such a dimeric system is the extent of delocalization across the two cofactors. The Mulliken atomic spin populations on the nitrogen atoms, N1 – N4, of Chl3A and Chl2A that were obtained from the DFT calculations of the A0·– state allow us to estimate the delocalization of electron spin density across the two Chl rings (Figure S13). Interestingly, the distribution of the electron spin density over the dimer of Chl is asymmetric, where the electron spin delocalization on Chl3A is favored by a factor of ∼3.0. This is similar to other known dimeric Chl pairs, such as P700, where previous DFT studies have suggested that while the SOMO of bare P700·+ sees virtually identical spin density distribution across Chl1A/1B (Sun et al., 2006), the P700·+ state within the native protein environment shows an asymmetry of the spin density distribution that arises from a hydrogen bond to one of the Chls in the dimer.

It is important to consider the possibility that the chemical treatment used to remove the phylloquinones (and antenna Chl) in wild-type PS I (see Method details section) could influence the primary charge separation reaction. However, inhibition of charge separation by such treatments is unlikely as previous studies using ether- (Ikegami et al., 1987) and ether/acetaldehyde-treated (Ikegami et al., 1995) PS I have demonstrated successful formation of the primary donor cation, P700·+, upon illumination. Moreover, we observed that the loss of antenna Chls does not have an impact on the charge-separation reaction as the ΔmenB variant of PS I with a mild detergent treatment (see STAR Methods) yields virtually identical spectral features in the A0·– state (Figure S5). The removal of the A1 cofactor by chemical treatment and/or interruption of phylloquinone biosynthesis could result in small perturbations of the protein environment. However, a recent study where forward ET was halted without altering the A1 binding pocket also yielded identical hyperfine interactions in the A0·– state and the spin density distribution of the A0·– state in the current DFT calculations is not impacted by the presence or absence of the phylloquinone cofactor, A1A (Gorka et al., 2021). Therefore, while there are inherent uncertainties associated with the chemical or genetic modification of PS I, it is clear that changes to the phylloquinone binding pocket do not impact the electronic properties of the A0·– state.

Comparison of the relative orientation of Chl3A and Chl2A in the X-ray crystal structure of PS I (Jordan et al., 2001) reveals a close resemblance with the Chl1A/1B dimer in the primary donor, P700. As can be seen in Figures 4A–4D, the ring planes of the Chl molecules in both cases are parallel to each other in a π-stacked orientation. The distance between the Chl3A and Chl2A ring in the direction perpendicular to the ring plane along the z axis is 3.9 Å (Figures 4A and 4B), which is nearly identical to the Chl1A/1B dimer of P700 (Figures 4C and 4D). The only substantive difference is the relative orientation of Chl2A and Chl3A, as there is a slight rotation in the ring plane which manifests as a change in the identity of the nearest pyrrole nitrogen atom in the neighboring ring. The N3 nitrogen from Chl1A in P700 is proximal to the N3 nitrogen of Chl1B (3.6 Å distance); however, the N3 nitrogen of Chl3A is nearly equidistant to the N3 and N2 nitrogens of Chl2A (5.8 and 5.7 Å, respectively). This difference in the relative orientation of Chl2A and Chl3A in A0 is best described as a “sliding” of the two Chl a rings away from each other along the X axis by ∼3 Å. This suggests that the coupling between the Chl3A and Chl2A rings of A0A should be weaker than the coupling of Chl1A/1B in the primary donor, P700.

It is important to note that while the DFT calculations were only performed on A0·– in the A-branch, the distance and relative orientation of the analogous Chl cofactors of the B-branch, Chl2B and Chl3B, are similar. The distance between the nearest ring nitrogen atoms is 5.7 Å and 5.4 Å for the A- and B-branch, respectively, and both display a parallel orientation. There is a minor difference between the relative orientation of the Chl2A/Chl3A and Chl2B/Chl3B dimers, as the rotation in the X-Y plane is less extensive (relative to P700) on the B-branch, resulting in the identity of the nearest nitrogen atom that more closely resembles P700, i.e. the N3 nitrogen of Chl2B is close to the N3 nitrogen of Chl3B. Despite the subtle differences in the placement of Chl2B and Chl3B, we expect that the delocalization of the electron spin is similar in the A0B·– state. It is currently unknown how much, if any, of the B-side is visible in our experiments and we plan on addressing the nature of the B-side ET in future studies.

The significant electron spin delocalization over the Chl2A/2B and Chl3A/3B cofactors in the A0·– state provides unique insights into the previously unknown role of the “accessory chlorophylls”, Chl2A/2B, of PS I. Our results clearly indicate that it is important to consider Chl2A/2B and Chl3A/3B in the A0A/0B·– state in the same light as the electronically coupled Chl1A/1B dimer in the primary donor cation, P700·+ (Madjet et al., 2009). The presence of an electronically coupled dimeric Chl motif has a pronounced influence on the nature of the electron donor and acceptors of PS I. ET between a given donor (D) and acceptor (A) in a RC was previously considered a discrete event where the kinetics are governed by factors such as the inter-cofactor distance, free energy difference, protein motion, and dielectric environment (Moser et al., 1992; Moia et al., 2020). However, based on our findings and those of others (Cherepanov et al., 2021) the initial light-driven charge separation in type I RCs may not be a discrete event that can be described by a simple D-A system. Instead, there is extended delocalization of electron density across a set of multiple coupled cofactors, or an exciplex, that is driven by downhill energetics. This provides a spatial separation of charges that occurs faster than de-excitation of the photo-excited primary donor, P700, alone. Previous calculations have suggested a process whereby the coupling of Chl2A/2B and Chl3A/3B may be the key to lowering the redox potential of the A0 electron acceptor so as to ensure that the formation of the charge-separated state is energetically favorable (Ptushenko et al., 2008). However, experimental evidence for such a coupling was lacking to date. It is important to note that the A1A cofactor (located 6.7 Å from Chl3A) does not have a discernable effect on the electron spin density distribution over the Chl2A and Chl3A rings individually or collectively, suggesting that the ET reactions after primary charge separation occur in a discrete fashion.

Efficient charge separation is not the sole requirement for attaining high quantum efficiency in RCs. To compete effectively with charge recombination, charge separation must be accompanied by rapid multi-step forward ET through one or more electron acceptors. The spatial separation afforded by the dimeric Chl motif of A0·–, as opposed to a Chl monomer located closer to P700, likely prevents ultrafast charge recombination with P700·+ thus allowing for forward ET to the A1 cofactor (Figure 1B). This is an important principle that could be incorporated in the design of artificial photosynthetic RCs that currently suffer from relatively low quantum yields and rapid charge recombination reactions (Llansola-Portoles et al., 2017). The incorporation of highly coupled donors and acceptors would allow for extensive delocalization that could prolong the lifetime of the charge-separated state with high quantum efficiency over a wide range of excitation wavelengths.

Photosynthetic RCs in a wide variety of organisms employ two symmetric branches of cofactors for light-driven charge separation and ET (Lancaster et al., 2007; Roy et al., 1995; Jordan et al., 2001; Umena et al., 2011; Gisriel et al., 2017) where, in addition to a primary donor (B)Chl, each branch also contains an accessory (B)Chl molecule. In light of our evidence for direct charge delocalization in the A0·– state of PS I, it is worth examining the nature and evolution of primary charge separation in other photosynthetic RCs. The homodimeric heliobacterial reaction center (HbRC) is considered to be the exemplar of ancestral PS I due to its structural and functional simplicity (Sattley and Swingley, 2013; Cardona, 2015; Ferlez et al., 2016; Orf et al., 2018). The ET pathway of the HbRC only contains two electron acceptors, namely, A0 and a single [4Fe–4S] cluster, FX (Ferlez et al., 2016; Gisriel et al., 2017). Similar to PS I, the core of the HbRC contains an “accessory” Chl molecule, BChl2A, which is positioned between BChl1A and Chl3A (where “B” refers to “bacterio”). It is interesting that the 81-OH Chl a and BChl g molecules of the HbRC are nearly identical in orientation to the analogous cofactors in PS I (Figures 5A and 5B), where the center-to-center distance between the two rings is 3.4 Å, but with a longer distance of ∼6.3 Å to the nearest nitrogen atom. This is in stark contrast to type II RCs (the bRC and PS II), where the relative orientation of the Chl2A and Chl3A pair deviates significantly from a parallel orientation (Figures 5C and 5D) and the chemical identity of the Chl3A acceptor is a (B)pheophytin ((B)Pheo) of the (B)Chl of the primary donor, while all known type I RCs contain a Chl a (or close derivative) in this site (Orf et al., 2018; Chen et al., 2020).

Figure 5.

Figure 5

Structural comparison of primary acceptor motifs in other type I and type II RCs

Relative orientation and distances of the (B)Chl3A and (B)Chl2A cofactors in the primary acceptor positions for the (A and B) heliobacterial reaction center (Gisriel et al., 2017) and (C and D) purple bacterial reaction center (Lancaster et al., 2007).

The differences in type I and type II RCs in terms of their primary acceptors likely reflects different strategies to control primary charge separation. It is important that charge separation in the RC take place rapidly and prior to de-excitation of the primary donor, which would require sufficiently downhill energy transfer. In the case of Type II RCs, a driving force of approximately 300 mV is gained by introducing (B)Pheo instead of (B)Chl as the primary acceptor (Fajer et al., 1975). Recent studies on PS II suggest that this gap is widened by protein-matrix effects, which redshifts the primary donor (suggested to be Chl2A) and blue shifts the primary acceptor (Sirohiwal et al., 2020). This is reflected in the findings of both pulsed EPR measurements and theoretical calculations on the reduced (B)Pheo·− state. The Aiso values for both monomeric Pheo·– in vitro (Lendzian et al., 1982) and photoaccumulated Pheo·– in PS II (Deligiannakis and Rutherford, 1997) that were determined through 14N ENDOR and ESEEM, respectively, are shown in Table 1. The Aiso values of for both Pheo·– (O'Malley, 2000b) and (B)Pheo·– (O’Malley, 1999) in conjunction with large scale calculations on the early acceptors in PS II (Sirohiwal et al., 2020) provide unambiguous evidence that the electron spin density on the primary acceptor (A0) of Type II RCs is localized entirely on the (B)Pheo cofactor. This is noteworthy as it indicates that: (1) electron delocalization across the primary acceptor with a dimeric Chl motif is not required for highly efficient energy transfer in these RCs, and (2) both type I and type II RCs utilize extensive protein-cofactor interactions (Srinivasan and Golbeck, 2009; Sirohiwal et al., 2020) to tune the energetic and electronic properties of its cofactors, albeit to different ends.

In contrast, type I RCs do not have the luxury of an inherent loss of free energy, as they are fixed with a Chl a as the primary acceptor, regardless of the identity of the primary donor. The reason for fixing Chl a as the acceptor is not fully understood; but regardless, Type I RCs must deal with the energetic restrictions of that Chl. To cope with this restriction, type I RCs evolved different methods to tune the energetics of their redox cofactors so as to ensure that both charge separation is favored and subsequent charge recombination is inhibited in varying landscapes of possible acceptors. For example, one major difference between the HbRC and PS I is the addition of a bound quinone (A1) in the acceptor chain in PS I, which is thought to prevent triplet state formation when recombining from Chl3A·–, albeit at the expense of marginally slower forward ET (reviewed in (Orf et al., 2018)). In PS I, A0 and A1 are poised at redox potentials sufficient to allow forward ET from A0·– to A1 and then from A1·– to FX, all while preventing recombination from these cofactors, not a trivial task when the identity of the Chl3 is fixed. However, not only has Chl dimerization been shown to have an impact on the redox potential (Ptushenko et al., 2008), but the coupling between monomers is tunable by altering the distance between the π-planes (Madjet et al., 2009). Thus, the employment of a dimeric motif to control the nature and extent of delocalization became a tool that nature used to precisely tune the energetics of the ET cofactors, and in this case, the efficiency of charge separation itself.

Limitations of the study

There are limitations of this study that should be considered when extending the model to other systems. For application of our model to other type I and type II RCs, it is important to consider that there can be significant variations in the electronic structure and protein-matrix influences of a certain cofactor in the same protein but from different species. Therefore, specific RCs would need tested before we can have a high confidence in the broad applicability of our model.

STAR★Methods

Key resources table

REAGENT or RESOURCE SOURCE IDENTIFIER
Bacterial and virus strains

Wild-type Synechocystis sp. PCC 6803 Lab of John Golbeck N/A
ΔmenB Synechocystis sp. PCC 6803 Lab of John Golbeck N/A

Biological samples

Wild-type Synechocystis sp. PCC 6803 Photosystem I Lab of John Golbeck N/A
ΔmenB Synechocystis sp. PCC 6803 Photosystem I Lab of John Golbeck N/A

Chemicals, peptides, and recombinant proteins

n-dodecyl β-d-maltopyranoside (β-DDM) Anatrace D310; CAS: 69227-93-6
Tris VWR 0497; CAS: 77-86-1
Glycerol VWR BDH1172; CAS: 56-81-5
Diethyl ether Millipore EX0190; CAS: 60-29-7
Ethanol VWR V1005; CAS: 64-17-5
Sucrose EMD SX1075; CAS: 57-50-1
Sodium Chloride EMD SX0420
Spectinomycin Dot scientific DSS23000; CAS 22189-32-9
Potassium Ferricyanide Mallinckrodt 6913; 13746-66-2
Sodium Carbonate VWR 0585; CAS: 497-19-8
Ferric Ammonium Citrate MP 158040; CAS: 1185-57-5
Citric Acid Aldrich 127; CAS: 77-92-9
Calcium Chloride Research organics 1005C; CAS: 10035-04-8
Magnesium Sulfate Heptahydrate Amresco 0662; CAS: 10034-99-8
Potassium Phosphate Dibasic J.T. Baker 3246; CAS: 7778-77-0
Sodium Nitrate J.T. Baker 3770; CAS: 7631-99-4
HEPES, Free Acid J.T. Baker 4018; CAS: 7365-45-9
Triton X-100 Sigma T9284; CAS: 9002-93-1
Sodium dithionite Sigma-Aldrich 157953; CAS: 7775-14-6

Software and algorithms

EasySpin 5.2 https://www.easyspin.org N/A
MatLab Mathworks.com N/A
ORCA 4.2 orcaforum.kofo.mpg.de N/A
Visual Molecular Dynamics (VMD) 1.9 NIH CENTER FOR MACROMOLECULAR MODELING & BIOINFORMATICS UNIVERSITY OF ILLINOIS AT URBANA-CHAMPAIGN N/A

Resource availability

Lead contact

Further information and requests for resources should be directed to and will be fulfilled by the Lead Contact, K. V. Lakshmi (lakshk@rpi.edu).

Materials availability

This study did not generate unique reagents. Requests for samples that were analyzed in this study can be directed to the Lead Contact, K. V. Lakshmi (lakshk@rpi.edu).

Data availability

Numerically simulated parameters in this study are available by contacting the lead contact, K. V. Lakshmi (lakshk@rpi.edu).

Method details

Preparation of photosystem I samples

PS I complexes were isolated and purified from wild-type (WT) and ΔmenB Synechocystis sp. PCC 6803 similar to previously published procedures (Rögner et al., 1990; Johnson et al., 2000). Cells were grown at 30°C in β-HEPES medium under continuous illumination while bubbling 1% CO2. The cells were harvested by centrifugation and lysed in a French pressure cell. The unbroken cells and debris were removed by centrifugation at 5,000 x g for 10 min at 4°C. Subsequently, the thylakoid membranes were pelleted by ultracentrifugation at 50,000 x g for 45 min at 4°C. Membranes were resuspended in 50 mM Tris 50 mM Tris.HCl buffer 10% glycerol at a pH of 8.3 and stored at –80°C until use. The thylakoid membranes were solubilized by adding 1% n-dodecyl β-d-maltopyranoside (β-DDM) in the dark at 4°C and the PS I complexes were isolated by a sucrose density gradient ultracentrifugation, using a gradient of 5% – 20% sucrose. The gradient was spun at rmax = 143K x g for 16 hours at 4°C. Trimers were collected from the first gradient from the dark green band at ∼15% sucrose, after which the detergent and sucrose were removed using ultrafiltration over a 100 kDa membrane. PS I particles were pelleted in a second gradient of 5% – 20% sucrose that was spun at rmax = 175K x g for 16 hours at 4°C. The PS I was resuspended in 50 mM Tris.HCl buffer containing 0.05% β-DDM and 10% glycerol at a pH of 8.3. The samples were stored at –80°C until use.

Trapping of the A0A· state

Cryogenic photo-accumulation of the A0A·– state requires inhibition of forward electron transfer to A1. This was accomplished by removal of the phylloquinone cofactor occupying the A1 site followed by chemical reduction of the [4Fe–4S] clusters of wild-type PS I. The phylloquinone in the A1 site was removed by the treatment of PS I with water-saturated diethyl ether as described previously (Ikegami and Kato, 1975; Itoh et al., 1987, 2001; Ikegami et al., 2000). PS I particles were washed over a 100 kDa membrane to remove detergent and subsequently lyophilized. The lyophilized protein was treated with water saturated diethyl ether three times at 4°C to remove the pigments, spinning down the protein after each treatment. The protein was then resuspended in in 50 mM Tris.HCl buffer containing 0.05% β-DDM and 10% glycerol. The reduction of the [4Fe–4S] clusters was accomplished by the addition of 50 mM sodium dithionite under anaerobic conditions (Lakshmi et al., 1999). The ΔmenB genetic variant of PS I was treated with 0.5% Triton X-100 to remove the plastoquinone in the A1 site (Johnson et al., 2001) and the A0·– state of this sample yielded virtually identical spectral features to the water-saturated diethyl ether treated wild-type PS I (Figure S5). However, the data presented in this paper was obtained from water-saturated diethyl ether treated wild-type PS I as the ΔmenB PS I sample treated with 0.5% Triton X-100 resulted in spectra with a lower signal-to-noise ratio due lower protein concentration. Prior to low-temperature illumination, the respective PS I samples were cooled in a dewar containing a dry ice/ethanol mixture at 220 K and after equilibration, the sample was illuminated for 1 min with 100 W of white light. The sample was rapidly frozen in liquid nitrogen in the dark after the illumination. The P700·+ state was generated by chemical oxidation of wild-type PS I by the addition of 0.05 mM potassium ferricyanide (Lakshmi et al., 1999). The inherent ratio of A:B:2:1 for ET along the A- and B-branches, combined with experimental studies showcasing the difficulty in reducing A1B·– compared to A1A·– (Poluektov et al., 2005), suggests that A0A·– is the primary species being generated in this study. Therefore, contributions by A0B·– are expected to be minimal.

Pulsed electron paramagnetic resonance spectroscopy

The EPR spectra were obtained on a custom-built continuous-wave (CW)/pulsed X-band Bruker ELEXSYS 580 EPR spectrometer using a dielectric flex-line ER 4118-MD5 probe (Bruker BioSpin, Billerica, MA) and a dynamic continuous-flow cryostat CF935 (Oxford Instruments, Oxfordshire, U.K.).

The 2D 14N and 1H HYSCORE spectra of the A0·– and P700·+ state of PS I were acquired at 80 K and 50 K, respectively, at a magnetic field position of 345.8 mT and 346.1 mT in the EPR spectrum (Figures 2 and S2, S3, S5, S7, and S8, respectively) which results in higher signal to noise of the 2D 14N HYSCORE spectrum. For the 4-pulse HYSCORE sequence, the echo amplitude was measured using the pulse sequence (π/2–τπ/2–t1πt2π/2–τ–echo) with a τ value of 136 ns. The pulse length was 8 ns and 16 ns for the π/2- and π-pulse, respectively. The delays in the pulse sequence are defined as the difference in the starting point of the pulses. The echo intensity was measured as a function of t1 and t2, where t1 and t2 were incremented in steps of 16 ns from an initial value of 40 ns and 32 ns, respectively. A total of 384 steps were used for each dimension. The 8 ns time difference between the initial value of t1 and t2 was set to account for the difference in length between the π/2- and π-pulse. This provided symmetric spectra in both dimensions. The unwanted echoes were eliminated by applying a 16-step phase cycling procedure.

The use of a 4-pulse HYSCORE sequence often results in the suppression of cross-peaks due to deep signal modulations from strongly coupled 14N nuclei and the presence of unwanted inter-nuclear combination peaks that hinder the observation of some of the hyperfine interactions in both ESEEM and 2D HYSCORE experiments. Therefore, a 6-pulse HYSCORE sequence was also used for the acquisition of the 2D 14N HYSCORE spectra of the P700·+ state of PS I. For the 6-pulse HYSCORE sequence, the echo amplitude was measured with the pulse sequence (π/2-τ1-π-τ1-π/2-t1-π-t2-π/2-τ2-π-τ2-echo). The inter-pulse delays were defined as the difference between the starting points of the pulses. Equal amplitude pulses of 8 ns for π/2 and 16 ns for π were used to record a 256 x 256 matrix. The 8 ns time difference between the initial value of the time delays, t1 and t2, and the π/2 and π pulses was used to account for the difference in length between the π/2 and π-pulses to obtain symmetric spectra. The echo intensity was measured as a function of t1 and t2, where t1 and t2 were incremented in steps of 24 ns. The blind spot pattern in a 6-pulse 2D HYSCORE spectrum is determined by the τ1 and τ2 delays. Hence, the τ1 and τ2 delay was set to 32 ns and 100 ns, respectively, to avoid blind spots and resonator ring down. The application of an 8-step phase cycling procedure was used to eliminate the unwanted echoes. The use of a 6-pulse HYSCORE sequence led to the observation of additional cross-peaks that were not detected in the 4-pulse HYSCORE experiment (see supplemental information for experimental and simulated spectra).

The time domain 4-pulse and 6-pulse 2D HYSCORE data was processed using MATLAB R2018b. A third order polynomial baseline was subtracted from the resulting time-domain spectra. The corrected spectra were zero-filled to obtain a [2048 x 2048] matrix and Fourier transformed using a Fast Fourier Transformation (FFT) algorithm. The frequency domain spectra were plotted as the amplitude (absolute value) of the 2D frequency components. Numerical simulations of the experimental 2D HYSCORE spectra using the hyperfine parameters obtained from the spectra analysis were performed using the “saffron” function of the EasySpin software package (Stoll and Schweiger, 2006).

DFT calculations

The DFT calculations of the primary acceptor anion, A0·–, and primary donor cation, P700·+, were performed with a B3LYP functional and EPR-II basis set using the software program, ORCA 4.0 (Neese, 2012, 2018). The coordinates for the atoms in each computational model were derived from the X-ray coordinates of Photosystem I from Thermosynechococcus elongatus (Jordan et al., 2001). The initial computational model for the A0·– state was comprised of the Chl3A·– anion radical with the constituent atoms of the Chl3A cofactor. The model was systematically expanded to include the axial Met ligand, Met688PsaA, that is coordinated to the central Mg2+ ion of the Chl3A cofactor and the hydrogen-bonded Tyr residue, Tyr696PsaA, A1A, Chl2A/2B, Chl1A/1B cofactors and their respective axial ligands, namely, water molecules and His residues. To limit the size of the computational model in the calculations, the isoprenoid substituent of the Chl3A, Chl2A/2B and Chl1A/1B cofactors was truncated in all of the calculations to contain 8 carbon atoms. The computational model for P700·+ contained the constituent atoms of the Chl1A/1B cofactors with the respective axial histidine ligand and the isoprenoid substituent of Chl1A/1B cofactor was truncated to contain 5 carbon atoms. All of the geometry optimizations were carried out at the B3LYP level of theory with the DZVP basis set to expand the Kohn-Sham orbitals. For the B3LYP single point calculations on the optimized structures, the EPR-II basis set was used for the lighter atoms in the models, while the 6-31G(d) basis set was incorporated for the Mg2+ ion. The RIJCOSX approximation was employed in conjunction with the CPCM solvent model with a dielectric (ε) of 4.0 to approximate the surrounding protein environment. The electron spin density distribution for each structure was visualized in the molecular modeling and visualization program, Visual Molecular Dynamics (VMD).

Acknowledgments

This study is supported by the Photosynthetic Systems Program, Office of Basic Energy Sciences of the U.S. Department of Energy under the contracts DE-FG02-07ER15903 (KVL) and DE-FG-05-05-ER46222 (JHG). The authors would like to acknowledge the Center for Computational Innovations (CCI) at Rensselaer Polytechnic Institute for computational resources.

Author contributions

M.G., P.C., V.K., and A.B. conducted experiments, analyzed the data, and performed calculations; M.G., K.V.L., and J.H.G. designed the experiments and wrote the paper.

Declaration of interests

The authors declare no competing interests.

Inclusion and diversity

While citing references scientifically relevant for this work, we also actively worked to promote gender balance in our reference list.

Published: July 23, 2021

Footnotes

Supplemental information can be found online at https://doi.org/10.1016/j.isci.2021.102719.

A video abstract is available at https://doi.org/10.1016/j.isci.2021.102719#mmc2.

Contributor Information

K.V. Lakshmi, Email: lakshk@rpi.edu.

John H. Golbeck, Email: jhg5@psu.edu.

Supplemental information

Document S1. Figures S1–S13 and Tables S1–S4
mmc1.pdf (1.1MB, pdf)
Video abstract
Download video file (41.5MB, mp4)

References

  1. Anderson J.M. The role of chlorophyll-protein complexes in the function and structure of chloroplast thylakoids. Mol. Cell Biochem. 1982;46:161–172. doi: 10.1007/BF00239665. [DOI] [PubMed] [Google Scholar]
  2. Ayers P.W. An elementary derivation of the hard/soft-acid/base principle. J. Chem. Phys. 2005;122:141102. doi: 10.1063/1.1897374. [DOI] [PubMed] [Google Scholar]
  3. Brettel K., Leibl W. Electron transfer in photosystem I. Biochim. Biophys. Acta Bioenerg. 2001;1507:100–114. doi: 10.1016/s0005-2728(01)00202-x. [DOI] [PubMed] [Google Scholar]
  4. Britt R.D., Campbell K.A., Peloquin J.M., Gilchrist M.L., Aznar C.P., Dicus M.M., Robblee J., Messinger J. Recent pulsed EPR studies of the photosystem II oxygen-evolving complex: Implications as to water oxidation mechanisms. Biochim. Biophys. Acta Bioenerg. 2004;1655:158–171. doi: 10.1016/j.bbabio.2003.11.009. [DOI] [PubMed] [Google Scholar]
  5. Cardona T. A fresh look at the evolution and diversification of photochemical reaction centers. Photosynth Res. 2015;126:111–134. doi: 10.1007/s11120-014-0065-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Charles P., Kalendra V., He Z., Khatami M.H., Golbeck J.H., van der Est A., Lakshmi K.V., Bryant D.A. Two-dimensional 67Zn HYSCORE spectroscopy reveals that a Zn-bacteriochlorophyll aP’ dimer is the primary donor (P840) in the type-1 reaction centers of Chloracidobacterium thermophilum. Phys. Chem. Chem. Phys. 2020;22:6457–6467. doi: 10.1039/c9cp06556c. [DOI] [PubMed] [Google Scholar]
  7. Chatterjee R., Milikisiyants S., Coates C.S., Lakshmi K.V. High-resolution two-dimensional 1H and 14N hyperfine sublevel correlation spectroscopy of the primary quinone of photosystem II. Biochemistry. 2011;50:491–501. doi: 10.1021/bi101883y. [DOI] [PubMed] [Google Scholar]
  8. Chauvet A., Dashdorj N., Golbeck J.H., Johnson T.W., Savikhin S. Spectral resolution of the primary electron acceptor A0 in photosystem I. J. Phys. Chem. B. 2012;116:3380–3386. doi: 10.1021/jp211246a. [DOI] [PubMed] [Google Scholar]
  9. Chen J.-H., Wu H., Xu C., Liu X.-C., Huang Z., Chang S., Wang W., Han G., Kuang T., Shen J.-R., Zhang X. Architecture of the photosynthetic complex from a green sulfur bacterium. Science. 2020;370:eabb6350. doi: 10.1126/science.abb6350. [DOI] [PubMed] [Google Scholar]
  10. Cherepanov D.A., Shelaev I.V., Gostev F.E., Aybush A.V., Mamedov M.D., Shuvalov V.A., Semenov A.Y., Nadtochenko V.A. Generation of ion-radical chlorophyll states in the light-harvesting antenna and the reaction center of cyanobacterial photosystem I. Photosyn Res. 2020;146:55–73. doi: 10.1007/s11120-020-00731-0. [DOI] [PubMed] [Google Scholar]
  11. Cherepanov D.A., Shelaev I.V., Gostev F.E., Petrova A., Aybush A.V., Nadtochenko V.A., Xu W., Golbeck J.H., Semenov A.Y. Primary charge separation within the structurally symmetric tetrameric Chl2APAPBChl2B chlorophyll exciplex in photosystem I. J. Photochem. Photobiol. B. 2021;217:112154. doi: 10.1016/j.jphotobiol.2021.112154. [DOI] [PubMed] [Google Scholar]
  12. Cohen R.O., Shen G., Golbeck J.H., Xu W., Chitnis P.R., Valieva A.I., van der Est A., Pushkar Y., Stehlik D. Evidence for asymmetric electron transfer in cyanobacterial photosystem I:  Analysis of a methionine-to-leucine mutation of the ligand to the primary electron acceptor A0. Biochemistry. 2004;43:4741–4754. doi: 10.1021/bi035633f. [DOI] [PubMed] [Google Scholar]
  13. Commoner B., Heise J.J., Townsend J. Light-induced paramagnetism in chloroplasts. Proc. Natl. Acad. Sci. U S A. 1956;42:710–718. doi: 10.1073/pnas.42.10.710. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Dashdorj N., Xu W., Cohen R.O., Golbeck J.H., Savikhin S. Asymmetric electron transfer in cyanobacterial photosystem I: Charge separation and secondary electron transfer dynamics of mutations near the primary electron acceptor A0. Biophys. J. 2005;88:1238–1249. doi: 10.1529/biophysj.104.050963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Davis I.H., Heathcote P., Maclachlan D.J., Evans M.C.W. Modulation analysis of the electron spin echo signals of in vivo oxidised primary donor 14N chlorophyll centres in bacterial, P870 and P960, and plant photosystem I, P700, reaction centres. Biochim. Biophys. Acta Bioenerg. 1993;1143:183–189. [Google Scholar]
  16. Deligiannakis Y., Hanley J., Rutherford A.W. Spin-lattice relaxation of the phyllosemiquinone radical of photosystem I. Biochemistry. 1998;37:3329–3336. doi: 10.1021/bi971297p. [DOI] [PubMed] [Google Scholar]
  17. Deligiannakis Y., Hanley J., Rutherford A.W. 1D- and 2D-ESEEM study of the semiquinone radical QA- of photosystem II. J. Amer Chem. Soc. 1999;121:7653–7664. [Google Scholar]
  18. Deligiannakis Y., Louloudi M., Hadjiliadis N. Electron spin echo envelope modulation (ESEEM) spectroscopy as a tool to investigate the coordination environment of metal centers. J. Coord. Chem. 2000;204:1–112. [Google Scholar]
  19. Deligiannakis Y., Rutherford A.W. One- and two-dimensional electron spin echo envelope modulation study of the intermediate electron acceptor pheophytin in 14N- and 15N-labeled photosystem II. J. Amer Chem. Soc. 1997;119:4471–4480. [Google Scholar]
  20. Deligiannakis Y., Rutherford A.W. Electron spin echo envelope modulation spectroscopy in photosystem I. Biochim. Biophys. Acta Bioenerg. 2001;1507:226–246. doi: 10.1016/s0005-2728(01)00201-8. [DOI] [PubMed] [Google Scholar]
  21. Fajer J., Brune D.C., Davis M.S., Forman A., Spaulding L.D. Primary charge separation in bacterial photosynthesis: Oxidized chlorophylls and reduced pheophytin. Proc. Natl. Acad. Sci. U S A. 1975;72:4956–4960. doi: 10.1073/pnas.72.12.4956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Fajer J., Davis M.S., Forman A., Klimov V.V., Dolan E., Ke B. Primary electron acceptors in plant photosynthesis. J. Amer. Chem. Soc. 1980;102:7143–7145. [Google Scholar]
  23. Ferlez B., Cowgill J., Dong W., Gisriel C., Lin S., Flores M., Walters K., Cetnar D., Redding K.E., Golbeck J.H. Thermodynamics of the electron acceptors in Heliobacterium modesticaldum: An exemplar of an early homodimeric Type I photosynthetic reaction center. Biochemistry. 2016;55:2358–2370. doi: 10.1021/acs.biochem.5b01320. [DOI] [PubMed] [Google Scholar]
  24. Forman A., Davis M.S., Fujita I., Hanson L.K., Smith K.M., Fajer J. Mechanisms of energy transduction in plant photosynthesis: ESR, ENDOR and MOs of the primary acceptors. Isr. J. Chem. 1981;21:265–269. [Google Scholar]
  25. Fromme P., Jordan P., Kraus N. Structure of photosystem I. Biochim. Biophys. Acta Bioenerg. 2001;1507:5–31. doi: 10.1016/s0005-2728(01)00195-5. [DOI] [PubMed] [Google Scholar]
  26. García-Rubio I., Martínez J.I., Picorel R., Yruela I., Alonso P.J. HYSCORE spectroscopy in the cytochrome b559 of the photosystem II reaction center. J. Amer. Chem. Soc. 2003;125:15846–15854. doi: 10.1021/ja035364g. [DOI] [PubMed] [Google Scholar]
  27. Gibasiewicz K., Ramesh V.M., Lin S., Redding K., Woodbury N.W., Webber A.N. Excitonic interactions in wild-type and mutant PSI reaction centers. Biophys. J. 2003;85:2547–2559. doi: 10.1016/s0006-3495(03)74677-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Gisriel C., Sarrou I., Ferlez B., Golbeck J.H., Redding K.E., Fromme R. Structure of a symmetric photosynthetic reaction center-photosystem. Science. 2017;357:1021–1025. doi: 10.1126/science.aan5611. [DOI] [PubMed] [Google Scholar]
  29. Golbeck J.H. Springer; 2006. Photosystem I: The Light-Driven Plastocyanin: Ferredoxin Oxidoreductase. [Google Scholar]
  30. Gorka M., Gruszecki E., Charles P., Kalendra V., Lakshmi K.V., Golbeck J.H. Two-dimensional HYSCORE spectroscopy reveals a histidine imidazole as the axial ligand to Chl3A in the M688HPsaA genetic variant of Photosystem I. Biochim. Biophys. Acta Bioenerg. 2021;1862:148424. doi: 10.1016/j.bbabio.2021.148424. [DOI] [PubMed] [Google Scholar]
  31. Harmer J., Mitrikas G., Schweiger A. Advanced pulse EPR methods for the characterization of metalloproteins. In: Berliner L., Hanson G., editors. High Resolution EPR: Applications to Metalloenzymes and Metals in Medicine. Springer; 2009. pp. 13–61. [Google Scholar]
  32. Höfer P., Grupp A., Nebenführ H., Mehring M. Hyperfine sublevel correlation (HYSCORE) spectroscopy: A 2D ESR investigation of the squaric acid radical. Chem. Phys. Lett. 1986;132:279–282. [Google Scholar]
  33. Hoff A.J., Lendzian F., Möbius K., Lubitz W. Proton and nitrogen electron nuclear double and triple resonance of the chlorophyll a anion in liquid solution. Chem. Phys. Lett. 1982;85:3–8. [Google Scholar]
  34. Holzwarth A.R., Müller M.G., Niklas J., Lubitz W. Ultrafast transient absorption studies on photosystem I reaction centers from Chlamydomonas reinhardtii. 2: Mutations near the P700 reaction center chlorophylls provide new insight into the nature of the primary electron donor. Biophys. J. 2006;90:552–565. doi: 10.1529/biophysj.105.059824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Huber M. The role of the electronic structure of the porphyrin as viewed by EPR/ENDOR methods in the efficiency of biomimetic model compounds for photosynthesis. Eur. J. Org. Chem. 2001;2001:4379–4389. [Google Scholar]
  36. Ikegami I., Itoh S., Iwaki M. Photoactive photosystem I particles with a molar ratio of chlorophyll a to P700 of 9. Plant Cell Physiol. 1995;36:857–864. [Google Scholar]
  37. Ikegami I., Itoh S., Iwaki M. Selective extraction of antenna chlorophylls, carotenoids and quinones from photosystem I reaction center. Plant Cell Physiol. 2000;41:1085–1095. doi: 10.1093/pcp/pcd033. [DOI] [PubMed] [Google Scholar]
  38. Ikegami I., Kato S. Enrichment of photosystem I reaction center chlorophyll from spinach chloroplasts. Biochim. Biophys. Acta. 1975;376:588–592. doi: 10.1016/0005-2728(75)90181-4. [DOI] [PubMed] [Google Scholar]
  39. Ikegami I., Sétif P., Mathis P. Absorption studies of photosystem I photochemistry in the absence of vitamin K-1. Biochim. Biophys. Acta Bioenerg. 1987;894:414–422. [Google Scholar]
  40. Ishikita H., Knapp E.W. Redox potential of quinones in both electron transfer branches of photosystem I. J. Biol. Chem. 2003;278:52002–52011. doi: 10.1074/jbc.M306434200. [DOI] [PubMed] [Google Scholar]
  41. Itoh S., Iwaki M., Ikegami I. Extraction of vitamin K-1 from photosystem I particles by treatment with diethyl ether and its effects on the A−1 EPR signal and system I photochemistry. Biochim. Biophys. Acta Bioenerg. 1987;893:508–516. [Google Scholar]
  42. Itoh S., Iwaki M., Ikegami I. Modification of photosystem I reaction center by the extraction and exchange of chlorophylls and quinones. Biochim. Biophys. Acta Bioenerg. 2001;1507:115–138. doi: 10.1016/s0005-2728(01)00199-2. [DOI] [PubMed] [Google Scholar]
  43. Jarvis J.A., Concistre M., Haies I.M., Bounds R.W., Kuprov I., Carravetta M., Williamson P.T.F. Quantitative analysis of 14N quadrupolar coupling using 1H detected 14N solid-state NMR. Phys. Chem. Chem. Phys. 2019;21:5941–5949. doi: 10.1039/c8cp06276e. [DOI] [PubMed] [Google Scholar]
  44. Johnson T.W., Shen G., Zybailov B., Kolling D., Reategui R., Beauparlant S., Vassiliev I.R., Bryant D.A., Jones A.D., Golbeck J.H., Chitnis P.R. Recruitment of a foreign quinone into the A1 site of photosystem I. I. Genetic and physiological characterization of phylloquinone biosynthetic pathway mutants in Synechocystis sp. pcc 6803. J. Biol. Chem. 2000;275:8523–8530. doi: 10.1074/jbc.275.12.8523. [DOI] [PubMed] [Google Scholar]
  45. Johnson T.W., Zybailov B., Jones A.D., Bittl R., Zech S., Stehlik D., Golbeck J.H., Chitnis P.R. Recruitment of a foreign quinone into the A1 site of photosystem I. J. Biol. Chem. 2001;276:39512–39521. doi: 10.1074/jbc.M104040200. [DOI] [PubMed] [Google Scholar]
  46. Jordan P., Fromme P., Witt H.T., Klukas O., Saenger W., Krauss N. Three-dimensional structure of cyanobacterial photosystem I at 2.5 A resolution. Nature. 2001;411:909–917. doi: 10.1038/35082000. [DOI] [PubMed] [Google Scholar]
  47. Kaess H., Rautter J., Boenigk B., Hoefer P., Lubitz W. 2D ESEEM of the 15N-labeled radical cations of bacteriochlorophyll a and of the primary donor in reaction centers of Rhodobacter sphaeroides. J. Phys. Chem. 1995;99:436–448. [Google Scholar]
  48. Kamerlin S.C.L., Haranczyk M., Warshel A. Progress in ab initio QM/MM free-energy simulations of electrostatic energies in proteins: Accelerated QM/MM studies of pKa, redox reactions and solvation free energies. J. Phys. Chem. B. 2009;113:1253–1272. doi: 10.1021/jp8071712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Käss H., Fromme P., Witt H.T., Lubitz W. Orientation and electronic structure of the primary donor radical cation in photosystem I:  A single crystals EPR and ENDOR Study. J. Phys. Chem. B. 2001;105:1225–1239. [Google Scholar]
  50. Katz J.J., Norris J.R., Shipman L.L., Thurnauer M.C., Wasielewski M.R. Chlorophyll function in the photosynthetic reaction center. Annu. Rev. Biophys. Bioeng. 1978;7:393–434. doi: 10.1146/annurev.bb.07.060178.002141. [DOI] [PubMed] [Google Scholar]
  51. Käβ H., Bittersmann-Weidlich E., Andréasson L.E., Bönigk B., Lubitz W. ENDOR and ESEEM of the 15N labelled radical cations of chlorophyll a and the primary donor P700 in photosystem I. Chem. Phys. 1995;194:419–432. [Google Scholar]
  52. Käβ H., Fromme P., Lubitz W. Quadrupole parameters of nitrogen nuclei in the cation radical P700.+ determined by ESEEM of single crystals of photosystem I. Chem. Phys. Lett. 1996;257:197–206. [Google Scholar]
  53. Käβ H., Lubitz W. Evaluation of 2D ESEEM data of 15N-labeled radical cations of the primary donor P700 in photosystem I and chlorophyll a. Chem. Phys. Lett. 1996;251:193–203. [Google Scholar]
  54. Kurashov V., Gorka M., Milanovsky G.E., Johnson T.W., Cherepanov D.A., Semenov A.Y., Golbeck J.H. Critical evaluation of electron transfer kinetics in P700-FA/FB, P700-FX, and P700-A1 photosystem I core complexes in liquid and in trehalose glass. Biochim. Biophys. Acta Bioenerg. 2018;1859:1288–1301. doi: 10.1016/j.bbabio.2018.09.367. [DOI] [PubMed] [Google Scholar]
  55. Lakshmi K.V., Jung Y.-S., Golbeck J.H., Brudvig G.W. Location of the iron−sulfur clusters FA and FB in photosystem I: An electron paramagnetic resonance study of spin relaxation enhancement of P700+ Biochemistry. 1999;38:13210–13215. doi: 10.1021/bi9910777. [DOI] [PubMed] [Google Scholar]
  56. Lancaster C.R., Hunte C., Kelley J., 3rd, Trumpower B.L., Ditchfield R. A comparison of stigmatellin conformations, free and bound to the photosynthetic reaction center and the cytochrome bc1 complex. J. Mol. Biol. 2007;368:197–208. doi: 10.1016/j.jmb.2007.02.013. [DOI] [PubMed] [Google Scholar]
  57. Lendzian F., Geβern C., Bönigk B., Plato M., Möbius K., Lubitz W. 15N- and 1H-ENDOR/TRIPLE resonance of the primary donor cation radical D+• in isotopically labelled reaction centers of Rhodobacter sphaeroides. In: Murata N., editor. Research in Photosynthesis, Proceedings of the IXth International Congress on Photosynthesis. Kluwer; 1992. pp. 57–68. [Google Scholar]
  58. Lendzian F., Huber M., Isaacson R.A., Endeward B., Plato M., Bönigk B., Möbius K., Lubitz W., Feher G. The electronic structure of the primary donor cation radical in Rhodobacter sphaeroides R-26: ENDOR and TRIPLE resonance studies in single crystals of reaction centers. Biochim. Biophys. Acta Bioenerg. 1993;1183:139–160. [Google Scholar]
  59. Lendzian F., Mobius K., Lubitz W. The pheophytin a anion radical. 14N and 1H ENDOR and triple resonance in liquid solution. Chem. Phys. Lett. 1982;90:375–381. [Google Scholar]
  60. Li Y., Est A., Lucas M., Vm R., Gu F., Petrenko A., Lin S., Webber A., Rappaport F., Redding K. Directing electron transfer within photosystem I by breaking H-bonds in the cofactor branches. Proc. Natl. Acad. Sci. U S A. 2006;103:2144–2149. doi: 10.1073/pnas.0506537103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Llansola-Portoles M.J., Gust D., Moore T.A., Moore A.L. Artificial photosynthetic antennas and reaction centers. Comptes Rendus Chim. 2017;20:296–313. [Google Scholar]
  62. Lubitz W., Lendzian F., Bittl R. Radicals, radical pairs and triplet states in photosynthesis. Acc. Chem. Res. 2002;35:313–320. doi: 10.1021/ar000084g. [DOI] [PubMed] [Google Scholar]
  63. Mac M., Bowlby N.R., Babcock G.T., McCracken J. Monomeric spin density distribution in the primary donor of photosystem I as determined by electron magnetic resonance:  Functional and thermodynamic implications. J. Amer Chem. Soc. 1998;120:13215–13223. [Google Scholar]
  64. Mac M., Tang X.-S., Diner B.A., McCracken J., Babcock G.T. Identification of histidine as an axial ligand to P700+ Biochemistry. 1996;35:13288–13293. doi: 10.1021/bi961765f. [DOI] [PubMed] [Google Scholar]
  65. Madjet M.E.-A., Müh F., Renger T. Deciphering the influence of short-range electronic couplings on optical properties of molecular dimers: application to “special pairs” in photosynthesis. J. Phys. Chem. B. 2009;113:12603–12614. doi: 10.1021/jp906009j. [DOI] [PubMed] [Google Scholar]
  66. Meyer T.E. Evolution of photosynthetic reaction centers and light harvesting chlorophyll proteins. Biosystems. 1994;33:167–175. doi: 10.1016/0303-2647(94)90002-7. [DOI] [PubMed] [Google Scholar]
  67. Milanovsky G.E., Ptushenko V.V., Cherepanov D.A., Semenov A.-Y. Mechanism of primary and secondary ion-radical pair formation in photosystem I complexes. Biochemistry. 2014;79:221–226. doi: 10.1134/S0006297914030079. [DOI] [PubMed] [Google Scholar]
  68. Moia D., Abe M., Wagner P., Saguchi H., Koumura N., Nelson J., Barnes P.R.F., Mori S. The effect of the dielectric environment on electron transfer reactions at the interfaces of molecular sensitized semiconductors in electrolytes. J. Phys. Chem. C. 2020;124:6979–6992. [Google Scholar]
  69. Moser C.C., Keske J.M., Warncke K., Farid R.S., Dutton P.L. Nature of biological electron transfer. Nature. 1992;355:796–802. doi: 10.1038/355796a0. [DOI] [PubMed] [Google Scholar]
  70. Müller M.G., Niklas J., Lubitz W., Holzwarth A.R. Ultrafast transient absorption studies on photosystem I reaction centers from Chlamydomonas reinhardtii. 1. A new interpretation of the energy trapping and early electron transfer steps in photosystem I. Biophys. J. 2003;85:3899–3922. doi: 10.1016/S0006-3495(03)74804-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Müller M.G., Slavov C., Luthra R., Redding K.E., Holzwarth A.R. Independent initiation of primary electron transfer in the two branches of the photosystem I reaction center. Proc. Natl. Acad. Sci. U S A. 2010;107:4123–4128. doi: 10.1073/pnas.0905407107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Neese F. The ORCA program system. Wires Comput. Mol. Sci. 2012;2:73–78. [Google Scholar]
  73. Neese F. Software update: the ORCA program system, version 4.0. Wires Comput. Mol. Sci. 2018;8:e1327. [Google Scholar]
  74. Norris J.R., Scheer H., Druyan M.E., Katz J.J. An electron-nuclear double resonance (ENDOR) study of the special pair model for photo-reactive chlorophyll in photosynthesis. Proc. Natl. Acad. Sci. U S A. 1974;71:4897–4900. doi: 10.1073/pnas.71.12.4897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Norris J.R., Uphaus R.A., Crespi H.L., Katz J.J. Electron spin resonance of chlorophyll and the origin of signal I in photosynthesis. Proc. Natl. Acad. Sci. U S A. 1971;68:625–628. doi: 10.1073/pnas.68.3.625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. O’Malley P.J. The electronic structure of the bacteriopheophytin a anion radical, in vivo. J. Amer. Chem. Soc. 1999;121:3185–3192. [Google Scholar]
  77. O'Malley P.J. The effect of oxidation and reduction of chlorophyll a on its geometry, vibrational and spin density properties as revealed by hybrid density functional methods. J. Amer Chem. Soc. 2000;122:7798–7801. [Google Scholar]
  78. O'Malley P.J. Hybrid density functional studies of pheophytin anion radicals:  Implications for initial electron transfer in photosynthetic reaction centers. J. Phys. Chem. B. 2000;104:2176–2182. [Google Scholar]
  79. Olson J.M., Pierson B.K. Origin and evolution of photosynthetic reaction centers. Orig Life Evol. Biosph. 1987;17:419–430. [Google Scholar]
  80. O’Malley P.J., Collins S.J. The effect of axial Mg ligation on the geometry and spin density distribution of chlorophyll and bacteriochlorophyll cation free radical models: A density functional study. J. Amer. Chem. Soc. 2001;123:11042–11046. doi: 10.1021/ja010522u. [DOI] [PubMed] [Google Scholar]
  81. Orf G.S., Gisriel C., Redding K.E. Evolution of photosynthetic reaction centers: insights from the structure of the heliobacterial reaction center. Photosynth. Res. 2018;138:11–37. doi: 10.1007/s11120-018-0503-2. [DOI] [PubMed] [Google Scholar]
  82. Pandey P.K.K., Chandra P. Theoretical study of isotropic hyperfine coupling constants in small radicals by MINDO/3 method. Can. J. Chem. 1979;57:3126–3134. [Google Scholar]
  83. Parusel A.B.J., Grimme S. A theoretical study of the excited states of chlorophyll a and pheophytin a. J. Phys. Chem. B. 2000;104:5395–5398. [Google Scholar]
  84. Pearson R.G. Hard and soft acids and bases. J. Amer. Chem. Soc. 1963;85:3533–3539. [Google Scholar]
  85. Plato M., Kraus N., Fromme P., Lubitz W. Molecular orbital study of the primary electron donor P700 of photosystem I based on a recent X-ray single crystal structure analysis. Chem. Phys. 2003;294:483–499. [Google Scholar]
  86. Poluektov O.G., Paschenko S.V., Utschig L.M., Lakshmi K.V., Thurnauer M.C. Bidirectional electron transfer in photosystem I: Direct evidence from high-frequency time-resolved EPR spectroscopy. J. Amer. Chem. Soc. 2005;127:11910–11911. doi: 10.1021/ja053315t. [DOI] [PubMed] [Google Scholar]
  87. Ptushenko V.V., Cherepanov D.A., Krishtalik L.I., Semenov A.Y. Semi-continuum electrostatic calculations of redox potentials in photosystem I. Photosyn Res. 2008;97:55. doi: 10.1007/s11120-008-9309-y. [DOI] [PubMed] [Google Scholar]
  88. Rabinowitch E.I., Govindjee The role of chlorophyll in photosynthesis. Sci. Amer. 1965;213:74–83. doi: 10.1038/scientificamerican0765-74. [DOI] [PubMed] [Google Scholar]
  89. Rigby S.E.J., Muhiuddin I.P., Santabarbara S., Evans M.C.W., Heathcote P. Proton ENDOR spectroscopy of the anion radicals of the chlorophyll primary electron acceptors in type I photosynthetic reaction centres. Chem. Phys. 2003;294:319–328. [Google Scholar]
  90. Rögner M., Nixon P.J., Diner B.A. Purification and characterization of photosystem I and photosystem II core complexes from wild-type and phycocyanin-deficient strains of the cyanobacterium Synechocystis PCC 6803. J. Biol. Chem. 1990;265:6189–6196. [PubMed] [Google Scholar]
  91. Roy C., Lancaster D., Ermler U., Michel H. The structures of photosynthetic reaction centers from purple bacteria as revealed by X-ray crystallography. In: Blankenship R.E., Madigan M.T., Bauer C.E., editors. Anoxygenic Photosynthetic Bacteria. Springer; 1995. pp. 503–526. [Google Scholar]
  92. Sattley W.M., Swingley W.D. Properties and evolutionary implications of the heliobacterial genome. Adv. Bot. Res. 2013;66:67–97. [Google Scholar]
  93. Savitsky A., Gopta O., Mamedov M., Golbeck J.H., Tikhonov A., Möbius K., Semenov A. Alteration of the axial met ligand to electron acceptor A0 in photosystem I: effect on the generation of P700·+A1·− radical pairs as studied by W-band transient EPR. Appl. Magn. Reson. 2009;37:85. [Google Scholar]
  94. Schnegg A., Dubinskii A.A., Fuchs M.R., Grishin Y.A., Kirilina E.P., Lubitz W., Plato M., Savitsky A., Möbius K. High-field EPR, ENDOR and ELDOR on bacterial photosynthetic reaction centers. Appl. Magn. Reson. 2007;31:59–98. [Google Scholar]
  95. Şener M., Strümpfer J., Hsin J., Chandler D., Scheuring S., Hunter C.N., Schulten K. Förster energy transfer theory as reflected in the structures of photosynthetic light-harvesting systems. Chem. Phys. Chem. 2011;12:518–531. doi: 10.1002/cphc.201000944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Shelaev I.V., Gostev F.E., Mamedov M.D., Sarkisov O.M., Nadtochenko V.A., Shuvalov V.A., Semenov A.Y. Femtosecond primary charge separation in Synechocystis sp. PCC 6803 photosystem I. Biochim. Biophys. Acta. 2010;1797:1410–1420. doi: 10.1016/j.bbabio.2010.02.026. [DOI] [PubMed] [Google Scholar]
  97. Shuvalov V.A., Dolan E., Ke B. Spectral and kinetic evidence for two early electron acceptors in photosystem I. Proc. Natl. Acad. Sci. U S A. 1979;76:770–773. doi: 10.1073/pnas.76.2.770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Shuvalov V.A., Bacon K.E., Dolan E. Kinetic and spectral properties of the intermediary electron acceptor A1 in photosystem I: Subnanosecond spectroscopy. FEBS Lett. 1979;100:5–8. doi: 10.1016/0014-5793(79)81119-9. [DOI] [PubMed] [Google Scholar]
  99. Shuvalov V.A., Klevanik A.V., Sharkov A.V., Kryukov P.G., Bacon K.E. Picosecond spectroscopy of photosystem I reaction centers. FEBS Lett. 1979;107:313–316. doi: 10.1016/0014-5793(79)80397-x. [DOI] [PubMed] [Google Scholar]
  100. Sinnecker S., Koch W., Lubitz W. Bacteriochlorophyll a radical cation and anion—calculation of isotropic hyperfine coupling constants by density functional methods. Phys. Chem. Chem. Phys. 2000;2:4772–4778. [Google Scholar]
  101. Sinnecker S., Koch W., Lubitz W. Chlorophyll a radical ions:  A density functional study. J. Phys. Chem. B. 2002;106:5281–5288. [Google Scholar]
  102. Sirohiwal A., Neese F., Pantazis D.A. Protein matrix control of reaction center excitation in photosystem II. J. Amer. Chem. Soc. 2020;142:18174–18190. doi: 10.1021/jacs.0c08526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Sproviero E.M., Gascón J.A., McEvoy J.P., Brudvig G.W., Batista V.S. QM/MM models of the O2-evolving complex of photosystem II. J. Chem. Theor. Comput. 2006;2:1119–1134. doi: 10.1021/ct060018l. [DOI] [PubMed] [Google Scholar]
  104. Srinivasan N., Chatterjee R., Milikisiyants S., Golbeck J.H., Lakshmi K.V. Effect of hydrogen bond strength on the redox properties of phylloquinones: a two-dimensional hyperfine sublevel correlation spectroscopy study of photosystem I. Biochemistry. 2011;50:3495–3501. doi: 10.1021/bi102056q. [DOI] [PubMed] [Google Scholar]
  105. Srinivasan N., Golbeck J.H. Protein-cofactor interactions in bioenergetic complexes: The role of the A1A and A1B phylloquinones in photosystem I. Biochim. Biophys. Acta. 2009;1787:1057–1088. doi: 10.1016/j.bbabio.2009.04.010. [DOI] [PubMed] [Google Scholar]
  106. Srinivasan N., Karyagina I., Bittl R., van der Est A., Golbeck J.H. Role of the hydrogen bond from Leu722 to the A1A phylloquinone in photosystem I. Biochemistry. 2009;48:3315–3324. doi: 10.1021/bi802340s. [DOI] [PubMed] [Google Scholar]
  107. Srinivasan N., Santabarbara S., Rappaport F., Carbonera D., Redding K., van der Est A., Golbeck J.H. Alteration of the H-bond to the A1A phylloquinone in photosystem I: influence on the kinetics and energetics of electron transfer. J. Phys. Chem. B. 2011;115:1751–1759. doi: 10.1021/jp109531b. [DOI] [PubMed] [Google Scholar]
  108. Stoll S., Calle C., Mitrikas G., Schweiger A. Peak suppression in ESEEM spectra of multinuclear spin systems. J. Magn. Reson. 2005;177:93–101. doi: 10.1016/j.jmr.2005.07.012. [DOI] [PubMed] [Google Scholar]
  109. Stoll S., Schweiger A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 2006;178:42–55. doi: 10.1016/j.jmr.2005.08.013. [DOI] [PubMed] [Google Scholar]
  110. Sun J., Hao S., Radle M., Xu W., Shelaev I., Nadtochenko V., Shuvalov V., Semenov A., Gordon H., van der Est A., Golbeck J.H. Evidence that histidine forms a coordination bond to the A0A and A0B chlorophylls and a second H-bond to the A1A and A1B phylloquinones in M688HPsaA and M668HPsaB variants of Synechocystis sp. PCC 6803. Biochim. Biophys. Acta Bioenerg. 2014;1837:1362–1375. doi: 10.1016/j.bbabio.2014.04.004. [DOI] [PubMed] [Google Scholar]
  111. Sun Y., Dai Z., Wang W., Sun Y. A DFT study on the electronic character of P700+ J. Theor. Comput. Chem. 2006;05:733–741. [Google Scholar]
  112. Taguchi A.T., Baldansuren A., Dikanov S.A. Basic and combination cross-features in X- and Q-band HYSCORE of the 15N labeled bacteriochlorophyll a cation radical. Z. Phys. Chem. 2017;231:725–743. [Google Scholar]
  113. Thornber J.P. Chlorophyll-proteins: Light-harvesting and reaction center components of plants. Annu. Rev. Plant Physiol. 1975;26:127–158. [Google Scholar]
  114. Umena Y., Kawakami K., Shen J.R., Kamiya N. Crystal structure of oxygen-evolving photosystem II at a resolution of 1.9 A. Nature. 2011;473:55–60. doi: 10.1038/nature09913. [DOI] [PubMed] [Google Scholar]
  115. van der Est A. Electron transfer involving phylloquinone in photosystem I. In: Golbeck J.H., editor. Phototosystem I: The Light-Induced Plastocyanin:Ferredoxin Oxidoreductase. Springer; 2006. pp. 387–411. [Google Scholar]
  116. van Doorslaer S. Understanding heme proteins with hyperfine spectroscopy. J. Magn. Reson. 2017;280:79–88. doi: 10.1016/j.jmr.2017.01.008. [DOI] [PubMed] [Google Scholar]
  117. Wasielewski M.R., Fenton J.M., Govindjee The rate of formation of P700+–A0- in photosystem I particles from spinach as measured by picosecond transient absorption spectroscopy. Photosyn Res. 1987;12:181–189. doi: 10.1007/BF00047947. [DOI] [PubMed] [Google Scholar]
  118. Webber A.N., Lubitz W. P700: The primary electron donor of photosystem I. Biochim. Biophys. Acta Bioenerg. 2001;1507:61–79. doi: 10.1016/s0005-2728(01)00198-0. [DOI] [PubMed] [Google Scholar]
  119. Xu W., Chitnis P.R., Valieva A., van der Est A., Brettel K., Guergova-Kuras M., Pushkar Y.N., Zech S.G., Stehlik D., Shen G.Z. Electron transfer in cyanobacterial photosystem I - II. Determination of forward electron transfer rates of site-directed mutants in a putative electron transfer pathway prom A0 through A1 to FX. J. Biol. Chem. 2003;278:27876–27887. doi: 10.1074/jbc.M302965200. [DOI] [PubMed] [Google Scholar]
  120. Xu W., Chitnis P.R., Valieva A., van der Est A., Pushkar Y.N., Krzystyniak M., Teutloff C., Zech S.G., Bittl B., Stehlik D. Electron transfer in cyanobacterial photosystem I - I. Physiological and spectroscopic characterization of site-directed mutants in a putative electron transfer pathway from A0 through A1 to FX. J. Biol. Chem. 2003;278:27864–27875. doi: 10.1074/jbc.M302962200. [DOI] [PubMed] [Google Scholar]
  121. Yin S., Dahlbom M.G., Canfield P.J., Hush N.S., Kobayashi R., Reimers J.R. Assignment of the Qy absorption spectrum of photosystem I from Thermosynechococcus elongatus based on CAM-B3LYP calculations at the PW91-optimized protein structure. J. Phys. Chem. B. 2007;111:9923–9930. doi: 10.1021/jp070030p. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Download video file (41.5MB, mp4)
Document S1. Figures S1–S13 and Tables S1–S4
mmc1.pdf (1.1MB, pdf)
Video abstract
Download video file (41.5MB, mp4)

Data Availability Statement

Numerically simulated parameters in this study are available by contacting the lead contact, K. V. Lakshmi (lakshk@rpi.edu).


Articles from iScience are provided here courtesy of Elsevier

RESOURCES