.143 A. .51..“ o . a fig? Ln...» .. 1:... :50.“ .1 . .x. a .r u . 5%.?- AS. .333..{..... “W...n.r.a¢a.?kt 3.32....I .! ’f #5533... 2.. 3v 2 .s a!!! a I . :in {hum ”flrfifig .5... Ash”-.. ”2 Hi to! .Ilar. 1’55... 2...; .3. 1.1.2: .I (:32. SJ... .6931: 1.. J: a .9...!'....u.ddfl. Jr :3?! V 1.3).: r, 11—... :7: 4 s) 1...... . v -.K-lv}..‘ 52.9.. C. A: . 7133......293‘ :2. .t 10...: C 2...... .I :: .II... I. It... :2 a: t. I...) x...|\ '|b:l . >1L1F.rw 1‘. THESIS /.r I‘ "3-) LVU/ This is to certify that the dissertation entitled THE EFFECTS OF SR2+-SUBSTITUTION ON THE REDUCTION RATES OF YZ' AND ON THE HYDROGEN ATOM ABSTRACTION MECHANISM IN PHOTOSYSTEM 11 presented by Kristi L. Westphai has been accepted towards fulfillment of the requirements for Ph.D. degree in Chemistry fl 7 477 xfy/wé Major professor / Date fly”? M 2/07? J MS U i: an Affirmative Action/Equal Opportunity Institution 0-12771 n mar-mm T Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 CJCIRC/Dateouop65-p. 15 THE EFFECTS OF SRZT-SUBSTITUTION ON THE REDUCTION RATES OF Yz' AND ON THE HYDROGEN ATOM ABSTRACTION MECHANISM IN PHOTOSYSTEM II By Kristi L. Westphal A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2000 ABSTRACT THE EFFECTS OF SRZT-SUBSTITUTION ON THE REDUCTION RATES OF Yz° AND ON THE HYDROGEN ATOM ABSTRACTION MECHANISM IN PHOTOSYTEM II By Kristi L. Westphal Photosystem II is a photosynthetic enzyme that catalyzes the light-driven oxidation of water to molecular oxygen. A total of four sequential light-induced turnovers of Photosystem II are required to form molecular oxygen. Each stable intermediate produced by a flash of light is called an Sn-state, n = 0-4. A general consensus has not yet been reached concerning the mechanism of this reaction or the structure of Photosystem ll, although much progress has been made over the last few years. It is generally agreed upon that the oxygen-evolving complex of Photosystem ll consists of a manganese cluster made up of four manganese atoms, a redox-active tyrosine called Y2, histidine 190, and Ca2+ and Cl' ions. Recently, several models have been proposed that directly involve Yz in the water oxidation chemistry as a hydrogen atom abstractor. Perhaps the most elusive subject is the role of Ca2+ in PSII. Two main proposals have been set forth: 1) Ca” plays a structural role and a binding site for CI' and, 2) Ca2+ is directly involved in the oxygen-oxygen bond forming reaction. One way to acquire more information on the role of Ca” is to perform biochemical treatments that perturb this cofactor and then monitor the kinetic efficiency of the oxygen- evolving complex. Several groups have recently investigated the effects of biochemical treatments, site-directed mutagenesis, or substitution of essential cofactors on the kinetics of the stepwise, water-oxidizing chemistry catalyzed by Photosystem ll. Consistently, these studies showed evidence for a slowing of the final, oxygen-releasing step, 83-)80, of the catalytic cycle. To a degree, some of this work also showed a slowing of the earlier S-state transitions. To test the roles that have been proposed for Ca”, PSII membranes were depleted of Ca2+ and subsequently reconstitution with Sr”. The rates of the S-state transitions in these samples were monitored by using time-resolved electron paramagnetic resonance spectroscopy. The results show a slowdown of the last transition in the cycle, consistent with an earlier report from Boussac and Rutherford (Boussac, A., Setif, P. and Rutherford, AW. (1992) Biochemistry 31, 1224-1234), and of the earlier S-state transitions as well. Our observations that both the lower and higher S-state transition rates decrease suggest that a common molecular mechanism is at work and that Sr” is less effective than Ca2+ in supporting it. The time-resolved electron paramagnetic resonance spectroscopy results were also analyzed in the context of the hydrogen atom abstraction mechanism. The retarding effect caused by Sr” substitution is additional evidence to support this mechanism. Furthermore, the hydrogen atom abstraction mechanism is best able to explain the unusual characteristics of the S-state transitions: slow rates, low activation energies, low pre-exponential factors, and small kinetic isotope effects. In Memory of My Father, Bruce G. Westphal ACKNOWLEDGMENTS I would like to begin by thanking my research advisor, Jerry Babcock, for his guidance, patience, and support. I couldn’t have asked for a better advisor! I would like to thank my mother, Diana Westphal, for her words of encouragement when I needed them most and for her never-ending faith in my becoming Dr. Westphal. I would also like to thank my boyfriend, Marcus Sims, for putting up with me when I knew I was unbearable to be around and for always being able to make me laugh! Also, for his words of encouragement when l was pulling my hair out in frustration (especially writing this thingl). I would lastly like to thank my lab mates and friends for being there when I needed help or just needed a good laugh (some of you probably didn’t even know you were supplying the latter!) I would particularly like to thank three of these people. First, my partner in crime, Heather Sackett, without you I would have had no one to talk stupid with or to surf the net for questionable web sites! Second, my second partner in crime and softball buddie, Shannon Haymond, may we never have another wine night like the first! Thirdly, Hans Schelvis, my desk neighbor for the past five years. You were always there to answer my questions and to make the best coffee! I will miss you all dearly. Thanks to everyone who have made my graduate school experience one that I will never forget! vi TABLE OF CONTENTS List of Tables ....................................................................................... vii List of Figures ..................................................................................... viii Chapter 1 .............................................................................................................. 1 1 Introduction ..................................................................................................... 1 1.1 The Electron Transport Chain in Photosynthesis ..................................... 1 1.2 Photosystem ll ......................................................................................... 3 1.2.1 Polypeptide Composition and Electron Transfer Reactions ................ 3 1.2.2 S-state Cycle in Water Oxidation ........................................................ 6 1.2.3 The Energetics of Water Oxidation ..................................................... 8 1.2.4 The Manganese Cluster ...................................................................... 8 1.2.5 Y0 and Y2 .......................................................................................... 12 1.2.6 A Model for Water Oxidation ............................................................. 13 1.2.7 The Reduction Rates of Y2' ............................................................... 15 1.2.8 Calcium ............................................................................................. 19 1.2.9 Proposed Roles for Ca2+ in Water Oxidation ..................................... 20 1.3 Project Goals ......................................................................................... 22 Chapter 2 ............................................................................................................ 25 2 Electron Paramagnetic Resonance Spectroscopy ....................................... 25 2.1 Basic Principles ...................................................................................... 25 2.2 The EPR Spectrometer .......................................................................... 28 Chapter 3 ............................................................................................................ 33 3 Materials and Methods ................................................................................. 33 3.1 Preparation of PSII Membranes ............................................................. 33 3.2 Calcium Depletion of PSII Membranes .................................................. 35 3.3 Sample Preparation for EPR Experiments ............................................. 35 3.4 Time-Resolved EPR Experiment Setup ................................................. 36 Chapter 4 ............................................................................................................ 39 4 Results of Time-Resolved EPR Experiments ............................................... 39 4.1 S-State Mixed Kinetics ........................................................................... 39 4.2 S-State Resolved Kinetics ...................................................................... 41 4.3 Effects of pH and Isotope Exchange ...................................................... 45 Chapter 5 ............................................................................................................ 53 5 Hydrogen Atom Abstraction .......................................................................... 53 5.1 Electron Transfer ................................................................................... 53 5.2 Proton Coupled Electron Transfer: Sequential or Concerted? ............... 56 5.3 Concerted Electron and Proton Transfer ................................................ 62 Bibliography ........................................................................................ 71 vii Table 1-1 Table 4-1 Table 5-1 Table 5-2 Table 5-3 Table 5-4 LIST OF TABLES Half-times for S-state Advance in Various Enzyme Preparation Types .............................................................. 18 Rates of Y2' Reduction for Ca2+- or Sr2+-Reconstituted PSII Membranes Determined by Time-Resolved EPR Spectroscopy .................................................................... 48 Reduction Potentials and pKa Values Used in Determining the Viability of the ET/PT and PT/ET Pathways ............................. 58 Kinetic and Thermodynamic Parameters for the S-state Transitions ....................................................................... 61 Kinetic and Thermodynamic Parameters for H-atom Abstraction Reactions From an Oxygen Atom by an Oxy-based Radical ............................................................. 63 Kinetic and Thermodynamic Parameters for Hydrogen Atom Abstraction by Radicals ............................................... 64 viii Figure 1-1 Figure 1-2 Figure 1-3 Figure 1-4 Figure 1-5 Figure 1-6 Figure 1-7 Figure 2-1 Figure 2-2 Figure 2-3 Figure 3-1 Figure 4-1 Figure 4-2 Figure 4-3 Figure 4-4 Figure 4-5 LIST OF FIGURES Flow of electrons through the electron transport chain in higher plants ...................................................................... 2 Schematic representation of the major polypeptides and redox components of PSII ............................................... 4 A period-four oscillation of oxygen evolution as a function of flash number in spinach chloroplasts ........................ 7 Redox potentials of the cofactors of PSII .................................. 9 The 9 = 4.1 (a) and multiline (b) EPR signals observed in PSII membranes ............................................................. 11 Tyrosine Yz oxidation and reduction in the 80 state .................... 14 A model for the S-state cycle in PSII ....................................... 16 Energy level scheme for a free electron as a function of applied magnetic field B ...................................................... 27 Block diagram for a continuous wave EPR spectrometer ............ 29 Continuous-wave EPR spectrum of Y0' ................................... 31 Setup for the time-resolved EPR experiments .......................... 37 Time-resolved EPR steady state kinetic traces (S-states are mixed) for Y; reduction in salt-washed PSII membranes reconstituted with (a) CaClz or (b) SrClz .................................. 40 Three-flash S-state resolved EPR kinetic traces ........................ 42 Time-resolved EPR kinetic traces for the S1Yz' -) 8ng transition triggered by the first flash in a three flash experiment....44 Time-resolved EPR kinetic traces for the SzYz' -) 8ng transition triggered by the second flash in a three flash experiment ........................................................................ 46 Time-resolved EPR kinetic traces for the Sng' -> SOYz transition triggered by the third flash in a three flash experiment ........................................................................ 47 ix Figure 4-6 Figure 4-7 Figure 5-1 Figure 5-2 Figure 5-3 Time-resolved EPR kinetic traces for the S1Yz'-) SzYz, SzYz' -> Sng, Sng' -> SOYz transitions in salt-washed, Srztreconstituted PSII membranes resuspended in buffers at pH = 5.9, 6.5, 7.1 ............................................................ 49 Time-resolved EPR steady state kinetic traces (S-states are mixed) for Y; reduction in salt-washed PSII membranes reconstituted with SrClz in either D20 or H20 ........... 52 Potential energy surfaces of the donor and acceptor .................. 55 Possible reaction paths for proton and electron transfer ............. 57 Calculated activation energies (Ea) for paths 3 and b ................. 60 Chapter 1 1 Introduction Two to three billion years ago photosynthetic cyanobacteria gained the capability to extract electrons from water and produce molecular oxygen as a byproduct. These organisms flourished on an inexhaustible supply of water, transforming the earth’s atmosphere from a composition of reduced gases to one rich in 02. Prior to this event, all organisms on earth were adapted to an anaerobic environment. As the atmosphere became enriched in oxygen, many organisms became extinct while the others adapted to the new environment. As a consequence, the highly efficient process of respiration evolved and higher organisms developed. Today, members of the higher plants, algae, and cyanobacteria families are all capable of carrying out oxygenic photosynthesis. 1.1 The Electron Transport Chain in Photosynthesis Photosynthesis in green plants is carried out by two photosystems, Photosystem l (PSI) and Photosystem ll (PSII) which undergo photochemical charge separation. The two photosystems are located in the stromal region of the chloroplast in membranous structures called thylakoids, which are located within the leaf of the plant. Two additional membranous components are also involved in the electron transport chain: the cytochrome bef complex (cyt bet) and adenosine 5’-triphosphate (ATP) synthase. The flow of electrons through these components is illustrated in Figure 1.1 and is initiated by a light-induced charge separation in PSII, which takes electrons from water and delivers them to the mobile electron carrier plastoquinone (PQ). Next, PQ travels to cyt bef, which is 1 ADP + Pi ATP NADPH A . . ~ 5 ATPZ PSI synthase 2 H26 .><7__J 02 + 4H+ E. Figure 1-1 Flow of electrons through the electron transport chain in higher plants. This figure was kindly provided by Wanrvick Hillier. involved in proton translocation, and then the electrons are taken up by a second mobile electron carrier called plastocyanin (PC) which delivers the electrons to PSI. Here the electrons are utilized in the terminal step of the electron transport chain where a molecule of nicotinamide adenine dinucleotide phosphate (NADP") is reduced to NADPH. In addition to the utilization of electrons for NADPH production, electron flow within each photosystem and between them induces a membrane potential through the translocation of protons. The electrochemical potential produced is then dissipated by ATP synthase in the production of ATP. The photosynthetic end products, NADPH and ATP, are both ultimately used by the plant to drive the synthesis of complex biomolecules essential for growth and reproduction. 1.2 Photosystem ll 1.2.1 Polypeptide Composition and Electron Transfer Reactions The photochemistry of PSII is performed on a protein scaffold that binds pigments and cofactors involved in the reaction. A schematic representation of the primary photochemical events that occur at PSII is depicted in Figure 1-2 (for reviews see (1, 2)). Light is first absorbed by the light-harvesting complex located peripherally to the PSII core, which consists of non-covalently bound pigment molecules including chlorophyll a, chlorophyll b and carotenoids (not shown in figure). The absorbed photon is then funneled into the reaction center to a specialized chlorophyll complex referred to as P680. Excitation of P580 results in formation of the radical pair, P530'+Pheo“, which is stabilized by rapid electron CP43 cyt b-559 / 2 m1 l7 Figure 1-2 Schematic representation of the major polypeptides and redox components of PSII. The arrows show the flow of electrons. transfer to a bound plastoquinone molecule, GA. A second plastoquinone molecule, 03, oxidizes QA' within 100-200 ps, and after an additional charge- separation and subsequent protonation, the QBHz plastoquinol (PQ) formed leaves the reaction center and is replaced by an oxidized quinone from the plastoquinone pool. The electron hole that is produced at P530 is quickly reduced in nanoseconds by a redox-active tryosine residue, Yz at D1-161. Subsequently, Yz is reduced on the microsecond to millisecond time-scale by a Mn cluster consisting of four Mn atoms. The (Mn)4 cluster and Y2, along with the necessary cofactors Ca2+ and Cl‘, make up what is called the oxygen-evolving complex (OEC) of PSII, and it is at this site where the oxidation of water is catalyzed. In vivo, PSII contains at least 20 polypeptides, but the minimal oxygen- evolving complex consists of seven major polypeptides: D1, DZ, CP43, CP47, the or and B polypeptides of cyt b-559, and the extrinsic polypeptide of 33 kDa (for a recent review see (3)). The D1 and DZ polypeptides, which each contain five transmembrane helices, form the photochemical reaction center by binding P580, Pheo, the non-heme iron, GA, 03, and contain the tyrosine Y2, as well as a second redox active tyrosine referred to as YD. Both the D1 and DZ proteins exhibit a very high degree of sequence homology and appear to form a heterodimer with C2 symmetry. It is estimated that there are 6 chlorophyll 3 molecules and 2 fl-carotene molecules per DI/D2 reaction center. An oxygen- evolving reaction center requires, in addition to D1 and DZ, two other chlorophyll- containing polypeptides known as CP47 and CP43. Both peptides contain about 15 bound chlorophyll 3 molecules and 2-3 molecules of B-carotene; they are believed to act as inner antennae that transfer energy from the light-harvesting complex to the reaction center P580. Cytochrome b-559, another necessary component of PSII, consists of two smaller polypeptides known as the a and B polypeptides, and has been proposed to protect PSII from photooxidative damage. The seventh major polypeptide of PSII has a molecular weight of 33 kDa and is located extrinsically to PSII. It is referred to as the manganese stabilizing protein (MSP) because it acts to protect the (Mn). complex from exogenous reductants and optimizes the complex’s catalytic efficiency. Two other extrinsic polypeptides are associated with PSII and have molecular weights of 17 and 23 kDa. They are implicated in the efficient production of oxygen by acting as barriers that prevent Ca2+ from migrating out of the site. Besides the major polypeptides described above, there are several small peptides associated with functional PSII complexes whose exact roles are unknown. 1.2.2 S-state Cycle in Water Oxidation One of the first ground-breaking experiments in understanding water oxidation was performed in 1969 by Joliot and coworkers who studied oxygen yields from dark-adapted chloroplasts as a function of single-turnover flashes (4). This work, shown in Figure 1-3, demonstrated that after the third flash and every fourth flash thereafter, a maximum in oxygen yield occurs. This pattern of oxygen release indicated that a total of four sequential light-induced turnovers of PSII are required to form molecular oxygen. Shortly after, Kok and coworkers interpreted these results by formulating a kinetic model where each stable intermediate produced by a flash of light is called an S-state (5). Each S-state is designated 6 Relative 02 Yield vs. Flash Number i 4 4. O . 12 3- 8 >- - , C >- ~ I T T I l 2 216 81D12141618 Flash Number Independent PSII Units: 4'0" mp hv hv h v hv SO—TSl—T 82—> S3"> SK ‘T l 02+ 4H+ Figure 1-3 A period-four oscillation of oxygen evolution as a function of flash number in spinach chloroplasts. This figure was kindly provided by Gerald T. Babcock. .2 Relative 02 Yield vs. Flash Number T I l I I 5316 81D12141618 Flash Number Independent PSII Units: 49" 2H20 hv hv h v hv sO—+s,—+ s —+ s3—>s‘< 2 02 + 4H+ Figure 1-3 A period-four oscillation of oxygen evolution as a function of flash number in spinach chloroplasts. This figure was kindly provided by Gerald T. Babcock. by Sn, where n is the number of oxidizing equivalents stored in the OEC. The 81 state predominates in dark-adapted samples and the 8.; state is a transient intermediate that spontaneously reverts to the 80 state with the concomitant release of 02. 1.2.3 The Energetics of Water Oxidation Water is an extremely stable molecule as evidenced by the fact that it covers over 75% of the earth’s surface. Thus, in order to oxidize water to molecular oxygen very high reduction potentials are required. P586, a product of the initial charge-separation reaction of PSII, is one of the strongest oxidants known in biology (6) and is sufficient enough to oxidize water (7). Figure 1-4 shows the reduction potentials of the major components of the donor and acceptor sides of PSII. A total of 3.72 V must be accumulated by the OEC in order to oxidize water (8). With a reduction potential of about 1.12 V, P530+ oxidizes Y2, which has a reduction potential of about 0.97 V. Therefore, about 4 V of energy is supplied by PSII to oxidize water, which leaves little surplus energy available to support any potential fluctuations as the OEC proceeds through the S-state cycle. This weak coupling of the cofactors leads to the highly efficient conversion of about two thirds of the photon energy into chemical potential (9-11). 1.2.4 The Manganese Cluster The PSII complex contains four manganese atoms per P530. From extended X-ray absorption fine structure (EXAFS) measurements, a model was 8 - 1.0— - 0.8—- Pm" \. T PhCO — 0.6 — - 0.4 —- E - 0.2— CIA "a .5 O __ QBHZ 3.3 hv a? 0 2 _ Ea g 0.4 — °‘ 0.6 - r a 0.8 ~— HZO“ 32/3,,33/52 Y 1.0 " 80/83 J Z \ 1.2— K ”38%de 02 + proton gradient Figure 1-4 Redox potentials of the cofactors of PSII. suggested for the local structure of the Mn complex. The model was proposed to consist of a pair of p-oxo bridged dimers with Mn-Mn distances of 2.7 A linked together by a mono-,u-oxo bridge with a Mn-Mn separation of 3.3 A arranged in a C-shaped structure (12, 13). There are two well-established low temperature EPR signals refered to as the multiline and the g = 4.1 signals that have been attributed to the 82 state of the (Mn). cluster (14, 15). The multiline signal, shown in Figure 1-5 (b), displays 18-20 partially resolved hyperfine lines and has a linewidth of about 1500-1800 G. The signal is generated by illumination of 02- evolving PSII samples at 160-200 K (Dismukes, 1980 #211; Hansson, 1982 #180; Brudvig, 1983 #179). Two other forms of the multiline signal can be generated by either Cay-depletion (16-18) or ammonia treatment (19-21) of PSII samples. Both treatments affect the number and splittings of the hyperfine lines, and have therefore been proposed to directly affect the structure of the (Mn)4 cluster. The Q = 4.1 signal, shown in Figure 1-5 (a), is due to a high-spin, ground state of the (Mn); cluster (22) and is generated by illumination of PSII samples at 140 K in the presence of high concentrations of cryoprotectant (15, 23). This signal has a linewidth of 320-360 G and lacks resolved hyperfine structure. Upon warming in the dark, the g = 4.1 signal converts to the multiline signal. Inhibition of PSII samples by incubation with acetate, which causes perturbations in the Ca2+ and Cl' sites, traps the OEC in a state in which the EPR spectrum from Y; is split by a magnetic interaction with the manganese 32 state. This split EPR signal was first discovered by Rutherford and co-workers (24) and 10 dX"/d H a) b) I r 3000 4000 Field (G) I 1000 2000 Figure 1-5 The 9 = 4.1 (a) and multiline (b) EPR signals observed in PSII membranes. This spectrum was kindly provided by Pierre Dorlet. 11 was later identified as coming from the SzYz' state (17, 25, 26). This signal has more recently been simulated by the Babcock (27) Britt (28), and Brudvig (29) groups in order to determine the distance between (Mn). and Yz°. From the results of the simulation studies, it was estimated that (Mn). and Y; are separated by a center-to-center, point-dipolar distance of 8-9 A. The oxidation state of each Mn atom in the cluster has been investigated by using X-ray absorbance near-edge spectroscopy (XANES). This technique has been used to detect oxidation state changes of many transition-metal compounds. The results of the study performed on PSII showed significant shifts in the XANES edge positions on going from Soto S1 and from S1 to $2 and have been interpreted as manganese-centered oxidations (30-32). Oxidation of the manganese cluster on the $2 to $3 transition is more controversial with two groups claiming a significant edge shift (30, 32) and one group reporting only a minimal shift in the edge position upon this transition (31). From XANES studies on model compounds and PSII, the oxidation states of the four manganese atoms are suggested to be: 80, II, III, IV, IV; S1, III, III, IV, IV; 82, III, IV, IV, IV; 83, IV, IV, IV, lV. 1.2.5 Y0 and Y2 There are two redox active tryosine residues found in PSII, YD and Y2, which are located in position 161 of the D2 and D1 polypeptides, respectively. The oxidized form of Y0, gives rise to a dark-stable electron paramagnetic resonance (EPR) signal known as Signal IIS, (5 = slow) due to its slow relaxation kinetics that are on the order of minutes (for reviews see (33)). It is a stable 12 neutral radical and is believed to be well shielded from solvent (26, 34). The tyrosine YD does not have a known role in the oxidation of water and will not be discussed further here. A second tyrosine residue, Yz', also gives rise to an EPR signal at room temperature in intact systems called Signal ”Vf (vf = very fast) with fast decay kinetics on the order of microseconds (35, 36). In contrast to Y0, Y2 has a very specific role in water oxidation by functioning as an intermediate cofactor between P630 and the manganese cluster. In light of experimental evidence and the role of tyrosine radicals in the general class of metalloradical enzymes (Frey, 1990 #215; Sigel, 1994 #216), it was suggested that Yz functions by using a hydrogen-atom abstraction mechanism (37, 38). The details of the mechanism for water oxidation will be discussed below. 1.2.6 A Model for Water Oxidation A metalloradical mechanism for water oxidation, proposed by the Babcock laboratory (39, 40), invokes Yz' as a hydrogen atom abstractor from manganese- bound water on each S-state transition. In this model, the two terminal manganese atoms are catalytically active and bind the two substrate water molecules in the So state. Another important feature of this model is charge neutrality. No charge is accumulated on the (Mn).; cluster, thereby avoiding the high energy situation of creating a charged species in regions of low dielectric. Figure 1-6 illustrates the important details of this mechanism for the 80 state. Following formation of P586, Yz is oxidized in the nanosecond time domain 13 Figure 1-6 Tyrosine Yz oxidation and reduction in the So state. The dashed arrows illustrate the flow of the electron and proton during Yz oxidation. The solid arrow shows abstraction of a hydrogen atom by Yz' during its reduction. 14 (41, 42) and simultaneously deprotonates to a nearby histidine residue (43-47), identified by mutagenic and kinetic data as D1-H190 (48-52). Subsequently, Yz° abstracts a hydrogen atom, on the microsecond to millisecond time scale, from a water molecule ligated to the (Mn), cluster. This sequence of steps is proposed to occur for each S-state transition as shown in Figure 1-7. Functions for the Ca2+ and CI' ions are also proposed in this model (38). The CI' ion binds to a Ca2+ atom that is carboxylate-bridged to the (Mn). cluster. After the formation of S1Yz', Cl‘ migrates to the upper catalytic Mn atom, bringing with it a proton from the lower position. This ensures that an abstractable hydrogen atom is available near Yz' in the higher S-states. Upon hydrogen atom abstraction in the 82 -) $3 transition, the oxo/hydroxyl species is formed. On the S3 -) 8.; transition, Yz' abstracts the final hydrogen atom in concert with the formation of an oxygen-oxygen bond and the reduction by one electron of the (Mn). cluster. (37). Molecular oxygen is released as product and then Cl' migrates back to the Ca2+ site and two water molecules bind to the (Mn)4 cluster to regenerate the 80 state. This model maintains overall cluster electroneutrality and minimizes nuclear motion, which are both important in terms of the low driving force available to oxidize water and the high turnover numbers (up to 200 electrons per second) that the OEC attains. 1.2.7 The Reduction Rates of Y2' Upon each S—state transition, Y; is formed and is subsequently reduced by the Mn cluster. Central to understanding the S-state mechanism is the kinetics of Yz' reduction. The rate of Yz° reduction for each S-state transition was 15 H . 0., . ., \OIIOIe.’ O .(. ., 3. . oo o o This figure was kindly Figure 1-7 A model for the S-state cycle in PSII. provided by Gerald T. Babcock. 16 first measured by Babcock in 1976 by using time-resolved electron paramagnetic resonance spectroscopy (EPR) (53) and more recently by Razeghifard (54). These rates have also been measured by optical absorption spectroscopic techniques in the UV region (Dekker, 1984 #212; Renger, 1986 #213; Rappaport, 1994 #191). The optical approach provides better signal to noise, but requires the use of spectral deconvolution techniques to avoid kinetic interferences from the acceptor side of PSII. Time-resolved EPR, on the other hand, is more convenient since it does not require spectral deconvolution, but is more time consuming because it requires extensive signal averaging to attain adequate signal to noise. In both cases, the decay rates of Y; were shown to occur on the us to ms time-scale depending upon the oxidation state of the (Mn). cluster. The rate of Yz' reduction by the (Mn)4 cluster has also been studied in PSII samples where the OEC has been chemically, biochemically, or genetically modified. The results of these studies are reported in Table 1 and show a retarding effect on some or all of the S-state transitions. For example, the depletion/repletion of Ca2+ (55' 56’ or the removal of the 33 kDa protein (57) has been shown to slow the 839(84)-)So transition. In addition, Razeghifard et al. showed that Y; reduction was slowed to approximately 6 ms on the 839(84)-)So transition in an engineered strain of cyanobacteria lacking the Mn-stabilizing protein (58). All of the above conditions led to a slowdown of the 02 releasing step in the S-state cycle. 17 - n 4 ‘-- I n . Table 1-1 Half-times for S-state Advance in Various Enzyme Preparation Types - Taken from (58-61) Sample Type Yz'S1-)Y282 Yz'Sz-9Y283 Yz'S3-)YzSo Thylakoids 65-86 as 140-245 us 750 us;1.3 ms PSII membranes 30-70 us 55-110 us 1.2-1.4 ms PSII core particles 75-95 us 225-380 us 4.1-4.6 ms D61N 240 us 520 us n.d. -MSP n.d. n.d. 6.0 ms Cl“ n.d. n.d. 3.3 ms Br" n.d. n.d. 4.9 ms N03” n.d. n.d. 15.6 ms N02“ n.d. n.d. 8.2 ms I‘ n.d. n.d. 11.7 ms 18 Ir bl Evidence for perturbations affecting the earlier S-state transitions was also found. For example, the rates of all the S-state transitions are dependent upon the degree to which the system is perturbed in the preparation of PSII membranes. Razeghifard et al. has shown that the treatment of thylakoids to prepare PSII membranes leads to retardation of all S-state transitions, with the largest effect seen on the last transition (54, 58). The mutation work done by Hundelt et al. on D1-Asp61 also demonstrated slower Yz' reduction rates (61). When this aspartate residue was mutated to the uncharged species, asparagine, oxygen evolution was retained, but the S-state transitions were retarded. The oxygen- evolving step was slowed by a factor of 9-10. So not only is the 83-)(84)-)So transition affected by sample perturbations, but the earlier S-state transitions can be affected as well. 1.2.8 Calcium In the early 19803 it was shown that Ca2+ is needed for efficient oxygen- evolution (62). When the 17 and 23 kDa polypeptides were removed from PSII by using a high salt treatment, a substantial amount of oxygen-evolution activity was lost. Successful reactivation of oxygen-evolution activity did not occur upon reconstitution of these polypeptides, but also required long incubation in the presence of excess Ca2+. These experiments ultimately led to the conclusion that the 17 and 23 kDa polypeptides were needed in order to retain Ca2+ in its native binding site, and without these polypeptides, excess concentrations of Ca2+ were needed in order to support high rates of oxygen-evolution. Without 19 C3” the system is unable to advance past the Sz'Yz' state, and more than 90- 95°/o of oxygen-evolution activity is suppressed. A study involving the examination of the ability of other metal atoms to replace Ca2+ and concurrently support efficient oxygen-evolution has shown that only Sr2+ can effectively perform this duty, reactivating about 40% of the original activity (62). Substitution of Sr” for Ca2+ results in a modified multiline EPR signal that has narrower hyperfine line splittings (63, 64). This result led to the proposal that the Ca2+ binding site is close to the Mn cluster and the substitution of Sr2+ for Ca2+ caused perturbations near the Mn cluster due to its larger ionic radius (1.13 A, compared to 1.0 A for Ca”). Additional evidence to support this proposal has come from EXAFS studies (65) and FTIR results showing a Ca2+ that is carboxylate-bridged to Mn (66). Although it is generally agreed upon that Ca2+ is required for efficient oxygen evolution, it is still not known exactly how Ca2+ participates in the water oxidation chemistry. Some possible roles for Ca2+ have been proposed and are discussed in the following section. 1.2.9 Proposed Roles for Ca2+ in Water Oxidation Although a general consensus for the function of Ca2+ has not been reached, several recently proposed models assign a possible role for Ca”. In general, these roles for Ca” fall into two categories. In one case, direct involvement of Ca2+ in the formation of the oxygen molecule has been suggested (67-69). For the second, a more general role for Ca2+ involves assisting in the catalytic efficiency of the OEC (37, 38). These proposed roles will be discussed along with an experiment devised to test these roles in the function of PSII. 20 In the first case, the removal of Ca2+ would be expected to deter the formation of the O-O bond on the 83-)(84)-)So transition. Pecoraro et al. and Limburg et al. postulated that Ca2+ presents a hydroxide nucleophile close enough to a putative Mn"=O species to form the critical O-O bond in the 84 state (67, 70). Siegbahn and Crabtree have also employed Ca2+ in the O-O bond forming step to provide a solution to the energetic difficulty of forming the reactive oxyl radical (68). The proposed five-coordinate, Mn-oxo complex formed in S3 is unreactive and incapable of forming an O-O bond until it chelates with the calcium complex to become six-coordinate. This allows the reactive Mn'V-O' complex to form, which subsequently leads to O-O bond formation. In these models, O-O bond formation cannot occur without Ca2+ in the site, but the cycle of events prior to this step is not expected to be influenced significantly by the depletion of this cofactor. This is especially so if some or all S-state transitions correspond to pure electron tunneling (71), as there appears to be consensus that little structural alteration accompanies Ca2+ replacement by Srz+ in the OEC (13, 65, 72). Likewise, there is no indication of a dramatic change in the driving force for S—state advance when cofactor substitution is carried out. If Sr2+ substitution did alter potentials significantly, oxygen evolution would not be possible since the potential span between P680*/P580 and water is only about 0.2 V. For the second class of models, proton-coupled electron transfer provides the mechanistic basis for each S-state advance (37, 38). Here, small structural changes will have significant effects on kinetics owing to the much greater 21 sensitivity of the proton potential surface to distance (7, 73). Thus, although Ca2+ is necessary for only the higher S-state transitions (38), slight perturbations to its geometry or replacement by 8er may affect all S-state transitions. This model places Ca2+ in close proximity to the Mn cluster but does not directly invoke it for O-O bond formation (38). Calcium’s function, here, is to act as a docking site for Cl‘, so that Cl' diffusion into the site does not become rate limiting when it is utilized for charge migration on the S1 to $2 transition, (38). The oxygen-oxygen bond is formed, in this model, by the concerted reaction between Yz', Mn'V=O, and Mn'V-OH (37). Siegbahn’s concerns about the energetics of Mn'V-O' formation disappear in the Hoganson et al. model, as an “energetic boost” is provided by the concerted O-O bond formation event. Calcium is important for efficient 02 production, but does not play a central role in product formation. Strontium replacement for Ca2+ provides an excellent probe to address these issues. For the first class of models, we would expect the substitution of Sr2+ for Ca2+ to hinder the O-O bond forming step, thereby causing a pronounced effect on the reduction kinetics of Y; on the 83-)(84)-)So transition. While for the second class of models, a retardation of all S—state transitions is expected because, in this model, Ca” plays more of a general role in water oxidation. 1.3 Project Goals The goal of my thesis research was to investigate the mechanistic predictions of the above models in regards to the role of Ca2+ in efficient 02- evolution, and to subsequently gain more information on the hydrogen atom abstraction mechanism. 22 U1 Ci tr. IE th ei E) To investigate the role of Ca2+, the reduction rate of Y; was monitored by using time-resolved EPR spectroscopy for each S-state transition, S1982, 82-)83, and 83—)(84)-9So, to determine if each transition was affected by Sr2+ substitution or if only the oxygen evolving step was retarded. If the role of Ca2+ is to assist in the formation of the O-O bond that ultimately leads to the release of molecular oxygen, then its depletion will hinder the formation of this bond. We expect, then, that only the rate of the 839(84)-)So transition would be significantly altered, particularly if some, or all, of the earlier S-state transitions involve pure electron transfer. On the other hand, if Ca2+ does not play a direct role in bond formation but rather is integral to the active site and plays another catalytic role by, for example, binding Cl', then its replacement by Sr”, which affects the integrity of the OEC only slightly, is likely to affect all S-state transitions. The second part of my research was to use the results obtained for the reduction of Y; in Sr2+-substituted PSII membranes to investigate the hydrogen atom abstraction mechanism proposed for water oxidation. If reduction of Y; takes place by a pure electron transfer process, any slight perturbations to the OEC caused by the substitution of Sr2+ would have very minute effects, if any, on the rate of the electron transfer process. But, if instead, both a proton and an electron are transferred as a hydrogen atom during Yz' reduction, then we would expect to see much larger rate alterations, due to the greater dependence of this type of process on donor and acceptor distance and orientation. Thus, the 23 results of the time-resolved EPR experiments obtained for the Ser-substituted PSII samples can lend some insight into the mechanism of Yz' reduction. 24 k\ k! T) SI Chapter 2 2 Electron Paramagnetic Resonance Spectroscopy 2.1 Basic Principles Electron paramagnetic resonance (EPR) is based upon the interaction of radiation of microwave frequency with magnetic moments arising from unpaired electrons. Therefore, the application of EPR spectroscopy is limited to paramagnetic systems. Information regarding the properties of the paramagnetic system is gained by measuring the attenuation versus frequency of a beam of electromagnetic radiation as it passes through the sample. The lines in the spectrum represent transitions between energy levels of the radical species, and their frequency values are a measure of the energy separation of the two levels. The identity of the radical as well as the nature of its surroundings can be gained from an analysis of the pattern and shape of the lines obtained. Electromagnetic radiation is made up of coupled electric (E1) and magnetic (B1) fields perpendicular to the direction of propagation, oscillating at some frequency value. Electromagetic radiation can also be represented as a stream of particles called photons, which have an energy value equal to hv. Different energy states arise from the interaction of the unpaired electron spin moment (n1s = i 1/2) with the magnetic field component, B1. This phenomenon, called the Zeeman effect, lifts the degeneracy of the energy levels for the m8 = +‘/2 and —‘/2, or as commonly called the or and [3 states, respectively. The splitting of the energy levels can be varied by changing the static magnetic field, B; absorption occurs when the energy (hv) of the microwave radiation incident on 25 the Fig the wh me m0 Spe the sample, equals the splitting of the or and [5 energy levels of the molecule (see Figure 2-1). This energy value is then used to calculate the Zeeman 9 factor of the radical species giving rise to the observed transition according to the equafion: h V = 916193 (1) where h = Plank constant v = resonant frequency ,6. = Bohr magneton B = static magnetic field The 9.3 factor for a free-electron is equal to 2.0023. Deviation of the measured 9 factor from this free-electron value arises from orbital angular momentum effects and can be used to aid in the identification of the radical species giving rise to the signal. In addition to the interaction of an electron with an external magnetic field, there is another interaction called nuclear hyperfine coupling. The electron experiences a second field generated by magnetic nuclei in its vicinity. These local fields 3.003. add vectorially to the external field to produce the total fieldBerr at the electron being probed. The neighboring nuclei may possess an intrinsic spin angular momentum with possible nuclear spin quantum number 1 values of 0, 1/2, 1, 3/2, 2, etc., with corresponding nuclear spin states given by 21+ 1. 26 E = + 1/2 gefieB B = O resonance field Br B ——+ Figure 2-1 Energy level scheme for a free electron as a function of applied magnetic field B. Absorption occurs when the incident radiant energy equals the energy-level separation. 27 Th SP hr 50 of The result of electron interaction with neighboring nuclear-dipole moments is a splitting of the resonance, referred to as hyperfine splitting. The nuclear hyperfine interaction splits each of the electron Zeeman energy levels into 21 + 1 sublevels. Hyperfine interactions assist in the study of paramagnetic species by offering more detailed information in addition to g-factor information gained from interaction of the electron with the external magnetic field. 2.2 The EPR Spectrometer A block diagram of a typical EPR spectrometer is shown in Figure 2-2. The source of microwave radiation is usually a klystron and the frequency typically used in most EPR spectrometers is approximately 9.5 GHz. The klystron is a vacuum tube that can produce microwave oscillations centered on small frequency ranges. The frequency of the radiation is determined by mechanical tuning of the klystron cavity and by the voltage applied to the klystron. Microwaves are transmitted from the klystron to the cavity through waveguides of a particular size. The most common is the “X band” which uses 12.7 x 25.4 mm OD rectangular brass pipe for the frequency range of 8.2-10.9 GHz. The frequency of the source is tuned to the appropriate resonant frequency of the cavity, which is determined by the dimensions of the cavity. In order to obtain the highest EPR signal sensitivity, the cavity should hold the sample at a maximum of B1 (the magnetic component of the microwave radiation) and at a minimum of E1 (the electric component of the of the microwave radiation), and have B1 perpendicular to the static magnetic field B. The source of the static magnetic field B is an electromagnet and the magnitude of B is measured and 28 Field Modulator Klystron lsolator Directiona ‘ Attenuator . coupler Resonant .:':::::3:‘: - «'g;‘77:?f.§2§=?5£4=“ *i ""3155; t O f, Amplifier AO’ J cavity ‘v 5:3: I! T [Will Computer .1153 "’ Scope . . as; Electromagnet Recorder Monitoring Detector ,3,- microwave frequency and power ,_ Figure 2-2 Block diagram for a continuous wave EPR spectrometer. Adapted from Electron Paramagnetic Resonance, J.A. Weil, J.R. Bolton, and J.E. Werlz, 1994, John Wiley 8. Sons. Reprinted by permission of John Wiley 8. Sons, Inc. 29 controlled by a Hall-effect detector. Stability in the static magnetic field B is important in order to get accurate measures of the energy level separation AE. The microwave radiation that is produced at the klystron travels through the isolator, which prevents any microwave energy from reflecting back to the klystron, to the attenuator where the microwave power incident on the sample is adjusted. The circulator directs the microwave power to the cavity and simultaneously directs the power reflected from the cavity to the detector. Most often, the frequency of the microwave power is held constant while the magnetic field B is varied until the resonant condition is met. At resonance, energy is absorbed by the sample and a change in the power level reflected from the cavity is produced at the detector. Electron paramagnetic resonance spectroscopy employs a technique to increase the signal-to-noise ratio called phase sensitive detection where the magnetic field B is modulated at 100 kHz. Iy modulating B the noise-contributing components of frequencies close to or below the modulation frequency are limited. The amplitude of the detected signa! will be proportional to the slope of the absorption peak. The output signal appears as the first derivative of the absorption signal where the inflection points of the absorption curve have maximum amplitudes in the first derivative signal and at the resonant field the amplitude is zero. For time-resolved EPR experiments, the magnetic field B is set to a resonant field value of the species under study, in this case Yz'. For these experiments, the field value chosen is the low-field peak, as indicated in Figure 2- 3 by the arrow, and the amplitude of the signal is monitored over time. This field 30 Figu field setting tortime-resolved experim ents I I ' I ' I ' I ' I ' I 3300 3320 3340 3360 3380 3400 Field (Gauss) Figure 2-3 Continuous-wave EPR spectrum of Yo'. 31 value minimizes interference from chlorophyll absorption (Babcock, #222). The principal g-values of the tyrosine radical in PSII have been measured by using high-field EPR spectroscopy and have been reported as gx=2.0074, gy=2.0045, and gz=2.0021. The maximum difference between 9)( and 92 corresponds to about nine Gauss at X-Band. The overall bandwidth of Y; is much greater than this as the result of hyperfine coupling (Babcock, 1997 #223). Thus, we are confident that setting the field at the low-field peak provides a homogenous sampling of all g-values. Using this technique, the formation and decay of Y; can be monitored as the S-states are cycled and its lifetime can be determined. 32 Chapter 3 3 Materials and Methods 3.1 Preparation of PSII Membranes Oxygen-evolving PSII membranes can be isolated by detergent solubilization of the chloroplast thylakoid membranes (74). These particles, called BBYs, consist of the PSII core with the Mn cluster, Ca2+, Cl', the extrinsic polypeptides, and the light harvesting complex, and are nearly devoid of PSI. Further refinement to make core particles with oxygen evolving capacity is possible, where most of the light harvesting complex is stripped away from the OEC (75). Both of these preparations have made possible significant advances in the study of the OEC due to the high rates of oxygen evolution activity that can be achieved. Photosystem ll membranes were prepared from fresh market spinach according to the method of Berthold et al. (74) with the modifications described in (76). The final product is capable of high rates of oxygen evolution and contains very little contamination from PSI. The procedure used in these experiments is as follows: 0 About 750 g of market spinach was deveined and washed. The sample was kept at 4° C throughout the isolation procedure. 0 The spinach was homogenized in a blender and filtered through a 6-layer cheesecloth. For every 300 g of spinach, 250 mL of grinding buffer containing 0.4 M NaCl, 2.0 mM MgCl2'6H20, 1.0 mM EDTA, and 2.0 g/L BSA, 33 20.0 mM N-2-hydroxyethylpiperazine-N-Z-ethanesulfonic acid (HEPES), pH 7.5 was used. The filtrate was centrifuged in a GSA rotor until 1,000 RPM was reached in order to remove large fragments. The supernate was centrifuged for 12 min. at 10,000 RPM. The pellet was resuspended by using a paint brush into the first resuspension buffer containing 0.15 M NaCl, 0.40 mM MgCl2'6H20, and 20.0 mM 2-(N- morpholino)ethanesulfonic acid (MES), pH 6.0 homogenized in a glass-teflon homogenizer, and centrifuged for 12 min. at 10,000 RPM. The pellet was resuspended by using a paint brush into the second resuspension buffer containing 15.0 mM NaCl, 0.50 mM MgCl2'6H20, 1.0 mM ascorbic acid and 50.0 mM MES, pH 6.0 and homogenized. The chlorophyll concentration was determined by measuring the absorbance values at 645 nm and 663 nm in 80% acetone and then adjusted to 2 mg/mL. A 25% Triton X-100 solution was added dropwise to the stirring sample in the dark to reach a detergent to Chl ratio of 30:1. After a 30-minute incubation period in complete darkness, the sample was centrifuged for 30 minutes at 20,000 RPM. The pellet was resuspended by using a paint brush into a buffer called SMN containing 15.0 mM NaCl, 0.40 M sucrose, and 50.0 mM MES, pH 6.0 homogenized, and centrifuged for 30 min. at 20,000 RPM. The pellet was resuspended by using a paint brush into SMN buffer to a final chl concentration of about 4 mg/mL. 34 The final PSII membrane sample was saved in 1 mL aliquots and stored at - 80° C. 3.2 Calcium Depletion of PSII Membranes Salt-washed membranes were prepared as in Boussac and Rutherford (63) as follows: PSII membranes were diluted to a chl concentration of 1 mg/mL and incubated in room light for 30 minutes at 4°C in a SMN buffer containing 0.30 M sucrose, 2.4 M NaCI, and 25 mM MES, pH 6.5 at a 1:1 ratio. Following incubation 50 pM EGTA was added. The sample was centrifuged for 20 min. at 20,000 RPM. The pellet was resuspended in a SMN buffer containing 0.30 M ultra-pure sucrose (Aristar, BDH Laboratory Supplies, United Kingdom), 30 mM NaCl, 50 pM EGTA, and 25 mM MES, pH 6.5. After homogenization, 50 uM EGTA was added and the sample was centrifuged for 20 min. at 20,000 RPM. The pellet was resuspended in a SMN buffer containing 0.30 M ultra-pure sucrose, 15 mM NaCl, and 25 mM MES, pH 6.5, to a final chl concentration of 2 mg/mL. 3.3 Sample Preparation for EPR Experiments Reconstitution of the Cay-depleted samples was carried out by adding CaClz or SrClz to a final concentration of 15 mM. Prior to flash illumination 100 pM PPBQ in DMSO and 1 mM ferricyanide were added to the sample. For the pH studies, the salt-washed PSII membranes were resuspended in the same 35 buffer, but at a pH of 5.9, 6.5 or 7.1. For the istope exchange studies, salt- washed PSII membranes were exchanged into a MN buffer containing 15 mM NaCI, 25 mM MES (pD 6.5) in 99.9% D2O with a final addition of 15 mM SrCl2. The preparation of the sample was identical to that above, except that the final washing steps were performed in a D2O-MN buffer. The rate of isotope exchange has been shown to occur in a matter of minutes or less, well within the time it takes to carry out the final washing steps (77-80). 3.4 Time-Resolved EPR Experiment Setup EPR spectroscopy was performed at room temperature with a Bruker ESP 300E spectrometer equipped with a TM011 cavity. The spectrometer was tuned to the low-field peak of the stable YD radical (D2-Y160) and the kinetic traces obtained for Y2 oxidation and its subsequent reduction were recorded at this field value. The experiments under steady state flashing light conditions were carried out by using a 350 pL capacity suprasil quartz flat cell (Wilmad, Buena, NJ). Each sample received 100 flashes before being replaced by a fresh sample. For the S-state resolved studies, a home-built, continuous-flow system that is depicted in Figure 3-1 was used to provide fresh samples to the flat cell mounted in the cavity. Each sample received a train of three flashes before being replaced. A pump (Gilson, lnc., Middleton, W!) was used to control the flow rate so that a fresh, dark-adapted sample was supplied for each flash sequence. Following flash illumination, the sample was recycled to a reservoir that was held at 4 °C and kept in absolute darkness for 15 minutes. Each experiment was completed in approximately five hours with about 2000 scans averaged. 36 HI—— flat cell ~1 r ‘— EPR cavity pump sample reservoir Figure 3-1 Setup for the time-resolved EPR experiments. Flowing the sample allows the formation and decay of Y2 to be monitored for each 8- state transition. Sample averaging is possible by recycling the enzyme. 37 Photoexcitation of the sample at 532 nm was accomplished by using a Quanta-Ray DCR-11 pulsed Nd-YAG laser (Spectra-Physics, Mountain View, CA). A personal computer equipped with a Metrabyte WAAG board and a Metrabyte CTM-05 timing board was used to trigger both the EPR spectrometer and the laser (81). The EPR conditions for each experiment are given in the figure legends. A least-squares fitting routine (Origin 5.0, Microcal Software, Inc., Northampton, MA) was used to analyze the kinetic traces. Oxygen evolution activity was measured with a Clark type electrode under continuous illumination with saturating white light from two projector lamps of 250 W each that are filtered through a CuSO4 solution. Oxygen evolution rates for intact BBYs measured at a concentration of 1 mg/mL chl in SMN buffer, pH 6.0 were typically between 650-750 umol O2/mg chl'h while salt-washed, Ca2+ and Sr” reconstituted samples measured at a concentration of 10 pg/mL chl in SMN buffer, pH 6.5 were typically 450-550 and 200-300 pmol O2/mg Chl'h, respectively. Ferricyanide and DCBQ, at concentrations of 600 pM and 300 pM, respectively, were used as the electron acceptors for these measurements. 38 Chapter 4 4 Results of Time-Resolved EPR Experiments 4.1 S-State Mixed Kinetics The formation and decay of Y2° was monitored under steady state flashing light conditions by setting the magnetic field to the low-field peak of the stable YD radical (see Figure 2-3) and inducing the oxidation and reduction of Y2 by using 532 nm laser flashes. Figure 4-1 shows the kinetic EPR traces of the decay of Y; in salt-washed PSII membranes reconstituted with either Ca2+ (trace a) or Sr2+ (trace b). The number of spins was determined by comparing the amplitudes of each kinetic trace to one obtained under the same experimental conditions for a tris-washed sample, which generates one Yz' spin per reaction center (82). Using this procedure, we estimate that 0.5 spin per reaction center for the Ca2+- reconstituted sample was produced on each flash, which is in agreement with values reported by this laboratory (81) and reports on intact PSII membranes by other labs (36, 58). The Sr2+-reconstituted sample, however, showed a 20% increase in spins due to the slowdown of Y2' reduction in the Sig->81 transition, which is sufficiently slowed in the Sr2+ sample that its detection is now possible (see below). The lower apparent Yz' spin count in O2-evolving samples relative to tris-washed material may be due to broadening of its EPR signal by the (Mn). cluster. In Cay-depleted samples, Y; is in close association with Mn, which broadens its EPR signal at lower temperatures (25, 27, 83). Similar effects could 39 Mixed S-states a) CaCI2 b) SrClQ Signal Amplitude |5msl Time Figure 4-1 Time-resolved EPR steady state kinetic traces (S-states are mixed) for Y,’ reduction in salt-washed PSII membranes reconstituted with (a) CaCl2 or (b) SrCl2. Experimental conditions: Time constant 40 us, modulation amplitude 4 G, gain 1 x 105, microwave power 3 dB, frequency 9.78 GHz, magnetic field locked at low field peak (3470 6), room temperature, flash frequency 0.66 Hz. 2000 events were averaged. 40 operate in O2-evolving samples and cause an apparent decrease in spin concentration. Each kinetic trace in Figure 4-1 shows a biphasic decay. For the Ca”- reconstituted sample, the fast phase, which is approximately 50% of the total amplitude, is attributed to the reactions S1Yz' —)S2Yz and S2Y2' -)S3Yz, while the slow phase, which is approximately 35% of the total amplitude, is produced from the Sng' -)(S4)-)SOY2 reaction. The remaining 15% of the total amplitude is attributed to the decay of Y; in damaged centers that behave like tris-washed PSII membranes. A laser flash artifact is detected on each flash and does not completely cancel with subtraction of an off-resonance trace obtained at 3400 G. This should not greatly interfere with the rate determination of the decaying phases of the kinetic traces. The Ca2+-reconstituted sample (trace a) shows half- times of ~300 us and ~4 ms for the fast and slow phase, respectively. The Sr”- reconstituted sample (trace b) has a greater amplitude due to contributions to the fast phase by So-)S1. The half-times are ~900 us and ~15ms for the fast and slow phase, respectively. Comparisons of the two kinetic traces shown in Figure 4-1 show that Sr2+-reconstitution apparently slows the rate of Y2' reduction on all S-state transitions and also shows an increase in the signal amplitude due to the additional detection of So-)S1. 4.2 S-State Resolved Kinetics In order to study the effects of SET-reconstitution on the rates of Y2“ reduction in more detail, we resolved the S—state transitions individually as demonstrated in Figure 4-2. To perform this experiment, a sample that was 41 a)CaCb WM o 'o 3 a E < b) SrCl2 E a SnYz. H Sntle .0.) \Yz —’ p680. U s,—>s2 s,—>s3 s,—>s,, T T T Time flash 1 flash 2 flash 3 I———‘I 50 ms Figure 4-2 Three-flash S-state resolved EPR kinetic traces. Flash one induces the formation of Yz'. Decay of the signal represents the reduction of Y,’ by the (Mn). cluster. 42 poised in the 81 state by dark adaptation was flowed into the flat cell and given three saturating flashes. The flow rate was adjusted so that each sample received three flashes spaced 100 ms apart before it was replaced by the next dark-adapted aliquot. The first flash initiated the transition S1Y2' -)S2Y2, while the second flash generated primarily S2Yz' 98ng. The final flash caused Yz' reduction on the Sng' ->(S4)->SOY2 transition, with a small contribution from the other states that is caused by double hits and misses. Figure 4-3 shows kinetic traces and best fits for the first flash given to salt-washed PSII membranes reconstituted with either Ca2+ (trace a) or Srz+ (trace b). Both traces show biphasic behavior with the fast phase being attributed to the decay of Yz° upon the S1Y2' --)S2Yz transition and the slow phase to reduction of Y2' in damaged centers. The reduction of Y; on the first flash for the Ca2+-reconstituted sample (trace a) has a half-time of ~200 ps. The Sr2+-reconsitituted sample displayed slower reduction kinetics on the SrYz’ -)S2Yz transition (trace b) with a half-time of ~900 us. The spin counts for each of the above kinetic traces were 0.40 and 0.90 spins for the Cay-reconstituted and the SrZI-reconstituted samples, respectively. The greater number of spins from the Sr2*-reconstituted sample was in part due to the additional detection of the So-)Sr transition, to Yz sites in centers deficient in 02 evolution, and to YD' generation on the first flash. The remaining increase is most likely due to S-state dependent effects caused by magnetic interaction between the tyrosyl radical and the manganese cluster that produce small differences in the Y; lineshape. These later effects will not 43 1st Flash S1 8'2 (D is. .0 . 3 t E. E < 6 c .9 (I) 5ms l J l I Time Figure 4-3 Time-resolved EPR kinetic traces for the S1Y2' 9 S2Y2 transition triggered by the first flash in a three flash experiment. Salt-washed PSII membranes reconstituted with (a) CaCl2 or (b) SrCl2 were used. Experimental conditions: Time constant 20 us in (a) and 40 us in (b), modulation amplitude 4 G, gain 1 x 105, microwave power 3 dB, frequency 9.78 GHz, magnetic field locked at low field peak (3470 G), room temperature. 4000 events were averaged. 44 substantially alter the fast kinetics that are the primary focus of this study and will not be discussed further. The kinetic traces for the second transition are shown in Figure 4-4. The S-state advance, S2Yz' 98%, showed a slow down in the SrZI-reconstituted sample that was similar to that of the first flash. The half-times calculated for the Cay-reconstituted and Sr2+-reconstituted samples were ~450 us and ~1.3 ms, respectively. The spin count, 0.50, is nearly the same for both sample types. Figure 4-5 shows the kinetic traces for Ca2+- and Sr2+-reconstituted samples on the S3Y2'9(S4)9SOY2 transition. Due to mixing of the S—state transitions by the third flash, the decays also had a small contribution from the earlier and faster 8- state transitions. In addition, the amplitude of these traces increased, compared to the second flash, to a spin count of about 0.55. The half-times were ~5.0 ms and ~18 ms for Ca2*- and Sr2*-reconstituted samples, respectively. The results of the steady-state and S-state resolved experiments are presented in Table 4-1. 4.3 Effects of pH and Isotope Exchange The effects of pH on the reduction rates of Y; in the salt-washed, Sr”- reconstituted samples were investigated to determine if a pH optimum for Y2° reduction becomes apparent in any of the S-state transitions upon metal substitution. The effects of pH were studied by carrying out the S-state resolved, three-flash experiment at pH values of 5.9, 6.5, and 7.1 in Ser-reconstituted samples. The results of this experiment, shown in Figure 4-6, indicate that there is no substantial effect of pH on the kinetics of Y2' reduction by the Mn cluster in the 5.9 to 7.1 range. A similar lack of pH effects in the physiological range have 45 2nd Flash 82-» S3 0) 'o :3 :0: E. E <1: 5 c F.” a) b) SrCI2 5 ms l l I I Time Figure 4-4 Time-resolved EPR kinetic traces for the S2Y2' 9 S2Yz transition triggered by the second flash in a three flash experiment. Salt-washed PSII membranes reconstituted with (a) CaCl2 or (b) SrCl2 were used. Experimental conditions: Time constant 20 us in (a) and 40 us in (b), modulation amplitude 4 G, gain 1 x 105, microwave power 3 dB, frequency 9.78 GHz, magnetic field locked at low field peak (3470 G), room temperature. 4000 events were averaged. 46 3rd Flash 83 80 a) 'o - fl 3 I E I o- . E <( E C .9 (D b) SrCI2 5ms I 1' Time ‘ Figure 4-5 Time-resolved EPR kinetic traces for the Sng' 9 $er2 transition triggered by the third flash in a three flash experiment. Salt-washed PSII membranes reconstituted with (a) CaCl2 or (b) SrCl2 were used. Experimental conditions: Time constant 20 us in (a) and 40 us in (b), modulation amplitude 4 G, gain 1 x 105, microwave power 3 dB, frequency 9.78 GHz, magnetic field locked at low field peak (3470 6), room temperature. 4000 events were averaged. 47 Table 4-1 Rates of Y2' Reduction for Ca23 or Sr”-Reconstituted PSII Membranes Determined by Time-Resolved EPR Spectroscopy. S-state spin .11/2 (slow) 31/2 (fast) Steady-state Ca2+ 0.50 4 ms (35%) 300 ps (50%) Sr2+ 0.60 15 ms (35%) 900 ps (50%) s,-> 32 Ga” 0.40 40 ms (15%) 200 ps (85%) Sr2+ 0.90 40 ms (15%) 900 ps (85%) 52 -> 53 Ca2+ 0.50 40 ms (15%) 450 ps (85%) Sr2+ 0.50 40 ms (15%) 1.3 ms (85%) $3 —> 80 Ga” 0.55 40 ms (15%) 5 ms (85%) s? 0.55 40 ms (15%) 18 ms (85%) * The rates here are reported as half-times to be consistent with the Photosystem ll literature. 48 0 pH 7.1 ‘O 3 E. E I , | I _ I .r g I! II ‘ I ‘ I) 1' ‘ . i I I. I. g), I ‘I t‘ .‘ Time a) pH 6.5 “O B '5. E < ‘I E I 'i“ I ‘i I!) ii Time pH 5.9 Signal Amplitude I 1 ‘ I—-'"-’l Time Figure 4-6 Time-resolved EPR kinetic traces for the S1Yz'9 S2Y2, S2Y2' 9 5ng, Sng' 9 $erz transitions in salt-washed, Sr2*-reconstituted PSII membranes resuspended in buffers at pH = 5.9, 6.5, 7.1. Experimental conditions: Time constant 40 us, modulation amplitude 4 G, gain 1 x 105, microwave power 3 dB, frequency 9.78 GHz, magnetic field locked at low field peak (3470 G), room temperature. 2000 events were averaged. 49 been reported earlier for the unperturbed, O2-evolving system (84). In addition, a hydrogen/deuterium isotope effect study was performed. Deuterium kinetic isotope effects on Y2“ reduction have been carried out in oxygen-evolving PSII membranes by using optical (60) and EPR (85) methods. These experiments were performed in light of the proposal that Y; reduction is accompanied by proton transfer from substrate water. This proposal suggests the possibility of significant kinetic H/D isotope effects, but the results of both the optical and EPR studies showed moderate deuterium kinetic isotope effects on the reduction of Yz'. The kH/ko values obtained range from 1.3 to 2.9, depending on the S-state transition, which suggests that the reduction of Y; is not severely limited by proton transfer in the intact system. Deuterium kinetic isotope effect studies on the reduction rate of Y2' in salt- washed, Sr2+-reconstituted PSII membrane fragments were conducted in this work in efforts to test suggestions that perturbations to the OEC, caused by Sr”- substitution, hinders proton transfer during Yz' reduction. The rate of proton transfer is dependent upon donor and acceptor distance to a much greater degree than electron transfer, and therefore is expected to rate limit Y2' and lead to large deuterium kinetic isotope effects (86). In order to test this theory, salt- washed PSII membrane fragments were exchanged into a MN buffer containing 15 mM NaCI, 25 mM MES (pD 6.5) in 99.9% D2O with a final addition of 15 mM SrCl2. The preparation of the sample was identical to that of the experiments performed above on samples in H2O, except that the final washing steps were performed in a D2O-MN buffer. The rate of isotope exchange has been shown to 50 occur in a matter of minutes or less, well within the time it takes to carry out the final washing steps (77-80). Analysis of the flash-induced kinetic traces that are shown in Figure 4-7 for a steady state experiment shows a kinetic isotope effect similar to that reported for the intact system (60, 85, 87) with a maximal value of about 1.4. 51 Signal Amplitude .3 . , u‘ . ' ‘ K v. 7‘ T. . A ‘.‘~. <3. If s. ‘- If .. t a - 'I'? '5' ..\.:l‘ :_ :' .‘ .' J ‘;.’ ..‘#{:."I ' ‘I" ‘ ' ' ('1' :-T',.."u.‘: I. n , 2 [:0 ' ' .r- T " 'Sj . Ii: .13; .' ‘Iv, ' - _| '1‘.‘ '. _“." ate 1., '.-\ ‘- - C Time Figure 4-7 Time-resolved EPR steady state kinetic traces (S-states are mixed) for Y,’ reduction in salt-washed PSII membranes reconstituted with SrCl2 in either D20 or H2O. The experimental conditions were the same as in Figure 4-2. 52 Chapter 5 5 Hydrogen Atom Abstraction 5.1 Electron Transfer The exact function of Ca2+ in efficient water oxidation by PSII is currently unknown and was the focus of the time-resolved EPR experiments presented in Chapter 4. The results of these experiments suggest that Ca2+ takes on a more generalized role that remains constant throughout the S-state cycle since the kinetics of Y2' reduction on each S-state transition is retarded by a comparable extent when Ca2+ is replaced by Sr”. Thus, the direct involvement of Ca2+ in the oxygen-oxygen bond forming step in water oxidation is unlikely. In addition, similar results have been presented where one or more of the S-state transitions are slowed upon point mutation or biochemical treatment of PSII (Table 1-1). Therefore, in order to show by what mechanism the reduction of Y2' is altered upon Sr2+-substitution in PSII membranes, we must also take into account the results shown in Table 1-1. To do this, we look at how electron transfer reactions occur, and specifically, how the rates of electron transfer between a donor and acceptor can be affected. Electron transfer reactions and their rate characteristics have been extensively studied (for a review see (88)). Two parts, electronic and nuclear, determine the rate of electron tunneling. The electronic part arises from the dependence on the coupling strength between the electron donor and acceptor wavefunctions; as the distance, R, between donor and acceptor increases, the rate decreases exponentially owing to the uncoupling of the wavefunctions. The 53 nuclear part depends on the driving force, AG, and the reorganization energy, 1.. The driving force term is the ground state free energy difference between the donor and acceptor while the reorganization energy is the amount of energy needed to reorganize the nuclear coordinates for electron tunneling. These parameters are defined pictorially in Figure 5-1. A kinetic expression for the estimation of rates of intraprotein, non-adiabatic electron tunneling has been formulated by Page et al. (71) as follows: log k.t = 13 — ((1.2 — 0.8p)(R - 3.8)) - 3.1 ((AG + 102/1] where ket represents the exergonic electron tunneling rate (s'1), p is the specific packing density of the inter-cofactor protein volume, R is the edge-to-edge distance between the donor and acceptor (A), AG is the driving force (eV), and I. is the reorganizational energy (eV). We can use this expression to predict the electron transfer rate for the reduction of Y2' by the Mn cluster to gain a better understanding of the mechanism of this process. In order to use this kinetic expression for the electron transfer between the Mn cluster and Yz', we use values to substitute into the equation as follows. The dipolar distance between the Mn cluster and Y2' has been reported as approximately 8 A from magnetic-resonance studies (27-29, 83, 89). The driving force for this reaction is approximately 60 meV (90) and the reorganizational energy is taken to be about 0.8 eV, which is a reasonable estimate for a protein environment (91-93). The packing density is determined by the van der Waals 54 Reaction Path Figure 5-1 Potential energy surfaces of the donor and acceptor. AG is the free energy difference between donor and acceptor, 7. is the reorganization energy, and R is the distance between donor and acceptor. 55 radii of the atoms between the donor and acceptor molecules and has an average value of about 0.75 in the proteins analyzed by Page, et al (71). When these values are substituted into the equation, rates on the ns time scale are predicted, which are orders of magnitude faster than what has been experimentally observed for Y; reduction in O2-evolving systems. Therefore, we conclude that the rate limiting step in the overall 8,, 9 Sn.) process is not the tunneling of the electron (71 ). As Y; is reduced it becomes protonated (94, 95). Thus, a more precise approach to studying this process is within the context of a proton coupled electron transfer (PCET) reaction. We expect that this approach will give additional insight into the rates that we and others (54, 58, 60, 87) have measured. 5.2 Proton Coupled Electron Transfer: Sequential or Concerted? For PCET reactions, three possible mechanistic pathways can be considered, as shown in Figure 5-2. In order to test the viability of each pathway, the energetics of each must be taken into account. The values that were used to determine the energetics of paths a and b are listed in Table 5-1. The first path involves the transfer of the electron, followed sequentially by the transfer of the proton, ET/PT (path a). For the process under investigation here, the transfer of an electron results in the reduction of Yz° to form the tyrosinate (state II). The reduction potential of Y'IY‘ is about 0.68 V aqueous solution (a = 80) (96) and the potential required to oxidize water is about 0.93 V (8); therefore the transfer of an electron from substrate water/Mn cluster to Yz' is an uphill process with a barrier 56 H+ w (ll)YZO MnTHZO (F0) . (IV) YZOH Mn+OH “- (a) (Y. S...) A {I e e e.t. e.t. W (a) (b) W (F0) (F0) (FC) (FC) (I) Ygo Mn H20 “5"” - (Ill)Yz'OH+ MnOH- H (Y: 8.) (F0) Figure 5-2 Possible reaction paths for proton and electron transfer. The four species that are shown are identified as follows: (I) the reactant state, (IV) the product state, and the two charged intermediate states, (II) and (III). In path (a), electron transfer is followed by proton transfer, while in path (b), proton transfer is followed by electron transfer. In path (c), the electron and proton transfer in a concerted process. 57 Table 5-1 Reduction Potentials and pKa Values Used in Determining the Viability of the ETIPT and PTIET Pathways of Figure 5-1. Redox couple Reduction potential Reference Fwy/P680 E... = 1.12 a: 0.05 v (6) ” Yz'le Em = 0.97 :t 0.02 V (90) 81/80 Em s 0.70 V (90) 82/81, 83/32 Em = 0.90-0.95 V (90) Y'IY‘ Em = 0.68 V (aqueous) (7) 02/H2O Em = 0.93 V (aqueous) (8) Molecule pKa value Reference “YT _.- -2 (aq ueous) W” __..__ WW- ( 7) Mn(H2O) 2 7 (38) 58 of at least 8 kcaI/mol. Furthermore, the difference in reduction potentials between the two cofactors will increase in the low dielectric of a protein environment, resulting in an even larger barrier. We can estimate the effect of the protein dielectric on these potentials by using the Born model: AG = 14.397/8r where AG 0!) is the electrostatic penalty for introducing a charged species with radius r (A) in a medium with a homogenous dielectric constant 8 (97). Assuming an average radius of 3 A for the tyrosine aromatic ring and a dielectric constant of about 8-10, which has been recently suggested for PSII (98), the Y°/Y’ potential will decrease from an aqueous value of 0.68 V to a value of 0.26 V. This leads to an activation barrier of at least 16 kcal/mol as illustrated in Figure 5- 3 (path a), which includes the energy that would be regained from the stabilizing interaction between the positive and negative charge produced. Activation energies have been measured for each of the S-state transitions (see Table 5-2) and are much lower than this estimated barrier. Thus, electron transfer followed by proton transfer is not supported experimentally and this first mechanistic possibility can be discarded. The second possibility is the transfer of a proton followed by the sequential transfer of an electron, PT/ET, Figure 5-2 (path b). Energetic considerations must also be taken into account to test the viability of forming the cationic tyrosine radical (state Ill). The pKa’s of the proton donor and acceptor 59 Path 8: Yzo‘ Mn+H,o t 9- YZOH Mn+0H" p. . (I YZO- Mn+H,o e.t. / Ea(et)2 16 £133: 028.) YZO- Mn H20 “25"“) Path b: + + YZOH Mn 0H- Y;OH MnOH / A Ea(pt)_>_ 12 $3! e.t. (Y5 5") (Y. S...) YZO- Mn H20 p.t. - YZOHT MnOH‘ Figure 5-3 Calculated activation energies (Ea) for paths a and b. This figure kindly provided by Gerald T. Babcock. Table 5-2 Kinetic and Thermodynamic Parameters for the S-state Transitions - Taken from (7, 59, 60) Reaction Ea (kcal M") A (s") KIE )r,°so—>)rzs1 1.2 Yz'S19YZS2 3.0 4.0x10‘5 1.3; 2.9 Yz'szevzsg 9.0 5.4x1o9 1.3 v;s,—>vzso 5.0T 8.9x105i 1 4-1.6 92* 2.9x10‘“ Ea = activation energy A = Arrhenius preexponential factor KIE = kinetic isotope effect Yz'le = reduction potential of Yz'le Sn+1/Sn = reduction potential of SWISn IT>280K *T<280K 61 are now important in determining the energetics of the process. The pKa of a Mn-l-I2O complex is expected to be 2 7 (99), while the pKa of Yz° measured in water is —2 (100). Thus, the ApKa for the proton transfer is -9, which as Figure 5- 3 (path b) illustrates, results in a barrier of at least 12 kcallmol, with this value reaching greater than 23 kcallmol in the lower dielectric protein environment. Consequently, the transfer of a proton prior to that of the electron leads to the formation of a barrier at least 3 kcallmol higher than has been observed for the S-state transitions, which allows us to discard this possibility. The general characteristics of the S-state transitions are summarized in Table 5-2 and, in many respects, are remarkable. Renger and coworkers have carefully characterized the activation energy EE, and pre-exponential factor A for the $1982, $2983, and 839(84)9So transitions (60). These values are substantially lower than those typically found for an enzyme catalyzed reaction; textbook values are usually taken to be about 10-12 kcallmol and 1013 M"s‘1 for E. and A, respectively (101, 102). The Iow A values, in particular, are striking. The low rates for Y; reduction and the small kinetic isotope effects that have been measured are also unusual considering the small distance between donor and acceptor (27-29) and that proton tunneling has been suggested in Y; reduction (59). Whether the concerted ETPT mechanism can explain these characteristics now becomes the issue at hand. 5.3 Concerted Electron and Proton Transfer Typical kinetic parameters that are observed for H-atom abstraction by radicals are reported in Tables 5—3 and 5-4. Considering first the Ea values, the 62 Table 5-3 : Kinetic and Thermodynamic Parameters for H-atom Abstraction Reactions From an Oxygen Atom by an Oxy-based Radical Reaction Ea log A KIE k Ref (kcal M") (M‘s‘) (M‘s‘) 4-OMe-Phenol + PhO' n.d. n.d. n.d. 4.75 x 105 (103, 104) 4-H-Phenol + PhO' 4.9a n.d. n.d. 3.5 x 105 (103, 104) 4-t-butyl-Phenol + PhO' 4.8 5.5 n.d. 51.3 (Mahoney, 1969 #210) 3,5-dimethyl-Phenol + PhO' 6.8 6.4 n.d. 16.3 (104) or-tocopneroi + PhO' 2.0 10.0 1.17 3.1 x 10° (105) a-naphthol + PhO' 2.2 8.9 n.d. 2.3 x 107 (105) ubiquinoI-O + PhO' 3.5 10.5 n.d. 9.1 x 107 (105) o-tocopnerol + (C5H5)2CO' n.d. n.d. 1.09 5.1 x 109 (Evans, 1992 #217) a-tocopherol + (CH3)3CO' n.d. n.d. 1.31 3.8 x 109 (Evans, 1992 #217) (CH3)3CO-H + .OC(CH3)3 2.6 6.4 n.d. n.d. (106) 3 Calculated by the BEBO method. 63 Table 5-4: Kinetic and Thermodynamic Parameters for Hydrogen Atom Abstraction by Radicals. Kinetic and Thermodynamic Parameters for Hydrogen Atom Abstraction by Radicals. Reaction E. (kcal M") log A (M"s") KIE k (M"s") Ref CH3CH2-H + 'CH3 11.59 11.7 4.04 2.57 x 105 (107) H3C-H +‘c113 14.51 11.8 3.61 7.41 x 103 (107) SiH3-H + 'CH3 6.98 11.9 3.10 1.20 x 108 (107) NH2-H + 'CH3 9.72 10.8 5.20 2.82 x 105 (107) SH-H + 'CH3 2.50 11.6 3.85 1.44 x 1010 (107) SiH3-H + 'CH2CH3 7.24 11.7 1.08 5.89 x 107 (107) PhS-H + 'CH2Ph 1.5-1.7 9.1-9.4 n.d. 1.3 x 108 (108) (CH3)3Sn-H + 'CH3 3.23 9.4 2.3 1.06 x 107 (109) (CH3)gSn-H + 'CH2CH3 3.8 9.1 1.9 2.3 x 106 (109) CchH2-H + 'OOCH3 14.9 11.5 n.d. n.d. (110) CH3-H + 'OOCH3 18.48 11.3 n.d. n.d. (110) 64 tabulated data span a broad, almost 20 kcal M'1 range. However, for reactions in which a hydrogen atom is abstracted from an oxygen atom by an oxy-based radical, which is the situation we postulate in PSII, the situation simplifies and becomes more homogeneous. The Ea values measured for these types of reactions are most often in the range of 2—7 kcallmol, and are unusually small compared to the values shown in Table 5-4 for other H-atom abstraction reactions, especially those in which C-H or N-H bonds are the hydrogen-atom donor. This phenomenon has been rationalized by Zavitsas and Chatgilialoglu as being due to triplet repulsion in the transition state (TS) (106). As the parallel spins on the donor and acceptor molecules move closer together during formation of the TS, a repulsive energy term becomes more significant to the overall activation energy barrier for the reaction. This term is much smaller for H- atom abstraction between RO' and —O-H bonds, thereby allowing the oxygen atoms to form a tighter transition state, which lowers the activation energy barrier. On the other hand, as Table 5-4 shows, H-atom abstraction between two carbon atoms, for example, occurs with much higher activation energies. According to Zavitsas and Chatgilialoglu, this is due to the stronger triplet repulsion experienced by carbon compared to oxygen. In the case of PSII, reduction of Y2’ has been proposed to occur via hydrogen atom abstraction by the tyrosine radical from Mn-ligated water or hydroxide ligands. The low activation energies that have been measured in PSII fit well in this context as reflecting what has been measured in other systems in which H-atom transfer between an oxygen-based radical and an -0-H bond occurs. Other theoretical treatments 65 have reached similar conclusions with respect to low Ea values, when hydrogen bonding allows relatively close approach of reactant species in a hydrogen-atom transfer process (81, 124). Textbook values for the Arrhenius pre-exponential factor, A, for hydrogen- atom abstractions are generally low; A s 108'5 M'1s'1 is often cited as a typical number (101, 102). These values are for second order processes, but with effective PSII concentrations of 1-10 M, the values reported below for first-order PSII processes become essentially directly comparable to the second-order solution reactions. The fact that the S-state transitions also proceed with low A factors provides support for our conclusion that the PSII reactions proceed along the concerted pathway in Figure 5-2. We can gain physical insight into the meaning of these values for A as follows. The classic interpretation of an A factor is as an “attempt frequency”. A low A value implies an unfavorable entropy of activation, which indicates that the system can position itself to enter the transition state only infrequently. For pure electron tunneling at the short Yz'Sn distance and for pure proton transfer much higher, typically 1013 M‘s", A values are expected (71, 111), which indicates a much higher attempt frequency. The fact that these larger A values are not found in PSII reinforces our conclusion (112) that the sequential pathways, a and b in Figure 5-2, are not likely in the 8- state advances. Figure 5-2 rationalizes the A-values in terms of the Franck-Condon (FC) factors for proton and electron transfer and follows closely arguments that have been advanced by Cukier (113). A rate constant expression for concerted 66 PCET can be formulated in analogy to the Marcus-Levich pure ET rate constant as(114x 2 e-( ).\'+:L\(;0+&'fi, .cin' )2 /4/1.Sk”'l‘ kerpr : 71 2.87 2” 2W" The parameters in Eq. (1) that come from the coupling of the reactant and product charge distributions to the surrounding medium are the solvent reorganization (As) and reaction-free energies (AG0); V6, is the electronic coupling. These are the standard parameters that determine an ET rate constants value (115, 116). The new ingredients are Franck-Condon factors of the initial (final) state proton vibronic wavefunctions 1,2,1 (12,). The initial states are summed over the equilibrium initial state proton distribution, pin». The “effective” activation energy appearing in the exponent of Eq. (1) involves the energetic difference of the proton eigenstates, am - an, and arises from the requirement of overall energy conservation between the initial and fine! electron-proton states. The tunneling of the proton is manifested in the Franck-Condon (FC) factors in Eq. (1). These factors tend to be small as they are determined by the overlaps of proton wavefunctions localized around the initial and final proton equilibrium positions, respectively. Thus, concerted PCET is limited by a “Franck-Condon drag” that is a reflection of the requirement that both the electron and the proton are tunneling species. As tunneling is a difficult process, and in concerted PCET there are two tunneling species, the tendency is to obtain relatively small rate constants. However, concerted PCET still can be the dominant reaction channel if the sequential processes are rate limited. As discussed above, both ET/PT 67 and PT/ET sequential processes are indeed rate limited and that leaves the concerted process as the viable channel. Use of reasonable choices for all the parameters in Eq. (1) does lead to rate-constant values compatible with the magnitudes displayed in Table 2. An important feature of Eq. (1) is that it does not, strictly speaking, have the form k = Aexp(—Ea/RT). Nevertheless, over typical temperature ranges accessible to biological studies, rate constants based on Eq. (1) will approximately obey this form. Thus, we may consider that effective activation energies and pre-exponential factors can be defined for concerted PCET reactions. Evaluation of these factors for appropriate parameter ranges produce values for E, and A consonant with those in Table 2. In particular, unfavorable entropy of activation factors (low A values) are to be expected for this mechanism. These considerations provide additional support for the notion that the S-state transitions occur by concerted hydrogen-atom transfer. The rates and kinetic parameters for overall hydrogen-atom transfers can now be rationalized. If the energetics are not too unfavorable, a sequential pathway is more likely, as the Franck-Condon factors do not hinder the reaction appreciably. This appears to be the case in the bacterial reaction center, where the proton-coupled Q2 to Q‘B reaction follows a sequential route (117). If, however, the energetics are too unfavorable, as we have argued is the case for the S-state transitions in the OEC, then the concerted pathway is favored. This was also the conclusion drawn by Roth et al. (118) and by Sjodin et al. (119) for their work on small-molecule model compounds. In these two studies, the 68 authors concluded that the electron transfer/deprotonation process occurred via a concerted mechanism by using similar arguments as those presented above. Thus, for the concerted reaction, the activation energies are generally lower than for the sequential process, but the Franck-Condon drag discussed above lowers the A factors and slows the overall rate considerably. The net result is that rates on the order of 103-105 8'1 are observed for both the S-state advances and for RO' + R'OH 9 ROH + R0" in small organic systems. In concerted PCET, a substantial isotopic effect can occur, as the deuteron FC factors are smaller than those for a proton. However, these FC factors are strongly dependent on the distance between the flanking heavy atoms that form the hydrogen/deuterium/tritium (H/DfT) bond, with the FC factor increasing for decreasing heavy-atom separation. This increase is more dramatic for T vs D vs H, and studies conducted by Krishtalik on model compounds have shown that the optimum tunneling distance for H, D, and T are different (120, 121). This result leads to the conclusion that it is more favorable for heavier isotopes to spend more energy against the repulsive forces of bringing the flanking atoms together, since tunneling is more difficult than for lighter isotopes. The net result is a decrease in the heavy-atom separation for which tunneling is most effective in the series H, D, T, and this reduces the isotope effect. Furthermore, as the heavy-atom separation decreases, the FC factors saturate to values that are not strongly dependent on distance (122). Thus, once the heavy-atom separation is sufficiently small, the distinction between H. D, and T tunnel factors disappears. The rate constants that result 69 from averaging over the heavy-atom separations then lead to kinetic isotope effects that are much more modest than would othenrvise be predicted. Thus, small isotope effects for hydrogen-atom abstractions are expected and are confirmed by theoretical calculations with the result of a proton/deuteron rate ratio of no more than a factor of 2, well in agreement with experimental values of 1.7 and 2.3 reported by the lngold and Mayer groups, respectively (105, 118). In conclusion, small disturbances of the cluster caused, for example, by the substitution of the larger Sr2+ ion, could cause a change in the OEC geometrically and/or electrostatically. Any small change upon the substitution of Sr”, or any of the other modifications mentioned above, can lead to modification of the activation barrier for hydrogen atom transfer. Since the donor and acceptor species are intimately coupled in these processes, any change that leads to a higher barrier will slow the rate of hydrogen atom transfer. This behavior can explain the reduction in rates seen upon depletion of Ca2+ or substitution of Sr” in the OEC, as well as explain the results of the other modifications to the intact system. For PSII, these effects will be seen on all S- state transitions by causing a slowdown of the reduction of Y2' by the substrate water/Mn cluster. 70 BIBLIOGRAPHY 71 10. 11. 12. Bibliography Debus, R. J. (1992) Biochim Biophys Acta 1102, 269-352. Barber, J., and Kt‘Jthrandt, W. (1999) Curr. Opin. Struc. Biol. 9, 469-475. Debus, R. J. (2000) in Metal Ions in Biological Systems (Sigel, A., and Sigel, H., Eds.) pp 657-711, Marcel Dekker, Inc., New York. Joliot, P., Barbieri, G., and Chabaud, R. (1969) Photochem. Photobiol. 10, 309-329. Kok, B., Forbush, B., and McGolin, M. (1970) Photochem. Photobiol. 11, 457-475. Klimov, V. V., Allakhverdiev, S. I., Demeter, S., and Krasnovskii, A. A. (1979) Oak]. Acad. Nauk. SSSR 249, 227-230. Tommos, C., and Babcock, G. T. (2000) Biochim. Biophys. Acta 1458, 199-219. Krishtalik, L. I. (1986) Biochim. Biophys. Acta 849, 162-171. Duysens, L. N. M. (1958) in Photochemical Apparatus, Brookhaven Symposium in Biology pp 18, Brookhaven National Laboratory, Upton, NY. Ross, R. T., and Calvin, M. (1967) Biophys. J. 7, 595-614. Diner, B. A., and Babcock, G. T. (1996) in Oxygenic Photosynthesis: The Light Reactions (Ort, D. R., and Yocum, C. F., Eds.) pp 213-247, Kluwer Academic Publishers, Dordrecht. Yachandra, V. K., Sauer, K., and Klein, M. P. (1996) Chem. Rev. 96, 2927-2950. 72 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. Cinco, R. M., Robblee, J. H., Rompel, A., Fernandez, C., Yachandra, V. K., Sauer, K., and Klein, M. P. (1998) J. Phys. Chem. 102, 8248-8256. Zimmermann, J. L., Boussac, A., and Rutherford, A. W. (1986) Biochemistry 25, 4609-4615. Zimmermann, J.-L., and Rutherford, A. W. (1984) Biochim. Biophys. Acta 767, 160-167. Sivaraja, M., Tso, J., and Dismukes, G. C. (1989) Biochemistry 28, 9459- 64. Hallahan, B. J., Nugent, J. H. A., Warden, J. T., and Evans, M. C. W. (1992) Biochemistry 31, 4562-73. van Vliet, P., Boussac, A., and Rutherford, A. W. (1994) Biochemistry 33, 12998-13004. Beck, W. F., de Paula, J. C., and Brudvig, G. W. (1986) J. Am. Chem. Soc. 108, 4018-4022. Andreasson, L.-E., Hansson, 0., and von Schenck, K. (1988) Biochim. Biophys. Acta 936, 351-360. Boussac, A., Rutherford, A. W., and Styring, S. (1990) Biochemistry 29, 24-32. Haddy, A., Dunham, W. R., Sands, R. H., and Aasa, R. (1992) Biochim. Biophys. Acta 1099, 25-34. Casey, J. L., and Sauer, K. (1984) Biochim. Biophys. Acta 767, 21-28. Boussac, A., Zimmermann, J.-L., Rutherford, A. W., and Lavergne, J. (1990) Nature 347, 303-306. Gilchrist, M. L., Jr., Ball, J. A., Randall, D. W., and Britt, R. D. (1995) Proc. Natl. Acad. Sci. USA. 92, 9545-9549. 73 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. Tang, X. S., Randall, D. W., Force, D. A., Diner, B. A., and Britt, R. D. (1996) J. Am. Chem. Soc. 118, 7638-7639. Dorlet, P., Di Valentin, M., Babcock, G. T., and McCracken, J. L. (1998) J. Phys. Chem. B 102, 8239-8247. Peloquin, J. M., Campbell, K. A., and Britt, R. D. (1998) J. Am. Chem. Soc. 120, 6840-6841. Lakshmi, K. V., Eaton, S. 8., Eaton, G. R., Frank, H. A., and Brudvig, G. W. (1998) J. Phys. Chem. B 102, 8327-8335. Ono, T.-a., Noguchi, T., lnoue, Y., Kusunoki, M., Matsushita, T., and Oyanagi, H. (1992) Science 258, 1335-1337. Roelofs, T. A., Liang, W. C., Latimer, M. J., Cinco, R. M., Rompel, A., Andrews, J. C., Sauer, K., Yachandra, V. K., and Klein, M. P. (1996) Proc. Natl. Acad. Sci. USA 93, 3335-3340. luzzolino, L., Dittmer, J., Dorner, W., Meyer-Klaucke, W., and Dau, H. (1998) Biochemistry 37, 17112-17119. Barry, ,8. A., and Babcock, G. T. (1987) Proc. Nat. Acad. Sci, USA 84, 7099-7103. Diner, B. A., Force, D. A., Randall, D. W., and Britt, R. D. (1998) Biochemistry 37, 17931-17943. Blankenship, R. E., Babcock, G. T., Warden, J. T., and Sauer, K. (1975) FEBS Lett. 51, 287-293. Warden, J. T., Blankenship, R. E., and Sauer, K. (1976) Biochim. Biophys. Acta 423, 462-478. Hoganson, C. W., and Babcock, G. T. (1997) Science 277, 1953-1956. Tommos, C., and Babcock, G. T. (1998) Accounts Chem. Res. 31, 18-25. 74 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. Babcock, G. T. (1995) in Photosynthesis: from Light to Biosphere (Mathis, P., Ed.) pp 209-215, Kluwer Academic Publishers, Netherlands. Hoganson, C. W., Lydakis-Simantiris, N., Tang, X.-S., Tommos, C., Warncke, K., Babcock, G. T., Diner, B. A., McCracken, J., and Styring, S. (1995) Photosynthesis Res. 46, 177-184. Brettel, K., Schlodder, E., and Witt, H. T. (1984) Biochim. Biophys. Acta 766, 403-415. Meyer, 8., Schlodder, E., Dekker, J. P., and Witt, H. T. (1989) Biochim. Biophys. Acta 974, 36-43. Rappaport, F ., and Lavergne, J. (1997) Biochemistry 36, 15294-15302. Schilstra, M. J., Rappaport, F., Nugent, J. H. A., Barnett, C. J., and Klug, D. R. (1998) Biochemistry 37, 3974-3981. Christen, G., and Renger, G. (1999) Biochemistry 38, 2068-2077. Christen, G., Seeliger, A., and Renger, G. (1999) Biochemistry 38, 6082- 6092. Hays, A.-M., Vassiliev, I. R., Golbeck, J. H., and Debus, R. J. (1999) Biochemistry 38, 11851 -1 1865. Diner, B. A., Nixon, P. J., and Farchaus, J. W. (1991) Curr. Opin. Struct. Biol. 1, 546-554. Roffey, R. A., Kramer, D. M., Govindjee, and Sayre, R. T. (1994) Biochim. Biophys. Acta 1 185, 257-270. Chu, H.-A., Nguyen, A. P., and Debus, R. J. (1995) Biochemistry 34, 5839-5858. Hays, A. M. A., Vassiliev, l. R., Golbeck, J. H., and Debus, R. J. (1998) Biochemistry 37, 1 1352-11365. 75 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. Mamedov, F., Sayer, R. T., and Styring, S. (1998) Biochemistry 37, 14245-14256. Babcock, G. T., Blankenship, R. E., and Sauer, K. (1976) FEBS Lett. 61, 286-289. Razeghifard, M. R., Klughammer, C., and Pace, R. J. (1997) Biochemistry 36, 86-92. Dekker, J. P., Demetrios, F. G., Plijter, J. J., Van Gorkon, H. J., and Babcock, G. T. (1984) Biochim. Biophys. Acta 767, 515-523. Boussac, A., Setif, P., and Rutherford, A. W. (1992) Biochemistry 31, 1224-34. Razeghifard, M. R., Wydrzynski, T., Pace, R. J., and Burnap, R. L. (1997) Biochemistry 36, 14474-8. Razeghifard, M. R., and Pace, R. J. (1997) Biochim. Biophys. Acta 1322, 141-150. Tommos, C., Hoganson, C. W., Di Valentin, M., Lydakis-Simantiris, N., Dorlet, P., Westphal, K., Chu, H.-A., McCracken, J., and Babcock, G. T. (1998) Curr. Opin. Chem. Biol. 2, 244-252. Karge, M., Irrgang, K. D., and Renger, G. (1997) Biochemistry 36, 8904- 8913. Hundelt, M., Hays, A.-M., Debus, R. J., and W., J. (1998) Biochemistry 37, 14450-14456. Ghanotakis, D. F., Babcock, G. T., and Yocum, C. F. (1984) FEBS Lett. 167, 127-130. Boussac, A., and Rutherford, A. W. ( 1988) Biochemistry 27, 3476-3483. Boussac, A., and Rutherford, A. W. (1988) Chem. Scripta 28A, 123-126. 76 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. Latimer, M. J., Derose, V. J., Mukerji, I., Yachandra, V. K., Sauer, K., and Klein, M. P. (1995) Biochemistry 34, 10898-10909. Noguchi, T., Ono, T., and lnoue, Y. (1995) Biochimica Et Biophysica Acta- Bioenergetics 1228, 189-200. Pecoraro, V. L., Baldwin, M. J., Caudle, M. T., Hsieh, W.-Y., and Law, N. A. (1998) Pure & Appl. Chem. 70, 925-929. Siegbahn, P. E. M., and Crabtree, R. H. (1999) J. Am. Chem. Soc. 121, 117-127. Limburg, J., Vrettos, J. S., Liable-Sands, L. M., Rheingold, A. L., Crabtree, R. H., and Brudvig, G. W. (1999) Science 283, 1524-1527. Limburg, J. Szalai, V. A, and Brudvig, G. W. (1999) J. Chem. Soc, Dalton Trans. 9, 1353- 1361. Page, C. C., Moser, C. C., Chen, X., and Dutton, P. L. (1999) Nature 402, 47-52. Riggs-Gelasco, P. J., Mei, R., Ghanotakis, D. F., Yocum, C. F., and Penner-Hahn, J. E. (1996) J. Am. Chem. Soc. 118, 2400-2410. Scheiner, S. (1994) Acc. Che. Res. 27, 402-408. Berthold, D. A., Babcock, GT, and Yocum, CF. (1981) FEBS Lett. 134, 231-234. Ghanotakis, D. F., Demetriou, D. M., and Yocum, C. F. (1987) Biochim. Biophys. Acta 891, 15-21. Ghanotakis, D. F., Topper, J., Babcock, G. T., and Yocum, C. F. (1984) Biochim. Biophys. Acta 767, 524-531. Hillier, W., Messinger, J., and Wydrzynski, T. (1998) Biochemistry 37, 16908-16914. 77 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. Tommos, C., McCracken, J., Styring, S., and Babcock, G. T. (1998) J. Am. Chem. Soc. 120, 10441-10452. Qian, M., Dao, L. A., Debus, R. J., and Burnap, R. L. (1999) Biochemistry 38, 6070-6081. Hillier, W., and Wydrzynski, T. (2000) Biochemistry in press. Hoganson, C. W., and Babcock, G. T. (1988) Biochemistry 27, 5848-5855. Babcock, G. T. (1987) in New Comprehensive Biochemistry (Amesz, J., Ed.) pp 125-158, Elsevier, Amsterdam. Lakshmi, K. V., Eaton, S. S., Eaton, G. R., and Brudvig, G. W. (1999) Biochemistry 38, 12758-12767. Rappaport, F., BIanchard-Desce, M., and Lavergne, J. (1994) Biochim. Biophys. Acta 1 184, 178-192. Lydakis-Simantiris, N., Ghanotakis, D. F., and Babcock, G. T. (1997) Biochim. Biophys. Acta. 1322, 129-140. Cha, Y., Murray, C.J., & Klinman, JP. (1989) Science 243, 1325-1330. Haumann, M., Bogershausen, 0., Cherepanov, D., Ahlbrink, R., and Junge, W. (1997) Photosynth. Res. 51 , 193-208. Barbara, P. F., Meyer, T. J., and Ratner, M. A. (1996) Journal of Physical Chemistry 100, 13148-13168. MacLachlan, D. J., Nugent, J. H. A., Warden, J. T., and Evans, M. C. W. (1994) Biochim. Biophys. Acta 1188, 325-334. Vass, |., and Styring, S. (1991) Biochemistry 30, 830-839. Winkler, J. R., and Gray, H. B. (1992) Chemical Reviews 92, 369-379. 78 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. McLendon, G., and Hake, R. (1992) Chemical Reviews 92, 481-490. Moser, C. C., Keske, J. M., Warncke, K., Farid, R. S., and Dutton, P. L. (1992) Nature 355, 796-802. Un, S., Tang, X.-S., and Diner, B. A. (1996) Biochemistry 35, 679-684. Berthomieu, C., Hienerwadel, R., Boussac, A., Breton, J., and Diner, BA. (1 998) Biochemistry 37, 10547-10554. Clark, W. M. (1960) Oxidation-Reduction Potentials of Organic Systems, The \Mlliams & Wilkins Company, Baltimore, VA. Rich, P. R., Meunier, B., Mitchell, R., and Moody, A. J. (1996) Biochim. Biophys. Acta 1275, 91-95. Mulkidjanian, A. Y. (1999) Biochim. Biophys. Acta 1410, 1-6. Caudle, M. T., and Pecoraro, V. L. (1997) J. Am. Chem. Soc. 119, 3415- 3416. Dixen, W. T., and Murphy, D. (1976) J. Chem. Soc. Faraday Trans. ll 72, 1221-1230. Segel, l. H. (1975) Enzyme Kinetics, Behavior and Analysis of Rapid Equilibrium and Steady-State Enzyme Systems, John VVIley & Sons, New York. Chance, 8., DeVault, D. C., Frauenfelder, H., Marcus, R. A., Schrieffer, J. R., and Sutin, N. (1979) Tunneling in Biological Systems, Academic Press, New York. Mahoney, L. R., and DaRooge, M. A. (1975) J. Am. Chem. Soc. 97, 4722- 4731. Mahoney, L. R., and DaRooge, M. A. (1972) J. Am. Chem. Soc. 94, 7002- 7009. 79 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. Foti, M., lngold, K. U., and Lusztyk, J. (1994) J. Am. Chem. Soc. 116, 9440-9447. Zavitsas, A. A., and Chatgilialoglu, C. (1995) J. Am. Chem. Soc. 117, 10645-10654. Arthur, N. L., and Bell, T. N. (1978) Rev. Chem. Intenned. 2, 37-74. Franz, J. A., Bushaw, B. A., and Alnajjar, M. S. (1989) J. Am. Chem. Soc. 111, 268-275. Chatgilialoglu, C., lngold, K. U., and Scaiano, J. C. (1981) J. Am. Chem. Soc. 103, 7739-7742. Mims, C. A., Mauti, R., Dean, A. M., and Rose, K. D. (1994) J. Phys. Chem. 98, 13357-13372. Silverrnan, D. N. (2000) Biochim. Biophys. Acta 1458, 88-103. Westphal, K. L., Tommos, C., Cukier, R. l., and Babcock, G. T. (2000) Current Opinions in Plant Biology 3, 236-242. Cukier, R. I. (1999) J. Phys. Chem. A 103, 5989-5995. Cukier, R. |., and Nocera, D. G. (1998) Annu. Rev. Phys. Chem. 49, 337- 369. Ulstrup, J. (1979) Charge Transfer Processes in Condensed Media, Springer, Berlin. Marcus, R. A., and Sutin, N. (1985) Biochim. Biophys. Acta 811, 265. Okamura, M. Y., Paddock, M. L., Graige, M. S., and Feher, G. (2000) in press. Roth, J. P., Lovell, 8., and Mayer, J. M. (2000) J. Am. Chem. Soc. ASAP. 80 119. 120. 121. 122. Sjodin, M., Stenbjorn, S., Akermark, 8., Sun, L., and Hammarstrom, L. (2000) J. Am. Chem. Soc. 122, 3932-3935. Krishtalik, L. I. (2000) Biochim. Biophys. Acta 1458, 6-27. Krishtalik, L. I. (1979) J. Electroanalyt. Chem. 100, 547-561. Cukier, R. |., and Zhu, J. (1997) J. Phys. Chem. 101, 7180-7188. 81 1((111)(11111111111111).1