Iowa State University Capstones, Theses and Retrospective Theses and Dissertations Dissertations

2000 Phylloquinone biosynthetic pathway Thomas Wade Johnson Iowa State University

Follow this and additional works at: https://lib.dr.iastate.edu/rtd Part of the Biochemistry Commons, and the Genetics Commons

Recommended Citation Johnson, Thomas Wade, "Phylloquinone biosynthetic pathway " (2000). Retrospective Theses and Dissertations. 12336. https://lib.dr.iastate.edu/rtd/12336

This Dissertation is brought to you for free and open access by the Iowa State University Capstones, Theses and Dissertations at Iowa State University Digital Repository. It has been accepted for inclusion in Retrospective Theses and Dissertations by an authorized administrator of Iowa State University Digital Repository. For more information, please contact [email protected]. INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the text directly from the original or copy sulxnitted. Thus, some thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the copy submitted. Broken or indistinct print, cotored or poor quality illustrations and photographs, print t>leedthrough, substarxiard margins, and improper alignment can adversely affect reproduction.

In the unlikely event that the author dkl not send UMI a complete manuscript and there are missing pages, these will t>e noted. Aiso. if unauthorized copyright material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, t>eginning at the upper left-hand corner and continuing from left to right in equal sections with small overiaps.

Photographs included in the original manuscript have been reproduced xerographically in this copy. Higher quality 6' x 9' blade and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to order.

Bell & Howell Information and Leaming 300 North Zeeb Road, Ann Arbor, Ml 48106-1346 USA 800-521-0600

Phylloquinone biosynthetic pathway

by

Thomas Wade Johnson

A dissertation submitted to the graduate faculty

in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY

Major: Biochemistry

Major Professor: Parag R. Chitnis

Iowa State University

Ames. Iowa

2000

Copyright ©Thomas Wade Johnson, 2000. All rights reserved UMI Number 9990460

Copyright 2000 by Johnson, Thomas Wade

All rights reserved.

UMI ®

UMI Microfomn9990460 Copyright 2001 by Bell & Howell Information and Leaming Company. All rights reserved. This microform edition is protected against unauthorized copying under Title 17, United States Code.

Bell & Howell Information and Leaming Company 300 North Zeeb Road P.O. Box 1346 Ann Arbor, Ml 48106-1346 ii

Graduate College

Iowa State Universit>'

This is to certify that the Doctoral dissertation of

Thomas Wade Johnson has met the dissertation requirements of Iowa State University

Signature was redacted for privacy.

Major Professor

Signature was redacted for privacy.

For the Major Program

Signature was redacted for privacy.

For the Scaduate College Ill

TABLE OF CONTENTS

ABSTRACT

CHAPTER L GENERAL INTRODUCTION

CHAPTER 2. PHYLLOQUINONE BIOSYNTHESIS AND BIOENERGETICS OF QUINONES IN THE A, SITE OF PHOTOSYSTEM I

CHAPTER 3. THE menD AND menE HOMOLOGS IN CYANOBACTERIA CODE FOR ESSENTIAL ENZYMES IN PHYLLOQUINONE BIOSYNTHETIC PATHWAY

CHAPTER 4. RECRUITMENT OF A FOREIGN QUTNONE INTO THE A, SITE OF PHOTOSYSTEM I I. GENETIC AND PHYSIOLOGICAL CHARACTERIZATION OF PHYLLOQUINONE BIOSYNTHETIC PATHWAY MUTANTS IN SYNECHOCYSTIS SP. PCC 6803

CHAPTER 5. RECRUITMENT OF A FOREIGN QUINONE INTO THE A, SITE OF PHOTOSYSTEM I IV. IN VIVO REPLACEMENT OF THE NATIVE QUINONE IN THE PHYLLOQUINONE-LESS MUTANTS OF SYNECHOCYSTIS SP. PCC 6803 BY EXTERNALLY SUPPLIED NAPHTHOQUINONES

CHAPTER 6. RECRUITMENT OF A FOREIGN QUINONE INTO THE A, SITE OF PHOTOSYSTEM I V. IN VIVO REPLACEMENT OF THE NATIVE QUINONE IN THE PHYLLOQUINONE-LESS MUTANTS OF SYNECHOCYSTIS SP. PCC 6803 BY 1,2-NAPHTHOQUINONE IV

CHAPTER 7. GENERAL CONCLUSIONS 156

APPENDIX A. DEVELOPED PROCEDURES AND PROTOCOLS SPECmC TO DISSERTATION EXPERIMENTS 158

APPENDIX B. RECRUITMENT OF A FOREIGN QUINONE INTO THE A, SITE OF PHOTOSYSTEM I II. STRUCTURAL AND FUNCTIONAL CHARACTERIZATION OF PHYLLOQUINONE BIOSYNTHETIC PATHWAY MUTANTS BY ELECTRON PARAMAGNETIC RESONANCE AND ELECTRON-NUCLEAR DOUBLE RESONANCE SPECTROSCOPY (SUMMARY) 166

APPENDIX C. THE RED-ABSORBING CHLOROPHYLL a ANTENNA STATES OF PHOTOSYSTEM 1: A HOLE-BURNING STUDY OF SYNECHOCYSTIS SP PCC 6803 AND ITS MUTANTS (SUMMARY) 168

APPENDIX D. ULTRAFAST PRIMARY PROCESSES IN PHOTOSYSTEM I REACTION CENTERS WITH A FOREIGN QUINONE IN THE A, SITE (SUMMARY) 169

APPENDIX E. ULTRAFAST PRIMARY PROCESSES IN A CYANOBACTERIAL PHOTOSYSTEM I REACTION CENTER (SUMMARY) 170

APPENDIX F. LIST OF UNREFEREED PUBLISHED PAPERS 171

ACKNOWLEDGMENTS 172 V

ABSTRACT

Photosynthesis converts solar energy into chemical energy to drive the biological processes on our planet. Phoiosystem I (PSI) is one of the two reaction centers of oxygenic photosynthesis in cyanobacteria and higher plants. PSI is a multisubunit membrane bound protein complex that catalyzes the photooxidation of plastocyanin in the thylakoid lumen and the photoreduction of ferredoxin in the cytoplasm. The PsaA/PsaB subunits form the heterodimeric core that harbors the primary electron donor P700 and acceptors Aq, A,, and Fx- Two electron acceptors. and Fg, occupy PsaC one of three peripheral proteins bound to PSI. The secondary electron acceptor. A,, contains a phylloquinone, a molecule utilized only in PSI. The object of this dissertation is to probe phylloquinone biosynthesis and the bioenergetic aspects of phylloquinone, plastoquinone, and other naphthoquinones in the A, site of PSI. To investigate phylloquinone's role in PSI, we substituted other naphthoquinones and plastoquinone the A| site to observe the physiological changes of the cells and energetic changes in PSI. Menaquinone biosynthetic genes have been identified in Esherichia coli with the phylloquinone homologs elucidated from the Synechocystis sp. PCC 6803 genome. Four phylloquinone biosynthetic genes in Synechocystis have been mutationally disabled. mcnA. meriB. menD. and menE, leading to a complete loss of phylloquinone in the cells. The -A.I site of the mutant contains plastoquinone, allowing a reduced electron transfer in PSI rate. The orientation and distance of the Ai quinone to P700 is the same as wild type. Physiological changes in the mutants include decreased growth rate, oxygen evolution, and chlorophyll content. Addition of the direct product of naphthoate synthase (menB). 2- carboxy-l,4-naphthoquinone. restored phylloquinone content of the PSI complexes to the wild-type levels. Mutant cells also use free phylloquinone from the growth medium to displace plastoquinone in PSI. Addition of vitamin K3, 1.4-naphthoquinone. 2-hydro.xy-1.4- naphthoquinone, or 1.2-naphthoquinone allowed the cells to synthesize phytylated naphthoquinones, which were inserted into the Ai site in a fraction of the PSI complexes. The results in this dissertation determined the biosynthetic pathway of phylloquinone and alternative phytylated quinones in the Ai site are capable of electron transfer from Ao to Fx. 1

CHAPTER 1. GENERAL INTRODUCTION

Photosynthesis is an extremely important process that converts solar energy into chemical energy, thereby fueling the living world directly or indirectly. In oxygenic photosynthesis, cyanobacteria and the chloroplasts of higher plants and algae sequester light through photosynthesis to release oxygen from water and produce carbohydrates from carbon dioxide reduction. In the light driven reactions of photosynthesis, photosystem I (PSI) and photosystem II (PSII) capture photons to utilize the energy for electron transfer from water to NADP"' through three membrane bound complexes: PSIL cytochrome btf and PSI. This process produces NADPH and a proton gradient across the thylakoid membranes that is used by ATP synthase to produce ATP. These products. ATP and NADPH. are used to reduce carbon dioxide in the production of carbohydrates. Photosystem I (PSI) is a well-characterized photosynthetic reaction center in cyanobacteria and higher plants (Golbeck, 1993; Nitschke and Rutherford. 1991: Nugent.

1996). PSI catalyzes the photooxidation of plastocyanin in the thylakoid lumen and the photoreduction of ferredoxin in the cyanobacterial cytoplasm (or chloroplast stroma). PSI contains at least eleven proteins, with PsaA and PsaB as a heterodimeric core. Surrounding PsaA/PsaB are 6 small subunit proteins, that serve various identified or unknown functions

(Chitnis. 1996). In addition to light-harvesting chlorophyll a (Chi a) molecules. PsaA/PsaB is imbedded with the P700 reaction center and a series of cofactors (Ao. Ai. Fx) that act as intermediate electron acceptors and provide a charge separation from the initial P700 reaction center (Brettel. 1997). P700 is comprised of a dimer of Chi a molecules. The electron acceptor cofactors include a molecule of chlorophyll a. AQ. phylloquinone. A\ and a [4Fe-4S] iron sulfur protein. Fx- PSI also contains three overlapping peripheral proteins (PsaC. PsaD and PsaE). PsaC is a prerequisite for PsaD and PsaE binding and contains the two terminal electron transfer [4Fe-4S] iron sulfur clusters FA and FB(YU et al.. 1993:

Antonkine et al.. 2000; Yu et al., 1995). Fa and Fb act as an electron shuttle from the PsaA/PsaB core cofactors to docked ferredoxin. PsaD provides a ferredoxin docking site (Kruip et al.. 1997; Chitnis et al., 1996), with PsaE assisting in the interaction of PSI with ferredoxin (Xu et al.. 1994). 1

The secondary electron acceptor site. A|. contains phylloquinone (PhQ. vitamin K|. 2-methyl-3-phytyl-1.4-naphthoquinone). There are two bound molecules of phylloquinone per molecule of P700 in PSI (Malkin. 1986; Schoeder and Lockau, 1986: Takahashi et al.. 1985: Biggins and Mathis. 1988). This molecule acts as an intermediate in the electron transfer from the primary acceptor Ao- a chlorophyll a to Fx- an iron-sulfur cluster (Kumazaki et al.. 1994: Brettel. 1988). The Ao to Fx electron transfer is an estimated 250 mV (Semenov et al.. 2000) to 320 mV (Brettel. 1997) event. PhQ acts as both a physical and energetic bridge for the electron between the primar>' acceptor and the terminal iron sulfur clusters. From the literature, several of non-native quinones were inserted into cyanobacterial PSI with varied and contradictory amounts of success towards Ai reduction. Phylloquinone can be extracted from PSI with a drj' or water saturated solution (Itoh et al.. 1987: Itoh and Iwaki. 1989) or by a hexane containing 0.3% methanol (Rustandi et al., 1990: Setif et al.. 1987). Both of these methods remove a significant amount of antenna Chi. caroteniods and lipids. Both extraction procedures blocked the electron from traveling past -Ao (Biggins and Mathis, 1988: Itoh et al.. 1987). Iwaki and Itoh reported then function of PhQ could be replaced by a quinone of an appropriate redox potential (Iwaki and Itoh. 1994). However, these compounds did not restore reduction of the terminal iron sulfur clusters. Biggins suggested structural constraints imposed by an alkyl tail in addition to energetic considerations (Biggins. 1990). Forward electron transfer to Fx was only achie\ ed \\ ith quinone compounds that had a hydrophobic side chain that determined the orientation of the quinone in the A| site. Discrepancy between these results could have arrived from a combination of the harsh PhQ extraction procedures and the lack of resolution of the instruments. The genes encoding the synthesis of phylloquinone have not been described in cyanobacteria. However, the biosynthetic genes of menaquinone, a phylloquinone-likc molecule, have been cloned in other bacteria, such as Esherichia coli. (Palaniappan et al.. 1992: Sharma et al.. 1992: Sharma et al.. 1993: Daruwala et al.. 1996: Sharma et al.. 1996: Suvama et al.. 1998). The nucleotide sequence of the Syncchocystis sp. PCC 6803 genome shows the existence of homologues of the menA. menB, menC. menD. menE, and mcnF J

{entC), and tnenG genes which code for enzymes involved in menaquinone biosynthesis in other bacteria. It is likely that the phylloquinone biosynthetic pathway is similar to menaquinone. given that the difference between the molecules is a mostly saturated C-20 phytyl tail and an unsaturated C-45 a nine unit isopreniod like side chain, respectively. The phytyl transferase enzyme. MenA. has low homology compared to the menaquinone analog. The MenB enzyme, that closes the second ring in the naphthoquinone, is the last step in the construction of the naphthoquinone head group with the following two steps consisting of attaching the phytyl tail and a C-2 methylation of naphthoquinone. The menD and menE genes encode for enzymes towards the beginning of the PhQ biosynthetic pathway. With the current understanding of how phylloquinone is utilized in PSI and the existence of homologs of the menaquinone biosynthetic genes in Synechocystis sp. FCC 6803. we propose to systematically disable genes in the biosynthetic pathway of phylloquinone to achieve two goals: 1) Create a series phylloquinone-less mutants, thus confirming the biosynthetic pathway. 2) Monitor the physiological changes within the cell and the energetic changes within PSI of the phylloquinone-less mutants. We would then be able to supplement non- native quinones to probe the structural and energetic requirements of the Ai site of PSI. 1. Organization of Dissertation This dissertation contains four manuscripts and a literattire review of phylloquinone synthesis and function in PSI. To confirm the genes involved in proposed phylloquinone biosynthetic pathway, the menA. menB. menD, and menE genes were mutationally disabled. This work has been divided into two manuscripts. The first half of this work concerning the menD and menE gene disruptions, which are towards the beginning of the PhQ pathway, has been summarized in the manuscript "The menD and menE homologues in cyanobacteria code for essential enzymes in phylloquinone biosynthetic pathway" to be submitted to the Plant Physiology. The work covering the men.A and menB gene disruptions is titled "Recruitment of a foreign quinone into the A, site of photosystem I I. genetic and physiological characterization of phylloquinone biosynthetic pathway mutants in Synechocystis sp. PCC 6803'" and is published in the Journal of Biological Chemistry. 275. 4

8523-8530. (2000). The four gene disruption mutations revealed that plasloquinone-9 occupies the A| site of PSI in place of the absent phyiloquinone (Semenov et al.. 2000; Zybailov et al.. 2000; Johnson et al.. 2000). The next logical step is to insert alternative naphthoquinones into the A| site to inonitor the physiological adaptations of the cells and the energetic changes within PSI. This study has been described in the manuscript "Recruitment of foreign quinones in the A| site of cyanobacterial photosystem 1 complex. IV. In vivo replacement of the native quinone in the phylloquinone-less mutants of Synechocystis sp. PCC 6803 by externally supplied naphthoquinones" to be submitted to Journal of Biological Chemistry. Further investigation of the addition of 1.2- naphthoquinone to the menB mutant cells was performed and organized in the manuscript "Recruitment of foreign quinones in the Ai site of cyanobacterial photosystem I V in vivo replacement of the native quinone in the phylloquinone-less mutants of Synechocystis sp. PCC 6803 by 1.2-naphthoquinone". Following will be a general conclusions section highlighting the developments of the research. For future graduate students who will be expanding this research, there is a section of protocols and procedures that were developed or adapted specifically for these experiments and conditions. The remaining sections are summaries of second author papers published or in preparation. The appendices will refer the reader to related topics concerning the phyiloquinone less mutants and experiments that are beyond the scope of this dissertation. 2. References

.Antonkine. M. L.. D. Bentrop. I. Bertini, C. Luchinat. G. Shen. D. A. Br\'ant. D. Stehlik. and J. H. Golbeck. 2000. Paramagnetic IH NMR spectroscopy of the reduced, unbound photosystem 1 subunit PsaC: sequence-specific assignment of contact-shiftcd resonances and identification of mixed- and equal-valence Fe-Fe pairs in [4Fc-4Sl centers FA- and FB-. J Biol Inorg Chem. 5:381-92. Biggins, J. 1990. Evaluation of selected benzoquinones. naphthoquinones, and anthraquinones as replacements for phyiloquinone in the Al acceptor site of the photosystem 1 reaction center. Biochemistry. 29:7259-64. Biggins. J., and P. Mathis. 1988. Functional role of vitamin K in photosystem 1 of the cyanobacterium 6803. Biochemistry. 27:1494-500. Brettel. K. 1988. Electron transfer from Al- to an iron-sulfur center with tl/2 = 200 ns ai room temperature in photosystem I. Characterization by flash absorption spectroscopy. In FEBS Lett. 93-8. Bretiel. K. 1997. Electron transfer and arrangement of the redox cofactors in photosystem I. Biochim. Biophys. Acta. 1318:322-373. Chitnis. P. R. 1996. Photosystem I. Chitnis. V. P.. Y. S. Jungs. L. Albee. J. H. Golbeck, and P. R. Chitnis. 1996. Mutational analysis of photosystem I polypeptides. Role of PsaD and the lysyl 106 residue in the reductase activit>' of the photosystem \. J Biol Chem. 271:11772-80. Daruwala, R., O. Kwon, R. Meganathan. and M. E. Hudspeth. 1996. A new isochorismate synthase specifically involved in menaquinone (vitamin K2) biosynthesis encoded by the menF gene. FEMS Microbiol Lett. 140:159-63. Golbeck. J. H. 1993. Shared thematic elements in photochemical reaction centers. Proc Natl AcadSci USA. 90:1642-6. Itoh. S., and M. Iwaki. 1989. Vitamin K sub(l) (phyiloquinone) restores the turnover of FeS centers in the ether-extracted spinach PS I particles. FEES LETT. 243:47-52. Itoh. S.. M. Iwqaki, and I. Ikegami. 1987. Extraction of vitamin K-1 from photosystem I particles by treatment with diethyl ether and its effect on the AfEPR signal and system I photochemistry. Biochim. Biophys. Acta. 893:508-516. Iwaki. M.. and S. Itoh. 1994. Reaction of reconstituted acceptor quinone and dynamic equilibration of electron transfer in the photosystem I reaction center. Plant Cell PhysioL 35:983-993. Johnson. T. W., G. Shen, B. Zybailov. D. Kolling. R. Reategui. S. Beauparlant. I. R. Vassiliev. D. A. Bryant. A. D. Jones. J. H. Golbeck, and P. R. Chitnis. 2000. Recruitment of a foreign quinone into the A(l) site of photosystem I. I. Genetic and physiological characterization of phyiloquinone biosynthetic pathway mutants in Synechocystis sp. pcc 6S03. J Biol Chem. 275:8523-30. Kruip. J.. P. R. Chitnis. B. Lagoutte. M. Rogner. and E. J. Boekema. 1997. Structural organization of the major subunits in cyanobacterial photosystem 1. Localization of subunits PsaC, -D, -E, -F. and -J. J Biol Chem. 272:17061-9. Kumazaki. S.. M. Iwaki. I. Ikegami, H. Kandori. K. Yoshihara, and S. Itoh. 1994. Rates of primar>' electron transfer reactions in the photosystem I reaction center reconstituted with different quinones as the secondary acceptor. J. Phys. Chem. 98:11220-11225. Malkin. R. 1986. On the function of two vitamin K molecules in the PSI electron acceptor complex. FEBS Lett. 208:343-346. Nitschke. W.. and A. Rutherford. 1991. Photosynthetic reaction centres: Variations on a common structural theme? Trends Biochem. Sci. 16:241-245. Nugent. J. H. 1996. Oxygenic photosynthesis. Electron transfer in photosystem I and photosystem II. Eur. J. Biochem. 237:519-531. Palaniappan. C.. V. Sharma. M. E. Hudspeth, and R. Meganathan. 1992. Menaquinone (vitamin K2) biosynthesis: evidence that the Escherichia coli menD gene encodes 6

both 2-succinyl-6-hydroxy-2.4-cyclohe.\adiene-I- carboxylic acid synthase and alpha-ketoglutarate decarboxylase activities. J Bacteriol. 174:8111-8. Rustandi. R. R.. S. W. Snyder. L. L. Feezel. T. J. Michalski. J. R. Norris, M. C. Thumauer. and J. Biggins- 1990. Contribution of vitamin K1 to the electron spin polarization in spinach photosystem I. Biochemistry. 29:8030-2. Schoeder. H. U.. and W. Lockau. 1986. Phylloquinone copurifies with the large subunit of photosystem I. FEES Lett. 199:23-27. Semenov. A. Y.. I. R. Vassiiiev. A. van Der Est. M. D. Mamedov. B. Zybailov. G. Shen. D. Stehlik. B. A. Diner. P. R. Chitnis. and J. H. Golbeck. 2000. Recruitment of a Foreign Quinone into the A1 Site of Photosystem I. Altered Kinetics of Electron Transfer in Phylloquinone Biosynthetic Pathway Mutants Studied by Time-Resolved Optical, EPR and Electrometric Techniques. J Biol Chem. 275:23429-23438. Setif, P., I. Ikegami. and J. Biggins. 1987. Light-induced charge separation in Photosystem I at low temperature is not influenced by vitamin K-1. Biochim. Biophys. Acta. 894:146-56. Sharma, V.. M. E. Hudspeth, and R. Meganathan. 1996. Menaquinone (vitamin K2) biosynthesis: localization and characterization of the menE gene from Escherichia coli. Gene. 168:43-8. Sharma, V.. R. Meganathan, and M. E. Hudspeth. 1993. Menaquinone (vitamin K2) biosynthesis: cloning, nucleotide sequence, and expression of the menC gene from Escherichia co\i. J Bacteriol. 175:4917-21. Sharma. V.. K. Suvama. R. Meganathan, and M. E. Hudspeth. 1992. Menaquinone (vitamin K2) biosynthesis: nucleotide sequence and expression of the menB gene from Escherichia coli. J Bacteriol. 174:5057-62. Suvarna. K.. D. Stevenson. R. Meganathan. and M. E. Hudspeth. 1998. Menaquinone (vitamin K2) biosynthesis: localization and characterization of the menA gene from Escherichia coli. J Bacteriol. 180:2782-7. Takahashi. Y.. K. Hirota. and S. Katoh. 1985. Multiple forms of P700-chlorophyll ^-protein complexes from Synechococcus sp.: The iron, quinone. and carotenoid contents. Phoiosynth. Res. 6:183-192. Xu. Q.. Y. S. Jung. V. P. Chitnis. J. A. Guikema. J. H. Golbeck, and P. R. Chitnis. 1994. Mutational analysis of photosystem I polypeptides in Synechocystis sp. PCC 6803. Subunit requirements for reduction of NADP"*" mediated by ferredoxin and flavodoxin. J. Biol. Chem. 269:21512-21518. Yu. J.. L. B. Smart. Y. S. Jung. J. Golbeck. and L. Mcintosh. 1995. Absence of PsaC subunii allows assembly of photosystem I core but prevents the binding of PsaD and PsaE in Synechocystis sp. PCC6803. Plant Mol Biol. 29:331-342. Yu. L., J. Zhao. W. Lu. D. A. Bryant, and J. H. Golbeck. 1993. Characterization of the [3Fe- 4S] and [4Fe-4S] clusters in unbound PsaC mutants C14D and C51D. Midpoint 7

potentials of the single [4Fe-4S] clusters are identical to FA and FB in bound PsaC of photosystem I. Biochemistry. 32:8251-8. Zybailov. B.. A. van der Est. S. G. Zech. C. Teutloff. T. W. Johnson. G. Shen. R. Bittl. D. Stehlik. P. R. Chitnis. and J. H. Golbeck. 2000. Recruitment of a foreign quinone into the A(l) site of photosystem I. II. Structural and functional characterization of phylloquinone biosynthetic pathway mutants by electron paramagnetic resonance and electron-nuclear double resonance spectroscopy. J Biol Chem. 275:853 1-9. 8

CHAPTER 2. PHYLLOQUINONE BIOSYNTHESIS AND BIOENERGETICS OF QUINONES IN THE A, SITE OF PHOTOSYSTEM I

1. Format of review This review highlights two related topics: phyiloquinone (PhQ. vitamin K| (K). 2- methyl-1.4-naphthoquinone) biosynthesis and the energetic properties of phyiloquinone in photosystem 1 (PSI). To adequately describe and place each subject in context, several secondar>' foci will be addressed. PhQ. and related compounds, in bacteria, plants, and animals are described for production and functions within the organisms. Plastoquinone (PQ) will also be described, since this molecule can mimic some of the functions of phyiloquinone. The genes involved in the biosynthetic pathway of phyiloquinone will be compared to menaquinone (vitamin Ki) biosynthesis genes in bacteria and higher plants, given the structural similarities between the molecules. Phyiloquinone is also a critical cofactor in PSI. hence the PSI complex and its subunit components are described. How PSI interacts with the electron donor and acceptor proteins is also mentioned. A brief description of the entire photosynthetic electron shuttling process from photosystem II (PSII) through Cyt hef through PSI will be given. The photosynthetic energy relationship of PSI is described in greater detail, with a focus on phyiloquinone. 2. The use of phyiloquinone in bacteria, plants, and animals Phyiloquinone or vitamin K is a fat soluble essential nutrient found in vegetables and certain oils. Fig 2.3 (Booth and Suttie. 1998). The recommended dietar>' allowance (RDA) of vitamin K is 1 >ig/(kg body wt • d) with the average American consumption exceeding the RD.A. at 300-500 fig/d (Monsen. 1989). Green leafy vegetables such as collard greens, spinach, and salad greens contain 40-50% of the daily intake of vitamin K. Oils like canola and soybean make an appreciable but minor contribution to the total vitamin K intake, at close to 15% (Bolton-Smith et al.. 2000). Menaquinone. referred to as vitamin K;. contributes a relatively small amount to satisfy the human requirement for the vitamin (Reedstrom and Suttie. 1995). Both molecules are absorbed in the gut for utilization by the body. Regardless of the structural similarities between the two. menaquinone cannot be converted into phyiloquinone. at least not in rats (Ronden et al.. 1998). 9

Vitamin K. Figure 2.1 and Figure 2.3. is a well known molecule that is essential for blood clotting (Voet and Voet. 1995). In vertebrates and invertebrates, the only known use for phylloquinone is as a cofactor for a single enzymatic reaction: the conversion of to y-carboxyglutamic acid (Gla) in select proteins (Furie et al.. 1999). Vitamin K is involved in activating several blood clotting proteins (factor VII, factor IX. and factor X) (Dam. 1935) and creating the y-carbo.xyglutamic acid which then acts as a metal binding amino acid side chain that confers chelating properties to proteins (Sperling et al.. 1978). These metal chelating proteins derive a structural rigidity or conformational change upon addition of or calcium ions. As an example, a derived from a venomous cone snail contains the polypeptide conantokin G (Rigby et al.. 1997). This inactive sleeper contains 17 residues with 5 y-carboxyglutamic acid residues spaced every three to four amino acids. Upon addition of Mg""^ or Ca""^, the polypeptide forms an a- helix and the Gla side chains orient into a linear array with each ion being chelated by two adjacent carboxyl groups from different residues. The molecule is then toxic. Recently the vitamin K-dependant carboxylase (carboxylase) was purified and cloned (Wu et al.. 1991). Carboxylase is a single chain protein of 758 amino acids (94 kD) and is dominated by hydrophobic amino acids. The enzyme has at least five known functional properties: a carboxylase active site, an epoxidase active site, a substrate recognition site, a polypeptide stimulation site for carboxylase and epoxidase activity, and a phylloquinone binding site. Each of the properties is employed in the mechanism for y- carboxylation. Carboxylation of glutamic acids is a multistep process in which the premodified amino acid is recognized and modified by carboxylase with a cocontaminant cycling of phylloquinone from they hydroquinone to epoxide state. Figure 2.1. The native state phylloquinone is first reduced to the active hydroquinone by a separate enzyme, vitamin K epoxide reductase (Suttie. 1985). Then phyllohydroquinone is converted, by the carboxylase, to the alkoxide which acts as a strong base for the removal of the y-proton on specific glutamic acid residues creating an anion and releasing hydroxide. Subsequently. CO: is added to the anionic carbon creating the y-carboxyglutamic acid side chain. The activated vitamin K species convert into vitamin K epoxide. The epoxide is reduced back 10

into the phyllohydroquinone by vitamin K epoxide reductase (Wallin and Martin. 1987). A class of vitamin K antagonists, such as warfarin, were discovered as anticoagulation agents that inhibit the conversion of the epoxide back into the phyllohydroquinone (Campbell and Link. 1941).

CARBOXYLASE

OH Cr vitamin KH? vitamin K alkoxide glutamic acid WEAK BASE

vitamin K

Waffann

vitamin K apoxide rcarbozyglutamic acid Figure 2.1 Function of phylloquinone in vitamin K-dependant carbo.xylase. Vitamin K (PhQ) acts as a strong base in the alkoxide form of the molecule, removing a y-proton from glutamic acid. The resulting carbanion is then carboxylated to form y-carboxyglutamic acid. The resulting vitamin K epoxide is recycled. From Ref. (Furie et al.. 1999)

Vitamin K is not synthesized naturally in animals but extracted in the digestive system from consumed green leafy parts of plants. The function of vitamin K is significantly different in animals from in plants and photosynthetic bacteria, with the only known use being an electron transfer intermediate cofactor of PSI. PSI along with photosystem II (PSII). the cytochrome b(f complex (Cyt b(J). and .A.TP synthase are the protein complexes imbedded in the thylakoid membranes of cyanobacteria and chloroplasts of higher plants. PSI is the last segment of the electron transport pathway that reduces ferredoxin for the production of NADPH. Previous segments II in the electron transport chain oxidize water to dioxygen thereby creating a proton gradient which is ultimately used to produce ATP by ATP synthase. 3. The use of plastoquinone in living systems Plastoquinone (PQ. 4.5-methyl-3-isoprenylq-L4-benzoquinone), Figure 2.2. is a dimethyl substituted p-benzoquinone with a nine unit isopreniod tail that imparts a high degree of hydrophobicity. It is present in the inner and outer envelope membrane of chloroplasts (Roshchina, 1979) and within the thylakoid membranes of cyanobacteria (Tiemann et al.. 1979). PQ undergoes a two electron (and two proton) reduction to the dihydroplastoquinol by PSII. PQ links PSII to the cy\b(f complex. Reduced plastoquinone then diffuses through the membrane until it encounters the Cyt b(f complex binding site. This will be described in greater detail in section 6. It will also be shown in later chapters that plastoquinone mimics phylloquinone functions in PSI of the phylloquinone-less generated mutant.

9

o

Plastoquinone

Figure 2.2 Structure of plastoquinone.

4. Genes involved in phylloquinone biosynthesis The biosynthesis of phylloquinone has not been described before in cyanobacteria. However, the biosynthetic pathway of an analog compound menaquinone has been determined. Figure 2.3. The differences between the molecules resides in the "tair* with phylloquinone having a mostly saturated side chain and menaquinone an unsaturated alkyl group. Given the similarity of the molecules and that there is no known function for menaquinone in cyanobacteria, most of the genes describing menaquinone biosynthesis should also describe phylloquinone biosynthesis. In Esherichia Coli as with some other 12

bacteria, vitamin Kt (menaquinone) is used during fumarate reduction in anaerobic respiration (Unden and Bongaerts. 1997). Since several partial and full genomes are knowTi. there is now a method for doing a direct gene comparison among organisms. Also described are strategies used to choose which genes are disrupted in the phylloquinone biosynthetic pathway.

CH,

CH3 CH3 O CHj

Phylloquinone Menaquinone

Figure 2.3 Structures of phylloquinone and menaquinone.

In the biosynthesis of menaquinone (Figure 2.3), the 2-methyl group is derived from L-methionine (Threlfall, 1971; Kaipling et al.. 1984) and the aromatic ring and the C-4 quinone carbonyl have been shown to be derived from shikimic acid, a precursor of chorismate (Cox and Gibson, 1966; Campbell et al., 1971). The three remaining carbons (C-

1 through C-3) of the naphthalene ring are from 2-oxoglutaric acid (Campbell et al.. 1971). Both quinone o.xygens are derived from water, not from molecular oxygen (Snyder and Rapoport. 1970). The phytyl tail synthesis was thought to be generated from Acetyl-CoA through the mevalonate pathway to form geranylgeranyl pyrophosphate (Rohmer et al.. 1993). This molecule is then hydrogenated to phytyl pyrophosphate with NADPH as the contributing electron donor (Schultz et al.. 1985). The synthesis of the phylloquinone phyt\'l tail and the menaquinone isoprenoid tail is treated in greater detail in Ricardo Reategui's Masters thesis entitled "Phylloquinone in the photosynthetic activity of Synechocystis sp. FCC 6803" (Reategui, 1998). Given that both molecules are closely related in structure and synthesis, it is likely that the sources of atoms that comprise phylloquinone are similar to menaquinone. 13

The genes encoding the synthesis of phylloquinone have not been described in cyanobacteria. However, the biosynthetic genes of menaquinone have been cloned in other bacteria, such as E. coli. (Palaniappan et al., 1992; Sharma et al.. 1992; Sharma et al.. 1993: Daruvvala et al.. 1996; Sharma et al.. 1996; Suvama et al., 1998). It is likely that the phylloquinone biosynthetic pathway is similar to menaquinone. given that the difference between the molecules is the phytyl tail and isopreniod side chain, respectively. With this exception, the synthesis of the naphthalene nucleus in phylloquinone and menaquinone is expected to include similar steps. The genome database for Synechocystis sp. PCC 6803 (Kaneko et al., 1996) contains homologs for several genes that encode enzymes for menaquinone biosynthesis: menF {entC) (isochorismate synthase), menD (2-succinyl-6- hydroxy-2,4-cyclohexadiene-l-carboxylate synthase). menE (O-succinylbenzoic acid-Co A ligase). menB (DHNA synthase) and menA (identified as a "menaquinone biosynthesis protein"). Possible homologs of menC and ORF241 (the DHNA thioesterase) have also been identified in database searches. This proposed pathway is illustrated in Figure 2.4. This review will focus on the likely dedicated steps to phylloquinone synthesis: isochorismate to phylloquinone, involving genes menD, menC. menE. menB, menA. and menG. The menF gene is not likely to be a dedicated step to phylloquinone biosynthesis given that isochorismate also participates in the synthesis of and enterochelin (www .genome.ad.jp. Website address). Assuming that phylloquinone biosynthesis is comparable to menaquinone biosynthesis, the starting molecule for the naphthoquinone nucleus is isochorismate. Figure 2.4. The first committed step (SHCHC synthase) involves the loss of pyruvate and carbon dioxide with the attachment of 2-oxogluterate to the ring. The next two steps involve stripping of the aromatic ring of the hydroxyl group (o-succinyl benzoic acid synthase) and subsequent attaching a CoA to the aromatic carboxylate group (o- succinylbenzoate:CoA synthetase). The succinyl group provides the C-2 through C-4 carbons in the formation of the second aromatic ring which is catalyzed by DHNA synthase with subsequent removal of the CoA by a thioesterase (ORF241). At this point the naphthoquinone head group of phylloquinone is essentially formed. The remaining step involves attaching a phytyl group in the C-3 position (phytyl transferase) with loss of 14

pyrophospate and carbon dioxide from the pyrophosphate-phytyi group and carboxyiate. respectively. Finally the demethylphylloquinone is methylated the C-2 position (demthylphylloquinone methyltransferase).

HO 7^ ItannEntCMOBir pynnua COa 2-suceiny(-e-(«ydro9(y-2.4- ••ocnarumic ads cycWiMdian»-i-carfiaylate

H]0 SynVaM

1) MMBMIIIZZDHNAsyndnee n •iirniiiTmiiiuM-- Tun liiinumi

ZJOtWACoA o-«uodnyi banzoyl-CoA o-«ucciny< tianzaic aba

Phytyt-PP DHNA pnytyl nanfeiair MenA/Ur1S18

privfut ^metrryTpnyCcauincne WVtfivttTBn«4cf«9*f ' ManC/GwCZMISSS

S-«denotyl-t- S-adano«yH.- o mMiionin* homocydiina 2-pnytyt-l .4-napmnoqunona PliyHoq«anana

Figure 2.4 Proposed biosynthetic pathway of phylloquinone biosynthesis in Synechocvsiis PCC 6803.

The search for the proteins involved in the phylloquinone biosynthetic pathway in Synechocystis sp. PCC 6803 starts with the identified menaquinone producing genes. The genes encoding enzymes involved in the conversion of the chorismate to menaquinone ha\ e been cloned in E. coli. (Palaniappan et al., 1992; Sharma et al., 1992; Sharma et al.. 1993; Daruwala et al.. 1996; Sharma et al., 1996; Suvama et al., 1998; Wu et al., 1992) and Bacillus subtilis (Driscoll and Taber, 1992; Hill et al.. 1990; Palaniappan et al.. 1994; Taber 15

et al.. 1981). Specific genes involved in the pathway have been determined in spinach (Kaipling et al.. 1984) and Bacillus stearothermophilus (Koike-Takeshita et al.. 1997). -Menaquinone biosynthetic genes have also been identified by homology in a variety other genomes presented in Table 2.1: Haemophilus Influenzae and Arabidopsis Thaliana (wAw.genome.ad.jp, Website address). Each gene in the PhQ biosynthetic pathway has been examined individually for domain/active site homologies between Synechocysiis sp. PCC 6803 and four other organisms ranging from bacteria to higher plants. The phylogenetic tree has been determined for each of them, it suggests the evolutionary lies among organisms, see Figures 2.5-2.10.

Table 2.1. Percent identity and percent homology of enzymes of proposed hom.ologous gene in the phylloquinone biosynthetic pathway of Synechocysiis sp. PCC 6803 to menaquinone genes in four other organisms: Escherichia coli. Bacillus Subtilis, Haemophilus Influenza. and Arabidopsis Thaliana. Percentage Identity and Homology compared to Synechocysiis sp. PCC 6803 (%I / %H) Escherichia Bacillus Haemophilus Arabidopsis Gene coli Sublilis Influenza Thaliana menD 32/49 33/52 29/42 28/42 menC 22121 23/40 25/42 26/46 menE 25/42 24/42 27/40 29/46 menB 64/78 67/79 70/82 54/67 orf24I 28/42 25/45 * 25/43 menA 26/44 20/42 24/41 30/46 menG 30/50 33/51 * 32/44

* No homologous proteins identified.

As suggested by the homology comparisons with other organisms, the menD. menC. menE. orf24l. and menB genes of Synechocysiis sp. PCC 6803 are thought to be involved in 1,4-dihydroxy-2-naphthoate synthesis with the product of a menA homologue catalyzing the addition of phytyl chain. The product of the phytyl transferase, 2-phytyl-1.4- naphthoquinone, requires a methylation step to complete the phylloquinone biosynthesis. The gene originally identified as gerC2 (sill653. menG) in the Synechocysiis sp. PCC 6803 16

database codcs for the 2-phytyl-1.4-naphthoquinone methyl transferase enzyme that catalyzes this reaction (Shen and Bryant. 2000). In general, the phylogenetic trees in Figures 2.5 through 2.10. indicate in the biosynthetic pathway that Synechocystis (Syn) and Bacillus Subtilis {BSU) tend to be paired first. £. Coli (Eco) and Haemophilus Influeza {HIN) also tend form a first pair throughout the pathway. Those two first pairs {SynlBSU and Eco/HIN) are then usually joined before Arahidopsis Thaliana (Art) intersects. For many of the steps Syn is more closely conserved with other bacteria and Art shows the most deviation in protein alignment, for obvious reasons. Nearly all Art proteins are significantly longer than the bacterial proteins. This is probably attributed to Art containing introns. This reduces the homology of the Art proteins to the bacterial. Some discussions will be given to likely common active sites or domains within an alignment. The menD (SHCHC synthase) and menC (o-succinyl benzoic acid synthase) genes are the first two genes coding for enzymes that are likely involved in phylloquinone biosynthesis. The enzyme SHCHC synthase product is 2-succinyl-6-hydro.\y-2.4- cyclohexadiene-l-carboxylate (SHCHC) with the subsequent step of o-succinyl benzoic acid synthase producing o-succinyl benzoic acid. Both menD genes from £. coli (Palaniappan et al.. 1992) and Bacillus subtilis (Palaniappan et al.. 1994) have been cloned. The menC gene has also been cloned in both £. coli (Sharma et al., 1993) and Bacillus subtilis (Hill et al., 1990; Guest, 1977). Comparisons with Syn the SHCHC synthase show a steady 42-52% homology and slightly less at 37-46% for the menC gene product. Table 2.1. Protein alignments for SHCHC synthase. Figure 2.5, show a highly conserved N-terminus domain that is likely to be the active site of the enzyme. Both pairs of organisms (Syn'BSC and EcoiHIN') show additional homologies within each pair. This domain could be pan of the active site or a possible substrate binding site for 2-oxogluterate which is consumed in the reaction with pyruvate and carbon dioxide released as side products. Removed from Figure 2.5 is the first 400 amino acids on the N-terminus of Art. The o-succinyl benzoic acid synthase {menC) has no well defined identical domains that can be considered an acti\ e site. Figure 2.6. Yet, there are several sets of residues that are have complete identity towards the C-terminus end. Given that the reaction involves only a loss of a hydroxyl from the substrate 17

menO—flnt2 CTFUJS -TVFC UffCBT^kC RH fiMH PLTTCLRCf-P ER^ILRFhf! I 397 m«nO_BSU. t. F«_S 3 ITDfiU UC »QSRS TP . »UI_C nHH POISUHUQIp RGFIP F» . 3 63 menO^yn. t l_FRL S .QQflU c posns rURL HRH OCIDCUUSILp ER^RSFPFl q.3 65 menD-£co.t LTRI:H 3 Lirhic IP POSRS rp u rURR HEM SRFIHHTHFb ERblII .GHIIL F). 3 64 menO—HIM.t LURiG 3 J'SHLC IP POSRS rp . rUERURLQNR GRUTCHTHPP ERpi(.CFPtF) 67 U.R. ipeosasTPt r.RR n.H ....CH Pn-EBB-CFPtal 400

menO_fli-12 SLI 11iTseorn 5 EDFUPLU <447 menO^BSU.t .P < ^Ql u i citsomfvi H VSRUPI 11 I 13 men0.i^yn . 1 IG<^"TGi JR UVCISOTA^A iSi -I VSQVPLH 1 15 menO^co. t < ^Kl JR UIU1 SOTA ^ 3 LTCEKUII 1 14 menO.JH IN . t f^< ^tq; /R 1 lui ramn ^ QTGUHUF^ 1 17 feUR . .-.UPL. 450

men0^nt2 -fSjlN HFOSFURFFF r«.PFnrou IP -URflULTTUO SRLHURTCSR 496 menO-BSU.t Ih a LFONFUKFFT OSI EESPQ riLRV IRTLRS RRRGERQKRP 163 fn«nD_Syn . t ]3 TK UVOHVPQWQT EI.I PMTIO VCHVLRQTRL HSUQKCFUPG 165 menO_£co.t G 3 ^ 3 'C MFRSHPTHS I SCIPP o fiQOIP -RRULUSTIO HRL GTLHR 16 1 menD_HIN.t C 3 ^ 11. 3 3M riFGQVPURNU ML PI'K 9 HRDVS -RQWLISU-E QRMFQQKQQG 166 al . . FO. .P . . . . .LP. -R.L..T.. .R 500

menO—firsts RCpCbGSPTIi WSSHCLMG LOnunSNREP FTKVFQUQSH 544 menO-^SU . t 'IIH RE»L1P —OLSO—EP FGRMRTGRHU SUKTGTQSUD 205 menO^^yn. t J iJH DE »L ^LEHR RESLQHUREE FSKEHFVRCU TTFTTMNRGE 215 menO—Eco.t - 3 : JH i:p- RE n. I'CEMOD TGLSUQQR LGOUUQDOKP ULR-ERPRLE 207 menD—HIN.t • 3 H J P- RE I'DRTOE EUNSHSULQP LQRUL I QNKS U I MUERQQIiE 2 15 . 3 btci .... L. .U E 550

menO—Rr't2 KSOGU-TTGQ I TE t LQU IKE RKKGLLl. ICR IHTEOE1- -W RSI.1.-LRKEU 590 menO^SU. t RESLS OU -REriLRE REKCniUCGE LHSOROK- — EMIIRLSKRL 245 menO^Syn.t RISLJ>UHFSI FSFPSPQUOF SPt-GUlUVGU MPOOETPSU- TOIURIRRGI- 265 menO—Eco.t SEKQR-OUFF URQKRCUUVR CRMSREEGKK URLURQTUGU PI.IGDUt_SQT 256 menD_HIN.t ULJIHE'MUOH URTKRGUUUU GQLPREQRMG IMSURSRMGU ULLTDIQSGU 264 - U G. . . .G R. . . .U . . I L 600

menD_Rr-t2 MWPUUROULS GURLRKL^KP FVEKLTHUFU DHL—OHRLF SDSURMLIEF 638 menO^SU . t QS^ ILROPLS MLRMGUHOKS TU I DRVOSFL KOOELKRKLR POUUIRFGPM 295 menO-Syn.t GVPVLCORLC SLRMVODGQT RLITMVOFLI RCPRUREQLU PEQI IQ t GEL 315 menO-Eco.t GQPLPCROLU LGMRKRTSEL QQRQIUUQLG SSLTGKRLLQ UQRSCEPEEV 306 menO-H I ti . t UPTTPVEOI U LRMQTUREKL LQRO 1U IQFG RRF I SKRI MQ FLQRFK-GEF 313 ..P...D.L. ..R K F KR.L. GE . 650

menO—^^12 DUUIQUGSRI TSKRUSQ M LEKCFPF-RV ILUOKHPCRH DPSHLUTHRU 685 menO—BSU.t PUSKPUFLWL KDDPTIQQIU IOEOGGUROP TQRSRHnIHC NRSUFREEIM 345 me.nO-JSyn. t PTSKRLRHUL GTIOCPRVIL MFHGEMLDPL QGQTIVLSRS URQUREVIHN 365 menO_Eco.K M i UDDIEGRL OPRHHRGRRL i RMIROULEL HPREKRQPUC UEIPRLREQR 356 menD_HIN.t ULUEQSGKRL OPVHHSLTRF MRKUHHULRR HPPLRQKPUL LEPLRLSKFC 363 - .U L W. . . 700

menD_Rr-t2 QSHIUQFRNC ULKSRFPWRR SKLHGHLQRL DGRIRREMSF QISRESSLTE 735 menO_BSU.t RGLTRRTRSS EULEKUQFUM GRFREHLQTI SSEDUSFEGM LVRILOHLUP 395 menD_Syn.t QGFIPOREQK MVRMSU-LEK QR-OSQTI I I SRLRDRHTPL MUROLRHCLP 4 13 menD_Eco.t nORUIRRRDR -FGERQLRHR t COVLPEQGQ LFUGIiSLUUR LIORLSQLPR 405 menO—HIM.t RTF IEQQUGG NLTERSLRLR LPTLLPVNGU LFLGfiSLLUR LUDRLTQLPE 413 ... I L..RL..L.. 750

f menD_Ar-t2

5820 f menD—BSU.t CD 2131 m f menD—Syn.t m 2702 f menD—Eco.t rn 1396 f menD—HIN.t

Figure 2.5 Homology and phylogenetic tree of the menD gene. Abbreviation for organism: An2 is Arabidopsis Thaliana, BSU is Bacillus Subtilis, Eco is E. Coli, HIN is Haemophilus Influenza, and Syn is Synechocystis sp. PCC 6803. 18

A«nC.BSU2 ni E lEKlTLVHLS riMLKKPFKNS lETLOeflKFL 33 »«nC.Sgn t fO K PVHUICFVSVW FPLPKPUUTR HCWtyWRGL 33 MnC^IN t URLKIRJWFD RTRJLCftSHR EKSFMLVfiVS IPVCSQLILB ORFL»C«flEGL 50 AcnC^co. 1 nR SflOUYfUJO iPnORGUULR ORRUCTWDGU 30 -t t. n D »^GOLOLULH -LnMEflKFSV 2^ Consensus .V JJL. R-CL R«nC-BSU2 lUERlOTSCU TCUC£USRFS SPUVTEETIC TCLHnxOFF iPtmuCRCFN 83 A«nC-S^ t lUeUTOQRCn RGRC£tAPL- -PUFGTEDLfl SRTDFCftS-i- KCRITPRQIE eo ••nC-Hlh t lURU-S-CSR OGUCEIRPL- -PCFSETTLO QRQCORiCU. TTUCNRSCOfl ee •LLn-csr SSU-IUSEL- -GITRSSIFP SURCGLEHRL Lr«*V«JRHOS 67 Consensus lU L ouceisPL- -p F rr. 100 ••nC-BSU2 »#>SCUPOSLR RVKCrmRKR CLESRmAtV RKKKGUSLRC RLCGTROKUP 133 •«nC-Syn t KI COfiLPC-C CFRFOS-REU ELESCS-VIE r«PWSLSVCQ LL—PTCERR 125 ••nC-HIN.t PRUPLOGTVP SUfiPCISCflfl OCrtCGVLQRE CMVHTflPLCV —COPOCLV 143 •«nC_Eco I PQfl P SURFGUSCRL RELTOTLPQR ANVARRPLCM —GOPOOLI 118 ••nC-Rrt t SLL C ILHVQICEENC SROPHSUQIC RLLDSECTPL EURVUR 107 Consensus RFC \ LC 150 ««nC.BSU2 RCUWCLRPU OOnucEIESV QKEGVQRIKI KIOPCQOUEL UKRIFIRSl^l 'PT 183 A«nC.i.S^. t LDVLRSr^OL CRSTF»aJ-KI GUCOFR-UER OLVOQU*«. PECCQI 173 tt«nC^ I n. I RKLRSneCOC UflWlCUO-tV ERhR 00 LIROrrUERI -POLOLOL 187 •enC-^co.1 LKLROnPCEK URKVKUG-LV ERUR OC nUUHLLLERI -POLHLhLOR 162 ••nC-ftrt.t RKLUQ-eOFS RIKUCUCRRU SSUO DR LUKDEURRRU GUOIEL 4 V)R 152 Consensus .KL. . . C. PK. KUG- V . 0 . . . R. 200 a«nC,.BSU2 IPLnRORnSS VEUCOiSRUC ELOOVHLnni Sfl QROOIU 0HRKLX9CHLK 233 ••nC-Syn I rCGLSUOim OULTLLOROR ensgpkiqvl E: » ppgeio EnFRLQEOFR 223 A«nC-HIn t rAHUSLEKRL O-FRRKUICLO —RIQFL Ei'»Da(QRLS REFRUOTDIR 234 •enC^co. t rmuTPLKOQ O-FRKYUf^ VRO—RIRFL Ei»X:troos RRFRRerCIR 20« •enC^J^t. t. rCRUTFEERR E-FGLLUM— -SC—ttJCYI XlhKOOL IRFHEETCLP 196 Consensus M. U. . R Q-F. y. . —.1 .L .0. F OT.-fi 250

•enC.BSU2 TR—ICLDES (CStJOORRRR lELGSCKIlN ( TPWJfSpL-T ERLKIfCLCK 281 ••nC_S^. 1 TS—IRLOER UCSLRQLRQL VDKGUSOLVt i

menC_BSU2 1738 menC^yn.t 2206 4 :: menC-HIN.t CO 840 menC-Eco.t 1542 3 menC_Art.t

Figure 2.6 Homology and phylogenetic tree of the menC gene. Abbreviation for organism: Art is Arabidopsis Thaliana. BSU2 is Bacillus Subtilis, Eco is E. Coli, HIN is Haemophilus Influenza, and Syn is Synechocystis sp. PCC 6803. 19 without a cofactor suggests that a smaller active site is plausible. Again Art has had the remaining 300 residues removed from the C-terminus end. The only change between protein alignments with and without the truncated protein is seen in the phylogenetic tree, with Art aligning more closely with Eco/HIN. To determine if genes are coding for enzymes in the phylloquinone biosynthetic pathway, the menD gene was mutationally disabled (Johnson et al.. 2000a). The menD gene was chosen over the menC gene because it had a higher homology and identity between Syn and Eco, Table 2.1. The largest average enzyme in the pathway is o-succinyl benzoaterCoA synthetase which is encoded by menE, Figure 2.7. The homology comparison to the other organisms is low compared to the other protein homologies in this pathway. Unlike menC which also has a low homology, in menE there is a C-terminal region of identity that could be the active site domain. Many of the other residues in this domain have more highly conserved hydrophobic side chains and possible kinks or turns in the 350 - 400 amino acid region of Eco. The enzyme performs a relatively complex task of attaching a CoA while utilizing the energy of hydrolysis of ATP to AMP. This suggests that there may be several domains that perform different aspects of the enzyme function. The Art protein is again longer than the homologous bacterial proteins, but only by approximately 100 amino acids. Given that in Art additional amino acids are contained in 4-8 residue stretches within the alignment of the other proteins, these additional residues could be additional loops or helices contained within the protein or possible introns. From the identification of this gene in both Ecu (Sharma et al.. 1996) and BSU (Driscoll and Taber. 1992), the menE gene was chosen for gene disruption in Syn (Johnson et al., 2000a). The menB gene product (DHNA synthase) is the most conserved protein within the biosynthetic pathway of phylloquinone. Table 2.1. Art has the lowest identity at 54% w hich possibly can be attributed to an intron in the N-terminal region. Figure 2.8. The overall homology of bacterial proteins is 80% and upon visual inspection the homology Art is similar, once the excess residues (possible introns) are removed. The C-terminus half is ver> highly conserved. Given the cross organism homology, the second ring closure can be considered a critical step turning a substituted benzene ring into a naphthoquinone. DHNA synthase is the only step that does not produce byproducts or consume energy as it 20

•«?•£-£ CO FSOUP« (Jjnce Tl- -fl Lfl—uecQL 28 ••r^co t «flrT-y»OCO PL£»1 U « •«rO*in VP-4JQ0P fliQ—PCFSO Kl- Lft-TTOCOTL 27 t flCLH-^FEDO FVRSL L ICT ••re..ssu —LTEQPfM. rOW-OLTPf: R»- HI LIV-€OCTVT » mr£JSJ t rtTW-LK3RF SWStX>—L ncum: -sn QfTCTGryOL Lfep-<1RIRS flU- -€ FSLHOCLCI 34 t CJ^S-fLOVS LCKOS ft PPFP- POWT tnrmxSF nVLSLirVA. rtttSVAlLSS PLLPA5»«>n5 TSFS^CFPF 50 Mr£_A-U ftSLUCFLPKF SUSCIimA E5VTMOEKT tCSl L -Q I «L - C .

*T£_gcO uoeuscFnu O-CUOCDSCU ruwhnrpor L L 67 ••f£_Eco t UlfSS USL IP U€LTEO-«£ QClRCri 270 T—UtELTTK IMQTVWFU* IC-GW«S«U flFUCWiSBtl L r 66 MrC_Hin t —CMXC^ ISFHT K Hli IP TELTQW-fUC VClETVSl^'C 25^ *«r£«Ba; F—flELfflfiS iCFnRCQUWI SUWCCOTRHJ LLH»#WEnu V R 69 ewC-BSU t CETW «C—Ci: LPLLEE-OC ICCFP«JF« 287 »«nC_S«/^ L--«CLPLD UQtlCETrCQ 9.TGIE5PQI LUCTMUOCF U A 74 mar£^£i^ t OnC m LXTTR Lll PflLLAK-^ ftSLM^li^271 FOSTM^SSQ SGSUCUTO RFSEASr«CS RIAIKK0CS VSVGQLTSSA <00 MrC_fr-t2 vovocsnoo Gflnrnnnuji Rpurvocsi TMJ.£7fti

»vi£_£co t POJV—<-L0 CC WIULP W^j-pqpu. EELLPrt-TL 100 wC-£co I Frergc- —«cfCCL CUKHW cm 306 •«r£jth • LYLfi—no LG »CJLC ir*W>®Cl flKLCETM]! 99 t jwsTwr fVKSOnc •cn EvcLuro eiu 300 »er£_BSj t ^«*:~FLL- —CUK T»(LST>€FLF -OLEOSCSGF 103 •w£_SSU I TCSQIVT LSPEFSICCL SC CllCIEnOCQV CEPvEHGCin 337 •«r£_S«^ t UGfi ——COFl flrmXNPCU OOUWFLQPO l09 t rmOiAT LTPfCFLACQ XTiTroo)- -cnufiPCTyc 343 U^ISkLFLKC OrmCCOETK ICVCCFQSLKC RRIGIIJ<«(PS REFOnCULST 150 Fvnm- LADVm EnciKEQEn* -OAhOJCeiC 441 A—PI, G n LL. M 0 .L ISO tftS . - - Pfl e Ki r« - - Gci *ic

•4<-£_£co t QP-4^

•er£^eo t F?XSnTL-TS OSTGLP KfWUMT V Oft H.ASA—-Q 164 MrC^Eco t QUFiU •«r£_Hm t TRTL-TS OSTCLP ICfMJW U——QH »1M K 154 MrCJiiri.i -niQ KZ ••r£_BSJJ t flTUf/TS GTTOP K-GUQO TT G-f« V^S«U S 175 MTC-BSU t -«iv LiCFPTC OSSGTI RFflLMS PQ TLTH MJOGL K 183 •arC^^ t •^10 PEO ••r£_^«rl2 RHtlPPUWI STSETUBCMQ FQDOSFEFEG KFLDOPfCIU VTSCTTGKPK 250 nr£J^U U ;>TLL£ »ff> IIflCSA TS OSTC P K . UH K 250

GULS-CrPFG GiriMI.VfC 200 MrC.£co t TLPP ELJCMCCI GUOl-U*FO GIUUM.VCO 190 MrCJ^m t n£K&-l »er£,BSU SfiLn—U3IT EODR- •tr£JSU t OFLGPLLPflM OQI •anCJS^ t Cl^JWTWlSiN SOUlrtTEflU EVTSIOrLH MrCJ^t2 RKKfVEDESX PUI GW. Q U L

1339

3401 menE_BSU.t 2129 3131

Figure 2.7 Homology and phylogenetic tree of the menE gene. Abbreviation for organism: Art2 is Arahidopsis Thaliana. BSU is Bacillus Subtilis. Eco is £. Coli. HIN is Haemophilus Influenza, and Syn is Synechocystis sp. PCC 6803. 21

t. t ^OSMELGSfl SRRLSUUTNH LIPIGFSPflR flOSUBLCSflS SflOORFHKW 50 nenB_6SU.t A£ 3 Acne...Syn. t 2 Attr3-H I rt. 1 iJMPKO —OULVfiP— 14 T n«r\6—Eco t 1Ivpoe- —flflLVRP— 14 Consensus — . .P— -ue- SO

•i«n8_^rl. 1 GEUPTHEUcCr OCTOFFCeCO NKEFUC fHl KRLiaiHKi inir«Anf 100 «)«n6-BSU. t -KT KRTV06 •t T VH- 5 lAKI innp£u» 34 AttnS^^. t +41 RKHVOC K--flG- 3 lAKl 33 VkenB-H t M . 1 DH SEGVSC KSTI ilflKt 47 Tfn«n6-£co. t OC SEGFEC Z KSTIO- 9IIMCI ir«P3uf 47 Consensus v oia ^ K. . D-ClfllU-I1 100 nenB^r t. t K sv ^KflFCSGCO (J^TQCXSvn 150 •t«nB-BSU. t pt9H- -SOCK st^C$CCCX}S 80 m«n6^ I N . t p^®3rvcE*i 7^?SCMf 3 1 Ni 3£K- ^CSCCOQK 89 Tni«n6-Eco t nj(iiI dRLfldwl 3) m iLr3^ 3DK- 'CSCCOQK 89 Consensus eF.OBa.DL Ml . JrcR 31. K- •CSGGDQK 150 8>enB_^l. t OPfO' .NU LOLQUQl I flriLUF EUOLFOFGLS m«n6-BSU. t. URIOHC 3- V UG-OOOiPl lULOLQRLlR UtP—KPUUft nenB^S^n.t URGEi V ID-OQGTPI 1UL0LQRLIR Snp—KUUIft m«n6—H i M.t UROO'ivfif—GV KO-OSC< IULOFQROIR SCP—KPUVR TmenB^co. t URIlOOV 3 —«OV KO-OSGiUHI- . 1VL0FQRQIR TCP—KPUUR il jQO Consensus UROO jH- .V .0-0.G. , «^.gRLIR ..P—KPUMR

menB^r t. t TTOTTSSROBSFDh 9^ t1^. A ELTITftno 250 •tenB-BSU. t FOQ TGPKUOSFD 4 3 > 5 5C'iVLf ^ I II 5H 163 •tenB—Syn t L :OL.T IFOQ TGPKU0SF0 3 3: SSSVLiM J 50 167 menB-JH IN. t IFOQ TGPKUOSFD > J fO- 177 TmenB.£co.t IFCQ TOPKUOSFO i J ^ 177 Consensus HBmRimn jaaajoBEDD ICQ- 250 •i«n0_Arl. t VPKSMKHRUI KUVECIQUGP nenBJBSU.t a^rn m«nB_Syn.t vtenB^ IH. t TnenB—Eco.t Consensus menB—fir t.t CMSPT ni«ijewajrT*^5Cl 3LG 5 nen6—BSU.t r flnx DFP 3 nen&^yn. t 4 R DC 3CC menB^ IM. t RDC 3C TmenS^co. t flscx: Consensus

•lenB-Rr t. t TflVWlflffBTbKFHRFp 378 nenB-BSU.t OSTKEK q ( »0 F »FPRF » 271 *»en6-Syn. l / - r i E 3S EO C ORFLEK ^:PO F >QVPUL ^ 275 nenB-JH IM t V - r; E » EO) NRFHEK H • PO F>KFRRfi» 285 Tmen6-^co i 1 r ! E 3G EG ) NArNQ»M|>0HQ»< V»0 FSKFKR^» 285 Consensus raBEb.Bw RF .E«ea.EEKF R.e 378

menB-Art.t

1367 menB—BSU.t 13 409 menB—Syn.t 504

menB_HIN,t 136 I 1 I TmenB-£co.t

Figure 2.8 Homology and phylogenetic tree of the menB gene. Abbreviation for organism; Art is Arabidopsis Thaliana. BSU is Bacillus Subtilis, Eco is E. Coli. HIN is Haemophilus Influenza, and Syn is Synechocystis sp. PCC 6803. functions. This suggests that the enz>'nie mechanism is quite refined, hence imparting a high degree of homology. DHNA meets all the requirements of fitting into the Ai site, except the tail. For these reasons, this step was the first chosen to be mutationally disabled in Syn (Johnson et al.. 2000b; Reategui et al.. 1998; Zybailov et al.. 2000). This disallows the production of the head group of phylloquinone. By disabling the function of DHNA synthase there are two possible outcomes for the Aj: remain empty or find a substitute molecule to act as an electron intermediate. Interestingly, plastoquinone occupies and functions within the A| site of this mutant. All gene disruptions (menB, menD, and menE) in the phylloquinone biosynthetic pathway of Synechocystis have the same result. Given the homology of the genes and that plastoquinone is a possible natural rescue molecule, fijture analysis of the phylloquinone biosynthetic pathway in higher plants such as Arabidopsis suggests that the menB gene is a logical place to start. Given the little knowledge gathered concerning orf241. we will only really acknowledge that it is probably in the menaquinone/phylloquinone biosynthetic pathway and that the homology is relatively low among organisms. It is thought to be a thioesterase which releases CoA fi-om DHNA-CoA. The DHNA phytyl transferase encoded from menA in Syn. is clearly an important step. WTiile low. the homologies are surprisingly similar given that menA product in Syn is transferring a mostly C-20 saturated phytyl tail and in other bacterial organisms the menA product is attaching an unsaturated C-40 group to naphthoquinone. Figure 2.9. This step differentiates the production of phylloquinone and menaquinone. The synthesized DHNA meets all the requirements of fitting into the Ai site, except the tail, which led to the author's question, "was the tail necessary for Ai fiinction or was the naphthoquinone head group the only requirement?' Therefore, this step was the first chosen along with menB to be mutationally disabled in Syn (Johnson et al., 2000b; Reategui et al., 1998; Zybailov et al.. 2000). It was found that DHNA was not incorporated into the Ai site, instead plastoquinone was utilized similar to gene disruptions earlier in the pathway. Therefore, the head group and alkyl tail are both requirements for stable interaction. The menG. methyl transferase, is the most illusive of the genes in the pathway to identify decisively. Methylation is a common reaction in cells and is performed by many nIGCRSCUGW L'^URftOU— flTCfWUPflKSKENSLUnCIC QOOTUMLMLR 4€ »«nfi^co2 riTEQQ I —SR TQ«-WLES— —URPKTLPU flFRftI IUOTR LRU-UQGH— 4C I«.t URSKKnihEK LKM-UUET— —flRPKTLPt- RLRSIFTGSR LGV-URfiPQG 44 m*nA_SSU. t nHQTNKOeCQ TflP-QKESnC QIUJOLTRPH TLTRSRJPUU LGT-VLflriFV 4€ (••nA^gn t nressPLflPs TfiPRTRKLWL ftfllKPPnVTi;RUUPITVCSR URVGUTGQUH 5C T E — . BP.TP. fl-..I.UG.R 5C

m«nR«Rr t.t QESBKPI^^ TLFEW-ROO LQR-UlOHLl. —SlUORENP ULlS- 8€ »«nfl_Eco2 FD-PLURURL - LITRCL-t-QI LSr*-LRfCVC -ORUKCSOKP onic- 8C m. t FH-CUO U-TTIL-LQU LSN-FRfCVC 'OHQKGSOTE ERIC- 84 ri»«nR^SU . t mCUOLLL rtLFSCLWiQr RTrt-LFMEW -OFKRGLOTR ESUO- 9C Nk^nR—S^n. 1 GOUFTI RlRllRUtML SMOUTOSOTG lOURKRHSUU MLTGHMLUF IOC L. LL .L. -L- Q L.N-L.MOVC -O.-KCO. IG- toe

n»«nfl-Rrt t RREQ—IFSR GGICRnRPGLU FLUSRRTREL RGLKELTUEH RRLGEI lEHI 13C «»«nR-Eco2 PLRC—HOKG UlTQQEnKRR LIlTUULICL SCLRLURUflC HTLROFUGFL 12€ n«nR^ i M . 1 PLRG— IQKG RISRKEUOIC LI LfKJnRSFU SGSFL IGIRV EliLSOLFRFR \3Z m«nR.BSU.t IGGR—lURH GrtCPKTILOL RLRSVGlRlL LGUVtCRSSS UULR-LIGUU t3T ShenR-Syn.t. LISHFFU-RG ULGLftSnSWR RQDUTULEUI GURIFLGVTV QCPPFRLGVU 15C . — I . G U G L. . . G 15C

•enR-Rn1. t, HTRSLIWOU LOeSDnRRGR ETUHELFGTR URULRGOFHT RQRSUVLRTt. I8C

»«nR_Rrt.t EMUEUIKLiS QUIKOFRSGE IKQRSSLFOC i*«nR_Eco2 HTUIPRLI LP RTRCCLLRTR ULMlNHLROt (MnR^ i M . t KSIOSHIiLP RIGSGLLRSR ULMINHLROI MnR.3SU. t OlCIMnOSiLI SIPIRILUGR INLSfffilRO •«nR«S

»enR_Rr t.t RSTKGRRIFS KUESKUREOn VOIiFfSkKLl OLS FQUUOOILOF TQSTEQLGKP 28e m«nR-£co2 NGKHTLW— -RL 3 EUNRRR VHRC 231 n«rtR_Hm t RGKMTLRU— -RL 3^1:ORU VHCI 23f MnR^SU. 1 GGRKTLRI— GRUT LLRR 23t ««nfl—Syn.t RGKICSPIU -RL. BtKlKLCSQ ULTL 23C ROCICTLRU 0 . 30C

ncnR^Rrl.t RRNOLRKGNI TRPUIFRLEN EPRLREllES EFCEPGSLEE RIEIURmOG 33C ii>«nR_£co2 —LLTIGSLU CL-RLFHLFS LHSLUGUU- -FLLRRPL-L UlCQRRVUnRE 273 »«nR_HIN.t LLSURRU CV-LRFRURT RISUTNVL- -FULRflPL-L RKHRIFUVCS 277 m«nR_BSU.t SFRURVI UU-OGLUITG RRS—PUL FUUFLSUPK PVQRUKGFUQ 27C m«nR-Syn.t SUUSLVL IT-RI GULCH QRPWOTLL- -IIRSLPURU QLIRHUGQVH 272 - -F.L. -F. ..PL. 3SC

menfi^r 1. t IKKRQELRKE KRELRLKNLM CLPRSGFRSR CfeOnunFMLE RIO 37^ ««nR_£co2 nO-PURnRPtl LERTUKGRL- LTML .GIFLSQ URR 30£ Ih . t QQ-PSELPPI LROnsniSL LIMl "SLGLLIG 30€ »«nfl_BSU-t NErtPrtMniUR nKSTRQTPfT- - Li.SlGLLlS- VFR 31 1 »«nR_Syn.t DQ-PEQUSMC KFIRIJMLHF -FSGn 5VCUR- GLG 307 292

menA^Art.t

men A_Eco2 m 772 GD 2217 men A_mr*4.t 1259

men A_BSU.t 1597

men A-JSy n.t

Figure 2.9 Homology and phylogenetic tree of the menA gene. Abbreviation for organism; An is Arabidopsis Thaliana, BSU is Bacillus Subtilis. Eco2 is E. Coli. HIN is Haemophilus Influenza, and Syn is Synechocystis sp. PCC 6803. 24

enzymes on a variety of substrates (Voet and Voet, 1995). Finding the specific enz\'me responsible for demethylphylloquinone methylation by homologies would be time

consuming and not necessarily conclusive. It is also possible that another enzyme may perform the same function. The homologies amount organisms are comparable to other genes. However, no comparable domains for the possible active site are found. Figure 2.10. Yei, in Syn the gene, sill653. has been identified and mutationally disabled (Shen and Bryant. 2000). The result of this mutant is that only demethylphylloquinone is produced with a complete lack of phylloquinone. This confirms not only the gene responsible for this enzymatic reaction, but that there is no alternative pathway for demethylphylloquinone methylation. Demethylphylloquinone is apparently utilized in PSI in place of phylloquinone and preferentially over plastoquinone. However, demethylphylloquinone in the Ai site reveals some kinetic differences have been observed and are being fully explored (Shen and Bryant. 2000). Homologous genes involved in phylloquinone biosynthesis of Synechocysfis sp. PCC 6803 have been identified. The regions of homology within organisms vary significantly among genes. Several of the genes in which high homology between Synechocystis sp PCC 6803 and E. coli, which have been cloned, have been disrupted. Five genes {menA. menB. menD. menE. and menG) of the seven in the pathway have been have been disrupted. 5. Physical description of Photosystem I Detailed biochemical, biophysical, and molecular biological studies have shov\Ti that the PSI complex of cyanobacteria is structurally and functionally equivalent to that of higher plants (Golbeck. 1994; Chitnis. 1996; Nexhushtai et al., 1996: Malkin. 1996). However, the greater stability of the cyanobacterial complex (Tsiotis et al.. 1993). an improving X-ray structure that has reached atomic level resolution (Schubert et al.. 1997; Krauss et al.. 1996; Klukas et al.. 1999a; Klukas et al., 1999b). as well as the relative ease ol" genetic manipulations in cyanobacteria (Thiel, 1994) are reasons for studying the cyanobacterial model complex. The fundamental organization of most of the photochemical electron transport components is similar in the Type I (PSI, green bacterial RCs) and Type II 25

Syn_menG .t V FS flTKT PIKPfifiM 14 BSU_nenG.t nQOSKEQRUH GUFEKIVKNV DQMNSUISFQ QHKKURDKTn RIMMUKEGflK 50 Eco I i menG M N REKS 6 flrab_menG. nVNFHSRGDV SOKSULMnSP ESLflFPSDVQ HLLCSSRGEN RUSDUFGSOE 50 Consensus M 50

Syn_menG . t GX WT-RLflNPTS R-RNCTCLG PUG-K RSK 40 BSU_nenG.t RLDUCCGTHO UTIRLRKHflG KSGEIKGUDF SEmLSUGEQ KyKDGGFSQI 100 Ecoli menG P-U NHMUDH 14 flrab-menG. LLSUflUSRLS SEHflSlflPEl RRMDDMUSLT UIKRKIRCHP SVPRLLQRVI 100 Consensus R. . .P 100

Syn_flienG . t LGCR—X RVSSU NR H 53 BSU_nenG.t ELLMGNHHEL PFDDDTFDVU TIGFGLRMUP DVUTULKEHR RUUKPGGQUU 150

Flrab-menG. DCQKUGRPPE IRCLLEEIQR ESDUVKQEUU PSSCFGRDPE LDEFtlETVCD 150 Consensus 150

Syn_menG .t CLRMGTTTL R—RVQ RXKLR—RRR LPGUXK—SR MRTF LPG 89 BSU_nenG.t CLETSQPEMF GFRQRVFnVF KVI flPFFCKL FRKSVKEVSW LQESRROFPG 200

flrab-menG. ILUKVKSDLR RPFDERTCFL NKIEMOLRNL CTGUESRRGU SEDGUISSDE 200 Consensus . L 200

Syn_;nenG.t LSTRCNSySfi WGILPTFRRP XPIURPSMSL SEKCNflRflSP CSHCnOKURV 139 BSU_MenG.t HKE LRGL F EERG 212

flrab-menG. ELSGGOHEUR EDGRORCEDR DLKDRLLRKF GSRISTLKLE FSKKKKKGKL 250 Consensus 250

Syn_jnenG.t LLLCFPRRIR NNSRQKSSPT TFPLyPTTFR PUKRKSPLPQ HTSNRRSPGO 189 BSU-lienG. t LKNUKV HS 220 EcoIi menG WUDLEGEFKP LHRIHPLRLG VIRERRGGLF GKKULDUGCG 66 flrab-menG. PREflRQRLLD UUNLH—VK U—P VPTE GDKI 277 Consensus P 300

Syn-jnenG.t TPLHFTHnflF 01UCUPKUKN ULRSSVUGRH RRKICCTLSX UGWUNRELL 238 BSU_nenG.t FT GGU RR TH IGUK 233 Ecoli menG GGILREStlRR EGRTUTGLDH GFEPLQURKL HRLESGIQUD VUQETUEEHR 116 flrab-menG. —RLRDfl TGLD QK QIN 291 Consensus L fl 349

Syn—menG.t 2410 e- BSU_MenG.t 3020 f~r

Ecoli menG 3641 Ar.ab—menG.'

Figure 2.10 Homology and phylogenetic tree of the menG gene. Abbreviation for organism: Arab is Arabidopsis Thaliana, BSU is Bacillus Subtilis, Ecoli is E. Coli, and Syn is Synechocyslis sp. PCC 6803. 26

(PSII, purple bacterial RC) reaction centers (Schubert et al.. 1998). A comparison of the structure of PSI (presently at a resolution of ~3 A (Surrmiers. 1980) with that of the bacterial

RC (at a resolution of —2.3 A (Lancaster and Michel, 1999)) shows that the organization of the primar>- donor (special pair), accessory Chls, primary acceptor, and secondar\' quinone acceptor is arranged in a manner similar to the equivalent cofactors in the purple bacterial. Figure 2.11. PSI is a membrane-bound multisubunit protein complex that catalyzes the photooxidation of plastocyanin in the thylakoid lumen and the photoreduction of ferredoxin in the cyanobacterial cytoplasm (or chloroplast stroma) (Manna and Chitnis . 1998; Malkin. 1996). Purified cyanobacterial PSI complexes contain at least eleven proteins: core proteins (PsaA and PsaB), six integral membrane proteins (PsaF, Psal, PsaJ, PsaK, and PsaL), and three peripheral proteins (PsaC, PsaD, PsaE). PsaM is an integral membrane protein found only in cyanobacteria. There are three other proteins contained in chloroplasts: one stromal peripheral protein (PsaH). one lumenal protein (PsaN). and one integral membrane protein (PsaG). The genes and characteristics of the PSI proteins are shown in Table 2.2. The PsaA and PsaB proteins form the hydrophobic core of PSI. The proteins are similar in mass (-82 kDa) and homology. Because of the similar sizes, the PsaA and PsaB proteins usually co-migrate as a diffiise band in a SDS-PAGE gel. The primary sequences of the proteins from various organisms contain higher than 78% and 76% amino acid identities for Psa.A. and PsaB. respectively (Sun et al., 1998). A direct comparison of PsaA and PsaB primar\' sequences to each other shows a 42% amino acid identitv^ and an additional 15% conservative replacement (Sun, 1999). PsaA and PsaB may have evolved by gene duplication, this is supported from the similar structural features. Structurally. PsaA and PsaB have a total of 22 transmembrane helices, arranged around a psuedo-2-fold axis of symmetry (Schubert et al.. 1997; Krauss et al.. 1996). The cofactors involved with electron transfer from the primar>' reaction center P700 to Fx are located in ten transmembrane helices within the central region of the Psa.A/PsaB subunits (five helices from each subunit). see Fig. 2.11. The helices farther away provide stabilit>' to the central core of helices and interact with other PSI subunits (Xu et al., 1994a; Chitnis et al.. 1993a; Chitnis et al., 1995; Chitnis et al., 1997). 27

Table 2.2 Polypeptide subunits of PSI. The molecular mass is determined from primary- sequences of Synechocystis sp. PCC 6803 and apparent mass observed from PAGE is in parenthesis. PsaG. PsaH, and PsaN molecular masses are the range from higher plants. Subunit Gene Mass Properties Functions Psa.A. PsaA 82.9 11 transmembrane, 3 stromal and I lumenal Light-harvesting; (66) surface helices each; together, they bind ~100 charge separation; PsaB PsaB Chi a, P700 Chi a dimer. 12-16 b-carotene, 2 electron transfer 81.2 PhQ, and one [4Fe-4S] cluster (F.\) photoprotection; charge (66) separation PsaC PsaC 8.8 Peripheral on stromal (cytoplasmic) side; Terminal acceptors; (8) binds two [4Fe-4S] clusters (FA. FB); 2 donates electrons to single-turn a-helices ferredoxin PsaD PsaD 15.6 Peripheral on stromal (cytoplasmic) side; 1 a- Ferredoxin-docking; (17.7) helix assembly of PsaC; normal EPR properties of FA and FB PsaE PsaE 8.1 Peripheral on stromal (cytoplasmic) side; fa- Facilitates interaction (8.8) barrel structure with ferredoxin; essential for cyclic electron transport PsaF PsaF 15.7 2 transmembrane helices; large exposure on Plastocyanin-docking (15.8) lumenal side; may bind 1-3 Chi a for fast electron transfer stabilize core antenna? PsaG PsaG 10- 2 putatative transmembrane helices; only in Interaction with LHCI 10.8 chloroplasts PsaH PsaH 10.2- Peripheral on stromal side; only in O 11 chloroplasts Psal Psal 4.4 1 transmembrane helix Normal organization of (3.4) PsaL PsaJ PsaJ 4.5 1 transmembrane helix Normal organization of (3.0) PsaF PsaK PsaK 8.6 2 transmembrane helices; may bind 1-2 Chi a Interaction with LHCI; (5.1) stabilize core antenna? PsaL PsaL 16.6 2 transmembrane helices; may bind 2-4 Chi a PSI trimerization; (14.3) stabilize core antenna? PsaM PsaM 3.4 I transmembrane helix; not in chloroplasts Stabilize trimer in (2.8) cooperation with PsaK? PsaN PsaN 9 Peripheral on lumenal side; only in 7 chloroplasts

Of the remaining membrane bound subunits PsaL is of note. The function of PsaL was determined by inactivation of the PsaL gene. Trimeric PSI complexes cannot be obtained from the PsaL-deficient mutants of Synechocystis sp. PCC 6803. hence PSI is in the monomeric form. PsaL is also accessible to proteolytic digestion in the monomers but not in the PSI trimers from wild type membranes (Chitnis et al.. 1993b; Chitnis and Chitnis, 28

1993). It was concluded that PsaL is responsible for the formation of trimers. which has been confirmed in another PsaL-less mutant of cyanobacterium Synechococciis sp. PCC 7002 (Schluchter. 1996). Typically the analysis of purified PSI comple.xes is done only on the trimers. since this form is what is commonly found in vivo.

FB FA 22

FX 14 14 22 24 8.S AO 26

10 17

P700

Figure 2.11 Arrangement of the PSI cofactors. Distances are depicted in juigstroms. The cofactors are identified as the primary donor P700 (a Chi a dimer). the primary acceptor .Ao. (a Chi a monomer), phylloquinone QK, and the three iron-sulfur clusters. Fx, FA. and FB- From Ref. (Klukas et al., 1999b).

The core proteins PsaA/PsaB bind and contain numerous cofactors. The electron carriers from P700 to Fx and 12-16 5-carotenes are bound within the heterodimer. The P700 reaction center is a dimer of Chi a and a ' molecules with a histidine from each subunit coordinating to one Chi a (Fromme et al., 1994). There are two symmetrical Chi monomers identified as the Ao primary electron acceptor, adjacent to P700. The phylloquinones in the secondar\' electron acceptor site A| have been identified in the cr\'stal structure (Klukas et al.. 1999a) and orientation of the quinones further was refined by EPR techniques (Bittl et a!.. 1997; Van der Est et al.. 1997). The final electron acceptor is a [4Fe-4S] iron sulfur cluster Fx that is located on the psuedo-2-fold symmetrical axis on the stromal side. Fig. 2.11. It is coordinated by both PsaA and PsaB with conserved cystienes. The P-carotenc molecules are not well defined within the crystal structure of PSI. 29

The PSI complex carries its integral core antenna system consisting of-100 antenna

Chi a molecules. In the crystal structure of PSI at 4 A resolution. 89 chlorophyll a molecules have been identified (Schubert et al., 1997). Six of those chlorophylls are cofactors of the electron transfer chain with the remaining 83 constituting the core antenna system. Most of these antenna chlorophylls are ligated to histidines from conserved amino acids within PsaA and PsaB. It is also thought that PsaL, PsaK. and PsaF may also coordinate antenna chlorophyll (Schubert et al.. 1997). The overall organization of the peripheral subunits attached to PsaA/PsaB shows PsaC oriented on top of the C2 psuedo-rotation of axis of PsaA/PsaB with PsaD overlapping

PsaC towards the PSI trimeric C3 axis and PsaE also overlapping PsaC away from the PSI trimeric C3 axis. Figure 2.12 (Klukas et al., 1999b). Little of PsaC protein seems to be solvent exposed, being surrounded by PsaD and PsaE. PsaC is a prerequisite for PsaD and PsaE binding (Mannan et al.. 1994; Yu et al.. 1995b). The peripheral protein formation is not necessary for stable PsaA and PsaB assembly in cyanobacteria. but is in other organisms like C. reinhardtii (Takahashi et al., 1991).

Figure 2.12 Stromal ridge of PSI shown form the stromal side onto the membrane plane. PsaC is depicted as light grey, PsaD in vertical lines, and PsaE in horizontal lines with the proposed ferredoxin binding site being the diagonal lines. Adapted from Ref. (Klukas et al., 1999b) 30

The peripheral proteins PsaC, PsaD. and PsaE are critical in shuttling the electron from Fx in the PsaA and PsaB heterodimer core to soluble ferredoxin. PsaC contains the two terminal electron transfer [4Fe-4S] iron sulfiir clusters FA and FB (YU et al., 1993; Antonkine et al., 2000; Yu et al., 1995a). FA and FB act as an electron sink from the PsaA/PsaB core cofactors to docked ferredoxin. PsaD provides a ferredoxin docking site (Kruip et al.. 1997; Chitnis et al., 1996). Inactivation of PsaD in Synechocysiis indicated that PsaD is essential for efficient function of cyanobacterial PSI and for stable assembly of PsaE (Chimis et al.. 1989). The PsaD:ferredoxin docking involves electrostatic interactions between the basic PsaD protein and electronegative surfaces of ferredoxin. Figure 2.12 (Lelong et al.. 1994). PsaE assisting in the interaction of PSI with ferredoxin (Xu et al., 1994b). Membranes of the PsaE-less mutant were severely deficient in the reduction of ferredoxin. It is suggested that both PsaD and PsaE together form a suitable docking site for ferredoxin. The A| site of PSI is of particular interest in this review. In the crystal structure, phylloquinone is fairly well identified in the location of PSI but does not include the interaction of amino acids (Klukas et al.. 1999a). However, a highly conserved primar>' amino acid sequence in both PsaA and PsaB containing the

A_686j_^FSGRGYWQELIE®'^ and BJ^^LISWRGYWQELIE^'® in the connections between transmembrane helices m and m' and peripheral helices n and n' are proposed to bind phylloquinone (Schubert et al.. 1997). Derivations of this fragment peptide are also partially conserved in PshA of the P800 reaction center of heliobacterium and PscA of the P840 reaction center of green sulfur bacterium Chlorobrium. The protein-phylloquinone interaction, by electron spin echo envelope modulation (ESEEM), suggests that there are two protein nitrogen nuclei coupled to the phyllosemiquinone radical AT (Hanley et al.. 1997). One nitrogen is thought to be from an indole tryptophan and the other from a histidine. , or asparagine. Thus, the conserved amino acids A_W693/A_Q694 or B_W668/B_Q669 may bind phylloquinone. On the cyanobacterium Anabaena variabilis (Rigby et al.. 1996) and Synechococcus (Klughammer et al.. 1999). ENDOR and related techniques determined that each quinone oxygen contained a hydrogen bond. However, positive identification of the proton donor has not been established. Point mutations have been performed on various amino acids within and near the Ai binding site to determine 31

their interaction with phylloquinone (Xu and Chitnis. 2000). However, final analysis has not been compiled. Further localization of quinones within the Ai site of Synechocystis sp. PCC 6803 has been determined by non-crystallographic techniques. Pulsed EPR has determined that

the distance from the reduced phylloquinone in the A| site to PTOO"^ is 25.4 A and 27 ± 5" out of line from P700 to Fx (Bittl and Zech, 1997). The vector of the carbonyl oxygens in phylloquinone point at P700 (Zech et al., 1997) with the carbonyls oriented 63° from the membrane plane (MacMillan et al., 1997). However, 1,4-naphthoquinone contained within the Al site is oriented by 90° with the carbonyl oxygens nearly perpendicular to P700 (Zech et al., 1997). This indicates that the alkyl 'tail* of the quinone is important for proper orientation within the Ai site. 6. A general summary of the bioenergetics of PSII through ATPase Photosynthesis is a complex series of steps that converts solar energy into the high energy form of cofactors (Ort et al.. 1996). In cyanobacteria and higher plants, two separate complexes trap photons of light creating a charge separation initiating a electron transfer through the protein. Several additional complexes either shuttle the generated electrons or siphon off a portion of the potential energy to chaimel it into the high energy form of a cofactor. .Most of the complexes and electron transports are lipid bound in the thylakoid membranes in cyanobacteria or chloroplasts. The soluble proteins act as the electron transport agents between the lipid bound protein complexes. The initial event is the absorption of a photon by a Chi molecule in the light harvesting complex 11 (LHC 11). The energy is transferred as an exciton among the Chls to the P680 reaction center (RC). Within the P680 a charge separation occurs and the high energy electron is then transferred into the electron transfer chain. This is composed of a series of intermediate pheophytin electron receptors and the QA and QB site, both of which are occupied with plastoquinone (PQ). PQ is a lipid soluble molecule that serves as a mobile carrier between PSII and the Cyt complex. The QB site is essentially a loading dock for PQ. Two electrons from two turnovers of P680 reduce the PQ, which through binding two protons from the stroma, creates PQH2. The PQH2 has a lower affinity to the QB site, thus dissociates into the thylakoid membrane. The Cyt complex performs many functions. starting with stripping the electrons from the plastoquinol. In the process, the protons are dumped into the lumen. Plastoquinone is then released from the Cyt b(f complex to be recycled. A second proton discharging site is the oxygen evolving complex (OEC) which is attached to the lumenal side of PSII. The OEC generates dioxygen from water using the four electrons for the reduction of P680 with the four proton being released into the lumen. This, in turn creates a proton gradient between the lumen and stroma. The generated electrons are transferred through Cyt b(f to reduce plastocyanin. The oxidized P700'' is reduced by the electron provided from the plastocyanin. The energy from a second photon is used to excite the P700 RC electron of PSI so that it can enter the second of the electron transport chains. P700 is the RC that is the primary electron donor in PSI. The PSI electron moves across the membrane through a series of prosthetic groups to a docked ferredoxin. The reduced ferredoxin transfers the electron via ferredoxin-NADP"' generating NADPH or back to the Cyt b(f complex, depending on energy needs. The previously generated proton gradient drives ATP synthesis while cocontaminantly releasing protons back into the stroma. The specific energetic mechanisms of PSI will be further discussed. 7. Photosystem I energetic schemes After absorption of a photon by an antenna Chi and migration of the exciton to the P700 trap, primary charge separation occurs between the primary reactants P700 and .A.o leading to a short-lived P700' Ao' radical pair. This charge-separated state is stabilized by rapid, secondary' electron transfer events though the intermediar>' acceptors Ai and Fx to the terminal FA/FB acceptors. This stabilization allows the soluble electron transport proteins plastocyanin (or cyt C(,) and soluble [2Fe-2S] ferredoxin to interact with the primar\' donor and terminal acceptors, respectively, in diffusion-limited processes. The final step in non- cyclic electron transport is the reduction of NADP^ by reduced ferredoxin. catalyzed b> the enzj'me ferredoxinrNADP"^ oxidoreductase (Knaff, 1996). An indispensable review b\ Klaus Brettel covers in brilliant detail electron transfer within PSI, therefore this review will be restrained to post 1997 publications, except for applicable background material (Brettel. 1997). In Synechocystis sp. PCC 6803 photon of light absorbed by an antenna Chi undergoes a complex equilibrium among the bulk Chi and red-absorbing Chls before it is trapped by the P700 reaction center in approximately 20-25 ps kinetics (Hastings et al.. 1994; Savikhin et al., 1999). Recent work suggests that the formation of the initial radical pair of (P700*Ao') is 9-10 ps (Savikhin et al.. 2000), which is slower than the 1-3 ps. as suggested by Brettel earlier in . review (Brettel, 1997). The reoxidation of Ao and presumed reduction of A| occurs with 50 ps (Hecks et al.. 1994). Direct reduction of Ai has not been well described due to, in part, of inadequately fast spectroscopic instruments that operate in the blue region at which PhQ absorbs. Flash absorption kinetics monitoring PhQ at 380 nm show that the PhQ reduction was faster than the 5 ns time resolution (Brettel, 1988). The reoxidation of the Ai phylloquinone is in the tens and hundreds of ns time range depending on the experiment (Brettel. 1997). Determining the reoxidation is further complicated by the organism, variations in preparation of the PSI particles, and the state of the iron-sulfur clusters. For both spinach and Synechocystis, in general the reoxidation of PhQ takes place at around 200 ns. with a 15-25 ns component that may be either a heterogeneity in the sample or a possible fast redox equilibrium between Ai and Fx. However, more recently Joliot has developed an in vivo technique for monitoring the PSI PhQ and determined that for Chlorella sorokininia there is a biphasic oxidation of PhQ with times of 18 and 160 ns (Joliot and Joliot, 1999). He attributes the two phases to two possible hypotheses '^1) Photosystem I reaction centers are present under two conformational states which differ by the reoxidation rate of AT. (2) The two phylloquinones corresponding to the two branches of the PS 1 heterodimer are involved in the electron transfer." (Joliot and Joliot, 1999). Hypotheses 1 suggests that there is a fast (18 ns) equilibrium between A| and Fx and the slow phase (160 ns) being attributed to the electron being bled off onto FA and FB- Evidence to argue against hypothesis 1 includes, extraction of FA and FB does not remove the 160 ns phase. Support for hypothesis 2 includes the similar amplitudes of the two phases implies that the rates of electron transfer from P700 to each PhQ is roughly equal. The two different rate constants measured both AT oxidation suggests some asymmetry between PhQ and Fx- This however may not be concluded until site directed mutants of the Ai site generated in Synechocystis and other organisms are completely analyzed (by this lab and others). 34

P700» PS I 10-30 ps

'P700 3-10 in / SOO«n

10 ^ Fid 100 ;«

10 rh: SO ms

ir.-v

1S

Figure 2.12 Energetics of PSI. The back reaction rates in the wild type refer to conditions where the succeeding electron acceptor has been removed either biochemically or genetically. The values in bold represent the dominant kinetic phase. The back reaction pathways are depicted as direct to P700 for the sake of clarity; actual pathway likely proceeds, at least in part back though the electron acceptor chain. Figure from Ref. (Semenov et al.. 2000).

The electron transfer rates from Ai to Fx and between FA and FB are complex. Given the similar spectroscopic properties of the clusters, discerning the electron transfer rates among the sites are an ongoing process. A variety of studies have determined that the sequence of forward electron transfer was Ai to Fx to (FAFB) (Brettel. 1997). Recently evidence by Golbeck has suggested that the correct arrangement of the iron-sulfur clusters is

FX to FA to FB (Shinkarev et al., 2000). This was in support that FB was found to be the iron- sulfur cluster that interacts with flavodoxin and ferredoxin (Vassiliev et al.. 1998). One last topic with the electron transfer within PSI concerns charge recombination kinetics between a secondary electron acceptor site and the P700" RC. As given in Fig. 2.11. the dotted lines that connect a secondary site with P700 is the back reaction time reversing through the acceptor chain. The primary pair P700^Ao' decays with a half time in the order of 30 ns, forming two products: a singlet ground state of P700 and a triplet state "PTOO. The "TTOO then decays to the ground state on a fis time scale (Golbeck and Brv'ant. 1991). The Ai site back reacts at about 100 (is. when the subsequent iron-sulfur cluster acceptors are either removed or prereduced (Brettel. 1990). The backreaction of the terminal iron-sulfur clusters is the most pertinent of the recombination kinetics with regards to the

rest of the dissertation. Normally the FAFB back reaction in the wild type is close to 80 ms and 12 ms (Vassiliev et al.. 1997). The biphasic decay is considered to be a result of the PSl complexes being in different confirmations. The microsecond decay is attributed to damaged reaction centers. However, this backreaction rate from FAFB is sensitive to the quinone in the Al site (Johnson et al.. 2000b). Hence, this is an effective spectroscopic tool for probing the electron intermediates within PSI. 8. Alternative quinones in the A| site of PSI The latter part of dissertation concerns replacing phylloquinone with alternative quinones into the Ai site in vivo. Previous work from the late 1980"s and early 1990"s. places the quinones into PSI in vitro upon extraction with organic solvents. It has been shown that phylloquinone can be extracted from PSI with a dry or water saturated diethyl ether solution (Itoh et al., 1987; Itoh and Iwaki, 1989) or by a hexane containing 0.3% methanol (Rustandi et al., 1990: Setif et al., 1987). However, approximately half of the antenna Chi and all the caroteniods are also removed. Both procedures demonstrated that charge separation was essentially blocked at the level of Ao (P700"Ao") (Biggins and Mathis. 1988; Itoh et al.. 1987). There were a variety of studies focusing on the reconstitution of electron transfer in phylloquinone depleted PSI by incubating with non-native quinones. li has been shown that many benzoquinones. naphthoquinones, anthraquinones. and even anthrones and fluorenones can be inserted into A| and accept electrons from AD" in the phylloquinone depleted cells (Rustandi et al.. 1992; Iwaki and Itoh. 1994b). Iwaki and Uoh reported the function of the A| site could be reactivated by a quinone of an appropriate redox potential (Iwaki and Itoh. 1994a). This included reduction of the iron-sulfur clusters by monitoring the absorbance changes of FAFB" and P700" at 470 nm and 430 nm. respectively. However, these results conflicted with other quinone reconstitution experiments where the phylloquinone was removed by a different organic solvent mixture (Biggins. 1990). Work by Biggins (Biggins, 1990) and Thumauer (Rustandi el al., 1992) 36

showed by both optical and EPR experiments that a the assortment of benzo-. naphtho-. and antliro- quinones used in the reconstitution experiments integrated into the Ai site with \ ar\ ing degrees of success and only the quinones with a "taiF actually passed electrons onto lo the iron-sulfur clusters. Time resolved EPR also showed that upon reconstitution with phylloquinone a charge separation of P700^FeS" was restored (Sieckman et al.. 1991). The remaining quinones (Q = duroquinone, dueterated duroquinone, and 1,4-naphthoquinone) contained a charge separation of P700^Q' with the electron being blocked from proceeding further. Biggins suggested structural constraints imposed by an alkyl tail in addition to energetic considerations (Biggins. 1990). Discrepancy between these results could have arrived from a com.bination of the harsh PhQ extraction procedures and the lack of resolution of the instruments. This was one motivation for the author to attempt to remove phylloquinone from PSI in vivo. Then in vivo or in vitro PSI complexes, alternative quinones can be reconstituted into the A| site of PSI complexes that have not been damaged or modified by organic solvents. 9. Summary Phylloquinone plays very specific roles in biological processes, yet critical ones. A lack of vitamin K in an animal would have a lethal result from the inabilit>' to clot blood. But the use of phylloquinone as a regulatory cofactor for the carboxylase reaction, converting specific glutamic acid residues to y-carboxyglutamic acid, simplifies the regulation of the whole pathway. Additionally, the phylloquinone is consumed either directly or indirectly in adequate levels in the diet of the animals. Within plants phylloquinone has only one known use as an intermediate electron acceptor in PSI. How phylloquinone became the evolutionary target as a specific electron transfer agent is unknown. Why plastoquinone is not used or has not evolved as a replacement for phylloquinone within PSI is also unknown. The author muses that phylloquinone is being used over plastoquinone in PSI as a possible response to some previous evolutionary pressure. Once the stress was past, PSI was quite optimized for phylloquinone. both in the electron intermediate head group and the binding of the shorter mostly unsaturated tail. 37

10. References

Antonkine. M. L.. D. Bentrop. I. Bertini. C. Luchinat. G. Shen. D. A. Bryant, D. Stehlik. and J. H. Golbeck. 2000. Paramagnetic IH NMR spectroscopy of the reduced, unbound photosystem I subunit PsaC: sequence-specific assignment of contact-shifted resonances and identification of mixed- and equal-valence Fe-Fe pairs in [4Fe-4S] centers FA- and FB-. J Biol Inorg Chem. 5:381-92. Biggins. J. 1990. Evaluation of selected benzoquinones, naphthoquinones, and anthraquinones as replacements for phylloquinone in the A1 acceptor site of the photosystem I reaction center. In Biochemistr>'. 7259-64. Biggins. J., and P. Mathis. 1988. Functional role of vitamin K1 in photosystem I of the cyanobacterium Synechocystis 6803. In Biochemistry. 1494-500. Bittl, R.. and S. G. Zech. 1997. Pulsed EPR study of spin coupled radical pairs in photosynthetic reaction centers: measurement of the distance between Pyoo"'

Ai'- in Photosystem 1 and between Ps65^' Qa"' bacterial reaction center../. Phys. Chem. B101:1429-1436. Bittl, R.. S. G. Zech, P. Fromme. H. T. Witt, and W. Lubitz. 1997. Pulsed EPR structure analysis of photosystem I single crystals: localization of the phylloquinone acceptor. Biochemistry. 36:12001-4. Bolton-Smith. C., R. J. Price, S. T. Fenton, D. J. Harrington, and M. J. Shearer. 2000. Compilation of a provisional UK database for the phylloquinone (vitamin Kl) content of foods. BrJNutr. 83:389-99. Booth. S. L.. and J. \V. Suttie. 1998. Dietary intake and adequacy of vitamin K. J Sutr. 128:785-8. Brettel. K. 1988. Electron transfer from Al- to an iron-sulfur center with tl/2 = 200 ns at room temperature in photosystem I. Characterization by flash absorption spectroscopy. In FEBS Lett. 93-8. Brettel. K. 1990. Electron transfer in photosystem I from Synechococcus sp. under reducing conditions. In Curr. Res. Photosynth.. Proc. Int. Conf. Photosynth.. 8th. 627-30. Brettel. K. 1997. Electron transfer and arrangement of the redox cofactors in photosystem 1. Biochim. Biophys. Acta. 1318:322-373. Campbell. H. A., and K. P. Link. 1941. Studies on the hemorrhagic sweet clover disease 1\'. The isolation and cr>'stallization of the hemorrhagic agent. J. Biol. chem. 138:21. Campbell. I. M., D. J. Robins, M. Kelsey, and R. Bentley. 1971. Biosynthesis of bacterial menaquinones (vitamins K. 2 ). Biochemistry. 10:3069-78. Chitnis. P. R. 1996. Photosystem L 38

Chitnis. P. R.. V. P. Chitnis. Q. Xu. Y. S. Jung. L. Yu. and J. H. Golbeck. 1995. Mutational analysis of photosystem I polypeptides. In Photosynth.: Light Biosphere. Proc. Int. Photosynth. Congr.. 10th. 17-22. Chitnis, P. R.. P. A. Reilly. and N. Nelson. 1989. Insertional inactivation of the gene encoding subunit II of photosystem I from the cyanobacterium Synechocystis sp. PCC 6803. J. Biol. Chem. 264:18381-18385. Chitnis. V. P., and P. R. Chitnis. 1993. PsaL subunit is required for the formation of photosystem I trimers in the cyanobacterium Synechocystis sp. PCC 6803. FEBS Lett. 336:330-334. Chitnis. V. P.. Y. S. Jungs. L. Albee, J. H. Golbeck. and P. R. Chitnis. 1996. Mutational analysis of photosystem I polypeptides. Role of PsaD and the lysyl 106 residue in the reductase activity of the photosystem \. J Biol Chem. 271:11772-80. Chitnis. V. P., A. Ke, and P. R. Chitnis. 1997. The PsaD subunit of photosystem I. Mutations in the basic domain reduce the level of PsaD in the membranes. Plant Physiol. 115:1699-705. Chitnis. V. P.. Q. Xu, L. Yu, J. H. Golbeck, H. Nakamoto, D. L. Xie. and P. R. Chitnis. 1993a. Targeted inactivation of the gene psaL encoding a subunit of photosystem I of the cyanobacterium Synechocystis sp. PCC 6803. J Biol Chem. 268:11678-84. Chitnis. V. P.. Q. Xu. L. Yu. J. H. Golbeck. H. Nakamoto. D. L. Xie, and P. R. Chitnis. 1993b. Targeted inactivation of the gene psal- encoding a subimit of photosystem 1 of the cyanobacterium Synechocystis sp. PCC 6803. J. Biol. Chem. 268:11678- I1684. Co.x. G. B.. and F. Gibson. 1966. The role of shikimic acid in the biosynthesis of vitamin K2. Biochem J. 100:1-6. Dam. H. 1935. The antihemorrhagic vitamin of the chick. Occurrence and chemical nature. Nature. 135:652. Daruwala. R.. O. Kwon. R. Meganathan, and M. E. Hudspeth. 1996. A new isochorismate synthase specifically involved in menaquinone (vitamin K2) biosynthesis encoded by the menF gene. FEMS Microbiol Lett. 140:159-63. Driscoll. J. R.. and H. W. Taber. 1992. Sequence organization and regulation of the Bacillus subtilis menBE operon. J Bacteriol. 174:5063-71. Fromme. P.. W.-D. Schubert, and N. Krauss. 1994. Structure of photosystem I: Suggestions on the docking sites for plastocyanin. ferredoxin and the coordination of P700. Biochim. Biophys. Acta. 1187:99-105. Furie. B.. B. A. Bouchard, and B. C. Furie. 1999. Vitamin K-dependent biosynthesis of gamma-carboxyglutamic acid. Blood. 93:1798-808. Golbeck. J. H. 1994. Photosystem I in Cyanobacteria. In The Molecular Biology of Cyanobacteria. D. A. Bryant, editor. Kluwer Academic Publishers. Dordrecht. The Netherlands. 179-220. 39

Golbeck. J. H., and D. A. Bryant. 1991. Pholosystem I. Current Topics in Bioenergerics. 16:83-177. Guest, J. R. 1977. Menaquinone biosynthesis: mutants of Escherichia coli K-12 requiring 2- succinylbenzoate. J Bacterial. 130:1038-46. Hanley. J.. Y. Deligiannakis. F. MacMillan. H. Bottin. and A. W. Rutherford. 1997. ESEEM study of the phyllosemiquinone radical Al.- in 14N- and 15N- labeled photosystem 1. Biochemistry. 36:11543-9. Hastings. G.. F. A. M. Kleinherenbrink, S. Lin. T. J. McHugh. and R. E. Blankenship. 1994. Observation of the Reduction and Reoxidation of the Primary Electron Acceptor in Photosystem I. In Biochemistry. 3193-200. Hecks, B., K. Wulf, J. Breton, W. Leibl, and H. W. Trissl. 1994. Primary charge separation in photosystem I: a two-step electrogenic charge separation connected with P700'^Ao- and P700^A,- formation. Biochemistry. 33:8619-24. Hill. K. F., J. P. Mueller, and H. W. Taber. 1990. The Bacillus subtilis menCD promoter is responsive to extracellular pH. Arch Microbiol. 153:355-9. lioh. S.. and M. Iwaki. 1989. Vitamin K sub(l) (phylloquinone) restores the turnover of FeS centers in the ether-extracted spinach PS I particles. FEBS LETT. 243:47-52. Itch, S.. M. Iwqaki, and I. Ikegami. 1987. Extraction of vitamin K-1 from photosystem 1 particles by treatment with diethyl ether and its effect on the A|"EPR signal and system 1 photochemistry. Biochim. Biophys. Acta. 893:508-516. Iwaki. M.. and S. Itoh. 1994b. Reaction of reconstituted acceptor quinone and dynamic equilibration of electron transfer in the photosystem I reaction center. In Plant Cell Physiol. 983-93. Iwaki. M.. and S. Itoh. 1994a. Reaction of reconstituted acceptor quinone and dynamic equilibration of electron transfer in the photosystem I reaction center. Plant Cell Physiol. 35:983-993. Johnson. T. W.. S. Naithani, J. H. Golbeck. and P. R. Chitnis. 2000a. investigating the topic of genetically disabling the menD and menE genes in Synechocystis sp. PCC 6803. To be submitted to Plant Physiology. Manuscript in preparation. Johnson, T. W.. G. Shen. B. Zybailov, D. Kolling. R. Reategui. S. Beauparlant. I. R. Vassiliev. D. A. Bryant, A. D. Jones, J. H. Golbeck, and P. R. Chitnis. 2000b. Recruitment of a foreign quinone into the A(l) site of photosystem I. I. Genetic and physiological characterization of phylloquinone biosynthetic pathway mutants in Synechocystis sp. pcc 6S03. J Biol Chem. 275:8523-30. Joliot. P.. and A. Joliot. 1999. In vivo analysis of the electron transfer within photosystem 1: are the two phylloquinones involved? Biochemistry. 38:1 1130-6. Kaipling. S.. J. Soli, and G. Schultz. 1984. Site of methylation of 2-phytyl-1.4- naphthoquinol in phylloquinone (vitamin Kl) synthesis in spinach chloroplasts. In Phytochemistry. 89-91. 40

Kaneko. T.. S. Sato. H. Kotani. A. Tanaka. E. Asamizu. Y. Nakamura, N. Miyajima. M. Hirosavva. M. Sugiura. S. Sasamoto. T. Kimura. T. Hosouchi. A. Matsuno. A. Muraki. N. Nakazaki, K. Naruo, S. Okumura, S. Shimpo. C. Takeuchi, T. Wada. A. Watanabe. M. Yamada. M. Yasuda, and S. Tabata. 1996. Sequence analysis of the genome of the unicellular cyanobacterium Synechocystis sp. strain PCC6803. II. Sequence determination of the entire genome and assignment of potential protein- coding regions. DNA Res. 3:109-36. Klughammer, C.. B. Klughammer, and R. Pace. 1999. Deuteration effects on the in vivo EPR spectrum of the reduced secondary photosystem I electron acceptor A1 in cyanobacteria. Biochemistry. 38:3726-32. Klukas. O.. W. D. Schubert. P. Jordan. N. Krau. P. Fromme. H. T. Witt, and W. Saenger. 1999a. Localization of two phylloquinones, QK and QK'. in an improved electron densit>' map of photosystem I at 4-A resolution. J 5/o/ Chem. 274:7361-7. Klukas. O.. W. D. Schubert, P. Jordan. N. Krauss. P. Fromme. H. T. Witt, and W. Saenger. 1999b. Photosystem I. an improved model of the stromal subunits PsaC. PsaD. and PsaE. J5/o/ Chem. 274:7351-60. Knaff. D. B. 1996. Ferredoxin and Ferredoxin-dependent enzymes. In Oxygenic Photosynthesis: the Light Reactions. D. R. Ort and C. F. Yocum. editors. Kluwer. Dordrecht. 333-361. Koike-Takeshita. A., T. Koyama, and K. Ogura. 1997. Identification of a novel gene cluster participating in menaquinone (vitamin K2) biosynthesis. Cloning and sequence determination of the 2- heptaprenyl-1.4-naphthoquinone methyltransferase gene of Bacillus stearothermophilus. y^/o/ Chem. 272:12380-3. Krauss. N.. W. D. Schubert. O. Klukas. P. Fromme. H. T. Witt, and W. Saenger. 1996. Photosystem 1 at 4 A resolution represents the first structural model of a joint photosynthetic reaction centre and core antenna system. Nat Struct Biol. 3:965-73. Kruip. J.. P. R. Chitnis. B. Lagoutte. M. Rogner, and E. J. Boekema. 1997. Structural organization of the major subunits in cyanobacterial photosystem 1. Localization of subunits PsaC. -D. -E. -F. and -i. J Biol Chem. 272:17061-9. Lancaster. C. R.. and H. Michel. 1999. Refined crystal structures of reaction centres from Rhodopseudomonas viridis in complexes with the herbicide atrazine and two chiral atrazine derivatives also lead to a new model of the bound carotenoid. J Mol Biol. 286:883-98. Leiong. C.. P. Setif. B. Lagoutte, and H. Bottin. 1994. Identification of the amino acids involved in the functional interaction between photosystem I and ferredoxin from Synechocystis sp. PCC 6803 by chemical cross-linking. J. Biol. Chem. 269:10034- 10039. MacMillan. F.. J. Hanley. L. van der Weerd. M. Knupling, S. Un, and A. W. Rutherford. 1997. Orientation of the phylloquinone electron acceptor anion radical in photosystem 1. Biochemistry. 36:9297-303. 41

Malkin. R. 1996. Photosystem I electron transfer reaction -- components and kinetics. In Oxygenic Photosynthesis: the Light Reactions. D. R. Ort and C. F. Yocum. editors. K-Iuwer. Dordrecht. 313-332. Manna. P.. and P. R. Chitnis . 1998. Function and molecular genetics of photosystem I. In Concepts in Photobiology: Photosynthesis and Photomorphogenesis. G. S. Singhal. G. Renger. and S. K. Sapory. editors. 212-251. Mannan. R. M., H. B. Pakrasi. and K. Sonoike. 1994. The PsaC protein is necessar>' for the stable association of the PsaD, PsaE. and PsaL proteins in the photosystem I complex: analysis of a cyanobacterial mutant strain. Arch Biochem Biophvs. 315:68- 73. Monsen. E. R. 1989. The 10th edition of the Recommended Dietar>' Allowances: what's new in the 1989 RD\sl J Am Diet Assoc. 89:1748-52. Nexhushtai, R.. A. Eden, Y. Cohen, and J. Klein. 1996. Introduction to Photosystem I: reaction center, function, composition, and structure. In Oxygenic Photosynthesis: the Light Reactions. D. R. Ort and C. F. Yocum. editors. Kluwer. Dordrecht. 289- 311. Ort. D. R.. C. F. Yocum, and I. F. Heichel. 1996. Oxygenic photosynthesis : : the light reactions. Kluwer, Dordrecht; Boston. Palaniappan, C.. V. Sharma. M. E. Hudspeth, and R. Meganathan. 1992. Menaquinone (vitamin K2) biosynthesis: evidence that the Escherichia coli menD gene encodes both 2-succinyl-6-hydroxy-2,4-cyclohexadiene-l- carboxylic acid synthase and alpha-ketoglutarate decarboxylase activities. J Bacterial. 174:8111-8. Palaniappan. C.. H. Taber. and R. Meganathan. 1994. Biosynthesis of o-succinylbenzoic acid in Bacillus subtilis: identification of menD mutants and evidence against the involvement of the alpha-ketoglutarate dehydrogenase complex. J Bacieriol. 176:2648-53. Reategui. R.. W. Johnson, W. Xu, B. Zybailov. L Vassiliev, G. Shen. J. H. Golbeck. and P. R. Chitnis. 1998. Mutational analysis of photosystem I Function of phylloquinones. In Photosynth.: Mech. Eff., Proc. Int. Congr. Photosynth., 11th. 515-520. Reategui, R. F. 1998. Phylloquinone in the photysnthetic activity of Synechocystis sp. PCC 6803. Iowa State Univsersity, Ames. Reedstrom. C. K.. and J. W. Suttie. 1995. Comparative distribution, metabolism, and utilization of phylloquinone and menaquinone-9 in rat liver. Proc Soc Exp Biol Med. 209:403-9. Rigby. A. C.. J. D. Baleja, L. Li, L. G. Pedersen, B. C. Furie, and B. Furie. 1997. Role of gamma-carboxyglutamic acid in the calcium-induced structural transition of conantokin G, a conotoxin from the marine snail Conus geographus. Biochemistry. 36:15677-84. Rigby. S. E.. M. C. Evans, and P. Heathcote. 1996. ENDOR and special triple resonance spectroscopy of Al.- of photosystem 1. Biochemistry. 35:6651-6. 42

Rohmer. M.. M. Knani. P. Simonin. B. Sutler, and H. Sahm. 1993. Isoprenoid biosynthesis in bacteria: a novel pathway for the early steps leading to isopentenyl diphosphate. BiochemJ. 295:517-24. Ronden, J. E.. M. J. Drittij-Reijnders, C. Vermeer. and H. H. Thijssen. 1998. Intestinal flora is not an intermediate in the phylloquinone- menaquinone-4 conversion in the rat. Biochim Biophys Acta. 1379:69-75. Roshchina. V. V. 1979. [Effect of dibromothymoquinone on photosynthetic electron transport]. Biokhimiia. 44:477-81. Rustandi. R. R.. S. W. Snyder, J. Biggins. J. R. Norris. and M. C. Thumauer. 1992. Reconstitution and exchange of quinones in the A1 site of photosystem I. An electron spin polarization electron paramagnetic resonance study. In Biochim. Biophys. Acta. 311-20. Rustandi. R. R.. S. W. Snyder, L. L. Feezel, T. J. Michalski, J. R. Norris, M. C. Thumauer. and J. Biggins. 1990. Contribution of vitamin K1 to the electron spin polarization in spinach photosystem I. Biochemistry. 29:8030-2. Savikhin. S.. W. Xu. P. R. Chitnis. and W. S. Struve. 2000. Ultrafast primary processes in PS I from Synechocystis sp. PCC 6803: roles of P700 and AO. In Biophys. J. 1573- 1586. Savikhin. S.. W. Xu. V. Soukoulis, P. R. Chitnis. and W. S. Stmve. 1999. Ultrafast primary processes in photosystem I of the cyanobacterium Synechocystis sp. PCC 6803. In Biophys. J. 3278-3288. Schluchter. W. M.. Shen, G.. Zhao. J. &. Bryant, D. A. 1996. Characterization of psal and psaL mutants ofSynechococcus sp. strain PCC 7002: a new model forstate transitions in cyanobacteria. Photochem. Photobiol. 64:53-66. Schubert. W. D.. O. Klukas. N. Krauss. W. Saenger. P. Fromme. and H. T. Witt. 1997. Photosystem I of Synechococcus elongatus at 4 .A. resolution: comprehensive structure analysis. J Mol Biol. 272:741-69. Schubert. W. D.. O. Klukas. W. Saenger. H. T. Witt, P. Fromme, and N. Krauss. 1998. A common ancestor for oxygenic and anoxygenic photosynthetic systems: a comparison based on the stmctural model of photosystem I. J Mol Biol. 280:297- 314. Schultz, G.. J. Soil. E. Fiedler, and D. Schultz-Siebert. 1985. Synthesis of prenylquinones in chloroplasts. Physiol. Plant. 64:123-129. Semenov. A. Y., 1. R. Vassiliev. A. van Der Est. M. D. Mamedov, B. Zybailov, G. Shen. D. Stehlik. B. .A,. Diner. P. R. Chitnis. and J. H. Golbeck. 2000. Recruitment of a Foreign Quinone into the A1 Site of Photosystem I. Altered Kinetics of Electron Transfer in Phylloquinone Biosynthetic Pathway Mutants Studied by Time-Resolved Optical. EPR and Electrometric Techniques. J5/o/ Chem. 275:23429-23438. 43

Setif, P.. I. Ikegami, and J. Biggins. 1987. Lighi-induced charge separation in Photosystem I at low temperature is not influenced by vitamin K-1. Biochim. Biophys. Acta. 894:146-56. Sharma. V.. M. E. Hudspeth, and R. Meganathan. 1996. Menaquinone (vitamin K2) biosynthesis: localization and characterization of the menE gene from Escherichia coli. Gene. 168:43-8. Sharma, V., R. Meganathan, and M. E. Hudspeth. 1993. Menaquinone (vitamin K2) biosynthesis: cloning, nucleotide sequence, and expression of the menC gene from Escherichia coli. J Bacteriol. 175:4917-21. Sharma, V., K. Suvama. R. Meganathan. and M. E. Hudspeth. 1992. Menaquinone (vitamin K2) biosynthesis: nucleotide sequence and expression of the menB gene from Escherichia coli. J Bacteriol. 174:5057-62. Shen, G.. and D. Bryant. 2000. Confirmation of the menG knockout in Synechosystis sp. PCC 6803. Shinkarev. V. P., I. R. Vassiliev, and J. H. Golbeck. 2000. A kinetic assessment of the sequence of electron transfer from F(X) to F(A) and further to F(B) in photosystem I: the value of the equilibrium constant between F(X) and F(A). Biophys J. 78:363-72. Sieckman. I., A. van der Est, H. Bottin, P. Setif. and D. Stehlik. 1991. Nanosecond electron transfer kinetics in photosystem I following substitution of quinones for vitamin K1 as studied by time resolved EPR. FEBS Lett. 284:98-102. Snyder, C. D., and H. Rapoport. 1970. Biosynthesis of bacterial menaquinones. Origin of quinone oxygens. Biochemistry. 9:2033-8. Sperling. R.. B. C. Furie, M. Blumenstein, B. Keyt, and B. Furie. 1978. Metal binding properties of gamma-carboxyglutamic acid. Implications for the vitamin K- dependent blood coagulation proteins. J5/o/ Chem. 253:3898-906. Summers. L. A. 1980. The bipyridinium herbicides. Academic Press, London : New York. Sun. J. 1999. Oxidizing side of photosystem I. Iowa State Universit>', Ames. Sun. J.. A. Ke. P. Jin. V. P. Chitnis, and P. R. Chitnis. 1998. Isolation and functional study of photosystem I subunits in the cyanobacterium Synechocystis sp. PCC 6803. .Methods Enzymol. 297:124-39. Suttie. J. W. 1985. Vitamin K-dependent carboxylase. Annu Rev Biochem. 54:459-77. Suvarna. K.. D. Stevenson, R. Meganathan, and M. E. Hudspeth. 1998. Menaquinone (vitamin K2) biosynthesis: localization and characterization of the menA gene from Escherichia coli. J Bacteriol. 180:2782-7. Taber. H. W., E. A. Dellers, and L. R. Lombardo. 1981. Menaquinone biosynthesis in Bacillus subtilis: isolation of men mutants and evidence for clustering of men genes. J Bacteriol. 145:321 -7. 44

Takahashi. Y.. M. Goldschmidt-Clermont, S.-Y. Soen, L. G. Franzen, and J.-D. Rochaix. 1991. Directed chloroplast transformation in Chlamydomonas reinhardtii : Insertional inactivation of the psaC gene encoding the iron sulfur protein destabilizes photosystem I. EMBOJ. 10:2033-2040. Thiel, T. 1994. Genetic analysis of cyanobacteria. In The molecular biology of cyanobacteria. D. A. Bryant, editor. Kluwer academic. Dodrecht. 581-611. Threlfall. D. R. 1971. The biosynthesis of vitamins E and K and related compounds. Viiam Horm. 29:153-200. Tiemann, R., G. Renger, P. Graber. and H. T. Witt. 1979. The plastoquinone pool as possible hydrogen pump in photosynthesis. Biochim Biophys Acta. 546:498-519. Tsiotis. G., W. Nitschke, W. Haase, and H. Michel. 1993. Purification and crystallization of photosystem I complex from a phycobilisome-less mutant of the cyanobacterium Synechococcus PCC 7002. Photosynthesis Res. 35:285-297. Unden, G., and J. Bongaerts. 1997. Alternative respiratory pathways of Escherichia coli: energetics and transcriptional regulation in response to electron acceptors. Biochim Biophys Acta. 1320:217-34. Van der Est, A.. T. Prisner, R. Bittl, P. Fromme, W. Lubitz, K. Moebius. and D. Stehlik. 1997. Time-resolved X-, K-. and W-band EPR of the radical pair state P700.bul.+Al.bul.- of photosystem I in comparison with P865.bul.+QA.bul.- in bacterial reaction centers. In J. Phys. Chem. B. 1437-1443. Vassiliev. I. R.. Y. S. Jung. M. D. Mamedov, A. Semenov, and J. H. Golbeck. 1997. Near- IR absorbance changes and electrogenic reactions in the microsecond-to-second time domain in Photosystem I. Biophys J. 72:301-15. Vassiliev. I. R.. Y.-S. Jung. F. Yang, and J. H. Golbeck. 1998. PsaC subunit of photosystem I Is oriented with iron-sulfur cluster FB as the immediate electron donor lo ferredoxin and flavodoxin. In Biophys. J. 2029-2035. Voet. D.. and J. G. Voet. 1995. Biochemistry. J. Wiley & Sons. New York. Wallin. R.. and L. F. Martin. 1987. Warfarin poisoning and vitamin K antagonism in rat and human liver. Design of a system in vitro that mimics the situation in vivo. Biochem J. 241:389-96. Wu, G.. H. D. Williams. M. Zamanian, F. Gibson, and R. K. Poole. 1992. Isolation and characterization of Escherichia coli mutants affected in aerobic respiration: the cloning and nucleotide sequence of ubiG. Identification of an S-adenosylmethionine- binding motif in protein. RNA. and small-molecule methyltransferases. J Gen Microbiol. 138:2101-12. Wu. S. M.. W. F. Cheung, D. Frazier, and D. W. Stafford. 1991. Cloning and expression of the cDNA for human gamma-glutamyl carboxylase. Science. 254:1634-6. www.genome.ad.jp. Website address. http://www.genome.ad.jp/dbget- bin/www_bget?path:synOO130. 45

Xu. Q.. Y. S. Jung, V. P. Chitnis. J. A. Guikema. J. H. Golbeck. and P. R. Chitnis. 1994b. Mutational analysis of photosystem I polypeptides in Synechocystis sp. PCC 6803. Subunit requirements for reduction of NADP"^ mediated by ferredoxin and flavodoxin. Biol. Chem. 269:21512-21518. Xu. Q.. Y. S. Jung. V. P. Chitnis, J. A. Guikema. J. H. Golbeck. and P. R. Chitnis. 1994a. Mutational analysis of photosystem I polypeptides in Synechocystis sp. PCC 6803. Subunit requirements for reduction of N.ADP+ mediated by ferredoxin and flavodoxin. j5/o/ Chem. 269:21512-8. Xu. W., and P. R. Chitnis. 2000. Unpublished results. Yu. J., L. B. Smart. Y. S. Jung, J. Golbeck. and L. Mcintosh. 1995a. Absence of PsaC subunit allows assembly of photosystem I core but prevents the binding of PsaD and PsaE in Synechocystis sp. PCC6803. Plant Mol Biol. 29:331-342. Yu, J.. L. B. Smart, Y. S. Jung, J. Golbeck. and L. Mcintosh. 1995b. Absence of PsaC subunit allows assembly of photosystem I core but prevents the binding of PsaD and PsaE in Synechocystis sp. PCC6803. Plant Mol Biol. 29:331-42. Yu. L.. J. Zhao, W. Lu, D. A. Bryant, and J. H. Golbeck. 1993. Characterization of the [3Fe- 4S] and [4Fe-4S] clusters in unbound PsaC mutants C14D and C51D. Midpoint potentials of the single [4Fe-4S] clusters are identical to FA and FB in bound PsaC of photosystem I. Biochemistry. 32:8251-8. Zech, S. G.. A. J. van der Est. and R. Bittl. 1997. Measurement of cofactor distances between P700.+ and Al.- in native and quinone-substituted photosystem I using pulsed electron paramagnetic resonance spectroscopy. Biochemistry. 36:9774-9. Zybailov. B., A. van der Est, S. G. Zech. C. Teutloff, T. W. Johnson. G. Shen. R. Bittl. D. Stehlik. P. R. Chitnis. and J. H. Golbeck. 2000. Recruitment of a foreign quinone into the A(I) site of photosystem I. II. Structural and functional characterization of phylloquinone biosynthetic pathway mutants by electron paramagnetic resonance and electron-nuclear double resonance spectroscopy. 75/o/ Chem. 275:8531-9. 46

CHAPTER 3. THE menD AND menE HOMOLOGUES IN CYANOBACTERIA CODE FOR ESSENTIAL ENZYMES IN PHYLLOQUINONE BIOSYNTHETIC PATHWAY'

A paper to be submitted to the Journal of Plant Physiology T. Wade Johnson^ Sushma Naithani^ Boris Zybailov^ A. Daniel Jones'^. John H. Golbeck* and Parag Chitnis^'"

Summarj- The genome of the cyanobacterium Synechocystis sp. PCC 6803 contains homologues of the Escherichia coli genes that encode enzymes of the menaquinone biosynthetic pathway. We inactivated the homologues of menD and menE, which code for 2-succinyl-6-hydroxyI-2,4-cyclohexadiene-l-carboxylate (SHCHC)^ synthase and O- succinylbenzoic acid-CoA ligase, respectively. The membranes of mutant strains do not contain phylloquinone. In Photosystem I complexes, the only known cyanobacterial enz\'me that utilizes phylloquinone in the wild type cells, contain plastoquinone-9 (PQ) in the mutant cells. The menD and menE mutants grow photoautotrophically under 40 (iE*m'"*s"' light intensity, with doubling times longer than the wild type. However, the menD mutant and wild type have similar doubling times under heterotrophic growth conditions, with menE remaining slower. Low temperature fluorescence show that the ratio of photosystem 1 to photosystem II is reduced in the menE and increased in menD relative to wild type. Electron transport activities are reduced in both mutants relative to wild type. Photoaccumulated EPR and 810 nm optical backreaction studies of mutant PSl trimers indicate that PQ is contained

' This work was supported by the National Science Foundation Grants {MCB0078 to PRC and MCB 9723661 to J. H. G.). This is Journal Paper No. J-XXXX of the Iowa Agriculture and Home Economics Experiments Station (Ames. I A), Project No. 3416 and supported by Hatch Act and State of Iowa funds.

" 'Department of Biochemistry. Biophysics and Molecular Biology. Iowa State University. Ames. lA 50011 •Department of Biochemistry and Molecular Biology. The Pennsylvania State Universit>', University Park. PA 16802°°Department of Chemistry. The Pennsylvania State Universit>'. University Park. PA 16802

Abbreviations; PS I. photosystem 1; PS II. photosystem II; EPR, electron paramagnetic resonance; DPIP. 2.6- dichlorophenolindophenol; Tris. PCR. polymerase chain reaction; SOS. sodium dodecyl sulfate; PAGE, polyacrylamide gel electrophoresis; Tricine, N-[2-hydroxy-l.I-bis(hydroxymethyl)ethyl]; PhQ. phylloquinone; PQ Plastoquinone 47 and active within the Ai site. In sunimar>% we demonstrate that the menD and menE homologues in cyanobacteria code for enzymes in phylloquinone biosynthetic pathway. We propose that there is no bypass for the SHCHC synthase and O-succinylbenzoic acid-CoA ligase in the PhQ biosynthetic pathway and that PQ is recruited as a cofactor into the .A.| site ofPSI. Introduction Phylloquinone, (PhQ. vitamin Ki, 2-methyl-3-phytyl-l,4-naphthoquinone). in animals is used as a cofactor for a single known enzymatic reaction, which converts specific glutamic acid residues to y-carboxyglutamic acid (Gla) by carboxylase (1). The Gla residue has the ability to chelate metals and impart structural changes in proteins upon exposure to calcium. Typically several Gla residues within a protein are clustered together and form a complex network of calcium chelated to multiple Gla residues (2). For example, these specific structural changes are responsible for activating several blood clotting factor proteins (VII. IX. X) and activating conantokin neuroactive peptides found in poisonous snails (3). The utilized phylloquinone is not synthesized by the animals but absorbed in the gui by consuming primarily green leafy' parts of plants (4). Phylloquinone's only known purpose in plants and cyanobacteria is to act as an electron acceptor in photosystem I (PSl). PSI is a photosynthetic reaction center (RC) that catalyzes the photooxidation of plastocyanin in the thylakoid lumen and the photoreduction of ferredoxin in the cyanobacterial cytoplasm (or chloroplast stroma) (5.6). PSI is a multisubunit protein, with PsaA and PsaB as a heterodimeric core surrounded by 6 small proteins (7). In addition to light-harvesting chlorophyll a molecules imbedded in PsaA/PsaB. there is the P700 RC. a dimer of Chi, and a series of cofactors (Ao, A|, Fx)- PSI also contains three stromal peripheral proteins; PsaD and PsaE acts as a ferredoxin docking site and PsaC contains two [4Fe-4S] iron-sulfur clusters (FA and FB). These cofactors act as electron acceptors which provide a charge separation from P700^ and act as an electron shuttle from the P700 on the lumenal side of the membrane to the terminal iron-sulfur clusters FA and FB on cytoplasmic side. The secondary electron acceptor site. A|. contains phylloquinone. This molecule acts as an intermediate in the electron transfer from the primary acceptor Ao. a chlorophyll a 48 molecule, to Fx, a [4Fe-4S] iron-sulfur cluster. From isolated PSI complexes, there are two bound molecules of phylloquinone per molecule of P700 (8-11); however only one molecule is considered to actually participate in the electron transfer (8,9,12). Although phylloquinone is an important photosynthetic cofactor, the biosynthetic route has not been determined for cyanobacteria. Many prokaryotes contain the metabolic pathway for phylloquinone biosynthesis. Fig. 1. In some bacteria, menaquinone (vitamin Ki). is used during fumarate reduction in anaerobic respiration (13,14). The genes encoding many enzymes involved in the conversion of the chorismate to menaquinone have been cloned in Esherichia coli. (15-20), Bacillus subtilis (21-24), and in Bacillus stearothermophilus (25). Menaquinone differs from phylloquinone in having a partially unsaturated C-40 side chain rather than a mostly saturated C-20 phytyl side chain, respectively. Therefore, we expect the synthesis of the naphthalene rings to involve similar steps. The genome of Synechocystis sp. FCC 6803 (26) contains homologs for several genes that encode enzymes in menaquinone biosynthesis: menF {entO) (isochorismate synthase), menD (2-succinyl-6-hydroxy-2,4-cyclohexadiene-l- carboxylate (SHCHC) synthase), menE (O-succinylbenzoic acid-CoA ligase), menB (DHNA synthase) and menA (phytyl transferase). Possible homologs of menC and ORF241 (the DHN.A. thioesterase) are also present in the genome. Previously we have shown phylloquinone is not synthesized in the cells by mutationally disabling the menA and menB homologs. therefore those genes are involved in phylloquinone biosynthesis (27). Two open reading frames sll0603 and slr0492 in Synechocystis sp. PCC 6803 genome code for proteins that have 28% and 29% sequence identity to the deduced amino acid sequences of the SHCHC synthase (menD gene) and O-succinylbenzoic acid-CoA ligase (menE gene) enzymes of Arabidopsis Thaliana. respectively. Therefore, we propose that the menD and menE genes code for the enzymes in the beginning of the phylloquinone biosynthetic pathway. In this paper we describe generation and characteristics of the mutant strains and properties of the PSI complexes in them. We also show that the Ai site of PSI is occupied by plastoquinone. 49

Material and Methods Generation of the menD and menE mutant strains of Synechocystis sp. FCC 6803 To generate a recombinant DNA construction for inactivation of the menD gene. 1.8 kb DNA fragment was amplified from Synechocystis sp. PCC 6803 genomic DNA by polymerase chain reaction using primer menDSNl (5'-GGATCAAGCTTTCAGCACAGC- 3 ) and menDSN2 (5"-ACCTCAGCCTAGGTAATGGC-3'). The 1.8 kb fragment containing the full menD gene and flanking sequences was cloned at EcoRV blunt end in pBIuescript II KS(+) and validity of the insert was confirmed by sequencing. The chloramphenicol resistance gene (CAT) from pUC4C was cloned at a unique EcoRV site within menD coding region. The resultant plasmid was designated menDC and used for transformation of the wild-type strain of Synechocystis sp. PCC 6803. Recombinants were selected on BGl 1 plates containing chloramphenicol C20 ^ig/ml) and isolation of segregated mutants was performed according to the previously published methods(28). Recombinant clones were confirmed for insertional inactivation of menD gene by PCR. For PCR confirmation primers menDSN3 (5'-ACCGCCGTTATTGTTGAGC-3') and menDSN4 (5"- GGTGGTGCATCTTAATTGTCC-3') were used. To generate a recombinant DNA construction for inactivation of the menE gene. 996 bp DNA fragment was amplified from Synechocystis sp. PCC 6803 genomic DNA by polymerase chain reaction using primer menESN3 (5'-TAATGTTGGCAACAGCAGTGG- 3 ) and menESN4 (5"-AGGGTGGAAACATCATTCCC -3') and was cloned at EcoRV blunt end in pBIuescript II KS(+). This clone, designated as menEC. was digested with .Mscl enzyme which has two sites within menE coding region. A 260 bp region was replaced by inserting the chloramphenicol resistance gene at blunt end. The chlorsimphenicol resistance gene (CAT) from pUC4C was cloned at Mscl and Smal. The resultant plasmid was designated as menEC3 and used for transformation of the wild-type strain of Synechocystis sp. PCC 6803. Recombinants were selected on BGll plates containing Chloramphenicol (20 p.g/ml) and isolation of segregated mutants was performed according as above. Recombinant clones were confirmed for insertional inactivation of menE gene by PCR using menSNE3 and menSN4 primers. 50

DNA isolation. PCR, and Southern blotting Genomic DNA from Synechocystis sp. PCC 6803 was prepared as described (29). Hybridization probes were generated with the DIG High-Prime DNA labeling system (Roche Molecular Biochemicals). Hybridization and detection were performed according to the manufacturer's protocols. Cyanobacterial strains and growth A glucose-tolerant strain of Synechocystis sp. PCC 6803 was used as the wild type strain. The wild-type, menD and menE mutant cells were grown in the BG-11 medium, with 5mM glucose and chloramphenicol antibiotic for the mutants (30). Agar plates for the growth of the stock cells were kept at low light intensity (2 to 10 p.E m " s '). Liquid cultures of the wild type and mutant strain were grown heterotrophically under normal light conditions (40 to 60 (lE m " s"') and were bubbled with sterile filtered air. Cell growth was monitored by measuring the optical density at 730 nm (A730) with a Shimadzu spectrophotometer. Cells from liquid cultures in the late exponential phase of growth (A730 = 0.8-1.2) were harvested by centrifiigation at 5000 x g for 15 min. Growth on a large scale was achieved by starting a 3 1 culture at 0.050 A730 and adding 2.5 mM glucose fmal concentration in addition to chloramphenicol for the mutants. Growth rates of the wild-type and mutant cells Cyanobacterial cultures in the late exponential phase were pelleted by centrifugation. washed twice with BGll medium, and suspended in BGll and chloramphenicol at approximately 10 A730 concentration. For estimating growth rates, the cultures were grov\ n in 6 well culture plates with 8-ml liquid medium in each well. All cultures were adjusted to the same initial cell density (A730 = 0.08 - 0.1). Final concentrations of glucose and DCML' were. 5mM. and lOfiM. respectively. The cells were shaken on an orbital shaker at 110 rpm under a bank of fluorescent lights at normal light conditions (40 to 60 p,E m " s '). Isolation ofThylakoid Membranes and PS I Complexes Thylakoid membranes were prepared from cells as described by Sun et al. (31). The thvlakoid membranes were pelleted by centrifugation at 50,000 x g for 90 min and were resuspended in SMN buffer (0.4 M sucrose. lOmM MOPS. lOmM NaCl) for storage. Chlorophyll was extracted from thylakoid membranes and PSI trimers with 80% acetone 51

and determined according to (32). For the isolation of PS I complexes, thylakoid membranes were incubated in SMN buffer with 20 mM CaCU for 0.5 to I.O hr at room temperature in the dark to enhance trimerization of PSI. To the mixture, «-dodecyl-B-D-maltoside (DM) was added to a final concentration of 1.5% (w/v) of chlorophyll amount and incubated in the dark on ice with occasional gentle mixing for 0.5 to 1.5 hrs. The non-solubilized material was removed by centrifugation at 10.000 x ^ for 15 minutes. The trimeric and monomeric PS I particles and PSII were separated by centrifugation in 10 to 30% (w/v) sucrose gradients with 0.04% DM in 10 mM MOPS, pH 7.0. Chlorophyll Analysis and Oxygen Evolution Measurements Chlorophyll was extracted from whole cells and thylakoids with 100% methanol. Chlorophyll concentrations were determined according to (33). Oxygen evolution measurements were p)erformed using a Clark-type electrode as described in (34). Cells were prepared by starting cultures at the same time and grown photomixotrophically with 5mM glucose for two subcultures. The final liquid culture was made without addition of glucose and grown until mid log growth phase (0.4 - 0.6 A730) and then analyzed. The temperature of the electrode chamber was maintained at 25°C by a circulating water bath. Cells were pelleted and washed twice with BGl 1 and resuspended in 40 mM HEPES/NaOH. pH 7.0

buffer in a concentration of 8 - 10 A730 ml '. Measurements were done on 5 A of cells. Whole-chain electron transport (H2O to CO2) measurements were determined after the addition of 5 mM NaHCO,; oxygen evolution mediated by PS n only was determined after addition of 4 mM p-benzoquinone (BQ). Light intensity was 1840 fiE m'" s '. 77 K Fluorescence Emission Spectra The low temperature fluorescence emission spectra were measured using a SLM 8000C spectrofluorometer as described (35). Cells from the exponential phase of growth were harvested by pelleting and resuspended in 25 mM HEPES/NaOH, pH 7.0 buffer. Cells (5 M-g Chi.) were diluted in 25 mM HEPES/NaOH, pH 7.0 to a final volume of 30 ^1 and dark adapted for 30 minutes on ice. To the solution, 70 jil of neat glycerol was added, mixed, prior to quickly freezing in liquid nitrogen. The excitation wavelength was 435 nm. The excitation slit width was set at 4 nm and the emission slit width was set at 2 nm. The emission was scanned from 600 nm to 800 nm twice and averaged. 52

Analysis of phylloquinone using HPLC and mass spectrometry On an equal chlorophyll basis, the samples were prepared and quinones were analyzed as previously described (36). with the following changes. Membranes containing 0.025 mg chlorophyll were centrifuged at 10000 x g for 60 min and the supernatant was removed. PSI trimers were concentrated by a Centracon and then lyophilized to dryness. The pigments were sequentially extracted with 1 ml methanol. 1 ml 1:1 (v/v) methanol;acetone. and I ml acetone, and the three extracts were combined. The resulting solution was concentrated by vacuum at 4°C in the dark to dr>'ness. Dr>' pigment samples were stored at -80 C until used. The pigments were resuspended in a mixture of 1:1 methanol:isopropanol at ca. 0.8 mg Chi ml"'. HPLC separations were monitored with photodiode array UV-visible detection using a Hewlett Packard (Agilent Technologies, Palo Alto, CA) model 1100 quaternary pump and model G1316A photodiode array detector. Sample injections (35 fiL) were made on a 4.6 mm X 25 cm Ultrasphere Cig column (4.6mm x 250mm) with 5 |im packing (Beckman Instruments. Palto Alto, CA) using a gradient elution at (solvent A = methanol; solvent B = isopropanol; 100% A from 0-10 min to 3% A/97% B at 30 min. hold until 40 min.) at 1.0 mL/min. The column was washed after each run (from 100% B to 100% A at 10 min. 0.5 mL/min. 100% A for 5 min at 1.0 mL/min.). A solution of phylloquinone (40 mM) was prepared in absolute ethanol and kept at -20 C as a standard for calibration. Extracts were also analyzed by LC/MS using a Perseptive Biosystems Mariner time-of-flight mass spectrometer using electrospray ionization in negative mode with a needle potential of —3500 V and a nozzle potential of -80 V. A post-column flow splitter delivered column eluent to the electrospray ion source at 10 pL min"'. Spectroscopic characterization of photosystem I complexes Optical kinetic spectroscopy of P700 in purified PSI trimers was performed as described in (37). Q-band EPR Spectroscopy was used to detect photoaccumulated A| signal in PS 1 Trimers as described in (38). Results Genotype of the mutant strains The genotypes of the menD' mutants strain were confirmed by PCR amplification of the appropriate genomic loci and then sequenced. In Fig. 2A (left panel), the genomic region 53

of menD is shov\Ti for the wild type and mutant. A 1.1 kb chloramphenicol resistance cartridge was inserted in the EcoRV site of the menD gene. Using primers within the menD coding sequence containing the EcoRV site. PGR amplification of the menD locus of the wild type produced the expected 990 bp fragment (Fig. 2A. right panel). PGR amplification of the menD mutant locus produced the 2.1 kB fragment, accounting for the integration of the chloramphenicol-resistance cartridge. There was no amplification of the 990 bp fragments in the mutant strain indicating complete segregation of the mutant allele. Southern blot hybridization analysis of the menD mutant confirmed complete segregation in the mutant strain (data not shown). Insertional inactivation of the menE gene was verified by PGR amplification of the menE locus and by DNA sequencing. The left panel of Fig. 2B shows the restriction map of the wild t> pe and mutant menE gene. Using primers within the menE coding sequence, PGR amplification of the menE locus of the wild type produced the expected 980 bp fragment. A 260 bp fragment of the menE gene was deleted and replaced wath a 1.2 Kb chloramphenicol cassette. PGR amplification of the mutant strain produced the expected 1.9 Kb fragment. The wild-type 980 bp fragment was not observed in the mutant. These results indicate that segregation of the alleles had occurred completely in the menE mutant. Absence of phylloquinone in the menD' and menE mutant strains The phylloquinone content of the thylakoid membranes and PSI trimers of the menD' and menE mutants was determined using HPLG with a photodiode array UV-Visible detector. By co-injecting standards and by interpreting the UV-Visible spectra, chlorophyll a was identified at 19.0 min and |5-carotene was identified at 28.5 min. With this solveni gradient, phylloquinone standard elutes at 20.7 min and shows the distinctive doublet peak at 248 nm and 269 nm. The 254 nm chromatogram of the solvent extracted pigments from PSI trimer complexes of the mutant {menE) indicate plastoquinone with an elution time 29.7 min and peak at 256 nm (Fig. 3). In the PSI extracted pigments only plastoquinone was observed in the menD' and menE mutants (Fig. 3). However, the wild type PSI pigments contains only phylloquinone and lack plastoquinone. To determine if phylloquinone was present elsewhere, the thylakoid membranes were analyzed. The menD' and menE mutant thylakoid membrane solvent extracted pigments also show an absence of phylloquinone (Fig 54

4. inset and Fig. 4. respectively). Inset of Fig. 4. wild type pigments contained both the phylloquinone and plastoquinone peak. In the wild type, phylloquinone is contributed from PSI and plastoquinone from PSII and Cy\b(J' which are also thylakoid membrane bound comple.xes that utilize plastoquinone as a lipid soluble electron transporter (39). Calculating the concentrations of chlorophyll to quinone for wild type and mutant yields a ratio close to the known appro.ximateiy 50:1 ChlrPhQ (40). The wild type has a ratio of 42:1 ± 3.6 Chl:PhQ while the meriD' and menE' mutants have a slightly higher ratio of 51:1 ±2.1 and 54:1 ± 4.1 ChhPQ, respectively. The higher ratio could indicate that a small number of the mutant Ai sites are without a quinone. Wild type showed a minute amount (<2%) of plastoquinone in the extracted pigments of the purified trimers. Wild t>'pe PSI trimers that are washed with 3-10 volumes of 10 mM MOPS and 0.04% DM and concentrated do not contain plastoquinone. yet have stiochiometric quantities of phylloquinone (data not shown). Therefore, the detected plastoquinone is probably attributed being captured in the lipid membranes and detergent during PSI purification from the plastoquinone rich thylakoid membranes. UV-Visible detection has a minimum detection limit of phylloquinone for the mutants of approximately 15% that of wild type (27). To achieve smaller detection limits LC-.MS was also used since it has several orders of magnitude greater sensitivity (41). Phylloquinone and plastoquinone eluted at the same times as with the UV-Visible detector with the molecular ion at m/z = 450 and m/z = 748. respectively. Pigments extracted from both thylakoid membranes and from purified PSI trimers of the menD' and menE mutants show only plastoquinone. Phylloquinone was not detected in either membrane or PSI trimer preparation (data not shown). However, this indicates that plastoquinone is only present in the mutant trimers. not actually being utilized by PSI in electron transport. We address this by directly probing the AI site by monitoring the back reaction kinetics from the iron sulfur clusters to P700^ and photoaccumulation EPR. EPR spectroscopy of photoaccumulated A / The EPR spectra of the photoaccumulated quinones have significant differences between the wild type and mutants (Fig. 6). In both mutants, the prominent phylloquinone 55 methyl hyperfine lines are missing and there is a larger g-anisotropy of Q' than in phylioquinone. menE shown. As shown previously, these changes in the spectrum are consistent with a loss of phylioquinone. which is replaced with plastoquinone. in the Ai site (38). Not only is plastoquinone found in stiocheometric amounts in PSI, but it appears to be able to be reduced by Ao- P~00' optical recombination kinetics In PSI trimers isolated from the wild type, the reduction of the P700" RC is multiphasic after a saturating flash (37.42). When measured by 810 nm near-IR optical spectroscopy and in the absence of external electron acceptors, the t>'pical wild type spectrum shows a majority of the P700* being reduced with a lifetime of 86 ms and a minor 12 ms phase. The biphasic decay is considered to be a result of the PSI complexes being in different confirmations (12,22). The microsecond decay is attributed to damaged reaction centers. There is also a long-lived (>1200 ms) kinetic phase of P700"^ reduction from direct reduction of the P700" by reduced DCPIP that contributes to 5 - 9 % of the total absorbance change. In PSI trimers isolated from the menD' and menE mutants, the reduction of P700^ is also multiphasic. Fig. 5. When menE is measured in the absence of external electron acceptor. P700~ is reduced with lifetimes of approximately 2.6 ms and 10 ms in a similar ratio to the wild t>'pe. Lifetimes of the menD' mutant strain is similar, at 2.7 ms and 9.9 ms. Both mutants have a relative contribution of the 3 ms to 10 ms phases of 65:30. The mutants also exhibit a minor long-lived kinetic phase attributed to electron donation of DCPIP to the oxidized P700^. The kinetic back reaction phases for both the menD' and menE mutants are consistent with plastoquinone being in the A| site of PSI and acting as an electron intermediate from An to Fx. (37). Growth of the mutant strains E.xchanging plastoquinone from phylioquinone in PSI has cellular effects primarily involved in growth rates and high light sensitivity'. Focusing on the phenotype of the menD' and menE mutant strain of whole cells reveals interesting differences to the wild type strain. Under normal light intensity conditions (40 - 60 p.E m"" s"'). the menE mutant cells grew photoautotrophically but a slower rate than wild type cells. The 80 h doubling time of the menE is significantly slower 56

than the 26 h wild type (Table 1). Yet menD' mutant was only slightly slower at 30 h. At higher light levels (120 - 160 (lE) the mutant menE is incapable of growing. However. menD' mutant does grow at the same high light. Photomixotrophic growth rates of the mutants were determined in the presence of 5 mM glucose, which allows both respiration and photosynthesis to provide energy for growth. The 14 h wild type doubling time was 42 h for menE' and 15 h for menD' mutant strains. Photoheterotrophic. respiration with 10 (iM DCMIJ to inhibit PSII. shows doubling times of both menD' and menE mutants were similar to those of the wild type strain. Only the photosynthetic energy conversion has changed leaving respiration to function normally. Clearly menD' mutant strains growth rates behaved more like wild type than a phylloquinone deficient mutant (27). These studies indicate that the mutant is capable of photoautorophic growth with high light sensitivity in menE mutant was alleviated by inhibiting PSII activity. However, the menD' mutant seemed to display more wild-t>'pe phenotype resulting from differences in photosynthetic electron transfer rates and the ratio of PSI to PSII. described below. Electron transfer rates in whole cells To determine if the differences in growth rates is a product of photosynthetic ability, we measured the activities of PSI and PSII in the mutant strains for whole chain and partial chain electron transfer activitv'. Whole-chain electron transfer from water to bicarbonate was measured in cells grown photomixotrophically under normal light conditions. On an equivalent cell number basis, the rates of whole chain oxygen evolution was 73% and 77%. to the wild type strain for the menD' and menE mutants, respectively (Table 1.). PSII activity was monitored by adding a soluble quinone electron acceptor. 1.4-benzoquinone. The PSII activity of menE was 81% of wild type. However, the PSII rates in menD was significantly reduced (27%). The menD' and menE mutant chlorophyll levels per cell were 85% and 65% of wild type, respectively. Given the reduced amount in chlorophyll level the electron transport is suitably decreased in whole chain and PSII activity in menE mutant. However, the relatively high amounts of PSII activity in the mutant may prove to be stressful to the cells as indicated by slower growth rates high light sensitivity. However in the menD' mutant the decreased PSII activity was in balance with PSI resulting in light tolerance and growth rates approaching wdld type. 57

Relative content of PS! per cell Observing the whole cell 77 K fluorescence emission spectra on an equal chlorophyll basis for the wild type strain and menE mutant reveals a change (Fig. 6). The PSII chlorophyll fluorescence emission at 685 and 695 nm show similar intensities between wild type and the mutant with changes occurring in the 721 nm PSI emission. The PSI to PSII v\as reduced in the menE mutant (2.2) relative to wild t>'pe (3.3). The relative increase of PSII in the mutant produces an excess of reductant overwhelming PSI. At higher light intensities this can prove toxic (27). Yet the menD' mutant has an increased PShPSII ratio of 4.0. This supports the reduced PSII activity in whole cells. Discussion Previously, latter steps in the phylloquinone biosynthetic pathway (menA. menB) in cyanobacteria has been confirmed by mutationally disabling the genes in the Synechocystis sp. PCC 6803 genome from compared homologue genes which code for enzymes involved in menaquinone biosynthesis in other bacteria (Fig. 1) (27). Disabling the menA and menB genes resulted in a complete loss of phylloquinone. The loss of the menB gene, naphthoate synthase. pre\ents the second ring closure and menA gene, phytyl transferase, disallows phytylation of the naphthoquinone ring, both critical steps. Because steps at the end of the phylloquinone biosynthetic pathway have been established, this study focuses on the earlier stages of phylloquinone production. These genes were first determined in £. coli. (15,19) and by analogy the homologs were found in the genome of Synechocystis sp. PCC 6803. The menD' gene codes for the enz\'me. 2- succinyl-6-hydroxy-2.4-cyclohexadiene-l-carboxylate (SHCHC) synthase, which catalyzes isochorismic acid to 2-succinyl-6-hydroxy-2,4-cyclohexadiene-l-carboxylate while concurrently releasing pyruvate and carbon dioxide. Precursor to naphthoate synthase enz\ me. the menE gene codes for the enzyme. O-succinylbenzyol-CoA synthetase, which attaches a CoA to O-succinyl benzoic acid. To determine if the enzymes encoded by the menD' and menE genes function in phylloquinone biosynthesis, we engineered mutants by targeted inactivation of the menD and menE genes in Synechocystis sp. PCC 6803. PCR amplification and sequencing were performed on both samples, with southern blot analysis on menD', to confirm the absence of 58

a complete menD and menE gene in the mutants (Fig. 2). HPLC-MS and HPLC-UV.'Vis showed that neither thylakoid membranes (Fig. 4) and purified PSI trimers (Fig. 3) of the menD and menE contained detectable levels of phylloquinone. This concludes that the menD and menE genes in Synechocystis sp. PCC 6803 genome code for essential enzymes in phylloquinone biosynthesis. It also suggests that there are no other biosynthetic routes to producing phylloquinone which circumvent SHCHC synthase or 0-succinylbenz\'ol-CoA synthetase. Although it has not been confirmed by disabling all the genes in the biosynthetic pathway, we believe that SHCHC synthase is the first enzyme dedicated to phylloquinone biosynthesis. Precursors to the products of SHCHC synthase, chorismate and isochorismate are common molecules employed in not only phylloquinone biosynthesis, but production of ubiquinone, aromatic amino acids, and serine. Therefore a toxic buildup in the cells of these two molecules is unlikely. Both menD' and menE mutants contain plastoquinone in place of phylloquinone within the PSI complex. Extracts of the PSI pigments of the mutants show a ratio of ChlrPQ of about 50:1. same as the known ChlrPhQ values from Synechococcus (40). The slight difference in wild type and mutant ratio of Chl:Q may be attributed to a small number of empty or damaged Ai sites in the mutants. Plastoquinone's unsaturated C-40 tail may also not as readily fit into a C-20 phytyl pocket in PSI. therefore keeping a portion of the A| sites free of plastoquinone. However, actual identification of plastoquinone being utilized by PSI in the Ai site was determined by EPR and monitoring the kinetic back reactions. Photoaccumulated EPR of PSI trimers of the menD' and menE mutants revealed that the spectrum is identical to the plastoquinone containing menA' and menB' mutants (38). This indicates that the electron is being passed from Ao to Ai, reducing PQ to PQ". To check whether the electron is transferred through the A| site into the iron-sulfur clusters (Fx. FA, FB), we obser\'ed the kinetic back reaction by monitoring the reduction of P700"^. Observed at 810 nm. the reduction of P700* in the menD' and menE mutants showed a multiphasic backreaction consistent with an electron originating from FA/FB and passing through PQ (37). A mixture of plastoquinone-containing PSI complexes (~3 ms) and a small amount of damaged RCs (-10 ms) was observed (Fig. 5). The contribution of the plastoquinone and damaged RC for 59

both mutants are similar (31%) to the menA and menB mutants. The mutants contain a greater number of damaged PSI RCs than in the wild type (10%). This suggests that the PSI RCs are operating with piastoquinone in the Ai site, similar to the menA' and menB' mutants. The imbalance in electron transport rates is seen markedly in the differences in 77 K fluorescent PSIrPSIl ratios. Wild type had a 3.3 PSIiPSII ratio with menE lower at 2.1 (Fig. 7). The menE mutant is similar to the menA' and menB" phylloquinone-less mutants in the PSI:PSII ratio and high light sensitivity (27). The resulting physiological changes in grovvih rates and oxygen evolution are consistent with the menA' and menB' mutants. The menD' mutant has a PSIrPSII of 4.0. The excess PSI is more capable of compensating for the PSII production of reductant. Stepping back, the whole cells were monitored for whole chain oxygen evolution (water to bicarbonate) and for PSII activity. On an equivalent cell number basis, the menB mutant rates of whole chain oxygen evolution and PSII activity were 77% and 81% of wild r\ pe. respectively. Whereas, the menD' has a similar whole chain activity (73%) to menE. the PSII activity is significantly reduced (27%) to the wild type. The chlorophyll levels per cell of the menE mutant is reduced. 67% of the wild type. Given the lower chlorophyll content of the menE mutant the PSII activity is relatively high compared to the wild type. The menD' mutant, however, has a whole chain activity comparable to PSII activity not three times greater as in the wild type and menE strains. Given piastoquinone reduced efficiency to act as an electron pathway intermediate in PSI (27). menD' has prevented an over production of reductant by decreasing the PSII levels, as indicated by fluorescence. This imbalance is translated into phenotypic changes in the cell. To study the physiology and growth characteristics of the cells, the menD' and menE mutant strain were grown under several conditions. The photoautotrophic and photomixotrophic growth rate is significantly slower than wild type when grown at normal light levels (40 nE m'* s"'). The menE. but not menD'. cells are incapable of growing at high light (>120 |jE m** s"'). Both mutants have similar doubling times to wild type when grov\Ti photoheterotropically with glucose and DCMU to block PSII function. This suggests that slower growth rate and high light phototoxicity in menE strain is an indirect effect caused by an imbalance of the rates of electron transport between PSI and PSII. The menD' strain 60

has compensated by adjusting the PSI to PSII levels. The possible secondan' site mutation responsible for this adaptation is being investigated. Both the ntenD' and menB deletion mutants prevents phylloquinone synthesis. These results parallel the deletion mutations in the menA and menBT genes later in the biosynthetic pathway of phylloquinone (27). Essentially, for the phylloquinone deletion mutants the end result is the same: plastoquinone replaces phylloquinone in the Ai site and is active although at a reduced rate and cells are light sensitive from a relative increase in PSII activity as seen from oxygen evolution measurements and fluorescent ratios of PSI:PSII. References

1. Furie. B., Bouchard. B. A... and Furie. B. C. (1999) Blood93(6), 1798-808 2. Li. L., Darden. T. A.. Freedman, S. J.. Furie, B. C.. Furie. B., BaJeja. J. D., Smith. H.. Hiskey. R. G.. and Pedersen. L. G. (1997) Biochemistry 36(8), 2132-8 3. Rigby. A. C.. Baleja, J. D.. Li. L., Pedersen. L. G.. Furie. B. C.. and Furie. B. (1997) Biochemistry 36(50), 15677-84 4. Booth. S. L.. and Suttie. J. W. i\99S)JNutr 128(5), 785-8 5. Manna. P.. and Chitnis . P. R. (1999) in Concepts in Photobiology: Photosynthesis and Photomorphogenesis (Singhal. G. S.. Renger. G.. and Sapory. S. K.. eds). pp. 212-251 6. .Malkin. R. (1996) in Oxygenic Photosynthesis: the Light Reactions (Ort. D. R.. and Yocum. C. F.. eds). pp. 313-332. Kluwen Dordrecht 7. Chitnis. P. R. (1996) in Plant Physiol. Vol. 111. pp. 661-669 8. Malkin. R. (1986) FEBS Lett. 208, 343-346 9. Schoeder. H. U., and Lockau. W. (1986) FEBS Lett. 199, 23-27 10. Takahashi. Y.. Hirota. K.. and Katoh, S. (1985) Photosynth. Res. 6, 183-192 11. Biggins. J., and Mathis. P. (1988) Biochemistry 27(5), 1494-500 12. Yang. F.. Shen. G.. Schluchter. W. M...; Zybailov. B. L.. Ganago. A. O.. Vassiliev. I. R.. Brvant. D. A., and Golbeck. J. H. (1998) J. Phys. Chem. B 102(42), 8288-8299 13. Iverson. T. M.. Luna-Chavez. C.. Cecchini. G., and Rees, D. C. (1999) Science 284(5422), 1961-6 14. Luna-Chavez. C.. Iverson. T. M.. Rees, D. C.. and Cecchini. G. (2000) Protein Expr Purif 19(\), 188-96 15. Palaniappan. C.. Sharma, V.. Hudspeth. M. E.. and Meganathan. R. (1992) J Bacteriol 174(24), 8111-8 61

16. Sharma. V.. Suvama. K.. Meganathan, R.. and Hudspeth, M. E. {1992) JBacteriol 174(15), 5057-62 17. Sharma, V.. Meganathan, R.. and Hudspeth, M. E. i\992) J Bacteriol 175(15), 4917- 21 18. Daruwala, R., Kwon, O., Meganathan. R.. and Hudspeth, M. E. (1996) FEMS Microbiol Lett 140(2-3), 159-63 19. Sharma, V.. Hudspeth. M. E., and Meganathan, R. (1996) Gene 168(1), 43-8 20. Suvama. K., Stevenson, D., Meganathan, R., and Hudspeth, M. E. {\99%) J Bacteriol 180(10), 2782-7 21. Driscoll. J. R.. and Taber. H. W. {1992) J Bacteriol 174(15), 5063-71 22. Hill. K. F.. Mueller. J. P.. and Taber, H. W. (1990) Arch Microbiol 153(4), 355-9 23. Palaniappan. C.. Taber. H.. and Megzinathan. R. {1994) J Bacteriol 176(9), 2648-53 24. Taber. H. W.. Dellers. E. A., and Lombardo, L. R. (1981) y Bacteriol 145( 1), 321 -7 25. Koike-Takeshita, A., Koyama, T., and Ogura, K. (1997) J Biol Chem 272( 19), 12380-3 26. Kaneko, T., Sato, S.. Kotani. H., Tanaka, A., Asamizu, E., Nakamura. Y.. Miyajima. N.. Hirosawa. M., Sugiura, M., Sasamoto, S., Kimura, T., Hosouchi, T., Matsuno. A.. Muraki, A., Nakazaki. N., Naruo, K., Okumura, S., Shimpo, S., Takeuchi, C., Wada, T.. Watanabe. A.. Yamada, M., Yasuda, M., and Tabata, S. (1996) DNA Res. 3, 109-136. 27. Johnson. T. W., Shen. G.. Zybailov, B., Kolling, D., Reategui. R.. Beauparlant. S.. Vassiliev. I. R., Bryant. D. A., Jones, A. D.. Golbeck. J. H.. and Chitnis. P. R. (2000) J Biol Chem 275(12), 8523-30 28. Chitnis. V. P.. Xu, Q., Yu, L., Golbeck, J. H., Nakamoto, H., Xie. D. L.. and Chitnis. P. R. {\992)JBiol Chem 268(16), 11678-84 29. Smart. L. B.. Anderson, S. L., and Mcintosh. L. (1991) EMBO J. 10( 11), 3289-96 30. Rippka. R.. Deruelies, J.. Waterbury, J. B.. Herdman, M., and Stanier. R. Y. (1979) J. Gen. Microbiol. Ill, 1-61 31. Sun. J.. Ke. A.. Jin, P., Chitnis, V. P.. and Chitnis. P. R. (1998) Meth. Enzymol. 297, 124-139 32. Arnon. D. (1949) Plant Physiol. 24, 1-14 33. MacKinney. G. (1941) J. Biol.Chem. 268, 11678-11684 34. Shen. G.. Eaton-Rye, J. J., and Vermaas, W. F. (1993) Biochemistry 32(19), 5109-15 35. Shen. G. Z.. and Bryant. D. A. (1995) Photosynth. Res. 44, 41-53 62

36. Johnson. T. W., Shen. G.. Zybailov, B.. Kolling. D., Reategui, R., Beauparlant. S.. Vassiliev. I. R.. Bryant. D. A., Jones, A. D.. Golbeck, J. H.. and Chitnis. P. R. (2000) J Biol Chem 275(12), 8523-30 37. Semenov. A. Y.. Vassiliev. I. R., van Der Est. A.. Mamedov. M. D.. Zybailov. B.. Shen. G.. Stehlik. D., Diner, B. A.. Chitnis. P. R.. and Golbeck. J. H. (iQOO) J Biol Chem 275(31), 23429-23438 38. Zybailov. B.. van der Est, A., Zech, S. G.. Teutloff. C., Johnson. T. W.. Shen. G.. Bittl. R.. Stehlik. D.. Chitnis. P. R., and Golbeck, J. H. {2QQQ) J Biol Chem 275(12), 8531-9 39. Hope, A- B. (2000) Biochim Biophys Acta 1456(1), 5-26 40. Schubert, W. D., Klukas, O.. Krauss. N., Saenger, W., Fromme, P.. and Witt, H. T. (1997) y Mol Biol 272(5), 741-69 41. Careri. M., Mangia. A., Manini. P.. and Taboni, N. (1996) in Fresertius' J. Anal. Chem. Vol. 355, pp. 48-56 42. Vassiliev. I. R.. Jung, Y. S., Mamedov. M. D., Semenov. A., and Golbeck. J. H. (1997) Biophys 7 72( 1), 301-15 63

TABLE I. Physiological Characteristics of the Synechocystis sp. PCC 6803 wild-type and menD' and menE mutant strains

wild type menD menE

Doubling Time (hrs) 40 |iE. m"-* s"'

Photoautotrophic Growth 26 ± 3.0 30 ± 1.2 80 ± 8.2 Photoheterotrophic Growth 14 ± 1.5 15 ± 0.8 42 ± 1.2 Photoheterotrophic Growth + DCMU 20 ± 2.9 20 ± 1.7 22 ± 1.2

Photosynthetic Rates

(fimol 02/(5 OD7,o* h *1) Whole Chain 2420±180 1770 ±530 1860±190 PS Il-mediated 7110 ±560 1960 ±660 5790 ± 450 (.(imol 0;/(Chl.* h *1) Whole Chain 708 ± 62 612 ±47 815 ± 83 PS Il-mediated 2070±130 676 ± 42 2450 ± 190 64

FIGURE LEGENDS FIG. I. Proposed biosynthetic pathway of phylloquinone biosynthesis in Synechocystis sp. PCC 6803. The gene products responsible for the biosynthesis of menaquinone were initially described in E. coli fl9). The homologs of these genes have been identified in the genome sequence of Synechocystis sp. PCC 6803 and menA, menB. (36) and menG were confirmed. FIG. 2. A. Left Panel. The restriction map of the Synechocystis sp. PCC 6803 genomic region around the menD wild type (top) and disrupted gene (below). A. Right Panel. Electrophoretic analysis of the DNA fragments amplified by PCR from genomic DNA of the wild type and mutant gene. Molecular weight marker left lane. B. Left Panel. The restriction map of the Synechocystis sp. PCC 6803 genomic region around the menE wild type (above) and disrupted gene (bottom). B. Right Panel. Electrophoretic analysis of the DNA fragments amplified by PCR from genomic DNA of the wild type and mutant gene. Molecular weight marker left lane. Each PCR fragment was sequenced to confirm its identity. FIG. 3. HPLC profiles of pigment extracts from lyophilized PS I complexes of the menE mutant strain of Synechocystis sp. PCC 6803. The pigments were separated on a 5.0 |j.m ultrasphere Cig reverse phase column and detected from 190 nm to 800 nm by a photodiode array. The detection wavelength shown was 254 nm. The inset spectrum from the mutant shows a peak that co-elutes with plastoquinone at 29.7 min. FIG. 4. HPLC profiles of pigment extracts from lyophilized thylakoid membranes of wild type, menD and menE mutant strain of Synechocystis sp. PCC 6803. The pigments were separated on a 5.0 (im ultrasphere Cig reverse phase column and detected from 190 nm to 800 nm by a photodiode array. The detection wavelength shown was 254 nm. The inset chromatograms compare wild type and menD to the full menE chromatogram in the region of the PhQ peak (20.5 min.). FIG. 5. P700* reduction kinetics in PS I complexes isolated from menE mutant. Flash induced optical transient measured at 810 nm after a single flash. The sample cuvette (10 mm X 10 mm) contained 3.0 ml of PSI complex at 50 ^g/ml Chi suspended in 25 mM Tris. pH 8.3 with 0.04% 3-DM, 10 (iM sodium ascorbate and 4|i,M DCPIP. Excitation 65

wavelength. 542 nm. excitation energy. 1.4 mJ. A 300 MHz bandwidth was used in the preamplifier to recover kinetics in the fis time range. FIG. 6. Photoaccumulated and Simulated Q-band CW EPR spectra of Ai' and Q in PS I complexes isolated from wild type and menD mutant. Top is compiled spectra and the bottom compares wild-type and menD. The photoaccumulation was carried out for 40 min at 205 K. Instrument settings: microwave power 1 mW. microwave frequency 34.056 GHz. modulation frequency 100 KHz. modulation amplitude 1 G. temperature 205 K. time constant 10 mS. conversion time 10 mS. 100 averaged scans. FIG. 7. Fluorescence emission spectra at 77 K of whole cells from Synechocystis sp. PCC 6803 wild type and menE mutant strains. Spectra were recorded at the same cell density and normalized relative to PS II. Each spectrum was the average of the four measurements. The excitation wavelength was 435 nm, which excites mostly chlorophyll. PS II and its accessory pigments exhibit emission maxima at 685 and 695 rmi; PS I has a maximum emission at 730 nm. 66

SHCHC »jngiw NO COOH COOH hacnoramaM Synnaw // 7^ ,COOM WeiMieCwoei? CH; 2 Vfvmtm CO2 2-«ucani^.6-nydroxy-2.4- CnonamaM isocnonimic add pMotiaxidlene-Kaffeexyute OHuoenyi bansDcacid HjO mtriGftMomSymwm 1)0HNA •M1127 exMceMmaimjn CoASiiOmn

ccr CoA II PPi*AMP CaA*ATP 2}0HNA^C^ O \.^-Oi*f^roK^-rm0OvH'C^ (OHNA) o-Mobnyl btnzoyt-CoA o-cuociny( benzoic abd Prrytyt-PP 1 DHNA pnyM MenA/tinsiS

pw^ 0«Trt«tnvionyttoQt;incAc> f^tfrynrerstcrMe M«nGXj«cC2Wl16S3

7^ S-adena«yi-L- s-«danosyM.- mMlMnine homocyiliim 2-phytyl-1.4-naptnnoqiiinone

Fig. 1 67

MW \VT m&tD' i4o^i u™.i Hindi Msc! Hpal bpei ECORV MSCI menD 2.1 kb

0.99 kb

MW \\T menE' EcoRl M-scI Stral

menE

1.9 kb

0.98 kb

Cam*"

Fie. 2 68

600 -

Plastoquinone n = 9 500 - 30-

20- 400 - 10- A (254 nm, n—I—I—I—r mAU) 300 - 250 300 350 400 450 Chi Wavelength (nm)

200 -

100 -

0 H TTTT ii|iiiiiiiii|iiiiiiiti|iiii|iiii|iii 16 18 20 22 24 26 28 30 32 Time (min)

Fig. 3 69

300-1

WT menD 250- PhQ

200-

20 20 22 24 (254 nm. 150- mAU) Chi ;i-carotene PQ 100-

50-

15 20 25 30 35 Time (min)

Fig. 4 70

0.2 -| 0.1 - 0 — -0.1 - -0.2 -

menE 2.7 ms

6 - 10.3 ms 4 -

2 - 1400 ms

-3 •2 .0 1 ,2 3 4 Time (ms)

0.2 n 0.1 - 0 — 0.1 -

10 - menD

8 - 2.6 ms

6 - 9.9 ms 4 - 3300 ms 2 -

•2 1 0 1 2 ,4 10 10 Time (ms)

Fig. 5 71

0.8 -

0.6 - menD

0.4 -

0.2 -

0.0 -

-0.2 -

-0.4 -

-0.6 -

2.010 2.008 2.006 2.004 2.002 2.000 1.998 9

Fig. 6 72

PS I

Wild Type menE" menD'

10

PSIl

5

600 650 700 750 nm

Fig. 7 73

CHAPTER 4. RECRUITMENT OF A FOREIGN QUINONE INTO THE A, SITE OF PHOTOSYSTEM I I. GENETIC AND PHYSIOLOGICAL CHARACTERIZATIONOF PHYLLOQUINONE BIOSYNTHETIC PATHWAYMUTANTS IN SYNECHOCYSTIS SP. PCC 6803'

A paper published in the Journal of Biological Chemistry*. T. Wade Johnson'''^, Gaozhong Shen^ \ Boris Zybailov^ Derrick Koiling^ Ricardo Reategui". Steve Beauparlant^, Ilya R. Vassiliev^ Donald A. Bryant^ A. Daniel Jones'", John

H. Golbeck^'" and Parag R. Chitnis"'"

Summary Genes encoding enzymes of the biosynthetic pathway leading to phylloquinone, the secondary electron acceptor of Photosystem (PS)^ I, were identified in Synechocystis sp. PCC 6803 by comparison with genes encoding enzymes of the menaquinone biosynthetic pathway in Escherichia coli. Targeted inactivation of the ntenA and metiB genes, which code for phytyl transferase and 1,4-dihydroxy-2-naphthoate synthase, respectively, prevented the synthesis of phylloquinone, thereby confirming the participation of these two gene products in the biosynthetic pathway. The metiA' and menB' mutants grow photoautotrophically under normal-light conditions (40 fiE m'^ s"') with doubling times twice that of the wild type, but are unable to grow under high-light conditions (160 fiE m*- s"'). The menA' and menB' mutants grow photoheterotrophically on media supplemented with glucose under low-light conditions with doubling times similar to that of the wild type but are unable to grow under high-light

This work is supported by grants from the National Science Foundation (MCB 9723661 to JHG, MCB 9723001 to PRC, MCB 9723469 to DAB, and DBI 9602232 to Penn State University). Journal Paper No. J-18740 of the Iowa Agriculture and Home Economics Experiment Station, Ames, Iowa, Project No. 3416 and supponed by Hatch Act and State of Iowa funds. " Reprinted with permission of the Journal of biological Chemistry, 275, 8523-8530, (2000) ' ^Dcpa|tmcni of Biochcmistr>'. Biophysics and Molecular Biology. Iowa State University, Ames. I.A 50011 Departm^i of Biochemistry and Molecular Biology, The Pennsylvania Slate University. Univer.sii\ Park. PA 16802 Department of Chemistry, The Pennsylvania State University, University Park, PA 16802 The authors thank Bruce Diner for his generous gift of authentic plastoquinone-9

^Abbreviations'. PS I, photosystem I; PS II, photosystem 11; EPR, electron paramagnetic resonance; HPLC. high performance liquid chromatography; GC, gas chromatography; MS, mass spectroscopy; DPIP, 2.6- dichlorophenol-indophenol; PCR, polymerase chain reaction; SDS, sodium dodecyl sulfate; PAGE, polyacrylamide gel electrophoresis; Tris, tris(hydroxymethyI)aminomethane; Tricine. N-[2-hydroxy-1,1- bis(hydroxymethyI)ethyl]gIycine; DPPH, a, a'-diphenyl-P-picryl hydrazyl; SAM, S-adenosyl methionine; SAH, S-adenosyl homocysteine. 74 conditions unless atrazine is present to inhibit PS II activity. The level of active PS II per cell in the menA' and menB' mutant strains is identical to that of the wild-type, but the level of active PS 1 is about 50 to 60% that of the wild type as assayed by low temperature fluorescence, P700 photoactivity, and electron transfer rates. PS I complexes isolated from the menA' and menB' mutant strains contain the full complement of polypeptides, show photoreduction of FA and FB at 15 K, and support 82 to 84% of the wild-type rate of electron transfer from cytochrome cg to flavodoxin. HPLC analyses show high levels of plastoquinone-9 in PS I complexes from the menA' and menB' mutants but not from the wild type. We propose that in the absence of phylloquinone, PS I recruits plastoquinone-9 into the A, site, where it functions as an efficient cofactor in electron transfer from A(j to the iron-sulfur clusters. Introduction All well-characterized photosynthetic reaction centers (RC) are known to contain a bound quinone molecule that participates in the early stages of photochemical charge separation and stabilization (1,2, 3). Type 11 reaction centers, such as Photosystem II (PS II) or those present in the purple non-sulfur bacteria, contain a bound benzoquinone or ubiquinone as the secondary electron acceptor. Type I reaction centers, such as Photosystem I (PS I) of cyanobacteria and green plants, contain a bound menaquinone. usually phylloquinone (Vitamin Ki, 2-methyl-3-phytyl-l,4-naphthoquinone) or less-commonly, 5'- monohydroxy-phylloquinone. as the secondary electron acceptor (4). (Whether green sulfur bacteria and heliobacteria. which have a PS Hike reaction center, contain a similar bound quinone is still under active investigation.) Two molecules of phylloquinone can be extracted per molecule of P700 from isolated PS I complexes (5, 6, 7, 8, 9); however, only one molecule of phylloquinone is considered to participate as an intermediate in electron transfer from Ao to Fx (5. 6. 10). One strategy to disallow A| fianction is to inactivate genes that code for enzymes involved in the proposed pathway of phylloquinone biosynthesis. Many prokaryotes as well as chloroplasts contain the metabolic pathway for phylloquinone (vitamin Ki) biosynthesis. In several bacteria, vitamin KT (menaquinone) is used during fumarate reduction in anaerobic respiration (11). The genes encoding the enzymes involved in the conversion of chorismate to menaquinone have been cloned in Escherichia coli (12, 13, 14) and Bacillus subtilis (15. 16). Although the route of phylloquinone biosynthesis has not been described in cyanobacteria. the pathway is likely to be similar to the pathway of menaquinone biosynthesis in other bacteria. Menaquinone differs from phylloquinone by the presence of a partly unsaturated. 75 predominantly C-40 side-chain rather than a mostly saturated, C-20 phytol side-chain. With this exception, the synthesis of the naphthalene nucleus in phylloquinone and menaquinone is expected to include similar steps (Figure 1). The genome database for Synechocystis sp. PCC 6803 (17) contains homologs for several genes that encode enzymes for menaquinone biosynthesis: menF {entO (isochorismate synthase), menD (2-succinyl-6-hydroxy-2.4- cyclohexadiene-1-carboxylate synthase), menE (O-succinylbenzoic acid-CoA ligase), menB (DHNA synthase) and menA (identified as 'menaquinone biosynthesis protein', but probably a phytyl transferase). Possible homologs of menC and ORF241 (the DHNA thioesterase) have also been identified in our database searches. We propose that menF/entC, menD, menE and menB o( Synechocystis sp. PCC 6803 are involved in 1,4-dihydroxy-2-naphthoate synthesis whereas the product of a metiA homologue catalyzes the addition of phytyl chain. The product of the phytyl transferase, 2-phytyl-1,4-naphthoquinone, requires a methylation step to become phylloquinone. The gene originally identified as gerC2 (sllI653) in the Synechocystis sp. PCC 6803 database probably codes for the 2-phytyl-1,4-naphthoquinone methyl transferase enzyme that catalyzes this reaction. There is no known function for menaquinone in cyanobacteria except to provide a precursor for phylloquinone biosynthesis. PS II and respiration require pIastoquinone-9. which is a benzoquinone derivative that is synthesized by an independent pa±way (18). There are two expected consequences of an interruption in the menaquinone pathway: 1) the participation of men genes in the biosynthetic pathway of phylloquinone in cyanobacteria would be confirmed; 2) the Ai site should be empty, thereby allowing a test of the requirement of phylloquinone in electron transfer from Ai to Fx- We initially generated mutants in which the nietiA and menB genes in Synechocystis sp. PCC 6803 have been inactivated by targeted mutagenesis. These two mutations were selected because it would additionally allow us to determine whether the phytyl chain is essential for function. If the phytyl chain is dispensable, then the menA deletion mutant may allow the head group 1,4-dihydroxy-2-naphthoate to be incorporated into the Ai site. In this paper we describe the construction as well as the genetic and functional characterization of menA' and menB' mutant strains of Synechocystis sp. PCC 6803. We propose that in the absence of phylloquinone the A, site does not remain empty. Instead, we suggest that PS I recruits plastoquinone-9 into the Ai site, and this quinone supports high-efficiency electron transfer from Ao to the iron-sulfur clusters. 76

Experimental Procedures Growth of Synechocystis sp. PCC 6803 Wild-type Synechocystis sp. PCC 6803 cells were grown in medium BG-11 (19). The metiA and menB' mutant strains were grown in medium BG-II-TES supplemented with 5 mM glucose and the appropriate antibiotic. Agar plates for the growth of mutant strains were kept at low light intensity (2 to 5 ^.E m-2 s"'); liquid cultures of wild type and mutant strains were grown under normal light conditions (30 to 50 ^lE m'2 s*') in presence of 5 mM glucose. Cell growth in liquid cultures was monitored by measuring the optical density at 730 nm using a Cary-i4 spectrophotometer that had been modified for computerized data acquisition by On- Line Instruments, Inc. (Bogart, OA). Cells from liquid starter cultures in the late exponential phase of growth (OD730 = 0.8-1.2) were harvested by centrifiigation and were washed once with BG-11 medium. All cultures were adjusted to the same initial cell density (OD730 =0.1) for growth experiments and bubbled with air as described (20). Generation of the menA and menB' mutant strains of Synechocystis sp. PCC 6803 To generate a recombinant DNA construction for inactivation of the menA gene, two DNA fragments were amplified from Synechocystis sp. PCC 6803 genomic DNA by polymerase chain reaction (Figure 2A). The Pstl restriction sites were incorporated in both fragments, whereas the Apa\ restriction site was also added to the downstream fragment. The first amplification product was digested with Eagl and Psth whereas the second fragment was digested with Pstl and Apal restriction enzymes. The fragments were ligated with the pBluescript SK (Stratagene) vector that had been digested with Eagl and Apal. The kanamycin-resistance gene from pUC4K was cloned in the Pstl site; this yielded recombinant plasmid pRRA that contained replacement of a 442-bp part of the menA gene by the resistance marker. Transformation of the wild-type strain of Synechocystis sp. PCC 6803 and isolation of segregated mutants was performed according to previously published methods (21). To generate a recombinant DNA construction for inactivating the menB gene, two 1.0- kb fragments from upstream and downstream of the menB gene were amplified by PCR (Figure 2B). The oligonucleotide primers included unique restriction sites (Spel and EcoRJ in the upstream fragment and EcoRJ and Xhol sites in the downstream fragment). The amplified fragments were cloned into pBluescript (Stratagene), and a 2.0-kb EcoRI fragment containing the streptomycin/ spectinomycin-resistance cassette of the Q. element of the recombinant plasmid pHP45Q (22) was inserted into the newly created unique EcoRJ site. The resultant 77

plasmid was designated pRRB and was used to produce the menB mutant strain according to previously described methods (21). DNA Isolation, PCR and Southern Blotting Genomic DNA from Synechocystis sp. PCC 6803 was prepared as described (20). Hybridization probes were generated with the DIG High-Prime DNA labeling system (Boehringer Mannheim-Enzo Diagnostics Inc., Indianapolis, IN). Hybridization and detection were performed according to the manufacturer's protocols. Chlorophyll Analysis and Oxygen Evolution Measurements Chlorophyll was extracted from whole cells and thylakoids with 100% methanol, and chlorophyll concentrations were determined as described (23). Oxygen evolution measurements were performed using a Clark-type electrode as described (24). The temperature of the electrode chamber was maintained at 30°C by a circulating water bath. Cells were resuspended in 25 mM HEPES/NaOH, pH 7.0 buffer in a final concentration of 1.0 OD730 ml"'. Whole-chain electron transport (H2O to CO2) measurements were determined after the addition of 5 mM NaHCOs. Oxygen evolution mediated by PS II only was determined after addition of 4 mM p-benzoquinone. Isolation ofThylakoid Membranes and PS I Particles Thylakoid membranes were prepared from cells in the late exponential growth phase as described (20). Cells were broken by two passages through a French pressure cell at 20.000 lb in-~ at 4°C. The thylakoid membranes were pelleted by centrifiigation at 50,000 x g for 45 min. The thylakoid membranes were resuspended in thylakoid buffer (50 mM HEPES/NaOH. pH 8.0; 5 mM MgCU, 10 mM CaCU, 0.5% (v/v) dimethyl sulfoxide and 15% (v/v) glycerol) for storage, and/or in 50 mM Tris/HCl, pH 8.0 for further PS I particle preparations. For the isolation of PS I complexes, thylakoid membranes were solublized in 1% (w/v) w-dodecyl-B- D-maltoside (DM) for 2 to 4 hrs at 4°C. The trimeric and monomeric PS I panicles were separated by centrifugation on 5 to 20% (w/v) sucrose gradients with 0.03% DM in 50 nxVl Tris. pH 8.0. Further purification was achieved by a second centrifugation on sucrose gradients in 50 mM Tris, pH 8.0 in the absence of DM (25). SDS-PAGE Analysis The methods used for SDS-polyacrylamide gel electrophoresis were identical to those previously described (24). To resolve the subunit compositions of the PS I preparations from the Synechocystis sp. PCC 6803 wild-type strain and the menA' and menS' mutants, the Tricine/Tris discontinuous buffer system was used (35). A 16% (w/v) acrylamide gel 78 containing 6 M urea was used as the separating gel. The resolved proteins were visualized by silver staining. 77 K Fluorescence Emission Spectra Low-temperature fluorescence emission spectra were measured using a SLM 8000C spectrofluorometer as described (24). Cells from the exponential phase of growth were harvested and resuspended in 25 mM HEPES/NaOH, pH 7.0 buffer. Cells were diluted in 25 mM HEPES/NaOH, pH 7.0 containing 60% (v/v) glycerol to a concentration of 1.0 OO730 ml"' prior to freezing in liquid nitrogen. The excitation wavelength was 440 nm. The excitation slit width was set at 4 nm and the emission slit width was set at 2 nm. PS I Activity Measurements Steady-state rales of electron transfer for isolated PS I complexes were measured using cytochrome c^ as electron donor and flavodoxin as electron acceptor as described (10). The measurement of P700 photooxidation in whole cyanobacterial cells was performed as described (27). EPR Spectroscopy of and FB Electron paramagnetic resonance (EPR) studies were performed using a Bruker ECS- 106 X-band spectrometer and a standard-mode resonator (ST 8615) which is equipped with a slotted pon for light entry. Cryogenic temperatures were maintained with a liquid helium cryostat and an ITC-4 temperature controller (Oxford Instruments. UK). The microwave frequency was measured with a Hewlett-Packard 5340A frequency counter, and the magnetic field was calibrated using DPPH as the standard. Sample temperatures were monitored by a calibrated thermocouple located 3 mm beneath the bottom of the quartz sample tube and referenced to liquid Ni. Samples were illuminated with a 150 W arc source (Oriel. Stratford. CT, Model 66057) passed through 5 cm of water and a heat-absorbing color filter to remove the near-IR light. Samples used for EPR measurements contained I mg Chi ml '. 1 mM sodium ascorbate. 30 M DCPIP in 50 mM Tris. pH 8.3. Analysis of phylloquinone using HPLC-UV/Vis and mass spectrometry Membranes containing 0.1 mg chlorophyll were centrifuged at 1000 x g for 60 min and the supernatant was removed. The membrane pigments were sequentially extracted with 1 ml methanol, 1 ml 1:1 (v/v) methanolracetone. and 1 ml acetone, and the three extracts were combined. The resulting solution was concentrated by vacuum at 10°C in the dark to ca. O.S mg Chi ml '. Chromatography with UV/Vis detection was performed on an ISCO dual pump HPLC system (Lincoln, NE). The pumps were operated by ISCO Chemresearch version 2.4.4 software, UV/VIS detection was performed with an ISCO V4 absorbance detector set at 79

255 nm, and data collection and processing was done using JCL6000 version 26 software (Jones Chromatography Limited, UK). HPLC separations were also monitored with photodiode array UV-visible detection using a Hewlett Packard (Palo Alto, CA) model 1100 quaternary pump and model G1316A photodiode array detector. Sample injections (20 nL) were made on a 4.6 mm x 25 cm Ultrasphere C|g column (4.6mm x 250mm) with 5 fim packing (Beckman Instruments, Palto Alto, CA, using gradient elution (solvent A = methanol; solvent B = isopropanol; 100% A from 0-10 min to 3% A/97% B at 30 min, hold until 40 min.) at 1.0 mL min '. A solution of phylloquinone (40 mM) was prepared in absolute ethanol and kept at -20C as a standard for calibration. Extracts were also analyzed by LC/MS using a Perseptive Biosystems Mariner time-of-flight mass spectrometer using electrospray ionization in negative mode with a needle potential of -3500 V and a nozzle potential of -80 V. A post- column flow splitter delivered column eluent to the electrospray ion source at 10 min '. GC/MS analyses were performed using a Hewlett Packard 5972 mass spectrometer coupled to a Hewlett Packard 5890 gas chromatograph. Split-less injections of 1.0 ^iL were made onto a 30 meter DB-5 column (J&W Scientific, Folsom, CA) using helium (35 cm sec ') as carrier gas. The column was programmed from 100* C to 300*C at a rate of 6* per min. Data were acquired in both full scanning mode and using selected ion monitoring of m/z = 450 for trace detection of phylloquinone. Results Analysis of the genotype of the menA" and menB' mutant strains The genotypes of the menA and menB' mutant strains were confirmed by Southern bloi hybridization analyses and by PCR amplification of the appropriate genomic loci. The left panel of Figure 2A shows restriction maps of the genomic regions surrounding the menA gene in the wild t>'pe and mutant strains. A 440-bp fragment in the menA gene was deleted and replaced by a I.3-kb kanamycin resistance cartridge encoding the aphll gene. Using primers flanking the coding sequence (small arrows in Figure 2A), PCR amplification of the menA locus of the wild type produced the expected fragment of 1070 bp (Figure 2A. right panel). PCR amplification of the menA locus of the mutant strain produced the expected 1.9-kb fragment (Figure 2A. right panel). Since no amplification of the 1070-bp fragment occurred when DNA from the mutant strain was used as a template, this result indicates that the wild type and mutant menA alleles had fully segregated. Southern blot hybridization analyses were also performed, and these experiments confirmed that the menA gene was interrupted as expected and that the menA' mutant strain was homozygous (data not shown). 80

Inseitional inactivation of the menB gene was also verified by both Southern blot hybridization analysis and by PCR amplification of the menB locus from the mutant strain. As shown in the left panel of Figure 2B. most of the menB gene was deleted and replaced with a 2-k;b spectinomycin-resistance cartridge. Using primers flanking the coding sequence (small arrows in Figure 2B), PCR amplification of the menB locus of the wild type produced the expected fragment of 930 bp (Figure 2B, right panel). However, PCR amplification of this locus in the mutant strain produced a 2.2-kb DNA fragment (Figure 2B. right panel), and no 930-bp fragment was observed. These results indicate that the menB" mutant is homozygous and that full segregation of alleles had occurred. The PCR amplification results were confirmed by Southern blot hybridization analyses which demonstrated that the menB mutant strain was homozygous and that the menB gene had been insertionally inactivated as shown in the left panel of Figure 2B (data not shown). Analysis of the phenotype of the menA" and menB" mutant strains We focused initially on the phenotypic analysis of whole cells of the menA' and menB' mutant strains. Photoautorotrophic growth rates of the menA and menB' mutants were measured in cells grown in BGl 1-TES medium, which contains minerals and bicarbonate as the sole carbon source. Under normal-light intensity conditions (40 |iE m"2 s*'), the menA' and menB' mutant cells grew photoautotrophically but at a slower rate than the wild-type cells. The 86-h doubling time of the menA mutant cells and the 77-h doubling time of the menB mutant cells were more than twice the 26-h doubling time of the wild-type cells (Table 1). Under high-light intensity conditions (160 (lE m"2 s"'), the doubling time of the wild-type cells was 29-h, but surprisingly the menA' and menB' mutant cells showed no measurable growth. Photomixotrophic growth rates of the menA' and men'B mutants were determined in the presence of 5 mM glucose, which allows both respiration and photosynthesis to provide energy for growth. Under low light intensity conditions, the 12-h doubling time of the wild- type was one-half that under photoautotrophic conditions, and the 26 to 15-h doubling times of the menA' and menB' mutants were fractionally greater than that of the wild-type (Table 1 ). Under high-light intensity conditions, the wild-type cells had a doubling time of 23-h. but again the menA' and menB mutant cells showed no measurable growth. Photoheterotrophic growth rates of the menA' and menB" mutants were determined in the presence of glucose to allow respiration but with 20 (iM atrazine to inhibit PS II function. Under both low and high light intensity conditions, the doubling limes of the menA' and menB' mutants were slightly greater than those of the wild-type strain (Table 1). These studies show that the menA' and menB' mutants are capable of photoautotrophic growth only under low light intensities. 81

However, the high-light sensitivit>' of the menA' and menB' mutants can be alleviated by inhibiting PS 11 activity. Complementation o/menA" and menB" mutant strains with wild-type genes To determine whether secondary mutations account for the observed high-light sensitivity phenotype, we complemented the menA' and menB' mutant strains with the corresponding wild-type genes. The menA' and menB' genes and their flanking regions were amplified by PCR and used for transformation of the corresponding mutant strains. After transformation of the mutant strains with the appropriate DNA fragment, the plates were incubated under non-selective, photoautotrophic conditions at an intermediate light intensity of 40 jiE m'- s*'. At this light intensity, the menA' and menB' mutant cells could grow under photoautotrophic. photomixotrophic, and photoheterotrophic conditions (data not shown). After three days, the plates were shifted to high light intensity (160 m-^ s*'), and colonies appeared after about two weeks. After restreaking under high light intensity growth conditions to ensure that segregation had occurred, selected independent transformants were tested for loss of antibiotic resistance and subjected to PCR analysis to verify that the menA or menB locus had been transformed to the wild-type genotype. Using this approach, antibiotic- sensitive. complemented strains of the menA' mutant (AWT) and menB' mutant (BWT) were isolated. These transformants, which were tolerant to high-light intensities, had growth rates and photosynthesis rates (measured by oxygen evolution) that were similar to those of the wild type (data not shown) and distinctly different from the menA' and menB' mutant strains. It was also possible to isolate suppressor mutants by omitting the transforming DNA. However, these suppressor mutants arose at a much lower frequency than was observed in the transformation experiments, retained their antibiotic resistance, and had growth and electron transport properties that differentiated these strains from the wild type (data not shown). Analysis of these suppressor mutants is in progress. We conclude that the phenotype observed for the menA' and menB' mutant strains is entirely due to the inactivation of the targeted genes. Electron transfer rates in whole cells We next explored the possibility that the high-light sensitivity of the menA' and nwnB mutant strains arises from a result of an imbalance between PS I and PS II electron transfer rates. The activities of PS I and PS II can be measured in the menA' and menB' mutant strains by whole chain and partial chain electron transfer assays. Whole-chain electron transfer from water to bicarbonate was measured in cells grown photomixotrophically under low-light conditions. On an equivalent cell-number basis, the rates of oxygen evolution in the menA and menB' mutants were 69% and 49%, respectively, those of the wild type strains (Table 1). This 82 result was as surprising as the ability of the mutants to grow photoautotrophically; if phylloquinone were absent, then room temperature electron transfer to NADP* should not occur (28, 29). Since their PS II activities are essentially unaffected, these results indicate that the decreased, whole-chain, electron transport activity of the menA' and menB' mutant cells is due to a PS I-related defect. This could be the result of a lower amount of PS I per cell, an impairment in PS I function, or both. Relative content of active PS I per cell Figure 3 shows the 77 K fluorescence emission spectra of whole cells on an equal cell- number basis for the wild-type strain and for the menA' and menB' mutants. In the menA and menB' mutant cells, the PS O-chlorophyll fluorescence emission at 685 nm and 695 nm (30) shows no obvious differences in intensity from the wild-type cells. However, the PS I- chlorophyll fluorescence emission at 721 nm was reduced in intensity in the two mutant strains relative to the wild type (Fig. 3). This result indicates that cells of the two mutant strains contain the same amount of PS II, but less PS I per cell, than the wild type. The absolute PS I content of whole cells can be determined by the light-induced absorbance increase at 832 nm due to the oxidation of P700 (31, 32). On the basis of equal ceil numbers, both the menA' and menB' mutant cells were found to contain between 50 and 60% of the photooxidizable P700 of the wild-type cells (data not shown). In Synechocystis sp. PCC 6803, ca. 100 Chi are associated with PS I and ca. 60 Chi are associated with PS II. If the smaller amount of photooxidizable P700 is due to fewer PS I complexes per cell, then the chlorophyll content should be lower in the menA' and menB' mutant cells than in the wild- type cells. Table I shows that, on an equivalent cell-number basis, the chlorophyll content of the menA' and menB' mutant cells is significantly lower than that of the wild-type cells. The reduced contents of PS I per cell in the mutant strains are therefore responsible, at least in part, for the lower whole-chain electron transfer rates in the menA' and menB' mutant cells. Polypeptide composition of isolated PS I complexes PS I complexes were solubilized from thylakoid membranes using /i-dodecyl-6.D- maltoside and purified by ultracentrifugation on two successive sucrose gradients. PS I isolated from the wild-type is 15% monomeric and 85% trimeric, whereas PS I isolated from the menA' and menB' mutants is 35% monomeric and 65% trimeric. Only PS I trimers were used in these studies. Analysis by SDS PAGE of the PS I trimers isolated from the mutant and wild type showed no differences in the polypeptide composition (data not shown). Also, no degradation fragments of PS I proteins were detected in the PS I trimers. These results show 83

thai the interruption of the phylloquinone biosynthetic pathway has no effect on the protein complement of PS I. Absence of phylloquinone in the menA" and menB" mutant strains The phylloquinone content of the PS I trimers was determined using HPLC with photodiode array UV-visible detection. As shown in Figure 4, multiple peaks are present in the 254-nm chromatogram of the solvent extracts from PS I complexes of wild-type cells. By co-injecting standards and by interpreting the UV-visible spectra, chlorophyll a was identified at 18.0 and 24.0 min (the former is missing the phytyl tail), a polar carotenoid (probably monohydroxylated but otherwise uncharacterized) was identified at 22.2 min, and B-carotene was identified at 28.8 min. Phylloquinone was identified by co-elution at 20.7 min with an authentic phylloquinone standard (Figure 4, top inset) and by its UV-visible spectrum (Figure 4 bottom inset). As shown in Figure 5, virtually identical chromatograms were obtained for solvent extracts of PS I complexes from the menA' and menB' mutants except that the phylloquinone peak at 20.7 min is missing. As expected, phylloquinone was present in the membrane pigment extracts of the complemented AWT and BWT strains at levels comparable to the wild type (data not shown, but similar to Figure 4). Based on a phylloquinone calibration curve, the minimum detection limit for phylloquinone using HPLC was estimated to be ca. 15% of the wild-type levels. GC/MS is capable of separating and detecting non-polar benzoquinones, ubiquinones, and naphthoquinones provided their molecular masses are less than about 600 Da. Using total ion current for detection, the crude solvent extract from wild-type PS I complexes showed a peak with a retention time of ca. 8 min. which matched that of authentic phylloquinone (data not shown). The molecular ion at m/z = 450 confirmed the identification of this molecule as phylloquinone. Sensitive selected-ion-monitoring analyses did not find a detectable amount of phylloquinone in solvent extracts of PS I complexes isolated from either the menA' or menB strains. The limit of detection using selected-ion-monitoring was determined from the calibration curve to correspond to ca. 0.1% of the wild-type level. Since there are ca. 100 Chl/P700 in cyanobacterial PS I complexes, the menA' and menB' mutant strains thus contain < 0.02 phylloquinones per P700. Presence of plastoquinone-9 in the menA" and menB" mutant strains The idea that a foreign quinone might be present in the A| site first came about when we discovered that a quinone-Iike EPR signal is present in whole cells of the menA' and menB mutants (see companion paper in this series (33)). To determine the identity of this quinone. solvent extracts of PS I trimers from the menA' and menB' mutants were analyzed by HPLC using photodiode array UV-visible detection. The search was initially complicated by the 84

absence of new peaks in chromatograms (?L = 270 nm) from the menA' and metiB' mutants when compared with the wild-type (Figure 5). We therefore sought evidence of a new component coeluting with another pigment by comparing the UV-visible spectra of peaks in chromatograms of the menA' and nienB' mutants with the corresponding peaks for the wild type. The only significant difference was in a component that coeluted with B-carotene at 29 min. The difference spectrum of the components eluting at 29 min showed a strong absorbance near 254 nm that was lacking in the wild-type (Figure 5, bottom inset). This is the spectral region in which the biologically-occurring benzoquinones, ubiquinones, and naphthoquinones absorb strongly, but in which B-carotene has relatively weak absorbance. We noted that the UV spectrum of the coeluting component was similar to pIastoquinone-9. a quinone that is present at 10-fold higher concentration than phylloquinone in thylakoid membranes (34). Indeed, we found that authentic plastoquinone-9 co-elutes with, and has a UV spectrum that matches, the peak at 29 min (Figure 5, top inset). Sensitive selected-ion-monitoring analyses of the HPLC eluate at the mass of plastoquinone-9 (m/e = 748) showed a peak at this retention time. We consistently found levels of plastoquinone-9 in trimeric PS I complexes from the menA' and menR mutants in amounts similar to phylloquinone in PS I complexes from the wild type. In contrast, we found none or a very small amount of plastoquinone-9 in PS I complexes from the wild type. Full-scan HPLC/MS analyses of the mutants showed that none of the other peaks displayed mass-spectral characteristics of related naphthoquinones such as biosynthetic precursors of phylloquinone. Low temperature reduction ofF^ and Fg The ability of PS I trimers from the menA' and menB' mutants to transfer electrons from F700 to the iron-sulfur clusters was determined at low temperature by EPR spectroscopy. When the samples were frozen in darkness and illuminated at 15 K, the relative spin concentrations of reduced FA {g = 2.05, 1.94, 1.85) and FB ig = 2.07, 1.92. 1.88) were identical to those for PS I complexes isolated from the wild-type (Figure 6). The ratio of FA to

FB reduced was also identical in the mutants and the wild type. Thus, the absence of phylloquinone in the Aj site does not effect low-temperature electron transfer from A,,' to the terminal iron-sulfur clusters. When PS I complexes from the menA' and menB' mutants were subjected to photoaccumulation conditions by freezing the sample during illumination. FA and

FB were completely reduced as shown by the presence of an interaction spectrum (^-values of 2.05. 1.94, 1.92, 1.88), with a total spin concentrations similar to that of the wild-type, li should be noted that the quantum yield cannot be determined in these studies because multiple turnovers of the PS I complex occur during continuous illumination. Nevertheless, the 85

quantitative reduction of FA and Fg does indicate that the entire population of mutant PS I complexes is competent in electron transport. Fluvodoxin reduction rates in PS I complexes Although the lower rates of whole-chain electron transfer in the metiA' and menB' mutant cells can be explained by a lower PS I content per cell, there still remains the possibility that the efficiency electron transfer in individual PS I complexes is altered by the absence of phylloquinone. Steady-state rates of electron transfer were determined in PS I trimers by measuring the rate of flavodoxin reduction with cytochrome cg as electron donor as a function of light intensity (Figure 7). The rates at saturating light intensity were determined by treating light as a substrate in a Michaelis-Menten kinetic analysis. The maximal rate of flavodoxin reduction that could be sustained was found to be 8460 ^imol mgChl ' h ' in the wild-type PS I complexes, 7128 ^imol mgChl ' h"' in the PS I complexes of the menA' mutant, and 6948 limol mgChl ' h ' in the PS I complexes of the menR mutant. Assuming 100 Chi per P700 in all PS I complexes, these maximal rales of electron transport correspond to 235 e' PS I "' s ' in the wild type, 198 e' PS I"' s"' in the menA' mutant and 193 PS F'S"' in the menB' mutant. As shown in Figure 7, the light-saturation dependence of the electron transfer rates in the PS I complexes of the menA' and menB' mutants strains is similar to that for the wild-type complexes, indicating that the relative quantum efficiencies of the PS I complexes are not affected by the mutations. These results show that, in spite of the absence of phylloquinone in the AI site, electron transfer throughputs in PS I complexes isolated from the menA' and menB' mutants are 82 to 84% as efficient as in PS I complexes isolated from the wild-type strain. Discussion Although the phylloquinone biosynthetic pathway in cyanobacteria has not been previously described, the nucleotide sequence of the Synechocystis sp. PCC 6803 genome shows the existence of homologues of the menA, menB, menC, menD, menE, and menF ieniC). and menG genes which code for enzymes involved in menaquinone biosynthesis in other bacteria. Because they encode enzymes that function near the end of the biosynthetic pathway, we focused exclusively on the menA and menB genes in this study. The menB gene of E. coli codes for 1,4-dihydroxy-2-naphthoic acid synthase, which catalyzes the formation of the bicyclic ring system by converting o-succinylbenzoyl-coenzyme A to 1.4-dihydroxy-2- naphthoic acid. Menaquinone differs from phylloquinone by the presence of a partly unsaturated, C-40 isoprenyl tail rather than a mostly saturated, C-20 phytyl side chain attached to the naphthoquinone nucleus. The menA gene of E. coli codes for l,4-dihydroxy-2- 86 naphihoate octaprenyl transferase, which catalyzes ligation of the C-40 isoprenyl chain to the C2 position of the naphthoate moiety. The low degree of identity (17%) in the primar>' sequences of the MenA proteins of E. coli and Synechococcus sp. PCC 6803 is consistent with the difference in the substrate specificity of these enzymes. Because of the structural similarity between menaquinones and phylloquinone, we worked from the premise that the menA and menB genes code for proteins that function in phylloquinone biosynthesis. To test this premise, we engineered mutants by targeted inactivation of the menA and menB genes in Synechocystis sp. PCC 6803. Southern blot hy bridization and PGR analyses were used to confirm the absence of a complete menA or menB gene in the mutants. HPLC/MS and GC/MS showed that the membranes of the menA and menB' mutant strains do not contain detectable levels of phylloquinone. Hence, one firm conclusion of this study is that the menA and menB homologues in the Synechocystis sp. PCC 6803 genome code for essential enzymes in the phylloquinone biosynthetic pathway. A corollary to this conclusion is that no other biosynthetic routes to phylloquinone exist beyond naphthoate synthase in Synechocystis sp. PCC 6803. The menA' and menB' mutants showed similar biochemical and physiological characteristics. Both lacked phylloquinone and both contained plastoquinone-9 in their PS I complexes. These results further suggest that the phytyl chain of phylloquinone is required for its stable assembly into PS I complexes in vivo. These observations confirm the participation of the menA and menB gene products in the phylloquinone biosynthesis pathway in Synechocystis sp. PCC 6803. To study their physiology and growth characteristics as well as the role of phylloquinone in photosynthetic electron transfer, the menA' and menB' mutant strains were grown under a variety of conditions. Both mutant strains grew photoautotrophically at low to moderate light intensities (30 m"- s"' and 50 |iE m'^ s"') but failed to grow either photoautotrophically or photo-mixotrophically when the light intensity exceeded 1(X) |iE m'- s* '. Because photoheterotrophic growth occurs in the mutants at high light-intensities when atrazine is present, excess reductant produced by PS II is proposed to be the cause of the failure to grow at high light intensities. Atrazine binds competitively to the QB site in PS II. blocks the high rate of damaging reductant and/or oxidant formation, and allows the cells to survive the toxic effect of light and to use glucose as the source of reduced carbon. We have found that while the amount of PS U is unchanged relative to the wild-type, the amount of functional PS I per cell in the menA' and menB' mutants is ca. 50% lower than in the wild- type. The observed phototoxicity may therefore be an indirect effect caused by an imbalance of the rates of electron transport between PS II and PS I. Indeed, we have recently obtained 87 several second-site suppressor mutants that allow a phylloquinone-less mutant to grow photoautotrophically under high light intensity (Chitnis P. and Golbeck, J. unpublished results). Most of these mutants have reduced PS 11 activity, thereby supporting our postulate that the PS II toxicity is responsible for the inability of the phylloquinone-less mutants to grow under high light intensity. The cause of reduction in the PS I level in the metiA' and menB' mutant cells could be a decreased rate of assembly or an increased turnover rate of PS I complexes. However, an examination of the degradation rate of the PS I apoproteins in the mutant and wild type strains did not show significant differences (data not shown). We therefore suggest that the absence of phylloquinone affects the rate of assembly of PS I complexes by influencing one or more steps involved in its biogenesis. The ability of the menA' and menB' mutants to grow at low to moderate light intensities agrees with the finding that the absence of phylloquinone did not abolish room temperature photosynthetic electron transfer activity through PS I complexes isolated from these mutants. The mutant strains contained less PS I than the wild type on a per cell basis; additionally, the steady-state rates of electron transfer from cytochrome c^ to flavodoxin in PS I complexes from the mutants were high but were 82 to 84% of the wild-type rate. Therefore, the lower rate of whole-chain electron transfer in the mutant cells is a combination of both effects. In PS I complexes isolated from the menA and menB mutants, electron transfer from P700 to the terminal iron-sulfur clusters is quantitative at cryogenic temperatures. However, in PS I complexes where phylloquinone has been partially extracted using solvents, the maximum amount of irreversible charge separation after a large number of flashes is independent of the number (0. 1, or 2) of phylloquinone molecules per PS I complex (28). Hence, single turnover optical studies of P7(X) turnover at room temperature will be necessary to confirm the slightly lower quantum efficiency of PS I electron transfer in the menA' and menB' mutants. The inescapable conclusion from this study is that phylloquinone is not required for efficient electron transfer in PS I at either room or cryogenic temperatures. We considered two explanations for the high rates of PS I activity. One is that the mutant PS I complexes differ from the solvent-extracted PS I complexes in allowing room temperature as well as low temperature electron transfer in the absence of phylloquinone. This bypass may be direct or it may involve a redox-active amino acid in the transmembrane domain of the PsaA and PsaB polypeptides. The second possibility is that a foreign quinone has been recruited into the A; site and that it may participate in electron transfer from Ao to Fx- This quinone, which may be identical to the plastoquinone-9 identified on the basis of the solvent extraction studies described here, may substitute for phylloquinone in the A| site, thereby promoting electron 88

transfer from Fx to FA/FB- Spectroscopic evidence supporting the presence of plastoquinone-9 in the A| site and its participation in forward electron transfer is provided in the second paper of this series (33). References 1. Nitschke, W. and Rutherford, A. W. (1991) TIBS 16. 241-245 2. Golbeck, J. H. (1993) Proc. Natl. Acad. Sci. U. S. A. 90, 1642-1646 3. Nugent, J. H. (1996) Eur. J. Biochem. 23, 519-531 4. Ziegier, K., Maldener. I. and Lockau, W. (1989) Z. Naturforsch. C: Biosci. 44, 468-472 5. Malkin. R. (1986) FEBS Lett. 208. 343-346 6. Schoeder. H. U. and Lockau, W. (1986) FEBS Lett. 199, 23-27 7. Lockau. W.. Schoeder. H. U., Nitschke, W. and Ziegier, K. (1987) in Prog. Photosynth. Res., Proc. Int. Congr. Photosynth., 7th, Meeting Date 1986, Volume 2, 37-40. Edited by: Biggins, John. Nijhoff: Dordrecht, Neth.. Regensburg 8400. Fed. Rep. Ger. 8. Takahashi. Y., Hirota, K. and Katoh, S. (1985) Photosyn. Res. 6, 183-192 9. Biggins, J. and Mathis. P. (1988) Biochemistry 21, 1494-1500 10. Yang, F.. Shen. G.. Schluchter, W.M., Zybailov, B.L., Ganago, A.O., Vassiliev, L. Bryant, D. and Golbeck, J. H. (1998) J. Phys. Chem. B. 102, 8288-8299 1 LIverson, T.M., Luna-Chavez, C., Cecchini, G. and Rees, D. C. (1999) Science 284. 1961-1966 12.Sharma. V.. Suvama, K., Meganathan, R. and Hudspeth. M. E. (1992) J. Bacterial. 174. 5057-5062 13. Palaniappan. C.. Sharma, V.. Hudspeth, M. E. and Meganathan, R. (1992) J. Bacteriol. 174, 8111-8118 14. Sharma. V.. Hudspeth, M. E. and Meganathan, R. (1996) Gene 168, 43-48 15. Driscoll. J. R. and Taber. H. W. (1992) J. Bacteriol. 174, 5063-5071 16. Hill, K. F.. Mueller, J. P. and Taber, H. W. (1990) Arch. Microbiol. 153, 355-359 17. Kaneko. T.. Sato. S.. Kotani. H.. Tanaka. A.. Asamizu, E., Nakamura, Y.. Miyajima. N.. Hirosawa. M.. Sugiura. M.. Sasamoto. S.. Kimura, T., Hosouchi, T.. Matsuno, A.. Muraki. A.. Nakazaki. N.. Naruo. K., Okumura, S.. Shimpo. S., Takeuchi, C.. Wada. T.. Watanabe. A.. Yamada. M.. Yasuda. M. and Tabata, S. 1996. DNA Res., 3. 109- 136 18. Schultz. G.. Soli. J.. Fiedler. E. and Schultz-Siebert. D. (1985) Physiol. Plant. 64. 123- 129. 19.Rippka. R., Deruelles, J., Waterbury, J.B., Herdman, M. and Stanier, R.Y. (1979) J. Gen. Microbiol. Ill, 1-61 89

20.Shen, G. Z., Eaton-Rye, J. J. and Vermaas, W. F. J. (1993) Biochemistry 32, 5109- 5115 21. Chitnis. V. P.. Xu. Q., Yu, L., Golbeck, J. H., Nakamolo, H., Xie, D. L. and Chitnis. P. R. (1993) J. Biol. Chem. 268, 11678-11684 22.Prentki. P. and Krisch. H.M. (1984) Gene 29, 303-313 23. MacKinney, G. (1941) J. Biol. Chem. 140, 315-322 24.Shen. G. Z. and Bryant, D. A. (1995) Photosynth. Res. 44, 41-53 25. Golbeck. J. H. (1995) in: CRC Handbook of Organic Photochemistry and Photobiology (Horspool, W. H. and Song, P. S.. eds), pp. 1407-1419, CRC Press, Bocca Raton. FL 26. Schagger, H. and von Jagow, G. (1987) Anal. Biochem. 166, 368-379 27. Redding, K.. Coumac, L., Vassiliev, I. R., Golbeck, J. H., Peltier, G. and Rochaix. J. D. (1999) J. Biol. Chem. 274, 10466-10473 28.Setif, P., Ikegami, I. and Biggins, J. (1987) Biochim. Biophys. Acta 894, 146-156 29. Biggins, J.. Tanguay, N. A. and Frank, H. A. (1989) FEBS Lett. 250, 271-442 30. Shen, G. Z. and Vermaas, W. F. J. (1994) J. Biol. Chem. 269, 13904-13910 31. Maxwell, P. C. and Biggins, J. (1976) Biochemistry 15, 3975-3981 32 Redding, K., Coumac, L., Vassiliev, I. R., Golbeck, J. H., Peltier, G. and Rochaix. J. D. (1999) J. Biol. Chem. 274, 10466-10473 33. Zybailov, B., van der Est, A., Zech, S., Teutloff, C.,. Johnson, W., Shen, G.. Bittl, R.. Stehlik. D.. Chitnis, P. and Golbeck, J. H. (2000) J. Biol. Chem. (in press) 34. Crane. P., Ehrlich, B. and Kegel, L. P. (1960) Biochem. Biophys. Res. Commun. 3. 37-41 90

Tabk 1

Physiological Characteristics of the Synechocystis sp. PCC 6803 wild-type and the menA' and menB' strains".

strain wild-type menA' menB'

Photoautotrophic Growth Doubling Time (hrs)

Normal light 26 ± 3.0 86 ± 5.0 77 ± 4.0

High b'ght 29 ± 1.4 n.m.'' n.m.

Photomixotrophic Growth Doubling Time (hrs)

Normal light 12±2.1 26 ± 3.8 15 ±2.3

High light 23 ± 0.8 n.m. n.m.

Photoheterotrophic Growth Doubling Time (hrs)

Normal light 18 ±2.2 19±3.1 20 ± 2.5

High light 19 ±2.8 22 ± 3.2 21 ±2.8

Oxygen Evolution" (fimol O2/OD730 h 1)

Whole chain 1077 ±55 750 ±71 529 ± 39

PS Il-mediated 1037 ±70 905 ±137 993 ± 88

Chlorophyll Content' (|ig Chl/dD73o ml) 4.21 ±0.18 3.36 ± 0.23 3.43±0.32

" Values are the average of four independent experiments (« = 4) '' not measurable " Cells were grown in BGl 1 plus 5 mM glucose under reduced light intensity; units are with reference to optical density of the whole cells at 730 nm. 91

Figure Legends FIG. 1. Proposed biosynthetic pathway of phylloquinone biosynthesis in Synechocystis sp. PCC 6803. The gene products responsible for the biosynthesis of menaquinone were initially described in Escherichia coli (see (23) for review). The homologs of these genes that have been identified in the genome sequence of Synechocystis sp. PCC 6803 are indicated. SAM is S-adenosyl methionine and SAH is S-adenosyl homocysteine. FIG. 2. Construction and verification of the menA' and menB' mutant strains of Synechocystis sp. PCC 6803. A. Inactivation of the menA gene. Left panel: Restriction maps of the genomic regions surrounding the menA gene in the wild-tyf)e and mutant strain. The small arrows indicate the position of the PCR primers used to amplify the menA coding sequence. Right panel: Electrophoretic analysis of the DNA fragments amplified from the genomic DNA of the wild-type and menA' mutant strains. B. Inactivation of the menB gene. Left panel: Restriction maps of the genomic regions surrounding the menB gene in the wild-type and mutant strains. The small arrows indicate the position of the PCR primers used to amplify the menB coding sequence. Right panel: Electrophoretic analysis of the DNA fragments amplified from the genomic DNA of the wild-typ>e and menB" mutant strains. FIG. 3. Fluorescence emission spectra at 77 K of whole cells from Synechocystis sp. PCC 6803 wild type and the menA' and menB' mutant strains. Spectra were recorded at the same cell density (1.0 OD730 ml"'). Each spectrum was the average of five measurements. The excitation wavelength is 435 nm, which excites mostly chlorophyll. PS II and its accessory pigments exhibit emission maxima at 685 and 695 nm: PS 1 has a maximum emission at 721 nm. FIG. 4. HPLC profiles of pigment extracts from lyophilized PS I complexes of the wild type strain of Synechocystis sp. PCC 6803. The pigments were separated on a 5.0 jxm Ultrasphere Cig reverse phase column. The detection wavelength was 270 nm. The extract from the wild type shows a peak that co-elutes with authentic phylloquinone at 29.7 minutes. Inset, Upper. UVA'^is spectrum of authentic phylloquinone. Inset, Lower. Near-UVA'^is spectrum of HPLC peak which elutes at 29.7 min. FIG. 5. HPLC profiles analysis of pigment extracts from lyophilized PS I complexes of the menB' mutant strain of Synechocystis sp. PCC 6803. The pigments were separated on a 5.0 (im Ultrasphere Cig reverse phase column. The peak at 37.2 min in the wild type co-elutes with 6-carotene, and shows a spectrum in the visible identical to B-carotene. The peak at 37.2 min in the menB mutant shows an additional UV-absorbing 92 component (Inset, Lower) which co-elutes with plastoquinone-9, and which shows a UV spectrum similar to plastoquinone-9 (Inset, Upper) and a m/e of 748 (not shown). The LC- MS analysis of the menA mutant was similar. FIG. 6. EPR spectra of and FG in PS I complexes isolated from the wild type and the menA' and menB' mutant strains. The samples were resuspended to a concentration of 800 jig Chi a ml ' in 25 mM Tris buffer (pH 8.3) with 15% glycerol. 0.02% DM. 10 mM sodium ascorbate and 4 (iM DCPIP. The spectra were obtained either upon illumination of the sample frozen in the dark (top three spectra) or upon illumination of the samples during freezing from 293 to 15.5 K (bottom three spectra). Background spectra were recorded in dark-adapted samples frozen to 15.5 K and subtracted from the light-induced spectra. Spectrometer settings: microwave power, 20 mW; microwave frequency, 9.478 GHz: receiver gain, 6.3 x lO'*; modulation amplitude, 10 G at 100 kHz: magnetic field: center field. 3480 G: scan width, of 1740 G: 3 scans averaged. FIG. 7. Steady-state rates of flavodoxin reduction in PS I complexes isolated from the wild type and the menA' and menB' mutant strains. Rates of electron transfer in units of (imol mg Chi ' h"' are plotted as a ftinction of light intensity. The rates at saturating light intensity were determined by least-squares fitting to a double reciprocal plot and by extrapolation to infinite light intensity (not shown). 93

Crwnsmate iKOcnorismate ISyritfiase MenFi'EnlC.'sJr0ei7 Isocfrarisinate 2-oxogluiarate

SHCHC Synthase Men0/sn0603

Pyruvate CO2

2-Succin(yl-€-hydroxy-2.4- cycioMexadiene-i -cartwicylate

0-succinylben2oc acid MenC/sfl0409 1 Synmase O-Sucdnylbenzoate 0-succinynBpthoate MenB/sii1127 1 (OMNA) Synthase 1. 4-Oinydroxy~2-na(>lhoyt-CoA DHNA-CoA ORF241/SIM916'' Thioest erase CoA

1. 4-0ihydroxy-2-napCha«: Acid

DHNA 1 Phylyltransferase MerWk/slr1518 DerrtothylptiydcKiuinonc

SAM Defne1hylp^>y1loqui^one Gert::2/MenG.'&01653 Melhyltransferaso

SAH

PhyUoqumooe

Fig. 1. 94

UuB ssr2554 menA hglK menA' Hild type

1900 bp

980 bp menA' 500 bp Pstl PstX

A/"

WT menB- pntA menB 's5r2062 gshB wild type ti/ ^ 2200 bp 920 bp ssr2062 gshB menB" £c<7R1 £roRI

Fig. 2. 95

5000 -i menA menB type

4000

3000 H

2000

-1

1000 H

650 700 750 Emission Wavelength (nm)

Fig. 3. 96

^toqunon* suneare 1200 - m>tf

1000 -

rr T I•' 'I 240 260 320 360 800 - \ fV"I vWd-«rP« 600 - V mAU

400 - •|iai|sii|iii| 240 280 320 360 20.7 200 - (nm

7V- 0-i 1 I I I I I T I I I I I 10 15 20 25 30 35 Retention lime (min)

Fig. 4. 97

800 -

— menB - • w id type

600 mAU

240 280 320 360 240 280 320 360 Wavefengtn (rwq WaveiengQi (nn)

200 - 29.8

OH ^ r T T 10 15 20 25 30 35 Retention time (min.)

Fig. 5. 98

uild type IJiCht

menA

menB

UjSht293 K ->15^ K wild typr

tnenA

menB

7 2,1 2.0 1.9 IJi 1.7 /^•value

Fig. 6. 99

6000

4000

menA menB 2000 -

0 20 40 60 80 100 Light intensity, ret. units

Fig. 7. 100

CHAPTER 5. RECRUITMENT OF A FOREIGN QUINONE INTO THE A, SITE OF PHOTOSYSTEM I IV. IN VIVO REPLACEMENT OF THE NATIVE QUINONE IN THE PHYLLOQUINONE-LESS MUTANTS OF SYNECHOCYSTIS SP. PCC 6803 BY EXTERNALLY SUPPLIED NAPHTHOQUINONES'

A paper to be submitted to the Journal of Biological Chemistry T. Wade Johnson^*'. Boris Zybailov*. A. Daniel Jones''. Stephan Zech. Dietmar Stehlik. John H. Golbeck*° and Parag R. Chitnis^°

Summary When the phylloquinone biosynthetic pathway is altered by interruption of the rnenA and menB genes in Synechocystis sp. PCC 6803 (Johnson, T. W. et al. (2000) J. Biol Chem. 275: 8523-8530; Zybailov, B. et al. (2000) J. Biol. Chem. 275: 8531-8538); Semenov et al. (2000) J. Biol. Chem. 275, 23429-23438), plastoquinone-9 (PQ)^ occupies the A, site and

functions as an efficient electron transfer cofactor from AQ to the FeS clusters in Photosystem I. We show here that phylloquinone can be reincorporated into the A, site by supplementing the growth medium of menB' and men A' mutant cells with authentic phylloquinone. The reincorporation also occurs in menB' and menA' cells that have been treated with protein synthesis inhibitors, as well as in PSI complexes isolated form the menA' and menB' mutants, indicating that phylloquinone displaces PQ in fully formed PSI complexes. This likely occurs because phylloquinone has a higher binding affinity for the A,

' This work is supported by grants from the National Science Foundation (MCB 9723661 to JHG. MCE 9723001 to PRC. MCB 9723469 to DAB, and DBI 9602232 to Penn State University) and to Deutsche Forschengsgemeinschaft SFB 312 Al to DS. Journal Paper No. J-18740 of the Iowa Agriculture and Home Economics Experiment Station, Ames, Iowa, Project No. 3416 and supported by Hatch Act and State of Iowa funds. * "Department of Biochemistry. Biophysics and Molecular Biology, Iowa State University, Ames, LA § 50011 Department of Biochemisiry and Molecular Biology, The Pennsylvania State University, University Park, PA 16802^ Department of Chemistry. The Pennsylvania State University, University Park, PA 16802 ' Abbreviations-. PQ. plastoquinone, plastoquinone-9: PSI, photosystem I; PSII, photosystem II; EPR, electron paramagnetic resonance; HPLC. high performance liquid chromatography; GC, gas chromatography; MS, mass spectroscopy; DPIP. 2.6-dichlorophenol-indophenol; PCR, polymerase chain reaction; SOS, sodium dodecyl sulfate; PAGE, polyacrylamide gel electrophoresis; Tris, tris(hydroxymethyl)aminomethane; Tricine, N-[2-hydroxy-1.1 -bis(hydroxymethyl)ethyl]glycine; 101

site than PQ. Phylloquinone is also reincorporated into the A, site by supplementing the growth medium of menB' mutant ceils with 2-carboxy-l,4-naphthoquinone and 2-methyl- 1.4-naphthoquinone. Both phylloquinone and demethyl phylloquinone are incorporated into the AI site by supplementing the growth medium of menB' mutant cells with 1,4- naphthoquinone and 2-hydroxy-1,4-naphthoquinone. This result indicates that the specificity of the phytyl transferase enzyme is relatively nonspecific in terms of the groups present at the 2- and 3-positions of the 1,4- naphthoquinone ring. Since phylloquinone and demethylphylloquinone is present after supplementing the growth medium of menB' mutant cells with any of the four naphthoquinones, the phytyl tail is considered necessary for displacement of PQ from the A, site. These finding open the possibility of incorporating novel quinones into the A, site in vivo by supplementing the growth medium of the menB' mutant cells with various quinone derivatives. Introduction Substituted quinones are frequently employed as electron and proton transfer cofactors in the electron transport chains (1-3). The quinones are comprised of a relatively polar ring, the so-called 'head group', which consists of either a benzoquinone (BQ) or a naphthoquinone (NQ), and a non-polar isoprenoid 'tail' of various chain lengths. Benzoquinones such as plastoquinone-9 and ubiquinone (UQ) function in membrane-bound protein complexes either as fixed or exchangeable electron/proton carriers during photosynthetic and respiratory electron transport. In photosystem II (PSII), PQ functions as a bound one-electron acceptor in the Q^ site and as an exchangeable 2-electron/2-proton acceptor in the Q^ site. The reduced PQH, is subsequently exchanged from the Qg site by an oxidized PQ, diffuses into the membrane, and is oxidized and deprotonated by the cytochrome complex. Photosynthetic reaction centers (RCs) of purple bacteria use UQ (e.g. Rhodobacter sphearoides) or menaquinone (e.g. Rhodosprillulum viridis) in a similar role. In photosystem I (PSI), phylloquinone (PhQ), a substituted NQ with a 20-carbon. largely saturated, phytyl tail, functions as bound one-electron acceptor in the A, site roughly analogous to that of Q;^ in PSII. There exist two molecules of PhQ in PSI, but to our knowledge, the second quinone does not function in a manner equivalent to the Q^ quinone in the bacterial RC and PSII. Quinones are therefore extremely versatile; they can function 102

as the interface between electron transfer involving organic cofactors and electron transfer involving iron-sulfur clusters (PSI), or between pure electron transfer and coupled electron/proton transfer involving a second organic cofactor (PSII). Each quinone displays equilibrium binding and redox properties that are different for each site of interaction (4) that are conferred largely by the protein environment. To understand the structural determinants that allow PhQ to function with high efficiency in the A, site of PSI, we embarked on a project aimed at biological replacement of the native PhQ with a variety of benzylquinone and naphthoquinone derivatives. Itoh and coworkers, and Biggins and coworkers used organic solvents to extract phylloquinone from lyophilized PSI complexes, and reconstituted the empty A, site with various benzoquinone, naphthoquinone, and anthraquinone head groups (5-9). The extraction procedure removes some of the chlorophyll and most of the carotenoid molecules along with the two molecules of PhQ, leading to the possibility that structural perturbations might have occurred to the PSI complexes. To circumvent this problem, we developed a genetic technique to accomplish the in vivo removal of the native PhQ. In a series of three papers(10-I2), we reported the generation and characterization of PQ-less mutants of the cyanobacterium Synechocystis sp. FCC 6803. We accomplished this by interruption mutagenesis of two genes, menA which codes for a phytyl transferase, and menB which codes for l,4-dihydroxy-2-naphthoate (DHNA) synthase. Our strategy was to preclude PhQ production with the intent of creating an empty A, site. To the contrary, we found that the A, site in the mutant PSI complex is occupied by a benzoquinone derivative, PQ. Detailed EPR and optical spectroscopic analyses show that the PQ occupies the site at the same distance from PTOO"^ and with the same orientation as phylloquinone in the wild-type, and functions as an efficient intermediate electron transfer carrier between A<, and Fx- The redox potential of PQ in the A, site is more oxidizing than phylloquinone, rendering electron transfer from Q" to thermodynamically unfavorable. However, forward electron transfer nevertheless occurs because of the large, favorable free energy change from Ao" to flavodoxin. In principle, it should be possible to introduce various quinones in the A, site in vivo by utilizing the unmodified phytyl transferase and methyl transferase enzymes in the remaining biosynthetic pathway of the menB' mutant cells. In addition, physiological studies 103

can be performed on living cells and detailed biophysical analyses can be performed on the purified PSI complexes. Material and Methods Cyanobacterial Strains and Growth A glucose-tolerant strain of Synechocystis sp. PCC 6803 was used as the wild type strain. The menA mutant strain lacks functional phytyl transferase and has been previously described (10). The menB' strain used here {menB26') was derived as a second site suppressor of the menB\%' strain reported earlier (10). Colonies of the menB\%' strain were replica-plated on BGll plates with 5 mM glucose. One set was grown at 'normal' light intensity (40 p,E m * s ') while the other was at 'high' light intensity (>100 fiE m'" s '). The menB26' was one of the isolates that could grow at high intensity. This strain showed no changes in the PSI and PSn activities compared to the menBlS' strain. The purified PSI complexes of the menB26' and menBlS' strains showed identical optical and EPR characteristics as well as identical protein and pigment composition. The molecular basis for the suppression of the original menB' phenotype is under investigation. Since the growth of the menB26' revertant was robust under a variety of different conditions, it was used in the supplementation studies described here. The wild-type, menA' and menB26' mutant cells were grown in the BG-11 medium, with kanamycin and spectinomycin added to the media of the menA'and menB26' mutant strains, respectively (13). Agar plates for the growth of the stock cells were kept at low light intensity (2 to 10 |iE m " s '); liquid cultures of the wild-type and mutant strains were grown autotrophically under normal light conditions (40 to 60 fiE m " s '). Cell growth in liquid cultures was bubbled with sterile filtered air and was monitored by measuring the optical density at 730 nm with a Shimadzu spectrophotometer (10). Cells from liquid cultures in the late exponential phase of growth (OD730 = 0.8-1.2) were harvested by centrifugation at 5000 x g for 15 min. Growth Rates of the Wild-Type and Mutant Cells with NQ Supplements Cyanobacterial cultures in the late exponential log phase were pelleted by centrifugation, washed twice with BGll medium, and suspended in BGll and spectinomycin at approximately 10 A730 concentration. For estimating growth rates, the cultures were grown in 6 well plates with 8 ml medium in each well. All cultures were 104 adjusted to the same initial cell density (A730 = 0.1). Final concentrations of naphthoquinone, glucose, and DCMU were 5 (iM, 5 mM, and 10 |iM, respectively. The cells were shaken on an orbital shaker at 110 rpm. Growth was monitored by measuring absorbance of cultures at 730 nm. Chlorophyll was extracted from whole cells with 100% methanol and the concentration was determined according to (14). The conditions for growing wild-type cells of Synechocystis sp. PCC 6803 in DiO have been described previously (10). Isolation ofThylakoid Membranes and PSI Particles Thylakoid membranes were prepared from ceils as described by (15). The thylakoid membranes were pelleted by centrifugation at 50,000 x g for 90 min and were resuspended in SMN buffer (0.4 M sucrose, lOmM MOPS, lOmM NaCI) for storage. Chlorophyll was extracted from thylakoid membranes and PSI trimers with 80% acetone and determined according to (14). For the isolation of PSI complexes, thylakoid membranes were incubated in SMN buffer with 20 mM CaCl, for 0.5 to 1.0 hr at room temperature in the dark to enhance trimerization of PSI. To the mixture, n-dodecyl-B-D-maltoside (DM) was added to a final concentration of 1.5% (w/v) and incubated in the dark on ice with occasional gentle mixing for 0.5 to 1.5 hrs. The non-solubilized material was removed by centrifugation at

10,000 X g for 15 minutes. The trimeric and monomeric PSI particles and PSII were separated by centrifugation in 10 to 30% (w/v) sucrose gradients with 0.04% DM in 10 mVI .MOPS. pH 7.0. Measurement of Photosystem I Activity Steady-state rates of electron transfer for isolated PSI complexes were measured using cytochrome as electron donor and flavodoxin as electron acceptor as described in (16). The measurement of P700 photooxidation in whole cells was based on the techniques developed by Harbison and adapted to cyanobacteria as described in (17). Analysis of Phylloquinone using HPLC and Mass Spectrometry PSI samples were prepared and quinones were analyzed on an equal chlorophyll basis, as previously described (10). However, the solvent gradient was modified to a gradient elution (solvent A = methanol; solvent B = isopropanol; 100% A from 0-10 min to 3% A/97% B at 30 min, hold until 40 min.) at 1.0 mL/min. A solution of PhQ (40 mM) was prepared in absolute ethanol and kept at -20° C as a standard for calibration. 105

Phylloqiiinone in vivo Incorporation Cells were grown in 31 flasks of BGl I with Fe and glucose (5mM) to a density of 0.550 to 0.675 at 730nm. For set AG, the cells were then sterile pelleted and resuspended in fresh media with Fe lacking glucose and grown for an additional day. For set 26. no additional preparation was performed other than a decreased amount of glucose (0.6 mM) was added to the media. At time minus 15 min. three antibiotics were added (132 (ig/ml chloramphenicol. 30 ^ig/ml kanomyocin. 75 ^ig/ml erythromycin) to half of the flasks. At time equal 0 min. lO^iM phylloquinone, final concentration, was added. At set time intervals, 400 ml liquid culture was pelleted and resuspended in SMN buffer, with EDTA (5mM) and PMSF (5fiM). Thylakoid membrane extractions were performed on a minipreparative scale. In a 2ml screw top vial, 0.250 ml glass beads were added to 1.6 ml of resuspended cells (400 O.D.). A small gap of air (-0.2 ml) was left for mixing. The vials were shaken eight times at 3400 rpm for 30 sec. on a mini-bead beater with a 8 vial adaptor (Biospec. Bartlesville OK). The remainder of the procedure is described in (10). PSl isolation was also done by the above described procedure. Isolated PSI timers were washed to remove sucrose and frozen. HPLC analysis of the pigments was done on an equal chlorophyll basis. Q-band EPR Spectroscopy of Photoaccumulated PSI Trimers Photoaccumulation experiments were carried out using a Bruker ER300E spectrometer outfitted with an ER 5106 QT-Wl resonator, which is equipped with a port for in-cavity illumination. Cryogenic temperatures were maintained with an ER4118CV liquid nitrogen cryostat. controlled with an ER412I temperature control unit. The microwave frequency was measured with a Hewlett-Packard 5352B frequency counter and the magnetic field was measured with a Bruker ER035M NMR Gaussmeter. The magnetic field wa.s calibrated using a.a' bisdiphenyiene-3-phenylallyl (BDPA) complexed 1:1 with benzene. Prior to the photoaccumulation, the sample was adjusted to pH 10 with glycine buffer to a final concentration of 233 mM, and sodium dithionite was added to a final concentration of 50 mM. After incubation for 20 min in the dark, the sample was placed into the resonator and the temperature was adjusted to 205 K. The sample was illuminated with a 20mW He- Ne laser at 630 nm for 40 minutes. The dark background was subtracted from the 106 photoaccumulated spectra. EPR spectral simulations were carried out on a Power Macintosh 7300/200 computer using a Windows 3.1 emulator (SoftWindows 3.0. Insignia Solutions. L'K) and SimFonia software (Broker Analytik GMBH). Optical Kinetic Spectroscopy in the nIR Region Optical absorbance changes in the near-IR were measured using a laboratory-built spectrophotometer (12). To assure resolution of kinetics in the microseconds time domain, a high-frequency roll-off amplifier described in the original specifications was not used, and the signal was fed directly into the plug-in (I1A33 differential comparator. 100 MHz bandwidth) of the Tektronix DSA601 oscilloscope. The sample cuvette contained the PSI trimers isolated from the menA or menB mutants at 50 |ig/ml Chi in 25 mM Tris-HCI, pH 8.3, 10 mM sodium ascorbate, 4 fiM DCPIP, and 0.04% n-dodecyl-6-D-maltoside (6-DM). Results Growth of Cells with Naphthoquinone Supplements Growth rates of the wild-type, menA', and menB26' mutants were measured in cells grown in BG-II medium which contains minerals including bicarbonate (Table 1). Under normal light conditions (40 (lE m"" s '), photoautotrophic doubling times for the wild type (26 h) were nearly three times faster than the mutants {menA' 86 h; and menB26', 77 h). The menA' mutant reached stationary growth at ~0.4 O.D, whereas the menB26' mutant and the wild type reached stationary phase at -0.8 and -1.0 A, respectively. Photomi.xotrophic growth that uses both respiration and photosynthesis for energy metabolism was achieved by adding 5 mM glucose. The wild type and menB26' mutant had similar doubling times of 14 h and 15 h, respectively, but the menA' strain grew slower with a doubling time of 26 h. The difference in the growth rates of the menB26' and menA' strains is consistent with the selection procedure used for isolating the former from the original menBlS' strain, which grows slower than the wild type strain under photomixotrophic growth conditions. Photoheterotrophic growth conditions included addition of glucose for respiration and of 10 |J.M DCMU to block PSII activity. The wild-type, menA' and menB26' mutant strains had similar doubling times under these conditions. Therefore, respiratory energy metabolism and cyclic photophosphorylation function normally in the mutants without phylloquinone. At 107 siationar>' phase in glucose-containing media, the wild-type and mutant cells attained a culture density between 2.75 and 3.5 A When the cells attained a culture density of -0.15 A, a variety of substituted naphthoquinones were added to the growth medium of the wild-type. menA' and menB26' cells to a final concentration of 5 to 10 (iM. Supplementing the medium of the wild-type cells produced a wide variety of responses in growth rates. Three naphthoquinone derivates, 5-OH-1.4-naphthoquinone, 2-methyl-5-OH-l,4-naphthoquinone, and 2.3-diCl-l.4- naphthoquinone, are known herbicides (18-20). Indeed, the wild-type cells died or ceased to grow upon addition of these naphthoquinones to the growth medium. Three naphthoquinone derivatives that are identical or closely related to the biosynthetic precursors of phylloquinone were found to slow the growth of the wild-type cells. The addition of 1.4- naphthoquinone, 2-C02-l,4-naphthoquinone. and 2-methyl-l,4-naphthoquinone (vitamin K,) to the growth medium led to an increase in the doubling times by 28%, 24% and 28%. respectively. Other 2-carboxy naphthoquinone derivatives such as 2-C02-3.5- naphthoquinone and l-COn-SJ- naphthoquinone cells led to an increase in the doubling times of wild-type cells by approximately 32%. Naphthoquinones that provided their own alkyl chain tail (menadione (Vitamin K,) and lacaphol (2-OH-3-prenyl-1.4- naphthoquinone)) also slowed the growth of the wild type. Wild-type doubling times were not affected by addition 2-OH-1,4-naphthoquinone to the growth medium. In general, naphthoquinone derivatives that are analogs or direct products of the naphthoate synthase enzyme appear to stress the wild-type cells and slow their growth, possibly by competing with cellular naphthoquinone for sites in the biosynthetic enzymes or in PSI. Addition of naphthoquinones that structurally resemble the product of naphthoate synthase (1.4-naphthoquinone. 2-OH-1.4-naphthoquinone. 2-C02-1.4-naphthoquinone and 2-methyl-l .4-naphthoquinone (vitamin K,)) accelerated the growth rates of the menB26' mutant cells (Table 1). The growth rate of menB26' increased by a factor of 3.7 and 1.8 when supplemented with 2-methyl-1,4-naphthoquinone (vitamin K,) and 1.4- naphthoquinone respectively. Indeed, the doubling time of the former approached that of the wild-type. Supplementing the direct product of naphthoate synthase, 2-C02-1.4- naphthoquinone improved the growth of menB26' by a factor of 2.1, but the doubling time 108

was still greater than the wild-type (Table 1). Vitamin K,, an analog phylloquinone compound with a C-40 unsaturated alkyl group, does decrease the doubling time by almost half. Another compound with a shorter alkyl tail, 2-OH-3-prenyl-1.4-naphthoquinone, also decreases the growth rate. Interestingly, the addition of naphthoquinones in which the head group does not structurally resemble 1,4-naphthoquinone have marginal effects on growth rates. Other 2- carboxy naphthoquinones C2-CO;-3.5-naphihoquinone and 2-C02-3,7-naphthoquinone) had virtually no effect on the growth rales of the menB26' mutant cells. Thus, the placement of the carbonyl oxygens on the naphthoquinone ring are as important as the presence of a carboxy group for affecting growth, which may be related to the ability of the quinone to become phytylated (Fig. 1). Those naphthoquinones that decreased the doubling times of the rnenA' and menB26' mutant cells were targeted for further study. In general, the men A' mutant was either unaffected or apparently stressed by the addition of naphthoquinone derivatives (Table 1). Since the menA' mutant lacks the ability to attach a phytyl tail to the naphthoquinone head group, these results would indicate that simple quinones either do not displace platoquinone-9 from the A, site, or alternatively cannot functional as an electron transfer cofactor if present in the A, site. Naphthoquinone derivatives that accelerated or did not affect growth rate in the menB26' mutant cells decreased the doubling times of the menA' mutant cells (Table 1). Only 2-OH- naphthoquinone allowed the cells to grow to normal stationar>' phase, although it did not affect the growth rate. Therefore, a 1,4-naphthoquinone may need to be phytylated before it can assemble and function in the A, site of PSI. Since the menA* strain was unable to use externally supplied quinones, all further work was performed using PSI complexes from the menB26' cells that had been grown with supplemented naphthoquinones. LC-MS Analysis of Pigment Extracts from PSI Complexes The presence of quinones in the extracts of purified PSI trimers was examined using HPLC coupled to a chemical ionization time-of-flight mass spectrometer. By co-injecting standards and by interpreting the spectra at selected masses, chlorophyll a (m/z 892) was identified with a retention time of 19.0 minutes. Phylloquinone (m/z 450) and PQ (m/z 748) had characteristics retention times of 19.9 and 29.1 min, respectively (Fig. 2). Virtually 109 identical chromatograms were obtained for pigment extracts of the wild-type. metiA' and menB26' mutants, except that the phylloquinone peak at 19.9 min was missing in the mutants (Table 2). Quantification by MS is somewhat difficult, given that each molecule has varying ionization characteristics. It is possible to determine the number of counts at a particular molecular weight, thereby calculating the area under the peak. However, without detailed controls and standards this MS data should be taken more as a trend and semi­ quantitative. More detailed quantitative data has been calculated from the UV/vis spectra, given that the extinction coefficients are well known for the observed molecules. The amount of PQ in the wild-type trimers appears to be very small ( < 1%). The low amount of PQ found in wild-type PSI has not been detected to be involved in electron transfer and is considered to carryover from the thylakoid membrane lipids. PSI complexes isolated from the nienB26' mutant cells supplemented with several naphthoquinone derivatives led to the restoration of the phylloquinone peak. The addition of authentic phylloquinone to nienB26' mutant cells revealed that the cells could harvest the highly nonpolar quinone from the liquid media and incorporate it into the A, site at a ratio (> 100:1 PhQrPQ) similar to the wild-type. Cells are also capable of harvesting menaquinone (m/z 444). the C-40 partially unsaturated tail group, an analog of phylloquinone. There is nearly complete displacement of plastoquinone as determined by MS. The menaquinone elutes earlier (13.3 min) than phylloquinone (20.5 min) in this solvent system. We also monitored a minute («1) but detectable amount of phylloquinone (m/z 450) in the pigment extracts that was not found in the neat menaquinone. Supplementing the menB26' mutant cells with the direct product of naphthoate synthase, 2-CO,-I,4-naphthoquinone. resulted in the presence of phylloquinone in the PSI complexes at a ratio greater than 100:1 PhQ:PQ. When the medium was supplemented with 2-methyl-1,4-naphthoquinone (vitamin Kj), phylloquinone was found in the PSI complexes but at a ratio (~ 30:1 PhQ:PQ) that was lower than the wild-type. Thus vitamin K, is probably not incorporated in the phylloquinone biosynthetic pathway as efficiently as is 2-C02-l,4-naphihoquinone. .A.nother "tailed' naphthoquinone used was 2-OH-3-prenyl-1.4-naphthoquinone (lacaphol). The molecule has not been identified in the pigment extracts in the present solvent system. Further investigation identifying the molecule is in progress. 110

When menB26' cells were grown with 1,4-naphthoquinone. the LC-MS analysis of pigments extracted from isolated PSI complexes showed the presence of phylloquinone (m/z 450). These samples also contained a species with a mass of 436 m/z that probably represents unmethylated 2-phytyl-1,4-naphthoquinone. The NQ-p/PQ ratio (6:1) was also lower than when the cells were grown with vitamin K, or I-CO,-1,4-naphthoquinone. The ratio of unmethylated 2-phytyI-naphthoquinone to the methylated phylloquinone (or derivative) is 2 times when the menBlt'- mutant was supplemented with 1.4- naphthoquinone. This suggests that the rate of methylation is slower than the rate of phytylation. and that the pool of unmethylated derivative is available for assimilation into the A, site. The presence of demethyl phylloquinone in the PSI complex also shows that the methyl group is not required for the assembly into the A, site. Phylloquinone was also found in significant amounts in the PSI complexes in the supplementation experiments with 1.4- naphthoquinone. The total phylloquinone derivative (methylated and unmethylated species) to plastoquinone ratios was 6:1 for 1,4-naphthoquinone. Although 2-OH-naphthoquinone supports an increase in the growth rate of the menB26' mutant, no peaks at m/z 450, corresponding to phylloquinone or m/z 436. corresponding to demethyl phylloquinone. were found. Rather, a new peak at 11 minutes corresponding to a mass of 453 was observed. We have tentatively assigned this peak to a 3- phytyl-1,4-naphthoquinone containing a 2-hydoxyl instead of 2-methyl group. The ratio of 2-hydroxy-3-phytyl-1.4-naphthoquinone to plastoquinone in PSI complexes is approximately 0.10:1. Therefore, plastoquinone is not replaced by the 2-OH derivative of phylloquinone to any significant degree in the mutant PSI complexes. Since the phytyl transferase has already been shown to function when a hydrogen or a carboxy group is present at ring position 2, it is not surprising that it also functions when a hydroxy group is present. However, the low amount of 2-hydroxy-3-phytyl-I.4-naphthoquinone in the PSI complexes may indicate that the specificity of the phytyl transferase for the 2-hydroxy-I.4- naphthoquinone substrate is low. HPLC Analysis of Pigment Extracts To confirm the results of the LC-MS analysis of the pigment extracts from PSI complexes, we examined the absorption spectra of the HPLC peaks with a UV-visible range Ill aiode array photospectrometer detector. In short, the distinct absorption spectra of pigments in the chromatograms complement the LC-MS data. However, quantitative data can be determined from the UV-visible data, since the extinction coefficients of the molecules are known. The retention limes in the HPLC analysis were consistent and close to those in the LC-MS studies, with slight variations that are attributed to different lengths of the connecting tubes between instruments. Coelution with other pigments, namely chlorophyll a at 18 minutes and (3-carotene at 29 minutes, allowed comparison of the wild-type chromatograms to the phylloquinone-less mutants. Phylloquinone, its derivatives and plastoquinone spectra are shown in Figure 3, and are consistent with the spectra of known standards. Elution and absorption properties of phylloquinone and plastoquinone have been described previously (10). The phylloquinone without methyl group has been identified to elute at 17.7 min. The spectrum of the vitamin K- supplement is identical to phylloquinone with absorptions at 248 nm and shoulder at 270 nm. Plastoquinone is only observed in the PSI trimers of phylloquinone-less mutants and some menB26- cells that were grown with naphthoquinone supplements. Plastoquinone could not be detected in any appreciable amount in the wild-type PSI timers (< 1-2%). In Figure 3, the plastoquinone inset is representative of the plastoquinone spectrum of a singlet at 270nm. Phylloquinone and derivatives were identified by a combination of Uv/Visible spectroscopy and mass spectrometry. When available, known standard solutions were used to confirm identifications. Demethylphylloquinone was identified by its mass and similarity to the phylloquinone spectrum. It also appeared before phylloquinone in the chromatogram.. The lack of a "greasy" methyl group is attributed to the decreased elution time of 2.3 minutes. The 2-OH-phylloquinone is only identified by its mass, thus tentative in assignment. The hydroxyl group significantly reduced the elution time to 11.0 minutes. The phylloquinone to plastoquinone ratio of the pigment extracts of PSI trimcr complexes varies significantly (Table 3). The wild-type shows a PhQ:PQ ratio of approximately 70:1, which corresponds to less than a 2% amount of PQ in the wild type trimer preparations. Several preparations have detected no plastoquinone in the trimers with other preparations have a plastoquinone content as high as 2% that of phylloquinone. This 112 suggests that any detected plastoquinone is carried from the lipid membranes over in the preparation of the trimer complexes. Plastoquinone has not been detected to be functioning within the A, site of the wild type by any of the experiments presented in this paper. Addition of phylloquinone to the growth media shows that the cells are capable of harvesting and incorporating phylloquinone in PSl in the menBlt' mutant strain. The calculated ratio is approximately 60:1 phylloquinone:plastoquinone. As described in the next experiment the amount of phylloquinone within the PSI complexes does change over time. Supplementing whole cells with menaquinone (vitamin K2) also shows incorporation into the PSI complexes. However, the ratio is not as high (8 MQ:PQ) as phylloquinone incorporation. The difference in displacement is attributed to the menaquinone alkyl group being a C-40 unsaturated "tail" instead of phylloquinone's C-20 phytyl group. Vitamin K, and 1,4-naphthoquinone is also incorporated into PSI after phytylation. The ratios of PhQ:PQ are lower than for directly adding phylloquinone or 2-CO,-1.4-naphthoquinone. This is thought to result from vitamin K, and 1.4-naphthoquinone being an analogous in structure to 2-CO;-l.4-naphthoquinone. Both the molecules lack the carboxy group, which may be important in the phytyl transferase mechanism or recognition. Considering that for most of the supplemented quinones the Chl/Qj ratio is close to the published norm (21). the A, site is not empty nor contains a non-phytylated quinone. Deviations arise in 1,4-naphthoquinone where is appears that the quinone content has dropped causing an increase in Chl:Qy ratio. There is some minor indication by pulsed EPR that a free 1.4-napthoquinone may be occupied within the A, site, which is undergoing further investigation (Stehlik D., personal communication). P^OO' optical recombination kinetics To determine if the quinone is within the Ai site, we probed its activity by observ ing the kinetic backreaction charge recombination of P700*. In PSI trimers isolated from the wild type, the reduction of the P700* RC is multiphasic after a saturating flash (12.22). When measured by 810 nm near-IR optical spectroscopy and in the absence of external electron acceptors, the typical wild type spectrum shows a majority of the P700' being reduced with a lifetime of 60 - 80 ms and an occasional minor faster phase (~10 ms) not seen here (Figure 4). The biphasic decay is considered to be a result of the PSI comple.xes 113 being in different confirmations (12.22). The microsecond decay is attributed to earlier damaged reaction centers. There is also a long-lived (>1200 ms) kinetic phase of P700' reduction from direct reduction of the P700^ by reduced DCPIP that contributes to 5 - 20 % of the total absorbance change. In PSI trimers isolated from the menB16' mutant, the reduction of P700^ is also multiphasic (Figure 4). When menBlS' is measured in the absence of external electron acceptor. P700* is reduced with lifetimes of approximately 3.0 ms and 19 ms in a 7:1 ratio. The 3 ms phase has been attributed to piastoquinone in the Ai site with the electrons originating on the iron sulfur clusters (12). The mutants also exhibit a minor long-lived kinetic phase attributed to electron donation of DCPIP to the oxidized P700~ that contribute the remainder of the absorbance change. The kinetic back reaction phases for both the menB26' mutant are consistent with piastoquinone being in the At site of PSI and acting as an electron intermediate from Ao to Fx (12). Monitoring the kinetic back reaction of the quinone supplemented cells indicates whether piastoquinone still remains active in the A, site or is replaced by a naphthoquinone (Table 4). The 1,4-naphthoquinone head group in the A, site causes a loss or decrease in contribution from the 3 ms phase. Phylloquinone reverts the ntenB26' PSI complexes back to wild-type spectra. Menaquinone also causes a loss of the 3 ms phase but does not restore wild type spectra. The 60 ms phylloquinone phase is replaced by a 30 ms phase. The difference may result in lower menaquinone to piastoquinone ratio yielding a mixture of two unresolved phases. It may also be attributed to a slightly different orientation of the quinone head group in the pocket given the significantly different alkyl 'tail'. Addition of the naphthoate synthase product DHNA, shows a complete loss of the 3 ms piastoquinone phase. The PSI complexes appear to be fully functional phylloquinone containing wild type. Both vitamin K, and l.4-naphthoquinone shows a decrease, but not loss, in the piastoquinone contribution, indicating that the phytylated quinone is active in at least a portion of the A, sites with piastoquinone occupying the remaining. Other quinones which have not been located in the PSI pigment extracts also show only piastoquinone in PSI. For 2-C02-3,5-naphthoquinone, 2-C02-3,5-naphthoquinone. and 2-OH-3-prenyl-1.4-naphthoquinone the 3 ms phase is dominate. 2-OH-1.4-naphthoquinone also shows a predominate piastoquinone peak. However, the lower contribution of 114 plastoquinone with the 2-OH-3-prenyl-I,4-naphthoquinone supplemented cells suggests that there may be partial replacement occurring. The single unit prenyl tail may have contributed to displacement some of the plastoquinone in the A, sites. This technique confirms that not only are certain quinones contained in PSI. but also active within the A, site as well. This data coupled with the pigment extract analysis suggests that only the phytylated quinones can displace plastoquinone. The non-phytylated quinones are not detected in the neither the HPLC or the kinetic back reaction, as indicated by the 3 ms plastoquinone phase. CW EPR Spectroscopy at Q-Band of Photoaccumulated A-, To further probe the contents of the A, site, we photoaccumulated the A, signal for Q-band EPR. Plastoquinone has been previously identified in the A, site of the menB mutants (12). Figure 5 shows the Q-band EPR spectra of photoaccumulated A," in PSI complexes isolated from menB26' and menB16' with naphthoquinone supplements. For comparison the PQ containing menB26' mutant strain and wild type is shown. Addition of 2-C02-L4-naphthoquinone. vitamin K,, and phylloquinone supplements to the mutant reveals PhQ" is exclusively photoaccumulated. 1.4-naphthoquinone shows a predominant phylloquinone spectrum, with possible contributions from plastoquinone and demethylphylloquinone. The menaquinone spectrum reveals a 1,4-naphthoquinone head group. EPR spectra showed only PQ' being photoaccumulated of quinone supplements not been identified within PSI by HPLC or kinetic back reaction. Phylloquinone in vivo Incorporation Experiment We have monitored the rate of incorporation of phylloquinone in the Ai site o\ er time, to determine whether there was a preferential incorporation of PhQ as a function of PSI assembly or mass diffusion. The phylloquinone-Iess menB26' strain was grown to mid- log phase, at which point 10 (J,M phylloquinone was added to the cultures and antibiotics added to half the cultures. We then obser\'ed the incorporation of phylloquinone in the PSI complexes for normal growing cells and cells in which protein synthesis has been inhibited. This will provide a clue as to the stability of phylloquinone in the Ai site and whether or not phylloquinone diffuses in or is incorporated during PSI assembly. As seen in Figure 6. phylloquinone has completely displaced plastoquinone from PSI in both sets of cultures 115

wiihin the first measurement of 30 minutes. Over the course of three days, the level of phylloquinone drops as it is rereplaced by plastoquinone. Both traces, with and without protein inhibiting antibiotics, show a similar trend. Given that the antibiotic containing cultures cannot synthesize new PSI complexes as a process to incorporate phylloquinone. phylloquinone is diffusing in and expelling plastoquinone. The initial incorporation of phylloquinone is very rapid suggesting that the difference in binding stabilities for phylloquinone and plastoquinone within the Ai site are large. However, over time phylloquinone is being replaced by the now native plastoquinone. This probably occurs as a combined process of mass action and catabolism of phylloquinone. Chlorophyll content on a per cell basis does not change significantly, suggesting that the cells are not undergoing substantial changes during the experiment. Discussion To understand function of phylloquinone in photosynthetic organisms, we generated cyanobacierial mutant strains that are unable to synthesize phylloquinone (10). The PSI complexes of these mutants contain plastoquinone in their A, site (11). The structure of phylloquinone contains several distinct features that affect their function. The phylloquinone carbonyl oxygens have been found to hydrogen bond to amino acids within the A, pocket (23), while the whole ring n stacks to either a tryptophan or histidine (24). For a naphthoquinone to properly function within the A, site an alkyl tail must be present in the 3 position on the ring (25). Most significantly, the redox potential of phylloquinone is significantly different in the PSI environment as compared to being free in solution ((12), and references therein). To further probe the structure-function relations of quinones in the A, site of PSI, we added different naphthoquinone supplements during growth of phylloquinone-less mutants. Once inside the cell, the supplemented quinone could undergo the remaining natural steps in phylloquinone synthesis: attachment of the phytyl tail by phytyl transferase (the menA product) and possible methylation by methylase (the product of menG). The newly synthesized phylloquinone or derivative competed with and replaced the plastoqumone in the A, site of the PSI complexes in the mutants. Much effon has been spent in extracting the phylloquinone from the A, site to insert alternative quinones in vitro (26,27,28,29). These benzo- naphtha- and anthraquinones have 116

been shown to bind in A, and to accept electrons from AQ (26,30). But only 1.4- naphthoquinones with a hydrocarbon tail can transfer the electron forward to the [4Fe-4S] cluster Fx (31). From this summary two hypotheses were proposed: the electron transfer is dependent on an appropriate redox potential or that a specific orientation of the quinone is necessary, which is provided by the isoprene tail. Initially, we studied the growth characteristics of the menB26' mutant strain and wild type with naphthoquinone supplements (Table 1). Naphthoquinone supplements that were found phytylated and in PSI decreased the doubling times of the cells. The remaining unutilized naphthoquinones did not significantly change the growth rales. Wild type cell growth rate, however, slowed upon addition of nearly ail supplemented quinones. We suggest that there is oxidative stress or competition with analog compounds caused by the exogenous naphthoquinone. Quinones with a 1,4-naphthqouinone moiety are known to be acetolactate synthase inhibitors (19). Acetolactate is used in the synthesis of hydrophobic amino acids and several metabolism pathways (32). Note that for the menB26' mutant strain the addition of the utilized naphthoquinones does not decrease the doubling time to that of the wild type, but only to the averaged naphthoquinone fed wild type doubling time of 36 ± 7.7 h. The naphthoquinone compounds used in the present studies can be classified into five categories. Firsi. phylloquinone and menaquinone can be added directly to the growth media and restore phylloquinone or menaquinone in the Ai site, respectively. The cells are apparently capable extracting the two quinones from the media, draw it through the cell membrane and insert it into PSI. Second, compounds those are analogous to the substrate of naphthoate synthase produce phylloquinone. which was incorporated into the A| site to the exclusion of plastoquinone. Vitamin K3 and 2-C02-l,4-naphthoquinone are converted into phylloquinone and less than 18% and 6% plastoquinone is left occupying the Ai site, respectively. Third, some naphthoquinones produced phylloquinone derivatives, which competed with plastoquinone for the A| site. The partial replacement could result from decreased availability due to limiting rates of synthesis or less binding affinity than phylloquinone to the A| site. 1.4-naphthoquinone produced demethylphylloquinone in addition to phylloquinone. In the fourth set of compounds, the phytylated naphthoquinone 117 deri\ative was not detected in the PSI complexes. These compounds. 2-CO:-3.5- naphthoquinone and 2-C02-3,7-naphthoquinone, do not affect the growth rates of the cells, are not identified in the Ai site, or affect the activity of PSI. These compounds may not be imported into the cell or are unable to get phytylated. thereby hindering their assembly into PSI. Consequently, plastoquinone remains active in the Ai site. Alternatively, the phytviated quinones have lower binding affinity for the Ai site than plastoquinone. This is unlikely, given that the phytylated quinones would have been identified in thylakoid membrane by LC-.MS. Finally, some naphthoquinones were toxic to the cells. 5-OH-1.4-naphthoquinone. 2-Me-5-OH-1.4-naphthoquinone, and 2,3-Cl-l,4-naphthoquinone are known phytotoxic herbicides (18.33). Addition of naphthoquinones that resemble the product of naphthoate synthase, exclusively produce phylloquinone, albeit not necessarily to the wild type levels as quantified by HPLC-UVA/^isible. Adding 2-C02-I,4-naphthoquinone, and vitamin K, to the menB26' mutant generate different levels of phylloquinone found in PSI. Wild type and phylloquinone supplemented mutant cells have a PhQ:PQ ratio of 70:1 and 59:1. respectively. However, the addition of 2-CO2-1,4-naphthoquinone has a lower ratio (16:1 PhQ:PQ). This difference may arise from the difficulty of a relatively water soluble quinone passing through lipid layers or arriving at the location where the next enzymatic step occurs, phytyl transferase. Vitamin k,, which lacks the 2-carboxy group has a ratio of 7:1 PhQ:PQ. The decreased ratio may be accounted for by the lack of the carboxyl group in the 2 posiiion. which is lost as carbon dioxide during the phytyl tail attachment. The methyl group on vitamin K, suggests that the phytyl transferase can apparently function without the high energy carboxyl group. However, that vitamin K, produces phylloquinone does reveal methylation is not necessarily the final step in phylloquinone synthesis. Naphthoquinones that are more dissimilar to the direct product of naphthoate synthase produce phylloquinone, demethylphylloquinone, and derivatives. Addition of 1,4- naphthoquinone to the growth media, produced phylloquinone and demethylphylloquinone (Table 3). The lower ratio of PhQ^iPQ may be accounted by I.4-naphthoquinone has to be altered by two enzymatic steps instead of just one, as with vitamin Kv The loss of the methyl group may also signal lesser affinities of the naphthoquinone head group for the A, 118

site as seen by Iwaki and Iioh with I.4-naphthoquinone and Vitamin K-, (2-Me-I,4- naphthoquinone) (34). Demethylphylloquinone was in greater proportion than phylloquinone since it has one less step. We believe that the methylation product phylloquinone is not necessary for function in PSI. This has been also confirmed by genetically disabling the menG gene (methyl transferase) and finding that demethylphylloquinone is operational in the A, site (35). Menaquinone is also incorporated successfully within the A, site, albeit at levels lower than phylloquinone. The difference between the two molecules resides in the alkyl side chain. The cost of the C-40 unsaturated prenyl group is a ten fold less ratio of NQ/PQ. However, it does show that a C-40 group, like the C-45 PQ 'tail', was also capable of fitting the quinone head group into the A, site. The binding pocket that houses the alkyl side chain of the quinones must be fairly accommodating and offers a new line of research: that of the function of the quinone alkyl binding site. The activity PSI. as monitored by the flash induced P700 back reaction and by photoaccumulated EPR mirrors the content of the Ai site. Compounds that contained appreciable amounts of phylloquinone or their derivatives show a lack of the characteristic 3 ms back reaction attributed to plastoquinone (Table 4). Vitamin K3 and 1.4-naphthoquinone supplements, which contained up to 20% plastoquinone. show the 3 ms phase, but at reduced contribution from the normal 70% for menB26'. Compounds not incorporated into PSI show the normal contribution by plastoquinone. This is a quick diagnostic determination of plastoquinone content. It also suggests that 2-OH-3-prenyI-l,4-naphthoquinone may be occupying some of the Ai sites, given that the 3 ms phase is present by at about half the normal contribution. EPR is less quantitative than P700 measurement, yet still show reduced naphthoquinone signals. The pulse chase experiment proved to be very valuable in the strategy for the feeding experiments. Phylloquinone was inserted into PSI very rapidly demonstrating that phylloquinone was capable of diffusing into an already constructed PSI complex and displacing plastoquinone. Initially, we did not expect a loss of phylloquinone over time. However the loss was justified by catabolism of phylloquinone and mass action by plastoquinone to kinetically reoccupy the site. 119

As reported earlier, plastoquinone occupies the A, site of PSI in the phylloquinone- less mutants for menB' and menA of Synechocystis sp. PCC 6803 (10.11). We show that the addition of particular naphthoquinones the cells complete the synthesis by attaching a phytyl tail producing phylloquinone and derivatives. These phytylated compounds compete with plastoquinone during PSI assembly and also are capable of replacing the resident plastoquinone in the existing PSI complexes. We used HPLC and LC-MS to detect the presence of phylloquinone analogs in PSI complexes. Inside the PSI complex, the phytylated naphthoquinone derivatives could function as an intermediate in the electron transfer from A(, to Fx. Optical and EPR spectroscopy was used to study activity of the phylloquinone analogs in the PSI complexes. When the foreign quinones replaced plastoquinone partially, proportion of kinetic phases in the P700 reduction kinetics was used to indicate relative contributions of the resident plastoquinone and foreign quinone. References 1. Nugent. J. H. (1996) Eur. J. Biochem. 237(3), 519-531 2. Golbeck. J. H. (1993) Proc. Natl. Acad. Sci. USA 90(5), 1642-6 3. Nitschke. W.. and Rutherford, A. (1991) Trends Biochem. Sci. 16, 241-245 4. Trumpower, B. L. (1982) Function of Quinones in Energy conserving Systems. Trans. 0-12-701280-x (Trumpower, B. L.. Ed.). Academic Press, New York 5. Iwaki. M.. and Itoh. S. (1994) Plant Cell Physiol. 35, 983-993 6. Kumazaki, S., Iwaki. M., Ikegami, I.. Kandori. H., Yoshihara. K., and Itoh. S. 0994) J. Phys. Chem. 98, 11220-11225 7. Itoh. S.. and Iwaki, M. (1991) Biochemistry 30(22), 5340-6 8. Biggins. J. (1990) Biochemistry 29(31), 7259-64 9. Iwaki, M.. and Itoh. S. (1989) FEBS Lett. 256, 11-16 10. Johnson, T. W.. Shen. G., Zybailov, B., Kolling, D., Reategui, R.. Beauparlant. S.. Vassiliev. I. R., Bryant, D. A., Jones, A. D., Golbeck. J. H., and Chitnis. P. R. (2000) J Biol Chem 275(12), 8523-30 11. Zybailov, B.. van der Est. A.. Zech, S. G., Teutloff. C.. Johnson. T. W.. Shen. G.. BittI, R.. Stehlik. D., Chitnis, P. R., and Golbeck, J. H. (2000) J Biol Chem 275(12), 8531-9 120

12. Semenov. A. Y., Vassiliev. I. R.. van Der Est. A., Mamedov. M. D., Zybailov. B.. Shen, G., Stehiik. D.. Diner, B. A.. Chitnis. P. R., and Golbeck, J. H. (2000) J Biol Chem 275(31), 23429-23438 13. Rippka, R.. Deruelles, J.. Waterbur>', J. B.. Herdman, M.. and Stanier. R. Y. (1979) J. Gen. Microbiol. Ill, 1-61 14. Amon. D. (1949) Plant Physiol. 24, 1-14 15. Sun. J.. Xu, Q., Chitnis. V. P.. Jin, P.. and Chitnis, P. R. (1997) J. Biol. Chem. 272(35), 21793-21802 16. Yang. F.. Shen. G., Schluchter. W. M.. Zybailov. B. L., Ganago. A. O., Vassiliev. I. R.. Bryant. D. A., and Golbeck. J. H. (1998) J. Phys. Chem. B 102(42), 8288-8299 17. Redding, K.. Coumac, L., Vassiliev. I. R.. Golbeck, J. H., Peltier. G.. and Rochaix. J. D. (1999) J Biol Chem 274(15), 10466-73 18. Oettmeier, W., Dostatni, R., and Santel, H. J. (1987) in Z. Naturforsch.. C: Biosci. Vol. 42, pp. 693-7 19. Haraguchi. H.. Yamano, K., Kusunoki. N., and Fukuda. A. (1997) in J. Agric. Food Chem. Vol. 45, pp. 2784-2787 20. Alegria. A. E.. Ferrer, A., Santiago, G.. Sepulveda, E., and Flores. W. (1999) in J. Photochem. PhotobioL. A Vol. 127, pp. 57-65 21. Schubert. W. D., Klukas, O., Krauss. N., Saenger, W., Fromme. P., and Witt. H. T. (1997) y Mo/ Biol 272(5), 741-69 22. Vassiliev, I. R.. Jung, Y. S., Mamedov. M. D., Semenov. A., and Golbeck. J. H. (1997) Biophys J 72( 1), 301-15 23. Klughammer, C., Klughammer. B., and Pace. R. (1999) Biochemistry 38(12), 3726- 32 24. Hanley. J., Deligiannakis. Y., MacMillan, F.. Bottin. H.. and Rutherford. A. W. (1997) Biochemistry 36(39), 11543-9 25. Biggins, J. (1990) in Biochemistry Vol. 29. pp. 7259-64 26. Itoh, S., Iwqaki, M.. and Ikegami. I. (1987) Biochim. Biophys. Acta 893, 508-516 27. Itoh. S.. and Iwaki. M. (1989) FEBS LETT. 243, 47-52 121

28. Rustandi, R. R.. Snyder, S. W., Feezei, L. L.. Michalski, T. J., Norris, J. R.. Thumauer, M. C.. and Biggins. J. (1990) Biochemistry 29(35), 8030-2 29. Setif. P.. Ikegami. I., and Biggins. J. (1987) Biochim. Biophys. Acta 894(2), 146-56 30. Biggins. J., and Mathis. P. (1988) in Biochemistry Vol. 27. pp. 1494-500 31. Rustandi. R. R., Snyder, S. W., Biggins, J., Norris, J. R., and Thumauer. M. C. (1992) in Biochim. Biophys. Acta Vol. 1101. pp. 311-20 32. (ww^v.genome.ad.jp website) 33. Dombos, D. L., Jr.. and Spencer, G. F. (1990) in J. Chem. Ecol. Vol. 16, pp. 339-52 34. Iwaki. M.. and Itoh. S. (1991) in Biochemistry Vol. 30. pp. 5347-52 35. Shen. G.. and Bryant. D. (2000) 122

Table 1

Growth of wild-type and mutant strains of Synechocysiis sp. PCC 6803 with supplemented naphthoquinones

Doubling Time (h) Wild type menA menB26' None 26 ± 3.0 86 ± 5.0 77 ± 4.0 Glucose 14 ±2.5 26 ± 8.7 15 ±3.4 DCMU + Glu 17 ±2.0 19 ±4.3 19 ±2.3 2-Me-1.4-NQ (Vit K,) 36 ±4.0 105* 21 ± 1.0

menaquinone (Vit K,) 49 ± 10 — 46 ± 4.9 1.4-NQ 34 ± 4.2 58* 41 ±3.4 5-OH-NQ died died died 2-Me-5-OH-l,4-NQ died died died 2-OH-NQ 26 ±3.8 83 32 ± 0.9

2-OH-3-prenyl-1,4-NQ 46 ± 4.9 — 61 ±4.0 2.3-Cl-NQ 48* died 115* 2-CO,-1.4-NQ 38 ±3.7 61 ±5.5* 36 ± 1.0 2-CO,-3,5-NQ 36 ±5.1 60* 70 ± 13 2-CO,-3,7-NQ 38 ±5.0 60* 57 ±6

Light intensity was 50 (lE, normal light. Error is in the fit linear slope and was between 5-15%. with an n = 3 or more. Concentrations of naphthoquinones, DCMU. and glucose were 5 [iM, 10 |iM. and 5 mM. respectively. Initial cell density was 0.10 at 730 nm. The "none" column is basal growth rate in BGII medium. * Indicates that the cells doubled once and then died. N.D. = Not done. Numbers without a statistical error were derived from less than three observations. Table 2

Quinone analysis of the PSI pigment extracts by LC-MS of the wild-type and menB26' mutant with naphthoquinone supplements

Sample Mass Ratios DeMePhQ/ PhQ/PQ NQT/PQ PO wild-type nd > 100 > 100 menB26' nd nd nd menB26' -r PhQ nd > 100 > 100 menB26- + 2-C02-1.4-NQ nd > 100 > 100 menBlS' + K3 nd 29 ± 1.3 29 ± 2.3 menB26' + 1.4-NQ 4 ±2.7 2± 1.3 6 ±3.2 menBia -r K2 nd > 100 (K2 only) >100

Table: Chromatograms were run on 0.026 mg chlorophyll content of the PSI trimers. Sample injections (35 jiL) were made on a 4.6 mm x 25 cm Ultrasphere Cis column (4.6 mm x 250 mm) with 5 |im packing (Beckman) using gradient elution (solvent A = methanol; solvent B = isopropanol; 100% A from 0-10 min to 3% A/97% B at 30 min. hold until 40 min.) at 1.0 mL/min. Mass 436. 450, and 748 represents PhQ minus a methyl group (or isomer). PhQ. and PQ n = 9. respectively. Mass ratios are based on the intensities of the selected peaks. The peak areas used for the ratios are the m/z ± 1.0 mass units. For a given standard deviation there are 4 to 6 independent results (n = 4 to 6). nd = not detected. 124

Table 3

Quinone analysis of the PSI pigment extracts by HPLC-UV/vis of the wiid-type and menB26' mutant with naphthoquinone supplements

Sample Chl/Qx PhQ/PQ

wild type 42 ±3 71 ±40 menBia 55 ±2 — menBia -f- PhQ 44 ±4 59 ±30 menB26' + 2-C02-1.4-NQ 45 ±3 15.9 ±6.7 menB26' -t- K3 50 ±5 5.4 ±3.5 menB26' + 1.4-NQ 250 (4.1) menB26' -t- K2 60 ±4 8± 1.8

Table: Chromatograms were run on 0.026 mg chlorophyll content of the PSI trimers. Sample injections (35 |iL) were made on a 4.6 mm x 25 cm Ultrasphere Cis column (4.6 mm x 250 mm) with 5 |j.m packing (Beckman) using gradient elution (solvent A = methanol; solvent B = isopropanol; 100% A from 0-10 min to 3% A/97% B at 30 min. hold until 40 min.) at 1.0 mL/min. PhQ-CHs. PhQ, and PQ n = 9, represents PhQ minus a methyl group. PhQ. and plastoquinone n = 9. respectively. Differences in retention times relative to LC-MS are attributed to slightly different lengths of the connecting tubes. Retention times for the components are PhQ-OH = 11.0 min, demethylPhQ = 17.7 min, PhQ = 19.9 min. PQ = 29.1 min. For a given standard deviation there are 4 to 6 independent results (n = 4 to 6). 125

Table 4

Flash induced P700 Back Reaction at measured at 810 nm on PSI trimers

Sample ms back % of A, content reaction contribution wild t>'pe 62 40 PhQ wild t\'pe anaerobic 7.2 4 PhQ tnenB26' 2.8 ± 1.5 65± 18 PQ menB26' + PhQ 10 ±4.5 20 ±5 PhQ menB26' K2 32 ± 5.4 20 ±2 MQ menBie- + 2-C02-K4-NQ 31 ± 14 9 ± 2 PhQ menB26' + K3 2.7 ± 0.9 23 ± 11 PhQ menB26' + 1.4-NQ 5.3 ± 0.47 28 ±3 Mix menB26' + 2-C02-3.5-NQ 3.3 78 PQ menB26' + 2-C02-3,7-NQ 3.4 76 PQ menB2e- + 2-OH-NQ 4.2 56 PQ menB!2(> + 2-OH-3-prenyI-1.4-NQ 2.4 31 PQ/Mix?

TABLE. P700'' reduction kinetics in PS I complexes isolated from the menB mutant. Flash induced optical transient measured at 811 nm during a single flash. Sample conditions: 4- sided fluorescence cuvette containing sample at 10 (ig/ml in Chi 25 mM Tris. pH 8.3. 0.04% B-DM. 10 mM ascorbate and 4 (iM DPIP. Excitation wavelength. 532 nm. excitation energy. 1.4 mJ. 126

Figure Legends Fig. 1. Naphthoquinone supplements used in this study. NQ and CO2 are naphthoquinone and carboxy, respectively. Fig. 2. LC-MS analysis of the pigments extracted from PSI complexes of the ntenB26' cells that had been grown with 1,4-naphthoquinone supplement. From top to bottom the chromatograms represent the total ion count at m/z 436, m/z 450, and mJz 748 which correspond to the mass of unmethylated PhQ (t = 17.6 min), PhQ (t = 19.9 min). and plastoquinone n = 9 (t = 29.1 min), respectively. Fig. 3. HPLC chromatogram of the pigment extracts of the menB26' cells that had been grown with 1,4-naphthoquinone supplement. The peaks were detected with a UV-visible photodiode detector. Labeled inset spectra at 17.7 min. 20.1 min, 29.1 min represent PhQ minus methyl group (PhQ-Me). phylloquinone (PhQ). and plastoquinone (PQ), respectively. The chlorophyll and 3-carotene peaks are at 19 min and 28.9 min. respectively. FIG. 4. P700^ Reduction Kinetics in PSI Complexes Isolated from the menB26' mutant. Laser-flash induced optical transient measured at 811 nm. Time is plotted on a logarithmic scale, in which a deviation from the horizontal represents a kinetic phase. The experimental data is shown as dots and the computer-generated exponential fits are shown as solid lines, with the lifetimes of each phase shown. The relative contributions of each kinetic phase can be judged by the intersection of the fit line with the abscissa. The error between the experimental data and the fits are shown above each plot. The samples A-F are wild- type, menB26\ menB26' + 2-Me-1,4-naphthoquinone, menB26' + 2-CO3-I.4- naphthoquinone. menB26' + 2-CO,-3,5-naphthoquinone, and menB26' + menaquinone (Vit. K,). respectively. FIG. 5. Photoaccumulated Q-band CW EPR spectra of A|' and Q in PSI complexes isolated from wild type, ntenB" mutant, and melts' mutant with naphthoquinone supplements. The photoaccumulation was carried out for 40 min at 205 K. Instrument settings: microwave power 1 mW, microwave frequency 34.056 GHz. modulation frequency 100 KHz, modulation amplitude 1 G. temperature 205 K. time constant 10 mS. conversion time 10 mS. 100 averaged scans. 127

Fig. 6. Pulse E.xperiment to monitor in vivo incorporation of externally supplied phyiloquinone into PSI. Percentage of PhQ to total quinone (PhQ + PQ) extracted from PSI trimer pigments. At time equal zero 10 fiM final concentration PhQ was added to the liquid media. Aliquots of cells were removed at intervals, pelleted and frozen. Frozen cells were later broken and PSI trimers extracted. Chlorophyll concentration and A730 were also recorded to determine if the cells were undergoing significant changes. Each value represents an average of two determinations. Data from one of two experiments is shown in the figure. 128

CH3

1.4-naphthoquinone 2-methyl- 2-hydroxy- 1.4-naphthoquinone 1.4-naphthoquinone (vitamin K3) (Lacaphol)

CO2 CO2 CO2

2-cart)oxy- 2-cartoxy- 2-cartx)xy- 1.4-naphthoquinone 3.5-naphthoquinone 3.7-naphthoquinone

CI (CH3)

CI

2.3-dichloro- 5-hydroxy- 1.4-naphthoquinone 1.4-naphthoquinone (2-methyl = plumbagin)

Fig. I 129

TIC 2100 T2.5(A662479.6) I.IEt « 80 ' T3.1(A386909.9) £ I 6040 - 20 ae 0 — ^ 0 1.0 7.2 13.4 19.6 25.8 32.0

Mass 436.011.0 ^00 T17.6(A4230.1) 452 lA T2.6(A1819.2) £ so T11.7(A1208.0) _£ se 0 A_. , . 0 1.0 13.4 19.6 25.8 32.0

Mass 450.011.0 T19.9(A5850.0) 673 2100M £ 50 c T2.5{A699.2) sP 0 - 0 1.0 7.2 13.4 19.6 25.8 3Z0

Man 748.011.0 ^00 T29.1(A1016Jii22 (A £ 50 c 0 0 1.0 13. 13.4 19.6 25.8 32.0 Ratantion Tinw (Min)

Fig. 2 130

2.0 - 1.5 - PhQ-Me 3 1.0 - CO.S - 0.0 - -0.5 ' I ' I ' I 1-' 240 280 320 240 280 320 240 280 320 Wavelength (nm) Wavelength (nm) Wavelength (nm)

A I I I I I I I I 10 15 20 25 30 35 Time (min)

Fig. 3 131

B

0.04 -

-0.04 J SIS' ms Wild type 0.8- menB26 •.V B; -I ms 3 2 ms r-0.6 -

3914 3 ms --0.4- 19 3 ms 2330 ms

.

10' 10° 10' 10- 10- 10" 10'^ 10"' 10° 10' 10^ 10^ lo" Time, ms Time, ms

0.04 H 0.04 - 0 nr -0.04 H -0.04 - 4 5 ms 5 "<5 menB26 + 55 0ms 2.0 - 2.5 - -.2.0 <-• 15 — 43 ms O ^1.5 H 1720 ms <1 2320 ms 0.5 menB26 + 2-C02-1,4-NQ I 0.5 H 0,0 -i 0.0 1 i 1 r 1 I

10' 10° 10^10^ 10^ 10^ 10* 10" 10^ 10^ 10 10" Time, ms Time, ms

0.6 -1 0.04 A 0.4 - 0.2 - 0 4- 0 -0,04 ^ -0.2 H -05 0 51 ms menB26 + 2-C02-3,5-NQ menB26 + K-, 2.5 - „~2.0 - 3 3 ms 38 ms O "o 2 •X X,1.5 H K 2390 ms 9 9 ms % 1.0 H 2120 ms 5 n 05 - 0.0 —1 i i 1 1—'I r 0 -I 10"' 10° 10' 10^ 10^ 10' 10^ 10'^ 10'^ 10"' 10° 10' 10^ 10^ 10" Time, ms Time, ms Fig. 4 132

menB26 + K, menB26 + 2-CO2-1.4-NQ menB26 + K3 menB26 + 1,4-NQ menB2&

I I [ I I I I I I I II I I I I I I ''' I I 'I' I I t 1 I I I I I I I I I I M I I M I I I I I I I I 1 I I [ I 2.010 2.008 2.006 2.004 2.002 2.000

Fig. 5 133

100 -

80

a. 60 7 -n Q_ 6 -

Si, 40 -

20 - PhQ + Antibiotic PhQ 1000 2000 3000 4000 Time (min.) 0 H

0 1000 2000 3000 4000 Time (min.)

Fig. 6 134

CHAPTER 6. RECRUITMENT OF A FOREIGN QUINONE INTO THE A, SITE OF PHOTOSYSTEM I IV. IN VIVO REPLACEMENT OF THE NATIVE QUINONE IN THE PHYLLOQUINONE-LESS MUTANTS OF SYNECHOCYSTIS SP. PCC 6803 BY 1.2- N APHTHOQUINONE'

A paper to be submitted to the Journal of Biological Chemistry T. Wade Johnson^'", Boris Zybailov^, A. Daniel Jones*. Stephan Zech. Dietmar Stehlik. John H. Golbeck*° and Parag R. Chitnis'°

Summary The phylloquinone production has been short circuited by the interruption of the menA and rnenB genes, which are pan of the phylloquinone biosynthetic pathway in Synechocystis sp. PCC 6803 (Johnson. T. W. et al. (2000) /. Biol. Chem. 275: 8523-8530; Zybailov, B. et al. (2000) J. Biol. Chem. 275: 8531-8538); Semenov et al. (2000) J. Biol. Chem. 275, 23429- 23438). It has been determined that for the phylloquinone-less mutants that, plastoquinone-9

(PQ) • occupies the A, site and functions as an electron transfer cofactor from AQ to the FeS clusters in Photosystem I (PSI). This is the first example of a phytylated naphthoquinone derivative, with the carbonyls in the 1.2 position instead of the 1,4 position, being incorporated into the A, site in vivo. The quinone also is utilized as an electron transfer intermediate from AQ to Fx- This was achieved by supplementing the growth medium of inenB' mutant cells with 1.2-naphthoquinone. This result indicates that the specificity of the

' This work is supported by grants from the National Science Foundation (MCB 9723661 to JHG. MCB 9723001 to PRC. MCB 9723469 to DAB. and DBI 9602232 to Penn State University) and to Deutsche Forschengsgemeinschaft SFB 312 Al to DS. Journal Paper No. J-I8740 of the Iowa Agriculture and Home Economics Experiment Station. Ames, Iowa, Project No. 3416 and supported by Hatch Act and State of Iowa funds. " 'Department of Biochemistry. Biophysics and Molecular Biology. Iowa State University. Ames, lA § 50011 Department of Biochemistry and Molecular Biology. The Pennsylvania State University. University Park. PA 16802^"^ Department of Chemistry. The Pennsylvania State University. University Park, PA 16802 " Abbreviations: PQ, plastoquinone, plastoquinone-9; PSI, photosystem I; PSII, photosystem 11; EPR, electron paramagnetic resonance; HPLC, high performance liquid chromatography; GC, gas chromatography; MS. mass spectroscopy; DPIP. 2,6-dichIorophenol-indophenol; PCR, polymerase chain reaction; SDS, sodium dodecyl sulfate; PAGE, polyacrylamide gel electrophoresis; Tris, tris(hydroxymethyI)aminomethane; Tricine. N-[2-hydroxy-l.l-bis(hydroxymethyl)ethyl]glycine; 135

phyiyl transferase enzyme is relatively nonspecific in terms of the groups present at the 3- and 4-positions of the 1,2-naphthoquinone ring. There is a population of phytyl- 1.2naphthoquinone and methyl-phytyl-1,2-naphthoquinone. Acknowledging that phytyled naphthoquinone derivatives are present upon supplementing the growth medium of menB mutant cells with 1,2-naphihoquinone, the phytyl tail is considered a necessary part of the molecule for displacement of plastoquinone-9 from the A, site. Introduction Substituted benzo- and naphtho- quinones are utilized as electron and proton transfer cofactors in electron transport chains (1-3). These quinones have two distinct characteristics a relatively polar aromatic ring which and a non-polar isoprenoid 'tail' of various chain lengths. Benzoquinones such as plastoquinone-9 (PQ) is comprised of a single substituted aromatic ring with a nine unit isoprenoid 'tail' and Ubiquinone (UQ) another substituted single aromatic ring with a long isoprenoid tail both function in membrane-bound protein complexes. Their purpose is to act as a electron/proton carrier during photosynthetic activity and/or as respiratory electron transport cofactor. PQ has multiple functions in photosystem II (PSII). It acts as a bound one-electron acceptor in the Q^ site and as an exchangeable 2- eIectron/2-proton acceptor in the QB site. The reduced PQH; then dissociates from the QB site, diffusing into the membrane, which is then replaced by an oxidized PQ. Photosynthetic reaction centers (RCs) of purple bacteria use UQ or menaquinone in a similar manner. In photosystem I (PSI), phylloquinone (PhQ), a methyl-substituted NQ with a largely saturated phytyl tail. It functions as bound one-electron redox intermediate in the A, site. Each quinone displays redox properties that are different for each site of interaction in PSI and PSII (4) which is conferred largely by the protein environment. We hope to lake advantage of protein induced altered redox properties upon replacement of the PhQ with alternative exogenously supplemented quinones. This will allow us understand the structural determinants that allow PhQ to function with high efficiency in the A, site of PSI. The goal was the biological replacement of the native PhQ with a variety of benzylquinone and naphthoquinone derivatives. Itoh and coworkers, and Biggins and coworkers used organic solvents to extract phylloquinone from lyophilized PSI complexes, and reconstituted the empty A, site with various benzoquinone, naphthoquinone. 136

and anthraquinone head groups (5-9). The extraction procedure removes up to 90% of the chlorophyll and most of the carotenoid molecules along with the two molecules of PhQ, leading to the significant possibility that structural perturbations might have occurred to the PSI complexes. To circumvent this problem, we developed a genetic technique to accomplish the in vivo removal of the native PhQ. In a series of three papers (10-12). we reported the generation and characterization of PQ-less mutants of the cyanobacterium Synechocystis sp. PCC 6803. We accomplished this by interruption mutagenesis of two genes, men A which codes for a phytyl transferase, and menB which codes for 1.4- dihydroxy-2-naphthoate (DHNA) synthase. Our strategy was to preclude PhQ production with the intent of creating an empty A, site. However, we found that the mutant PSI complex is occupied by a PQ, supposedly recruited from the thylakoid membrane. By EPR and optical spectroscopic analyses PQ was show to occupy the A, site at the same distance from PyOO" and with the same orientation as phylloquinone in the wild type. It also functions as an intermediate electron transfer carrier between Ao and Fx. We have shown that is it possible to introduce various quinones in the A, site in vivo by utilizing the unmodified phytyl transferase and methyl transferase enzymes in the remaining biosynthetic pathway of the menB' mutant cells, chapter 5. However, this is the first example of a quinone with the carbonyls in the 1,2 positions being phytylated and incorporated into the A1 site of the menB" mutant. Material and Methods Cyanohacterial Strains and Growth. A glucose-tolerant strain of Synechocystis sp. PCC 6803 was used as the wild type strain. The menA mutant strain lacks functional phytyl transferase and has been previously described (10). The menR strain used here imenB26) was derived as a second site suppressor of the menBlS' strain reported earlier (10). Description of the menB26' mutant is in chapter 5 of this dissertation. The wild type, menA and menB26' mutant cells were grown in the BG-11 medium under condition described in Ref. (13) and in Chapters. Growth Rates of the Wild type and Mutant Cells with 1.2-naphthqouinone Supplement. Monitoring growth rates and determining chlorophyll concentrations are described elsewhere in Chapter 5 and in Ref. (14), respectively. 137

Isolation of Thylakoid Membranes and PSI Particles Thylakoid membranes were prepared from cells and stored as described by (15), with the subsequent preparation of the PSI complexes described in Chapter 5. Measurement of Photosystem I Activity Steady-state rates of electron transfer for isolated PSI complexes were measured using cytochrome as electron donor and flavodoxin as electron acceptor as described in (16). The measurement of P7(X) phoiooxidation in whole cells was based on the techniques developed by Harbison and adapted to cyanobacteria as described in (17). Analysis of Phylloquinone using HPLC and Mass Spectrometry PSI samples were prepared and quinones were analyzed on an equal chlorophyll basis, as previously described (10). However, the solvent gradient was modified to a gradient elution (solvent A = methanol; solvent B = isopropanol; 100% A from 0-10 min to 3% A/97% B at 30 min, hold until 40 min.) at 1.0 mL/min. A solution of PhQ (40 mM) was prepared in absolute ethanol and kept at -20° C as a standard for calibration. Q-band EPR Spectroscopy of Photoaccumulated PSI Trimers Photoaccumulation experiments were carried out as described in (10). Optical Kinetic Spectroscopy in the nIR Region Optical absorbance changes in the near-IR were measured using a laboratory-built spectrophotometer (12), following the same sample preparation procedures. Results Growth of Cells with Naphthoquinone Supplements Growth rates of the wild type, menA', and menB26' mutants were measured in cells grown in BG-11 medium containing bicarbonate (Table 1). Photoautotrophic doubling times for the wild type (26 h) were nearly three times faster than the mutants {menA' 86 h: and menB26\ 77 h). Photomixotrophic growth uses both respiration (5 mM glucose) and photosynthesis for energy metabolism. The wild type and menB26' mutant had similar doubling times of 14 h and 15 h, respectively, but the menA' strain was sluggish with a doubling time of 26 h. Photoheterotrophic growth included addition of glucose for respiration and of 10 ^iM DCMU to block PSII activity. This effectively prevented 138 photosynthesis. The wild type, nieiiA and menB26' mutant strains had similar doubling times indicating that the respiratory energy metabolism function normally. When the cells attained a culture density of ~0.15 O.D.. 1,2-naphthoquinone was added to the growth medium of the wild type. menA and menB26' mutant cells to a final concentration of 5 to 10 p.M. Wild type doubling times were not affected significantly by addition of 1,2-naphthoquinone to the growth medium. The metiA' mutant was stressed by the addition of the naphthoquinone derivative and subsequently died (Table 1). The menA' mutant can not attach the phytyl tail to the naphthoquinone head group, these results would indicate that the 1,2-naphthoquinone does not displace PQ. Adding 1,2-naphthoquinone to the menB26' mutant had only a marginal effect on growth rates. Compared with the decreasing doubling times of other supplemented naphthoquinones that are incorporated into PSI, the 1,2-naphthoquionone was unlikely to be phytylated and incorporated into PSI (Chapter 5). Vitamin K, and 2-C02-l,4- naphthoquinone supplements have doubling times less than 50% that of basal growth rate. The 1,2-naphthoquinone supplement decreases the doubling time by only 21%. This is more on the order of other naphthoquinones that are not phytylated (Chapter 5). LC-MS Analysis of Pigment Extracts from PSI Complexes The presence of quinones in the extracts of purified PSI trimers was examined using HPLC coupled to a chemical ionization time-of-flight mass spectrometer. Chlorophyll a (ni/z 892) was identified with a retention time of 19.0 minutes. Phylloquinone (m/z 450) and plastoquinone-9 (m/z 748) had characteristics retention times of 19.9 and 29.1 min. respectively (Fig. 2). Known standards were used to confirm each of these peaks. Nearly identical chromatograms were obtained for organic solvent extracts of the PSI trimer preparations of the wild type. menA' and menB26' mutants, except that the PhQ peak at 19.9 min was missing in the mutants (Table 2). Quantification by MS is difficult, given that each molecule has slightly varying ionization characteristics. We determined the number of counts at a particular molecular weight ion, thereby crudely calculating the area under the peak. The amount of plastoquinone-9 in the wild type trimers appears to be very small ( < 139

1%) and has not been detected to be involved in electron transfer discussed funher in Chapter 5. When menB26' cells were grown with 1.2-naphthoquinone, the LC-MS analysis of pigments extracted from isolated PSI complexes showed a peak at the same retention time and with the identical mass (m/z 450) as phylloquinone (Figure 2). It is likely that this peak represents the analog of phylloquinone. the methyl derivative of a methyl-phytyl-1.2- naphthoquinone (1,2-phylloquinone). Accordingly, the positions of the carbonyls apparently do not appear to affect the retention times of the 1.2- and 1.4- substituted naphthoquinones. Spiking the 1.2-phylloquinone derivative with authentic phylloquinone showed one peak at 19.9 minutes with some line broadening; however, the peak was not shifted nor did it create a doublet, data not shown. These samples also contained a species with a mass of 436 m/z and retention time of 17.6 min. This species likely represents unmethylated phytyl-1.2- naphthoquinone. The ratio of phytyl-1,2-naphthoquinone to the 1,2-phylloquinone derivative is 2 in the menB26' mutant with supplemented 1,2-naphthoquinone. This suggests that the rate of methylation is slower than the rate of phytylation. and that the pool of unmethylated derivative is available for assimilation into the A, site. The presence of demethyl 1.2- phylloquinone in the PSI complex also shows that the methyl group is not required for the assembly into the A, site. The total 1,2-phylloquinone derivative (methylated and unmethylated species) to plastoquinone ratio is 3:1 for 1.2-naphthoquinone. The location of the phytyl tail and methyl group on 1,2-naphihoquinone are not known, but were likely to be on the 3 and 4 positions. However, the location of substitutions are not known. Investigation by a collaborator. Dr. Gobleck, is underway. HPLC Analysis of Pigment Extracts To complement the LC-MS analysis of the pigment extracts from PSI complexes, u e examined the absorption spectra of the HPLC peaks with a UV-visible range diode array detector. Quantitative data can be estimated from the UV-visible. since the extinction coefficients of the molecules are known. The retention times in the HPLC analysis were consistent to those in the LC-MS studies, with slight variations that are attributed to different lengths of the connecting tubes between instruments. 140

Coelution with other pigments, namely chlorophyll a at 19 minutes and ji-carotene at 28.5 minutes, allowed comparison of the wild type chromatograms to the phylloquinone-less mutants. Elution and absorption properties of phylloquinone and plastoquinone have been described previously in Chapter 5 and Ref. (10). Plastoquinone is only observed in the PSI trimers of phylloquinone-less mutants. Identification of the phytylated 1,2-naphthoquinone compounds was complicated by coelution with chlorophyll and other molecules. 1,2-phyIloquinone has been identified at 20 min.. However. phytyl-l,2-naphthoquinone identified at 17 minutes by LC-MS but not confirmed by absorption spectroscopy. The phytylated-1,2-naphthoquinone was not identified as it was masked under the chlorophyll peak and that absorption detection is three orders of magnitude less sensitive (18). The 1.2-phylloquinone to plastoquinone ratio is 1.5:1 with the chlorophyll to total quinone ratio at 63 based on absorption spectroscopy. Supposedly, the remainder of the A, sites are either empty or occupied with the phytyl-l,2-napthoquinone. P~0(f Optical Recombination Kinetics To determine if the quinone is within the Ai site, we probed its activity by obser\'ing the kinetic backreaction charge recombination of PTOO"^. In PSI trimers isolated from the wild type, after a saturating flash the reduction of the PTOG"^ RC is multiphasic (12,19). When measured by 810 nm near-IR optical spectroscopy and in the absence of external electron acceptors, the typical wild type spectrum shows a majority of the P700' being reduced with a lifetime of 60 - 80 ms (12,19). There is also a long-lived (>2000 ms) kinetic phase of P700* reduction from direct reduction of the P700" by reduced DCPIP that contributes to 20 % of the total absorbance change. In PSI trimers isolated from the menB26 mutant, the reduction of P700'^ is multiphasic (Figure 3). When menB16' is measured in the absence of external electron acceptor, P700^ is reduced with lifetimes of approximately 3.0 ms and 19 ms in a 7:1 ratio. The 3 ms phase has been attributed to plastoquinone in the Ai site with the electrons originating from the iron sulfur clusters (12). The mutants also e.xhibit a minor long-lived kinetic phase attributed to electron donation of DCPIP to the oxidized P700* that contribute the remainder of the absorbance change. The 141 kinetic back reaction phases for both the menB26' mutant are consistent with plastoquinone being in the Ai site of PSI and acting as an electron intermediate from Ao to Fx (12). Monitoring the kinetic back reaction of the quinone supplemented cells indicates whether plastoquinone still remains active in the A, site or is replaced by a naphthoquinone. Table 4. The 1,4-naphihoquinone head group in the A, site causes a loss or decrease in contribution from the 3 ms phase as seen in chapter 5. The 1,2-naphthoquinone supplement was fit to four exponentials instead of the usual three. In fitting the curve, three exponentials was not adequate in defining all the phases. With four exponentials there is a clearly defined minor 3 ms phase (8%) attributed to plastoquinone. The 34 ms (27%) and 10 ms (25%) phases may be the active phytylated 1,2-naphthoquinone within the A, site. Two phases may be the result of PSI being in different confirmations (12). An alternative is that both pathways through PSI (and A,) are accepting electrons. The latter option has not been shown in cyanobacteria. The long lived decay of 2060 ms (40%) is attributed the direct reduction of P700^ by DPCIP (12). Given that the plastoquinone phase is present in reduced amounts indicates that there is another quinone also in the active site. The longer phases (2060 ms) are in higher contributions than normal, indicating more damaged reaction centers or that 1.2-naphthoquionone is not allowing reduction of P700*. This result confirms that not only are the naphthoquinone is contained in PSI. but also active within the A, site as well. This data coupled with the pigment extract analysis suggests that the phytylated quinone has displaced plastoquinone. CW EPR Spectroscopy at Q-Band of Photoaccumulated A, To further probe the contents of the A, site, we photoaccumulated the A, signal for Q-band EPR. Plastoquinone has been previously identified in the A, site of the menB mutants (12). Figure 4 shows the Q-band EPR spectra of photoaccumulated A," in PSI complexes isolated from wild type, menB26', and menB26' with 1,2-naphthoquinone supplemented. The menB26' mutant strain shows the previously identified plastoquinone photoaccumulated in the A, site with wild type shown for comparison. The nienB26' + 1.2- naphthoquinone spectrum contains a naphthoquinone group that is being photoaccumulated. Some hyperfine splitting is evident indicating the 1,2-naphthoquinone ring is also at least 142 panially methylated. This spectrum complements the PTOO" back reaction rates indicating that there is an active naphthoquinone within the A, site. Discussion We generated cyanobacterial mutant strains that are unable to synthesize phylloquinone, to understand function of phylloquinone in photosynthetic organisms (10). The PSI complexes of these mutants contain plastoquinone in their A, site (11). The redox potential of phylloquinone is significantly different in the PSI environment as compared to being free in solution ((12), and references therein). To further probe the structure-function relations of quinones in the A, site of PSI. we added 1.2-naphthoquinone as a supplement during growth of phylloquinone-less mutants. Once inside the cell, the supplemented quinone could undergo the remaining natural steps in phylloquinone synthesis; attachment of the phytyl tail by phytyl transferase (the menA product) and possible methylation by methylase (the product of menG). The newly synthesized phylloquinone or derivative competed with and replaced the plastoquinone in the A, site of the PSI complexes in the mutants. Much effon has been spent in extracting the phylloquinone from the A, site to insert alternative quinones in vitro (20,21) (22,23). These benzo- naphtha- and anthraquinones have been shown to bind in A, and to accept electrons from AQ (20,24). But only 1.4- naphthoquinones with a hydrocarbon tail can transfer the electron forward to the [4Fe-4S] cluster Fx (25). Two hypotheses have been proposed: the electron transfer is dependent on an appropriate redox potential or that a specific orientation of the quinone is necessary, which is provided by the isoprene tail. Initially, we studied the growth characteristics of the menB26' mutant strain and wild type with naphthoquinone supplement (Table 1). In chapter 5. several of the phytylated naphthoquinone supplements resulted in decreased the doubling times of the cells. It was further shown that those phytylated-naphthoquinones were utilized in PSI. The remaining unutilized naphthoquinones did not significantly change the growth rates. Wild type cell growth rate slowed upon addition of supplemented quinones, except for 1.2- naphthoquinone. It appears at first glance that 1,2-naphthqouinone is not being utilized by 143 the ceils given that the growth rate does not significantly increase for menB26'. But the compound was apparently toxic to the menA' strain. The pigment extracts separated by HPLC and analyzed by both mass spectrometry (Table 2) and ultraviolet absorbance (Table 3) detect molecules identified as phytyl-1,2- naphthoquione and methyl-phytyl-l,2-naphthoquinone. The locations of the phytyl and methyl groups are not known but likely arranged on the 3 and 4 carbon positions. By mass spectroscopy the area under the curve shows that there is about twice as much unmethyiated derivative as the 1,2-phylIoquinone. It shows that both molecules are capable of inserting into the A, site. However, we can not speculate as to the cause of the ratio. It is possible that the order of the enzymatic reactions becomes important given that op>en sites for substitution are now different than with the 1,4-naphthoquionone carbonyls. The distribution of substitutions are not resolvable by this type of spectroscopy. The amounts of the phytylated compounds are too small to perform NMR or other structural determining techniques. The ratios of phytylated naphthoquinone to plastoquinone is the low, indicating that either 1.2- naphthoquinone is not efficiently phytylated by phytyl transferase or that the head group does not fit well into the A, site decreasing the binding affinity. The activity PSI, as monitored by the flash induced P700 back reaction measures the content of the Ai site. Table 4. The 1,2-naphthoquinone supplemented cells still have a distinct 3 ms phase attributed to plastoquinone (Figure 3). However, the contribution of the phase has decreased from approximately 65% in menB26' to 8%. The contribution of the longer phases of 10 ms and 2060 ms are greater also in this sample than in the controls. The 10 ms phase may be attributed to damage reaction centers (12). The 2060 ms phase is the direct reduction of P700+ by the external donor. The high amount of contribution from this phase may be from the electron being very slow unable to overcome the thermodynamic barrier or that the electron is essentially trapped in AT with the external electron donor reducing P700^. Measuring the forward electron transfer from Ai to the iron sulfur clusters

(FX/FA/FB, monitored at 430/480 nm) has yielded inconclusive results as to the electron transfer from Af to FX, data not shown. The identity of the 34 ms phase has not been determined, but could be the back reaction from the iron sulfur clusters through the 1.2- naphthoquinone. This is plausible given that phylloquinone has a kinetic back reaction time 144 of about 60 - 85 ms (observed and Ref. (12)) and the demethylphylloquinone (generated from the menG methyl transferase gene disruption) has a time of about 50 ms (26). The 34 ms phase is comparable to a properly aligned naphthoquinone head group occupying A|. We have not yet determined if it is from the methylated or unmethylated 1,2-naphthoquinone head group. Photoaccumulated EPR reveals that menB26 with 1,2-naphthoquinone supplement does not contain plastoquinone (Figure 4). Wild type and menB26' differ in both g-tensor principle values and linewidths (II). The 1.2-naphthoquinone supplement generates a photoaccumulated spectrum closer aligned with a naphthoquinone ring than benzoquinone. Some hyperfine details are also visible, albeit at low resolution. Thus some form of phytylated 1.2-naphthoquinone is being incorporated into the Ai site. As reported earlier, plastoquinone occupies the Ai site of PSI in the phylloquinone- less mutants for menR and menA' ofSynechocystis sp. PCC 6803 (10,II). We show that the addition 1.2-naphthoquinone. the cells complete the synthesis by attaching a phytyl tail producing phylloquinone derivatives. These phytylated compounds out compete and replace the resident plastoquinone in the existing PSI complexes. We used HPLC and LC-MS to detect the presence of phytylated-1,2-naphthoquinone molecules in PSI complexes. Inside the PSI complex, the phytylated naphthoquinone derivatives did apparently function as an intermediate in the electron transfer from Ao to Fx. Optical and EPR spectroscopy was used to study activity of the phytylated-1,2-naphthoquinone molecules in the A| site of the PSI complexes. When the foreign quinones replaced plastoquinone partially, proportion of kinetic phases in the P700 reduction kinetics was used to indicate relative contributions of the resident plastoquinone and foreign quinone. This is the first example of a utilized quinone with the carbonyls in the 1,2- position. This confirms that a variety of quinones that contain a phyt\'I tail can be incorporated in vivo into the Ai site of PSI and perform electron transfer. References 1. Nugent. J. H. (1996) Eur. J. Biochem. 237(3), 519-531 2. Golbeck. J. H. (1993) Proc. Natl. Acad. Sci. USA 90(5), 1642-6 3. Nitschke, W., and Rutherford, A. (1991) Trends Biochem. Sci. 16, 241-245 145

4. Trumpower, B. L. (1982) Function of Quinones in Energy conserving Systems. Trans. 0-12-701280-x (Trumpower. B. L.. Ed.). Academic Press, New York 5. Iwaki, M.. and Itoh. S. (1994) Plant Cell Physiol. 35, 983-993 6. Kumazaki. S., Iwaki, M., Ikegami, I., Kandori, H., Yoshihara, K., and Itoh. S. (1994) J. Phys. Chem. 98, 11220-11225 7. Itoh. S.. and Iwaki, M. (1991) Biochemistry 30(22), 5340-6 8. Biggins, J. (1990) Biochemistry 29(31), 7259-64 9. Iwaki, M., and Itoh. S. (1989) FEBS Lett. 256, 11-16 10. Johnson. T. W.. Shen. G.. Zybailov, B., Kolling, D., Reategui, R., Beauparlant. S.. Vassiliev, I. R., Bryant. D. A., Jones, A. D.. Golbeck, J. H.. and Chitnis, P. R. (2000) J Biol Chem 275(12), 8523-30 11. Zybailov, B.. van der Est, A., Zech, S. G., Teutloff, C., Johnson, T. W., Shen, G.. BittI, R., Stehlik, D., Chitnis, P. R., and Golbeck, J. H. {2Q0Q) J Biol Chem 275(12), 8531-9 12. Semenov. A. Y., Vassiliev, I. R.. van Der Est. A., Mamedov. M. D.. Zybailov, B.. Shen, G.. Stehlik, D., Diner. B. A.. Chimis, P. R.. and Golbeck, J. H. (2000) J Biol Chem 275(31), 23429-23438 13. Rippka, R., Deruelles, J., Waterbury. J. B., Herdman, M., and Stanier, R. Y. (1979) J. Gen. Microbiol. Ill, 1-61 14. Arnon. D. (1949) Plant Physiol. 24, 1-14 15. Sun. J., Xu. Q.. Chitnis. V. P., Jin. P.. and Chitnis. P. R. (1997) J. Biol. Chem. 272(35), 21793-21802 16. Yang. F.. Shen. G.. Schluchter. W. M.. .; Zybailov, B. L., Ganago. A. O.. Vassiliev. I. R.. Bryant. D. A., and Golbeck, J. H. (1998) J. Phys. Chem. B 102(42), 8288-8299 17. Redding, K.. Coumac, L., Vassiliev, I. R.. Golbeck, J. H., Peltier. G., and Rochai.x. J. D. {\999) J Biol Chem 274(15), 10466-73 18. Careri. M.. Mangia, A.. Manini. P., and Taboni, N. (1996) in Fresenius' J. Anal. Chem. Vol. 355, pp. 48-56 19. Vassiliev. I. R.. Jung, Y. S., Mamedov, M. D.. Semenov, A., and Golbeck. J. H. (1997) Biophys 7 72( 1), 301 -15 146

20. Itoh. S.. Iwqaki. M.. and Ikegami. I. (1987) Biochim. Biophys. Acta 893, 508-516 21. Itoh. S.. and Iwaki. M. (1989) FEBS LETT. 243, 47-52 22. Rustandi. R. R.. Snyder, S. W.. Feezei, L. L.. Michalski. T. J.. Norris, J. R-. Thumauer. M. C., and Biggins. J. (1990) Biochemistry 29(35), 8030-2 23. Setif. P.. Ikegami. I., and Biggins. J. (1987) Biochim. Biophys. Acta 894(2), 146-56 24. Biggins. J., and Mathis. P. (1988) in Biochemistry Vol. 27. pp. 1494-500 25. Rustandi, R. R.. Snyder, S. W., Biggins. J., Norris. J. R.. and Thumauer, M. C. (1992) in Biochim. Biophys. Acta Vol. 1101. pp. 311-20 26. Shen. G.. and Bryant. D. (2000) 147

Table 1

Growth of wild type and mutant strains of Synechocystis sp. PCC 6803 with supplemented naphthoquinones

Doubling Time (h) Supplement Wild-type menA' menB26' None 26 ± 3.0 86 ± 5.0 77 ±4.0 Glucose 14 ±2.5 26 ± 8.7 15 ±3.4 DCMU + Glu 17 ±2.0 19 ± 4.3 19 ±2.3 l,2-NQ 25 ± 2.2 died 61 ±5.5

Light intensity was 50 |iE, normal light. Error is in the fit linear slope and was between 5-15%. with an n = 3 or more. Concentrations of naphthoquinones. DCMU. and glucose were 5 (iM, 10 (iM, and 5 mM, respectively. Initial cell density was 0.10 O.D. at 730 nm. The "none" column is basal growth rate in BGl I medium. * Indicates that the cells doubled once and then died. N.D. = Not done. Numbers without a statistical error were derived from less than three observations. 148

Table 2

Quinone analysis of the PSI pigment extracts by LC-MS of the wild type and menB26 mutant with naphthoquinone supplements

Sample Mass Ratios DeMePhQ/ PhQ/PQ NQT/PQ PO wild type nd > 100 > 100 menB26' nd nd nd menB26' + PhQ nd > 100 > 100 menB26'+ 1.2-NO 2 1 3

Table: Chromatograms were run on 0.026 mg chlorophyll content of the PSI trimers. Sample injections (35 |iL) were made on a 4.6 mm x 25 cm Ultrasphere Cig column (4.6 mm x 250 mm) with 5 fim packing (Beckman) using gradient elution (solvent A = methanol; solvent B = isopropanoi; 100% A from 0-10 min to 3% A/97% B at 30 min, hold until 40 min.) at 1.0 mL/min. Mass 436, 450, and 748 represents PhQ minus a methyl group (or isomer), PhQ. and PQ n = 9, respectively. Mass ratios are based on the intensities of the selected peaks. The peak areas used for the ratios are the m/z ±1.0 mass units. For a given standard deviation there are 4 to 6 independent results (n = 4 to 6). nd = not detected. 149

Table 3

Quinone analysis of the PSI pigment extracts by HPLC-UV/vis of the wild type and inenB mutant with naphthoquinone supplements

Sample Chl/Qj NQj/PQ

wild type 42 71 menB26' 55 — menB26' + PhQ 44 59 menB26'+ 1.2-NQ 63 2

Table: Chromatograms were run on 0.026 mg chlorophyll content of the PSI trimers. Sample injections (35 |iL) were made on a 4.6 mm x 25 cm Ultrasphere Cig colurmi (4.6 mm x 250 mm) with 5 (im packing (Beckman) using gradient elation (solvent A = methanol; solvent B = isopropanol; 100% A from 0-10 min to 3% A/97% B at 30 min, hold until 40 min.) at 1.0 mL/min. PhQ-CHj, PhQ, and PQ n = 9, represents PhQ minus a methyl group. PhQ, and piastoquinone n = 9. respectively. Differences in retention times relative to LC-MS are attributed to slightly different lengths of the connecting tubes. Retention times for the components are PhQ-OH = II.0 min, demethylPhQ = 17.7 min, PhQ = 19.9 min. PQ = 29.1 min. For a given standard deviation there are 4 to 6 independent results (n = 4 to 6). 150

Table 4

Flash induced P700 Back Reaction at measured at 810 nm on PSI trimers

Sample ms back % of A, content reaction contribution wild t\'pe 62 40 PhQ menB26' 2.8 65 PQ menB26' + PhQ 10 20 PhQ menB26' + 1,2-NO 2.9 9 Mix

TABLE. P700^ reduction kinetics in PS I complexes isolated from the menB mutant. Flash induced optical transient measured at 811 nm during a single flash. Sample conditions: 4- sided fluorescence cuvette containing sample at 10 |ig/ml in Chi 25 mM Tris, pH 8.3. 0.04% B-DM. 10 mM ascorbate and 4 (iM DPIP. Excitation wavelength, 532 nm. excitation energy. 1.4 mJ. 151

Figure Legends Fig. 1. Figure of phylloquinone, plastoquinone and 1,2-naphthoquinone. Fig. 2. LC-MS analysis of the pigments extracted from PSI complexes of the menB26 cells that had been grown with 1 ^-naphthoquinone supplement. From top to bottom the chromatograms represent the total ion count at m/z 436, m/z 450, and m/z 748 which correspond to the mass of unmethylated 1.2-PhQ (t = 17.6 min), 1,2-PhQ (t = 19.9 min), and plastoquinone n = 9 (t = 29.1 min), respectively. FIG. 3. P700^ Reduction Kinetics in PSI Complexes Isolated from the menB2(t mutant. Laser-flash induced optical transient measured at 811 nm. Time is plotted on a logarithmic scale, in which a deviation from the horizontal represents a kinetic phase. The e.xperimental data is shown as dots and the computer-generated exponential fits are shown as solid lines, with the lifetimes of each phase shown. The relative contributions of each kinetic phase can be judged by the intersection of the fit line with the abscissa. The error between the e.xperimental data and the fits are shown above each plot. The samples A-C are wild type, menB26'. and menB26' + 1,2-naphthoquinone. respectively. FIG. 4. Photoaccumulated Q-band CW EPR spectra of AF and Q in PSI complexes isolated from wild type, menB26' mutant, and menB26' mutant with naphthoquinone supplements. The photoaccumulation was carried out for 40 min at 205 K.. Instrument settings: microwave power 1 mW. microwave frequency 34.056 GHz. modulation frequency 100 KHz, modulation amplitude 1 G, temperature 205 K. time constant 10 mS, conversion time 10 mS, 100 averaged scans. 152

O CH3

1,2-naphthoquinone Phylloquinone

6

Plastoquinone

Fig. 1 % Intensity % Intensity % Intensity % Intensity o^gSiS oS^SSi oS^SSS

hOH N u>N > CJ (£> 00 CD (D ! s ro (D n I CJ f 2 AN tou>

r 'H <0 I. O) T > oB f» T) { tiS" I fI (O L/iUJ to I bo

S y i s "d <0

fO b

CO ' i fw I k>'

& ft eo bo bo 154

0.10 0.04- 0.05 0 -005 0.04 - -0.10 3:I37ms Wild tvpe menB26 3 - 3 2 ms r»

33K 2 m: ? 2^ •0.4- K < 19 3 ms < 1 H 2330 ms

0 4 "I ^ 1 ^ r iIIii 10' 10"' 10° 10' 10- 10" 10' 10'^ 10"' 10° 10' 10^ 10^ lO"® Time, ms Time, ms

0.04 - 0

-0 04 -I menB26 • 1.2-NQ 2.5

2.0 -i "iisi 1, 0 .j

05 T

0.0 -1. 10" 10 10 10' 10' 10 10 Time, ms

Fig. 3 155

1.5 - —- wild type menB26 - - menB26 + 1.2-NQ 1.0 -

0.5 -

0.0 -

-0.5 -

-1.0 -

2.010 2.008 2.006 2.004 2.002 2.000 1.998 1.996

Fig. 4 156

CHAPTER 7. GENER.\L CONCLUSIONS

Both plants and animals depend on phylloquinone to function in distinctly different roles. In animals, phylloquinone is essential for initiating blood clotting and covalently modifying specific glutamic acid residues. There also has been much effort in understanding the physiological effects of a lack of phylloquinone in the diets of animals and molecules that inhibit phylloquinone's actions. In plants phylloquinone is only utilized as an electron transport intermediate in photosystem I. Each of these uses for phylloquinone is necessar>' for the optimum survival of the organism. But how critical phylloquinone is to the function of photosystem I has not been probed. Therefore we have undertaken a project to understand how phylloquinone is biosynthetically produced. We also strove to understand how photosystem I operated, if at all, when phylloquinone can not be synthesized. Initially we genetically disallowed phylloquinone biosynthesis by disrupting four individual genes in the pathway {menA, menB, menD, and menE^. The menD and menE genes are relatively early in the pathway and code for enzymes that modifies chorismate into a naphthoquinone precursor. The menB gene codes for an enzyme that produces the naphthoquinone head group in phylloquinone. The menA gene codes for an enzyme that attaches the phytyl tail onto the ring. These genes were chosen disruption for different reasons. The menD and menE genes were disrupted to understand where in the biosynthetic pathway is the production of phylloquinone dedicated by these enzymes. The menB gene product was disabled because it prevented phytylation of the ring in the next step. The menA gene product was disabled because it prevented the phytylation of the intact naphthoquinone head, which was thought to be a possible candidate for Ai site, given the lack of phylloquinone. In all cases, phylloquinone was not produced. The implications here are thai four of the seven genes have no alternative pathway for the production of the specific molecules. The menD gene codes for an enzyme in the first dedicated step in the pathua\. The menA disruption shows that the head group DHNA is not the most suitable replacement of phylloquinone in the mutants. However, instead of finding an empty Ai site, plastoquinone was determined to be present in vivo in the phylloquinone-less mutants. Plastoquinone is a fair substitute for 157

phylloquinone activity in PSI. The orientation of plastoquinone is similar to phylloquinone. however there are distinctly different energetic properties. Plastoquinone is less reducing than phylloquinone and that change in redox potential has significant effects on the electron transfer rates through Ai. Efforts in a variet>' of collaborating laboratories are continuing to probe these differences between plastoquinone and phylloquinone. We also determined that phylloquinone and derivatives can be reinserted into the menB mutant in vivo by the addition of phylloquinone or the product of the menB reaction (DHNA). Addition of naphthoquinone analogs of DHNA (vitamin K3, I.4-naphthoquinone) also produces phylloquinone and, possibly, derivatives. The remaining machinery for production of phylloquinone is not affected, allowing us to supplement the cells with alternative quinones which the cells, in turn phytylate and insert into PSI. The phytylated quinones displace plastoquinone. We also show ±at a 1,2-naphthoquinone is also phytylated and methylated and inserted into the A] site. This is the first example of a 1.2 substituted quinone in PSI in vivo. The menB mutant that has been 'fed' these quinones show distinctly different physiological characteristics and changes in the activity of PSI. In the cases where phylloquinone like molecules are in PSI, wild type characteristics emerge. The use of the phylloquinone-less mutants is continuing. We are now able to. at will, substitute quinones into the A] site. We also know what the requirements are for displacing plastoquinone. This leads us in new directions for synthesizing tailor made quinones that meet a particular structure or redox potential. Alternative uses include probing the electron transport from Ao through the desired quinone in Ai to Fx- This will not only allow us a better understanding of the Ai site, but its interactions with its electron transport partners. 158

APPENDIX A. DEVELOPED PROCEDURES AND PROTOCOLS SPECIHC TO DISSERTATION EXPERIMENTS

Al. Cyanobacterial strains and growth. A glucose-tolerant strain of Synechocystis sp. PCC 6803 was used as the wild type strain. The cells were grown in the BG-Il medium, with 5 mM glucose and l)ig/ml antibiotic for the mutants (Rippka et al., 1979). Agar plates for the growth of the stock cells were kept at low light intensity (2 to 10 |i.E m " s ') on a 5 mM glucose and 2 fig/ml antibiotic. Liquid cultures of the wild type and mutant strains were grown heterotrophically under normal light conditions (40 to 60 ^E m " s ') and were bubbled with sterile filtered air. Usually two sets of cells were kept growing. The first set was at normal light levels and the second set at was low light levels to keep the growth rate reduced. Less glucose or no glucose can also be used to keep a liquid stock of cells growing slowly without having to recuiture and potentially contaminate the cells as frequently as cells grown with glucose or at normal light levels. Cell growth was monitored by measuring the optical density at 730 nm (A730) with a Shimadzu spectrophotometer. It is important to be consistent with taking the reading to prevent artificial error. Cells from liquid cultures in the late exponential phase of growth (A730 = 0.8-1.2) were harvested by centrifugation at 5000 x g for 15 min.

Growth on a large scale was achieved by starting a 3 L culture at 0.050 O.D.730 and adding 2.5 mM glucose final concentration in addition to antibiotics for the mutants. The mutants contain antibiotic cassettes that impart resistance to specific antibiotics. Addition of quinones was previously done at around 0.20 - 0.30 O.D.730- Presently quinones were added in the late stage log phase of the growing cells and incubated for 1-3 hours before har\'esting. A2. Growth rates of the wild-type and mutant cells. Fresh cyanobacterial cultures were grown twice with glucose and once or twice, depending on the growth rates of the strain, without glucose to deplete any exogenous glucose stored in the cells. In the late exponential phase of the third liquid culture, the cells were pelleted by centrifugation, washed twice with BGl 1 medium, and resuspended in

BGll and the appropriate antibiotics for the mutants at approximately 10 O.D.730 concentration. For estimating growth rates, the cultures were grown in 6 well culture plates 159

with 8 ml total liquid medium in each well. .A.1I cultiires were adjusted to the same initial cell density (O.D.730 = 0.08 - 0.1). A test well was done to determine the amount of concentrated cells and BGl 1 that were needed to dilute in the 8 ml liquid culture. From this

1 ml of liquid is drawn and O.D.730 was determined. This reading is the initial O.D.730 for the cells in the growth experiment. Final concentration of glucose and DCMU was 5 mM and 10 |iM. respectively. The order of addition to the well was: 1) cells; 2) glucose; 3) DCMU; 4) quinone or other additives; 5) balance of BGI1. Make sure that the cells do not come into contact with the other added ingredients before adding the BGll. Adding the BGl I liquid last will ensure thorough mixing so as not to spike the cells with overly concentrated solutions. The 6-well plates were sealed with parafilm and put on an orbital shaker at 110 rpm under a bank of fluorescent lights at normal light conditions (40 to 60 ^.E m"" s '). Plates were deposited in only one layer on the shaker. Given the differences in shakers, find a speed that the liquid approaches the top of the well but does not go over or up to the lip. Tape the wells down if on the edge of the shaker. Each mutant strain or wild type must be in their own 6-well plate. If you mix strains in an individual plate you will see a growth rate cross over effect. Duplicates are highly advised for each run. The experiments were completed 3-7 times in duplicate. Cell density was measured by removing 200 nl of culture add SOOjil of water (total 1 ml) and measuring the O.D.730 and multiplying by 5. When the cells are thicker (-0.50 O.D.730). 100 |il of culture was drawn with the balance water.

To determine the doubling time, record the O.D.730, time and date of each reading starting when the cells were mixed. Convert the time into minutes and plot time verses the natural log of the O.D.730 of each reading. Determine the slope only for the linear part of the graph. The slope of the line is in units of 1/minutes which should be on the order of lO""* - 10 ^ Dividing 115 by the first two digits in the slope (y.zx 10"*) will give the doubling time of the cells in hours. The number 115 is from 160

The advantage of this method is that you can do two or three times as many samples at once and be assured of nearly homogeneous conditions. It is reasonable to do 12 6-well plates at once, given enough shaker space. .A.3. Chlorophyll Analysis and Oxygen Evolution Measurements. Chlorophyll was extracted from whole cells and thylakoids with 100% methanol. Chlorophyll concentrations were determined according to Ref (MacKinney, 1941). Oxygen evolution measurements were performed using a Clark-type electrode as described in (Shen et al.. 1993). Cells were prepared by starting cultures at the same time and grown photomixotrophically with 5mM glucose for two subculturing. The final liquid culture was made without addition of glucose and grown until mid log growth phase (0.4 -

0.6 OD) and then analyzed. From these cells the initial O.D.730 was recorded and the chlorophyll concentration by methanol extraction of chlorophyll. The cells were pelleted and washed with small amounts of 40mM HEPES/NaOH buffer pH 7.0. The cells were resuspended at about 6-12 O.D.730 ml'' in 40mM HEPES/NaOH buffer pH 7.0. Increasing the concentration past 12 O.D.730 will cause the cells to pellet out. The cells were kept in the dark during the experiments and are only good for 4 - 8 hours after the initial prep. Preperation of the cells were staggered during the experimental runs. The temperature of the electrode chamber was maintained at 25°C by a circulating water bath. Measurements were done on 5 O.D. of cells. Whole-chain electron transport (HjO to CO,) measurements were determined after the addition of 5 mM NaHCO,; oxygen evolution mediated by PS II only was determined after addition of 4 mM p-benzoquinone (BQ). More BQ was used if the Slock solution was not fresh or has significant amounts of black oxidized quinone. Note that an excess of BQ is not going to affect the rate significantly, since there must be an excess of electron acceptor for PSII to operate maximally. The light intensity was 1840 ^E m" s ' as created by neutral density filters. Experiments were repeated in triplicate and averaged. Also one of the initial runs was repeated at the end to determine that the instrument readings have not drifted over time. The final experiment was adding sodium dithionite to distilled air saturated with bubbled water. Add the powder until the oxygen readings do not drop. The difference between the starting point and bottom is your delta (A), which will be used later to determine the rate of oxygen evolution. 161

The units are properly expressed as jimol O2/A730 h ml. The number 0.253 mmol/ml is the concentration of oxygen in air saturated water at 25°C. There is a conversion of 3600 s/h to convert from readings in seconds to hours. Delta (A) is from above. The data was converted to ASCII and imported into a math spreadsheet. Igor Pro 3.14 (Wavemetrics, Inc.. Lake Oswego, OR). By linear regression the slope of linear part of the oxygen evolution line was fitted. It was entered into the following calculation.

0.253";;;?' Y 3600; - (slope) = A j(^ 1 hr y

A4. 77 K Fluorescence Emission Spectra. The low temperature fluorescence emission spectra were measured using a SLM 8000C spectrofluorometer as described (Shen and Bryant, 1995). Cells from the exponential phase of growth were harvested by pelleting and resuspended in 25 mM HEPES/NaOH. pH 7.0 buffer or water. Cells (5 fig Chl.) were diluted in 25 mM HEPES/NaOH. pH 7.0 to a final volume of 30 |il and dark adapted for 30 minutes on ice. To the solution. 70 ^1 of neat glycerol was added, mixed, prior to quickly freezing in liquid nitrogen. The excitation wavelength was 435 nm. On the monochromator. the excitation slit width was set at 4 nm and the emission slit width was set at 2 nm. The emission was scanned twice from 600 nm to 800 nm and averaged. Sodium fluoroscene is a recommended internal standard for maize chloroplasts. It is difficult to use with cyanobacteria, since it intercalates into the cell walls and produce two peaks, a bound molecule spectrum and a free molecule spectnun. Sample cuvettes are long stemmed (10 in.) disposable borosilicate glass pipettes from Fisher. Flame tip closed by a Bunsen burner adjusted to a hot flame by sticking 2-3 mm of the glass tip into the flame until orange hot. Take care to use pipettes that are relatively straight. A curved pipette will not align prop>erly in the beam path of the fluorometer. To set up the fluorometer properly, the following items are needed: fluorometer low temperature kit (face plate, sample holder bracket, top plate with dewer holder and cover, quanz dewer. sample holder which fits into dewer, screws, hex wrench), two 2 L liquid 162

nitrogen dewers. 200 p.! pippetmen, 20 ^il pippetmen, tips, neat glycerol, distilled water, 100% ethanol. 12 in small gauge needle with syringe adaptor. 0.5 ml syringe, flashlight, small mirror, ring stand with clamp for quartz dewer, long thin wooden (not metal) stick, sample cuvettes, kimwipes, tube filled with dryrite and connecting hoses, and icebucket for samples. First assemble the low temperature kit onto fluorometer and attach the air hose with dryrite tube to the face plate. The air blown through the dryrite tube prevent condensation from forming on the liquid nitrogen filled quartz dewer during a scan. Add to the quartz dewer enough liquid nitrogen to fill the clear glass section. Remove bubbles coming from the lower section of the dewer by touching the wooden rod to the bubble at the point of origin. Fill the dewer the rest of the way up with liquid nitrogen. Keep all liquid nitrogen dewers covered at all times to minimize condensation build up. Snow introduced into the dewers will collect onto the tip of the samples cuvette and block the beam path. Third, prepare sample and inject about 2 in. of sample into bottom of cuvette. Shake liquid to bottom. Air bubbles will shatter the cuvette when frozen. Put the sample in the sample holder. Freeze the sample quickly in one of the two 2 L dewers of liquid nitrogen. The spectrum of the glycerol/cell mixture will change if the mixture is left unfrozen. Put the sample in the quartz dewer and align it so that the sample is in the middle of the round flask. Place the dewer in the fluorometer. It may be necessary to align the sample in the beam paths to achieve a reasonable signal. The sample cuvette is a single use piece. The cuvette is non-recoverable and usually breaks upon melting. A5. .\nalysis of phylloquinone using HPLC and mass spectrometry. On an equal chlorophyll basis, the samples were prepared and quinones were analyzed as previously described (Johnson et al., 2000). with the following changes. Membranes containing 0.025 mg chlorophyll were centrifuged at 10,000 x g for 60 min and the supernatant was removed. PSI trimers were concentrated by a Centracon and then lyophilized to dr>'ness. Care must be taken to remove all sucrose from the sucrose gradient first. The pigments were sequentially extracted with 0.5 ml methanol. 0.5 ml 1:1 (v/v) methanol:acetone, and 0.4 ml acetone, and the three extracts were combined. The resulting solution was concentrated by vacuum at 4°C in the dark to dryness. Dry pigment samples were stored at -80 °C until used. If the samples remain dry, dark, and in a -20 °C or colder. 163 the samples will remain stable for several weeks. Solutions will decompose within hours, faster with light. The pigments were resuspended in a 60 ^.1 mixture of 1:1 methanol:isopropanol at ca. 0.8 mg Chi ml''. HPLC separations were monitored with photodiode array UV-visible detection using a Hewlett Packard (Agilent Technologies. Palo Alto. CA) model 1100 quaternary pump and model G1316A photodiode array detector. Sample injections (35 ^,1) were made on a reverse phase 4.6 mm x 25 cm Ultrasphere Cig column (4.6 mm x 250 mm) with 5 |im packing (Beckman Instruments. Palto Alto. CA. using a gradient elution at (solvent A = methanol; solvent B = isopropanol: 100% A from 0- 10 min to 3% A/97% B at 30 min, hold until 40 min.) at 1.0 ml/min. The column was washed after each run (from 100% B to 100% A at 10 min. 0.5 ml/min, 100% A for 5 min at 1.0 ml/min.). A solution of phylloquinone (40 mM) was prepared in absolute ethanol and kept at -20 °C as a standard for calibration and peak identification. Extracts were also analyzed by LC/MS using a Perseptive Biosystems Mariner time-of-flight mass spectrometer using electrospray ionization in negative mode with a needle potential of -3500 V and a nozzle potential of -80 V. A post-column flow splitter delivered column eluent to the electrospray ion source at 10 nL min"'. For the UV/Vis detector, the peak spectrum was analyzed and compared to known standards or analogs. Phylloquinone has a nicely resolvable doublet at 248 run and 270 rmi by the diode array. Plastoquinone has a singlet at 256 nm. Other compounds like chlorophyll and p-carotene are readily identifiable. The spectra are readily quantifiable. .Vlorton's book on Biochemical Spectroscopy (John Wiley and Sons, 1975) contains useful quantitative data. Extinction coefficients used are. PQ-9, 256 nm, 15.200 cm*l/mol: PhQ. 248 nm. 18.600 cm*l/mol; Chi., 615 nm, 13.800 cm*l/mol. Mass detection is much more sensitive than UV/Vis but less quantifiable. The MS spectrum was scanned for a particular mass peak to help determine its position on the absorption spectrum. It is also used to determine if there is any amount of a particular molecule that is not concentrated enough for UV/Vis detection. The following masses were searched for. PhQ. m/z = 450; PQ-9, m/z = 748. Chlorophyll appears at different masses depending on vvhat kind of adduct it forms (Na* or CI'). 164

A6. Optical Kinetic Spectroscopy in the Visible Region. Optical absorbance changes in the visible were measured using a laboratory-built spectrometer consisting of a 400-watt tungsten-halogen source (Oriel), a -meter monochromator (Jarrell Ash Model 82-410) prior to the sample cuvette (to select a measuring wavelength), a sample compartment for a 1 cm x I cm fluorescence (4-sided clear) cuvette, a second _ meter monochromator (Jarrell Ash Model 82-410) after the sample cuvene (to reject the flash and fluorescence anifacts), and a PIN-10 photodiode detector (UDT). Suitable lenses were placed to focus the light on the monochromator slits and to provide a collimated beam through the sample cuvette. The photocurrent was converted to a voltage with a 30 kOhm resistor, the voltage was amplified using a Model 100 amplifier (EG&G) set to a DC bandwidth of 100 kHz. and a signal was digitized using a Model 100 digital oscilloscope (Nicolet Instruments, Austin, TX) in a Power Macintosh 7100/80 computer. The data were transferred to the computer and stored as binary files using LabView 4.1 (National Instruments, Austin, TX) and further processed in Igor Pro 3.14 (Wavemetrics. Inc.. Lake Oswego, OR). Actinic flashes were supplied using a frequency doubled ND-YAG laser (Spectra Physics) operating as 532 nm at a flash energy of 1.4 mJ. Each kinetic trace represents 8 to 32 averages within the digital oscilloscope or on the computer, depending on the setup. The sample cuvette (10 mm x 10 mm) contained 1.0 ml of PSI complex at 50 fig/ml Chi suspended in 25 mM Tris, pH 8.3 with 0.04% P-DM. 10 |iM sodium ascorbate and 4 ^iM DCPIP. The samples were prepared anaerobically from stock solution mi.xtures. A7. Q-band EPR Spectroscopy of Photoaccumulated PS I Trimers. Photoaccumulation experiments were carried out using a Bruker ER300E spectrometer outfitted with an ER 5106 QT-Wl resonator, which is equipped with a port for in-cavity illumination. Cryogenic temperatures were maintained with an ER4118CV liquid nitrogen crv'ostat. controlled with an ER4121 temperature control unit. The microwave frequency was measured with a Hewlett-Packard 5352B frequency counter and the magnetic field was measured with a Bruker ER035M NMR Gaussmeter. The magnetic field was calibrated using y,Y bisdiphenylene-p-phenylailyl (BDPA) complexed 1:1 with benzene. 165

Prior to the photoaccumulation. the pH of the sample (60 |il final) was adjusted with a glycine buffer pH 10.0, final concentration 233 mM. and sodium dithionite was added to a final concentration of 50 mM. This is done in an anaerobic glove box. After incubation in the dark for at least 20 min, the sample was frozen in the dark and placed into the resonator and the temperature was adjusted to 205 K. The sample was illuminated with a 20 mW He- Ne laser at 630 nm for a variable amount of time, usually around 40 minutes. The purpose of a \'ariable illumination time is to remove the Ao peak while fully reducing the Ai site. The dark background was subtracted fi-om the photoaccumulated spectra. EPR spectral simulations were carried out on a Power Macintosh 7300/200 computer using a Windows 3.1 emulator (SoftWindows 3.0. Insignia Solutions, UK) and SimFonia software (Bruker Analytik GMBH). A8. References Johnson. T. W., G. Shen, B. Zybailov, D. Kolling, R. Reategui, S. Beauparlant, I. R. Vassiliev, D. A. Bryant, A. D. Jones, J. H. Golbeck, and P. R. Chitnis. 2000. Recruitment of a foreign quinone into the A(l) site of photosystem 1.1. Genetic and physiological characterization of phylloquinone biosynthetic pathway mutants in synechocystis sp. pcc 6803 [In Process Citation]. J Biol Chem. 275:8523-30. MacKinney. G. 1941. J. Biol.Chem. 268:11678-11684. Rippka. R.. J. Deruelles, J. B. Waterbury, M. Herdman, and R. Y. Stanier. 1979. Generic assignments, strain histories and properties of pure cultures of cyanobacteria. J. Gen. Microbiol. 111:1-61. Shen, G.. J. J. Eaton-Rye. and W. F. Vermaas. 1993. Mutation of histidine residues in CP47 leads to destabilization of the photosystem II complex and to impairment of light energy transfer. Biochemistry. 32:5109-15. Shen. G. Z., and D. A. Bryant. 1995. Pholosynth. Res. 44:41-53. 166

APPENDIX B. RECUITMENT OF A FOREIGN QUINONE INTO THE A, SITE OF PHOTOSYSTEM I II. STRUCTURAL AND FUNCTIONAL CHARACTERIZATION OF PHYLLOQUINONE BIOSYNTHETIC PATHWAY MUTANTS BY ELECTRON PARAMAGNETIC RESONANCE AND ELECTRON-NUCLEAR DOUBLE RESONANCE SPECTROSCOPY

A paper published in the Journal of Biological Chemistry

Boris Zybailov, Art van der Est, Stephan G. Zechi, Christian Teutloffi, T. Wade Johnson. Gaozhong Shen, Robert Bittl. Dietmar Stehlik. Parag R. Chitnis, and John H. Golbeck

Summarj' Electron paramagnetic resonance (EPR) and electron-nuclear double resonance studies of the photosystem(PS) I quinone acceptor, Ai, in phylloquinone biosynthetic pathway mutants are described. Room temperature continuous wave EPR measurements at X-band of whole cells of menA and menB interruption mutants show a transient reduction and oxidation of an organic radical with a ^-value and anisotropy characteristic of a quinone. In PS I complexes, the continuous wave EPR spectrum of the photoaccumulated Q'radical, measured at Q-band, and the electron spin-polarized transient EPR spectra of the radical pair P7001Q", measured at X-. Q-. and W-bands, show three prominent features: (i) QTias a larger g-anisotropy than native phylloquinone, (ii) Q'does not display the prominent methyl hyperfme couplings attributed to the 2-methylgroup of phylloquinone, and (iii) the orientation of Q'in the A| site as derived from the spin polarization is that of native phylloquinone in the wild type. Electron spin echo modulation experiments on PTOO'^Q'show that the dipolar coupling in the radical pair is the same as in native PS I, /.e. the distance between PTOO'^and Q" (25.3 ± 0.3 A) is the same as between P700*and Ai'in the wild type. Pulsed electron-nuclear double resonance studies show two sets of resolved spectral features with nearly axially symmetric hyperfine couplings. They are tentatively assigned to the two methyl groups of the recruited plastoquinone-9. and their difference indicates strong inequivalence among the two groups when in the Ai site. These results show that Q (i) functioning accepting an electron from Ao' 167 and in passing the electron forward to the iron-sulfur clusters, (ii)occupies the A1 site with an orientation similar to that of phylloquinone in the wild type, and (iii) has spectroscopic properties consistent with its identity as plastoquinone-9. 168

APPENDIX C. THE RED-ABSORBING CHLOROPHYLL a ANTENNA STATES OF PHOTOS YSTEM 1: A HOLE-BURNING STUDY OF SYNECHOCYSTJS SP PCC 6803 AND ITS MUTANTS

A paper published in the Journal Physical Chemistry B

M. Ratsep, T. W. Johnson, P. R. Chitnis, G. J. Small

Summary' Low temperature (4.2 K) absorption and hole-bumed spectra are presented for the trimeric (wild-type. WT) photosystem 1 complex of the cyanobacterium Synechocystis sp. PCC 6803. its monomeric form, an mutants deficient in the PsaF, K. L, and M protein subunits. High-pressure- and Stark-hole-buming data for the WT trimer are presented as well as its temperature dependent Qy-absorption and -fluorescence spectra. Taken as a whole, the data lead to assignment of a new and lowest energy antenna Qy-state located at 714 run at low- temperatures. It is this state that is responsible for the fluorescence in the low-temperature limit and not the previously identified antenna Qy-state near 708 rmi. the data indicate that the 714 nm state is associated with strongly coupled chlorophyll a molecules (perhaps a dimer) and possesses significant charge transfer character. The red chlorophylls absorbing at 708 and 714 nm do not appear to be directly bound to any of the above protein subunits. The results are consistent with a location close to the interfacial regions between PsaL and M and the PsaA/B heterodimeric core. It is likely that the red chlorophylls are bound to PsaA and/or PsaB. 169

APPENDIX D. ULTRAFAST PRIMARY PROCESSES IN PHOTO YSTEM I REACTION CENTERS WITH A FOREIGN QUINONE IN THE A, SITE

A paper to be submitted to Journal of Physical Chemistry B

Sergei Savikhin, T. Wade Johnson, Peter Martinsson, Parag R. Chitnis, and Walter S. Struve

Summary Ultrafast primary' processes have been studied in photosystem I core anterma - reaction center complexes from a menB mutant of the cyanobacterium Synechocystis sp. PCC 6803. in which the native phylloquinone secondary electron acceptor Ai has been replaced by a foreign quinone (plastoquinone-9). The kinetics of antenna excitation equilibration and trapping at the P700 reaction center are essentially the same as in native PS I. The steady- state (P700" - P700) absorption difference spectrum in the menB mutant virtually coincides with the wild type, indicating that replacement of the phylloquinone does not strongly perturb the electronic structure of the special pair chlorophylls. However, the difference between time-resolved transient absorption profiles for menB PS I with open and closed reaction centers (excited at 660 nm and probed at 690 nm) is strikingly different from the wild type. The latter (open - closed) absorption difference profile, which isolates the kinetics of reaction center processes, indicates that the Ao"—> Q electron transfer in the menB mutant requires 1.4 ± 0.4 ns. This is -45 times slower than the uppier limit of -30 ps that has been inferred for the time scale of the corresponding Ao"—>A| transfer in native PS I. 170

APPENDIX E. ULTRAFAST PRIMARY PROCESSES IN A CYANOBACTERIAL PHOTOSYSTEM I REACTION CENTER

A paper to be submitted to Journal of Physical Chemistry B

Sergei Savikhin, T. Wade Johnson, Peter Martinsson, Parag R. Chitnis, and Walter S. Struve*

Summar>' The kinetics of primary events in photosynthetic reaction centers have previously been well characterized only in purple photosynthetic bacteria; these have provided a basis for widely used theories of electron transfer in proteins. The primary charge separation in photosystem I of the cyanobacterium Synechocystis sp. 6803 is found to be slower by a factor of 2-3, and the subsequent electron transfer from the primary to secondary electron acceptor is faster by two orders of magnitude, than the corresponding processes in purple bacteria. These differences are intriguing, because they are not in accord with current predictions based on the known reaction center geometries. 171

APPENDIX F. LIST OF UNREFEREED PUBLISHED PAPERS

F.l MUTATIONAL ANALYSIS OF PHOTOSYSTEM I Function of Phylloquinones

A paper published in Photosynthesis: Mechanisms and Effects, Vol. L 515-520. 1998. Kluwer Academic Publishers, Netherlands. Ricardo Reategui, Wade Johnson, Wu Xu, Boris Zybailov, Ilya R. Vassiliev, Gaozhong Shen, John H. Golbeck, and Parag R. Chitnis

F.2 MUTATIONS IN THE PHYLLOQUINONE BIOSYNTHETIC PATHWAY: .A. Foreign Quinone is Recruited into the A1 Binding Site after Inactivation of the menA and menB Genes in Synechocystis sp. PCC 6803

A paper published in Photosynthesis: Mechanisms and Effects, Vol. I, 647-650. 1998. Kluwer Academic Publishers, Netherlands. Boris Zybailov, Gaozhong Shen, Ilya R. Vassiliev, Donald A. Bryant. Ricardo Reategui. Wade Johnson, Wu Xu, Parag R. Chitnis, and John H. Golbeck 172

ACKNOWLEDGMENTS

The road towards a doctorate degree has been a long one filled with forks and roundabouts. I ended up in a traveling a different path than I originally set out on. but am much better for it. Dr. Parag R. Chitnis will always have my thanks for being my doctorate advisor. His willingness to take on a chemist in a biochemistry laboratory has proved to be very beneficial. I have enjoyed learning molecular biology while being nurtured to think independently and encouraged to run with those ideas and experiments. To the rest of my advisory committee, thank you. Dr. David Metzler and Dr. Dennis Johnson have been on my POS committee from the beginning and have proved to be both critical and supportive of my research. Dr. Walter Struve is more than a committee member, but a valuable collaborator. Dr. Mark Hargrove has also been helpful in data analysis and theory of my research. The members of the Chitnis lab are great teachers. Dr. Don Heck, Dr. Jun Sun, and Wu Xu were patient in teaching me not only the protocols used in molecular biology but the theory behind the techniques also. I would like to also thank Huadong Tang, Yingchun. Wang, Maria Pantelidou for being great peers. Thank you mom, Mariann Kaye, for your continued support and belief that I could accomplish this degree. Thank you dad. John Johnson, for your advice and quiet encouragement after attaining my Masters degree. My sister. Hetty Johnson, has always been in my thoughts and is part of my motivation. I would also like to thank Courtney Thomas for her immense support and patience during the final push to finish. It would have been much more difficult without you.