V x V..V. . V . .. . ..., .V .... V. . . V .V. . V. . I .. . . W. «I. w A ‘ “Wm? .va V V T . . ., j. H . ....V V . .V .. VFW a t ...h... GWOWLg. . V .x .u VVm 3....3mmfi 5.1 ”6.2.51, wngfif. H. V . V. V. V . ,V V , V. V V _ ..V.‘...1V.VV.. «1. L .m.... a.» aura .,.V..V. V .. . .. V . . .... .13 uerJV£Rr v.,.r¢..0. . .9, .3... EV? .. .1. ~ 1.7.9? Marv... . 3...? E V. ...». . as»... . 16.. I, \p. .l ... V. $.H v ufim V . V .V .. . I» v V, V . V .. V .. mum. 1.1.. fl \mfi. ... “mgumnzfi fiawipfw ..utnmwfiz : .....V.........V.. . V . . . . , V V 1 V . . V _ V . V . $_H.v.a..hknwfln:..h... 4‘ ..I. a I... ... ..Ufi ..quV v . . V «4%? ..vurn a}. V V . V . V. . V . . .V. . V . lflPflctvd. Run“ “......” ..Wwfiykufi qia,{r..me :aymar .. V. V .V . .. . .. V .V _ V. V. .V V ..V -..-.wflfifimmm Pa. . . .. a L . . . V . V . V V V . . . V V V .. $3 Law. V . . v Avluo .1, «05¢...V 01h. ‘. a, o y :I Mr, I . V. V . . .VIVVafiQ... .. V ..VVV V . . V V V. V .. V V V V V V V. VllVlvf.-!‘§-IAI§1 . V . . . V . V. .V V . V . . ., V V V V .V V, . . .V . 1.1(.V..u.....a«I,(..1$. magma... . ctrmusfiaulbfl. V . . .. V . V . . .V V . . . V V. V V 13m... C. V V . V. _ . 3...: .. .V . . ”u Knowlyoafifl ...: ...r {Iniivvfi . .. .. . . . V V . .V ‘vau.... . 2 3:. ....o ..... S. V . . c.,HV»41,VVMVHN. udfnv: --.:V ...)” !V .V 1. “vs-Hun. I .OII ,ll.'h .. V V . JMUMW 11% V V . . VV .....IVD H..ol.u.eroV!¢t.V‘¥bVHPnh‘ .l. . 3.13.1.1?! not. V. . V ..wfi....1-.v.. . . . . V . . . V V . 7...}...Svlfitl . ,. .. Av. .. V V V V V V .V . . . V 215 ”£60..qu V .21”! .. .MV‘W u . bid :1. u.nr.1.V...m:2:....bl . Ya I. 5. \ , hr. -..-hunk . ...aTfivfl. . .V .. ...»). . V . V, in V 2V V ,. V ,.V . .V ....yI :1V3cl.un.;....: _ “.....H . V V . eff}. .8 ~|o| .1 .[ vol... .IV ... 'A u V lnnl‘llt alt! l’l?¢a|.1lt.l V. 11V” . t .. . I .. Judah..- Huh} . LEI-tr 1’...) ’vlu‘.'¢.» .A . .cl DI! Vlvlll‘ 31“; p v ' t V!!! .' ... b. .0 turn} ..I l iVJ.¢.|.\IIA. ..irch' .... ‘2 vi?!- . ..v..1.u.v.l. . ..1V . H ( ..V!» TfloEDH JEJ. 1 V . (V \V n )u. .E‘. :3. b I osLlr Vin» . I! v v .5: ..r. , ...aiiluih‘. L .7... a ..lVLVu VV k tsz V; a». ‘ L: V- _ V r t: fl! .1 r L» n d H I |' lHESlS l ANESTAT UN IV lllllllllllllllll lllllllllllllllllllllllllll 3 1293 01563 39093 This is to certify that the dissertation entitled The Allosteric Regulation Sites of ADP-Glucose Pyrophosphorylase from Cyanobacterium Anabaena PCC 7120 presented by Jun Sheng has been accepted towards fulfillment of the requirements for Ph.D. degree in Biochemistry Major professor Date 8/7 /77Z / MSU is an Affirmatiw Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE N RETURN BOX to roman this checkout from your rocord. TO AVOID FINES Mum on or odor. duo duo. DATE DUE DATE DUE DATE DUE D MSU In An Afflnnotlvo Adlai/EM Opportunity Institution WM! THE ALLOSTERIC REGULATION SITES OF ADP-GLUCOSE PYROPHOSPHORYLASE FROM CYANOBACTERIUM ANABAENA PCC 7120 By Jun Sheng A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1997 ABSTRACT THE ALLOSTERIC REGULATION SITES OF ADP-GLUCOSE PYROPHOSPHORYLASE FROM CYANOBACTERIUM ANABAENA PCC 7120 By Jun Sheng ADP-glucose pyrophosphorylase is a key enzyme in the biosynthesis of glycogen in bacteria and starch in plants. This enzyme catalyzes the conversion of glucose-l-P and ATP to ADP-glucose and pyrophosphate. It is subject to allosteric regulation. ADP- glucose pyrophosphorylase from cyanobacterium Anabaena PCC 7120 is mainly activated by 3-P—glycerate and inhibited by phosphate. Previous chemical modification studies with pyridoxal-P suggested that Lys382 of Anabaena ADP-glucose pyrophosphorylase is involved in the binding of 3-P-glycerate, the allosteric activator. Site-directed mutagenesis of Lys382 of the Anabaena enzyme was performed to determine the role of this residue. Replacing Lys382 with either arginine, alanine, or glutamine produced mutant enzymes that were still activated by 3-P-glycerate but with apparent affinities 10- to 160-fold lower than that of the wild-type enzyme. In contrast, the glutamic acid mutant enzyme was inhibited by 3-P—glycerate. These mutations had lesser impact on the kinetic constants for the substrates and inhibitor, phosphate, and on the thermal stability. These results indicate that both the charge and size of the residue at position 382 influence the binding of 3-P-glycerate. Site-directed mutagenesis was also performed to obtain a K382R-K419R double mutant. The apparent affinity for 3-P- glycerate of this double-mutant enzyme was 104-fold lower than that of the wild-type enzyme, and the specificity for the activator of this mutant enzyme was altered. Treatment of Anabaena ADP-glucose pyrophosphorylase with phenylglyoxal resulted in a time- and concentration-dependent loss of enzyme activity. Phosphate, the inhibitor, protected the enzyme from inactivation most effectively, while 3-P-glycerate, fructose-1,6-P2, pyridoxal-P, and ATP plus magnesium were also good protectors. After incubation with 2 mM phenylglyoxal for 1 hour, the modified enzyme had a 10-fold lower apparent affinity for phosphate in the absence of the activator, 3-P—glycerate, than that of the wild—type enzyme. This result has implicated the involvement of an arginine residue at the allosteric sites, most probably the inhibitor-binding site. In order to identify the arginine residue, five arginine residues, which are conserved in all higher-plant and cyanobacterial enzymes but not in enteric bacterial enzymes, were individually converted to alanine by site-directed mutagenesis. The mutant enzymes were purified and the properties of these mutants were compared with the wild- type enzyme. Substitution of only Arg294 with alanine resulted in an enzyme with more than lOO-fold or 40-fold lower affinity for the inhibitor, phosphate, in the absence or presence of 3-P-glycerate, respectively. This mutation had no or lesser impact on the kinetic constants for the substrates and the activator, 3-P-glycerate. It is concluded that Arg294 is a critical residue in phosphate binding and inhibition. To my family iv ACKNOWLEDGEMENTS I would like to thank my supervisor, Dr. Jack Preiss, for his direction and valuable support. During the past five years, I have gained significant experience in biochemistry and molecular biology and learned how to conduct scientific research under Dr. Preiss' guidance. I also thank my committee members, Drs. Michael Bagdasarian, Shelagh Ferguson-Miller, Lee Kroos, and Honggao Yan, for their help through my graduate study. Dr. David G. McConnell gave me tremendous help when I came to the United States and did my first rotation in his laboratory and he also gave me spiritual support during these years. Drs. Yee-yung Charng and Chris Meyer, with whom I began my research in this laboratory, are not only my mentors but also my friends. They gave me many ideas and technical help and encouraged me whenever I had a problem. Dr. Miguel A. Ballicora, Dr. Minxian Wu, Brian Smith-White, and Yingbin Fu contributed their time in discussion with me and provided many intriguing ideas. I am also grateful to other members in the laboratory for their friendship which made me feel like I had a big family at Michigan State University. My parents and my brothers give me support through my education and without their love I would not have a chance to gain my doctoral degree. I would like to thank my newly-wedded wife, Xin, for her love and encouragement which made the final stage of my graduate study easier than otherwise it would be. I also want to thank the friends I met here during the past several years, Bing Ren, Bo Zhang, Qing Yang, Xiaoyu Wu, Jing Wang, Lei Lei, Bing Lian, Bing Li, Jie Qian, and many others. They gave me their hands whenever I needed. TABLE OF CONTENTS Page LIST OF TABLES .................................................................................... x LIST OF FIGURES ................................................................................... xi ABBREVIATIONS ................................................................................... xiii INTRODUCTION ..................................................................................... 1 CHAPTER I LITERATURE REVIEW ........................................................... 6 1. Starch and Glycogen ...................................................................... 7 2. Synthesis of Starch and Glycogen ...................................................... 8 3. The Role of ADP-Glucose Pyrophosphorylase in Biosynthesis of Starch and Glycogen .................................................................... 8 3.1. ADP-Glucose Pyrophosphorylase in Starch Synthesis .................... 9 3.2. ADP-Glucose Pyrophosphorylase in Glycogen Synthesis ............... ll 4. Regulation of the ADP-Glucose pathway ............................................. 11 4.1. Regulation of the ADP-Glucose Pathway in Starch Synthesis ......... 12 4.2. Regulation of the ADP-Glucose Pathway in Bacterial Glycogen Synthesis ............................................................. 13 5. Subunit and Primary Structure of ADP-Glucose Pyrophosphorylase ............ 17 6. Structures of the Substrate and Regulator Binding Sites of ADP-Glucose Pyrophosphorylase ....................................................... 24 6.1. The Primary Structure of the Substrate Binding Sites ................... 26 6.2. The Primary Structure of the Regulator Binding Sites .................. 30 7. Prediction of the Secondary Structure of ADP-Glucose Pyrophosphorylase. . . .39 References ...................................................................................... 44 vi CHAPTER II SITE-DIRECTED MUTAGENESIS OF LYSINE382, THE ACTIVATOR-BINDING SITE, OF ADP-GLUCOSE PYROPHOSPHORYLASE FROM ANABAENA PCC 7120 ................. 51 Abstract ......................................................................................... 52 Introduction .................................................................................... 53 Material and Methods ........................................................................ 55 Reagents ............................................................................... 55 Bacterial Strains and Media ........................................................ 55 Site-Directed Mutagenesis .......................................................... 55 Expression and Purification of the Wild-type and Mutant Enzymes .............................................................. 56 Enzyme Assay ........................................................................ 56 Kinetic Characterization ............................................................ 59 Protein Assay ......................................................................... 60 Protein Electrophoresis and Immunoblotting .................................... 60 Molecular Mass Determination .................................................... 61 Thermal Stability ..................................................................... 61 Reductive Phosphopyridoxylation ................................................. 61 Results .......................................................................................... 62 Expression and Purification of Lys382 Mutant Enzymes ....................... 62 Kinetic Characterization of Lys382 Mutant Enzymes ........................... 62 Kinetic Characterization of the K382R-K419R Enzyme ...................... 66 Effector Specificity .................................................................. 66 Molecular Masses of the Wild-Type and Mutant Enzymes .................. 67 Thermal Stability of Lys382 Mutant Enzymes ................................... 71 Reductive Phosphopyridoxylation of the K382R-K419R Mutant Enzyme ................................................. 71 Discussion ...................................................................................... 72 References ...................................................................................... 76 CHAPTER III INVOLVEMENT OF ARGININE AT THE ALLOSTERIC SITES OF ADP-GLUCOSE PYROPHOSPHORYLASE FROM ANABAENA PCC 7120 .......................................................... 78 Abstract ......................................................................................... 79 vii Introduction .................................................................................... 80 Material and Methods ........................................................................ 82 Reagents ............................................................................... 82 Expression and Purification of ADP-glucose Pyrophosphorylase ................................................................. 82 Enzyme Assay ........................................................................ 83 Kinetic Characterization ............................................................ 84 Modification of ADP-glucose Pyrophosphorylase with Phenylglyoxal ................................................................ 84 Incorporation of [“C]phenylglyoxal to ADP-glucose Pyrophosphorylase ................................................ 84 Results .......................................................................................... 86 Inactivation of ADP-glucose Pyrophosphorylase by Phenylglyoxal ................................................................... 86 Protection of Enzyme Activity by Substrates and Allosteric Effectors ........................................................... 86 Effect of Modification with Phenylglyoxal on Enzyme Kinetics ................................................................... 86 Stoichiometry of Modification ..................................................... 94 Discussion ...................................................................................... 99 References .................................................................................... 101 CHAPTER IV ARGININE294 IS ESSENTIAL FOR THE INHIBITION OF ANABAENA PCC 7120 ADP-GLUCOSE PYROPHOSPHORYLASE BY PHOSPHATE .............................. 103 Abstract ....................................................................................... 104 Introduction ................................................................................... 105 MATERIALS AND METHODS ......................................................... 107 Reagents .............................................................................. 107 Bacterial Strains and Media ....................................................... 107 Site-Directed Mutagenesis ......................................................... 107 Expression and Purification of the Mutant Enzymes ......................... 109 Enzyme Assay ....................................................................... 109 Kinetic Characterization ........................................................... 110 Protein Assay ........................................................................ 1 10 Protein Electrophoresis and Immunoblotting .................................. 110 viii Thermal Stability ................................................................... 111 Results ......................................................................................... 112 Expression of the Mutant Enzymes ............................................. 112 Screen for Alanine-Substituted Mutant Enzymes with altered Pi Inhibition ........................................................ 112 Purification of the Mutant Enzymes ............................................. 112 Kinetic Characterization of the R294A Enzyme .............................. 115 Kinetic Characterization of the R66A, the R105A, and the R385A Enzymes ......................................... 121 Thermal Stability of the Wild-type and the Mutant Enzymes .............. 121 Discussion ..................................................................................... 122 References .................................................................................... 125 CHAPTER V SUMMARY AND PERSPECTIVES .......................................... 127 l. Lys382 of the cyanobacterium Anabaena ADP-glucose pyrophosphorylase is involved in the binding of the activator, 3PGA ................................................ 128 2. An arginine residue(s) is involved in the allosteric regulation of Anabaena ADP-glucose pyrophosphorylase ....................................................... 129 3. Arg294 of the Anabaena ADP-glucose pyrophosphorylase is important for the inhibition by Pi ............................................................................ 130 4. Future Research .......................................................................... 13 1 References .................................................................................... 133 ix LIST OF TABLES Page Chapter I 1. Activator specificity of ADP-glucose pyrophosphorylases from various bacterial species ............................................................ 15 Chapter II 1. Kinetic Parameters for Activation of the Wild-Type and Mutant ADP-Glucose Pyrophosphorylases ........................................ 63 2. Kinetic Parameters of the Wild-Type and Mutant ADP-Glucose Pyrophosphorylases ....................................................... 65 3. Specificity of Allosteric Effectors of the Wild-Type and Mutant ADP-Glucose Pyrophosphorylases ........................................ 68 Chapter III 1. Effect of substrates and allosteric effectors on modification of Anabaena ADP-glucose pyrophosphorylase by phenylglyoxal .................. 93 Chapter IV 1. Conserved arginine residues in Anabaena ADP-glucose pyrophosphorylase and theoligonecleotides used for conversion to alanine ...................................................................... 108 2. Kinetic parameters of the Anabaena wild-type, R66A, R105A, R294A, and R385A enzymes ................................................. 116 LIST OF FIGURES Chapter I 1. Alignment of the primary structure of ADP-glucose perphosphorylases ............................................... 2. Amino acid sequence of the phosphopyridoxylated peptide, which was protected by ADP—glucose plus Mg“, of E. coli ADP-glucose pyrophosphorylase and the alignment with other known sequences .............................. 3. Amino acid sequence of the major azido-ADP-[l‘C] glucose incorporated peptide of E. coli ADP-glucose pyrophosphorylase and the alignment with other known sequences .............................. 4. Amino acid sequence of the phosphopyridoxylated peptide, which was protected by fructose-1,6-P2, of E. coli ADP-glucose pyrophosphorylase and the alignment with other known sequences ........................................ 5. Amino acid sequence of the phosphopyridoxylated peptides of spinach leaf and Anabaena ADP-glucose pyrophosphorylases and the alignment with other known sequences of higher-plant and cyanobacterial enzymes ...................................................... 6. Profile Neural Network (PHD) prediction of the secondary structure of ADP-glucose pyrophosphorylase ................................. Chapter II 1. Nucleotide sequence and encoded protein sequence of the Anabaena ADP-glucose pyrophosphorylase gene in the region of Lys382 and the synthetic oligonucleotides used for site-directed mutagenesis at position 382 ........................... Page ......... 19 ......... 28 ......... 31 ......... 33 ......... 36 ......... 41 ......... 57 2. Activation of the wild-type and K382R-K419R enzymes by 3PGA, FBP, and AMP ..................................................... 69 Chapter 111 1. Inhibition of Anabaena ADP-glucose pyrophosphorylase by phenylglyoxal .................................................... 87 2. Kinetics of inhibition of ADP-glucose pyrophosphorylase activity by phenylglyoxal (enzyme activity was measured under activated conditions) ................................................................ 89 3. Kinetics of inhibition of ADP-glucose pyrophosphorylase activity by phenylglyoxal (enzyme activity was measured under unactivated conditions) ............................................................. 91 4. Inhibition of ADP-glucose pyrophosphorylase by Pi ................................ 95 5. Incorporation of [”C]phenylglyoxal into ADP-glucose pyrophosphorylase ......................................................... 97 Chapter IV 1. Pi inhibition of the wild-type, the R66A, and the R294A enzymes in the crude extract ......................................... 113 2. The effect of Pi on the enzyme activity of the wild-type and the R294A enzymes ................................................. 117 3. The effect of Pi and KC] on the enzyme activity of the wild-type and the R294A enzyme .............................................. 119 xii ABBREVIATIONS 3PGA, 3-P—glycerate cpm, counts per minute EDTA, ethylenediamine tetraacetic acid FBP, fructose-1,6-P2 FPLC , fast performance liquid chromatography kb, kilo-base pair(s) kDa, kilo-dalton LB medium, Luria-Bertani medium PAGE, polyacrylamide gel electrophoresis Pi, orthophosphate PLP, pyridoxal—P PPi, pyrophosphate SDS, sodium dodecyl sulfate xiii INTRODUCTION Starch is the major reserve carbohydrate of algae and higher plants, deposited as a normal consequence of active photosynthesis. Glycogen is an energy reserve compound in many bacteria, accumulated particularly when growth is limited and an excess carbon source is available. It is believed that the biosynthesis of starch in algae and higher plants and of bacterial glycogen occurs predominantly via the ADP-glucose pathway (Preiss & Romeo, 1989; Preiss, 1991; Preiss, 1996; Preiss & Sivak, 1996). In this pathway, ADP-glucose is synthesized from ATP and glucose-l-P in a reaction catalyzed by ADP-glucose pyrophosphorylase (ATPza-D-glucose-l-P adenylyltransferase, EC. 2.7.7.27). The glucosyl moiety of the sugar nucleotide is then transferred to a maltodextrin, glycogen or starch in an a-1,4-glucosidic linkage by glycogen synthase or starch synthase (EC. 2.4.1.21). Branching enzyme (BE; EC. 2.4.1.18) catalyzes the synthesis of the a-l ,6-glucosidic linkages found in amylopectin and in glycogen. ADP-glucose pyrophosphorylase is known as an allosterically regulated enzyme. The enzyme from higher plants is mainly activated by 3-P—glycerate (3PGA) and inhibited by phosphate (Pi) (Ghosh & Preiss, 1965; Sanwal et a1. , 1968), whereas the enzyme from enteric bacteria is activated by fructose-1,6-P2 (FBP) and inhibited by AMP (Preiss et a1. , 1966). The bacterial enzyme is homotetrarneric in structure (I-Iaugen et a1. , 1976), while 2 the higher-plant enzyme is more complex, being heterotetrameric with two different subunits (Preiss, 1991). Cyanobacteria are prokaryotic organisms that have the chloroplast type of photosynthesis (Aitken, 1988). However, cyanobacteria synthesize glycogen as the major carbon and energy reserve (Shively, 1988), in a similar manner to that in other bacteria. The cyanobacterial ADP-glucose pyrophosphorylase is homotetrameric similar to the enteric bacterial enzyme, but is regulated by 3PGA and Pi like the higher-plant enzyme (Levi & Preiss, 1976; Iglesias et al., 1991). To understand the structure—function relationships of the allosteric sites of the cyanobacterial ADP-glucose pyrophosphorylase would be of great interest. Previous studies have shown that the activator-binding site of the E. coli enzyme, Lys”, is close to the N-terminus (Parsons & Preiss, 1978; Gardiol & Preiss, 1990), while that of the spinach leaf enzyme is near the C-terminus (Morell et al., 1988; Ball & Preiss, 1994). Chemical modification and site-directed mutagenesis studies on the cyanobacterial “9 is involved in ADP-glucose pyrophosphorylase from Anabaena have shown that Lys binding of 3PGA (Charng et al., 1994). Chemical modification studies of the K419R enzyme with pyridoxal-P (PLP), where residue 419 is no longer reactive, have shown that the phosphopyridoxylated K419R enzyme is less dependent on the presence of 3PGA for activation and less sensitive to Pi inhibition. Lys382 of the mutant enzyme was found to be labeled by PLP. Kinetic and protection studies suggested that Lys382 was also involved in binding of the allosteric activator (Charng et al., 1994). Site-directed mutagenesis of Lys382 of Anabaena ADP- glucose pyrophosphorylase 3 was performed to determine the role of this residue. Lys382 was replaced with either arginine, alanine, glutamine, or glutamic acid. The mutant enzymes were purified and characterized by kinetic studies and compared with the wild-type enzyme. The results indicate that Lys382 is involved in the binding of the activator, 3PGA. This work was published in Biochemistry (1996) 9;, 3115-3121. Chapter H of this dissertation is derived from this publication. Previous studies on E. coli, synechocystis, and spinach leaf enzymes using an arginine-specific reagent, phenylglyoxal, suggested that there is also an essential arginine residue involved in the allosteric activation and inhibition (Carlson & Preiss, 1982; Iglesias et al., 1992; Ball & Preiss, 1992). Treatment of Anabaena ADP-glucose pyrophosphorylase with phenylglyoxal resulted in a time- and concentration-dependent loss of enzyme activity. Pi, the inhibitor, protected the enzyme from inactivation most effectively, while 3PGA, FBP, PLP, and ATP plus magnesium were also good protectors. The modified enzyme had lower apparent affinities for Pi and 3PGA than those of the wild-type enzyme. Thus, phenylglyoxal appears to interfere with the allosteric regulation, as well as the catalysis, of Anabaena ADP-glucose pyrophosphorylase. This work is presented as Chapter III of this dissertation. In order to identify the arginine residue implicated by chemical modification (Chapter III), five conserved arginine residues were individually converted to alanine by site-directed mutagenesis. The mutant enzymes were purified and the properties of these mutants were compared with the wild-type enzyme. Substitution of arginine294 with alanine resulted in an enzyme with more than 100-fold or 40-fold lower affinity for the inhibitor, 4 Pi, in the absence or presence of 3PGA, respectively. This mutation had no or lesser impact on the kinetic constants for the substrates and the activator, 3PGA. These results suggest that Arg294 is involved in the binding of the inhibitor, P,. This work is included in Chapter IV of this dissertation. REFERENCES Aitken, A. (1988) Methods Enzymol. 167, 145-154. Ball, K., & Preiss, J. (1992) J. Protein Chem. 11, 231-238. Ball, K., & Preiss, J. (1994) J. Biol. Chem. 269, 24706-24711. Carlson, C.A., & Preiss, J. (1982) Biochemistry 21, 1929-1934. Charng, Y., Iglesias, A. A., & Preiss, J. (1994)]. Biol. Chem. 269, 24107-24113. Gardiol, A., & Preiss, J. (1990) Arch. Biochem. Biophys. 280, 175-180. Ghosh, H. P., & Preiss, J. (1965) J. Biol. Chem. 240, 960-961. Haugen, T. H., Ishaque, A., & Preiss, J. (1976) J. Biol. Chem. 251, 7880-7885. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1991) Plant Physiol. 97, 1187-1195. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1992) J. Protein Chem. 11, 119-128. Levi, C., & Preiss, J. (1976) Plant Physiol. 58, 753-756. Morel], M., Bloom, M., & Preiss, J. (1988) J. Biol. Chem. 263, 633-637. Parsons, T., & Preiss, J. (1978) J. Biol. Chem. 253, 6197-6202. Preiss, J ., Shen, L., Greenberg, E., & Gentner, N. (1966) Biochemistry 5, 1833-1845. Preiss, J., & Romeo, T. (1989) Adv. Microb. Physiol. 30, 183-238. Preiss, J. (1991) in Plant Molecular and Cell Biology (Mifflin, B., ed.) Vol. 7, pp 59-114, Oxford University, Oxford. Preiss, J. (1996) Biotech. Annu. Rev. 2, 259-279. Preiss, J, & Sivak, M. (1996) in Photoassimilate Distribution in Plants and Crops (Zamski, E., Schaffer, A. A., ed.) pp 63-96, Marcel Dekker, Inc., New York. Sanwal, G. G., Greenberg, E., Hardie, J., Cameron, E. C., & Preiss, J. (1968) Plant Physiol. 43, 417-427. Shively, J. M. (1988) Methods Enzymol. 167, 195-203. 5 CHAPTER I LITERATURE REVIEW LITERATURE REVIEW 1 . Starch and Glycogen: Starch, a polysaccharide, is the major carbon and energy reserve of most green plants. It is the end-product of carbon fixation by photosynthesis and is a water-insoluble granule present in every type of tissue: leaves, fruit, pollen grains, roots, shoots, and stems. The size and shape are varied depending on the plant source and the stage of development. The starch granule contains two types of polysaccharides, amylose and amylopectin. Amylose is mainly found as a linear chain of a-1,4 linked a-D- glucopyranosyl residues of about 1000 units long, with occasional secondary chains attached by OH ,6-glucosidic linkage. In contrast, amylopectin is more highly branched with an average of one branch point, a-l,6—glucosidic linkage, every 20 to 26 units (Manners, 1985; Morrison & Karkalas, 1990). Because of the use of starch in both food and non-food industries, there is a desire to increase the content of starch in some plants as well as manipulate the quality of starch. Glycogen has been found as one of the major carbon and energy reserves in many bacteria as well as in eukaryotic organisms. It is a branched polysaccharide that contains about 90 % of its glucose in a-1,4-glucosidic linkages and the rest in a-1,6-linkages. Glycogen usually accumulates in the bacterial cell when there is excess of carbon in the media and when the nutrition supply for growth is limited (Preiss & Romeo, 1989). It is considered as a storage compound which can be degraded to an energy form that is utilized by the cell, enabling it to survive better under certain conditions (Preiss, 1984). The 8 observations that glycogen-containing bacterial cells survived longer than glycogen-less bacterial cells under starvation conditions (Strange etal., 1961; Strange, 1968; Van Houte & Jansen, 1970), and the utilization of glycogen during spore formation (Strasdine, 1968, 1972; Mackey & Morris, 1971; Slock & Stahly, 1974; Brana etal., 1980), are consistent with the role of glycogen as an energy reserve. 2. Synthesis of Starch and Glycogen: The biosynthesis of starch in algae and higher plants and of glycogen in bacteria share common reactions. The ADP-glucose pathway is the predominant pathway (Preiss & Romeo, 1989; Preiss, 1991, 1996; Preiss & Sivak, 1996) which is shown below. 1. ATP + oc-glucose-l-P ~ ADP-glucose + pyrophosphate 2. ADP-glucose + a-1,4-glucan ~ ADP + a-1,4-glucosyl-a-1,4-glucan 3. Elongated a-1,4-glucan chain - Branched a-1,6- a-1,4-glucan First, the sugar nucleotide, ADP-glucose, is synthesized from ATP and glucose-l-P in a reaction catalyzed by ADP-glucose pyrophosphorylase (ATPza-D-glucose-l-P adenylyltransferase, EC. 2.7.7.27). In the next reaction, the glucosyl moiety of the sugar nucleotide is transferred to a maltodextrin, glycogen or starch in an a-1,4-glucosidic linkage by glycogen synthase or starchsynthase (BC. 2.4. 1.21). In the third reaction of the pathway, branching enzyme (BE; EC. 2.4.1.18) catalyzes the synthesis of the a-1,6- glucosidic linkages found in amylopectin and in glycogen. 9 3. The Role of ADP-Glucose Pyrophosphorylase in Biosynthesis of Starch and Glycogen: ADP-glucose pyrophosphorylase catalyzes the conversion of glucose-l-P and ATP to ADP-glucose and pyrophosphate (PPi), the first step in the biosynthesis of starch in plants (Preiss, 1991, 1996; Preiss & Sivak, 1996) and glycogen in bacteria (Preiss & Romeo, 1989, Preiss, 1996). 3.1. ADP-Glucose Pyrophosphorylase in Starch Synthesis: In 1962, Espada first described the enzymatic synthesis of ADP-glucose by ADP- glucose pyrophosphorylase in soybean. ADP-glucose pyrophosphorylase was later found distributed in chloroplasts and amyloplasts in which starch synthesis occurs (Okita et a1. , 1979; ap Rees et al., 1984; Lin etal., 1988). The activity of this enzyme is sufficient to account for the rates of starch synthesis according to the studies on various plant tissues (Ozbun et al., 1973; Heldt et al., 1977; Okita etal., 1979; ap Rees et al., 1984; Edwards et al., 1988). It is believed that ADP-glucose pathway is the main pathway leading to starch synthesis. Data obtained from a number of biochemical and genetic studies support this view. Mutants of maize endosperm having deficient levels of ADP-glucose pyrophosphorylase (5-10 % of wild-type) also have deficient level of starch content (25-30 % of wild-type) (Tsai & Nelson, 1966; Dickinson & Preiss, 1969). Smith et a1. (1989) have shown that one pea line with a mutation at the rb locus, which controls the level of ADP-glucose pyrophosphorylase activity in developing pea embryos, contained 3-5 % of the ADP-glucose pyrophosphorylase activity and 38-72 % of the starch found in the pea 10 line having the dominant genes. Lin and colleagues (1989) first showed that starch biosynthesis in the chloroplast is entirely dependent on the pathway involving ADP- glucose pyrophosphorylase. A mutant of Arabidopsis thaliana containing less than 2 % of the leaf starch content found in the normal strain had only 2 % of the normal leaf ADP- glucose perphosphorylase activity (Lin et a1. , 1988a). A chimeric gene encoding antisense ribonucleic acid for the ADP-glucose pyrophosphorylase was expressed in potato tuber, causing a decrease in enzymatic activity to 2-5 % of the wild-type level and a commensurate decrease in starch content (Mfiller—Rober et al. , 1992). Stark et a1. (1992) provided persuasive evidence for the role of ADP-glucose pyrophosphorylase in starch synthesis. They increased ADP-glucose pyrophosphorylase activity by transformation of the potato plant with a mutant Escherichia coli (E. coli) ADP-glucose pyrophosphorylase gene insensitive to the regulators that affect the plant enzyme and found that the starch content of potato tubers was increased by 30-60 %. Other data demonstrating a direct relationship between activity of the ADP-glucose pyrophosphorylase and starch accumulation in several plant species (Preiss & Levi, 1980; Preiss, 1988; Okita, 1992) also support the hypothesis that the first reaction catalyzed by ADP-glucose pyrophosphorylase is the major regulatory step in starch synthesis. There has been no relationship noted between starch synthesis and phosphorylase or UDP-glucose pyrophosphorylase activity. Many studies have shown that the kinetic properties of the enzymes in the ADP-glucose pathway (80.5 and Vm values) are consistent with the known concentrations of substrates and effector metabolites present in plant cells. In contrast, for the UDP—glucose-specific starch synthases and plant phosphorylases, the 11 high $0.5 values for the respective substrates (U DP-glucose and glucose-l-P) compared to the actual cellular levels, strongly argues that these two enzymes do not play a significant role in starch biosynthesis. 3.2. ADP-Glucose Pyrophosphorylase in Glycogen Synthesis: In mammalian tissues and eukaryotic micro-organisms, UDP—glucose was considered to be the glycosyl donor for glycogen synthesis (Leloir et al., 1959; Stalmans & Hers, 1973; Krebs & Preiss, 1975). In 1964, Sigal et al., reported that many UDP- glucose pyrophosphorylase—deficient mutants of E. coli were. still able to produce normal amounts of glycogen. It was shown that all bacterial glycogen syntheses that have been tested utilize ADP-glucose as a substrate much better than UDP-glucose (Greenberg & Preiss, 1964). It was also shown that extracts of several bacteria contained both an ADP- glucose pyrophosphorylase (Shen & Preiss, 1964) and an ADP-glucose-specific glycogen synthase (Greenberg & Preiss, 1964). Mutants of E. coli and Salmonella typhimurium (S. typhimurium), affected in ADP-glucose pyrophosphorylase and/or glycogen synthase, are either glycogen-deficient or glycogen-hyperproducing (Preiss, 1984; Preiss & Romeo, 1989). These studies suggested that bacterial glycogen synthesis utilizes ADP-glucose, instead of UDP-glucose, as the glycosyl donor. ADP-glucose is produced from glucose-1- P and ATP in the reaction catalyzed by bacterial ADP-glucose pyrophosphorylase. 4. Regulation of the ADP-Glucose pathway: Regulation of the ADP-glucose pathway is believed to occur at the ADP-glucose synthetic step in both starch synthesis (Preiss, 1984; Kruger, 1990; Preiss & Sivak, 1996) 12 and bacterial glycogen synthesis (Preiss & Romeo, 1989; Preiss, 1991 , 1996). The activity of ADP-glucose pyrophosphorylase is allosterically regulated by metabolites. This is in agreement with the concept that regulation of a biosynthetic pathway occurs at its first committed step. 4.1. Regulation of the ADP-Glucose Pathway in Starch Synthesis: There is much evidence suggesting the major activator of ADP-glucose pyrophosphorylase is 3PGA and the major inhibitor is Pi. It was first reported by Ghosh and Preiss (1965) that the spinach leaf enzyme is activated by 3PGA and inhibited by Pi. This was also seen for ADP-glucose pyrophosphorylases from other higher plants (Sanwal et al., 1968), green algae (Sanwal & Preiss, 1967; Ball et al., 1991), and blue-green bacteria (Levi & Preiss, 1976; Iglesias et al., 1991). ADP-glucose pyrophosphorylases from other species (Preiss, J. , 1982) had been shown to be activated by 3PGA. 3PGA also caused the enzyme to have higher affinity for the substrates, ATP and glucose-l-P, and lower sensitivity to the inhibitor, Pi. Both in vitro and in vivo experiments suggested that the activation of ADP-glucose pyrophosphorylase by 3PGA and inhibition by Pi are physiologically important. Many of these experiments (Preiss 1982, 1988, 1991; Preiss & Levi, 1980) showed a direct correlation between the concentration of 3PGA and starch production, and an inverse one between the concentration of Pi and starch content. Pettersson and Pettersson (1989) applied modern control theory (Kacser & Burns, 1973; Kacser, 1987) to develop a kinetic model to determine the extent that metabolites, which affect in vitro leaf ADP-glucose pyrophosphorylase activity, controlled the rate of 13 starch production under conditions of light and CO2 saturation. They reached the conclusion that 3PGA and Pi play an important role in the regulation of starch synthesis. In another Kacser-Burns analysis, it was shown that in Arabidopsis thaliana ADP-glucose pyrophosphorylase is a major site of regulation for starch synthesis (Neuhaus & Stitt, 1990) and regulation of ADP-glucose pyrophosphorylase by 3PGA concentration and 3PGA/P, ratio is an important determinant of the rate of starch synthesis in vivo (Neuhaus et a1. , 1989). i In 1991, Ball et a1. isolated a starch-deficient mutant of Chlamydomonas reinhardii in which ADP-glucose pyrophosphorylase could not be effectively activated by 3PGA. This work shows that activation of ADP-glucose pyrophosphorylase by 3PGA is important physiologically. 4.2. Regulation of the ADP-Glucose Pathway in Bacterial Glycogen Synthesis: Regulation of the ADP-glucose pathway occurs at the first step, the ADP-glucose synthetic step, in bacterial glycogen synthesis (Preiss & Romeo, 1989; Preiss, 1991, 1996) as in starch synthesis. The activity of bacterial ADP-glucose pyrophosphorylase is also allosterically regulated by metabolites (Preiss, 1984; Preiss & Romeo, 1989). For most bacterial ADP-glucose pyrophosphorylases tested, glycolytic intermediates, FBP, fructose-6-P, pyruvate, or 3PGA are the major activators while AMP, ADP, or P, are the major inhibitors (Preiss & Romeo, 1989; Preiss, 1996). The activators increase the apparent affinity of the enzyme for its substrates, ATP and glucose-l-P. In addition, the activators may reverse the inhibition of the enzyme by AMP, ADP, or P,. The activators are required for optimal activity, and may be considered as 14 indicators of carbon in excess of energy needs (Preiss, 1984). The bacterial ADP-glucose pyrophosphorylase is more diverse than the plant enzyme with respect to the specificity for activators. The activator specificity of ADP-glucose pyrophosphorylases from various bacterial species has been classified into seven groups (Preiss, 1984; Preiss & Romeo, 1989). It was postulated that the activator specificity seen in Table 1 is correlated with the type of carbon degradation or assimilation pathway in the organisms (Preiss, 1984; Preiss & Romeo, 1989). For example, when bacteria, such as E. coli and other enterics, obtain their energy mainly through glycolysis, ADP-glucose pyrophosphorylases are usually activated by FBP (Preiss, 1984, 1996; Preiss & Romeo, 1989). Regulation of the glycolytic pathway is at the phosphofructokinase step, the synthesis of FBP. For those organisms with Entner- Doudoroff pathway as their predominant pathway, the major activators for ADP-glucose pyrophosphorylases are fructose-6-P and pyruvate (Preiss, 1969, 1984, 1996; Preiss & Romeo, 1989). These organisms have no phosphofructokinase but phosphoglucoisomerase. For bacteria such as Rhodospirillum rubrum, pyruvate is the only activator of the ADP- glucose pyrophosphorylases (Furlong & Preiss, 1969). These bacteria grow anaerobically on pyruvate or lactate or on C02, and pyruvate is a product of CO2 fixation under anaerobic photosynthesis. For the photosynthetic cyanobacterial ADP-glucose pyrophosphorylases, 3PGA, the initial CO2 fixation product of oxygenic photosynthesis, is the major activator (Preiss & Levi, 1980; Preiss, 1988, 1989, 1996). Thus, ADP- glucose pyrophosphorylases, which are involved in synthesizing one of the final products of photosynthesis, glycogen, are stimulated by the first product of CO2 fixation, 3PGA. 15 Table 1: Activator specificity of ADP-glucose pyrophosphorylases from various bacterial species“ bacteria activator(s) Clostridium pasteurianum, Enterobacter hafniae, Serratia spp. (Ser. liquifaciens, Ser. marcescens) Citrobacter freundii, Edwardsiella tarda, Enterobacter aerogenes, Enterobacter cloacae, Escherichia aurescens, Escherichia coli, Klebsiella pneumoniae, Salmonella enteriditis, Salmonella typhimurium, Shigella dysenteriae Rhodospirillum spp. (Rh. fitlvum, Rh. molischianum, Rh. photometricum, Rh. rubrum, Rh. tenue), Rhodocyclus purpureus Anabaena 7120, Aphanocapsa 6308, synechococcus 6301 , Synechocystis 6803 Aeromonas hydrophila, Micrococcus luteus, Mycobacterium smegmatis, Rhodobacter viridis Agrobacterium tumefaciens, Arthrobacter viscosus, Chlorobium limicola, Chromatium vinosum, Rhodobacter spp. (R. acidophila, R. blastica, R. capsulata, R. palustris), Rhodomicrobium vannielii Rhodobacter spp. (R. gelatinosa, R. globiformis, R. sphaeroides) none fi'uctose-l ,6-P2 pyruvate 3-P-glycerate W'l ,6'P2 fructose-6-P pyruvate fructose—6-P fructose-l ,6-P2 pyruvate fructose-6-P “ This table is based on Table 2 in a review (Preiss & Romeo, 1989). 16 ADP-glucose pyrophosphorylase from Rhodobacter spheroides is effectively activated by either FBP, fructose-6-P, or pyruvate (Y ung & Preiss, 1982). This is consistent with the view that there seems to be a correlation between the nature of the activator and the major assimilation pathway, since Rhodobacter spheroides can metabolize glucose by both glycolytic pathway and Entner-Doudoroff pathway and can assimilate CO2 during anaerobic photosynthesis. It is also shown in Table 1 that various groups may have an overlapping activator specificity. This suggests that the structures of the activator sites for different groups are related. It would be of interest to determine the natures of the activator sites from different classes and compare them with each other. AMP, ADP, and Pi are the major inhibitors for many bacterial ADP-glucose pyrophosphorylases (Preiss, 1973, 1978; Preiss & Walsh, 1981). For the bacteria which undergo glycolysis or the Entner-Doudoroff pathway, ADP-glucose pyrophosphorylases are mainly inhibited by AMP and ADP or AMP alone. For cyanobacteria, the enzymes are mainly inhibited by P,. Increasing concentrations of activator either reverse or prevent the inhibition caused by the inhibitors (Preiss, 1969, 1973, 1978; Preiss & Walsh, 1981). The activators may be considered as signals of carbon excess in the cell while the inhibitors as indications of the low energy charge. This suggests that optimal ADP-glucose and glycogen synthesis only occur with excess carbon and at high energy charge (Preiss, 1984). It has been shown that allosteric regulation of bacterial ADP-glucose pyrophosphorylase is important for regulation of glycogen synthesis in vivo. The ability 17 to accumulate glycogen in the mutant cells of E. coli (Cattaneo et a1. , 1969; Govons et al. , 1969, 1973; Preiss et al., 1971, 1975, 1976) and S. typhimurium LT-2 (Steiner & Preiss, 1977) can be correlated to the altered regulatory properties of the mutant ADP-glucose pyrophosphorylases. The mutants containing ADP-glucose pyrophosphorylases with higher affinity for the activator, FBP, and/or lower affinity for the inhibitor, AMP, accumulate more glycogen than the wild-type strain, while those containing ADP-glucose pyrophosphorylases with lower affinity for the activator accumulate less glycogen (Preiss & Romeo, 1989). 5. Subunit and Primary Structure of ADP-Glucose Pyrophosphorylase: ADP-glucose pyrophosphorylase from all sources is tetrameric in structure. For all the bacterial enzymes studied to date, the enzyme is a homotetramer, with a molecular mass of about 50 to 55 kDa for each subunit (Preiss, 1984). The higher-plant enzyme is more complex, being heterotetrameric with two related but different subunits having molecular masses in the 50 to 60 kDa range (Preiss, 1996). The most studied higher-plant ADP-glucose pyrophosphorylase with respect to structural properties is the spinach leaf enzyme (Morell et al., 1987; Preiss et al., 1987a,b). This enzyme is composed of two different subunits with molecular masses of 51 and 54 kDa. The subunits can also be distinguished by amino acid composition, N- terminal amino acid sequences, tryptic peptide mapping on high-performance liquid chromatography (HPLC), and antigenic properties. The 51 kDa and 54 kDa subunits are usually referred to as the small and large subunits, respectively. In irnmunoblotting experiments, the polyclonal antibody raised against the small subunit react strongly with 18 the small subunit but weakly with the large subunit, and vice versa. These studies suggest that the small and large subunits of the spinach leaf enzyme are structurally different (Morell etal., 1987) and coded by two different genes. ADP-glucose pyrophosphorylases from Arabidopsis thaliana leaf (Lin et al., 1988a,b), maize endosperm (Preiss et al., 1990), potato tuber (Okita et al., 1990), and Chlamydomonas reinhardtii (Iglesias et al., 1994) have been studied and they have been shown to be composed of two different subunits. Cyanobacteria are prokaryotic organisms that have metabolic properties similar to chloroplasts of higher plants (Aitken, 1988). Mereschowsky originally proposed that the chloroplast of eukaryotic cells originated by endosymbiosis of a cyanobacterium (Mereschowsky, 1905). However, the former organisms synthesize glycogen as the major carbohydrate reserve (Shively, 1988), in a similar manner to what is observed in other bacteria. The cyanobacterial ADP-glucose pyrophosphorylase is homotetrameric similar to the E. coli enzyme, but is regulated by 3PGA and P, like the higher-plant enzyme (Levi & Preiss, 1976; Iglesias et al., 1991; Charng et al., 1992). The structural genes of ADP-glucose pyrophosphorylases (glgC genes) of E. coli and S. typhimurium have been cloned (Okita, etal., 1981; Leung & Preiss, 1987a) and sequenced (Baecker et a1. , 1983; Leung & Preiss, 1987b). Amino acid sequences of these two glgC genes have been deduced from the nucleotide sequences and compared (Figure 1). The total amino acid compositions of the deduced amino acid sequences are in accordance with those determined by acid hydrolysis (Haugen et al., 1976; Lehmann & Preiss, 1980) and the deduced amino acid sequences are identical to the partial sequences 19 Figure 1: Alignment of the primary structure of ADP-glucose pyrophosphorylase from spinach leaf large subunit (31-54) and small subunit (sl-51), wheat leaf (wl), maize sh-2 and bt-2 loci, wheat endosperm (we7 and we3), potato tuber large subunit (pot-l) and small subunit (pot-s), Anabaena (ana), Synechocystis (syn), rice seed (rice), Arabidopsis thaliana (arabid), E. coli (ecoli) and Salmonella typhimurium (s.t.). The number indicated is corresponding to the sequence of the Anabaena enzyme. .n unshah 35E. 4E8 4% M2. a. ”Ma. 30% KOHfig§0fl09flb0§00§h§4§ (“490493. .g004ae. "B. a. E. 3040HNO>E0 09 09H>0§0QZHUZQA4hQHHEO§4§g 4n.—4. 900.5». .Eh #8252039. :H2m094mmg mfigégggHa§g4§flh ...» .:fiZfl magi. .0022049 agafiggggHflH§g4§A 5900.5»: 05mm 000E039. .ghagafigggfianHgg4gh 4n4900h.».. 3:0004EOE. .gfiagafigégngflHgg4gh 404900;. .Efim004g0g. .sgflggafiggngagg4gflh 5900.3». .Gggfl0. h. ..00h 2. NggfiggfiéHa§g4§ 4949009.». .Egflh.h. .005. 2. 93§E09§H HEZHUZNEHGHEE4 E (049005;. a. >0 §§a§§§a§H3§4 4949 “mg 40490.95». 04593094405589. ha. H0. ENNH :ggEEHKBH mgHflHgHAbagng 094900;. 04a§09§>00fi4. hz. H.0 0§§90E0§00§Ha§4fi9§§ 34900.0». .0 ....... §.§ME.H4. . . . .EODZHOEHQ.H400H§4Q. . . . . . . . Egg . E . A30§4§0>2 ENQUSH: . >04 . A30§§0§ N§§H Hgaéma S§0§§ ”Egan HOE4h0002> ......... 949040003" Hangman—Mamas» . . . HEEHHéfigmz . . 94904083” H§§m§9§m§a§g§ 9&904000flHH4AE ”Egan—"g §0§§§§§0§E9§ Hugging . 6003059904439 . 04Eg§m00m40m04940 . fi0§§flmm§ . . . 300090§HH40>§§0090§§9QH0009§§.KHOMKO0HOEHEHODOHUWNHON . . . . Egg. . . . . . .mangHEngE ....... _ éséiiiS-{i‘ééiéi‘i ««««« HHHHH vvvvv A . unoov .n Gun—bah {In} I o o 5: 4.2Idd. Hag 00000000000000 E I I o I gag“ 43:3 00000000000000 Ema hangar ............ on» . . §¢o>§oumagmgmohn . . 43:3]. . ASHE ESE ............ 8N . . fi‘GNrfiSNB‘UWEMUhQ . . . plll 43:3 0 ASHE 533E» ............ on». . 82053395 .3393 . . . ii .. > Saga. .ASHE Egan—Hz» ............ c8— . . Higgmagnnbumohn . . > .333. . 39.3 5.5585 ............ on» . . fidgggaomgmumn . . tag . . ASHES ESHEH ............ Ge . . .HH4OR . EQEQWHHHNUhQ . . . I iaiuh .§.§§s . EHUHE ............ nae. ..HN40>.Z.N§AHHHMO&Q. . . . 435.03.0Hgg EBGHE ............ E . . HN40> .ggfiHHflmohfl . . . III 0 . Nun . Eonafimflgflg ESHg ............ E. ogg>.§H§a. . Ila. .fifi.§§s Managua» ............ an... . .>h¢o>.§mguum0hnm . . -ltnn. .58 . Egan ESE ............ 2E. .Huédmnnpdmmqnnmounm. . - Eéasdngmgu .553ng ............ BC. .qhgiugmuunuohn. . .....- .Eu.»>aa§.§.nu ......... hm. .SHE ............ a. .g‘. 0 I I . .‘fiHHgn. C .fiag. . . . DE . _ O m N o o N 0.54 .§.a>h§.n I19): 133%. .4“. . .HNKHHEE 2% NE. Elsa 4. I." I ~ ...—333% . .4K . . .Nan-Hzg I .5333. ....r E. .ZN . .ggag liviU-n - III-H” I- -u .55. a E. E. . Eggnog -I:u .. .133 ..... r 5.5. an . Enron» I. nan-”In L ngu E Iggog bag—nu... .53! ...... E. Q.» . HEHHOE I .. . :55... a £5.39. .8». .nadfiauog ... E. g lip LL- :u-Iul I II. M“: I F Egg g i I. Jan F .E ........ Ila-r 1.3.3.... '1 ganflHaH" ...GEEOE H>BO.QQAA¢QKH>B>.HO..OKHhHUQBflZ‘dflObZQ‘ZflHM>ZDOH¢£ZKQHH‘MKHUMZMUHGHQHOQHH‘AAK HbBO.MfiHA¢GKH>B>.HO..MKHhHUQBMN¢¢HO>ZGm2HHMbZfiUHMdZHGHH‘dHHMMZOOHOfl>>MOM¢¢AAK HHHO.MQHA¢QKH>B?.HO..mKHhMDQBHM14HO>ZQMZHHK>ZHUHM¢ZKGHH‘fiflHflDZKUHOHfibMQM‘dAAK ddfiflhflafl.¢ddd¢flKH>B?.HO..MKHhHOQBfl¢¢4HO>ZD>ZHHK>ZDOHN¢2KDHH‘flflHflDZKOHGHmHOQKHOflAM .Q‘HBbZKH>>>.HO..2“HHhUAHHMZdHObZHMZbHEfiZKOHM‘ZMGHH‘flflHBBOmOHaddHflafiflflflBD .QBHbi2Kfl>>?.HO.cMMHNhOOmHNQ4MO>ZDKZHHM>DNOHM¢ZMQHH‘flflHHBGNOHDbNHQUKGHMUO H C o B .HO..NKHMHUHHmmndfiOHOMMZBbeZMUHK‘ZSHHHUBflHKHZMUHOHQbKQ‘AAHKm .ndem2KA>>B.HO..QNHNhOHm¢MG¢HOZUfl¢ZdHBbZKUHM‘ZKQHHDZOHNHZHUHOthKUH‘flOQG .DKH d2flhH>B.HO..m..H>DG...ndflabbflm2MHHbzo. M‘ZKDHH‘GKH lllll . Fl IFI-rl fl ?hna aavgvaa cetzaifiuaw £2; .o8>Hfifld>lflflH20hflBHfl>Mhhfl>mO>>>M ..... OMHHOOOh>HMKdBZUNmUNKNOth‘MfiAMHSMBMHfiBBODHBQflOfiB?.NdflKNZdMR .0.. ..BhflhflflhlflflHflUth>fl>fluhflbflfl>>>h ..... OnH>000ubflmzdhzummauflfiabhfldfififlmHBNBKHExagnhxndflfi>>.uddDQZdKS HHOO. MH‘HHHHHD‘UZHQMQHHHddflmHUm¢fl0>>h2flHHUBKH>UOHDHbmnfibnd ..... Qfl>uKMQQHMKAOBHHfidfiflnhhthmHmMKBHOQZGBH NIUA . . . .hflEHg .0.—"Al“! MH‘GBMMHD‘USHABQHHH‘GflmHOMMAUHbMNNHKUZKHbUOHUHbmnBHnfl.....ndm.HummflflflfiafihHfiflflflflflhflhflfibfiflflfiHOflZ‘Zfi HnIHI flflddflflNNDRUZHQDQHHH‘UfluHUNKAD>>MNNHKUBKH>UOMOH>hflB>Dfl ..... OAS.KQQQHHMAOHMHh‘flflfihhhhflhbhflflflHOflZ‘Zfl IIHOQ KH‘HBHHHD‘GSRAQQHHH‘UflmHOmflAD>>NNNHKUZKH>UOflDH>hnb>nfl ..... nfibuKMflfiQNKQOBNHfidflflflhhhhflfibhflkfiHOflBflZfl .Oflu “HumMHHMDZDS>QBQHHBUQMHHMMGHOQ>MNNHMUOKHZUONUHEMHBme...oo2H3oKfihfiflhmOMBHH“(KHZHthQmmmoflafldflzflflfl ah. Mfl>m¢dhhnddzuflmHHHSUOBHHNMM>OA>MMOHKUZMAHUDHUHHMHBbMU ..... nagoKBmmflNMdMBHHfidflflnflhmhmm2h008fl4HZRZH lfll HHMHBONHD£UZZhBQMdfl>OUQAMNMHD>HMNH>QUQKHMUDNMHH‘GKHKU ..... ZDH.Kfimmflhflmmflhhmfiflmnhhahflm.hflOfifldflMflZfl HIOOQ HDHQHHMD‘USZS‘ZKAHM mZAKmm>UHmBfiHHMUHMAhUONAHHdfluHMU..... Gm.KBNhAHMfiWthABKQQNhflhKQ.mOHUflm221QH MI? SMHHBHHMQ‘USSZ‘ZKAHMUMZANQo..htfiflflHKUflKflhUOflAHH‘HKHMU ..... Mam Kammflflmmmfihhmfiflmnhhflhflm.AOHUQGM21flh DI! . dflflflBflMHfldUSSPMQKQHUOWN>EMD>UH>MNHHZUMMQAUOGMthfiMEKU ..... OH OH hflUflfi‘ HhfiBKhflHhflhK.thHBfldflzdfifi Nlfil KHSEHHHh946Q3?BQKAN>ZmOHMNMHOEhNNHbMUKDAhUOMmHHMGBHKM ..... 2m fidZflflfiHHflhKfldflhhflhM.mmafldfl¢A2dflh #3 dflHflMBDdG AABQKHN>BQB?..... Mm mazmmmahflfiflfidfihhuhfloQQMHHfldAZdflh ind. 23 determined by Edman degradation (Baecker et al., 1983; Leung & Preiss, 1987b). There is 90 % identity in the deduced amino acid sequences of the E. coli and S. typhimurium glgC genes, and most of the changes are conservative (Preiss & Romeo, 1989). The structural genes of the higher-plant ADP-glucose pyrophosphorylases have been cloned from several sources and tissues (Figure 1), such as maize endosperm (Barton et al., 1986), rice seed (Preiss et al., 1987a; Anderson et al., 1989), wheat leaf and endosperm (Olive et al. , 1989), potato tuber (Miiller-Rober et al. , 1990; Anderson et al., 1990; Nakata et al., 1991), and spinach leaf (Smith-White & Preiss, 1992). A systematic comparison of the primary structures of ADP- glucose pyrophosphorylases from different sources has been done by Smith-White and Preiss (1992). The small subunit is highly conserved with an amino acid sequence identity varying from 85 to 95 % when the enzymes from different plant sources and tissues are compared. Both the identity among large subunits from different sources and that between the large and small subunits are about 50 to 60 %. The identity of the E. coli enzyme with the higher-plant enzymes is about 30 %. Analysis of the deduced amino acid sequences (Figure 1) indicated that the cyanobacterial enzyme is more similar to the higher-plant enzyme than to the other bacterial enzymes in primary structure (Kakefuda et al., 1992; Charng et al., 1992). Furthermore, the cyanobacterial enzyme is more similar to the small than to the large subunit of higher-plant ADP-glucose pyrophosphorylases (Kakefuda et al. , 1992; Charng et al. , 1992). This is in agreement with the fact that the cross-reaction of cyanobacterial enzyme with the antibody prepared with the spinach leaf large subunit was much weaker 24 than with the antibody prepared with the small subunit (Iglesias et al. , 1991). Recent studies on potato tuber ADP-glucose pyrophosphorylase have shown that the small subunit expressed by itself exhibited high enzyme activity , with lower affinity for activator 3PGA and higher sensitivity toward Pi inhibition (Ballicora et al., 1995). It is suggested that the small subunit of higher-plant ADP-glucose pyrophosphorylase functions as the catalytic subunit while the large subunit is involved in the allosteric regulation of the activity (Ballicora et al., 1995). Based on the above observations, it has been postulated that the higher-plant ADP-glucose pyrophosphorylase genes evolved from the bacterial enzyme gene by gene duplication and diversion (Smith-White & Preiss, 1992). 6. Structures of the Substrate and Regulator Binding Sites of ADP-Glucose Pyrophosphorylase: From steady-state kinetics (Paule & Preiss, 1971; Kleczkowski, et al., 1993) and equilibrium binding (Haugen & Preiss, 1979) studies of the reaction catalyzed by ADP- glucose pyrophosphorylase, substrates are known to bind in an ordered mechanism. ATP binds first, followed by the binding of glucose-l-P (Haugen & Preiss, 1979). Product release is ordered with PPi released first, followed by ADP-glucose (Paule & Preiss, 1971). According to Haugen and Preiss (1979), there are four binding sites for substrates, activators, and inhibitor per tetrameric E. coli enzyme. In binding experiments, CrATP, ATP, and FBP appear to bind to only half of the expected sites in the tetrameric enzyme, 25 while ADP-glucose, the activators, pyridoxal-P (PLP) and 1,6-hexanediol-P2, and the inhibitor, AMP, bind to four sites per tetrameric enzyme. Glucose-l-P does not bind to the enzyme unless CrATP and MgCl2 are present and binds to four sites per tetrameric enzyme. In the presence of glucose-l-P, CrATP binds to four sites per tetrameric enzyme. It was suggested that the enzyme first binds two ATP molecules per tetramer which allows the binding of glucose-l-P, then binds two more ATP molecules. The binding of ATP and FBP is synergistic. ATP and FBP must be present together to effectively inhibit AMP binding. Thus, there are binding sites for substrates, activators, and inhibitor located on each subunit and the binding sites can interact. Structure—function studies to determine the location of substrate and regulator binding sites of ADP-glucose pyrophosphorylase have been mainly performed by chemical modification using substrate or regulator analogs and site—directed mutagenesis of cloned ADP-glucose pyrophosphorylases. Characterization of the mutant E. coli enzyme with altered regulatory properties also provides very useful information. In chemical modification studies, several compounds have been used as the affinity labels for ADP-glucose pyrophosphorylase, PLP, 8—azido—ATP, 8-azido-ADP-glucose, 8- azido-AMP, and phenylglyoxal. They all have the following properties. First, the affinity label has to be shown as an analog of the substrate, activator, or inhibitor. Second, the compound can bind the enzyme covalently in chosen conditions. Then, the labeled enzyme is subjected to proteolysis and the labeled peptide(s) is isolated and sequenced (Preiss & Sivak, 1996). PLP is an activator of many bacterial and higher-plant ADP-glucose 26 pyrophosphorylases (Parsons & Preiss, 1978a; Preiss et al., 1987b; Iglesias et al., 1991) and is considered as an analog of 3PGA. It has been used as a site specific probe to identify lysyl residues in enzymes that have binding sites for sugar phosphates. A Schiff base is formed between PLP and the e-amino group of lysine and can be converted to a covalent bond by reduction with sodium borohydride (NaBH4). The photoaffinity labeling agents 8-azido-ATP and 8-azido-ADP-glucose are considered as substrate analogs (Lee et al., 1986) and 8-azido-AMP as an inhibitor analog (Larsen et al. , 1986). In the presence of UV light, these 8-azido compounds form a nitrene radical which can react with electron-rich residues, and therefore, become covalently incorporated into the enzyme. Phenylglyoxal is an arginine-specific reagent (Takahashi, 1968). It reacts with the guanido group of the arginine residue under mild conditions, pH 7 to 8 and 25 °C. The derivative contains two phenylglyoxal moieties per guanido group. 6.1. The Primary Structure of the Substrate Binding Sites: Reductive phosphopyridoxylation of E. coli ADP-glucose pyrophosphorylase at high concentrations of PLP (Parsons & Preiss, 1978a) cause a decrease in enzyme activity, and this inactivation could be prevented by the presence of ADP-glucose plus Mg“. After reductive phosphopyridoxylation in the presence of FBP or hexanediol-l ,6-P2, the labeled enzyme was cleaved with CNBr. The labeled peptide was isolated and sequenced. Lys‘” was found to be labeled with [’11] PLP. These experiments suggested that Lys‘” is required for either substrate binding or for catalytic activity. Site-directed mutagenesis experiments have been performed to further study the 27 role of Lys‘” (Hill et al., 1991). The apparent affinity for glucose-l-P of the mutant enzyme, which was generated by substitution of Lys‘” with glutamate, is 12,000-fold lower than that of the wild-type enzyme. The kinetic constants for ATP, Mg“, ADP- glucose, and FBP are relatively unchanged. The catalytic efficiency of the mutant enzyme is similar to that measured for the wild-type enzyme. According to the kinetic studies on a series of Lys‘95 mutants, both the charge and size of the this lysine are believed to be important for proper binding of glucose-l-P. Figure 2 shows the alignment of the E. coli amino acid sequence near Lys195 with known sequences from other sources. Lys195 is conserved in all the sequences. Recently, site-directed mutagenesis was performed on potato tuber ADP-glucose pyrophosphorylase and the corresponding lysine was shown to be involved in the glucose-l-P binding of the potato tuber enzyme (Fu et a1. , unpublished data). To obtain further information on the substrate binding sites of the E. coli ADP- glucose pyrophosphorylase, the photoaffinity labeling agent 8-azido-ATP and 8-azido- ADP-glucose were used (Lee et al., 1986; Lee & Preiss, 1986). The kinetic experiments indicated that these two agents specifically interact at the substrate site of ADP-glucose pyrophosphorylase. Photo-inactivation of the E. coli enzyme with either 8-azido-ATP or 8-azido-ADP—glucose, in the presence of FBP and Mg2+ , was effectively prevented by the presence of either ATP or ADP-glucose. The incorporation sites of [“C]8-azido-ADP- glucose into the E. coli enzyme were determined by tryptic digestion of the modified 28 Figure 2: Amino acid sequence of the phosphopyridoxylated peptide, which was protected by ADP-glucose plus Mg“, of E. coli ADP-glucose pyrophosphorylase and the alignment with other known sequences. The labeled lysine residue, Lys‘”, is in bold. The numbers indicated correspond to the sequence of the E. coli enzyme. Nomenclature is the same as in Figure 1. 51-54 wl sh—2 we7 we3 pot-l ana syn rice pot—s sl-Sl arabid bt-Z ecoli S.t. 29 182 195 200 I I I K ...... VLSFSEKPKGDD KIDDTGRVISFSEKPRGAD KIDHTGRVLQFFEKPKGAD KFDSSGRVVQFSEQPKGDD KFDSSGRVVQFSEKPKGDD KIDSRGRVVQFAEKPKGFD KIDNSGRVIDFSEKPKGEA KIDAQGRITDFSEKPRGK. KIDEEGRIVEFAEKPKGEQ KIDEEGRIIEFAEKPQGEQ KIDETGRIIEFAEKPKGEQ KIDEEGRIIEFAEKPKGEH KIDEEGRIIEFAEKPKGEQ AVDENDKIIEFVEKP.ANP AVDESDKIIDFVEKP.ANP Figure 2 30 enzyme and isolation of the labeled peptides via HPLC (Lee & Preiss, 1986). Amino- terminal sequence analysis showed that about 65 % of the label was in the region of 108 to 128 and Tyr“4 was believed to be the acceptor of the nitrene radical. This region is highly conserved in all sequences known to date (Figure 3). About 20 % of the label was incorporated between Cys163 and Lys208 and Lys195 is in that region. Another region with 15 % incorporation is between Arg381 and Mg”. Tyr114 was found in a predicted Rossman—fold supersecondary structure, a structure consisting of three consecutive strands of parallel B—pleated sheet and two joining a-helices (Larsen et al., 1986). Study of the 114 mutant enzyme with a Tyr to phenylalanine substitution via site-directed mutagenesis indicated that Tyr‘“ is involved directly or indirectly in enzyme catalysis (Kumar et al. , 1988). 6.2. The Primary Structure of the Regulator Binding Sites: Previous studies have shown that the activator-binding site of the E. coli enzyme is near the N-terminus (Parsons & Preiss, 1978a,b; Gardiol & Preiss, 1990). Reductive phosphopyridoxylation of the E. coli ADP-glucose pyrophosphorylase in the presence of ADP-glucose plus Mgl+ led to formation of an enzyme which no longer required FBP for high activity (Parsons & Preiss, 1978a). The incorporation of PLP was inhibited by the presence of other activators, such as FBP and hexanediol- l ,6-P2. These results suggest that reductive phosphopyridoxylation occurred at the activator binding site (Parsons & Preiss, 1978a). The E. coli enzyme labeled with PLP was degraded with CNBr and the labeled peptide was isolated and sequenced (Parsons & Preiss, 1978b). Lys39 was found to be the labeled lysine residue. The region around Lys39 (Figure 4) is highly conserved in all the 31 Figure 3: Amino acid sequence of the major azido-ADP-P‘C] glucose incorporated peptide of E. coli ADP-glucose pyrophosphorylase and the alignment with other known sequences. Tyr‘“ is in bold. The numbers indicated correspond to the sequence of the E. coli enzyme. Nomenclature is the same as in Figure 1. 81-54 sh-Z we7 pot—l ana syn rice pot—s sl-Sl arabid bt-2 ecoli s.t. 108 ....S. 32 129 .WFQGTADAVRQFGWLFED PEEPAG.WFQGTADSIRKFIWVLED PGEAAG.WFRGTADAWRK.IWVLED PGEAGKKWFQGTADAVRKFIWVFED PENPN. KDNPD. PDNPN. PENPD. PENPD. PENPN. PDNPN. MKGEN. MKGEN. .WFQGTADAVRQYLWMLQE .WFQGTADAVRQYLWLFRE .WFQGTADAVRQYLWLFEE .WFQGTADAVRQYLWLFEE .WFQGTADAVRQYLWLFEE .W ................. .WFQGTADAVRQYLWLFEE .WYRGTADAVTQNLDIIRR .WYRGTADAVTQNLDIIRR Figure 3 33 Figure 4: Amino acid sequence of the phosphopyridoxylated peptide, which was protected by fructose-1,6-P2, of E. coli ADP-glucose pyrophosphorylase and the alignment with other known sequences. The labeled lysine residue, Lys”, is in bold. The numbers indicated correspond to the sequence of the E. coli enzyme. Nomenclature is the same as in Figure 1. . sl—54 sh-2 we7 pot-l ana syn rice pot-s sl-Sl arabid bt-Z ecoli s.t. 34 33 I LFPL ...... PAV LFPLTSTRATPAV LFPLTSTRATPAV LFPLTSRTATPAV LYPLTKLRAKPAV LYPLTKLRAKPAV LYPLTKKRAKPAV LYPLTKKRAKPAV LYPLTKKRAKPAV LYPLTKKRAKPAV LYPLTKKRAKPAV LKDLTNKRAKPAV LKDLANKRAKPAV Figure 4 35 ADP—glucose pyrophosphorylase sequences known to date. Substitution of Lys39 with glutamic acid formed an enzyme with 23-fold lower apparent affinity for the activator, FBP (Gardiol & Preiss, 1990). Therefore, Lys39 is believed to be involved in the binding of FBP. Similar studies using PLP as the affinity label agent was conducted for the spinach leaf ADP-glucose pyrophosphorylase (Morell et al., 1988; Preiss et al., 1992; Ball & Preiss, 1994). When PLP is covalently bound, the enzyme no longer requires 3PGA for activation. It was shown that one lysine residue of the small subunit (Morell et al. , 1988) and three lysine residues of the large subunit (Preiss et al., 1992; Ball & Preiss, 1994) are involved in the binding of PLP (Figure 5). The modification of these lysine residues by PLP is prevented by the presence of 3PGA, while only the lysine residue of site 1 of the small subunit (Lys‘w) and site 2 of the large subunit are prevented by Pi from being modified. These observations indicate that the activator analog, PLP, binds at the activator site and the most important sites involved in the binding of activator are site 1 and site 2. Recent chemical modification and site-directed mutagenesis studies on the cyanobacterial ADP-glucose pyrophosphorylase from Anabaena have shown that Lys‘”, corresponding to Lys440 of the small subunit of spinach leaf enzyme (Figure 5), is involved in the binding of the activator, 3PGA (Charng et al. , 1994). Chemical modification with PLP resulted in an enzyme that no longer required activator for high activity and the modification was prevented by 3PGA and Pi. Lys“19 was found to be labeled by PLP. 419 Replacing Lys with either arginine, alanine, glutamine, or glutamic acid produced 36 Figure 5: Amino acid sequence of the phosphopyridoxylated peptides of spinach leaf and Anabaena ADP-glucose pyrophosphorylases and the alignment with other known sequences of higher-plant and cyanobacterial enzymes. The labeled lysine residues are in bold. The number indicated corresponds to the sequence of the small subunit of the spinach enzyme. Nomenclature is the same as in Figure 1. 81-54 wl sh-2 we7 we3 pot—l ana syn rice pot—s sl-Sl bt-2 Site 3 KWFQGTADAVRQ GWFQGTADSIRK GWFRGTADAWRK KWFQGTADAVRK NWFQGTADAVRQ DWFQGTADAVRQ NWFQGTADAVRQ DWFQGTADAVRQ DWFQGTADAVRQ NWFQGTADAVRQ 37 Site 2 IKDAIIDKNAR IQNCIIDKNAR IRNCIIDMNAR ISNCIIDMNAR ISNCIIDMNAR IRKCIIDKNAK IRRAIIDKNAR IRRAIIDKNAR IRRAIIDKNAR IKRAIIDKNAR IKRAIIDKNAR IRRAIIDKNAR Figure 5 Site 1 440 I SGITVIFKNATIKDGVV SGITVVLKNSVIADGLVI SGIVVILKNATINECLVI SGIVVIQKNATIKDGTVV SGIVVIQKNATIKDGTVV SGIIIILEKATIRDGTVI SGIVVVLKNAVITDGTII NGIVVVIKNVTIADGTVI SGIVTVIKDALLLAEQLY SGIVTVIKDALIPSGIII SGIVTVIKDALIPSGTVI GGIVTVIKDALLPSGTVI 38 mutant enzymes with apparent affinities for 3PGA 25-, 50-, 140-, or 150-fold lower than that of the wild-type enzyme, respectively. These mutant enzymes still could be activated at higher concentrations of 3PGA, which suggested that other residue(s) might be involved in the binding of 3PGA. Chemical modification studies of the K419R enzyme have shown that PLP can also be incorporated into this mutant enzyme. The phosphopyridoxylated K419R enzyme is less dependent on the presence of 3PGA for activation and less sensitive to Pi inhibition. Lys382 of the mutant enzyme was identified as the labeled residue by PLP and corresponds to the PLP-labeled lysine residue of site. 2 of the spinach leaf large subunit (Figure 5). Kinetic and protection studies suggested that Lys382 might be also involved in binding of the allosteric activator (Charng et al., 1994). The photoaffinity analog 8-azido-AMP was used as a site-specific probe for the inhibitor-binding sites of the E. coli ADP-glucose pyrophosphorylase (Larsen et al. , 1986). In the absence of light, the enzyme was inhibited by 8-azido—AMP and the inhibition is similar to that observed with the natural inhibitor, AMP. Tyr‘“, which is located within the major incorporation site of 8-azido-ADP-glucose, was identified as the amino acid modified by 8-azido-AMP. This suggests that the binding site for AMP might be overlapping with those for the substrates, ATP and ADP-glucose. The arginine-specific reagent phenylglyoxal has been used to covalently modify ADP-glucose pyrophosphorylases from E. coli (Carlson & Preiss, 1982), Synechocystis (Iglesias et al., 1992), and spinach leaf (Ball & Preiss, 1992). The results suggested that there was an essential arginine residue(s) involved in the allosteric regulation. However, the arginine residue(s) labeled by phenylglyoxal has not been identified. 39 Mutant E. coli stains with altered glycogen accumulation have been produced by chemical mutagenesis (Govons et al., 1969; Preiss et al., 1976; Creuzat-Sigal et al., 1972). ADP-glucose pyrophosphorylases from E. coli B mutant strain CL1136 (Preiss et al., 1969) and SGS (Govons et al., 1969, 1973) and E. coli K12 mutant strain 618 (Cattaneo et al. , 1969; Creuzat-Sigal et al. , 1972) have higher activity in the absence of the activator, FBP, and are less sensitive to the inhibition by AMP than the wild-type enzyme. It was found that Arg67 of mutant CL1136 (Ghosh et al. , 1992), Prom of mutant sos (Meyer et al., 1992), and my336 of mutant 618 (Kumar et al., 1939) were changed to cysteine, serine, and aspartate, respectively. The enzyme from E. coli B mutant strain S614, in which Ala44 was replaced with threonine, has lower affinities for the substrates, the activator, FBP, and the inhibitor, AMP (Meyer et al., 1993). These residues might also play important roles in the regulation of ADP-glucose pyrophosphorylase. 7. Prediction of the Secondary Structure of ADP-Glucose Pyrophosphorylase: The structures of ADP-glucose pyrophosphorylases from different sources have been predicted in the basis of their amino acid sequences by Ballicora et al. (1996) using hydrophobic cluster analysis (Lemesle—Varloot et al. , 1990) and PHD program (Rost & Sander, 1993). Hydrophobic analysis shows that ADP-glucose pyrophosphorylases are identical in the position of many clusters and in some others the differences are small. The small insertions seen among ADP-glucose pyrophosphorylases are not part of the "core" of the protein because there are no buried amino acids in these insertions, indicating that ADP-glucose pyrophosphorylases from different sources share a common folding pattern 40 though they have different quaternary structures (homotetrameric in bacteria and heterotetrameric in higher plants) and different specificities for the regulators. A general structure (Figure 6) that fits all of these protein was predicted using the PHD program (Rost & Sander, 1993). ADP-glucose pyrophosphorylase is an alpha/beta protein with some parts of it mainly beta, for example, the C-terminus and the domain denoted as 3 (Figure 6). Most of the conserved amino acids known to have roles in the binding of substrates (T yr114 and Lys‘” in E. call) and activators (Lys39 in E. coli and Lys419 in Anabaena) are located in 100ps, or very close to loops. The residues Prom and Gly336 that are important for the regulation of the E. coli enzyme are also in loops. A supersecondary structure seen in nucleotide-binding proteins (Rossmann et al. , 1974) is also present in this model. The glycine loop in the domain 1, which is thought to bind the phosphates of the ATP, and the region 2, with three beta sheets and alpha helices compatible with a Rossmann fold. The predicted secondary structure in region 1 and 2 is identical to the structure of the oncogenic protein H-Ras (p21) which is used as the folding model for nucleotide phosphate binding of GTP (Tong et al., 1991). The results from trypsin treatment of the E. coli and Anabaéna ADP-glucose pyrophosphorylases (Charng & Preiss, unpublished data) are in good agreement with the predicted model. Exposed loops would be more sensitive to proteolysis and the proteolysis studies show that trypsin cuts at sites predicted as loops by the model. Recent proteolysis studies of ADP-glucose pyrophosphorylases from E. coli (Wu et al. , unpublished data) and Anabaena (Sheng et al. , unpublished data) with proteinase K also confirmed that the 41 Figure 6: Profile Neural Network (PHD) prediction of the secondary structure of ADP-glucose pyrophosphorylase (Ballicora et al., 1996). Section 1 contains the FBP activator site KRAKPAV (E. call) in a loop as well as Arg‘”. Section 2 has the putative ATP binding site, Tyr‘“ in a loop area between a beta-strand and alpha-helix starting at GTAD. The glucose-l-P binding site is also seen in a loop among a series of predicted beta-strands. The topology between regions 1 and 2 cannot be ascertained and is drawn as a dotted line. GlPSite V mzu ,. O ..-...OOOOOIOOOOOOI......I.---O-" .00.... ......COOCCCCOU N—T (V' 42 Anahaennndplm Helmmaloop Figure 6 43 predicted model is valid, since all the cleavage sites are in the predicted loops. The cleavage sites close to the putative glucose-l-P binding site (Lys‘95 in E. coll) are protected by ADP-glucose, Mg“, and the activator. Interestingly, the E. coli enzyme treated with proteinase K in the presence of ADP-glucose, Mg“, and FBP, with 11 amino acids from the N -terminus and 2 amino acids from the C-terminus cleaved, retains maximal activity but it is no longer sensitive to AMP inhibition and does not require FBP for activation (Wu et al. , unpublished data). It was postulated that the N-terminal loop is involved in the allosteric regulation of the E. coli enzyme. REFERENCES Aitken, A. (1988) Methods Enzymol. 167, 145-154. Anderson, J. M., Hnilo, J., Larson, R., Okita, T. W., Morell, M., & Preiss, J. (1989) J. Biol. Chem. 264, 12238-12242. Anderson, J. M., Okita, T. W., & Preiss, J. (1990) in The Molecular Biology of the Potato (Vayda, M. E., & Park, W. D., eds) pp. 159-180, International, Wallingford. ap Rees, T. (1984) in Storage Carbohydrates in Vascular Plants (Lewis, D. H., ed.) pp 53-73, Cambridge University Press, Cambridge. Baecker, P. A., Furlong, C. E., & Preiss, J. (1983) J. Biol. Chem. 258, 5084-5088. Ball, K., & Preiss, J. (1992) J. Protein Chem. 11, 231-238. Ball, K., & Preiss, J. (1994) J. Biol. Chem. 269, 24706-24711. Ball, 8., Marianne, T., Dirik, L., Fresnoy, M., Delrue, B., & Decq, A. (1991) Planta 185, 17-26. Ballicora, M. A., Laughlin, M. J., Fu, Y., Okita, T. W., Barry, G. F., & Preiss, J. (1995) Plant Physiol. 109, 245-251. Ballicora, M. A., Fu, Y., Wu, M-X., Sheng, J ., Nesbitt, & Preiss, J. (1996) inAbstracts of the 5th NIAR/COE International Symposium (Nakamura, Y. ed.) pp 5-11, National Institute of Agrobiological Resources, Tsukuba, Japan. Barton, C., Yang, L., Galvin, M., Sengupa-Gopalan, C., & Borelli, T. (1986) in Regulation of Carbon and Nitrogen Reduction and Utilization in Maize (Shannon, J. C. , Kinevel, D. P., & Boyer, C. D., eds.) pp 363-365, American Society of Plant Physiologist, Rockville, MD. Brana, A. F., Manzanal, M. B., Hardisson, C. (1980) J. Bacterial. 144, 1139-1142. Carlson, C.A., & Preiss, J. (1982) Biochemistry 21, 1929-1934. Cattaneo, J ., Damotte, M., Sigal, N., Sanchez-Medina, G., & Puig, J. (1969) Biochem. Biophys. Res. Commun. 34, 694-701. Charng, Y., Kakefuda, G., Iglesias, A. A., Buikema, W. J ., & Preiss, J. (1992) Plant Mol. Biol. 20, 37-47. 44 45 Charng, Y., Iglesias, A. A., & Preiss, J. (1994) J. Biol. Chem. 269, 24107-24113. Creuzat-Sigal, N., Latil-Damotte, M., Cattaneo, J ., & Puig, J. (1972) in Biochemistry of the Glycosidic Linkage (Piras, R., & Pontis, H. G., eds.) pp 647-680, Academic Press, New York. Dickinson, D. B., & Preiss, J. (1969) Plant Physiol. 44, 1058-1062. Edwards, J ., Green, J. H., & ap Rees, T. (1988) Phytochemistry 27, 1615-1620. Espada, J. (1962) J. Biol. Chem. 237, 3577-3581. Furlong, C. E., Preiss, J. (1969) J. Biol. Chem. 244, 2539-2548. Gardiol, A., & Preiss, J. (1990) Arch. Biochem. Biophys. 280, 175-180. Ghosh, H. P., & Preiss, J. (1965) J. Biol. Chem. 240, 960-961. Ghosh, P., Meyer, C. R., Remy, E., Peterson, D., & Preiss, J. (1992) Arch. Biochem. Biophys. 296, 122-128. Govons, S., Vinopal, R., Ingraham, J ., & Preiss, J. (1969) J. Bacteriol. 97, 970-972. Govons, S., Gentner, N., Greenberg, E., & Preiss, J. (1973) J. Biol. Chem. 248, 1731- 1740. Greenberg, E., & Preiss, J. (1964) J. Biol. Chem. 239, 4314-4315. Haugen, T. H., Ishaque, A., & Preiss, J. (1976) J. Biol. Chem. 251, 7880-7885. Haugen, T. H., & Preiss, J. (1979) J. Biol. Chem. 254, 127-136. Heldt, H. W., Chon, C. J., Maronde, D., Herold, A., Stankovic, Z. S., Walker, D. A., Kraminer, A., Kirk, M. R., & Herber, U. (1977) Plant Physiol. 59, 1146-1155. Hill, M. A., Kaufmann, K., Okita, J., & Preiss, J. (1991) J. Biol. Chem. 266, 12455- 12460. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1991) Plant Physiol. 97, 1187-1195. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1992) J. Protein Chem. 11, 119-128. Iglesias, A. A., Charng, Y., Ball, 8. & Preiss, J. (1994) Plant Physiol. 104, 1287-1294. 46 Kacser, H., & Burns, J. A. (1973) symposium Soc. Exp. Biol. 27, 65-107. Kacser, H. (1987) in The Biochemistry of Plants (Davies, D. D., ed.) Vol 11, pp 39-67, Academic Press, New York. Kakefuda, G., Charng, Y., Iglesias, A. A., Mcintosh, L., & Preiss, J. (1992) Plant Physiol. 99, 359-361. Kleczkowski, L. A., Villand, P., Preiss, J., & Olsen, 0. -A. (1993) J. Biol. Chem. 268, 6228-6233. Krebs, E. G., & Preiss, J. (1975) in International Review of Science, Biochemistry of Carbohydrates (Whelan, W. J ., ed.) pp 337-389, University Park Press, Baltimore. Kruger, N. J. (1990) in Plant Physiology, Biochemistry and. Molecular Biology (Dennis, D. T., & Terpin, D. H., eds.) pp 59-76, John Wiley, New York. Kumar, A., Tanaka, T., Lee, Y. M., & Preiss, J. (1988) J. Biol. Chem. 263, 14634- 14639. Kumar, A., Ghosh, R, Lee, Y. M., Hill, M. A., & Preiss, J. (1989) J. Biol. Chem. 264, 10464-10471. Larsen, C. E., Lee, Y. M., & Preiss, J. (1986) J. Biol. Chem. 261, 15402-15409. Lee, Y. M., Mukherjee, S., & Preiss, J. (1986) Arch. Biochem. Biophys. 244, 585-595. Lee, Y. M., & Preiss, J. (1986) J. Biol. Chem. 261, 1058-1064. Lehmann, M., & Preiss, J. (1980) J. Bacteriol. 143, 120-127 . Leloir, L. F., Olivarria, J. M., Goldemberg, S. H., & Carminatti, H. (1959) Arch. Biochem. Biophys. 81 , 508. Lemesle-Varloot, L., Henrissat, B., Gaboriaud, C., Bissery, V., Morgat, A., & Momon, J. P. (1990) Biochimie. 72, 555-574. Leung, P. S. C., & Preiss, J. (1987a) J. Bacteriol. 129, 4349-4354. Leung, P. S. C., & Preiss, J. (1987b) J. Bacteriol. 129, 4355-4360. Levi, C., & Preiss, J. (1976) Plant Physiol. 58, 753-756. Lin, T. -P., Caspar, T., Somerville, C., & Preiss, J. (1988a) Plant Physiol. 86, 1131- 47 1135. Lin, T. -P., Caspar, T., Somerville, C., & Preiss, J. (1988b) Plant Physiol. 88, 1175- 1181. Mackey, B. M., Morris, J. G. (1971) J. Gen. Microbial. 66, 1-13. Manners, D. J. (1985) in Biochemistry of Storage Carbohydrates in Green Plants (Dey, ' P. M., & Dixon, R. A., eds.) pp 149-203, Academic Press, New York. Mereschowsky, C. (1905) Biol. Zentralbl. 25, 593. Meyer, C. R., Ghosh, P., Remy, E., & Preiss, J. (1992) J. Bacterial. 174, 4509-4512. Meyer, C. R., Ghosh, P., Nadler, S., & Preiss, J. (1993) Arch. Biochem. Biophys. 302, 64-71. Morell, M., Bloom, M., Knowles, V., & Preiss, J. (1987) Plant Physiol. 85, 182-187. Morell, M., Bloom, M., & Preiss, J. (1988) J. Biol. Chem. 263, 633-637. Morrison, W. R., & Karkalas, J. (1990) in Methods in Plant Biochemistry (Dey, P. M., ed.) Vol 2, pp 323-353, Academic Press, New York. Miiller-Rtiber, B. T., KoBmann, J ., Hannah, L. C., Wilmitzer, L., & Sonnewald, U. (1990) Mol. Gen. Genet. 224, 136-146. Muller-Robot, B. T., Sonnewald, U., & Wilmitzer, L. (1992) EMBO J. 11, 1229-1238. Nakata, P. A., Greene, T. W., Anderson, J. M., Smith-White, B. J., Okita, T. W., & Preiss, J. (1991) Plant Mol. Biol. 17, 1089-1093. Neuhaus, H. E., Kruckeberg, A. L., Feil, R., & Stitt, M. (1989) Planta 178, 110-122. Neuhaus, H. E., & Stitt M. (1990) Planta 182, 445-454. Okita, T. W., Greenberg, E., Kuhn, D. N., & Preiss, J. (1979) Plant Physiol. 64, 187- 192. Okita, T. W., Rodriguez, R. L., & Preiss, J. (1981) J. Biol. Chem. 256, 6944-6952. Okita, T. W., Nakata, P. A., Anderson, J. M., Sowokinos, J., Morell, M., & Preiss, J. (1990) Plant Physiol. 93, 785-790. 48 Olive, M. R., Ellis, R. J., & Schuch, W. W. (1989) Plant Mol. Biol. 12, 525-538. Ozbun, J. L., Hawker, J. S., Greenberg, E., Lammel, C., & Preiss, J. (1973) Plant Physiol. 51, 1-5. Parsons, T., & Preiss, J. (1978a) J. Biol. Chem. 253, 6197-6202. Parsons, T., & Preiss, J. (1978b) J. Biol. Chem. 253, 7638-7645. Paule, M.R., and Preiss, J. (1971) J. Biol. Chem. 246, 4602-4609. Pettersson, G., & Ryde-Pettersson, U. (1989) Eur. J. Biochem. 179, 169-172. Preiss, J. (1969) Curr. Topics Cell Reg. 1, 125-161. Preiss, J ., Sabraw, A., & Greenberg, E. (1971) Biochem. Biophys. Res. Commun. 42, 180-196. Preiss, J. (1973) The Enzymes 8, 73-119. Preiss, J ., Greenberg, E., & Sabraw, A. (1975) J. Biol. Chem. 250, 7631-7638. Preiss, J ., Lammel, C., & Greenberg, E. (1976) Arch. Biochem. Biophys. 174, 105-119. Preiss, J. (1978) in Advances in Enzymalagy and Related Areas of Molecular Biology (Meister, A., ed.) Vol. 46, pp 317-381, Wiley-Interscience, New York. Preiss, J., & Levi, C. (1980) in Biochemistry of Plant Carbohydrates (Preiss, J ., ed.) Vol 3, pp 371-423, Academic Press, New York. Preiss, J ., & Walsh, D. A. (1981) in Biology of Carbohydrates (Ginsburg, V., ed.) Chapter 5, pp 199-313, J. Wiley and Sons, New York. Preiss, J. (1982) Annu. Rev. Plant Physiol. 54, 431-454. Preiss, J. (1984) Annu. Rev. Microbial. 38, 419-458. Preiss, J., Bloom, M., Morell, M., Knowles, V. L., Plaxon, W. C., Okita, T. W., Larson, R., Harmon, A. C., & Putnam-Evans, C. (1987a) in Tailoring Genes for Crop Improvement (Bruening, G., Harada, J., Kosuge, T., & Hollaender, A., eds.) pp 132-152, Plenum, New York. Preiss, J ., Morell, M., Bloom, M., Knowles, V. L., & Lin, T. P. (1987b) in Proceeding of the V11 International Congress on Photosynthesis (Biggins, J ., ed.) Vol. 3, pp 693-700, 49 Martinus Nijhoff, Dordrecht, the Netherlands. Peiss, J. (1988) in Biochemistry of Plants (Preiss, J., ed.) Vol XIV, pp 181-254, Academic Press, New York. Preiss, J ., & Romeo, T. (1989) Adv. Microb. Physiol. 30, 183-238. Preiss, J., Danner, S., Summers, P. S., Morell, M., Barton, C. R., Yang, L., & Nieder, M. (1990) Plant Physiol. 92, 881-885. Preiss, J. (1991) in Plant Molecularand Cell Biology (Mifflin, B., ed.) Vol. 7, pp 59-114, Oxford University, Oxford. Preiss, J ., Ball, K., Charng, Y., & Iglesias, A. A. (1992) in Research in Photosynthesis (Murata, N., ed.) Vol. 3, pp 697-700, Kluwer Academic Publishers, Dordrecht, Netherlands. Preiss, J. (1996) Biotech. Annu. Rev. 2, 259-279. Preiss, J, & Sivak, M. (1996) in Photaassimilate Distribution in Plants and Crops (Zamski, E., & Schaffer, A. A., eds.) pp 63-96, Marcel Dekker, Inc., New York. Rossmann, M. G., Moras, D., & Olsen, K. W. (1974) Nature 250, 194-199. Rost & Sander (1993) Proc. Natl. Acad. Sci. 90, 7558-7562. Sanwal, G. G., & Preiss, J. (1967) Arch. Biochem. Biophys. 119, 454-469. Sanwal, G. G., Greenberg, E., Hardie, J., Cameron, E. C., & Preiss, J. (1968) Plant Physiol. 43, 417-427. Shen, L., & Preiss, J. (1964) Biochem. Biophys. Res. Commun. 17, 424-429. Shively, J. M. (1988) Methods Enzymol. 167, 195-203. Slock, J. A., & Stahly, D. P. (1974) J. Bacterial. 120, 399-406. Smith, A. M., Bettey, M., & Bedford, I. D. (1989) Plant Physiol. 89, 1279-1284. Smith-White, B. J ., & Preiss, J. (1992) J. Mol. Eval. 34, 449-464. Stalmans, W. & Hers, H. G. (1973) in The Enzyme (Boyer, P. D., ed.) 3rd ed, vol. 8, pp 309-361. Academic Press, New York. 50 Stark, D. M., Timmerman, K. P., Barry, G. F., Preiss, J., & Kishore, G. M. (1992) Science 258, 287-292. Steiner, K. E., & Preiss, J. (1977) J. Bacterial. 129, 246-253. Strange, R. E., Dark, F. A., & Ness, A. G. (1961) J. Gen. Microbial. 25, 61-76. Strange, R. E. (1968) Nature 220, 606-607. Strasdine, G. A. (1968) Can. J. Microbiol. 14, 1059-1062. Strasdine, G. A. (1972) Can. J. Microbiol. 18, 211-217. Takahashi, K. (1968) J. Biol. Chem. 243, 6171-6179. Tong, L. A., de Vos, A. M., Milburn, M. V., Kim, S. H. (1991) J. Mol. Biol., 217, 503- 516. Tsai, C. Y., & Nelson, 0. E. (1966) Science 151, 341-343. Van Houte, J ., & Jansen, H. M. (1970) J. Bacterial. 101, 1083-1085. Yung, S., & Preiss, J. (1982) J. Bacterial. 151, 742-749. CHAPTER II SITE-DIRECTED MUTAGENESIS OF LYSINEm, THE ACTIVATOR- BINDING SITE, OF ADP-GLUCOSE PYROPHOSPHORYLASE FROM ANABAENA PCC 7120 51 ABSTRACT Previous studies have shown that a highly conserved lysyl residue (Lysm) near the C-terminus of Anabaena ADP-glucose pyrophosphorylase is involved in the binding of 3PGA, the allosteric activator (Charng et al. , 1994). Phosphopyridoxylation of the K419R mutant enzyme modified another conserved lysyl residue (Lysm), suggesting that this residue might also be located within the activator-binding site (Charng et al. , 1994). Site- directed mutagenesis of Lys382 of the Anabaena enzyme was performed to determine the role of this residue. Replacing Lys382 with either arginine, alanine, or glutamine produced mutant enzymes that were still activated by 3PGA but with apparent affinities 10— to 160- fold lower than that of the wild-type enzyme. In contrast, the glutamic acid mutant enzyme was inhibited by 3PGA. These mutations had lesser impact on the kinetic constants for the substrates and inhibitor, Pi, and on the thermal stability. These results indicate that both the charge and size of the residue at position 382 influence the binding of 3PGA. Site- directed mutagenesis was also performed to obtain a K382R-K419R double mutant. The apparent affinity for 3PGA of this double-mutant enzyme was 104-fold lower than that of the wild-type enzyme, and the specificity for activator of this mutant enzyme was altered. The K382R-K419R enzyme could not be phosphopyridoxylated, suggesting that other lysine residues are not involved in the binding of 3PGA. 52 INTRODUCTION ADP-glucose pyrophosphorylase catalyzes the conversion of glucose-l-P and ATP to ADP-glucose and PP,. This allosterically regulated enzyme catalyzes the first committed step in the biosynthesis of glycogen in bacteria (Preiss & Romeo, 1989; Preiss, 1996) and starch in plants (Preiss, 1991, Preiss, 1996; Preiss & Sivak, 1996). ADP-glucose pyrophosphorylase from higher plants is mainly activated by 3PGA and inhibited by P, (Ghosh & Preiss, 1965; Sanwal et al., 1968), whereas the enzyme from enteric bacteria is activated by FBP and inhibited by AMP (Preiss et al., 1966). The bacterial enzyme is homotetrameric in structure (Haugen et al., 1976), while. the higher-plant enzyme is heterotetrameric with two different subunits (Preiss, 1991). Cyanobacteria are prokaryotic organisms that have metabolic properties similar to chloroplasts of higher plants (Aitken, 1988). However, the former organisms synthesize glycogen as the major carbohydrate reserve (Shively, 1988), in a similar manner to what is observed in bacteria. The cyanobacterial ADP-glucose pyrophosphorylase is homotetrameric similar to the E. coli enzyme, but is regulated by 3PGA and P, like the ' higher-plant enzyme (Levi & Preiss, 1976; Iglesias et al., 1991). Analysis of the deduced amino acid sequence indicated that the cyanobacterial enzyme is more similar to the higher-plant enzyme than to the enteric bacterial enzyme in primary structure (Smith- White & Preiss, 1992). It is of great interest to understand the structure—function relationships of the allosteric site of the cyanobacterial ADP-glucose pyrophosphorylase due to the properties of the cyanobacterial enzyme and the key position of these photosynthetic prokaryotes during evolution (Aitken, 1988). 53 54 Previous studies have shown that the activator-binding site of the E. coli enzyme, Lys”, is near the N-terminus (Parsons & Preiss, 1978; Gardiol & Preiss, 1990), while that of the spinach leaf enzyme, Lys”, is situated toward the C-terminus (Morell et al., 1988; Ball & Preiss, 1994). Chemical modification and site-directed mutagenesis studies on the cyanobacterial ADP-glucose pyrophosphorylase from Anabaena have shown that Lys”, corresponding to Lys“10 in spinach leaf enzyme, is involved in binding of the activator, 3PGA (Charng et al., 1994). Lys‘”9 was labeled by PLP, an analogue of the activator, 3PGA. Replacing Lys419 with either arginine, alanine, glutamine, or glutamic acid produced mutant enzymes with apparent affmities for 3PGA 25-, 50—, 140-, or lSO-fold lower than that of wild-type enzyme, respectively. These mutant enzymes still could be activated by 3PGA, which suggested that other residue(s) might be involved in the binding of 3PGA. Chemical modification studies of the K419R enzyme have shown that PLP can also be incorporated into this mutant enzyme. The phosphopyridoxylated K419R enzyme is less dependent on the presence of 3PGA for activation and less sensitive to P, inhibition. Lys332 of the mutant enzyme was found to be labeled by PLP. Kinetic and protection studies suggested that Lys382 was also involved in binding of the allosteric activator (Charng et al., 1994). Here, the results of further site-directed mutagenesis and chemical modification of the cyanobacterial ADP-glucose pyrophosphorylase from Anabaena are reported. Lys382 was replaced with either arginine, alanine, glutamine, or glutamic acid with site-directed mutagenesis to determine the role of this residue. A double-mutant enzyme, K382R- K419R, was also generated and characterized. MATERIALS AND METHODS Reagents [”P]PP, was purchased from Du Pant-New England Nuclear. [“C]Glucose—1-P was from ICN Pharmaceuticals Inc. [a-35S]dATP and the in vitro mutagenesis kit were from Amersham Corp. [4-3H]PLP was synthesized and purified as described previously (Morell et al. , 1988) according to the method of Stock et al. (1966) and Ahrens and Kortnyk (1969). Oligonucleotides were synthesized and purified by the Macromolecular Facility at Michigan State University. All other reagents were purchased at the highest quality available. . Bacterial Strains and Media E. coli strain TGl [K12, A(lac-pro), supE, thi, hsdD5/F'traD36, proA+B+, lacl“, lacZAMlS] was used for site-directed mutagenesis and grown in LB medium. E. coli mutant strain AC70R1-504, which is deficient in ADP-glucose pyrophosphorylase activity (Carlson et al., 1976), was used for expression of the Anabaena ADP-glucose pyrophosphorylase gene (Charng et al. , 1992) and grown in enriched medium containing 1.1% KzHPO4, 0.85% KHZPO4, 0.6% yeast extract, and 0.2% glucose, pH 7.0. Site-Directed Mutagenesis Plasmid pAnaE3a was used for both site-directed mutagenesis and gene expression. In pAnaE3a, a 5.5 kb EcoRI fragment of Anabaena genomic DNA containing the Anabaena ADP-glucose pyrophosphorylase gene and its putative promoter was ligated onto the EcoRI site of pUC119 plasmid (Charng et al., 1992). The orientation of the gene is opposite to the lac promoter of pUC119 and enables TGl cells to synthesize 55 56 single-stranded DNA containing the antisense strand of the gene using helper phage M13KO7 (Sambrook et al., 1982). Site-directed mutagenesis was performed according to the method of Sayers et al. (1988) using the in vitro mutagenesis kit from Amersham Corp. The Lys382 mutant enzymes with substitution of arginine, alanine, glutamine, and glutamate acid were designated as K382R, K382A, K382Q, and K382E, respectively. The 419 double mutant enzyme, in which both Lys382 and Lys were replaced with arginine, was obtained by performing site-directed mutagenesis on the K419R mutant gene (Charng et al., 1994) and designated as K382R-K419R. The oligonucleotides used to create the mutants are shown in Figure 1. The plasmids recovered in the last step of the mutagenesis were screened by dideoxy sequencing (Sanger etal., 1977) in the regions of the desired mutations. Prior to expression of the mutant enzymes, the entire coding regions of these mutant alleles were sequenced to verify that there were no unintended mutations. Expression and Purification of the Wild-Type and Mutant Enzymes The wild-type and mutant genes were expressed in AC70R1-504 cells. Cell extracts were prepared by sonication. The enzymes were purified by ion-exchange chromatography on DEAE-Sepharose, FPLC chromatography on Mono Q, and Phenyl-Superose columns as described previously (Iglesias et al., 1991; Charng et al., 1992). Enzyme Assay (A) Assay 1. Enzymatic activity was measured in the pyrophosphorylase direction at 37 °C according to Preiss et al. (1966) during enzyme purification. The reaction mixtures contained 80 mM Hepes—NaOH buffer (pH 7.0), 2 mM ADP-glucose, 8 mM 57 Figure 1: Nucleotide sequence and encoded protein sequence of the Anabaena ADP-glucose pyrophosphorylase gene in the region of Lys382 (upper) and the synthetic oligonucleotides used for site-directed mutagenesis at position 382 (lower). The position 382 codons are underlined. Arg-382 Ala-382 Gln—382 Glu-382 5'- ATC 3'- TAG 5'- ATC 5'- ATC 5'- ATC 5'- ATC ATC TAG ATC ATC ATC ATC 58 GAT CTA GAT GAT GAT GAT Figure 1 HEW: O C) r-l C) O '-1 O > C) O > > GCC CGG GCC GCC GCC GCC CGC -3' GCG -5' CGC -3‘ CGC -3' CGC -3' CGC -3' 59 MgClz, 2 mM [32P]PP, (about 1500—3000 cpm/nmol), 4 mM 3PGA, 4 mM NaF, 50 ug of bovine serum albumin, and enzyme in a total volume of 250 uL. For the K382Q and K382R-K419R enzymes, 16 mM 3PGA was included in the reaction. (B) Assay II. Enzymatic activity in the ADP-glucose synthesis direction at 37 °C was measured according to the method of Preiss et al. (1966) to determine kinetic constants. For assay of the wild-type enzyme in the presence of activator, reaction mixtures contained 100 mM Hepes—NaOH buffer (pH 8.0), 0.5 mM [”C]glucose-1-P (about 1000 cpm/nmol), 2.5 mM ATP, 10 mM MgC12, 2.5 'mM 3PGA, 50 ng of bovine serum albumin, 0.15 unit of inorganic perphosphatase, and enzyme in a final volume of 200 ptL. For assay of the mutant enzymes, the reaction conditions were identical to wild- type, except that the amount of 3PGA was altered to obtain maximal activity. For the K382R, the K382A, the K382Q, and the K382R-K419R enzymes, 5, 20, 30, and 20 mM 3PGA were used, respectively. The assay of the wild-type enzyme in the absence of activator was the same as described above except that 3PGA was omitted and the amount of ATP was 5 mM in the reaction mixture. The amounts of ATP and [“C]glucose-l-P were altered for some mutant enzymes to obtain maximal activity. The amount of ATP was increased to 8 mM for K382Q and K382E enzymes. For the K382R, the K382A, and the K382R-K419R enzymes, 3 mM [“C]glucose-l-P was used, while 2 mM was used for K382E. Kinetic Characterization Kinetic data were plotted as initial velocity versus substrate or effector concentration. Saturating concentrations of substrates and effectors were determined to 60 ensure that maximal velocity was attained. Data were replotted as double-reciprocal plots to determine Vmax. Kinetic constants from hyperbolic plots were also determined by double-reciprocal plots. Sigmoidal plots were replotted as Hill plots to obtain kinetic constants. Interaction coefficients, n,,, were also determined by Hill plots. Kinetic constants were expressed as A0,, 80.5, and 10.5, which correspond to the concentration of activator, substrate, or inhibitor giving 50% of maximal activation, velocity, and inhibition, respectively. All the kinetic parameters calculated from double-reciprocal or Hill plots were in good agreement with those obtained by using a computer program (Canellas & Wedding, 1980) which performed nonlinear iterative least-squares fitting to the Hill equation. Protein Assay Protein concentration was determined by using bicinchoninic acid reagent (Smith et al., 1985) with bovine serum albumin as the standard. Protein Electrophoresis and Immunoblotting Polyacrylamide gel electrophoresis with SDS (SDS—PAGE) was performed according to Laemmli (1970). Following electrophoresis, proteins on the gel were visualized by staining with Coomassie Brilliant Blue R-250 or electroblotted onto a nitrocellulose membrane according to Bumette (1981). After electroblotting, nitrocellulose membranes were treated with affinity-purified anti-spinach leaf ADP-glucose pyrophosphorylase IgG (Morell et al., 1987), and the antigen—antibody complex was visualized via treatment with alkaline phosphatase linked goat anti-rabbit IgG followed by staining with BM purple AP-substrate precipitating reagent (from Boehringer Mannheim 61 GmbH). Molecular Mass Determination The molecular masses of the purified wild-type and mutant enzymes were determined on sucrose density gradients according to Martin and Ames (1961). The markers used were rabbit muscle lactate dehydrogenase (MW 144,000) and rabbit muscle pyruvate kinase (MW 237,000) (Worthington, 1988). Thermal Stability The purified enzymes were diluted to the same concentration, 0.2 mg/mL, in 50 mM Hepes—NaOH (pH 7.0) containing 1 mg/mL bovine serum albumin. The samples were heated for 5 min in a 60 °C water bath and then immediately placed on ice. The activities of these heated enzymes were assayed in the pyrophosphorolysis direction as described above. The activity of the K382E enzyme was assayed in the absence of the activator, 3PGA. Reductive Phosphopyridoxylation One hundred micrograms of the K382R-K419R enzyme in 50 mM Hepes—NaOH (pH 8.0) was incubated with 0.1 mM or 0.5 mM [3H]PLP (about 38,000 cpm/nmol) in a final volume of 1 mL in the dark at room temperature for 30 min. Then 100 uL of NaBIL was added to a final concentration of '49 mM to reduce the Schiff base formed between PLP and lysine residues. The reaction mixture was incubated in the dark at room temperature for another 60 min. The incorporation of [3H]PLP into protein was measured by determining trichloroacetic acid precipitable counts as described previously (Charng et al., 1994). RESULTS Expression and Purification of Lys’” Mutant Enzymes Normal expression of the mutant enzymes was confirmed by resolving the crude extract proteins with SDS—PAGE. Anabaena ADP-glucose pyrophosphorylase was identified by immunoblotting with antibody prepared against the spinach leaf ADP-glucose pyrophosphorylase that has been shown to be reactive with the Anabaena enzyme (Iglesias et al., 1991; Charng et al., 1992). All the mutant enzymes were expressed at a level similar to the wild-type enzyme based on the result of immunoblotting. The apparent sizes of these mutant enzymes were the same as that of the wild-type enzyme. The mutant enzymes were purified to greater than 90% homogeneity as estimated by SDS—PAGE of about 5 pg of protein. Kinetic Characterization of Lys382 Mutant Enzymes The apparent affinity for 3PGA decreased dramatically when Lys382 was replaced with either arginine, alanine, or glutamine (Table 1). The A05 values for 3PGA of the K382R, the K382A, and the K382Q enzymes were about 10-, 40-, and 160-fold higher than that of wild-type enzyme, respectively. The interaction coefficients were changed from 1.0 for the wild-type to 1.7 for the mutant enzymes. This suggested that the binding of the activator to the mutant enzymes was cooperative. The K382E enzyme was not activated by 3PGA, but rather was inhibited with an 105 value of 5.9 mM. This is in contrast with the K419E mutant enzyme (Charng et al., 1994) which is activated by 3PGA with an A,5 about 6.0 mM. In the absence of activator, the VM value of the K382A enzyme was slightly 62 63 Table 1: Kinetic Parameters for Activation of the Wild-Type and Mutant ADP-Glucose Pyrophosphorylases“ Vmu(—-—3PGA) Vm(+3PGA) A0_5(3PGA)(mM) (units/mg)” (units/mg) wild-type 0.050 1 0.005 (1.0) 6.9 i 0.3 60 i 4 K382R 0.53 i 0.05 (1.7) 0.58 :1; 0.04 61 j; 6 K382A 1.9 i 0.2 (1.7) 8.8 i 0.7 57 j: 5 K382Q 7.8 i 0.5 (1.7) 0.87 :1; 0.03 5.9 :1: 0.3 K382E 5.9 :1: 0.2 (1.2)‘ 0.22 :1: 0.01 - K419Rd 1.0 :t 0.1 (1.8) 0.8 :1; 0.1 103 i 11 K419Ed 6.0 :1: 0.8 (1.9) 0.4 :t 0.1 21 :1: 3 K382R-K419R 5.2 :I: 0.2 (2.1) 0.50 :1: 0.01 1.9 i 0.1 " Reactions were performed in the synthesis direction, assay 11 as described under Materials and Methods. Data represent the mean :1: standard deviation of two independent experiments. The values in parentheses are the Hill interaction coefficients. ” One unit of enzyme activity is expressed as the amount of enzyme required to form 1 pmol of ADP- glucose/min at 37 °C assayed in the synthesis direction as described under Materials and Methods. ‘ The K382E enzyme is inhibited by 3PGA. Therefore, the constant is 105 instead of A05. ‘ Cited from Charng et al. (1994). 64 higher, about 1.3-fold, than that of the wild-type enzyme, while the Vmax values of the K382R, the K382Q, and the K382E enzymes were about 8, l3, and 3% of the wild-type enzyme Vm, respectively. In the presence of saturating 3PGA, however, the Vm, values of the K382R and the K382A enzymes were similar to that of wild-type, while the Vmax value of the K382Q enzyme was about 10% of the wild-type V,,,,,. The degree of activation increased from 8.7-fold for the wild-type enzyme to 105-fold for the K382R enzyme. However, those for the K382A and the K382Q enzymes were about 6.5- and 6.8-fold, respectively, similar to that of the wild-type enzyme (T able. 1). The apparent affinities for the substrates (ATP, glucose-l-P, Mg“) the and inhibitor (P,) were all relatively less affected by the mutations at position 382 (Table 2), indicating that the conformations of these ligand-binding sites were essentially unchanged. The only significant changes observed were the 4- or 5-fold increase in the 50.5 for ATP of the K382Q enzyme in the absence or presence of activator, the 6- and 5-fold increases in the S0, for glucose-l-P of the K382R and the K382E enzymes in the absence of activator, respectively, and the 11-fold increase in the 105 value for P, of the K382E enzyme in the absence of activator (Table 2). These changes are smaller than the 10-, 40-, and 160-fold increases seen in the A05 value for the arginine, alanine, and glutamine mutants, respectively. For the wild-type enzyme, 2.5 mM 3PGA increased the 105 value for P, from 55 rm to 1.0 mM. With the apparent affinity relatively unchanged for P, but largely decreased for 3PGA, the K382R, the K382A, and the K382Q enzymes were more sensitive to P, inhibition by having lower [05 in the presence of 2.5 mM 3PGA, the amount which was saturating for the wild-type enzyme (Table 2). 65 A30: .3 8 med—U So 08 38.35 mavv— 05 can «gem 05 we 880888: 0.:qu 25. o .28 n. N 83 wow: 888.:wd-m mwo 85858088 08 8883wdtm Mo 3888 05 E 05 mg? 856 . 858508 88838 :5 05 88 8.85288 5 828.» 05. 558885 88:38qu 9.8 no .8890: 85.53» H 508 05 68882 San— .80502 :5 88582 Sun: 80288.. «a a .288 838:: $8588 08 8 8088.68 803 8088M .. 8.: 88 H 888 8.: 8 H 88 2.: 88 H :8 8.8 88 H 8.8 28 on 8.: 88 H 88 8.: 28 H 8.4 88: 88 H 28 8.: 38 H 88 88 882338802 8. : 88 H 2 8 8.: 2 H 2 8.: 888 H 888 8.: 88 H 88 28 on 2. :888H 888 88: 88 H S. 8.: 288 H 288 8.: 48 H N8 28 88:82 8 e 88 H 84.8 2...: 8 H 88 2.: 888 H 388 8.: 88 H R8 28 2 G: 88 H :8 8.: «8 H 28 8.: 88 H 28 8.: 28 H 5 88 .8232 2 :88 H 88 8.: 8 H «8 88: 88 H 88 88: 28 H 88 88 “.8802 8. : 88 H 2 8 8.8 28 H 3. 2.: 888 H 688 8.: 88 H 88 28 8m 8. :88 H a 8 88: 28 H 88 8.: 888 H $88 8.: 8 H 28 88 0882 8.8 so H 88 8.: 28 H 5.. 8.: 388 H ~88 8.: 88 H 88 :8 on 6:88 H 28 2.2: 28 H 88 88: 88 H 28 8.: 8 H 28 25: <89 8.8 88 H 9.8 8.: 28 H 88 2 : 388 H 388 8.: 88 H 88 28 n 2.: 88 H n 8 88: 28 H 88 8.: 88 H 38 8.: 28 H 3 88 8802 8988 H 8.2 8.3 8 H 88 88: 388 H 488 8.: 88 H :8 :8 an 8.:888H «88 88: 28 H to 8.8 288 H 988 8.: 28 H 4.2 88 25.83 28: .8 he .32 285 a»: 9:5 8-2-9.828 38: 8k <98 .8 3: 3m .8388er 3820.83 8:2 ea 25.83 on. to 83.88 8252 8 2.8. 66 Kinetic Characterization of the K382R-K419R Enzyme When both Lysm and Lys419 were replaced with arginine, the A0,, value for 3PGA was about lOO-fold higher than that of the wild-type enzyme, while those of the K382R and the K419R enzymes were 10- and 20-fold higher than that of the wild-type enzyme, respectively (Fable 1). The interaction coefficient was changed from 1.0 for the wild-type to 2.1 for this double mutant enzyme. The Vm value was about 7% of the wild-type VM in the absence of 3PGA and about 3% of that of the wild-type enzyme in the presence of 3PGA. Thus, catalytic efficiency was substantially decreased compared to the wild-type enzyme. The mutations at residue 382 and 419 did not cause much alteration in the apparent affinities for the substrates, ATP, glucose-l-P, Mg“, and inhibitor, Pi (Table 2). The only significant changes observed were the 4-fold increase in the S05 for ATP in the presence of 3PGA, the 7-fold increase in the So.5 for glucose-l-P in the absence of activator, and the 4-fold increase in the 10.5 for Pi in the absence of activator (Table 2). Effector Specificity Previous studies have shown that a mutant Anabaena enzyme, in which Lys419 was replaced by a glutamine, has an altered activator specificity (Charng et al. , 1995). The mutant enzyme is activated more effectively by FBP than by 3PGA at lower concentrations (Charng et al., 1995). It was of interest to examine whether the specificities for the allosteric effectors of the Lys382 mutant enzymes had changed. Several compounds, some of which are known as the major activators or inhibitors of other bacterial enzymes (Preiss, 1991), were used to test their effects on the mutants (Table 3). Some compounds have much different effects on the mutant enzymes than on the wild-type enzyme. But 67 3PGA still is the major activator for the K382R, the K382A, and the K382Q mutant enzymes, while Pi still is the most effective inhibitor for all mutant enzymes. However, the K382E enzyme cannot be activated by 3PGA, nor by any other effectors used. The wild-type enzyme is activated by 3PGA more effectively than by either FBP or AMP (Figure 2A). For the K382R-K419R enzyme, FBP and AMP activate the enzyme 5.4— and 8.3-fold at 5 mM, respectively (T able 3), and more effectively than 3PGA (Figure 2B), which has only a 2.7-fold activation (Table 3). The A05 values for FBP and AMP of the double-mutant enzyme are 0.35 and 2.0 mM, respectively, which are similar to those of the wild-type enzyme, 0.11 and 1.8 mM. Molecular Masses of the Wild-Type and Mutant Enzymes The molecular mass of the wild-type enzyme determined on a sucrose density gradient is 197 kDa. The calculated molecular mass based on the deduced amino acid sequence is 193 kDa. This result is also in reasonable agreement with the molecular mass, 225 kDa, determined by gel filtration chromatography (Iglesias et al., 1991) and verifies a tetrameric structure for ADP-glucose pyrophosphorylase from Anabaena. To ascertain whether the loss of enzyme activity or the change of kinetic constants was perhaps due to the instability of the tetrameric complex, each mutant was subjected to sucrose density gradient sedimentation. The sedimentation patterns for the mutant enzymes were identical to that of the wild-type enzyme with a single peak where the apparent molecular masses ranged from 196 to 201 kDa, indicating that all the mutant enzymes retain the tetrameric StfllCtlli'C . 68 Table 3: Specificity of Allosteric Effectors of the Wild-Type and Mutant ADP-Glucose Pyrophosphorylases” ADP-glucose formed (nmol/10 min) effector wild—type K382R K382A K382Q K382E. K382R-K419R none 0.58 0.20 0.97 0.82 1.99 1.12 effector concn relative activity effector (mM) wild-type K382R K382A K382Q K382E K382R-K419R none 1.0 1.0 1.0 1.0 1.0 1.0 3PGA 2 9.8 5 10.7 89.5 6.4 3.1 0.6 2.7 fructose-6-P 2 5.7 5 6.5 7.5 1.9 1.7 1.0 1.1 fructose-1 ,6-P2 2 l .8 5 1.9 1.9 0.2 0.9 0.7 5.4 glucose—6-P 2 3.8 5 5.3 2.9 1.3 1.9 0.8 0.9 glucose-1,6-P2 2 0.4 5 0.4 1.8 0.2 0.2 0.9 4.6 pyruvate 2 1.0 5 1.1 1.3 0.9 2.3 0.9 1.0 PLP 2 2.0 5 2.0 16.8 0.3 0.2 0.1 1.9 P-enolpyruvate 2 6. 1 5 6.8 10.0 0.5 0.7 0.8 2.6 NADPH 0.5 0.8 1.0 1.0 0.9 0.8 1.0 2 0.4 0.9 0.6 0.7 0.7 0.9 hexane-1,6—diol-P2 0.5 0.6 0.5 0.2 0.5 0.5 0.7 2 0.5 0.3 0.1 0.2 0.3 0.6 ADP 2 0.9 1.8 0.6 1.6 0.8 2.1 5 0.5 0.3 0.3 0.4 0.2 1.2 AMP 2 2.5 4.3 0.7 2.0 0.9 5.9 5 3.3 1.6 0.7 1.3 0.5 8.3 Pi 2 0.1 0.3 0.1 0.2 0.1 0.3 ‘ Reactions were performed in the synthesis direction, assay 11 as described under Materials and Methods, with the presence of effectors as indicated. Data represent the average of two duplications with less than 10% deviation. The amounts of the wild-type, the K382R, the K382A, the K382Q, the K3825, and the K382R-K419R enzymes used were 0.01, 0.04, 0.01, 0.15, 0.90, and 0.24 pg, respectively. 69 Figure 2: Activation of the wild-type (A) and K382R-K419R (B) enzymes by 3PGA (O), FBP (O), and AMP (I). Initial velocities of the enzymes were determined in the synthesis direction, assay 11, as described under Materials and Methods with the concentration of activators being varied. The amounts of the wild-type and the K382R- K419R enzymes used were about 0.02 and 0.48 pg, respectively. 70 A A A O . 6 2 8 4 0 1. «I EEO—Sec... €253 oaooaado< [Activator], mM . . O 6 2 8 4 0 2 4| 4| EEoEoEc 6252 33336.0( 20 16 12 [Activator], mM Figure 2 71 Thermal Stability of Lys382 Mutant Enzymes After heat treatment at 60 °C for 5 min, the activity of the wild-type enzyme remained unchanged, while the K382R, the K382A, the K382Q, and the K382E enzymes retained 94, 84, 84, and 85 % activity, respectively. The result indicates that Lys382 is not required for the stability of the enzyme. The K382R-K419R enzyme retained about 78 % activity after the same heat treatment. Reductive Phosphopyridoxylation of the K382R-K419R Mutant Enzyme Reductive phosphopyridoxylation of the K382R-K419R enzyme was performed to determine if additional lysine residues could be labeled. The double-mutant enzyme was activated about 2-fold by PLP, which was similar to that of the wild-type enzyme (Charng et al., 1994). The A05 value of 0.20 mM was about 200-fold higher than that of the wild- type enzyme (Charng et a1. , 1994). Unlike the wild-type and the K419R enzymes (Chamg et al. , 1994), reductive phosphopyridoxylation had no effect on either enzyme activity or 3PGA activation of the K382R-K419R enzyme. In the presence of 0.1 or 0.5 mM [3H]PLP, no incorporation of PLP into the mutant enzyme was found. Thus, no other lysine residue appears to be involved in PLP binding. DISCUSSION Previous studies of Anabaena ADP-glucose pyrophosphorylase have shown that LysM9 is located within the activator-binding site (Charng et al., 1994). Another lysyl residue, Lysm, was modified by PLP when Lys419 was replaced by arginine (Charng et al. , 1994), suggesting that Lys382 is also part of the activator-binding site. Alignment of all the amino acid sequences of ADP-glucose pyrophosphorylases available has shown that Lys382 is conserved in the cyanobacterial enzymes, in the small subunit of the higher-plant enzymes, and in the large subunit of the spinach leaf, wheat leaf, and potato tuber enzymes (Smith-White & Preiss, 1992). Site-directed mutagenesis experiments have been performed to probe and verify the function of Lys382 of Anabaena ADP-glucose pyrophosphorylase. The large effects on the A05 values for 3PGA when Lys382 was replaced by other amino acids and the inability of 3PGA to activate the glutamic acid mutant enzyme are consistent with the view that Lys382 is involved in 3PGA binding. As substitutions of Lys382 go from basic to neutral, the apparent affinities for 3PGA decrease. When the substituted amino acid is acidic, the mutant enzyme cannot be activated by 3PGA, suggesting that 3PGA cannot bind to the 3PGA-binding site. Thus, the cationic property seems to be the most important factor in 3PGA binding at position 382. The arginine mutant enzyme has about a 10-fold lower apparent affinity for 3PGA, indicating that charge alone is insufficient for the proper binding of 3PGA. The glutamine mutant enzyme has a much lower apparent affinity for 3PGA than the alanine mutant enzyme though both glutamine and alanine are neutral amino acids. Therefore, the size of the amino acid may also be important due to steric interference with proper binding of the 72 73 activator, 3PGA. Interestingly, the K382E enzyme cannot be activated by 3PGA but rather is inhibited. It seemed that 3PGA bound to the activator site induces an abnormal conformational change when Lys382 was replaced by glutamic acid. But this is not the case at position 419. The K419E enzyme still can be activated by 3PGA for about 50-fold with an A05 value of 6.0 mM (Charng et al., 1994). Somehow, the enzyme can tolerate the negative charge at position 419 for the binding of 3PGA. All the mutant enzymes retain at least 80 % activities after heat treatment. Lys382 is obviously not critical to the stability of the native folded state. The kinetic constants for ATP, Mg“, and glucose-l-P are relatively unaffected for the five mutants compared to the dramatic changes in A05, showing the tolerance to amino acid substitution at position 382. The mutant enzymes are tetrameric in structure as is the wild-type enzyme, also indicating that Lys382 is not essential for maintaining the quaternary structure of the enzyme. Based on previous studies (Charng et a1. , 1994) and our results, we can conclude that both Lys382 and Lys419 are involved in the binding of 3PGA. Both charge and size of these two lysyl residues are important for the binding of activator. However, the charge seems to be more important, and this may be explained by the interactions between the two positively charged e—amino groups and the negatively charged carboxyl and phosphate groups of the activator, 3PGA. One possibility is that one of the two lysyl residues interacts with the carboxyl group and the other one with the phosphate group. The two negatively charged groups of 3PGA may interact with the two lysyl residues on the same 74 subunit or on two different subunits. In either case, the two lysyl residues that bind to the same 3PGA molecule must be very close in three-dimensional structure. The affinities for P, of these mutant enzymes are relatively unaffected compared to the dramatic changes of affinities for 3PGA. This suggests that P, has a different binding site, but the P, binding site may overlap with the 3PGA binding site and this may explain why increasing concentrations of 3PGA reverses P, inhibition and that P, at higher concentrations can reverse 3PGA activation (Charng et a1. , 1992, 1994). One amino acid substitution at position 382 changes the effects of almost all analogues of activator, 3PGA, on the Anabaena ADP-glucose pyrophosphorylase, suggesting that these analogues bind at the same site as 3PGA to activate the enzyme. Previous studies also have shown that FBP (Ghosh & Preiss, 1966; Charng et al., 1995) and P-enolpyruvate (Ghosh & Preiss, 1966) bind to the same site as 3PGA. The fact that no effectors activate the K382E enzyme can be explained by the electrostatic repulsion between anionic ligands and enzyme. FBP and AMP become more effective as activators than 3PGA for the K382R- K419R enzyme. The alteration of the activator specificity was also observed for the K419Q Anabaena mutant enzyme (Charng et al., 1995). For the K419Q enzyme, FBP is about as effective as 3PGA and activates the enzyme more than 3PGA at lower concentrations. For the K382R-K419R enzyme, FBP and AMP have much higher affinities than 3PGA and activate the enzyme more than 3PGA even at higher concentrations. It should be noted that FBP and AMP are the activator and inhibitor, respectively, of the E. coli ADP-glucose pyrophosphorylase. Both residues in the E. coli enzyme, corresponding 75 to Lys382 and Lys“19 in Anabaena enzyme, are arginine instead of lysine (Charng et al., 1992). It is suggestive that the differences of activator specificity relate to the difference of amino acids at the activator-binding site of the E. coli and Anabaena/higher-plant enzymes. REFERENCES Ahrens, H., & Kortnyk, W. (1969) Anal. Biochem. 30, 413-420. Aitken, A. (1988) Methods Enzymol. 167, 145-154. Ball, K., & Preiss, J. (1994) J. Biol. Chem. 269, 24706-24711. Bumette, W. W. (1981) Anal. Biochem. 112, 195-203. Canellas, P. F., & Wedding, R. T. (1980) Arch. Biochem. Biophys. 199, 259-264. Carlson, C. A., Parsons, T. F., & Preiss, J. (1976) J. Biol. Chem. 251, 7886-7892. Charng, Y. Kakefuda, G., Iglesias, A. A, Buikema, W. J. & Preiss, J. (1992) Plant Mol. Biol. 20, 37-47. Charng, Y., Iglesias, A. A., & Preiss, J. (1994)]. Biol. Chem. 269, 24107-24113. Charng, Y., Sheng, J ., & Preiss, J. (1995) Arch. Biochem. Biophys. 318, 476-480. Gardiol, A., & Preiss, J. (1990) Arch. Biochem. Biophys. 280, 175-180. Ghosh, H. P., & Preiss, J. (1965) J. Biol. Chem. 240, 960-961. Ghosh, H. P., & Preiss, J. (1966) J. Biol. Chem. 241, 4491-4504. Haugen, T. H., Ishaque, A., & Preiss, J. (1976) J. Biol. Chem. 251, 7880-7885. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1991) Plant Physiol. 97, 1187-1195. Laemmli, U. K. (1970) Nature 227, 680-685. Levi, C., & Preiss, J. (1976) Plant Physiol. 58, 753-756. Martin, R. G., & Ames, B. N. (1961) J. Biol. Chem. 236, 1372-1379. Morell, M., Bloom, M., Knowles, V., & Preiss, J. (1987) Plant Physiol. 85, 182-187. Morell, M., Bloom, M., & Preiss, J. (1988) J. Biol. Chem. 263, 633-637. Parsons, T., & Preiss, J. (1978) J. Biol. Chem. 253, 6197-6202. Preiss, J ., Shen, L., Greenberg, E., & Gentner, N. (1966) Biochemistry 5, 1833-1845. 76 77 Preiss, J ., & Romeo, T. (1989) Adv. Microb. Physiol. 30, 183-238. Preiss, J. (1991) in Plant Molecularand Cell Biology (Mifflin, B., ed.) Vol. 7, pp 59-114, Oxford University, Oxford. Preiss, J. (1996) Biotech. Annu. Rev. 2, 259-279. Preiss, J, & Sivak, M. (1996) in Photoassimilate Distribution in Plants and Crops (Zamski, E., & Schaffer, A. A., eds.) pp 63—96, Marcel Dekker, Inc., New York. Sambrook, J ., Fritsch, E. F., & Maniatis, T. (1982) Molecular Cloning: A Laboratory Manual, 2nd Ed., Cold Spring Harbor Laboratory, Cold Spring Harbor, NY. Sanger, F., Nicklen, S., & Coulson, A. R. (1977) Proc. Natl. Acad. Sci. U.S.A. 74, 5463-5467. Sanwal, G. G., Greenberg, E., Hardie, J., Cameron, E. C., & Preiss, J. (1968) Plant Physiol. 43, 417-427. Sayers, J. R., Schmidt, W., & Eckstein, F. (1988) Nucleic. Acids Res. 16, 791-802. Shively, J. M. (1988) Methods Enzymol. 167, 195-203. Smith, P. K., Krohn, R. I., Hermanson, G. T., Mallia, A. K., Gartner, F. H., Provenzano, M. D., Fujimoto, E. K., Geoke, N. M., Olson, B. J., & Klenk, D. C. (1985) Anal. Biochem. 150, 76-85. Smith-White, B. J ., & Preiss, J. (1992) J. Mol. Evol. 34, 449-464. Stock, A., Ortanderl, F., & Pfleiderer, G. (1966) Biochem. Z. 344, 353-360. Worthington, C. C. , (1988) Worthington Enzyme Manual: Enzymes and Related Biochemicals, Worthington Biochemical Corporation, Freehold, NJ. CHAPTER III INVOLVEMENT OF ARGININE AT THE ALLOSTERIC SITES OF ADP- GLUCOSE PYROPHOSPHORYLASE FROM ANABAENA PCC 7120 78 ABSTRACT Treatment of ADP-glucose pyrophosphorylase from the cyanobacterium Anabaena PCC 7120 with phenylglyoxal in 50 mM Hepes, pH 8.0, at 25 °C resulted in a time- and concentration-dependent loss of enzyme activity. P,, the inhibitor, protected the enzyme from inactivation most effectively, while 3PGA, FBP, PLP, and ATP plus magnesium were also good protectors. After incubation with 2 mM phenylglyoxal for 1 hour, the modified enzyme had a 10-fold lower apparent affinity for P, in the absence of the activator, 3PGA, than that of the wild-type enzyme. The apparent affinity for 3PGA of the modified enzyme was 4-fold lower than that of the wild-type enzyme. Thus, phenylglyoxal appears to interfere with the allosteric regulation, as well as the catalysis, of ADP-glucose pyrophosphorylase from Anabaena. 79 INTRODUCTION ADP-glucose pyrophosphorylase is a key enzyme in the biosynthesis of glycogen in bacteria (Preiss & Romeo, 1989; Preiss, 1996) and starch in plants (Preiss, 1991; Preiss, 1996; Preiss & Sivak, 1996). This enzyme catalyzes the conversion of glucose-l-P and ATP to ADP-glucose and PPi (Shen & Preiss, 1964). ADP-glucose pyrophosphorylase is subject to allosteric regulation (Preiss & Romeo, 1989; Preiss, 1996). There are significant differences between the bacterial and plant enzyme. The enzyme from enteric bacteria is activated by FBP and inhibited by AMP (Preiss et al., 1966), whereas the enzyme from higher plants is mainly activated by 3PGA and inhibited by P, (Ghosh & Preiss, 1965; Sanwal et al., 1968). In addition, the bacterial enzyme is homotetrameric in structure (Haugen et a1. , 1976), while the higher- plant enzyme is heterotetrameric with two different subunits (Preiss, 1991). Cyanobacteria have metabolic properties similar to chloroplasts of higher plants (Aitken, 1988), but synthesize glycogen as the major carbohydrate reserve (Shively, 1988), in a similar manner to what is observed in bacteria. The cyanobacterial ADP- glucose pyrophosphorylase is regulated by 3PGA and P, like the hi gher-plant enzyme (Levi & Preiss, 1976; Iglesias et al., 1991), but is homotetrameric similar to the E. coli enzyme (Iglesias et al., 1991). Due to the properties of the cyanobacterial enzyme and the key position of these photosynthetic prokaryotes during evolution (Aitken, 1988), it is of great interest to understand the structure—function relationships of the allosteric site of the cyanobacterial ADP-glucose pyrophosphorylase. Previous chemical modification with PLP and site-directed mutagenesis studies 80 81 have shown that the activator-binding site of the E. coli ADP-glucose pyrophosphorylase, Lys”, is near the N-terminus (Parsons & Preiss, 1978; Gardiol & Preiss, 1990), while 419 those of the cyanobacterial enzyme from Anabaena, Lys382 and Lys , are close to the C- terrninus (Sheng et al., 1996; Charng et al., 1994). Studies on the spinach leaf enzyme 419 ' have shown that Lys“), which corresponds to Lys 1n Anabaena, is involved in binding of the activator, 3PGA (Morell et al., 1988; Ball & Preiss, 1994). Studies on E. coli, synechocystis, and spinach leaf enzymes using an arginine- specific reagent, phenylglyoxal, suggested that there is also an essential arginine residue involved in the allosteric activation and inhibition (Carlson & Preiss, 1982; Iglesias et al. , 1992; Ball & Preiss, 1992). ADP-glucose pyrophosphorylase from cyanobacterium Anabaena PCC 7120 has been purified to homogeneity. The present study shows that the Anabaena enzyme is inactivated by phenylglyoxal and that the inactivation can be prevented by the presence of the activator 3PGA, the inhibitor P,, or the substrate ATP plus magnesium. The modified enzyme is less sensitive to 3PGA activation and Pi inhibition. Stoichiometry studies show that there is one arginine residue modified per enzyme subunit during inactivation. MATERIALS AND METHODS Reagents Phenylglyoxal was purchased from Sigma. Stock solutions were prepared in 50 mM Hepes-NaOH (pH 8.0). [7-”C]phenylglyoxal (27.0 mCi/mmol) was from Amersham. [”C]Glucose-l-P was from ICN Pharmaceuticals Inc. All other reagents were of the highest quality available. Expression and Purification of ADP-Glucose Pyrophosphorylase E. coli mutant strain AC70R1-504, which is deficient in ADP-glucose pyrophosphorylase activity (Carlson et al., 1976), was used for expression of the Anabaena ADP-glucose pyrophosphorylase gene (Charng et al., 1992) and grown in enriched medium containing 1.1% KZHPO4, 0.85% KH2P04, 0.6% yeast extract, and 0.2% glucose, pH 7.0. The enzyme was purified as previously described (Iglesias et al., 1991; Charng et al., 1992) with some modification. Frozen cells were thawed into the extraction buffer containing 20 mM K—P,, pH 7.5, 5 mM DTT, 1 mL EDTA (about 5 mL of buffer/g of cells). The cell paste was sonicated for four 1 min intervals (1 min on ice between two sonications). The homogenate was centrifuged at 10,000 rpm for 10 min. The supernatant was absorbed onto a DEAE-sepharose Fast-Flow column (about 0.03 mL bed volume/mg of protein) that had been equilibrated with 20 mM K—P, buffer, pH 7.5, containing 2 mM DTT. After washing, the enzyme was eluted with a linear gradient consisting of 5-bed volumes of the above buffer in the mixing chamber and S-bed volumes of 50 mM K-P,, pH 6.0, containing 2 mM D'l'I‘ and 0.3 M KCl in the reservoir chamber. Fractions of 9 mL were collected and those containing activity were pooled. 82 83 Ammonium sulfate was added to the sample with stirring to a final concentration of 50% saturation. After centrifugation at 8000 rpm for 20 min, the pellet was put on ice for 30 min and resuspended in 1.2 M ammonium sulfate, pH 7.5 (about 0.1 mL/mg of protein). After centrifugation, the new pellet was dissolved in Buffer A (about 0.5 mL/mg of protein) for mono Q (20 mM Bis-Tris-propane, 5 mM K-P,, 1 mM EDTA, 10% sucrose (w/v), 2 mM DTT, pH 7.0.) and dialyzed against the same buffer. After dialysis, the sample was applied to a Mono Q HR16/ 10 column equilibrated with buffer A. The column was washed with 100 mL of Buffer A and eluted with two linear KC] gradients (50 mL, 0-0.1 M; 200 mL, 0.1-0.3 M) in Buffer A. Fractions of 5 mL were collected and those containing activity were pooled. The pooled sample was concentrated to a final concentration of about 2 mg/mL in an Amicon concentrator fitted with a YM-30 membrane. Enzyme Assay Enzymatic activity was measured in the ADP-glucose synthesis direction at 37 °C according to the method of Preiss et al. (1966). (A) Activated conditions: The synthesis of ADP-[“C]glucose from [”C]glucose-1-P and ATP was measured in the presence of activator, 3PGA. The reaction mixture contained 100 mM Hepes—NaOH buffer (pH 8.0), 0.5 mM [”C]glucose-l-P (about 1000 cpm/nmol), 2.5 mM ATP, 10 mM MgC12, 2.5 mM 3PGA, 50 ug of bovine serum albumin, 0.15 unit of inorganic pyrophosphatase, and enzyme in a final volume of 200 pL. (B) Unactivated conditions: The assay in the absence of activator was the same as described above except that 3PGA was omitted and the amount of ATP was 5 mM in the reaction mixture. 84 Kinetic Characterization Kinetic data were plotted as initial velocity versus substrate or effector concentration. Saturating concentrations of substrates and effectors were determined to ensure that maximal velocity was attained. Kinetic constants from hyperbolic plots were determined by double-reciprocal plots. Sigmoidal plots were replotted as Hill plots to obtain kinetic constants. Kinetic constants were expressed as A0, and 105, which correspond to the concentration of activator and inhibitor giving 50 % of maximal activation and inhibition, respectively. All the kinetic parameters calculated from double- reciprocal or Hill plots were in good agreement with those obtained by using a computer program (Canellas & Wedding, 1980) which performed nonlinear iterative least-squares fitting to the Hill equation. Modification of ADP-Glucose Pyrophosphorylase with Phenylglyoxal ADP-glucose pyrophosphorylase (0.5 uM) in 50 mM Hepes-NaOH (pH 8.0) was incubated in the dark at room temperature with different concentrations of phenylglyoxal, as indicated in the figure legends. The reactions were started by the addition of enzyme and stopped by the addition of 10 mM arginine. In protection experiments, substrates or allosteric effectors were added before the addition of enzyme. When necessary, the reaction mixtures were passed through Sephadex G-50 before assaying for activity. Incorporation of [“Clphenylglyoxal into ADP-Glucose Pyrophosphorylase One hundred micrograms of ADP-glucose pyrophosphorylase in 50 mM Hepes—NaOH (pH 8.0) was incubated with 4 mM [”C]phenylglyoxal (11,000 cpm/nmol) in a volume of 100 uL. The incubation was carried out in the dark at room temperature. 85 At different times, aliquots of 10 al. were removed, mixed with 2 uL of 100 mM arginine and after 5 min put on ice. The volume of each sample was brought to 100 uL with 50 mM Hepes-NaOH (pH 8.0) and 10 aL was withdrawn to check the enzyme activity. One mL of cold 10% (w/v) trichloroacetic acid was added to the remaining 90 uL of sample followed by 1 h of incubation on ice. The precipitate was collected by centrifugation at 12,000 rpm for 15 min, washed twice with 1 mL of cold 10% trichloroacetic acid, and dissolved in 400 aL of 3% NaZCO3 in 0.1 N NaOH. Two hundred pl. of this sample was used to measure the protein concentration with a micro bicinchoninic acid reagent kit from Pierce. The radioactivity was determined by counting the other 200 aL of the sample in 5 mL of Safety Solve. RESULTS Inactivation of ADP-Glucose Pyrophosphorylase by Phenylglyoxal ADP-glucose pyrophosphorylase was inactivated by phenylglyoxal when the incubation was carried out in Hepes buffer (pH 8.0) at room temperature over a 2 hours period (Figure l). The 3PGA-activated and unactivated activity were affected by phenylglyoxal in a similar pattern. Loss of enzyme activity with time was measured at various concentrations of phenylglyoxal. The inactivations of both activated and unactivated enzyme activity followed pseudo-first-order kinetics (Figure 2A and 3A). The orders of inactivation were determined (Figure 2B and 3B) according to Levy et al., 1963. The n values were 0.95 and 0.90, respectively. This suggested that the loss of enzyme activity resulted from at least one molecule of phenylglyoxal binding per enzyme subunit. Protection of Enzyme Activity by Substrates and Allosteric Effectors The effects of substrates and allosteric effectors on the inactivation by phenylglyoxal are seen in Table 1. The protection against inactivation was indicated by the increase in the time required for half-inactivation. For both 3PGA-activated and unactivated activity, P, proved to be the most effective at protecting the enzyme from inactivation by phenylglyoxal, while 3PGA and other activators, fructose-1,6-bisphosphate and PLP, were also very good protectors. Substrate ATP plus Mg2+ had a similar protection against inactivation as 3PGA, while PP, had a lesser effect. All the other substrates had no or very little protection against the inactivation by phenylglyoxal. 86 87 Figure 1: Inhibition of Anabaena ADP-glucose pyrophosphorylase by phenylglyoxal. Enzyme (0.5 aM) was incubated for 2 h with phenylglyoxal. Activity was measured in the presence (O) and absence (O) of 2.5 mM 3PGA (Assay II) and 100 % activity was 8.6 and 6.4 nmol of ADP-glucose formed per 10 min, respectively. The data represents the mean of two separate experiments. 88 % Initial actuvnty s. a) . .° . ‘P 0 . , 0.0 1.0 ' 2:0 370 4:0 7 5:0 [Phenylglyoxal], mM Figure 1 89 Figure 2: Kinetics of inhibition of ADP-glucose pyrophosphorylase activity by phenylglyoxal. (A) The loss of activity with time when enzyme (0.5 uM) was incubated with the following concentrations of phenylglyoxal: 0, 0.4, 0.8, 1.5, and 2.4 mM. The activity was measured under activated conditions (Assay 11A) and 100 % activity was 8.9 nmol of ADP-glucose formed per 10 min. The data represents the mean of two separate experiments. (B) Determination of the order of the reaction. N 2&5 SE .mEC. ON? cor om 00 CV ON 0 — p p . _ s p . _ . - S... .383: 3.. r E vow . do «we ova No. 1.9. I. OP r'fio rfiu. . r v q N and... ' 5 in m] 601 2.5... m 1:. E md E Yo l.. o o o « . 02 < MARS? new % 91 Figure 3: Kinetics of inhibition of ADP-glucose pyrophosphorylase activity by phenylglyoxal. (A) The loss of activity with time when enzyme (0.5 PM) was incubated with the following concentrations of phenylglyoxal: 0, 0.4, 0.8, 1.5, and 2.4 mM. The activity was measured under unactivated conditions (Assay HE) and 100 % activity was 6.9 nmol of ADP-glucose formed per 10 min. The data represents the mean of two separate experiments. (B) Determination of the order of the reaction. m 23E SE .mEz. 92 cm? 2.: o.m o.m ow o.N o n n F - vo—ZENNnx no. N 03v n ..... o - .9, q ... nor E YN . a sum cad-... N ......n. .. 5:: m... ...- . m 1' I 2E wd . .28 v.0 . o ” o b I 2: I‘M/“139 IBM“! % 93 Table 1: Effect of substrates and allosteric effectors on modification of Anabaena ADP- glucose pyrophosphorylase by phenylglyoxal substrate/effector concentration (mM) to, (min) + 3PGA - 3PGA none - 36 39 MgCl2 2 41 41 ATP + 10 mM MgCl2 2 240 375 ADP-glucose + 10 mM MgCl2 2 85 70 glucose-l-P 2 50 60 PP, 2 180 . 70 P, 2 370 700 3-P—glycerate 2 220 300 fructose-1 ,6-bisphosphate 2 220 220 pyridoxal—P 0.05 210 145 Enzyme was incubated with 2 mM phenylglyoxal in the presence of different ligands at the stated concentrations. Aliquots were removed at different times and the modifications were stopped with 10 mM arginine. Activity of the modified enzyme was assayed in the presence and absence of 3PGA. 94 Effect of Modification with Phenylglyoxal on Enzyme Kinetics Modification with phenylglyoxal made the enzyme less sensitive to P, to inhibition (Fig. 4). When the enzyme was assayed in the presence of 3PGA, the 10,, increased from 0.93 mM for the unmodified enzyme to 1.7 mM for the enzyme incubated with 2 mM phenylglyoxal for 1 h (Fig. 4A). When the enzyme was assayed in the absence of activator, the modified enzyme had an 10, about 0.27 mM, which was 10-fold higher than that of the unmodified enzyme, 0.026 mM (Fig. 4B). Inactivation of the enzyme with 1 mM and 2 mM phenylglyoxal for 1 h resulted in an enzyme with A05 value for 3PGA of 0.12 mM and 0.22 mM, respectively, 2- and 4-fold higher than that of the unmodified enzyme, 0.058 mM (data not shown). Stoichiometry of Modification Stoichiometry of the modification by phenylglyoxal was examined by measuring the [”C]phenylglyoxal incorporation during enzyme inactivation. Figure 5 shows the incorporation of phenylglyoxal into the enzyme followed a linear correlation with the loss of enzyme activity. At full inactivation, about 8 mol of phenylglyoxal was incorporated per mol of enzyme. Since phenylglyoxal reacts with arginine in a 2:1 stoichiometry (Takahashi, 1968), and the enzyme is a tetramer, this corresponds to 1 arginine residue modified per subunit. 95 Figure 4: Inhibition of ADP-glucose pyrophosphorylase by P,. The enzyme was incubated for l h in the absence (O) or presence (O) of 2 mM phenylglyoxal. After incubation, enzyme activity was assayed under activated (A) or unactivated (B) conditions (Assay II) and 100 % activity was 9.1 and 8.3 nmol of ADP-glucose formed per 10 min, respectively. The data represents the mean of two separate experiments. 96 [P1] ,0 mM . . . . . 0 0 0 8 6 4. 33:8 $25 ea 20- 3:8 325 a. [Pl], mM Figure 4 97 Figure 5: Incorporation of [”C]phenylglyoxal into ADP-glucose pyrophosphorylase. Enzyme was incubated with 4 mM [”C]phenylglyoxal for different times. Aliquots were assayed for incorporated radioactivity and 3PGA activated activity (Assay 11A). The data represents the mean of two separate experiments and 100 % activity was 14.3 nmol of ADP-glucose formed per 10 min. % initial activity 98 . , . , . , . 1 0.0 2.0 4.0 6.0 8.0 14 mol [ Clphenylglyoxal I mol enzyme Figure 5 DISCUSSION Guanidinium groups are frequently involved in the binding of anionic ligands to proteins (Takahashi, 1968; Riordan, 1979). They are suited to bind phosphate groups due to their planar structure and hydrogen-bonding capability (Cotton et al., 1973; Riordan, 1979). The fact that 3PGA and P, are the physiological allosteric regulators of ADP- glucose pyrophosphorylase and the previous chemical modification studies with phenylglyoxal (Carlson & Preiss, 1982; Iglesias et al., 1992; Ball & Preiss, 1992) suggest that the allosteric region of the enzyme might contain arginine residue(s) involved in the binding of the effectors. ' Anabaena ADP-glucose pyrophosphorylase was inactivated by phenylglyoxal. At full inactivation, one arginine residue was modified per enzyme subunit. The activator 3PGA protected the enzyme from inactivation. However, the inhibitor P, proved to be more effective at protecting the enzyme than 3PGA. In addition, substrate ATP plus Mg“ also protected the enzyme from inactivation, which was not seen in previous studies with other ADP-glucose pyrophosphorylases (Carlson & Preiss, 1982; Ball & Preiss, 1992; Iglesias et al., 1992). Kinetic studies indicate that modification of the Anabaena enzyme by phenylglyoxal was not simply inactivating catalysis but also resulted from interference with normal allosteric regulation. The modification decreased the enzyme's sensitivity to inhibition by P, and activation by 3PGA. These results suggest that guanidinium groups are in the vicinity of the regulatory site and that the activator 3PGA and the inhibitor P, binding sites are overlapping or they are connected by a conformational change, where binding at one 99 100 site affected binding at another site. Although the enzymes from E. coli (Carlson & Preiss, 1982) and Synechocystis (Iglesias et al., 1992) were also inactivated by phenylglyoxal, enzyme activity in the absence of activator was not affected much by phenylglyoxal. The Anabaena and spinach leaf enzymes showed loss of both activated and unactivated activity when modified by phenylglyoxal. These studies suggest that either modification with phenylglyoxal affects the enzyme's catalysis or phenylglyoxal could mimic P, as the allosteric inhibitor. Further site- directed mutagenesis (random or alanine scanning) study might be useful to locate the position(s) of the arginine residue(s) involved in the allosteric regulation. REFERENCES Aitken, A. (1988) Methods Enzymol. 167, 145-154. Ball, K., & Preiss, J. (1992) J. Protein Chem. 11, 231-238. Ball, K., & Preiss, J. (1994) J. Biol. Chem. 269, 24706-24711. Canellas, P. F., & Wedding, R. T. (1980) Arch. Biochem. Biophys. 199, 259-264. Carlson, C. A., Parsons, T. F., & Preiss, J. (1976) J. Biol. Chem. 251 , 7886-7892. Carlson, C.A., & Preiss, J. (1982) Biochemistry 21, 1929-1934. Charng, Y., Kakefuda, G, Iglesias, A. A., Buikema, W. J, & Preiss, J. (1992) Plant Mol. Biol. 20, 37—47. Charng, Y., Iglesias, A. A., & Preiss, J. (1994)]. Biol. Chem. 269, 24107-24113. Cotton, F. A., Hazen, E. E., Day, V. W., Larsen, 8., Norman, J. G., Wong, S. T. K., & Johnson, K. H. (1973) J. Am. Chem. Soc. 95, 2367-2369. Gardiol, A., & Preiss, J. (1990) Arch. Biochem. Biophys. 280, 175-180. Ghosh, H. P., & Preiss, J. (1965) J. Biol. Chem. 240, 960-961. Haugen, T. H., Ishaque, A., & Preiss, J. (1976) J. Biol. Chem. 251, 7880-7885. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1991) Plant Physiol. 97, 1187-1195. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1992) J. Protein Chem. .11, 119-128. Levi, C., & Preiss, J. (1976) Plant Physiol. 58, 753-756. Morell, M., Bloom, M., & Preiss, J. (1988) J. Biol. Chem. 263, 633-637. Parsons, T., & Preiss, J. (1978) J. Biol. Chem. 253, 6197-6202. Preiss, J ., Shen, L., Greenberg, E., & Gentner, N. (1966) Biochemistry 5, 1833-1845. Preiss, J ., & Romeo, T. (1989) Adv. Microb. Physiol. 30, 183—238. Preiss, J. (1991) in Plant Molecular and Cell Biology (Mifflin, B., ed.) Vol. 7, pp 59-114, Oxford University, Oxford. 101 102 Preiss, J. (1996) Biotech. Annu. Rev. 2, 259-279. Preiss, J, & Sivak, M. (1996) in Photoassimilate Distribution in Plants and Crops (Zamski, E., Schaffer, A. A., ed.) pp 63-96, Marcel Dekker, Inc., New York. Riordan, J. F. (1979) Mol. Cell. Biochem. 26, 71-92. Sanwal, G. G., Greenberg, E., Hardie, J., Cameron, E. C., & Preiss, J. (1968) Plant Physiol. 43, 417-427. Shen, L., & Preiss, J. (1964) Biochem. Biophys. Res. Commun. 17, 424-429. Sheng, J., Charng, Y., & Preiss, J. (1996) Biochemistry 35, 3115-3121. Shively, J. M. (1988) Methods Enzymol. 167, 195-203. Takahashi, K. (1968) J. Biol. Chem. 243, 6171-6179. CHAPTER IV ARGININE“ IS ESSENTIAL FOR THE INHIBITION OF ANABAENA PCC 7120 ADP-GLUCOSE PYROPHOSPHORYLASE BY PHOSPHATE 103 ABSTRACT The chemical modification study with phenylglyoxal has implicated the involvement of an arginine residue at the allosteric sites, most probably the inhibitor- binding site, of ADP-glucose pyrophosphorylase from cyanobacterium Anabaena PCC 7120 (Chapter III). In order to identify the arginine residue, five arginine residues, which are conserved in all higher-plant and cyanobacterial enzymes but not in enteric bacterial enzymes, were individually converted to alanine by site-directed mutagenesis. The mutant enzymes were purified and the properties of these mutants were compared with the wild- type enzyme. Substitution of arginine294 with alanine resulted in an enzyme with more than 100-fold or 40-fold lower affinity for the inhibitor, P,, in the absence or presence of 3PGA, respectively. This mutation had no or lesser impact on the kinetic constants for the substrates and the activator, 3PGA. 104 INTRODUCTION ADP-glucose pyrophosphorylase catalyzes the reaction in which ADP-glucose, the glucosyl donor of plant starch and bacterial glycogen synthesis, is synthesized from ATP and glucose-l-P (Preiss & Romeo, 1989; Preiss, 1991; Preiss, 1996; Preiss & Sivak, 1996). The enzyme from higher plants and cyanobacteria is mainly activated by 3PGA and inhibited by P, (Ghosh & Preiss, 1965; Sanwal et al., 1968), whereas the enzyme from enteric bacteria is activated by FBP and inhibited by AMP (Preiss et al., 1966). Previous chemical modification with PLP and site-directed mutagenesis studies have shown that the activator-binding site of the E. coli ADP-glucose pyrophosphorylase, Lys”, is near the N-terminus (Parsons & Preiss, 1978; Gardiol & Preiss, 1990), while 419 those of the cyanobacterial enzyme from Anabaena, Lys382 and Lys , are close to the C- terminus (Sheng et al., 1996; Charng et al., 1994). Studies on the spinach leaf enzyme 419 have shown that Lys‘”, which corresponds to Lys in Anabaena, is involved in binding of the activator, 3PGA (Morell et al., 1988; Ball & Preiss, 1994). Previous studies on E. coli, Synechocystis, and spinach leaf enzymes using an arginine-specific reagent, phenylglyoxal, suggested that there was an essential arginine residue involved in the allosteric activation and inhibition (Carlson & Preiss, 1982; Iglesias et al., 1992; Ball & Preiss, 1992). The present study (Chapter III) shows that the Anabaena enzyme is inactivated by phenylglyoxal. The inactivation can be prevented by the presence of the activator 3PGA, the inhibitor P,, or the substrate ATP plus magnesium, with P, as the best protector. The apparent affinity for P, of the modified enzyme is 10—fold lower than that 105 106 of the wild-type enzyme in the absence of 3PGA. The modified enzyme also has a 4-fold lower affinity for 3PGA. This suggested that either modification affected the enzyme's catalysis or phenylglyoxal could mimic P, as the allosteric inhibitor. Alanine scanning of five conserved arginine residues which might be involved in P, inhibition was performed. There are five arginine residues, Arg“, Argws, Arg‘", Mg“, and Arg’“, conserved in the higher-plant and the cyanobacterial enzymes but not in the E. coli enzyme. These five arginine residues were considered as the candidates for P,-binding site, since P, inhibits the higher-plant and the cyanobacterial enzymes but it is not the inhibitor of the E. coli enzyme. Replacing Ar 9“ with alanine resulted in an enzyme with more than 100- or 40-fold lower apparent affinity for P, in the absence or presence of the activator, 3PGA, respectively. This result indicated that Arg294 was involved in the binding of the inhibitor, P,. MATERIALS AND METHODS Reagents [32P]PP, was purchased from Du Pont-New England Nuclear. [“C]Glucose-1-P was from ICN Pharmaceuticals Inc. [oz-”SMATP and the in vitro mutagenesis kit were from Amersham Corp. Oligonucleotides were synthesized and purified by the Macromolecular Facility at Michigan State University. All other reagents were purchased at the highest quality available. Bacterial Strains and Media E. coli strain TGl [K12, A(lac-pro), supE, thi, hsdD5/F‘traD36, proA+B+, lacl“, lacZAMlS] was used for site-directed mutagenesis and grown in LB medium. E. coli mutant strain AC70R1-504, which is deficient in ADP-glucose pyrophosphorylase activity (Carlson et al., 1976), was used for expression of the Anabaena ADP-glucose pyrophosphorylase gene (Charng et al. , 1992) and grown in enriched medium containing 1.1% KZHP04, 0.85% KHZPO4, 0.6% yeast extract, and 0.2% glucose, pH 7.0. Site-Directed Mutagenesis Site-directed mutagenesis was performed according to the method of Sayers et al. (1988) using the in vitro mutagenesis kit from Amersham Corp. Plasmid pAnaE3a was used for both site-directed mutagenesis and gene expression (Charng et al. , 1992; Charng et al., 1994; Sheng etal., 1996). The mutant enzymes with alanine substitution at position 66, 105, 171, 294, and 385 were designated as R66A, R105A, R171A, R294A, and R385A, respectively. The oligonucleotides used for conversion to alanine are shown in Table 1. The plasmids recovered in the last step of the mutagenesis were screened by 107 108 Table 1: Conserved arginine residues in Anabaena ADP-glucose pyrophosphorylase and the oligonecleotides used for conversion to alanine position amino acid sequence original sequence mutant sequence 66 SLNRHLS TCTCTCAATCGCCACATTGCC TCTCTCAATGCTCACATTGCC 105 DAVRQYL GATGCTGTACGTCAGTATCTC GATGCTGTAGCTCAGTATCTC 171 NSGRVID AACTCTGGACGAGTCATTGAT AACTCTGGAGCTGTCATTGAT 294 TRARYLP ACCCGCGCTCGTTACTTACCA ACCCGCGCTGCTTACTTACCA 3 8 5 KNARIGH AACAATGCCCGCATCGGTCAC AACAATGCCGCTATCGGTCAC 109 dideoxy sequencing (Sanger et al., 1977) in the regions of the desired mutations. The entire coding regions of these mutant alleles were sequenced to verify that there were no unintended mutations. Expression and Purification of the Mutant Enzymes The mutant genes were expressed in AC70Rl-504 cells. The R66A and the R294A enzymes were purified to homogeneity as previously described (Chapter III). The purification procedure for the R105A and the R385A enzymes was the same as that for the R66A and the R294A enzymes except that after the pellet was resuspended in 1.2 M ammonium sulfate, pH 7.5 and centrifugation, the new pellet was dissolved in 50 mM Hepes-NaOH, 2 mM DTT, pH 8.0 (about 0.5 mL/mg of protein). Enzyme Assay Assay I. Enzymatic activity was measured in the pyrophosphorolysis direction at 37 °C according to Preiss et al. (1966) during enzyme purification. The reaction mixtures contained 80 mM Hepes—NaOH buffer (pH 7.0), 2 mM ADP-glucose, 8 mM MgC12, 2 mM [32P]PP, (about 1500—3000 cpm/nmol), 4 mM 3PGA, 4 mM NaF, 50 pg of bovine serum albumin, and enzyme in a total volume of 250 pL. Assay II. Enzymatic activity was measured in the ADP-glucose synthesis direction at 37 °C according to the method of Preiss et al. (1966). (A) Activated conditions: The synthesis of ADP-[“C]glucose from [”C]glucose-1-P and ATP was measured in the presence of activator, 3PGA. The reaction mixture contained 100 mM Hepes—NaOH buffer (pH 8.0), 0.5 mM [“C]glucose-1-P (about 1000 cpm/nmol), 2.5 mM ATP, 10 mM MgC12, 2.5 mM 3PGA, 50 pg of bovine serum albumin, 0.15 unit of inorganic 110 pyrophosphatase, and enzyme in a final volume of 200 pL. (B) Unactivated conditions: The assay in the absence of activator was the same as described above except that 3PGA was omitted and the amount of ATP was 5 mM in the reaction mixture. For assay of the R66A and the R294A enzymes, the reaction conditions were identical to wild-type, except that 1.0 mM [“C]glucose-1-P was used to obtain maximal activity. Kinetic Characterization Kinetic data were plotted as initial velocity versus substrate or effector concentration. Saturating concentrations of substrates and effectors were determined to ensure that maximal velocity was attained. Data were replotted as double-reciprocal plots to determine Vmu. Kinetic constants from hyperbolic plots were also determined by double-reciprocal plots. Sigmoidal plots were replotted as Hill plots to obtain kinetic constants. Interaction coefficients, n", were also determined by Hill plots. Kinetic constants were expressed as A0,, 80.5, and 10.5, which correspond to the concentration of activator, substrate, or inhibitor giving 50% of maximal activation, velocity, and inhibition, respectively. All the kinetic parameters calculated from double-reciprocal or Hill plots were in good agreement with those obtained by using a computer program (Canellas & Wedding, 1980) which performed nonlinear iterative least-squares fitting to the Hill equation. Protein Assay Protein concentration was determined by using bicinchoninic acid reagent (Smith et al. , 1985) with bovine serum albumin as the standard. 111 Protein Electrophoresis and Immunoblotting SDS—PAGE was performed according to Laemmli (1970). Following electrophoresis, proteins on the gel were visualized by staining with Coomassie Brilliant Blue R-250 or electroblotted onto a nitrocellulose membrane according to Bumette (1981). After electroblotting, nitrocellulose membranes were treated with affinity-purified anti- spinach leaf ADP-glucose pyrophosphorylase IgG (Morell et al., 1987), and the antigen—antibody complex was visualized via treatment with alkaline phOsphatase linked goat anti-rabbit IgG followed by staining with BM purple AP-substrate precipitating reagent (from Boehringer Mannheim GmbH). Thermal Stability The purified enzymes were diluted to the same concentration, 0.2 mg/mL, in 50 mM Hepes—NaOH (pH 7.0) containing 1 mg/mL bovine serum albumin. The samples were heated for 5 min in a 60 °C water bath and then immediately placed on ice. The activities of these heated enzymes were assayed in the pyrophosphorolysis direction as described above. RESULTS Expression of the Mutant Enzymes The expression of the mutant enzyme was confirmed by resolving the crude extract proteins with SDS—~PAGE. Anabaena ADP-glucose pyrophosphorylase was identified by immunoblotting with antibody prepared against the spinach leaf ADP-glucose pyrophosphorylase that has been shown to be reactive with the Anabaena enzyme (Iglesias et al., 1991; Charng et al., 1992). The R66A, the R105A, and the R294A enzymes were expressed at a level similar to the wild-type enzyme, while the R385A enzyme was expressed less than the wild-type enzyme based on the result of immunoblotting. The apparent size of the mutant enzymes were the same as that of the wild-type enzyme. The R171A enzyme could not be identified by immunoblotting and the activity in the crude extract could not be detected, while the amount of the expression plasmid in the cells was similar to that of the wild-type enzyme. Screen for Alanine-Substituted Mutant Enzymes with Altered P, Inhibition Enzymatic activity in the crude extract was measured in the pyrophosphorolysis direction (Assay I). The R66A and the R294A enzymes were less sensitive to P, inhibition than the wild-type enzyme (Figure 1). The inhibition of the R105A enzyme and the R385A enzyme by P, was similar to that of the wild-type enzyme (data not shown). Purification of the Mutant Enzymes The R66A and the R294A enzymes were purified to greater than 90% homogeneity, while the R105A and the R385A enzymes were purified to greater than 50% and 80% , respectively, as estimated by SDS—PAGE of about 5 pg of protein. 112 113 Figure 1: P, inhibition of the wild-type (O), the R66A (I), and the R294A (v) enzymes in the crude extract. Enzyme activity was assayed in the pyrophosphorolysis direction (Assay I) and 100 % activity was 20.6, 7.7, and 5.9 nmol of ATP formed per 10 min, respectively. The data represents the mean of two separate experiments. 114 33:8 .55 as [Pi], mM Figure 1 1 15 Kinetic Characterization of the R294A Enzyme In the synthesis direction, the apparent affinity for P, decreased dramatically when Arg294 was replaced with alanine (Table 2). The 10.5 of the wild-type enzyme was 0.055mM and the enzyme was inhibited 100 % by 0.04 mM P, in the absence of 3PGA, while the remaining activity of the R294A enzyme was more than 60 % at 4 mM P, (Figure 2A). In the presence of 3PGA, the 10.5 of the wild-type enzyme was 1.0 mM and the enzyme was inhibited 100 % by 2 mM P,, while P, seemed to have no effect on the R294A enzyme up to 4 mM (Figure 213). ' Up to 100 mM of P, was used to see the effect of P, at higher concentration. The R294A enzyme was totally inhibited at 100 mM and the apparent affinity for P, was 5.2 mM or 38 mM in the absence (Figure 2A) or presence (Figure 2B) of 3PGA, respectively. Considering this effect may be due to the high ionic strength but not the specific inhibition by P,, KCl with the ionic strength was used to see the effect on enzyme activity (Figure 3). KCl had very little effect on enzyme activity of the wild-type enzyme compared to P, at lower ionic strength but inactivated the enzyme at high ionic strength (Figure 3A). The effect of KCl on the R294A enzyme activity was less than P, at low ionic strength but similar to P, at high ionic strength (Figure 3B). This suggested that the inhibition of P, at high concentration was due to the ionic strength. Thus the actual 10,, for P, of the R294A enzyme was more than 5.2 mM and 38 mM in the absence and presence of 3PGA, respectively (Table 2). In the presence of saturating 3PGA, the Vm value of the R294A enzyme was about 3-fold higher than that of wild-type, while in the absence of the activator, the VM value 116 .8528 $8588 05 S enigma 0.. mm “a abacus—win? .«o 38.: u Show 9 8:38 0850 Mo 2508a ca 3 332:5 s bran“ 0% he «E: 25 . .«2 «no: :2. «a: ..«8 + 86 a $5 _ a : 3 a 3. 3 a 3. no a me . cusses is e: s... a «35 6.: ~86 a 98... e: 85 « Sod ca: Sod « god 8.: “86 « Bod :5 .agafiaa 8.: 85 « Ba ad _ a 8 A 8.: 8.0 a as 2.: ad a and 6.3 86 a 3 + 6.: «85 « ~85 6.: a... a an A e: «85 a E; a: 85 a go 8.: «85 a mad . 25 ..m 6.3 3 a 1.. . 9.3 3 « tn 2.: ..o a 3 a: 3 a 3 as: 3 a 9n + G: 3 a 3 as: 3 a 3 ed 3 a 2. ed: 3 «.3 Ge 3 a to - as can: 8.: 89¢ a «85 8.: «85 a god 2.: Bod m E; 8.: 89¢ « ~36 ad: «86 a 235 + 3.8 nod « ”86 8.9 86 a :5 6.: ”85 a as... 8.: 485 a 485 8.9 8.:: a «86 - :5 disease 3.: ad a 26 a: 8... a n3 8.: So u :5 8.: 85 a $3 a: 86 a :6 + a: 3 a 3 9.: 3 a S 8.: 3 a 3 e: 3 a 3 a: 3 a I - :a ..E. 3:21. .3— .am 3583.9... <32 <32 <88 52 25.33 .acoatomxo 80:83qu 95 ...o dosage: 3%qu a use 05 some? 35 sends <22 as 632 £82 .éem .2533 8% as .8 masseuse Boa: a 23 117 Figure 2: The effect of P, on the enzyme activity of the wild-type (O) and the R294A (O) enzymes. One hundred percent activity was 6.5 and 12.0 nmol of ADP-glucose formed per 10 min, respectively, when enzyme activity was assayed in the absence (A) of 3PGA (Assay IIB). When enzyme activity was assayed in the presence (B) of 3PGA (Assay IIA), 100 % activity was 5.4 and 11.6 nmol of ADP-glucose formed per 10 min, respectively. The data represents the mean of two separate experiments. % initial activity % initial activity 118 100 80 60 1.0 2.0 3.0 4.0 50.0 100.0 [Pi], mM ll 20 - 0.0 II 119 Figure 3: The effect of P, (O) and KCl (I) on the enzyme activity of the wild-type (A) and the R294A enzyme (B). Enzyme activity was assayed in the absence of 3PGA (Assay IIB). One hundred percent activity was 6.5 and 12.2 nmol of ADP-glucose formed per 10 min for the wild-type enzyme and 13.2 and 14.1 nmol of ADP-glucose formed per 10 min for the R294A enzyme. The data represents the mean of two separate experiments. % initial activity % initial activity 120 J] II Jl ' é ' 5 "'160'263300 Ionic strength 0 ‘- 100 80 60- 40- 20- 0 so too ' 150 ' zoo ' 250 300 Ionic strength Figure 3 121 was about 2-fold higher. The apparent affinities for the substrates, ATP, glucose-l-P, Mg“ , and activator, 3PGA were all relatively unaffected by the mutation at position 294 (Table 2), indicating that the conformation of this P,-binding site was relatively unchanged. Kinetic Characterization of the R66A, the R105A, and the R385A Enzymes The apparent affinity for P, of the R66A enzyme decreased about 5-fold compared to the wild-type enzyme, while those of the R105A and the R385A enzymes were similar to that of the wild-type (Table 2). The affinities for substrates, ATP, glucose-l-P, Mg“, and activator, 3PGA were all relatively unaffected by the mutations (Table 2), indicating that the conformations of these ligand-binding sites are relatively unchanged. The only significant changes observed were the 4- and 7-fold increase in the 80.5 for ATP of the R66A and the R105A enzyme in the presence of activator, and the 10- and 5-fold decreases in the Vm of the R385A enzyme in the absence and presence of activator, respectively (Table 2). Thermal Stability of the Wild-Type and the Mutant Enzymes After heat treatment at 60 °C for 5 min, the activity of the wild—type, the R66A, R105A, R294A, and the R385A enzymes retained 92, 82, 79, 88, and 75% activity, respectively. The result indicates that these arginine residues are not required for the stability of the enzyme. DISCUSSION Anabaena ADP-glucose pyrophosphorylase was inactivated by phenylglyoxal. At full inactivation, one arginine residue was modified per enzyme subunit. The activator 3PGA protected the enzyme from inactivation. However, the inhibitor P, proved to be more effective at protecting the enzyme than 3PGA. In addition, substrate ATP plus Mg“ also protected the enzyme from inactivation, which was not seen in previous studies with other ADP—glucose pyrophosphorylases (Carlson & Preiss, 1982; Ball & Preiss, 1992; Iglesias et al., 1992). Alignment of all the amino acid sequences of ADP-glucose pyrophosphorylase available has shown that these five arginine residues are conserved in all the cyanobacterial and the higher-plant enzymes (Smith-White & Preiss, 1992). Considering the effect of the ionic strength on enzyme activity, the actual apparent affinity for P, of the R294A enzyme should be more than 100- and 40-fold lower than those of the wild-type enzyme in the absence and presence of 3PGA, respectively. This large effect on the 10,, value for P, when 294 Arg was replaced by alanine suggests that Arg“ is involved in P, binding. Although the 105 for P, of the R66A enzyme in the absence of 3PGA was 5-fold higher than that of the wild-type, this change was not significant compared to the more than 100-fold change for the R294A enzyme. It was also not specific considering the So, for ATP in the presence of 3PGA and the affinity for glucose-l-P in the absence of 3PGA were changed 4- and 2-fold, respectively. Thus, Arg66 seems not to be directly involved in P, inhibition. The R105A enzyme had a 7-fold lower affinity for ATP in the presence of 3PGA 122 123 while it had similar kinetic constants for other substrates, activator 3PGA, and inhibitor P,. This suggested that Arg‘°5 might be close to ATP-binding site. It should be noted that Arg105 is eight amino acids away in the alignment (Smith-White & Preiss, 1992) from Tyr‘” in E. coli enzyme, which is the putative ATP-binding site (Lee & Preiss, 1986). The apparent affinities for all substrates, activator 3PGA, and inhibitor P, of the R385A enzyme were similar to those of the wild-type enzyme. The Va,” of this mutant enzyme was 10- or 5-fold lower than that of the wild-type enzyme in the absence or presence of 3PGA, respectively, which suggests that mutation at Arg’” may result in a conformational change and affect the enzyme's catalysis. The R171A enzyme could not be identified by immunoblotting and no activity could be detected in the crude extract, though the amount of the expression plasmid in the cells was similar to that of the wild-type expression plasmid. Argm may be important for maintaining the proper structure of ADP-glucose pyrophosphorylase. All the other mutant enzymes were expressed at a level similar to the wild-type enzyme and retained more than 75% activities after heat treatment. Arg“, Argw’, Arg294, and Arg’“ are obviously not critical to the stability of the native folded state. Evidence of this was also observed in the relatively unchanged kinetic constants obtained for ATP, Mg“, glucose-l-P, and 3PGA for the four mutants, showing the tolerance to amino acid substitution at these positions. Due to the properties of the cyanobacterial ADP-glucose pyrophosphorylase and the key position of these photosynthetic prokaryotes during evolution (Aitken, 1988), it is of great interest to understand the structure—function relationships of the allosteric sites of the cyanobacterial ADP-glucose pyrophosphorylase. Previous studies have shown that 124 Lys382 and Lys419 of Anabaena enzyme were involved in the binding of the activator, 3PGA. The present study clearly indicated that Arg294 of Anabaena enzyme is involved in the binding of P,. The activator and the inhibitor obviously bind to different sites. It would be interesting to replace Arg294 with another positively charged amino acid, lysine, or other neutral amino acids with larger size than alanine, such as leucine or glutamine, or even a negatively charged amino acid, like glutamic acid. This would show if the charge is the most important factor for binding of P, and if the size of the amino acid at this position is also important. The fact that the corresponding amino acid at this position in the E. coli enzyme is glutamic acid makes it more interesting to see whether or not the inhibitor specificity would change when substituting Arg294 in the Anabaena enzyme with glutamic acid or replacing the corresponding glutamic acid in E. coli enzyme with arginine. Although more site-directed mutagenesis study at this P,-binding site would be very interesting, the most reliable way to study the structure-function relationships of ADP- glucose pyrophosphorylase is to crystallize the enzyme and determine the 3-dimensional structure . REFERENCES Aitken, A. (1988) Methods Enzymol. 167, 145-154. Ball, K., & Preiss, J. (1992) J. Protein Chem. 11, 231-238. Ball, K., & Preiss, J. (1994) J. Biol. Chem. 269, 24706-24711. Bumette, W. W. (1981) Anal. Biochem. 112, 195-203. Canellas, P. F., & Wedding, R. T. (1980) Arch. Biochem. Biophys. 199, 259-264. Carlson, C. A., Parsons, T. F., & Preiss, J. (1976) J. Biol. Chem. 251, 7886-7892. Carlson, C.A., & Preiss, J. (1982) Biochemistry 21, 1929-1934. ' Charng, Y., Kakefuda, G., Iglesias, A. A., Buikema, W. J ., & Preiss, J. (1992) Plant Mol. Biol. 20, 37-47. Charng, Y., Iglesias, A. A., & Preiss, J. (1994) J. Biol. Chem. 269, 24107-24113. Gardiol, A., & Preiss, J. (1990) Arch. Biochem. Biophys. 280, 175-180. Ghosh, H. P., & Preiss, J. (1965) J. Biol. Chem. 240, 960-961. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1991) Plant Physiol. 97, 1187-1195. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1992) J. Protein Chem. 11, 119-128. Laemmli, U. K. (1970) Nature 227, 680-685. Lee, Y. M., & Preiss, J. (1986) J. Biol. Chem. 261, 1058-1064. Morell, M., Bloom, M., Knowles, V., & Preiss, J. (1987) Plant Physiol. 85, 182-187. Morell, M., Bloom, M., & Preiss, J. (1988) J. Biol. Chem. 263, 633-637. Parsons, T., & Preiss, J. (1978) J. Biol. Chem. 253, 6197-6202. Preiss, J ., Shen, L., Greenberg, E., & Gentner, N. (1966) Biochemistry 5, 1833-1845. Preiss, J ., & Romeo, T. (1989) Adv. Microb. Physiol. 30, 183-238. Preiss, J. (1991) in Plant Molecular and Cell Biology (Mifflin, B., ed.) Vol. 7, pp 59- 125 126 114, Oxford University, Oxford. Preiss, J. (1996) Biotech. Annu. Rev. 2, 259-279. Preiss, J, & Sivak, M. (1996) in Photoassimilate Distribution in Plants and Crops (Zamski, E., Schaffer, A. A., ed.) pp 63-96, Marcel Dekker, Inc., New York. Sanger, F., Nicklen, S., & Coulson, A. R. (1977) Proc. Natl. Acad. Sci. U.S.A. 74, 5463-5467. Sanwal, G. G., Greenberg, E., Hardie, J., Cameron, E. C., & Preiss, J. (1968) Plant Physiol. 43, 417-427. Sayers, J. R., Schmidt, W., & Eckstein, F. (1988) Nucleic. Acids Res. 16, 791-802. Sheng, J., Charng, Y., & Preiss, J. (1996) Biochemistry 35, 3115—3121. Smith, P. K., Krohn, R. 1., Hermanson, G. T., Mallia, A. K., Gartner, F. H., Provenzano, M. D., Fujirnoto, E. K., Geoke, N. M., Olson, B. J., & Klenk, D. C. (1985) Anal. Biochem. 150, 76-85. Smith-White, B. J ., & Preiss, J. (1992) J. Mol. Evol. 34, 449-464. CHAPTER V SUMMARY AND PERSPECTIVES 127 SUMNIARY AND PERSPECTIVES 1 . Lys382 of the cyanobacterium Anabaena ADP-glucose pyrophosphorylase is involved in the binding of the activator, 3PGA. A site—directed mutagenesis study on the cyanobacterium Anabaena ADP-glucose pyrophosphorylase was conducted to determine the role of Lys382 in binding of the activator, 3PGA. This work was based on a previous chemical modification study. Phosphopyridoxylation of the K419R mutant enzyme modified Lysm, suggesting that this residue might also be located within the activator-binding site (Charng et al., 1994). The corresponding lysine residue on the large subunit of spinach leaf enzyme was labeled with PLP in a similar phosphopyridoxylation experiment (Ball & Preiss, 1994). Alignment of all the amino acid sequences of ADP-glucose pyrophosphorylases available has shown that Lys3m is conserved in the cyanobacterial enzymes, in the small subunit of the higher-plant enzymes, and in the large subunit of the spinach leaf, wheat leaf, and potato tuber enzymes (Smith-White & Preiss, 1992). Replacing Lys382 with either arginine, alanine, or glutamine produced mutant enzymes with apparent affinities for 3PGA 10—, 40-, or 160-fold lower than that of the wild-type enzyme. The glutamic acid mutant enzyme was inhibited by 3PGA and it may be due to an abnormal conformational change. The kinetic constants for ATP, Mg“, glucose-l-P, and P, are relatively unchanged for the five mutants compared to the dramatic changes in Am, suggesting that there are different binding sites for the substrates and the inhibitor, P,. 128 129 Site-directed mutagenesis was also performed to obtain a K382R-K419R double mutant. The apparent affinity for 3PGA of this double-mutant enzyme was about 100-fold lower than that of the wild-type enzyme, and the specificity for activator of this mutant enzyme was altered. FBP and AMP, the major activator and inhibitor of the E. coli ADP- glucose pyrophosphorylase, were more effective as activators than 3PGA. Both residues 419 in E. coli enzyme, corresponding to Lys382 and Lys in Anabaena enzyme, are arginine instead of lysine (Charng et al. , 1992). The differences of activator specificity may relate to the difference of amino acids at the activator-binding site of the E. coli and Anabaena/higher-plant enzymes. Chemical modification study on the K382R-K419R enzyme with PLP suggested that other lysine residues are not involved in the binding of 3PGA. Based on previous studies (Charng et al. , 1994) and our results, we can conclude that both Lys382 and Lys419 are involved in the binding of 3PGA. Both charge and size of these two lysine residues are important for the binding of the activator. The charge seems to be more important, which may be explained by the interactions between the two positively charged e-amino groups and the negatively charged carboxyl and phosphate groups of 3PGA. 2. An arginine residue(s) is involved in the allosteric regulation of Anabaena ADP- glucose pyrophosphorylase. Previous chemical modification studies with phenylglyoxal have shown that there is an essential arginine residue(s) involved in the allosteric regulation (Carlson & Preiss, 130 1982; Iglesias et al. , 1992; Ball & Preiss, 1992). Treatment of ADP-glucose pyrophosphorylase from the cyanobacterium Anabaena with phenylglyoxal resulted in a time- and concentration-dependent loss of both activated and unactivated activity when modified by phenylglyoxal. P,, the inhibitor, protected the enzyme from inactivation most effectively, while 3PGA, FBP, PLP, and ATP plus magnesium were also good protectors. The modified enzyme had lower apparent affinity for P, in the absence of the activator, 3PGA, than that of the wild-type enzyme. At full inactivation, one arginine residue was modified per enzyme subunit. These results suggested that either modification affected the enzyme's catalysis or phenylglyoxal could mimic P, as the allosteric inhibitor. 3. Argz“ of the Anabaena ADP-glucose pyrophosphorylase is important for the inhibition by P,. The chemical modification study with phenylglyoxal has implicated the involvement of an arginine residue(s) at the allosteric sites, most probably the inhibitor- binding site, of ADP-glucose pyrophosphorylase from cyanobacterium Anabaena. In order to identify the arginine residue(s), five conserved arginine residues were individually converted to alanine by site-directed mutagenesis. The properties of these mutant enzymes were compared with those of the wild-type enzyme. Substitution of arginine294 with alanine produced an enzyme with more than 100-fold or 40-fold lower affinity for the inhibitor, P,, in the absence or presence of 3PGA, respectively. This mutation had no or lesser impact on the kinetic constants for the substrates and the activator, 3PGA. This large effect on the 1,,5 value for P, when Argz94 was replaced by alanine suggests that Arg“294 is 131 involved in P, binding. The apparent affinity for P, of the R66A enzyme decreased about 5-fold compared to that of the wild-type enzyme, while those of the R105A and the R385A enzymes were similar to that of the wild-type. The affinities for the substrates, ATP, glucose-l-P, Mg“, and the activator, 3PGA were all relatively unaffected by the mutations at position 66, 105, and 385, indicating that the conformations of these ligand-binding sites are relatively unchanged. The only significant changes observed were the 4- and 7 -fold increase in the 5,5 for ATP of the R66A and the R105A enzyme in the presence of activator, and the 10- and 5-fold decreases in the Vm, of the R385A enzyme in the absence and presence of activator, respectively. The R171A enzyme could not be identified by immunoblotting and no activity could be detected in the crude extract, though the amount of the expression plasmid in the cells was similar to that of the wild-type expression plasmid. Arg171 may be important for maintaining the proper structure of ADP-glucose pyrophosphorylase. 4. Future Research: It is of great interest to understand the structure—function relationships of the allosteric sites of the cyanobacterial ADP-glucose pyrophosphorylase due to the properties of the cyanobacterial ADP-glucose pyrophosphorylase and the key position of these photosynthetic prokaryotes during evolution (Aitken, 1988). We have shown that Lysm, 419 in addition to Lys , of Anabaena enzyme is involved in the binding of the activator, 3PGA. The present study also indicates that Arg294 of the Anabaena enzyme is involved 132 in the binding of P,. The activator and the inhibitor obviously bind to different sites. It would be of interest to replace Arg294 with another positively charged amino acid, lysine, or other neutral amino acids with larger size than alanine, such as leucine or glutamine, or even a negatively charged amino acid, like glutamic acid. This would show if the charge is the most important factor for binding of P, and if the size of the amino acid at this position is also important. The fact that the corresponding amino acid at this position in the E. coli enzyme is glutamic acid makes it more interesting to see if the inhibitor specificity would change when substituting Arg294 in the Anabaena enzyme with glutamic acid or replacing the corresponding glutamic acid in E. coli enzyme with arginine. More site—directed mutagenesis studies on other cyanobacterial and higher-plant ADP-glucose pyrophosphorylases would show if the arginines corresponding to Arg294 are also important for P, inhibition. The combination of chemical modification and site-directed mutagenesis is very useful to study the structure-function relationships of ADP-glucose pyrophosphorylase, while determination of the 3-dimensional structure will be very helpful. Other biophysical techniques, such as NMR and circular dichroism, also may be very useful to provide information about the structure of ADP-glucose pyrophosphorylase. REFERENCES Aitken, A. (1988) Methods Enzymol. 167, 145-154. Ball, K., & Preiss, J. (1992) J. Protein Chem. 11, 231-238. Ball, K., & Preiss, J. (1994) J. Biol. Chem. 269, 24706-24711. Carlson, C.A., & Preiss, J. (1982) Biochemistry 21, 1929-1934. Charng, Y., Kakefuda, G., Iglesias, A. A., Buikema, W. J ., & Preiss, J. (1992) Plant - Mol. Biol. 20, 37-47. Charng, Y., Iglesias, A. A., & Preiss, J. (1994) J. Biol. Chem. 269, 24107-24113. Iglesias, A. A., Kakefuda, G., & Preiss, J. (1992) J. Protein Chem. 11, 119-128. Smith-White, B. J ., & Preiss, J. (1992) J. Mol. Evol. 34, 449-464. 133