mm . 14.‘ [23.1111 3...... .r . . iiufistvt. » 4.1 «753...»: ..'| . I ‘ . .«Lttv . . A "44... . J .14! ‘AI’I Atfi f zafiéfié g§§¢ o i. 5 ; F .W3 A at kl k) C _.Q LIBRARY Michigan State University This is to certify that the dissertation entitled X-RAY CRYSTALLOGRAPHIC STUDIES OF CELLULAR RETINOIC ACID-BINDING PROTEIN II MUTANTS DESIGNED AS RHODOPSIN MIMICS presented by XIAOFEI JIA has been accepted towards fulfillment of the requirements for the doctoral degree in chemistry Major Professor’s Signature V/ ll/ll/Og Date MSU is an affinnetive-action, equal-opportunity employer PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K'IProj/AccsPres/ClRC/DateDue.indd X-RAY CRYSTALLOGRAPHIC STUDIES OF CELLULAR RETINOIC ACID- BINDING PROTEIN II MUTANTS DESIGNED AS RHODOPSIN MIMICS By Xiaofei J ia A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2008 ABSTRACT X-RAY CRYSTALLOGRAPHIC STUDIES OF CELLULAR RETINOIC ACID- BINDING PROTEIN II MUTANTS DESIGNED AS RHODOPSIN MIMICS By Xiaofei Jia Color vision is one of the most fascinating mechanisms of the human anatomy. It is achieved by the collaboration of four visual pigments in the retina: the rod, red, green and blue rhodopsins. Upon binding of 11-cis-retinal, the four rhodopsins exhibit distinct absorption maxima, which altogether cover the entire visible spectrum and enable color vision. The underlying opsin-chromophore interactions within each rhodopsin are therefore of great interest and significance. Elucidation of these mechanisms, however, has been hampered by difficulties associated with the overexpression, purification, and routine crystallization of the membrane-bound rhodopsins. To overcome these difliculties and advance our understanding of the wavelength regulation mechanism in color vision, we have established a model system for rhodopsin, employing a mimic protein, cellular retinoic acid-binding protein II (CRABPII). Upon binding of a shorter retinal analog, Cls-aldehyde, several rationally designed CRABPII mutants exhibited altered absorption maxima of the bound chromophore. These mutants were characterized both biochemically and structurally. Chapter II of this dissertation focuses on the structural insights obtained for four such holo-CRABPII mutants based on the high-resolution crystal structures. The protonated Schiff base formation between Lysl32 and the chromophore has been confirmed in all structures. Two distinctive binding orientations have been observed in different C's-aldehyde- CRABPII complexes. A single mutation at 121 is shown to play a key role in determining the binding mode of the ligand. Based on the available high-resolution structures, further rational design of CRABPII has led to the discovery that placing negatively charged residues along the polyene chain of the chromophore leads to a clear trend in wavelength regulation. High-resolution crystal structures of six ape-CRABPII mutants have also been determined. Chapter III details the discussion of how different mutations affect the structural integrity of the protein. A water-mediated hydrogen-bonding network has been observed in all apo-CRABPII structures, including both wild-type and mutants. Mutations usually lead to altered water-mediated networks and yet do not disrupt the overall structure of the protein. Mutations on the helical portal, however, tend to have detrimental structural effects. Chapter IV presents an attempt toward elucidating the direct interaction between CRABPII and the retinoic acid receptor (RAR), which is responsible for the retinoic acid delivery. Chapter V discusses the efforts toward obtaining the first crystal structure of the cellular retinaldehyde-binding protein (CRALBP), another important protein in vision. CRALBP functions in the regeneration process of 11-cis-retinal from all-trans-retinol in the retina pigment epithelium. Initial efforts toward crystallizing the full length CRALBP were unsuccessful. Subsequent mutagenesis combined with biochemical studies have helped identify suitable truncations at the N-terminal of this protein. Recent crystallization trials with a truncated CRALBP complexed with a synthetic ligand have shown encouraging results. Copyright by XIAOFEI JIA 2008 To my father ACKNOWLEDGMENTS First and foremost, I would like to thank my two advisors, professor James H. Geiger and professor Babak Borhan. My journey of pursuing a doctoral degree in chemistry would have been interrupted if not for Jim and Babak, who provided me the opportunity to continue my study and advance myself in science. Jim and Babak share some of the finest qualities as scientists and mentors. They both have great passion for science and possess tremendous knowledge. Besides, what made them very special is their joyful attitude toward science. They both put their interests and passion to science in such a positive and contagious way that people around them are easily inspired. I consider myself very lucky to work in the labs of these two great scientists/mentors and learn a great deal from each of them. I also feel very grateful to my colleagues working in the same project. Dr. Chrysoula Vasileiou has been with this project since the very beginning and has been the biggest factor for its success. Chrysoula not only contributes excellent scientific insights into the research, but also coordinates this group of people in a very enjoyable, cordial way. I especially thank her for all the suggestions she gave to me. Dr. Soheila Vaezeslami worked on the structure determination of CRABPII mutants before me. Her success certainly paved the way for me. I owe a very special thanks to Kin-Sing Lee. Sing and I had the closest collaboration. We exchanged results and ideas on a regular basis. His intellectual inputs as well as his hard work have greatly aided the growth of the project and also helped me in doing my part of the research. Many of his data have been incorporated and discussed in this dissertation. I would also like to thank Calvin vi Grant, Wenjing Wang, Camille Watson and Rafida Nossori who are working on other aspects of this project. I appreciate all your collaboration and support over the years. I would like to thank my labmates in the Geiger group. Dr. Stacy Hovde was the one who helped me in everything when I first started in the lab. Her intelligence, diligence, friendliness and her super love for Spartan hockey all impressed me very much. It was very a pleasant experience working with her. Dr. Sara Cnude and Dr. Fang Sheng helped me a great deal on getting started with crystallography. Their kindness and patience had made my learning process very enjoyable. I also want to thank Lei Feng for the discussions and friendship over the years. I owe a special thanks to Susan Wortas for her taking so much time in proofreading my various writings. Her kindness and patience are deeply appreciated. I also want to thank Remie Touma for bringing her energy and positive attitude into the lab. I would also like to thank the Borhan group members, especially Dr. Tao Zheng, Dr. Jun Yan and Xiaoyong Li. I could not list all the names, but I certainly appreciate the their friendship and support over the years. I am deeply gratefirl to my friends, Dr. Jian Yi, Dr. Ningqing Ran, Dr. Yana Cen, Dr. Jinsong Yang, Dr. Gang Hu, Dr. Justas Jancauskas and Xianying Jiang. I am truly blessed by your friendship and support through all these years. Last but not least, I thank my family for their unconditional love and support throughout my life. My father, mother, grandma and my sister have always been there for me whenever I needed them. My life would not be the same without everything they have done for me. This dissertation is dedicated to my father, Ping J ia, who passed away on December 6'", 2008, two days before my defense. vii TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES ABBREVIATIONS CHAPTER I. Introduction _ - - - -1 1.1. The Mystery of Vision ............................................................................................... 1 1.2. Rhodopsins as mammalian visual pigments: structure and function in photoactivation. ............................................................................................................... 9 1.3. Other rhodopsins .................................................................................................... 12 1.4. Color vision ............................................................................................................. 15 1.5. Designing CRABPII mutants as rhodopsin mimics ................................................ 30 1.5.1. CRABPII general background. ........................................................................ 31 1.5.2. Designing CRABPII mutants as rhodopsin mimics. ....................................... 35 1.6. Cellular retinaldehyde-binding protein and the 11-cis-retinal regeneration. ....... 38 Reference ....................................................................................................................... 44 CHAPTER 11. Crystal Structures of Ligand-bound Cellular Retinoic Acid Binding Protein 11 Mutants 62 2.1. Introduction. ........................................................................................................... 62 2.1.1. Need for a ligand that can be totally buried inside the CRABPII mutants. ..... 62 2.1.2. Engineered CRABPII mutants as switchable colorimetric fusion proteins.....65 2.2. C I 5-aldehyde-bound CRABPII mutant structures ................................................... 67 2.2.3. Protonated Schiff base stabilization. ............................................................... 74 2.2.4. Cis versus trans protonated Schiff base. ......................................................... 77 2.2.5. Flexibility of the helix-tum-helix lid. .............................................................. 79 2.2.6. An evolutionary correlation between CRABPII and plasma retinol-binding protein? ...................................................................................................................... 82 2.2.7. Testing the point charge theory based on the C‘s-aldehyde-bound KLE- R59W-CRABPII structure. ........................................................................................ 86 2.2.8. Conclusion ....................................................................................................... 87 2.3. Merocyanin-bound CRABPII mutant structure. ..................................................... 89 References ...................................................................................................................... 93 CHAPTER 111. Crystal Structures of the Apo-CRABPII Mutants -_ 98 3.1. Introduction ............................................................................................................ 98 viii 3.2. Results ................................................................................................................... 100 3.2.1. Mutations of conserved and non-conserved residues .................................... 100 3.2.2. The structure ofapo-R132KzR111L2L121EzR59E-CRABPII (apo-KLE-R59E) ................................................................................................................................. 102 3.2.3. The structure ofapo-R132KzY134F:R111L:L121E-CRABPII (apo-KF LE) 105 3.2.4. The structure ofapo-Rl32K:Y134F:R111L:L121D-CRABPII (apo-KFLD) ................................................................................................................................. 107 3.2.5. The structure ofapo-R132K:Y134F:R111L:L121D:T54V-CRABPII (apo- KFLD-T54V) ........................................................................................................... 108 3.2.6. The structure ofapo-R132KzR111L:T54E-CRABPII (apo-KL-T54E) ........ 109 3.2.7. The structure of apo-R132KzR1 1 1L:A32E-CRABPII (apo-KL-A32E) ........ 111 3.3. Discussion ............................................................................................................. 112 3.3.1. Rigid overall fold of the apo-CRABPII mutant structures ............................ 112 3.3.2. A32E mutation is detrimental to the a2 helix in apo-KL-A32E ................... 114 3.3.3. Water-mediated network acts as a pillar inside the cavity ............................ 116 3.3.4. The hydrophobic cavity allows binding of BTP ............................................ 117 3.5. Conclusion ............................................................................................................ 118 Reference ..................................................................................................................... 1 19 CHAPTER IV. Investigation of the Interaction between Cellular Retinoic Acid Binding Protein 11 and Retinoic Acid Receptor Gamma Ligand Binding Domain.122 4.1. Background ........................................................................................................... 122 4.2 Results and discussion ........................................................................................... 128 4.2.6. Circular dichroism of RARy-LBD and fluorescence titration of RARy-LBD with retinoic acid. .................................................................................................... 133 4.2.7. Titration of KLE-CRABP II with retinal followed by UV. ........................... 135 4.2.8. In vitro binding assay using RARy-LBD directly from Ni-NTA resin. ........ 135 4.3. Conclusions .......................................................................................................... 13 7 References .................................................................................................................... 138 CHAPTER V. Preliminary X-Ray Crystallographic Studies of Cellular Retinaldehyde Binding Protein _ - - 140 5.1. Background ........................................................................................................... 140 5.2. Crystallization of the full-length CRALBP ........................................................... 140 5.2.1. His-tagged full-length CRALBP crystallization ............................................ 140 5.2.2. Non-fusion full-length CRALBP crystallization. .......................................... 142 5.2.3. Full-length CRALBP with a cleavable His-tag. ............................................ 144 5.3. Crystallization of the N-terminal-truncated CRALBP .......................................... 145 5.3.1. Truncation at 43/44 on the N-terminal of CRALBP. .................................... 145 5.3.2. Discovery of the ideal N-terminal truncation site .......................................... 147 5.3.3. Enhancing the expression and purification of functional CRALBP .............. 158 5.3.4. Crystallization trials. ...................................................................................... 159 ix 5.4. Conclusion. ........................................................................................................... 160 Reference ..................................................................................................................... 1 61 CHAPTER VI. Experimental - ...... -- ..... 162 6.1. Cloning, mutagenesis, expression and purification .............................................. 162 6.1.1. Cloning, mutagenesis, expression and purification of CRABPII mutants. ...162 6.1.2. Cloning, expression and purification of RARy-LBD. ................................... 164 6.1.3. Cloning, mutagenesis, expression and purification of CRALBP. ................. 165 6.2. Protein crystallization .......................................................................................... 1 70 6.2.1. Co-crystallization of CRABPII mutants bound with ligands. ....................... 170 6.2.2. Crystallization of apo-CRABPII mutants. ..................................................... 173 6. 3. Data processing and structure determination ...................................................... I 75 6.3.1. Structures of the Cls-aldehyde-bound CRABPII mutants. ............................ 175 6.3.2. Structure of the merocyanin-bound KLE-R59W-CRABPII mutant. ........... 176 6.3.3. Structures of the apo-CRABPII mutants. ..................................................... 177 6.4. Protein characterization ....................................................................................... 193 6.4.1. General procedure for UV-vis titrations of CRABPII/CRALBP with retinoids. ................................................................................................................................. 193 6.4.2. Determination of the extinction coefficients of proteins. .............................. 193 6.4.3. Reductive amination procedure. .................................................................... 195 6.6.4. Circular Dichroism of protein samples. ......................................................... 195 6.4.5. Fluorescence. ................................................................................................. 196 6.4.6. General protocol followed for Fluorescence Quenching and K, determination experiments. ............................................................................................................. 198 References .................................................................................................................... 203 LIST OF TABLES Table 1-1. Rhodopsins of various species. ......................................................................... 17 Table II-l. CRABPII mutants bound with C1 5-aldehyde. ................................................. 64 Table II-2. Residues around the ionone rings. ................................................................... 86 Table II-3. Absorption maxima observed for the KLE-X-CRABPII series of mutants ....87 Table II-4. UV maximum absorption of the merocyanin-bound CRABPII mutants. ....... 89 Table VI-l. X-ray data collection statistics of the ligand-bound CRABPII mutant crystals. ....................................................................................................................... 178 Table VI-2. Structure refinement statistics of the ligand-bound CRABPII mutant crystals. ....................................................................................................................... 179 Table VI-3. X-ray data collection statistics of the apo-CRABPII mutant crystals .......... 180 Table VI-4. Structure refinement statistics of the apo-CRABPII mutant crystals. ......... 18] xi LIST OF FIGURES Figure H. The human eye ................................................................................................... 2 Figure [-2. A. rod and cone cells in retina; B. rhodopsins embedded in the disks in photoreceptor cells. ............................................................................................ 3 Figure I-3. Schematic representation of the intermediates observed in the isomerization of 11-cis-retinal to all-trans-retinal in rhodopsin ................................................... 5 Figure I-4. Photoactivation cascade ..................................................................................... 7 Figure I-5. Crystal structure of bovine rod rhodopsin. ...................................................... 10 Figure I-6. Absorption profiles of different human rhodopsins. ....................................... 16 Figure I-7. Model compounds used to study wavelength shifts controlled by distance between the PSB and the counter anion. ......................................................... 20 Figure I-8. The point charge model ................................................................................... 21 Figure [-9. Hypothesis of steric interactions leading to wavelength regulation ................ 23 Figure I-10. Vicinity of retinal in cone pigments as shown in the homology models ....... 24 Figure I-l 1. Illustration of the sites of mutations for converting the red rhodopsin to a green-rhodopsin-like pigment .......................................................................... 27 Figure I-12. Illustration of the sites of mutations for converting the green rhodopsin to a red-rhodopsin-like pigment. ............................................................................ 27 Figure [-1 3. Illustration of the sites of mutations for converting the rod rhodopsin to a red-shifted pigment .......................................................................................... 28 Figure I-14. Crystal structure of retinoic acid bound in WT-CRABPII ............................ 32 Figure I-15. Water-mediated hydrogen bonding interactions observed in the molecule A of the apo-WT—CRABPII crystal structure ............................................ 34 Figure I-16. Crystal packing observed in the structure of apo-WT-CRABPII. ................. 35 Figure I-17. All-trans-retinal bound in KLE-CRABPII. ................................................... 37 Figure [-18 The visual cycle. ........................................................................................... 39 Figure I-19. Stereoview of the structure models built for CRALBP. ................................ 42 xii Figure II-l. Retinoic acid bound in wild-type-CRABPII. ................................................. 63 Figure II-2. Structures of the related compounds. ............................................................. 63 Figure II-3. Ovelay of the ligands bound in different binding orientations ....................... 69 Figure II-4. 2Fo-Fc electron density map calculated at 1.0 o for bound ligands. A. C15- aldehyde bound to Lysl32 (upper) and bis-tris-propane (lower) in KLE- R59W; B. Cis-aldehyde bound to Lysl32 in KL-A32E. ................................. 69 Figure II-5. Interactions between the PSB and the residues nearby in different C15- aldehyde-bound CRABPII mutants. ................................................................ 70 Figure II-6. A. The structures of KLE-Rt and Cls-aldehyde-KL-A32E are overlaid. Glul2l from KLE-Rt (magenta) and Cl5-aldehyde from hoIo-KL-A32E (green) are shown; B. Overlay of the C15-aldehyde bound in three CRABPII mutants: KL-A32E (magenta), KL-T54E (green), and KFLDV (yellow) ....... 72 Figure II-7. A. Proposed nucleophilic attack of lysine amino group to the aldehyde following the Bilrgi-Dunitz model; B. Surface representation of CI5-aldehyde (yellow) bound in KL-A32E (gray). ................................................................ 78 Figure II-8. Stereoview of the overlay between holo-KL-A32E (green) and halo-KLE- R59W (magenta) .............................................................................................. 80 Figure II-9. Overlay between C‘s-aldehyde-bound KLE-R59W (blue) and apo-KLE- R59E (green). A. Model of C15-aldehyde in apo-KLE-R59E; B. Crystal structure of Cis-aldehyde bound in KLE-R59W. ............................................ 81 Figure II-10. Stereoview of the overlay of Cls-aldehyde-bound KL-A32E (green) and retinol-bound RBP (pink) ................................................................................ 84 Figure II-l 1. C15-aldehyde bound in the KLE-R59W-CRABPII mutant. ......................... 87 Figure II-12. Structures of merocyanin and MES. ............................................................ 89 Figure II-13. Merocyanin bound in KLE-R59W-CRABPII mutant .................................. 91 Figure III-1. Stereoview of the mutation sites involved in these CRABPII mutants ...... 101 Figure III-2. Structure of theapo-KLE-R59E-CRABPII ................................................ 104 Figure III-3. Structure of the apo-KFLE-CRABPII ........................................................ 106 Figure III-4. Structure of the apo—KFLD-CRABPII. ....................................................... 107 Figure III-5. Structure of the apo-KFLD-T54V-CRABPII ............................................. 109 Figure III-6. Structure of the apo-KL-T54E-CRABPII ................................................... 110 xiii Figure III-7. Structure of the apo-KL-A32E-CRABPII .................................................. 111 Figure III-8. Stereoview of the overlay of all apo structures together ............................. 113 Figure IV—l. all-trans retinoic acid and 9-cis retinoic acid. ............................................ 122 Figure IV-2. Structural organization of nulear receptors ................................................. 123 Figure IV-3. Comparison of the crystal structures of the apo-RXRa (A, PDB: lLBD) and holo-RARy (B, PDB: 2LBD) ligand-binding domains reveals the ligand- induced conformational change ..................................................................... 124 Figure 1V-4. The synthetic retinoid analog CD270. ........................................................ 127 Figure IV-5.1n vitro binding assay .................................................................................. 129 Figure IV-6. Detergent titration of the protonated Schiff base formed between KLE- CRABPII and retinal ...................................................................................... 131 Figure IV-7. Fluorescence titration of RARy-LBD with all—trans-retinoic acid. ............ 133 Figure IV-8. Titration of KLE-CRABPII with retinal. .................................................... 134 Figure IV-9. In vitro binding assay using RARy-LBD directly from Ni-NTA resin. ..... 136 Figure V-l. Pure fractions of His-rCRALBP eluted after Ni—NTA purification. ............ 141 Figure V-2. Non-fusion CRALBP purified. .................................................................... 142 Figure V-3. UV titration of non-fusion CRALBP with ll-cis-retinal. ........................... 143 Figure V—4. Full length CRALBP expressed in pET28a and purified with his-tag cleaved ........................................................................................................... 144 Figure V-5. N-43/44-truncated CRALBP expressed from pET28a and purified by Ni- column ........................................................................................................... 147 Figure V-6. Expression of the truncated CRALBPs ........................................................ 148 Figure V-7. UV titration of full-length and truncated CRALBPs with l 1-cis-retinal.....l49 Figure V-8. Gel filtration separation of full-length-CRALBP. ....................................... 152 Figure V-9. UV titration of different fractions from gel filtration separation of A20- CRALBP. ....................................................................................................... 153 Figure V-lO. CD spectra of different fractions from gel filtration separation of A20- CRALBP ........................................................................................................ 155 xiv Figure V-l 1. Gel filtration purification of the A20-CRALBP grown at different temperatures ................................................................................................... 1 57 Figure V-12. Trypsin digestion performed for both full length and A20-CRALBP under 4°C. ................................................................................................................ 1 58 Figure V-13. Crystals of merocyanin-bound proteins. A. merocyanin-bound A20- CRALBP; B. merocyanin-bound KLE-R59W-CRABPII. ............................ 159 Figure VI-l. R132K:Y134F:R111L:L121E-CRABPII (KFLE) purified after source-ISQ anion exchange column. ................................................................................ 162 Figure VI-2. Elution fractions of RARy-LBD from Ni-NTA purification. ..................... 164 Figure VI-3. Pure fractions of the his-tagged rCRALBP eluted off the Ni column. ....... 166 Figure VI-4. Non-fusion CRALBP after purification. .................................................... 166 Figure VI-5. Primers used for constructing different N-tenninal deletion CRALBP ..... 167 Figure VI-6. Typical size-exclusion purification profile of CRALBP represented by A20- CRALBP grown. at 16°C. .............................................................................. 169 Figure VI-7. CRALBP after purification ......................................................................... 169 Figure VI-8. Pictures of crystals of the ligand-bound CRABPII mutants ....................... 171 Figure VI-9. Pictures of crystals of the apo-CRABPII mutants ...................................... 174 Figure VI-10. Ramachandran plot of the C15-aldehyde-bound-KLE-R59W-CRABPII. 182 Figure VI-l 1. Ramachandran plot of the C15-aldehyde-bound-KL-A32E-CRABPII. 1 83 Figure VI-12. Ramachandran plot of the C15-aldehyde-bound-KL-T54E-CRABPII. 1 84 Figure VI-13. Ramachandran plot of the C15-aldehyde-bound-KFLDV-CRABPII ........ 185 Figure VI-l4. Ramachandran plot of the merocyanin-bound-KLE-RS9W-CRABPII.... 186 Figure VI-15. Ramachandran plot of the apo-KLE-R59E-CRABPII. ............................ 187 Figure VI-16. Ramachandran plot of the apo-KFLE—CRABPII. ..................................... 188 Figure VI-17. Ramachandran plot of the apo-KFLD-CRABPII. .................................... 189 Figure VI-18. Ramachandran plot of the apo-KFLDV-CRABPII. ................................. 190 Figure VI-19. Ramachandran plot of the apo-KL-T54E-CRABPII. ............................... 191 XV Figure VI-20. Ramachandran plot of the apo-KL-A32E-CRABPII ................................ 192 Images in this dissertation are presented in color. xvi Amino Acids Ala, A Arg, R Asn, N Asp, D Cys, C Gln, Q Glu, E Gly, G His, H Ile, I Leu, L Lys, K Met, M Phe, F Pro, P Ser, S Thr, T Trp, W Tyr, Y Val, V ABBREVIATIONS Alanine Arginine Asparagine Aspartic acid Cysteine Glutamine Glutamic acid Glycine Histidine Isoleucine Leucine Lysine Methionine Phenylanaline Proline Serine Threonine Tryptophan Tyrosine Valine CRABPII Mutants Abbreviations KFLD KFLDV KFLE KFLEV KL KL-A32E KL-T54E KLE KLE-R59E KLE-R59W R132K:Y134F:R111L:L121D R132K:Y134F:R111L:L121D:T54V R132K:Y134F:R111L:L121E R132KzYl34F2R111L:L121E:T54V R132K1R111L R132K:R111L:A32E R132K2R111LzT54E R132K:R111L:L121E R132K:R111L:L121E:R59E R132K1R1 11L:L121E:R59W Other Symbols and Abbreviations A Abs Angstrom Absorption xvii BTP Cor CCP4 CD cGMP Chlor. CRABPII CRALBP CRBP C-terminal Da DEAE DNA dNTP DTT 8 E. coli FABP FPLC FTIR Gar GDP GMP GPCR GTP h HTP iLBP IPTG Kr) L LB Amax Ampicillin Advanced Photon Source Base pair Bis-tris-propane The alpha carbon in the peptide bond Collaborative computational project, number 4 Circulat dichroism Guanisine 3’, 5’-cyclic monophosphate Chlroamphenicol Cellular retinoic acid-binding protein 11 Cellular retinaldehyde-binding protein Cellular retinol binding protein Carboxy terminal Dmmn Diethylarninoethyl cellulose Deoxyribonucleic acid Deoxynucleotide triphosphate Dithiothreitol Extinction coefficient Escherichi coli Fatty acid binding protein Fast protein liquid chromatography Fourier transform infra red Transducin, activated Guanisine diphosphate Guanisine monophosphate G-protein coupled receptor Guanisine triphosphate Hour Hydroxyapatite Intracellular lipid binding protein Isopropyl-l-thio- B-D-galactopyranoside Dissociation constant Liter Luria broth Wavelength of the maximum absorption xviii MES ul ml MOPS MPD MW nm nM NMR N-terminal PCR PDB PDE PEG PSB R1! RA RAR RARE RBP R-factor Rmsd RNA RPE RXR SB SDS-PAGE Tris UV vdw vis WT Molar 2-Morpholinoethanesulfonic acid, monohydrate microliter milliliter millimeter 3-(N-Morpholino)-propanesulfonic acid 2-Methyl-2,4-pentanediol Molecular weight Nanometer Nanomolar Nuclear Magnetic Resonance Amino terminal Polymerase chain reaction Protein Data Bank Phosphodiesterase Polyethylene glycol Protonated Schiff base Rhodopsin, activated Retinoic acid Retinoic acid receptor RAR-response element Retinol-binding protein Reliability factor root mean square deviation Ribonucleic acid Retinal pigment epithelium Retinoid X receptor Schiff base Sodium dodecyl sulfate — polyacrylamide gel electrophoresis 2-Amino-2-(hydroxymethyl)-1 ,3-propanediol Ultraviolet light Van der Waals Visible light wild type xix CHAPTER I. Introduction 1.1. The Mystery of Vision Vision is a precious gift that deeply enriches our lives. Through vision, one is able to perceive the order and beauty of the world. The visual process is one of the most delicate and efficient mechanisms of the human anatomy. Given the fascinating nature of vision, it is not surprising that people have been intrigued by the question of how one is enabled to see since the early days of human history. Ancient Greek philosophers, Plato (ca. 450 BC), Democritus (ca. 350 BC), Aristotle (ca. 330 BC) and Euclid (ca. 300 BC), were the first known to seek the explanations of vision (1). However, only Aristotle's theory, which was later extended by his student Theophrastus (ca. 300 BC), was close to capturing the true nature of this process: light is reflected by the objects, travels through the medium (air), and is then received by the eye (2). Although not popular at its time, this theory was improved throughout the centuries and eventually revisited by Johannes Kepler (1571-1630) in 1604 (3). He postulated the first complete description of the optics of the eye, refuting the once-popular theory that vision is enabled when the objects are being touched by the emissions sent out from the eyes. In 1704, Isaac Newton (1642-1727) published the book Optic/rs, which revolutionized the understanding of the laws that govern light (4). Most of his insights on the physical and physiological nature of color remain valid today. In the 18303 it was discovered that the perception of light happens in the rod and cone cells of the retina, the inner part of the eye. In 1878 the photosensitive protein in the rod cells, now known as rhodopsin, was isolated by F. W. Kuhne (1837-1900) (5). The active research of the subject over many years finally led to the revolutionary discovery by Wald (1906-1997) and coworkers that greatly advanced our understanding of the first steps in the vision process (6-8). Vitreous gel Figure [-1. The human eye (9). In a simplified picture, the current model of the vision process can be depicted as follows. When light enters the eye through the pupil, it is immediately focused by the lens before being sensed by the photoreceptor cells in the retina (Figure I-l) (10). Through a cascade of complicated events, the absorption of photons is converted into specific electrical signals in the photoreceptor cells. These signals are then transmitted in specific patterns out of the photoreceptor cells, through the optic nerves, and onto certain parts of the brain. 2pm rod rhodopsins \ cone CHO Disc\ 1 1-cis-retinal diskb«~ __-- cytoplasm _-, 35 um (~1000 disks) membrane o n -——-————-—-—-———0 n * a ‘ r .1 ' 9. ‘Iv‘ ., .r-.v {HHH‘H ”Y1 rhodopsin molecule Figure [-2. A. rod and cone cells in retina (11); B. rhodopsins embedded in the disks in photoreceptor cells (12). Through decades of hard work by many scientists, a more thorough and complete picture of this process has been achieved. For vertebrates, in the retina where vision initiates, two types of photoreceptor cells, rod and cone, exist (Figure I-2A) (13). The human eye contains approximately 100 million rod cells and about 6 million cone cells (14). Both types of the photoreceptor cells contain the visual pigments, known as rhodopsins. Rhodopsins refer to the specific proteins, called opsins, when they are bound to retinal via a Schiff base linkage with the active site lysine. In human eyes, each rod cell uses rod rhodopsins as its sole visual pigments, which are responsible for dim vision, while each cone cell contains its own specific rhodopsins: blue, green or red opsins, respectively. The blue, green and red cones are responsible for color vision. In the dark state, all four opsin proteins bind to the same natural substrate, ll-cis-retinal. Although the four rhodopsins differ in their absorption spectrum, they share the same mechanism in the process called phototransduction (15-1 7). As shown in Figure I-2A, a photoreceptor cell, either rod or cone, contains an outer segment and an inner segment (18). Taking the rod cell for example, in the outer segment hundreds of phospholipid bilayer disks are stored in an ordered, parallel fashion. Rhodopsins, the 7-helical trans-membrane protein, are embedded inside these disks with an orientation that is mostly perpendicular to the disk (Figure I-2B). The rhodopsins occupy ~50% of the disk surface area. Although it varies between species, each rod cell typically contains 1700 disks and each disk contains up to 1.5 million rhodopsin molecules (19). The inner segment contains the mitochondria, the nucleus and the synaptic end. The human vision process starts with the photolysis of rhodopsin, which involves a Cis-trans isomerization of the bound ll-cis-retinal. This activation-deactivation process of rhodopsin involves several thermal/dark steps characterized by intermediates identified at very low temperatures (Figure I-3) (20-25). In the dark state, ll-cis-retinal binds within rhodopsin through a protonated Schiff base (PSB) with Lys296 (26-29). Within femtoseconds after radiation, photorhodopsin is formed as a highly distorted "all-trans" PSB moiety (30), which then relaxes to the first isolatable intermediate, bathorhodopsin (31, 32). The formation of lumirhodopsin immediately follows within nanoseconds, followed by metarhodopsin I and then metarhodopsin II. The formation of metarhodopsin II happens milliseond after light absorption. Metarhodopsin 11 turns out to be the activated form of rhodopsin, which triggers the subsequent phototransduction cascade. The Schiff base is found to be unprotonated at metarhodopsin II. Following the metarhodopsin II activation, the imine bond is protonated to afford metarhodopsin 111. Finally, dissociation of all—trans-retinal from rhodopsin is achieved through hydrolysis. all- Inns, 385 nm opsin 1 1 -cls, 498 nm N/H \ 6) {>0' C min. hv <200 is “*1“ meta-ll. 0' C. 380 nm . activates transducin Unprotonated —" begins the visual .11/334‘3/ Schiff 3859 transduction process +H+ ‘H+ rns Photo. 570 nm “all-trans“ highly distorted meta-l, -15° C 480 nm ps us 11 (+3 ns . 12 H 497 nm Bathe, -140' C, 543 nm “distorted trans” Figure [-3. Schematic representation of the intermediates observed in the isomerization of ll-cis-retinal to all-trans-retinal in rhodopsin (21). The electric signal is generated as follows (Figure 1-4). In vertebrate, a dark current carried by Na+ and Ca2+ enters the outer segment of the rod cell through cGMP-gated channels keeping the cytoplasmic membrane relatively depolarized (33-35). The activation of rhodopsin through photon absorption generates metarhodopsin II (36, 37), which is able to bind to transducin, a G-protein consisting of three subunits: Tor, TB and Ty. It is believed that this interaction is achieved when the a-subunit of transducin binds to amino acids located in the second and third cytoplasmic loops of metarhodopsin (38- 41). This interaction triggers a conformational change in Ta, which involves the opening of the nucleotide-binding pocket and release of the guanosine-5'-diphosphate (GDP) bound (42). Transducin is then activated through binding to guanosine-5'-triphosphate (GTP) and subsequently the dissociation of To from TB and Ty, which eventually enables the signal transduction going from transducin to the next entity, the cyclic guanosine-5'- monophosphate (cGMP)-specif1c phosphodiesterase (PDE). PDE is a membrane-bound protein consisting of four subunits: Pa, PB and 2Py. The catalytic activity of PDE is enabled by the two subunits, Pa and PB. In the inactive state, the two inhibitory Py subunits block the activity of PDE. Activation of PDE is achieved through binding with Tor-GTP and subsequent removal of the two inhibitory Py subunits (43, 44). The activated PDE then hydrolyzes the cGMP into 5'-GMP, reducing the cellular concentration of cGMP. The lowered concentration of cGMP signals the closing of the cGMP-gated ion channels, which then results in interruption of the dark current. In this way, the absorption of light is transformed by the rod cells to a hyperpolarized electrical signal which is subsequently transmitted to the synaptic end and sent out to the brain through the optic nerves. .annEoE «Emma 05 5 £8530 :2 cucumdfiOo mo 2320 05 E 3:52 33:55 3285:0280 A2200 mo _o>o_ 533 dong—588 came—m8? m: mics»: £200 Co max—83: 05 DRESS Aim—Q5 man— uoumzuoa BE. .3833. > 05 mo coufioowmmu E 330:8 .3833 > m: 55 mason—BE t3 mam magnum 382.333 *0 GPO a c8 mQOAVgon a mammfifioxo .3 9.0V 3338a 0583 WESESE 05 :25; com: £38355 3208:: 808:8 :05 .& scam 6.5 Emmovofi 8558 5 9 mauso— =Qofi mo 53.8%? AWE cage nouguoaSesm .vA 2:»:— l QEGU 12.0 Phototransduction in vertebrate involves a light-activated isomerization of a double bond, which triggers a cascade of events and eventually leads to vision. This process is one of the most efficient and sensitive biological processes known, with a quantum yield of ~O.67 (46). The energy amplification of the visual process is calculated to be in the order of 100,000 based on quantitative analysis of the electrical response to a single photon (47, 48). On the molecular level, the amplification is achieved in primarily two steps. The first step lies in the interaction between metarhodopsin II (R*) with transducin (T). Within a fraction of a second, one molecule of R* is able to contact hundreds of transducins. The second stage of amplification is provided by the activity of PDE which catalyzes ~1000 molecules of cGMP per second (49, 50). The total amplification therefore equals to ~100 x 1000 = 100,000, which is in agreement with the calculations based on the electrical response. This explains the excellent sensitivity of vision that allows us to see even in the darkest environments. The above steps are involved in the rising phase of the light-signal response. The inactivation of this response is also important and involves mainly three mechanisms. The first mechanism is the guanylate cyclase-catalyzed generation of cGMP from GTP. This process compensates the loss of cGMP caused by light-activation. The decreased cGMP concentration signals closing of the ion channels, which leads to a decrease in the concentration of calcium (35, 51). The activity of guanylate cyclase is stimulated by the 2+ . . . . lowered Ca concentration in the rod cell. The cGMP concentration then increases so that the cGMP-gated channels are open and the dark current is restored. The second mechanism of inactivation is through phosphorylation of rhodopsin. The major role of rhodopsin phosphorylation, mediated by rhodopsin kinase and protein kinase C, is to promote arrestin binding and decrease transducin binding (52, 53). The third mechanism involves the transducin-catalyzed hydrolysis of GTP to GDP (54, 55). The GTPase activity is greatly accelerated during the rising phase of the light activation, which later on leads to reduced concentration of activated transducin, Tor- GTP. Phototransduction in invertebrates shares a great deal of similarity with that of vertebrates in the overall strategy. Light activation of rhodopsin triggers a series of intermediates leading to metarhodopsin. Similarly, this metarhodopsin also triggers a cascade of events leading to an electrical response. One difference is that the metarhodopsin in invertebrates is stable and inter-convertible with rhodopsin. Another interesting difference of the phototransduction in invertebrates is that the electrical signal is generated by depolarization of the plasma membrane, which is caused by the light- activated closing of the ion channels (56). Overall, the vision process in invertebrates shares several important features and similar sensitivity with that in vertebrates (5 7, 58). 1.2. Rhodopsins as mammalian visual pigments: structure and function in photoactivation. Rhodopsin is a 39KD protein, consisting of 348 amino acids and two oligosacchride chains (59, 60). It belongs to the family of G protein-coupled receptors (GPCR), which is by far the largest class of cell surface signal transducing receptors. Similar to the other membrane-bound GPCR family proteins, rhodopsin features a typical 7-helical structure. Y268 E113 1 1 -cis-natinal Figure [-5. Crystal structure of bovine rod rhodopsin. A. Overall fold; B. Binding of 11-cis-retinal in rhodopsin. Our understanding of rhodopsin was greatly advanced with the determination of the crystal structure of bovine rod rhodopsin at 2.8 A in the year 2000 by Palczewski and coworkers (28). This first-ever crystal structure of rhodopsin was able to define 95% of the amino acids as well as the posttranslational modifications (Figure I-5A). Continued efforts from different research groups has since yielded rhodopsin crystal structures in different crystal forms and improved resolutions to 2.2 A (61, 62). As revealed by the crystal structure, ll-cis-retinal is bound to the transmembrane portion of rhodopsin through a covalent trans-imine bond with Ly5296 (Figure I-SB). This confirms the protonated Schiff base (PSB) formation between the chromophore and Ly5296 that was suggested previously based on other biochemical evidence (26, 27). Glu113 is identified as the counter anion that stabilizes the PSB 3.45 A away. This electrostatic interaction 10 between the PSB and the carboxylate counter anion is found to be conserved only among visual pigments and is considered to be important for wavelength shifi of the bound chromophore as well as for maintaining the inactive, dark state conformation of rhodopsin (63-66). Prior to the determination of the crystal structure of rhodopsin, the role of Glu113 acting as a counter anion had been identified through mutagenesis studies (65). Mutation of Glu113 to Gln113 leads to a dramatic change in the absorption maximum from 500 nm to 380 nm, which is accompanied by a decrease in pKa from over 16 to around 6. Rhodopsin also features a compact extracellular (intradiscal) portion, part of which folds inward to enclose the bound chromophore (67). The two oligosacchrides are also present in the extracellular part of the protein, linked through two Asn residues, Asn2 and AsnlS, at the N-terrninus. The cytoplasmic portion of rhodopsin is not as rigid as the extracellular portion. In fact, the B-factors for this region vary considerably among different ground—state crystal structures, indicating the flexible nature of this part of rhodopsin (68). The transmembrane (TM) portion of rhodopsin constitutes as much as 60% of its secondary structure (69, 70). This region also contains the majority of the conserved residues in rhodopsin and other GPCRs, many of which are suggested to play important roles in maintaining the stability of rhodopsin. As mentioned earlier, the initial light activation of rhodopsin leads to several photoproducts. The crystal structures of some of these photoproducts have been successfully obtained after trapping these photolyzed states at low temperatures (71-73). Three-dimentional structures have been obtained for bathorhodopsin and lumirhodopsin. A structural model for metarhodopsin I has also been obtained based on electron 11 crystallography of two-dimentional crystals. Even though these structures are only at moderate to low resolutions (2.7 - 5.5 A), comparison of them with the dark-state structure of rhodopsin reveals that the double bond between C11 and C12 in retinal adopts the trans conformation upon light-activation (74). From bathorhodopsin to lumirhodopsin, a positional/confonnational change of the [ii-ionone ring of retinal has been suggested, but only by low-resolution structures. A structural model of an intermediate containing a deprotonated Schiff base has been obtained from lumirhodopsin crystals (75). This metarhodopsin II-like photoproduct seems to be similar to the dark state rhodopsin in overall structure. However, parts of the second and third loops in the cytoplasmic portion are found to be disordered in this photoactivated rhodopsin. In addition, residues Phe212, Trp265, Leu266 and Tyr268, were found to make direct contact with the B-ionone ring of the chromophore, which is consistent with observations from solid state NMR studies. 1.3. Other rhodopsins Rhodopsins exist in all three domains of life, Archaea, Bacteria and Eukarya (76). However, not all rhodopsins fimction as visual pigments as they do in higher animals. In fact, microbial rhodopsins represent a diverse group of photochemically reactive proteins, whose functions involve converting light energy for either ion transport or photosensory movement of cells (77). A generally accepted hypothesis that unifies microbial rhodopsin transport and signaling is that the accessibility of the Schiff base is switched 12 from the periplasmic side to the cytoplasmic side through structural changes, which opens the cytoplasmic channel for ion pumping and contributes to signal relay in the sensors (78). Bacteriorhodopsin, isolated from the haloarchaea Halobacterium halobium, was the first microbial rhodopsin discovered (79). It functions as a proton pump through converting light energy into a proton gradient across the bacterial cell wall (80). Bacteriorhodopsin binds all-trans-retinal through a protonated Schiff base with Ly3216 (81). The PSB is stabilized by two Asp residues through water-mediated hydrogen bonding interactions. The pKa of this PSB was measured to be over 12 (82, 83). With the chromophore bound, bacteriorhodopsin is able to absorb at 570 nm (84). Similar rhodopsin proton pumps have been observed in systems other than Haloarchaea, including marine Archaea (euryarchaeotic rhodopsins) (85) and Bacteria (proteorhodopsin and xanthorhodopsin) (86, 87). A similar case is halorhodopsin which is found in haloarchaea (88). Halorhodopsin functions as a chloride pump, transporting chloride ions into the cytoplasm and hyperpolarizing the cell membrane (78). In halorhodopsin, the PSB is stabilized by a tightly bound chloride ion which is translocated upon photoactivation. This unique structural feature enables the specific function of halorhodopsin. Halorhodopsin has also been found in the halophilic bacterium Salinibacter ruber (89). Another major class of microbial rhodopsins consists of the sensory rhodopsins (90). The first sensory rhodopsin, sensory rhodopsin I (SR1), was discovered in a search for receptors responsible for phototaxis in the haloarchaeal prokaryote Halobacterium salinarum (91, 92). SR] can signal the cells to be either attracted or repelled by the light, 13 depending on the light-activation mechanisms. Sensory rhodopsin II (SRII) is functional only as a repellent phototaxis receptor in haloarchaea. Both SRI and SRII have been well characterized both structurally and functionally. As revealed in the atomic structure of SRII, the PSB in SRII is stabilized by water-mediated hydrogen bonding interactions with two adjacent Asp residues, Asp85 and Asp212 (93). SRII absorbs at ~500 nm, similar to the rod rhodopsin in humans. The photoactivated intermediates of SRII have also been characterized using Fourier transformed infrared spectroscopy (FTIR) (94-97). Upon activation, both SRI and SRII form 2:2 complexes in haloarchaeal membranes with their cognate transducers, HtrI and HtrII, respectively. Formation of the complexes initiates the signal relay and is now under intense investigation. A unique sensory rhodopsin, Anabaena sensory rhodopsin (ASR), has been identified from the freshwater cyanobacterium Anabaena sp. PCC7120 (98). The crystal structure of ASR has been solved at 2.0 A resolution (99). The overall membrane- embedded seven helical structure of ASR resembles the other haloarchaeal rhodopsins. However, there are also striking differences separating ASR from other haloarchaeal rhodopsins. First, the interior of the cytoplasmic side of ASR is rather hydrophilic, compared to other haloarchaeal rhodopsins. These hydrophilic residues are stabilized through a water-mediated hydrogen bonding network, which extends toward the cytoplasmic surface where ASR interacts with its transducer. Second, ASR exists in a mixture of both all-trans and 13-cis retinal isomeric forms, with absorption maxima at 550 nm and 537 nm, respectively. Each of the two forms exhibits efficient light-induced conversion to the other, which results in the ratio of the two being dependent on the wavelength of illumination. This mechanism resembles that of the phytochrome in 14 plants, which led to the suggestion that ASR might be responsible for modulating chromatic adaptation in photosynthesis by triggering differential biosynthesis of light- absorbing pigments based on the quality of light. An interesting discovery reported recently shed light on the relationship between the transport rhodopsins and the sensory rhodopsins (100). Spudich and coworkers successfully converted the bacteriorhodopsin (BR) to a SRII-like sensor by replacing three residues in BR with the corresponding residues in SRII. The resulting protein is capable of relaying the retinal photoisomerization signal to the transducer HtrII, which then induces robust phototaxis responses. This finding not only identified the key residues in determining the functions of the two proteins, but also demonstrated the elegant and yet simple design by nature to achieve functional diversity from the same protein scaffold and similar ligand-protein complexation. 1.4. Color vision Rod rhodopsin is the most studied and best understood rhodopsin among the over 1000 visual rhodopsins discovered so far due to its importance in animal vision and also its abundance in the retina. The cone rhodopsins, however, are less understood despite their equal or even greater importance in vision. As mentioned before, rod rhodopsins are responsible for dim vision and cone rhodopsins are responsible for color vision. 15 420 498 534 564 100- Q 0 C N 'E O 3 ,I g 50“ ’1” 0 ’a .2 , ' .5 .- E O z G r 11 IlirIirtrlrrrrlrlrilrrrrirrfiri l 400 500 600 700 Wavelength (nm) Figure [-6. Absorption profiles of different human rhodopsins (101). In the human vision process, cone rhodopsins share similar photoactivation cascades as rod rhodopsins. However, the maximal sensitivities of each rhodopsin differ at different wavelengths. Human rod rhodopsin absorbs at 500 nm. The human cone rhodopsins: red, green and blue, absorb at 560 nm, 530 nm and 420 nm, respectively (102-105). All four rhodopsins exhibit red-shifted maximal absorption wavelengths afier binding to the chromophore ll-cis-retinal, which by itself absorbs at 380 nm as a free aldehyde. The combination of the four pigments constitutes the whole 400-700 nm range, which is defined as the visible spectrum (Figure I-6). 16 Table [-1. Rhodopsins of various species (106). Cone sensor group (nm) Species SW81 SW82 M M/L L Human 420 530 560 Dog 430 555 Horse 430 540 Squid 470/485 500 Black bream 475 560 Gold fish 360 450 540 625 Chicken 420 455 510 560 - Peking robin 355 455 500 570 Pacific shearwater 405 450 505 565 Butterfly 360 400 450 520 610 Housefly 355 460/49 530 Honey bee 350 445 550 Within animals, the requirement for color vision has been met in different ways, suggesting that color vision has developed independently several times during evolution (107). Depending on the number of types of cone photoreceptors present, some degree of color vision becomes possible. Primates express three types of cone opsins, while many animals have two and some insects, birds and fish have four (108). These are described as being trichromats, dichromats and tetrachromats, respectively. As a general classification, the cone rhodopsins can be roughly grouped into the following categories: very short wavelength (i.e. UV)-sensitive (SWSl), short wavelength ("blue")-sensitive (SWS2), medium wavelength ("green")-sensitive (MWS), and long wavelength ("red")- sensitive (LWS) (106). Table [-1 shows how certain cone rhodopsins are classified based on their maximal absorption wavelengths. Apparently, most of the cone rhodopsins display red-shifted absorption maxima (Max): while other cone rhodopsins are blue- shifted. 17 Though to different extents, the red shift of kmax observed in most of the rhodopsins had been found to be partly due to the formation of the protonated Schiff base (PSB) between 11-cis-retinal and the active site lysine. The discovery of the PSB started with the isolation of retinal from rhodopsin in 1933 by Wald (109). In 1944, Morton and Goddwin showed that retinal is the aldehyde form of vitamin A (110), which was confirmed to be 11-cis-retinal in 1950 (111). As early as 1955, the formation of the PSB between retinal and a lysine residue in rhodopsin was proposed to account for the bathochromic shift in visual pigments. During the late 1960's, Bownds and Akhtar, through independent studies, provided evidences for the formation of a Schiff base between retinal and a lysine within rhodopsin (112, 113). It was not until 1974 that protonation of the Schiff base was unambiguously confirmed by Callender through resonance Raman studies (114). As a free aldehyde, ll-cis-retinal absorbs at 380 run. When the Schiff base is formed with n-butylamine, a slight blue shifl in absorbance can be observed and the deprotonated Schiff base absorbs at 365 nm. However, when the Schiff base is protonated there is a significant red shifi of the chromophore’s absorbance to 440 nm (115, 116). With controlled concentration of the chromophore and careful choice of solvent this value may shift up to ~500 nm (117, 118). Protonation of the Schiff base alone, however, is probably not responsible for the spectral tuning in color vision as 440 nm is just the beginning of the region of the electromagnetic spectrum that we can actually see (~400 —- 700 nm). The differences in the absorbance observed for different visual pigments as compared to the PSB of 11-cis-retinal in solution are defined as opsin 18 shifts and are believed to result from specific interactions of each individual opsin protein with the retinylidene chromophore (119, 120). Several hypotheses have been proposed to explain the opsin shifts. Although progress has been made, a complete and satisfactory explanation for color vision is still missing. The following is a brief description of the hypotheses and accomplishments achieved in this field of research. In 1958, Hubbard made the first attempt to explain the opsin shift (12]). He not only suggested formation of a protonated Schiff base, but also proposed that specific electrostatic interactions between the chromophore and the protein could induce a delocalization of the PSB charge along the polyene backbone, leading to altered spectroscopic behavior of the bound chromophore. Later, in the 1970's, theoretical calculation studies suggested that a pigment protein should be able to regulate the wavelength shifts by placing polar or charged residues around the chromophore (122- 126). In 1976, Honig proposed, through theoretical calculation, that the PSB is closely associated with a counterion and additional negatively charged or polar groups are positioned by the protein in the vicinity of the ring half of the chromophore (123). Later, Honig provided firrther support for this external point-charge model using theoretical calculations based on spectroscopic data of model compounds ([23). Despite the lack of absolute validity, additional support from Nakanishi and others made this model generally accepted at the time (127-129). The confirmation of the PSB counter anion was achieved in 1989 when three independent studies indicated that Glu113, a highly conserved residue within vertebrate visual pigments (130), functions as the counter anion in bovine rhodopsin (63-65). 19 Replacement of Glu113 with Gln113 resulted in two species that absorb at 490 nm and 380 nm, respectively. These two species were identified as the protonated and unprotonated forms of the Schiff base which exist in a pH-dependent equilibrium. The formation of the Schiff base between the chromophore and Lys296 and the presence of the Glu113 as counter anion were both confirmed with the determination of the crystal structure of bovine rhodopsin in 2000 (28). R R R T: h h N fcoort N 000” N COOH n . R] COOH h choorr N\/' N\/\/ \\ \ R=Wsfl Figure [-7. Model compounds used to study wavelength shifts controlled by distance between the PSB and the counter anion (131). The distance between the PSB nitrogen and the stabilizing counter anion was also shown to affect the red shift of the bound chromophore. In 1972, Blatz suggested that the different absorption maxima in different opsins were caused by differed distances between the PSB and the counter anion, with increased distance leading to greater red shifi (132). This hypothesis was supported by the discovery in 1989 that replacement of Glu113 with Asp resulted in a red shift of rod rhodopsin Xmax to 505 nm, a 5 nm more 20 red shified pigment than the native rod rhodopsin (64). Furthermore, the effect of varying both distance and angle between the PSB and the counterion was also investigated. Sheves and coworkers synthesized a series of model compounds (Figure I-7), each adopting a unique arrangement of the PSB and a nearby carboxylate (131). The measured wavelength shifts as well as pKa values were shown to vary according to each molecule. However, the maximum difference in red shifted )‘vmax among these synthetic model molecules was only 10 nm, which is much smaller than the maximum difference observed between the visual pigments (I33, 134). Therefore, although the position and orientation of the counterion could affect the Max, it could not be the cause for the whole spectrum shift in different rhodopsins. Figure [-8. The point charge model. Positioning negative charges at different places along the polyene chain of the chromophore may modulate the maximum absorption wavelength. As discussed earlier Honig suggested in 1976 that additional charged or polar residues are placed close to the polyene chain (123). Following this hypothesis Nathans 21 suggested, by examining and comparing the gene sequences of all four human opsins, that the placement of different negatively charged amino acids along the polyene backbone in different proteins could result in extended or disrupted conjugation of the positive charge, which could be nature’s way of tuning the wavelength (Figure I-8) (130). In addition to the point charge theories that mostly focus on the electrostatic or polar interactions between the chromophore and the opsin protein is another series of hypotheses that emphasize the steric interactions between the two parties. An early proposal by Blatz suggested that twisting the planes about the single bonds of the chromophore would lead to different levels of conjugation and therefore wavelength regulation (Figure [-9) (135). Although the polyene system in retinal tends to adopt the planar conformation in order to achieve maximum conjugation, steric constraints in 11- Cis-retinal in fact results in a non-planar molecule. NMR studies had shown that the C6- C7 single bond of 11-cis-retinal bound in rhodopsin adopts a Cis-like conformation, namely 6-S-cis (128). Further investigation of the system revealed that, due to the steric interactions between the C5-Me and C8-H, the 6-S-cis conformation is indeed twisted to be non-planar. A similar situation is suggested for the non-planar 12-S-trans conformation, where the steric interaction between C13-Me and CIO-H forces a twisting in the C12-C13 single bond (136, 137). According to this theory, red rhodopsin should adopt the retinal in a more planar conformation, optimizing the orbital overlap and leading to a more red-shifted kmax. However, theoretical studies combined with NMR have also shown that the induced degree of conjugation does not appear to be solely responsible for the opsin shifts (I38). 22 Figure [-9. Hypothesis of steric interactions leading to wavelength regulation. Twisting the single bonds will reduce the degree of conjugation and therefore lead to different absorption maxima. Another hypothesis in explaining the opsin shift involves excitonic coupling between the chromophore’s excited states and nearby aromatic residues. Based on mathematical analysis of the excited state energies of 11-cis-retinal, the bathochromic shift in visual pigments were proposed to be caused by excitonic coupling, which could lead to the formation of superconducting pairs in the pigments (9). Light-induced perturbation of aromatic residues in bovine rhodopsins and bacteriorhodopsins (139) as well as photoexcited energy transfer in frog rhodopsins (11) were observed in support of the existence of such coupling. Recent spectral studies demonstrated the presence of excitonic coupling between retinal and tryptophan residues in the visual pigments. The coupling induced by alignment of dipole moments of the chromophore and a nearby tryptophan has been suggested to be responsible for driving the tortional motion of the chromophore and leading to the Cis-trans isomerization (12). However, no correlation between the excitonic coupling and spectral tuning in the visual pigments has been observed. 23 Figure I- 10. Vicinity of retinal in cone pigments as shown in the homology models (140). Charged amino acids are indicated in red. (A) Blue cone pigment; (B) Green; (C) Red. The determination of the first crystal structure of bovine rod rhodopsin in 2000 by Palczewski marked one of the most remarkable milestones in the field (28). This 2.8 A structure not only provided absolute evidence for many previous results and suggestions, but also paved the way for future research on other visual pigments. Continued efforts on the crystallography of rhodopsin has since yielded improved resolutions as good as 2.2 A (62). As shown in Figure [-58, the formation of the PSB between ll-cis-retinal and Ly5296 is confirmed. Glu113 acts as the counter anion at 3.45 A away from the PSB nitrogen. A negatively charged residue is observed to be over 5 A away from the double bond between C11 and C12, which is further away than what is expected based on both the hypothesis of Honig and the results from related NMR studies. However, two polar residues, Thr118 and Tyr268, are found to be in the vicinity of the central part of the polyene. Twisting of the single bonds is also observed, suggesting the contribution from the steric-induced non-planarity of the chromophore on wavelength regulation. The double bond between C11 and C12 has also been shown to adopt a twisted conformation as revealed by the 2.2 A-resolution structure, which is consistent with the predictions based on theoretical calculations. Furthermore, Trp265 is found in the structure to be near the ionpne ring as well as the C13-Me of the chromophore, suggesting possible excitonic coupling effects. The determination of the crystal structure also led to construction of homology models for the three human cone pigments using the rod rhodopsin structure as a template (140). The amino acid sequence identities for the blue, green and red cone pigments with rhodopsin are 41%, 38% and 37%, respectively. According to the homology models, the 25 secondary structures of all three cone opsins are nearly identical to that of the rod rhodopsin, exhibiting a seven helical transmembrane portion together with one cytoplasmic helix and two beta strands. The only difference, the same for all three cone opsins, is the one-amino-acid extension at the N-tenninal l3 strand. Similar to rod rhodopsin, the retinal binding environment is highly hydrophobic in all three cone rhodopsins (Figure I-10). Red and green rhodopsins share almost identical binding probably due to their high homology (96% homologous, differ by only 15 amino acids) (130). Both contain a Trp182 in the vicinity of the central part of the polyene. The two opsins also share an additional glutamate at 102, besides the counter anion Glu129 that is analogous to Glu113 in rod rhodopsin. The blue opsin, however, displays a much different arrangement of amino acids around the retinal. As shown in the model, the central residue forming the cavity in the blue cone pigment is Tyr262. Because of its proximity to the retinal B-ionone ring, Tyr262 is believed to be the major factor in the blue shift of this pigment. The additional glutamate observed in red and green rhodospin models is missing in the blue pigment model structure, which is suggested to be responsible for additional blue shift of the Amax. In fact, prior to the rhodopsin crystal structure, mutational studies involving specific key residues in visual pigments also yielded very interesting results. Inspired by the high sequence homology between red and green rhodopsins, Oprian and coworkers showed that the red cone can be converted into a green-rhodopsin-like pigment by implanting seven specific mutations, S116Y:Sl80A:123OT:A233S:Y277F:T285A:Y309F (Figure I-1l) (102). The artificial pigment is able to constitute a nearly 40 nm shift in the Xmax. Within the seven sites, three were identified to play more significant roles in the spectrum 26 I“ 8180 $116 1 1-cis-retinal K... .2... )1 Y309 Y2" a< T285 Figure [-1]. Illustration of the sites of mutations for converting the red rhodopsin to a green-rhodopsin-like pigment (102). A233 ’ Figure [-12. Illustration of the sites of mutations for converting the green rhodopsin to a red-rhodopsin-like pigment (141). 27 tuning. A triple mutant of the green opsin, F 277Y:A285T:A180$ (Figure I-12), is able to account for a majority of the red shift, going from green rhodopsin to a red-rhodopsin- like pigment (141). Interestingly, all three mutations are in fact conversions from hydrophobic residues to polar residues around the beta-ionone ring of the chromophore. A similar trend was also observed in mutagenesis studies of the rod rhodopsin. A rod rhodopsin mutant, F 261Y:A269T:A164S (Figure I-13), red shifts to 520 nm, achieving an amazing 200 nm red shift (142). A164 E113 1 1 -cls-retinal K296 Figure [-13. Illustration of the sites of mutations for converting the rod rhodopsin to a red-shifted pigment (142). In recent years, quantum mechanics (QM) combined with molecular mechanics (MM) (denoted as QM/MM) techniques have been used intensively in studying the spectral tuning in the visual pigments (143). Electrostatic interactions have been suggested to affect the wavelength shifts between different rhodopsins due to the fact that 28 the HOMO-LUMO transition is actually an intramolecular electron transfer from the ionone ring to the PSB (144-147). The charge distributions within the ll-cis-retinilidene were shown to be different for the ground states and excited states (148). In a most recent study done by Nakatsuji and co-workers, QM/MM calculations were performed based on the homology models of the cone pigments. The results suggested that the color tuning is regulated by amino acids at specific positions (145). More specifically, a common Ser (186 in rod rhodopsin) was shown to discriminate human blue rhodopsins from other human rhodopsins presumably due to its unique orientation relative to the It system of the chromophore. In the human red rhodopsin, Tyr277 and Thr285 (replacing Phe and Ala, respectively, in other human rhodopsins), were shown to provide negative electrostatic potential around the ionone ring, which differentiates the red human rhodopsins from other rhodopsins. Studies from Goddard and co-workers, however, indicated that conformational twisting is what causes major blue shift of the absorption maxima in blue pigments (149). They also suggested the red-shifting role of dipolar side chains in the binding pocket as a fine adjustment to the opsin shift. Besides, electronic polarization effects of the protein environment were also suggested to play a predominant role in causing bathochromic shifts in bacteriorhodopsins (150, 151). In addition, a counterion switch mechanism was proposed by Mathies and co- workers (152). Glu113 was shown earlier to be the counterion that stabilizes the PSB in the ground state. Mutation E181Q resulted in nearly unchanged Raman vibrational specctra in the ground state. However, it was shown that the same E118Q mutation led to a dramatic shift of the pKa of the PSB in the metarhodopsin II state. Mathies then 29 proposed that a counterion switch occurs within the photoactivation process through a proton transfer from Glu18l to Glu113. This is accompanied by a specific conformational change of the rhodopsin protein, leading to its activation. This model has been validated by several theoretical calculations (153, 154). Such mechanism is possibly responsible for the shifted absorption maxima observed for certain photo intermediates. In conclusion, environmental differences around the ionone ring and the PSB of the bound chromophore among different visual pigments seem to play major roles in the wavelength regulation mechanism. Other factors, including distortion of the planar conformation of retinal as well as excitonic coupling, might also contribute to the wavelength changes. The real spectrum tuning mechanism very possibly is achieved by a combination of these factors. The wavelength regulation mechanism is indeed much more complicated than anticipated. Alternative approaches are therefore desired in trying to provide a clearer picture of the mechanism. 1.5. Designing CRABPII mutants as rhodopsin mimics As concluded above, the hypotheses regarding the wavelength regulation in visual pigments need to be further tested. However, difficulties associated with working with the membrane-bound protein rhodopsins, such as complication in mutagenesis, low heterologous expression level, and difficulties in obtaining 3-D structures, still pose 30 problems. This is why our research team opted for a model system that can be used to elucidate the wavelength regulation mechanism. Our goal is to develop a model protein that mimics rhodopsin in the binding of the chromophore, retinal. Through protein engineering techniques, we created different situations for the bound retinal, such as placing charged or polar residues at various positions along the chromophore backbone, introducing steric interactions to force the non-planarity of the molecule, or placement of aromatic residues in the vicinity of the chromophore. For this particular purpose, this protein mimic should first show good affinity for retinal and also be very tolerant towards mutagenesis. Additionally, this system should also be more friendly than rhodopsin in heterologous expression, purification and crystallization, which would facilitate the determination of the 3-D structures and therefore provide valuable information in terms of both binding details and directions for future rational redesign. 1.5.1. CRABPII general background. Retinoic acid (RA), the biologically active metabolite of vitamin A, functions as a morphogen during embryonic morphogenesis (155) and has been suggested to play an indispensable role in regulating cell differentiation and growth (156-158). RA and synthetic retinoids have been used successfully for treatment of acute promyelocytic leukemia (APL), skin disorders, and epithelial malignancies, such as skin cancer and cervical cancer (155, 158-1 63). 31 Retinoic acid R111 R132 Figure [-14. Crystal structure of retinoic acid bound in WT-CRABPII (164). A. Stereoview of the overall fold of the protein; B. Binding of retinoic acid in the active site. RA exerts its regulatory function in the nucleus by association with the transcriptional activator, retinoic acid receptor (RAR) (155, 165-170). Upon complexation with RA, the RAR/RXR (retinoid X receptor) heterodimer recognizes the RA-response elements (RARE) in the promotor region of certain genes and regulates their transcription. The association of RA to RAR is facilitated by cellular retinoic acid binding proteins, CRABPI and CRABPII (171-174). CRABPs belong to the intracellular Lipid Binding Protein (iLBP) family and are present in virtually all vertebrates (175). The 32 overexpression of CRABPs has been reported in a wide variety of human cancers, suggesting important roles in cancer development (176-183). CRABPs specifically bind to RA and modulate its cellular concentration in the cytoplasm. The RA delivery mechanism has been suggested to be different between CRABPI and CRABPII. While CRABPI shows no evidence of interacting with RAR, CRABPII shuttles RA to RAR through direct protein-protein interaction (184, 185). Nuclear localization of CRABPII, but not CRABPI, upon ligand binding has also been observed, providing further evidence for its interaction with the nuclear receptor RAR (186). Recently we reported the first crystal structure of apo-WT CRABPII (187). The crystal structure of WT-CRABPII-RA has been reported previously by other groups (188), however a higher resolution structure of this complex (Figure I-14), determined at 1.48 A, is now available (187). In the latter work, using three different data sets collected on apo-WT CRABPII we showed that apo- and holo-CRABPII share very similar structures. Binding of RA appeared to increase the overall rigidity of the structure, although the induced structural changes were not as pronounced as previously thought (1 89). The enhanced structural rigidity may be an important determinant for the enhanced nuclear localization of the RA-bound protein. Interestingly, water-mediated interactions are observed in the cavity of apo-WT-CRABPII MolA, connecting the Argl 11 deep inside the cavity through Arg132 to Ala36 and Val33 both located on the (12 helix (Figure I-15). The water-mediated network is not observed in MolB probably due to crystal packing. The difference in crystal packing environments between MolA and MolB at the (12 helix region is significant (Figure [-16), which is apparent from the number of intermolecular interactions that each a2 helix makes with the neighboring 33 molecules: MolA has 0 hydrogen bond and 8 vdw contacts and MolB has 7 hydrogen bonds and 43 vdw contacts. 2.9.. 2.7 .0‘ Acetate ,o' 204 R111 R132 Figure [-15. Water-mediated hydrogen bonding interactions observed in the molecule A of the apo-WT-CRABPII crystal structure (164). Similar water-mediated networks have been observed in other members of the iLBP family, such as CRABPI (190) and intestinal fatty acid-binding protein (191), and were suggested to play important roles in maintaining the stability of these proteins as well. In the rat intestinal fatty acid-binding protein (FABP) structure (IIFC, 1.2 A), the water- mediated network connects the Arg106 (conserved residue of Argl 11 in CRABPII) through Arg126 (conserved residue of Arg132 in CRABPII) to Asp34 and Ala32 on the (12 helix at the portal. In the CRABPI structure (lCBI), the network is observed in part, but the lower resolution of the structure (2.7 A) probably leads to an incomplete observation of all the ordered water molecules in the cavity. In the high resolution structure of cellular retinol-binding protein 11 (CRBPII, 1.2 A, PDB ID: 2RCQ), which is also an iLBP family member, the water-mediated network is again observed, which also 34 resembles the one in apo-WT-CRABPII. In the apo-CRBPII structure, the network starts at the Glu108 (conserved residue of Arglll in CRABPII based on similarity) and extends to Ala36 and Val33 (both conserved in CRABPII based on identity). These observations further emphasize the structural importance of these water-mediated networks for the iLBP family proteins. a Figure [-16. Crystal packing observed in the structure of apo-WT-CRABPII. A. Molecule A; B. Molecule B. 1.5.2. Designing CRABPII mutants as rhodopsin mimics. CRABPII was chosen as our model protein for mimicking rhodopsin. First, all- trans retinoic acid, the natural substrate of CRABPII, resembles retinal that is bound in rhodopsin. This similarity therefore eases our way in introducing retinal into our model systems, minimizing the risks in redesigning an entire binding site. Second, CRABPII is a small and soluble protein that is easily expressed and purified in large quantities using 35 recombinant expression systems, which represents a huge advantage over the membrane- bound rhodopsin. Third, CRABPII and other iLBP family proteins are well known for their excellent tolerance of multiple mutations (192). The stability of CRABPII in the presence of multiple mutations provides us additional freedom in introducing mutations for protein redesign. Fourth, CRABPII mutants can be readily crystallized routinely using established conditions, leading to three-dimensional structures usually at atomic resolution (189, 193, 194). This feature contrasts drastically with that of rhodopsin, which is more difficult to crystallize. The excellent tendency of CRABPII toward crystallization gives us routine access to detailed structural information, leading to more rapid data acquisition/analysis. Previously, through rational design, CRABPII mutants were obtained that were capable of forming protonated Schiff bases (PSB) between all-trans-retinal (Rt) and an active site lysine introduced through site-directed mutagenesis (195, 196). A triple mutant, R132K:R111L:L121E (KLE), is of particular interest due to its excellent ability to shift the absorption maxima of the chromophore as well as its nanomolar dissociation constant toward Rt (Figure I-17) (196). Besides the active site Lysl32, it has been proven that the Glu121 that provides the counterion stabilization for the PSB is also required for the bathochromic shift. The mutation of Arg to Leu at position 111 is also required in that an ordered water molecule is removed upon this mutation, which helps place the aldehyde group of Rt in a favorable position for nucleophilic attack and eventually leads to PSB formation. 36 All-trans-retinal Figure I-17. All-trans-retinal bound in KLE-CRABPII (196). The triple mutant KLE-CRABPII served as a great example, demonstrating the importance of formation of the PSB as well as a stabilizing counter anion nearby. However, although the red shift to 447 nm observed for the all-trans-retinal-bound KLE- CRABPII complex is impressive, the mechanism identified within it does not seem to fully explain the wavelength regulation. As shown earlier, the PSB formed between retinal and n-butylarnine absorbs at 440 nm and the red cone rhodopsin absorbs at 564 nm. Apparently, other mechanisms besides the PSB formation and the counter anion interaction must be fimctioning in order to achieve the dramatic red shift in red rhodopsin. Our next mission then is to further engage our model system into testing the hypotheses proposed for the wavelength regulation mechanism in color vision. Most recent work in this project features manipulation of the absorption maxima of rationally designed CRABPII mutants when bound to a shorter retinal analog, CI5-aldehyde. 37 Chapter II of this dissertation details about the X-ray crystallographic studies of such C15-aldehyde-CRABPII mutant complexes. Chapter II also includes brief descriptions of the results from the biochemical studies performed by other colleagues. Chapter [[1 describes the structural insights derived from the crystal structures of the holo-CRABPII mutants in terms of structural integrity upon mutations and the structural role of the water-mediated network. Chapter IV describes an effort that we took toward understanding the direct channeling of retinoic acid from CRABPII to RAR through protein-protein interactions. 1.6. Cellular retinaldehyde-binding protein and the 1 1-cis-retinal regeneration. Cellular retinaldehyde binding protein (CRALBP) is a water-soluble protein found in the retina and pineal gland (197). This 36 KD protein is remarkably stereoselective, carrying as endogenous ligands 11-cis-retinol and 11-cis-retinaldehyde (ll-cis-retinal) that are only known to function in vision. Notably, ll-cis-retinal bound to CRALBP is less susceptible to photo-isomerization than when bound to rhodopsin (198). Mouse knockout studies have shown that CRALBP gene defects impair regeneration of visual pigments (I 99). 38 \ \ \ \ \ \QRhodopsin O <—2— | H all tr ans- retinal all-trans-retinilidene all-trans- RDH/ \\photon W0, all- trans-retinol nod outeroegment E? Rhodopsin H E 2 \ \ \ 2 I \ OH g g all- trans- retinol 11-cis-retinal CHO ° \LRAT /f 1-cis-RDH '3. .7; npees 3 all- trans-retinyl ester 11-cis-retinol OH L J Figure I-18. The visual cycle. Retinoids undergo intercellular transport and several enzymatic transformations during the course of the visual cycle (Figure I-18). All-trans-retinal, the product of visual cycle bleaching, is reduced to all-trans-retinol within photoreceptor cells and transported through an extra-cellular compartment to the retinal pigment epithelium (RPE) where it is esterified by lecithin-retinol acyltransferase (LRAT) and then isomerized to 11-cis-retinol by the isomerohydrolase (RPE65). A dehydrogenase-mediated oxidation in RPE in mammals produces ll-cis-retinal, which must then traverse the extracellular compartment to gain access to the photoreceptors where regeneration of visual pigments occurs. CRALBP functions in retinal pigment epithelium (RPE) as a major acceptor of ll-cis-retinol in the isomerization step of the rod visual cycle. It also serves as a 39 substrate carrier for ll-cis-retinol dehydrogenase (RDH5), facilitating the oxidation of ll-cis-retinol to ll-cis-retinal (197, 198, 200-203). Direct CRALBP interactions have been demonstrated in vitro with ll-cis-retinol dehydrogenase and with ERM (ezrin, radixin, moesin)-binding phosphoprotein 50 (EBPSO), also known as sodium hydrogen exchanger regulator factor type 1 (NHERF-l) (204). Interactions with EBP50 have been suggested as a mechanism for localizing CRALBP to the apical RPE plasma membrane for export of 11-cis-retinal to the adjacent rod photoreceptor cells for visual pigment regeneration. Ongoing proteomic studies also support the existence of a RPE retinoid processing protein complex containing CRALBP (205). Due to the lack of crystal structures of this protein, the interaction of human recombinant CRALBP (rCRALBP) with retinoids has been investigated by other methods. Based on preliminary NMR studies (206), Crabb and co-workers first targeted Met and Trp residues as important sites for retinoid interaction (203). They created five mutants, M208A, M222A, M225A, W165F, W244F, and evaluated retinoid binding properties and substrate carrier functions. It turned out that all mutants bound ll-cis- retinal and 9-cis-retinal and therefore were not grossly misfolded. However, altered UV- visible spectra and lower retinoid binding affinities were observed, supporting modified ligand interactions. Altered kinetic parameters were also observed for RDH5 oxidation of 11-cis-retinol bound to rCRALBP mutants M222A, M225A, and W224F, suggesting impaired substrate carrier function. Heteronuclear single quantum correlation NMR analyses confirmed localized structural changes upon photoisomerization of rCRALBP- bound ll-cis-retinal and demonstrated ligand-dependent conformational changes for residues Met-208, Met-222, Trp-165, and Trp-244. From this study, all five residues 40 (Met-208, Met-222, Met-225, Trp-165, and Trp-244) have been implicated to be components of the rCRALBP ligand binding cavity. Later, Crabb and co-workers further investigated the binding cavity using photoaffrnity labeling and hydrogen/deuterium exchange (207). Photo-induced covalent labeling employing synthetic ligand 3-diazo-4-keto-1l-cis-retinal, followed by LC-MS analysis, successfully identified eight residues being photoaffinity-modified: Tyr-179, Phe-197, Cys-198, Met-208, Lys-221, Met-222, Val-223, and Met-225. Among these eight, three of them (Met-208, Met-222, and Met-225) had already been identified in their previous investigation. In addition, topological analysis of apo- and holo-rCRALBP by hydrogen/deuterium exchange and mass spectrometry demonstrated that residues 198~255 incorporated significantly less deuterium when the retinoid binding pocket is occupied with ll-cis-retinal (203). This further confirmed that the identified residues were inside the retinoid binding pocket of rCRALBP. Based on sequence information, CRALBP has been classified as a member of the CRAL-TRIO lipid-binding protein family (Pfam entry: PF00650) (208). Crystal structures have been determined for three CRAL-TRIO lipid-binding proteins, including human or-tocopherol transfer protein (ATTP) (209, 210), yeast Sec14, a phosphatidylinositol-binding protein (211), and human supernatant protein factor (SPF) (212). CRALBP shares 20~33% sequence identity and ~50% homology with these three proteins. Structure models of CRALBP have been built based on the available crystal structures of the three family members (207, 213). Both open and closed conformation models have been produced (Figure I-l9). The model of the closed conformation agreed well with the residues identified earlier, which are involved in interactions with retinoid. 41 It has also been suggested based on MD simulations of the model that the open conformation of CRALBP participates in the ligand association and dissociation processes; the closed form serves as a relatively stable carrier of retinoids, protecting them from the cellular hydrophilic environment. Figure [-19. Stereoview of the structure models built for CRALBP (213). A. Closed form; B. Open form. N—temrinal starts from 66 (blue) in the rainbow presentations. Despite all the efforts of elucidating the structural mechanism of CRALBP interacting with retinoids, the lack of a crystal structure still poses serious problems for research associated with CRALBP. More reliable answers are needed for questions like how retinoid is released from the high affinity CRALBP binding cavity, which remains an unresolved but important issue for understanding how ll-cis-retinal is exported from the RPE for visual pigment regeneration. Besides that, the function of CRALBP in tissues other than the RPE remain to be determined (i.e. in retinal Muller cells, ciliary 42 epithelium, iris, cornea, pineal gland, and a subset of oligodendrocytes of the optic nerve and brain). To answers these questions, determination of the three-dimensional structure of CRALBP will be essential. In addition to elucidating the physiological role of CRALBP, another reason that we are interested in determining the crystal structure of CRALBP is its potential as a model for rhodopsin. As discussed earlier, the qualifications of a good rhodopsin mimic should include affinity for retinoids, ease in molecular biology manipulations, high expression levels, great solubility, and routine access to high-resolution crystal structures. CRALBP fits in almost all categories except for the crystal structures. Beyond these features, CRALBP is superior over CRABPII in mimicking rhodopsin in that retinal in CRALBP is completely buried as in rhodopsin, while in CRABPII the ionone ring of the chromophore is partially solvent-exposed. With these two reasons, we are driven to pursue the crystallization of CRALBP. Chapter V details about our effort toward obtaining the first crystal structure of CRALBP. 43 Reference 10. 11. 12. 13. 14. 15. Wade, N. J. (1998) A Natural History of Vision, Bradford Books. Woodbridge, F. J. E. (1983) Aristotle's Vision of Nature, Greenwood Press Reprint. Kepler, J ., and Donahue, W. H. (2000) Optics, Green Lion Press. Newton, 1., Cohen, B. J., Einstein, A., and Whittaker, E. (1952) Opticks: Or a Treatise of the Reflections, Refractions, Inflections & Colours of Light-Based on the Fourth Edition London, 1 730, Dover Publications. Crescitelli, F. (1977) Friedrich Wilhelm Kuhne. The Centennial of Rhodopsin, Arch. Opthalmol. 95, 1766. Wald, G., and Brown, P. K. (1958) Human Rhodopsin, Science 127, 222-226. Wald, G. (1968) Molecular Basis of Visual Excitation, Nature 219, 800-&. Wald, G. (1968) Molecular Basis of Visual Excitation, Science 162, 230-&. Dunne, L. J. (1974) Excitonic Coupling in Visual Pigments, Phys Lett A A 48, 13- 13. Tiffany, J. M., Noorjahan, P., Abdel-Latif, A. A., Harding, J. J., Mayne, R., Brewton, R. G., and Ren, Z. X. (1997) Biochemistry of the Eye, Chapman and Hall, London, UK. Shirane, K. (1975) Conversion of Photoexcitation Energy in Rhodopsin, Nature 254, 722-723. Schenkl, S., van Mourik, F., van der Zwan, G., Haacke, S., and Chergui, M. (2005) Probing the ultrafast charge translocation of photoexcited retinal in bacteriorhodopsin, Science 309, 917-920. Crescitelli, F. (1972) The visual cells and visual pigments of the vertebrate eye., in Handbook of Sensory Physiology (Dartnell, H. J. A., Ed.), pp 245-363, Springer-Verlag, New York , NY. Wissinger, B., and Sharpe, L. T. (1998) New aspects of an old theme: The genetic basis of human color vision, Am J Hum Genet 63, 1257-1262. Kefalov, V., Fu, Y. B., Marsh-Armstrong, N., and Yau, K. W. (2003) Role of visual pigment properties in rod and cone phototransduction, Nature 425, 526- 531. 44 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. Sakurai, K., Onishi, A., Imai, H., Chisaka, O., Ueda, Y., Usukura, J., Nakatani, K., and Shichida, Y. (2007) Physiological properties of rod photoreceptor cells in green-sensitive cone pigment knock-in mice, J Gen Physiol 130, 21-40. Shi, G., Yau, K. W., Chen, J ., and Kefalov, V. J. (2007) Signaling properties of a short-wave cone visual pigment and its role in Phototransduction, J Neurosci 27, 10084-10093. Pugh, E. N. (2001) Rods are rods and cones cones, and (never) the twain shall meet, Neuron 32, 375-376. Hoffmann, R. (1995) The Same and Not the Same, Columbia University Press, New York, NY. Bennett, N., Michelvillaz, M., and Kuhn, H. (1982) Light-Induced Interaction between Rhodopsin and the th-Binding Protein Metarhodopsin-Ii Is the Major Photoproduct Involved, Eur J Biochem 127, 97-103. Borhan, B., Souto, M. L., Imai, H., Shichida, Y., and Nakanishi, K. (2000) Movement of retinal along the visual transduction path, Science 288, 2209-2212. Emeis, D., Kuhn, H., Reichert, J., and Hofrnann, K. P. (1982) Complex- Formation between Metarhodopsin-Ii and th-Binding Protein in Bovine Photoreceptor-Membranes Leads to a Shift of the Photoproduct Equilibrium, F ebs Lett 143, 29-34. Kliger, D. S., and Lewis, J. W. (1995) Spectral and kinetic characterization of visual pigment photointermediates, Israel J Chem 35, 289-307. Kliger, D. S., Lewis, J. W., and Szundi, I. (1999) Time-resolved spectroscopy of visual pigment intermediates, Abstr Pap Am Chem S 21 7, U323-U323. Souto, M. L., Borhan, B., and Nakanishi, K. (2000) Low-temperature photoaffinity labeling of rhodopsin and intermediates along transduction path, Vertebrate Phototransduction and the Visual Cycle, Pt B 316, 425-435. Collins, F. D. (1953) Rhodopsin and indicator yellow, Nature 1 71, 469-472. Morton, R. A., and Pitt, G. A. J. (1955) Rhodopsin. IX. pH and the hydrolysis of indicator yellow., Biochem. J. 59, 128-134. Palczewski, K., Kumasaka, T., Hori, T., Behnke, C. A., Motoshima, H., Fox, B. A., Le Trong, I., Teller, D. C., Okada, T., Stenkamp, R. E., Yamamoto, M., and Miyano, M. (2000) Crystal structure of rhodopsin: A G protein-coupled receptor, Science 289, 739-745. Rando, R. R. (1996) Polyenes and vision, Chem Biol 3, 255-262. 45 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. Shichida, Y., Matuoka, S., and Yoshizawa, T. (1984) Formation of Photorhodopsin, a Precursor of Bathorhodopsin, Detected by Picosecond Laser Photolysis at Room-Temperature, Photobioch Photobiop 7, 221-228. Shieh, T., Han, M., Sakmar, T. P., and Smith, S. O. (1997) The steric trigger in rhodopsin activation, J Mol Biol 269, 373-384. Yan, B., Nakanishi, K., and Spudich, J. L. (1991) Mechanism of Activation of Sensory Rhodopsin-I - Evidence for a Steric Trigger, P Natl Acad Sci USA 88, 9412-9416. Detwiler, P. (2002) Open the loop: Dissecting feedback regulation of a second messenger transduction cascade, Neuron 36, 3-4. Graykeller, M. P., and Detwiler, P. B. (1994) The Calcium Feedback Signal in the Phototransduction Cascade of Vertebrate Rods, Neuron 13, 849-861. Schnetkamp, P. P. M. (1995) How Does the Retinal Rod Na-Ca+K Exchanger Regulate Cytosolic-Free Ca2+, J Biol Chem 270, 13231-13239. Abdulaev, N. G., and Ridge, K. D. (1998) Light-induced exposure of the cytoplasmic end of transmembrane helix seven in rhodopsin, P Natl Acad Sci USA 95, 12854-12859. Farrens, D. L., Altenbach, C., Yang, K., Hubbell, W. L., and Khorana, H. G. (1996) Requirement of rigid-body motion of transmembrane helices for light activation of rhodopsin, Science 274, 768-770. Konig, B., Welte, W., and Hofinann, K. P. (1989) Photoactivation of Rhodopsin and Interaction with Transducin in Detergent Micelles - Effect of Doping with Steroid Molecules, F ebs Lett 25 7, 163-166. SHI, W., Osawa, S., Dickerson, C. D., and Weiss, E. R. (1995) Rhodopsin Mutants Discriminate Sites Important for the Activation of Rhodopsin Kinase and G(T), J Biol Chem 270, 2112-2119. Ernst, O. P., Hofrnann, K. P., and Sakmar, T. P. (1995) Characterization of Rhodopsin Mutants That Bind Transducin but Fail to Induce th Nucleotide Uptake - Classification of Mutant Pigments by Fluorescence, Nucleotide Release, and F lash-Induced Light-Scattering Assays, J Biol Chem 270, 10580-10586. Franke, R. R., Sakmar, T. P., Graham, R. M., and Khorana, H. G. (1992) Structure and Function in Rhodopsin - Studies of the Interaction between the Rhodopsin Cytoplasmic Domain and Transducin, J Biol Chem 267, 14767-14774. Onrust, R., Herzrnark, P., Chi, P., Garcia, P. D., Lichtarge, O., Kingsley, C., and Boume, H. R. (1997) Receptor and beta gamma binding sites in the alpha subunit of the retinal G protein transducin, Science 275, 381-3 84. 46 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. Granovsky, A. E., Natochin, M., and Artemyev, N. O. (1997) The gamma subunit of rod cGMP-phosphodiesterase blocks the enzyme catalytic site, J Biol Chem 272, 11686-11689. Stryer, L. (1996) Vision: From photon to perception, P Natl Acad Sci USA 93, 557-559. Leskov, I. B., Klenchin, V. A., Handy, J. W., Whitlock, G. G., Govardovskii, V. I., Bownds, M. D., Lamb, T. D., Pugh, E. N., and Arshavsky, V. Y. (2000) The gain of rod phototransduction: Reconciliation of biochemical and electrophysiological measurements, Neuron 27, 525-537. Birge, R. R. (1990) Nature of the Primary Photochemical Events in RhodOpsin and Bacteriorhodopsin, Biochim Biophys Acta 1016, 293-327. Pepe, I. M. (2001) Recent advances in our understanding of rhodopsin and phototransduction, Progress in Retinal and Eye Research 20, 733-759. Rieke, F ., and Baylor, D. A. (1998) Origin of reproducibility in the responses of retinal rods to single photons, Biophys J 75, 1836-1857. Lamb, T. D. (1996) Gain and kinetics of activation in the G-protein cascade of phototransduction, P Natl Acad Sci USA 93, 566-570. Miki, N., Keims, J. J., Marcus, F. R., Freeman, J., and Bitensky, M. W. (1973) Regulation of Cyclic Nucleotide Concentrations in Photoreceptors - Atp- Dependent Stimulation of Cyclic Nucleotide Phosphodiesterase by Light, P Natl Acad Sci USA 70, 3820-3824. Graykeller, M. P., and Detwiler, P. B. (1994) Intracellular Calcium Measurements in Isolated Rod Photoreceptors, Invest Ophth Vis Sci 35, 1486-1486. Palczewski, K., Buczylko, J., Vanhooser, P., Carr, S. A., Huddleston, M. J., and Crabb, J. W. (1992) Identification of the Autophosphorylation Sites in Rhodopsin Kinase, J Biol Chem 26 7, 18991-18998. Palczewski, K., Rispoli, G., and Detwiler, P. B. (1992) The Influence of Arrestin (48k Protein) and Rhodopsin Kinase on Visual Transduction, Neuron 8, 117-126. Lamb, T. D., and Matthews, H. R. (1988) External and Internal Actions in the Response of Salamander Retinal Rods to Altered External Calcium- Concentration, Journal of Physiology-London 403, 473-494. Sagoo, M. S., and Lagnado, L. (1997) G-protein deactivation is rate-limiting for shut-off of the phototransduction cascade, Nature 389, 392-395. Bacigalupo, J., and Lisman, J. E. (1983) Single-Channel Currents Activated by Light in Limulus Ventral Photoreceptors, Nature 304, 268-270. 47 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. Lott, J. S., Wilde, J. I., Came, A., Evans, N., and Findlay, J. B. C. (1999) The ordered visual transduction complex of the squid photoreceptor membrane, Molecular Neurobiology 20, 61-80. Oday, P. M., Bacigalupo, J ., Vergara, C., and Haab, J. E. (1997) Current issues in invertebrate phototransduction - Second messengers and ion conductances, Molecular Neurobiology 15, 41-63. Hargrave, P. A. (1977) Amino-Terminal Tryptic Peptide of Bovine Rhodopsin - Glycopeptide Containing 2 Sites of Oligosaccharide Attachment, Biochim Biophys A cta 492, 83-94. Hargrave, P. A., Mcdowell, J. H., Curtis, D. R., Wang, J. K., Juszczak, E., Fong, S. L., Rao, J. K. M., and Argos, P. (1983) The Structure of Bovine Rhodopsin, Biophysics of Structure and Mechanism 9, 23 5-244. Okada, T., Fujiyoshi, Y., Silow, M., Navarro, J., Landau, E. M., and Shichida, Y. (2002) Functional role of internal water molecules in rhodopsin revealed by x-ray crystallography, P Natl Acad Sci USA 99, 5982-5987. Okada, T., Sugihara, M., Bondar, A. N., Elstner, M., Entel, P., and Buss, V. (2004) The retinal conformation and its environment in rhodopsin in light of a new 2.2 angstrom crystal structure, J Mol Biol 342, 571-583. Nathans, J. (1990) Determinants of Visual Pigment Absorbency - Identification of the Retinylidene Schiff-Base Counterion in Bovine Rhodopsin, Biochemistry-Us 29, 9746-9752. Sakmar, T. P., Franke, R. R., and Khorana, H. G. (1989) Glutamic Acid-113 Serves as the Retinylidene Schiff-Base Counterion in Bovine Rhodopsin, P Natl Acad Sci USA 86, 8309-8313. Zhukovsky, E. A., and Oprian, D. D. (1989) Effect of Carboxylic-Acid Side- Chains on the Absorption Maximum of Visual Pigments, Science 246, 928-930. RAO, V. R., Cohen, G. B., and Oprian, D. D. (1994) Rhodopsin Mutation G90d and a Molecular Mechanism for Congenital Night Blindness, Nature 367, 639- 642. Schertler, G. F. X. (1998) Structure of rhodopsin, EYE 12, 504-510. Palczewski, K. (2006) G protein-coupled receptor rhodopsin, Annu Rev Biochem 75, 743-767. Unger, V. M., Hargrave, P. A., Baldwin, J. M., and Schertler, G. F. X. (1997) Arrangement of rhodopsin transmembrane alpha-helices, Nature 389, 203-206. 48 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. Unger, V. M., and Schertler, G. F. X. (1995) Low-Resolution Structure of Bovine Rhodopsin Determined by Electron Cryomicroscopy, Biophys J 68, 1776—1786. Nakamichi, H., and Okada, T. (2006) Local peptide movement in the photoreaction intermediate of rhodopsin, P Natl Acad Sci USA 103, 12729-12734. Nakamichi, H., and Okada, T. (2006) Crystallographic analysis of primary visual photochemistry, Angewandte Chemie-International Edition 45, 4270-4273. Ruprecht, J. J., Mielke, T., Vogel, R., Villa, C., and Schertler, G. F. X. (2004) Electron crystallography reveals the structure of metarhodopsin I, Embo J 23, 3609-3620. Ridge, K. D., and Palczewski, K. (2007) Visual rhodopsin sees the light: Structure and mechanism of G protein signaling, J Biol Chem 282, 9297-9301. Salom, D., Lodowski, D. T., Stenkamp, R. E., Le Trong, I., Golczak, M., Jastrzebska, B., Harris, T., Ballesteros, J. A., and Palczewski, K. (2006) Crystal structure of a photoactivated deprotonated intermediate of rhodopsin, P Natl Acad Sci USA 103,16123-16128. Spudich, J. L., Yang, C. S., Jung, K. H., and Spudich, E. N. (2000) Retinylidene proteins: Structures and functions from archaea to humans, Annu Rev Cell Dev Bi 16, 365-+. Sharma, A. K., Spudich, J. L., and Doolittle, W. F. (2006) Microbial rhodopsins: functional versatility and genetic mobility, Trends in Microbiology 14, 463-469. Spudich, J. L. (1998) Variations on a molecular switch: transport and sensory signalling by archaea] rhodopsins, Mol Microbiol 28, 1051-1058. Oesterhe.D, and Stoecken.W. (1973) Functions of a New Photoreceptor Membrane, P Natl Acad Sci USA 70, 2853-2857. Stoeckenius, W., and Bogomolni, R. A. (1982) Bacteriorhodopsin and Related Pigments of Halobacteria, Annu Rev Biochem 51, 587-616. Luecke, H., Schobert, B., Richter, H. T., Cartailler, J. P., and Lanyi, J. K. (1999) Structure of bacteriorhodopsin at 1.55 angstrom resolution, J Mol Biol 291, 899- 911. Doukas, A. G., Pande, A., Suzuki, T., Callender, R. H., Honig, B., and Ottolenghi, M. (1981) On the Mechanism of Hydrogen-Deuterium Exchange in Bacteriorhodopsin, Biophys J 33, 275-279. Ehrenberg, B., Lewis, A., Porta, T. K., Nagle, J. F., and Stoeckenius, W. (1980) Exchange Kinetics of the Schiff-Base Proton in Bacteriorhodopsin, P Natl Acad Sci-Biol 77, 6571-6573. 49 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. Degroot, H. J. M., Harbison, G. S., Herzfeld, J., and Griffin, R. G. (1989) Nuclear Magnetic-Resonance Study of the Schiff-Base in Bacteriorhodopsin - Counterion Effects on the N-15 Shift Anisotropy, Biochemistry-Us 28, 3346-3353. Frigaard, N. U., Martinez, A., Mincer, T. J., and DeLong, E. F. (2006) Proteorhodopsin lateral gene transfer between marine planktonic Bacteria and Archaea, Nature 439, 847-850. Balashov, S. P., Imasheva, E. S., Boichenko, V. A., Anton, J., Wang, J. M., and Lanyi, J. K. (2005) Xanthorhodopsin: A proton pump with a light-harvesting carotenoid antenna, Science 309, 2061-2064. Beja, 0., Aravind, L., Koonin, E. V., Suzuki, M. T., Hadd, A., Nguyen, L. P., Jovanovich, S., Gates, C. M., Feldman, R. A., Spudich, J. L., Spudich, E. N., and DeLong, E. F. (2000) Bacterial rhodopsin: Evidence for a new type of phototrophy in the sea, Science 289, 1902-1906. Matsunoyagi, A., and Mukohata, Y. (1977) 2 Possible Roles of Bacteriorhodopsin - Comparative-Study of Strains of Halobacterium-Halobium Differing in Pigmentation, Biochemical and Biophysical Research Communications 78, 237- 243. Mongodin, E. F., Nelson, K. E., Daugherty, S., DeBoy, R. T., Wister, J., Khouri, H., Weidman, J., Walsh, D. A., Papke, R. T., Perez, G. S., Sharma, A. K., Nesbo, C. L., MacLeod, D., Bapteste, E., Doolittle, W. F ., Charlebois, R. L., Legault, B., and Rodriguez-Valera, F. (2005) The genome of Salinibacter ruber: Convergence and gene exchange among hyperhalophilic bacteria and archaea, P Natl Acad Sci USA 102,18147-18152. Spudich, J. L. (2006) The multitalented microbial sensory rhodopsins, Trends in Microbiology 14, 480-487. Bogomolni, R. A., and Spudich, J. L. (1982) Identification of a 3rd-Rhodopsin- Like Pigment in Phototactic Halobacterium-Halobium, P Natl Acad Sci-Biol 79, 6250-6254. Spudich, J. A., Kuczrnarski, E. R., Pardee, J. D., Simpson, P. A., Yamarnoto, K., and Stryer, L. (1981) Control of Assembly of Dictyostelium Myosin and Actin- Filaments, Cold Spring Harbor Symposia on Quantitative Biology 46, 553-561. Ren, L., Martin, C. H., Wise, K. J., Gillespie, N. B., Luecke, H., Lanyi, J. K., Spudich, J. L., and Birge, R. R. (2001) Molecular mechanism of spectral tuning in sensory rhodopsin II, Biochemistry— Us 40, 13906-13914. Bergo, V., Spudich, E. N., Spudich, J. L., and Rothschild, K. J. (2003) Conformational changes detected in a sensory rhodopsin II-transducer complex, J Biol Chem 278, 36556-36562. 50 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. Furutani, Y., Kamada, K., Sudo, Y., Shimono, K., Kamo, N., and Kandori, H. (2005) Structural changes of the complex between pharaonis phoborhodopsin and its cognate transducer upon formation of the m photointermediate, Biochemistry- Us 44, 2909-2915. Hein, M., Radu, 1., Klare, J. P., Engelhard, M., and Siebert, F. (2004) Consequences of counterion mutation in sensory rhodopsin II of Natronobacterium pharaonis for photoreaction and receptor activation: An FTIR study, Biochemistry-Us 43, 995-1002. Sudo, Y., Furutani, Y. J., Shimono, K., Kamo, N., and Hidekori, K. (2003) Hydrogen bondin alteration of Thr-204 in the complex between pharaonis phoborhodopsin and its transducer protein, Biochemistry- Us 42, 14166-14172. Jung, K. H., Trivedi, V. D., and Spudich, J. L. (2003) Demonstration of a sensory rhodopsin in eubacteria, Mol Microbiol 47, 1513-1522. Vogeley, L., Sineshchekov, O. A., Trivedi, V. D., Sasaki, J., Spudich, J. L., and Luecke, H. (2004) Anabaena sensory rhodopsin: A photochromic color 0 sensor at 2.0 angstrom, Science 306, 1390-1393. Sudo, Y., and Spudich, J. L. (2006) Three strategically placed hydrogen-bonding residues convert a proton pump into a sensory receptor, P Natl Acad Sci USA 103, 16129-16134. Bowmaker, J. K., and Dartnall, H. J. A. (1980) Visual Pigments of Rods and Cones in a Human Retina, Journal of Physiology-London 298, 501-511. Asenjo, A. B., Rim, J., and Oprian, D. D. (1994) Molecular Determinants of Human Red/Green Color Discrimination, Neuron 12, 1131-1138. Merbs, S. L., and Nathans, J. (1992) Absorption-Spectra of Human Cone Pigments, Nature 356, 433-435. Oprian, D. D., Asenjo, A. B., Lee, N., and Pelletier, S. L. (1991) Design, Chemical Synthesis, and Expression of Genes for the 3 Human Color-Vision Pigments, Biochemistry- Us 30, 1 1367-1 1372. Stockman, A., Sharpe, L. T., Merbs, S., and Nathans, J. (2000) Spectral sensitivities of human cone visual pigments determined in vivo and in vitro, Vertebrate Phototransduction and the Visual Cycle, Pt B 316, 626-650. Dawson, T. L. (2006) Colour and colour vision of creatures great and small, Coloration Technology 122, 61-73. Bowmaker, J. K. (1991) The evolution of vertebrate visual pigments and photoreceptors, The Macmillan Press, Houndsmills, Basingstoke, Hampshire, UK. 51 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. Kelber, A., Vorobyev, M., and Osorio, D. (2003) Animal colour vision - behavioural tests and physiological concepts, Biological Reviews 78, 81-118. Wald, G. (1933) Vitamin A in the retina., Nature 132, 316-317. Morton, R. A., and Goodwin, T. W. (1944) Preparation of retinene in vitro., Nature 153, 405-406. Wald, G., and Brown, P. K. (1950) The Synthesis of Rhodopsin from Retinene, P Natl Acad Sci USA 36, 84-92. Akhtar, M., Blosse, D., and Dewhurst, P. (1968) Studies on vision. The nature of the retinal-opsin linkage, Biochem. J. 110, 693-702. Bownds, D. (1967) Site of attachment of retinal in rhodopsin, Nature 216, 1178- 1181. Oseroff, A. R., and Callende.Rh. (1974) Resonance Raman-Spectroscopy of Rhodopsin in Retinal Disk Membranes, Biochemistry- Us 13, 4243-4248. Blatz, P. E., Baumgartner, N., Balasubramaniyan, B., Balasubramaniyan, P., and Stedman, F. (1971) Wavelength regulation in visual pigment chromophore. Large induced bathochromic shifts in retinol and related polyenes., Photochem. Photobiol. 14, 531-549. Pitt, G. A. J., Collins, F. D., Morton, R. A., and Stok, P. (1955) Rhodopsin. VIII. N-Retinylidenemethylamine, an indicator yellow analog, Biochem. J. 59, 122- 128. Blatz, P. E., Mohler, J. H., and Ahmed, W. (1991) Spectroscopic Observation of Solvent Interaction with Selected Retinal Schiff-Bases, Photochem Photobiol 54, 255-264. Erickson, J. O., and Blatz, P. E. (1968) N-retinylidene-1-amino-2-propanol: a Schiff base analog for rhodopsin, Vision Res. 8, 1367-1375. Motto, M. G., Sheves, M., Tsujimoto, K., Baloghnair, V., and Nakanishi, K. (1980) Opsin Shifts in Bovine Rhodopsin and Bacteriorhodopsin - Comparison of 2 External Point-Charge Models, J Am Chem Soc 102, 7947-7949. Nakanishi, K., Baloghnair, V., Amaboldi, M., Tsujimoto, K., and Honig, B. (1980) An External Point-Charge Model for Bacteriorhodopsin to Account for Its Purple Color, J Am Chem Soc 102, 7945-7947. Hubbard, R. E. (1958) Paper 4, in Proc. Nat. Phys. Lab. Symp. No. 8, London. 52 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. 133. Blatz, P. E., and Mohler, J. H. (1975) Effect of Selected Anions and Solvents on Electronic Absorption, Nuclear Magnetic-Resonance, and Infrared-Spectra of N- Retinylidene-N-Butylammonium Cation, Biochemistry- Us 14, 2304-2309. Honig, B., Greenberg, A. D., Dinur, U., and Ebrey, T. G. (1976) Visual-Pigment Spectra - Implications of Protonation of Retinal Schiff-Base, Biochemistry- Us 15, 4593-4599. Kliger, D. S., Milder, S. J., and Dratz, E. A. (1977) Solvent Effects on Spectra of Retinal Schiff-Bases .1. Models for Bathochromic Shift of Chromophore Spectrum in Visual Pigments, Photochem Photobiol 25, 277-286. Suzuki, H., Komatsu, T., and Kitajirna, H. (1974) Theory of Optical Property of Visual Pigment, Journal of the Physical Society of Japan 3 7, 177-185. Waddell, W. H., Schaffer, A. M., and Becker, R. S. (1977) Visual Pigments .7. Experimental and Theoretical Investigations of Absorption Spectral Properties of Protonated Retinal Schiff-Bases and Implications for Bathochromic Shift in Visual Pigments, J Am Chem Soc 99, 8456-8460. Amaboldi, M., Motto, M. G., Tsujimoto, K., Baloghnair, V., and Nakanishi, K. (1979) Hydro-Retinals and Hydro-Rhodopsins, J Am Chem Soc 101, 7082-7084. Mollevanger, L. C. P. J., Kentgens, A. P. M., Pardoen, J. A., Courtin, J. M. L., Veeman, W. S., Lugtenburg, J ., and Degrip, W. J. (1987) High-Resolution Solid- State C-13-Nmr Study of Carbons C-5 and C-12 of the Chromophore of Bovine Rhodopsin - Evidence for a 6-S-Cis Conformation with Negative-Charge Perturbation near C-12, Eur J Biochem 163, 9-14. Nakanishi, K. (1985) Bioorganic Studies with Rhodopsin, Pure Appl Chem 57, 769-776. Nathans, J., Thomas, D., and Hogness, D. S. (1986) Molecular-Genetics of Hmnan Color-Vision - the Genes Encoding Blue, Green, and Red Pigments, Science 232, 193-202. Gat, Y., and Sheves, M. (1993) A Mechanism for Controlling the Pk(a) of the Retinal Protonated Schiff-Base in Retinal Proteins - a Study with Model Compounds, J Am Chem Soc 115, 3772-3773. Blatz, P. E., Mohler, J. H., and Navangul, H. V. (1972) Anion-Induced Wavelength Regulation of Absorption Maxima of Schiff-Bases of Retinal, Biochemistry-Us 11, 848-&. Livnah, N., and Sheves, M. (1993) The Schiff-Base Bond Configuration in Bacteriorhodopsin and in Model Compounds, Biochemistry- Us 32, 7223-7228. 53 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. Livnah, N., and Sheves, M. (1993) Model Compounds Can Mimic Spectroscopic Properties of Bovine Rhodopsin, J Am Chem Soc 115, 351-353. Blatz, P. E., and Liebman, P. A. (1973) Wavelength regulation in visual pigments, Exp. Eye Res. 17, 573-580. Tan, Q., Lou, J. H., Borhan, B., Karnaukhova, E., Berova, N., and Nakanishi, K. (1997) Absolute sense of twist of the C12-C13 bond of the retinal chromophore in bovine rhodopsin based on exciton-coupled CD spectra of 11,12-dihydroretinal analogues, Angew Chem Int Edit 36, 2089-2093. Verdegem, P. J. E., Bovee-Geurts, P. H. M., de Grip, W. J., Lugtenburg, J., and de Groot, H. J. M. (1999) Retinylidene ligand structure in bovine rhodopsin, metarhodopsin-I, and lO-methylrhodopsin from internuclear distance measurements using C-l3-labeling and 1-D rotational resonance MAS NMR, Biochemistry- Us 38, 1 1316-1 1324. Honig, B., Hudson, B., Sykes, B. D., and Karplus, M. (1971) Ring Orientation in Beta-Ionone and Retinals (Theoretical/Nmr/Semi-Empirical Approach/Overhauser Effect/S-Cis Conformation), P Natl Acad Sci USA 68, 1289-&. Rafferty, C. N. (1979) Light-Induced Perturbation of Aromatic Residues in Bovine Rhodopsin and Bacteriorhodopsin, Photochem Photobiol 29, 109-120. Stenkamp, R. E., Filipek, S., Driessen, C. A. G. G., Teller, D. C., and Palczewski, K. (2002) Crystal structure of rhodopsin: a template for cone visual pigments and other G protein-coupled receptors, Bba-Biomembranes 1 5 65 , 168-182. Yokoyama, R., and Yokoyama, S. (1990) Convergent Evolution of the Red-Like and Green-Like Visual Pigment Genes in Fish, Astyanax-Fasciatus, and Human, P Natl Acad Sci USA 87, 9315-9318. Chan, T., Lee, M., and Sakmar, T. P. (1992) Introduction of Hydroxyl-Bearing Amino-Acids Causes Bathochromic Spectral Shifts in Rhodopsin - Amino-Acid Substitutions Responsible for Red-Green Color Pigment Spectral Tuning, J Biol Chem 26 7, 9478-9480. Altun, A., Yokoyama, S., and Morokuma, K. (2008) Quantum mechanical/molecular mechanical studies on spectral tuning mechanisms of visual pigments and other photoactive proteins, Photochem Photobiol 84, 845-854. Fujimoto, K., Hasegawa, J. Y., Hayashi, S., and Nakatsuji, H. (2006) On the color-tuning mechanism of Human-Blue visual pigment: SAC-CI and QM/MM study, Chem Phys Lett 432, 252-256. 54 145. 146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. Fujimoto, K., Hasegawa, J. Y., and Nakatsuji, H. (2008) Origin of color tuning in human red, green, and blue cone pigments: SAC-CI and QM/MM study, Chem Phys Lett 462, 318-320. Birge, R. R., Murray, L. P., Pierce, B. M., Akita, H., Baloghnair, V., Findsen, L. A., and Nakanishi, K. (1985) 2-Photon Spectroscopy of Locked-1 l-Cis- Rhodopsin - Evidence for a Protonated Schiff-Base in a Neutral Protein-Binding Site, P Natl Acad Sci USA 82, 4117-4121. Mathies, R., and Stryer, L. (1976) Retinal Has a Highly Dipolar Vertically Excited Singlet-State - Implications for Vision, P Natl Acad Sci USA 73, 2169- 2173. Wanko, M., Hoffrnann, M., Frauenheim, T., and Elstner, M. (2006) Computational photochemistry of retinal proteins, J Comput Aid Mol Des 20, 511-518. Trabanino, R. J., Vaidehi, N., and Goddard, W. A. (2006) Exploring the molecular mechanism for color distinction in humans, J Phys Chem B 110, 17230-17239. Houjou, H., Inoue, Y., and Sakurai, M. (2001) Study of the opsin shift of bacteriorhodopsin: Insight from QM/MM calculations with electronic polarization effects of the protein environment, J Phys Chem B 105, 867-879. Warshel, A., and Chu, Z. T. (2001) Nature of the surface crossing process in bacteriorhodopsin: Computer simulations of the quantum dynamics of the primary photochemical event, J Phys Chem B 105, 9857-9871. Yan, E. C. Y., Kazrni, M. A., Ganim, Z., Hou, J. M., Pan, D. H., Chang, B. S. W., Sakmar, T. P., and Mathies, R. A. (2003) Retinal counterion switch in the photoactivation of the G protein-coupled receptor rhodopsin, P Natl Acad Sci USA 100, 9262-9267. Kholmurodov, K. T., Fel'dman, T. B., and Ostrovsky, M. A. (2007) Interaction of chromophore, ll-cis-retinal, with amino acid residues of the visual pigment rhodopsin in the region of protonated Schiff base: A molecular dynamics study, Russ Chem 3+ 56, 20-27. Martinez-Mayorga, K., Pitrnan, M. C., Grossfield, A., Feller, S. E., and Brown, M. F. (2006) Retinal counterion switch mechanism in vision evaluated by molecular simulations, J Am Chem Soc 128, 16502-16503. Giguere, V. (1994) Retinoic Acid Receptors and Cellular Retinoid-Binding Proteins - Complex Interplay in Retinoid Signaling, Endocrine Reviews 15, 61-79. Charnbon, P. (1996) A decade of molecular biology of retinoic acid receptors, Faseb J 10, 940-954. 55 157. 158. 159. 160. 161. 162. 163. 164. 165. 166. 167. 168. Lotan, R. (1980) Effects of Vitamin-a and Its Analogs (Retinoids) on Normal and Neoplastic-Cells, Biochimica Et Biophysica Acta 605, 33-91. DeLuca, L. M. (1991) Retinoids and their receptors in differeciation, embryogenesis, and neoplasia., Faseb J 5, 2924-2933. Chomienne, C., F enaux, P., and Degos, L. (1996) Retinoid differentiation therapy in promyelocytic leukemia, Faseb J 10, 1025-1030. Delva, L., Bastie, J. N., Rochette-Egly, C., Kraiba, R., Balitrand, N., Despouy, G., Chambon, P., and Chomienne, C. (1999) Physical and functional interactions between cellular retinoic acid binding protein [1 and the retinoic acid-dependent nuclear complex, Molecular and Cellular Biology 1 9, 7158-7167. Fisher, G. J ., and Voorhees, J. J. (1996) Molecular mechanisms of retinoid actions in skin, Faseb J 10, 1002-1013. Tallman, M. S. (1996) Differentiating therapy in acute myeloid leukemia, Leukemia 10, 1262-1268. Hong, W. K., Itri, L. M. (1994) Retinoids and human cancer, in The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. B., Roberts, A. B., Goodman, D. S., Ed.) 2nd ed., pp 597-630, Raven Press, New York. Vaezeslami, S., Mathes, E., Vasilelou, C., Borhan, B., and Geiger, J. H. (2006) The structure of apo-wild-type cellular retinoic acid binding protein [1 at 1.4 A and its relationship to ligand binding and nuclear translocation, J Mol Biol 363, 687-701. Kliewer, S. A., Umesono, K., Mangelsdorf, D. J ., and Evans, R. M. (1992) Retinoid X-Receptor Interacts with Nuclear Receptors in Retinoic Acid, Thyroid- Hormone and Vitamin-D3 Signaling, Nature 355, 446-449. Kreutz, M., Fritsche, J., Andreesen, R., and Krause, S. W. (1998) Regulation of cellular retinoic acid binding protein (CRABP 11) during human monocyte differentiation in vitro, Biochem Bioph Res Co 248, 830-834. Leid, M., Kastner, P., Lyons, R., Nakshatri, H., Saunders, M., Zacharewski, T., Chen, J. Y., Staub, A., Gamier, J. M., Mader, S., and Chambon, P. (1992) Purification, Cloning, and er Identity of the Hela—Cell Factor with Which Rar or Tr Heterodimerizes to Bind Target Sequences Efficiently, Cell 68, 377-395. Leroy, P., Krust, A., Kastner, P., Mendelsohn, C., Zelent, A., Chambon, P. (1992) Retinoic acid receptors, in Retinoids in Normal Development and T eratogenesis (Morris-Kay, G., ed., Ed.), pp 7-25, Oxford University Press, Oxford. 56 169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179. 180. Mangelsdorf, D. J., Thummel, C., Beato, M., Herrlich, P., Schutz, G., Umesono, K., Blumberg, B., Kastner, P., Mark, M., Chambon, P., and Evans, R. M. (1995) The Nuclear Receptor Superfamily - the 2nd Decade, Cell 83, 835-839. Morrisskay, G. (1993) Retinoic Acid and Craniofacial Development - Molecules and Morphogenesis, Bioessays 15, 9-15. Bailey, J. S., and SIU, C. H. (1988) Purification and Partial Characterization of a Novel Binding-Protein for Retinoic Acid from Neonatal Rat, J Biol Chem 263, 9326-9332. ONG, D. E., and Chytil, F. (1975) Retinoic Acid-Binding Protein in Rat Tissue - Partial-Purification and Comparison to Rat Tissue Retinol-Binding Protein, J Biol Chem 250, 6113-6117. Sani, B. P. H., D.L. (1974) Retinoic acid-binding-protein in chick-embryo metatarsal skin, Biochem. Biophys. Res. Commun. 61, 1276-1282. Willmert, L. J., and Noy, N. (2003) Involvement of CRABP-II in transcriptional activity of RAR, Faseb J 17, A315-A3l6. Ong, D. E., Newcomer, M. E., & Chytil, F. (1994) Cellular retinoid-binding proteins, in The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. B., Roberts, A. B., Goodman, D. S., Ed.) 2nd ed., pp 283-318, Raven Press, New York. Bertucci, F., Houlgatte, R., Benziane, A., et al. (2000) Gene expression profiling of primary breast carcinomas using arrays of candidate genes, Hum. Mol. Genet. 9, 2981-2991. Bertucci, F., Van Hulst, S., Bernard, K., Loriod, B., Granjeaud, S., Tagett, R., Starkey, M., Nguyen, C., Jordan, 8., and Birnbaum, D. (1999) Expression scanning of an array of growth control genes in human tumor cell lines, Oncogene 18, 3905-3912. Delva, L., Comic, M., Balitrand, N., Guidez, F., Miclea, J. M., Delmer, A., Teillet, F., Fenaux, P., Castaigne, S., Degos, L., and Chomienne, C. (1993) Resistance to All-Trans-Retinoic Acid (Atra) Therapy in Relapsing Acute Promyelocytic Leukemia - Study of in-Vitro Atra Sensitivity and Cellular Retinoic Acid-Binding Protein-Levels in Leukemic-Cells, Blood 82, 2175-2181. Hibbs, K., Skubitz, K. M., Pambuccian, S. E., Casey, R. C., Burleson, K. M., Oegema, T. R., Thiele, J. J., Grindle, S. M., Bliss, R. L., and Skubitz, A. P. N. (2004) Differential gene expression in ovarian carcinoma - Identification of potential biomarkers, American Journal of Pathology 165, 397-414. Li, C. M., Guo, M. R., Borczuk, A., Powell, C. A., Wei, M., Thaker, H. M., Friedman, R., Klein, U., and Tycko, B. (2002) Gene expression in Wilms' tumor 57 181. 182. 183. 184. 185. 186. 187. 188. 189. 190. mimics the earliest committed stage in the metanephric mesenchymal-epithelial transition, American Journal of Pathology [60, 2181-2190. Tsibris, J. C. M., Segars, J., Coppola, D., Mane, S., Wilbanks, G. D., O'Brien, W. F., and Spellacy, W. N. (2002) Insights from gene arrays on the development and growth regulation of uterine leiomyomata, Fertility and Sterility 78, 114-121. Vo, H. P., and Crowe, D. L. (1998) Transcriptional regulation of retinoic acid responsive genes by cellular retinoic acid binding protein-II modulates RA mediated tumor cell proliferation and invasion, Anticancer Research 18, 217-224. Zhou, D. C., Hallam, S. J., Lee, S. J., Klein, R. S., Wiemik, P. H., Tallman, M. S., and Robert, E. (1998) Constitutive expression of cellular retinoic acid binding protein 11 and lack of correlation with sensitivity to all-trans retinoic acid in acute promyelocytic leukemia cells, Cancer Research 58, 5770-5776. Budhu, A. S., and Noy, N. (2002) Direct channeling of retinoic acid between cellular retinoic acid-binding protein 11 and retinoic acid receptor sensitizes mammary carcinoma cells to retinoic acid—induced growth arrest, Molecular and Cellular Biology 22, 2632-2641. Dong, D., Ruuska, S. E., Levinthal, D. J., and Noy, N. (1999) Distinct roles for cellular retinoic acid-binding proteins I and II in regulating signaling by retinoic acid, J Biol Chem 274, 23695-23698. Sessler, R. J ., and Noy, N. (2005) A ligand-activated nuclear localization signal in cellular retinoic acid binding protein-II, Molecular Cell 18, 343-353. Vaezeslami, S., Mathes, E., Vasileiou, C., Borhan, B., and Geiger, J. H. (2006) The Structure of Apo-wild-type Cellular Retinoic Acid Binding Protein II at 1.4 A and its Relationship to Ligand Binding and Nuclear Translocation, J Mol Biol. Kleywegt, G. J ., Bergfors, T., Senn, H., Lemotte, P., Gsell, B., Shudo, K., and Jones, T. A. (1994) Crystal-Structures of Cellular Retinoic Acid-Binding Protein-I and Protein-Ii in Complex with All-Trans-Retinoic Acid and a Synthetic Retinoid, Structure 2, 1241-1258. Chen, X., Tordova, M., Gilliland, G. L., Wang, L. C., Li, Y., Yan, H. G., and Ji, X. H. (1998) Crystal structure of apo-cellular retinoic acid-binding protein type 11 (R1 11M) suggests a mechanism of ligand entry, J Mol Biol 278, 641-653. Thompson, J. R., Bratt, J. M., and Banaszak, L. J. (1995) Crystal-Structure of Cellular Retinoic Acid-Binding Protein-I Shows Increased Access to the Binding Cavity Due to Formation of an Intermolecular Beta-Sheet, J. Mol. Biol. 252, 433- 446. 58 191. 192. 193. 194. 195. 196. 197. 198. 199. 200. 201. Scapin, G., Gordon, J. I., and Sacchettini, J. C. (1992) Refinement of the Structure of Recombinant Rat Intestinal Fatty Acid-Binding Apoprotein at 1.2-a Resolution, .1. Biol. Chem. 26 7, 4253-4269. Wang, L. C., and Yan, H. G. (1998) NMR study suggests a major role for Arg] 11 in maintaining the structure and dynamical properties of type [1 human cellular retinoic acid binding protein, Biochemistry- Us 3 7, 13021-13032. Chaudhuri, B. N., Kleywegt, G. J., Broutin-L'Hermite, 1., Bergfors, T., Senn, H., Le Motte, P., Partouche, 0., and Jones, T. A. (1999) Structures of cellular retinoic acid binding proteins 1 and II in complex with synthetic retinoids, Acta Crystallogr D 55, 1850-1857. Kleywegt, G. J., Bergfors, T., Senn, H., Le Motte, P., Gsell, B., Shudo, K., and Jones, T. A. (1994) Crystal structures of cellular retinoic acid binding proteins I and II in complex with all-trans retinoic acid and a synthetic retinoid, Structure 2,1241-1258. Crist, R. M., Vasileiou, C., Rabago-Smith, M., Geiger, J. H., and Borhan, B. (2006) Engineering a rhodopsin protein mimic, J Am Chem Soc 128, 4522-4523. Vasileiou, C., Vaezeslami, S., Crist, R. M., Rabago-Smith, M., Geiger, J. H., and Borhan, B. (2007) Protein design: Reengineering Cellular Retinoic Acid Binding Protein 11 into a rhodopsin protein mimic, J Am Chem Soc 129, 6140-6148. Saari, J. C., Bredberg, L., and Garwin, G. G. (1982) Identification of the Endogenous Retinoids Associated with 3 Cellular Retinoid-Binding Proteins from Bovine Retina and Retinal-Pigment Epithelium, J Biol Chem 25 7, 3329-3333. Saari, J. C., and Bredberg, D. L. (1987) Photochemistry and Stereoselectivity of Cellular Retinaldehyde-Binding Protein from Bovine Retina, J Biol Chem 262, 7618-7622. Saari, J. C., Nawrot, M., Kennedy, B. N., Garwin, G. G., Hurley, J. H, Huang, J., Possin, D. E., and Crabb, J. W. (2001) Visual cycle impairment in cellular retinaldehyde binding protein (CRALBP) knockout mice results in delayed dark adaptation, Neuron 29, 739-748. Golovleva, 1., Bhattacharya, S., Wu, Z. P., Shaw, N., Yang, Y. W., Andrabi, H., West, K. A., Burstedt, M. S. I., Forsman, K., Holmgren, G., Sandgren, O., Noy, N., Qin, J., and Crabb, J. W. (2003) Disease-causing mutations in the cellular retinaldehyde binding protein tighten and abolish ligand interactions, J Biol Chem 278, 12397-12402. Stecher, H., Gelb, M. H., Saari, J. C., and Palczewski, K. (1999) Preferential release of ll-cis-retinol from retinal pigment epithelial cells in the presence of cellular retinaldehyde-binding protein, J Biol Chem 274, 8577-8585. 59 202. 203. 204. 205. 206. 207. 208. 209. 210. 211. 212. Winston, A., and Rando, R. R. (1998) Regulation of isomerohydrolase activity in the visual cycle, Biochemistry- Us 3 7, 2044-2050. Wu, Z. P., Yang, Y. W., Shaw, N., Bhattacharya, S., Yan, L., West, K., Roth, K., Noy, N., Qin, J., and Crabb, J. W. (2003) Mapping the ligand binding pocket in the cellular retinaldehyde binding protein, J Biol Chem 278, 12390-12396. Nawrot, M., West, K., Huang, J., Possin, D. E., Bretscher, A., Crabb, J. W., and Saari, J. C. (2004) Cellular retinaldehyde-binding protein interacts with ERM- binding phosphoprotein 50 in retinal pigment epithelium, Invest Ophth Vis Sci 45 , 393-401. Bhattacharya, S. K., Wu, Z., Jin, Z., Yan, L., Miyagi, M., West, K., Nawrot, M., Saari, J. C., and Crabb, J. W. (2002) Proteomic approach to identification of a mammalian visual cycle protein complex., Faseb J 16, Al4-A14. Crabb, J. W., Carlson, A., Chen, Y., Goldflarn, S., Intres, R., West, K. A., Hulmes, J. D., Kapron, J. T., Luck, L. A., Horwitz, J., and Bok, D. (1998) Structural and functional characterization of recombinant human cellular retinaldehyde-binding protein, Protein Sci 7, 746-757. Wu, Z. P., Hasan, A., Liu, T. Y., Teller, D. C., and Crabb, J. W. (2004) Identification of CRALBP ligand interactions by photoaffmity labeling, hydrogen/deuteritun exchange, and structural modeling, J Biol Chem 279, 27357- 27364. Panagabko, C., Morley, S., Hernandez, M., Cassolato, P., Gordon, H., Parsons, R., Manor, D., and Atkinson, J. (2003) Ligand specificity in the CRAL-TRIO protein family, Biochemistry- Us 42, 6467-6474. Meier, R., Tomizaki, T., Schulze-Briese, C., Baumann, U., and Stocker, A. (2003) The molecular basis of vitamin E retention: Structure of human alpha-tocopherol transfer protein, J Mol Biol 331, 725-734. Min, K. C., Kovall, R. A., and Hendrickson, W. A. (2003) Crystal structure, of human alpha-tocopherol transfer protein bound to its ligand: Implications for ataxia with vitamin E deficiency, P Natl Acad Sci USA 100, 14713-14718. Sha, B. D., Phillips, S. E., Bankaitis, V. A., and Luo, M. (1998) Crystal structure of the Saccharomyces cerevisiae phosphatidylinositol-transfer protein, Nature 391, 506-510. Stocker, A., and Baumann, U. (2003) Supernatant protein factor in complex with RRR-alpha-tocopherquuinone: A link between oxidized vitamin E and cholesterol biosynthesis, J Mol Biol 332, 759-765. 60 213. Liu, T. Y., Jenwitheesuk, E., Teller, D. C., and Samudrala, R. (2005) Structural insights into the cellular retinaldehyde-binding protein (CRALBP), Proteins- Structure Function and Bioinformatics 61 , 412-422. 61 CHAPTER II. Crystal Structures of Ligand-bound Cellular Retinoic Acid Binding Protein ll Mutants 2. 1. Introduction. 2.1.1 . Need for a ligand that can be totally buried inside the CRABPII mutants. The triple mutant KLE-CRABPII was a successful design in that it proved that the protonated Schiff base and the stabilizing counter anion are two essential factors in inducing the red shift of the bound retinal (1 ). However, there are still limitations of the KLE-CRABPII-all-trans-retinal system in terms of its application towards investigation of the point charge theory. According to the point charge theory, the placement of counter anions along the polyene chain would potentially stabilize the positive charge of the bound ligand at different positions and therefore lead to stabilization of different resonance structures, which would result in differed red shift abilities. To test this theory, the ideal situation is that the ligand is fully buried inside the protein and able to experience various protein-ligand interactions generated by binding site mutations. However, as shown in Figure II-l, the all-trans-retinal-bound CRABPII mutants all feature a binding mode with the ionone ring of the retinoid partially solvent-exposed. To circumvent this less-than-ideal situation associated with the Rt-bound CRABPII model systems, we decided to explore the possibility of introducing a shorter ligand into the CRABPII mutants. By docking simulations, we have identified a shorter retinal analog, 62 C15-aldehyde (Figure II-2), which is predicted to be totally buried inside the protein once bound to the CRABPII mutants. Figure [I-l. Retinoic acid bound in wild-type-CRABPII. OH H H OH I \ \ \O HO\%’N\/\’N]16) (35) and that of the bacteriorhodopsin (pKa 13.3) (31). The striking difference might be contributed to whether negatively charged residues are placed in the vicinity of the PSB. In all of the rhodopsin structures solved so far, one or two negatively charged residues are always present in the vicinity of the PSB and interact with the PSB nitrogen either directly (Glu113 in bovine rhodopsin and Asp238 in halorhodopsin) or through water-mediated hydrogen bonding (Asp85, Asp212 in bacteriorhodopsin and Asp75 in sensory rhodopsin). Although having a negatively charged residue next to the PSB does not guarantee a greatly enhanced pKa as demonstrated in the all-trans-retinal- bound KLE-CRABPII mutant (pKa 8.7), it is reasonable to rationalize that the lack of negatively charged residues next to the PSB could possibly make the protonation of the Schiff base harder and therefore results in a lower pKa of the PSB. Among the four C 15- aldehyde-bound CRABPII mutants, KL-A32E, KL-T54E and KFLDV all lack negatively charged residues in the vicinity of the PSB and are therefore not surprising to have much lower pKa values compared to the rhodopsins. In the C's-aldehyde-bound KLE-R59W-CRABPII mutant, the counterion, Glu121, is close to the PSB (4.8 A) and is making indirect hydrogen bonding with the PSB. However, similar to the all-trans-retinal-bound KLE-CRABPII mutant, the C15- aldehyde-bound KLE-R59W-CRABPII also shows relatively lower pKa (7.7). It has been proven that distances between the PSB nitrogen and the counter ion as well as the geometry of the hydrogen bonding also play important roles in determining the pKa of 76 the PSB (36-38). Having a negatively charged residue next to the PSB is therefore not a guarantee for increased pKa. While the relatively short distance between the counterion and the PSB (2.6 A) in the Rt-bound KLE-CRABPII is probably responsible for its relatively lower pKa (8.7), the apparent pKa of the C15-aldehyde-bound KLE-R59W- CRABPII (7.7) is harder to interpret. The pKa of the Cl5-aldehyde-bound KLE-R59W- CRABPII was measured in the absence of BTP. Attempts to measure the pKa in the presence of 100 mM BTP were not successful due to precipitation of the protein probably caused by the high concentration of BTP. 2.2.4. Cis versus trans protonated Schiff base. As revealed in the crystal structures of different rhodopsins, the PSB always adopts a trans conformation (20, 21, 24-26). In our KLE-R59W-C15-aldehyde complex, a trans double bond is also observed for the PSB. However, in all three of the CRABPII mutants where the C15-aldehyde adopts the alternative binding mode, a cis double bond was observed. A cis imine double bond has also been observed in our KLE-Rt complex previously reported (1 ). Given the fact that trans double bonds are thermodynamically more favored in solution, the formation of the cis PSB is apparently due to the positioning of the ligand by specific residues inside the binding cavities of these particular proteins. 77 Figure “-7. A. Proposed nucleophilic attack of lysine amino group to the aldehyde following the Biirgi-Dunitz model; B. Surface representation of C1 5-aldehyde (yellow) bound in KL-A32E (gray). In an overly simplified model, let us first assume that the nucleophilic attack happens following the Biirgi-Dunitz trajectory (39, 40), where the lysine nitrogen adopts a favorable 107° attacking angle with the terminal C=O bond of the chromophore and the N-C-O plane bisects the H-C(O)-C angle (Figure II-7A). We could also assume that the C15—aldehyde is not moving much from its original binding position due to the tight interactions of the residues around it (Figure lI-7B), which is supported by the three CRABPII mutant crystal structures with C15-aldehyde bound in the alternative binding pocket (Figure II-6B). Under these two assumptions we would see that during the reaction the lysine side chain would have to either swing in a way to facilitate the formation of a cis double bond or swing in another way facilitating formation of a trans PSB. Given that the energy difference between a cis and a trans protonated Schiff base is 7-12 kJ mol-I (15-17 kJ mol_1 for the unprotonated imine) (41), if the energy gained by the lysine swinging into the “cis” conformation versus “trans” is greater than 12 kJ mol- 78 I, the PSB formed would be cis. Otherwise, a trans double might be preferred. In the case where the chromophore is less tightly packed, the increased freedom of the ligand would probably favor the more thermodynamically stable trans double bond, as seen in the C15-aldehyde-bound KLE-R59W mutant structure. 2.2.5. Flexibility of the helix-turn-helix lid. When overlaying different CRABPII mutant structures, one can clearly see that the B-barrel portion of the protein is very rigid, whereas the helix-tum-helix shows flexibility. The flexibility of the helix-tum-helix portion of CRABPII has been previously observed (42). Based on the comparison of the crystal structures we have obtained for CRABPII mutants, the largest movement is observed between the C15- aldehyde-bound KL-A32E mutant and the C15-aldehyde-bound KLE-R59W mutant, with the longest distance being over 10 A between the same Ala35 C01 in the two different structures (Figure [1-8). As shown in holo-KLE-R59W, where the largest outward movement of 012 is observed, the C-terrninal of this helix (blue color) loses one helical turn and becomes a loop-like structure. Similar unwinding has been observed in several apo mutants as well and therefore is not induced by ligand binding. Unfortunately, up to this point, the apo structure of KLE-R59W has not been obtained probably due to the decreased structural stability of this protein when there is no ligand bound. This partial unwinding in the C 15- aldehyde-bound KLE-R59W results in the biggest outward movement among all our 79 CRABPII mutant crystal structures. Overlay of the C15-aldehyde-bound KLE-R59W and the apo structure of KLE-R59E that shares the same partial unwinding shows that the ligand would bump into the unwound portion of apo-KLE-R59E (Figure II-9A). In holo- KLE-R59W the bound C15-aldehyde indeed pushed out the unwound C-terminal portion of the 012 helix even more (Figure II-9B). In other words, the flexible helix lid allows the ligand to bind in this particular trajectory. Figure II-8. Stereoview of the overlay between holo-KL-A32E (green) and holo- KLE-R59W (magenta). Blue portion indicates the unwinding of the 012 helix. Distance (A) is measured between the C01 of Ala35 in the two structures. The same partial unwinding has been observed in M01 B of the C15-aldehyde- bound KFLDV mutant. However, the outward movement is more moderate compared to the holo-KLE-R59W. In Mol A of halo-KFLDV, the 0.2 helix is well maintained and adopts a “median” position in terms of movement by comparison with all the other CRABPII mutant structures solved so far. The fact that the two molecules in the same 80 asymmetric unit are displaying fairly large structural differences indicates that crystal packing could be playing a role in this partial unwinding of the helix. Both molecules in the asymmetric unit of the holo-KL-T54E mutant exhibit well-maintained helices. The positions of the helices in both molecules could also be described as “median” in the context of all the known CRABPII mutant structures. Figure [1-9. Overlay between C1 5-aldehyde-bound KLE-R59W (blue) and apo-KLE- R59E (green). A. Model of C15-aldehyde in apo-KLE-R59E; B. Crystal structure of C 1 5-aldehyde bound in KLE-R59W. 81 The most inward movement of the (12 helix is observed in the KL-A32E mutant, where the helical structure is well maintained (Figure 11-8). The inward movement of the helix is possibly achieved through hydrogen bonding interactions between the two backbone carbonyl oxygen atoms on the helix and the PSB nitrogen (Figure II-5B). Crystal packing effects from the tight arrangement of neighboring molecules in this crystal, which is not observed in the other three holo-CRABPII mutant structures, could also contribute to the movement in the structure of the C15-aldehyde-bound KL-A32E mutant. Nevertheless, the results from our crystal structures confirm and clearly demonstrate the great flexibility of the helix-turn-helix motif in CRABPII. 2.2.6. An evolutionary correlation between CRABPII and plasma retinal-binding protein? Plasma retinol-binding protein (RBP) is a 21kD protein with 184 amino acid residues and is the first structurally characterized lipocalin protein (16, 17, 43-47). RBP is synthesized in hepatocytes and transports vitamin A (retinol) from the liver, its storage site, to a variety of target tissues (44, 48-50). After being secreted out of the liver and into the extracellular fluid, RBP circulates the plasma by forming a complex with transthyretin (TTR), a thyroid-homone-transporting protein (51, 52). The complexation protects the relatively small RBP from being lost through glomerular filtration. RBP exhibits a typical lipocalin scaffold (43, 45-4 7, 53). A hydrophobic binding cavity is formed within an 8-strand B-barrel. Two helices are attached at each terminus of the polypeptide. Unlike CRABPs, the RBP helices are not positioned at the entrance of the 82 binding site, but instead swing outward into an almost parallel orientation with neighbouring B-barrel strands. Both CRABPs and RBP belong to the calycin superfamily, which includes fatty acid binding proteins, lipocalins, avidin, metalloproteinase inhibitors, and triabin (I 7). The calycin proteins all contain a basic B-barrel scaffold to form the binding cavity, which binds small, lipophilic ligands. An evolutionary correlation between cellular retinoid-binding proteins and RBP has been previously suggested, though the sequence homology between the two is very low (10% identity). Through site-directed mutagenesis, Schmidt and co-workers have created a point mutation, Q1171, deep inside the binding pocket of RBP (54). This Gln to Ile mutation established a better hydrophobic interaction with the ionone ring of retinol, which resulted in more than three-fold increase of the expression of the recombinant protein. The conversion into a hydrophobic residue from a hydrophilic one in a hydrophobic environment was believed to lead to further stability of the protein and therefore higher expression level. The same mutant also showed slightly enhanced binding affinity for retinol due to a higher association rate. The authors then raised the question of why nature would place a hydrophilic Gln residue at that critical position when it has better choices. They then compared RBP with cellular retinol-binding protein (CRBP). In CRBP, retinol binds in a completely opposite orientation and a Gln is also present inside the hydrophobic binding pocket, interacting with the hydroxy group of the ligand (55, 56). Interestingly, replacement of Gln with Arg in CRBP has been shown to diminish retinol affinity (57). They therefore speculated that “the Gln117 residue in RBP constitutes an evolutionary 83 remnant from ancient times when vitamin A might have been bound to the B-barrel in a different manner” (5 4). A. \7 '74:! v I a B. C-l a-Idehyde I 121 YIJ—l - . rL-rinol HUN 37 L l 157 \\ 2.1 Xvitl ‘\ 24 «4 Figure “-10. Stereoview of the overlay of C1 5-aldehyde-bound KL-A32E (green) and retinol-bound RBP (pink). A. Overall scaffold; B. Binding sites overlay. What we have observed in our C15-aldehyde-bound CRABPII system favors the same argument. Although both CRABPs and RBP are retinoid-binding proteins, they differ dramatically in the way that ligands sit inside the [ii-barrel pockets. Like its iLBP family member CRBP, WT-CRABPII positions its retinoid ligand, RA, with the carboxylate end going inside the cavity and the ionone ring portion solvent-exposed at the entrance. RBP, however, positions retinol in the opposite way with the ionone ring totally buried inside the cavity and the hydroxy group pointing outside, exposed to the 84 solvent. With the discovery of the alternative binding trajectory of C] 5-aldehyde in KL- A32E and the other two CRABPII mutants, we noticed the similarity between the way that C15-aldehyde is bound in KL-A32E and the retinol in RBP. By overlaying the two structures together through secondary structure alignment using the online program “secondary-structure matching” (SSM) (58), 70% of RBP and 58% of CRABPII are aligned, respectively, with an rrnsd of 2.62 A (Figure II-lOA). In the [ii-barrel region, the overall folds of the two proteins very much resemble one another. Helical portions of the two, however, do not overlay at all. When looking into the binding cavities of the two, we were happy to find that the ionone ring of C15-aldehyde and that of retinol occupy relatively similar positions (Figure 11-10). Certain residues around the ionone rings seem to be conserved (Figure II-lOB and Table II-2). Phe65 in CRABPII and Phe77 in RBP are close to each other in space and both are relatively close to the ionone rings of the ligands (6.2 A and 4.4 A, respectively). Tyr134 in CRABPII and Phe137 are also close (6.2 A and 4.6 A, respectively), approaching the ionone rings from another direction. Trp7 in CRABPII and Trp24 in RBP are finther away (beyond 8 A), but are very well conserved. Several other nearby residues, mostly hydrophobic, are also listed in Table 3, with each pair making similar interactions with the ionone rings. Based on this observation, we propose that an evolutionarily conserved, hydrophobic, yet under-used binding cavity is present in CRABPII. This observation favors the argument that an evolutionary correlation exists between the intracellular Lipid-Binding Proteins (iLBPs) and the lipocalin protein family. 85 Table [1-2. Residues around the ionone rings. Entry CRABPII RBP 1 F65 (6.21) F77 (4.44) 2 Y134 (6.25) F137 (4.62) 3 W7 (9.64) W24 (8.40) 4 T54 (3.74) A57 (3.45) 5 152 (3.66) A55 (3.99) 6 L121 (3.67) 9117 (4.60L 7 V41 (4.86) A43 (4.69) Numbers in parenthesis refer to the closest distances (A) between the specific residues to the corresponding ionone rings of the ligands. 2.2.7. Testing the point charge theory based on the C15-aldehyde- bound KLE-R59W-CRABPII structure. The high-resolution crystal structures of the C15-aldehyde-bound CRABPII mutants revealed detailed binding interactions between the chromophore and the nearby residues. Such information not only gave us confidence in interpreting the wavelength shifts results, but also helped us to identify options for firrther rational design. We then set out to design out experiments to test the wavelength shifts upon placing negative charged residues next to the polyene chain, referred to as the “point charge theory” (59). Table II-3 shows a series of KLE-X-CRABPII mutants. As shown in discussed earlier, the three mutations, KLE, not only ensure formation of the PSB between C15- aldehyde and Lysl32, but also render the ligand to adopt only one binding mode. In addition to the KLE mutations, these mutants were designed in a way that a negatively charged residue was introduced as a fourth mutation at places close to the polyene chain of the chromophore (Figure II-l 1). As shown in Table II-3, the greatest red shift was observed when a Glu is placed at position 59, which is right next to the ionone ring. As the negative charge moves gradually toward the PSB, the absorption maxima also drop 86 gradually to about 400 nm. This observation agrees well with the proposed theory in that negative charges placed further away from the PSB should stabilize the positive charge more remotely, leading to more extended conjugation and therefore greater red shift of the absorption maxima. Table [[-3. Absorption maxima observed for the KLE-X-CRABPII series of mutants Mutant Mu (mm) C] s-aldehyde (buffer) 340 C15-aldehyde PSB (EtOH) 380 R132K:R111L:L121E 330 R132K:R111L:L121E:R59E 424 R132K:R111L:L121EzA32E 416 R132K:R111L:L121EzA36E 410 R132K:R111L:L121EzF15D 397 R132K:R111L:L121E2V76D 399 * Results from Kin-Sing Lee. F15 6"" A32 N C15-aldOhYd9 W59 A36 \- K Figure II-ll. C1 5-aldehyde bound in the KLE-R59W-CRABPII mutant. 2.2.8. Conclusion The crystal structures of four different C15-aldehyde-bound CRABPII mutants were determined. Protonated Schiff base formation was observed between the active site 87 Lys at position 132 and the chromophore. Each of the four mutants contains a unique hydrogen bonding pattern that stabilizes the PSB. Two dramatically different binding orientations were observed in the different mutants with a single mutation at position 121 being key in determining which binding pocket the ligand occupies. Through steric interactions, Glu or Gln at position 121 forces the C15-aldehyde into the orthogonal binding mode where the ionone ring of the ligand extends towards the entrance of the cavity. A smaller residue (Leu or Asp) at position 121 allows the ligand to swing into an alternative, deeper, and more hydrophobic cavity. The constrained binding interactions in the alternative binding mode prefers the formation of a cis PSB, while the less tightly packed ligand environment in the orthogonal binding mode results in trans imine formation. Flexibility of the helix-turn-helix motif was also observed, featuring a dramatic 10 A movement of Ala35 C01 in two Cis-aldehyde-bound CRABPII mutants. By overlaying the alternative binding mode in the mutant CRABPII with the retinol- bound RBP through secondary structure alignment, we have observed structural resemblances between the two proteins both in the overall B-barrel fold and in the binding environments around the ionone rings of the ligands. Taken together with the previous literature, we proposed that a conserved, yet underused binding pocket is present in CRABPII and that an evolutionary correlation exists between the intracellular Lipid- Binding Proteins (iLBP) and the lipocalin protein family. A series of KLE-X-CRABPII mutants featuring negatively charged residues placed next to the polyene chain of the chromophore displayed interesting wavelength regulation properties. Negative charge placed further away from the PSB led to a greater red shift of the absorption maximum of the bound chromophore. This result agrees well 88 with the proposed point charge theory and presumably shed more light to the wavelength regulation mechanism in color vision. 2.3. Merocyanin-bound CRABPII mutant structure. H0.S.0 K‘ NW 0 O Merocyanin MES Figure II-12. Structures of merocyanin and MES. Table “-4. UV maximum absorption of the merocyanin-bound CRABPII mutants. Mutant Amax 1 Merocyanin in EtOH 461 2 Merocyanin in buffer 493 3 Merocyanin PSB in EtOH 578 4 Merocyanin PSB in buffer 573 5 R132K:R111L:L121E 588 6 R132K:R111L:L121E:R59E 596 7 R132K:R111L:L121E:R59W 596 8 R132K:R111L:L121EzT56W 591 9 R132K:R111L:L121E2T56R 591 10 R132K:R111L:L121EzV76E 593 11 R132K:R111L:L121E:V76W 570 12 R132K:R111L:L121EzV76R 586 * Results from Kin-Sing Lee. As an effort toward designing CRABPII mutants as fluorescence tags, Kin-Sing Lee and Calvin Grant in the Borhan lab have been working on synthesizing putative 89 chromophores. An interesting compound, merocyanin, was synthesized and showed excellent spectroscopic properties (Table II-4). By feeding this molecule to a series of CRABPII mutants, red shift of the Amax was observed in almost all mutants tested. The largest shifts were observed in mutants KLE-R59E and KLE-R59W, both affording an opsin shift of over 20 nm. However, the binding of merocyanin into CRABPII mutants appeared to be much slower compared to C15-aldehyde or all-trans-retinal. The crystallization of merocyanin-CRABPII mutant complexes was pursued. Structure of the KLE-R59W CRABPII mutant bound with merocyanin was determined at 2.60 A. The merocyanin is bound the orthogonal way with its aromatic ring extending toward the entrance and positioning itself right under the Trp59 (Figure 11-13). The Schiff base formed appears to be in a trans conformation and is stabilized by a weak hydrogen bonding interaction from a water molecule that is 3.4 A away (Figure II-13C). It is interesting to see the trans imine formation in this mutant especially when we compare this mutant to the C15-aldehyde-bound-KLE-R59W-CRABPII where the ligand binds also in the orthogonal way and also in trans imine form. The similarity of binding in the two complexes seems to suggest that shorter chromophores that bind the orthogonal way in CRABPII tends to favor a trans imine formation probably due to the ligands’ freedom in this spacious binding cavity. A buffer molecule, 2-(N- morpholino)ethanesulfonic acid (MES), is found to be bound inside the cavity as well (Figure II-13B), occupying a space approximately where BTP molecules are found in some other CRABPII mutant structures. The overall fold of this merocyanin-bound- 9O KLE-R59W-CRABPII resembles the C15-aldehyde-bound-KLE—R59W-CRABPII structure, with similar partial unwinding at the C-terminal end of the 012 helix. Figure [I-13. Merocyanin bound in KLE-R59W-CRABPII mutant. A. Stereoview of the overall fold of the protein; B. 2Fo-Fc map calculated at 1 o; C. Detailed binding interactions between merocyanin and the protein. This merocyanin molecule shows excellent fluorescence ability after being complexed with CRABPII mutants. In particular, the complex between merocyanin and KLE-R59W-CRABPII is capable to emit fluorescence at 614 nm after being excited at 91 570 nm (to be published by Kin-Sing Lee). This serves as the proof-of-concept for designing CRABPII mutants as ligand-inducible fluorescence fusion tags. The structural information obtained from this complex will also be instructional for directing our future ligand design and protein engineering efforts. 92 References 10. ll. 12. 13. Vasileiou, C., Vaezeslami, S., Crist, R. M., Rabago-Smith, M., Geiger, J. H., and Borhan, B. (2007) Protein design: Reengineering Cellular Retinoic Acid Binding Protein 11 into a rhodopsin protein mimic, J Am Chem Soc 129, 6140-6148. Gerisch, G., and Muller-Taubenberger, A. (2003) GFP-fusion proteins as fluorescent reporters to study organelle and cytoskeleton dynamics in chemotaxis and phagocytosis, Method Enzymol 361, 320-337. Hoffman, R. M. (2005) The multiple uses of fluorescent proteins to visualize cancer in vivo, Nat Rev Cancer 5, 796-806. Lippincott-Schwartz, J., Snapp, E., and Kenworthy, A. (2001) Studying protein dynamics in living cells, Nat Rev M01 Cell Bio 2, 444-456. Muller-Taubenberger, A., and Anderson, K. I. (2007) Recent advances using green and red fluorescent protein variants, Appl Microbiol Biot 77, 1-12. Shaner, N. C., Patterson, G. H., and Davidson, M. W. (2007) Advances in fluorescent protein technology, J Cell Sci 120, 4247-4260. Tsien, R. Y. (1998) The green fluorescent protein, Annu Rev Biochem 67, 509- 544. VanEngelenburg, S. B., and Palmer, A. E. (2008) Fluorescent biosensors of protein function, Curr Opin Chem Biol 12, 60-65. Heim, R., Prasher, D. C., and Tsien, R. Y. (1994) Wavelength Mutations and Posttranslational Autoxidation of Green Fluorescent Protein, P Natl Acad Sci USA 91, 12501-12504. Phillips, G. J. (2001) Green fluorescent protein - a bright idea for the study of bacterial protein localization, Fems Microbiol Lett 204, 9-18. Baumann, C. T., Lim, C. S., and Hager, G. L. (1998) Simultaneous visualization of the yellow and green forms of the green fluorescent protein in living cells, J Histochem Cytochem 46, 1073-1076. Chalfie, M., Tu, Y., Euskirchen, G., Ward, W. W., and Prasher, D. C. (1994) Green Fluorescent Protein as a Marker for Gene-Expression, Science 263, 802- 805. Cormack, B. P., Valdivia, R. H., and Falkow, S. (1996) FACS-optimized mutants of the green fluorescent protein (GFP), Gene 1 73, 33-38. 93 14. 15. 16. 17. l8. 19. 20. 21. 22. 23. 24. 25. 26. Hein, R., and Tsien, R. Y. (1996) Engineering green fluorescent protein for improved brightness, longer wavelengths and fluorescence resonance energy transfer, Curr Biol 6, 178-182. Akerstrom, B., Flower, D. R., and Salier, J. P. (2000) Lipocalins: unity in diversity, Bba-Protein Struct M 1482, 1-8. Flower, D. R. (2000) Experimentally determined lipocalin structures, Bba-Protein Struct M 1482, 46-56. Flower, D. R., North, A. C. T., and Sansom, C. E. (2000) The lipocalin protein family: structural and sequence overview, Bba-Protein Struct M 1482, 9-24. Skerra, A. (2000) Lipocalins as a scaffold, Bba-Protein Struct M 1482, 337-350. Lee, K. S., Jia, X., Vasileiou, C., Geiger, J. H., and Borhan, B., To be published. Okada, T., Sugihara, M., Bondar, A. N., Elstner, M., Entel, P., and Buss, V. (2004) The retinal conformation and its environment in rhodopsin in light of a new 2.2 angstrom crystal structure, Journal of Molecular Biology 342, 571-583. Palczewski, K., Kumasaka, T., Hori, T., Behnke, C. A., Motoshima, H., Fox, B. A., Le Trong, 1., Teller, D. C., Okada, T., Stenkamp, R. E., Yamamoto, M., and Miyano, M. (2000) Crystal structure of rhodopsin: A G protein-coupled receptor, Science 289, 739-745. Kholmurodov, K. T., Fel'dman, T. B., and Ostrovsky, M. A. (2007) Interaction of chromophore, ll-cis-retinal, with amino acid residues of the visual pigment rhodopsin in the region of protonated Schiff base: A molecular dynamics study, Russian Chemical Bulletin 56, 20-27. Sekharan, S., Sugihara, M., and Buss, V. (2007) Origin of spectral tuning in rhodopsin - It is not the binding pocket, Angewandte Chemie-International Edition 46, 269-271. Luecke, H., Schobert, B., Lanyi, J. K., Spudich, E. N., and Spudich, J. L. (2001) Crystal structure of sensory rhodopsin II at 2.4 angstroms: Insights into color tuning and transducer interaction, Science 293, 1499-1503. Facciotti, M. T., Rouhani, S., Burkard, F. T., Betancourt, F. M., Downing, K. H., Rose, R. B., McDermott, G., and Glaeser, R. M. (2001) Structure of an early intermediate in the M-state phase of the bacteriorhodopsin photocycle, Biophys J 81 , 3442-3455. Kolbe, M., Besir, H., Essen, L. O., and Oesterhelt, D. (2000) Structure of the light-driven chloride pump halorhodopsin at 1.8 angstrom resolution, Science 288, 1390-1396. 94 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. Brady, L., Brzozowski, A. M., Derewenda, Z. S., Dodson, E., Dodson, G., Tolley, S., Turkenburg, J. P., Christiansen, L., Hugejensen, B., Norskov, L., Thim, L., and Menge, U. (1990) A Serine Protease Triad Forms the Catalytic Center of a Triacylglycerol Lipase, Nature 343, 767-770. Jaeger, K. E., Dijkstra, B. W., and Reetz, M. T. (1999) Bacterial biocatalysts: Molecular biology, three-dimensional structures, and biotechnological applications of lipases, Annual Review of Microbiology 53, 315-+. Kraut, J. (1977) Serine Proteases - Structure and Mechanism of Catalysis, Annu Rev Biochem 46, 331-358. Menard, R., and Storer, A. C. (1992) Oxyanion Hole Interactions in Serine and Cysteine Proteases, Biological Chemistry Hoppe-Seyler 373, 393-400. Druckmann, S., Ottolenghi, M., Pande, A., Pande, J., and Callender, R. H. (1982) Acid-Base-Equilibriurn of the Schiff-Base in Bacteriorhodopsin, Biochemistry- Us 21 , 4953-4959. Koutalos, Y. (1992) High-Ph Form of Bovine Rhodopsin, Biophys J 61 , 272-275. Koutalos, Y., Ebrey, T. G., Gilson, H. R., and Honig, B. (1990) Octopus Photoreceptor-Membranes - Surface-Charge Density and Pk of the Schiff-Base of the Pigments, Biophys J 58, 493-501. Liang, J., Steinberg, G., Livnah, N., Sheves, M., Ebrey, T. G., and Tsuda, M. (1994) The Pk(a) of the Protonated Schiff-Bases of Gecko Cone and Octopus Visual Pigments, Biophys J 6 7, 848-854. Steinberg, G., Ottolenghi, M., and Sheves, M. (1993) Pk(a) of the Protonated Schiff-Base of Bovine Rhodopsin - a Study with Artificial Pigments, Biophys J 64, 1499-1502. GAT, Y., and Sheves, M. (1993) A Mechanism for Controlling the Pk(a) of the Retinal Protonated Schiff-Base in Retinal Proteins - a Study with Model Compounds, J Am Chem Soc 115, 3772-3773. Scheiner, S., and Hillenbrand, E. A. (1985) Modification of Pk Values Caused by Change in H-Bond Geometry, P Natl Acad Sci USA 82, 2741-2745. Taylor, R., Kennard, O., and Versichel, W. (1983) Geometry of the N-H=/=C Hydrogen-Bond .1. Lone-Pair Directionality, J Am Chem Soc 105, 5761-5766. Burgi, H. B., and Dunitz, J. D. (1983) From Crystal Statics to Chemical- Dynamics, Accounts of Chemical Research 16, 153-161. 95 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. Burgi, H. B., Dunitz, J. D., and Shefter, E. (1973) Geometrical Reaction Coordinates .2. Nucleophilic Addition to a Carbonyl Group, J Am Chem Soc 95, 5065-5067. Hammerum, S., and Solling, T. I. (1999) The proton affinities of imines and the heats of formation of immonium ions investigated with composite ab initio methods, J Am Chem Soc 121, 6002-6009. Wang, L. C., Li, Y., Abildgaard, F ., Markley, J. L., and Yan, H. G. (1998) NMR solution structure of type [1 human cellular retinoic acid binding protein: Implications for ligand binding, Biochemistry- Us 37, 12727-12736. Cowan, S. W., Newcomer, M. E., and Jones, T. A. (1990) Crystallographic Refinement of Human Serum Retinol Binding-Protein at 23 Resolution, Proteins 8, 44-61. Kanai, M., Raz, A., and Goodman, D. S. (1968) Retinol-Binding Protein - Transport Protein for Vitamin a in Human Plasma, Journal of Clinical Investigation 4 7, 2025—&. Newcomer, M. E., Jones, T. A., Aqvist, J., Sundelin, J., Eriksson, U., Rask, L., and Peterson, P. A. (1984) The 3-Dimensional Structure of Retinol-Binding Protein, Embo Journal 3, 1451-1454. Zanotti, G., Berni, R., and Monaco, H. L. (1993) Crystal-Structure of Liganded and Unliganded Forms of Bovine Plasma Retinol-Binding Protein, J Biol Chem 268, 10728-10738. Zanotti, G., Ottonello, S., Berni, R., and Monaco, H. L. (1993) Crystal-Structure of the Trigonal F orm of Human Plasma Retinol-Binding Protein at 2.5-Angstrom Resolution, Journal of Molecular Biology 230, 613-624. Blomhoff, R. (1994) Transport and Metabolism of Vitamin-A, Nutrition Reviews 52, 13-23. Blomhoff, R., Green, M. H., Berg, T., and Norum, K. R. (1990) Transport and Storage of Vitamin-A, Science 250, 399-404. Quadro, L., Blaner, W. S., Salchow, D. J., Vogel, S., Piantedosi, R., Gouras, P., Freeman, S., Cosma, M. P., Colantuoni, V., and Gottesman, M. E. (1999) Impaired retinal firnction and vitamin A availability in mice lacking retinol- binding protein, Embo Journal 18, 4633-4644. Goodman, D. S., and RAZ, A. (1972) Extraction and Recombination Studies of Interaction of Retinol with Human Plasma Retinol-Binding Protein, Journal of Lipid Research 13, 338-&. 96 52. 53. 54. 55. 56. 57. 58. 59. Monaco, H. L., Rizzi, M., and Coda, A. (1995) Structure of a Complex of 2 Plasma-Proteins - Transthyretin and Retinol-Binding Protein, Science 268, 1039- 1041. Grzyb, J., Latowski, D., and Strzalka, K. (2006) Lipocalins - a family portrait, Journal of Plant Physiology 163, 895-915. Schmidt, A. M., Bloss, 1., and Skerra, A. (1998) Improved folding of apo-retinol- binding protein in the periplasm of Escherichia coli: positive influences of dst coexpression and of an amino acid exchange in the vitamin A binding site, Protein Engineering 11, 601-607. Calderone, V., Folli, C., Marchesani, A., Berni, R., and Zanotti, G. (2002) Identification and structural analysis of a zebrafish apo and holo cellular retinol- binding protein, Journal of Molecular Biology 321 , 527-535. Cowan, S. W., Newcomer, M. E., and Jones, T. A. (1993) Crystallographic Studies on a Family of Cellular Lipophilic Transport Proteins - Refinement of P2 Myelin Protein and the Structure Determination and Refinement of Cellular Retinol-Binding Protein in Complex with All-Trans-Retinol, Journal of Molecular Biology 230, 1225-1246. Stump, D. G., Lloyd, R. S., and Chytil, F. (1991) Site-Directed Mutagenesis of Rat Cellular Retinol-Binding Protein - Alteration in Binding-Specificity Resulting from Mutation of Glutamine-108 to Arginine, J Biol Chem 266, 4622-4630. Krissinel, E., and Hendrick, K. Protein structure comparison service SSM at European Bioinformatic Institute (http://www.ebi.ac.uk/msd-srv/ssm). Nathans, J., Thomas, D., and Hogness, D. S. (1986) Molecular-Genetics of Human Color-Vision - the Genes Encoding Blue, Green, and Red Pigments, Science 232, 193-202. 97 CHAPTER III. Crystal Structures of the Apo-CRABPII Mutants 3. 1. Introduction Retinoic acid (RA) is the most important metabolite of vitamin A (retinol) in vertebrates. It has been proven to play indispensable roles during cell grth and differentiation by regulating the transcription of a variety of target genes (1-4). RA also possesses antitumor activity. RA and other retinoid analogs have been successfully employed in treatment again several types of cancers, including acute promyelocytic leukemia (2), skin cancer (5) and cervical cancer (6). The transcription effects of RA are exerted through binding of RA to the retinoic acid receptor (RAR). RAR is a nuclear transcriptional activator, which forms a heterodimer with retinoids X receptor (RXR) upon association with RA (7-10). The RAR/RXR dimer then recognizes the RA-response elements on specific target genes and regulates their transcription. The delivery of RA to RAR is facilitated by cellular retinoic acid binding proteins (CRABPS), members of the intracellular lipid-binding protein family (iLBP) (1 1-13). CRABPs are cytosolic, small (~16 kD), and soluble proteins that are found in all vertebrates (14). There are two types of CRABPs, CRABPI and CRABPII. CRABPI and II are highly conserved (both >90% identity) across species, respectively (15, 16). The identity between CRABPI and CRABPII in the same specie, however, is relatively lower (74% in human). Therefore, the two proteins probably diverged at a fairly early stage during evolution and have distinct cellular functions. CRABPs modulate the RA 98 effective concentration through multiple ways, including binding to excess RA and/or metabolizing the excess RA in the cell through interaction with metabolizing enzymes. CRABPI and II deliver RA to RAR through different mechanisms (17-19). While CRABPI shows no interaction with RAR, CRABPII has been shown to localize into the nucleus upon RA complexation and delivers RA to RAR through protein/protein interactions (16, 20). The crystal structures of the wild-type CRABPI and II have been solved earlier (21, 22). Both proteins adopt a typical iLBP fold with a rigid "ti-barrel" forming a huge binding cavity and a flexible helix-turn-helix motif functioning as a "lid" at the entrance (Figure III-1). Interestingly, water-mediated interactions are observed in the cavity of apo-WT-CRABPII MolA, connecting the Argl 11 deep inside the cavity through Arg132 to the Ala36 and Val33 on the 012 helix. The water-mediated network is not observed in MolB probably due to the crystal packing. The difference in crystal packing environments between MolA and MolB at the (12 helix region is dramatic, which is apparent from the number of the intermolecular interactions that each 012 helix makes with the neighboring molecules: MolA has 0 hydrogen bond and 8 vdw contacts and MolB has 7 hydrogen bonds and 43 vdw contacts. Similar water-mediated networks have been observed in other members of the iLBP family, such as CRABPI (23) and intestinal fatty acid-binding protein (24), and suggested to play important roles in maintaining the stability of these proteins. In the rat intestinal fatty acid-binding protein (FABP) structure (lIF C, 1.2 A), the water-mediated network connects the Arg106 (conserved residue of Argl 11 in CRABPII) through Arg126 (conserved residue of Arg132 in CRABPII) to Asp34 and Ala32 on the (12 helix 99 at the portal. In the CRABPI structure (lCBI), the network is observed, but not in a continuous manner. This is probably a result of certain water molecules not being observed due to the relatively low resolution of the structure (2.7 A). In the high resolution structure of cellular retinol-binding protein [1 (CRBPII, 1.2 A, PDB ID: 2RCQ), which is also an iLBP family member, the water-mediated network is again observed, which also resembles the one in apo-WT-CRABPII. In the apo-CRBPII structure, the network starts at the Glu108 (conserved residue of Arglll in CRABPII based on similarity) and extends to Ala36 and Val33 (both conserved in CRABPII based on identity). These observations further emphasize the structural importance of these water-mediated networks for the iLBP family proteins. In the course of our study, we have generated many different CRABPII mutants. The crystallization trials of these mutants have resulted in both ligand-bound and apo- structures of certain mutants. In this chapter, the apo-CRABPII mutant structures will be reported. Interesting structural information that is obtained from these structures will be discussed. 3.2. Results 3.2.1. Mutations of conserved and non-conserved residues Through site-directed mutagenesis, we have created a series of CRABPII mutants, each of which carry multiple mutations involving both conserved and non-conserved residues (25). All of the mutation sites that are involved in these mutants are shown in Figure 111-1. 100 Arg132 resides deep inside the binding pocket and is a fully conserved residue in the basis of identity in the intracellular binding protein family. In the WT-CRABPII, Arg132 directly interacts with the incoming carboxylate group of the retinoic acid (RA). The R132K mutation therefore greatly decreases the binding affinity of CRABPII toward Figure [II-1. Stereoview of the mutation sites involved in these CRABPII mutants. Tertiary structure of CRABPII is represented by retinoic acid-bound CRABPII (PDB code: 2FR3). Tyr134 and Arglll are also positioned deep in the pocket and are conserved residues based on similarity. Y134F is a very conserved mutation and ~20% of the iLBP family members actually have Phe at 134 as the native residue. R1 1 1L mutation, however, introduces a dramatic change from the hydrophilic Arg to a hydrophobic Leu, which enhances the hydrophobicity of the cavity and also creates an extra space due to the size differences of the two residues. Arg59, Ala32 and Thr54 are also conserved residues based on similarity. Among the three, Arg59 and Ala32 are fully conserved residues and Thr54 is conserved in 46 out 101 of 52 iLBP family members. These three residues are not deep inside the cavity and are instead at positions where the linear part of RA interacts with. R59E mutation, which happens at the entrance, switches the positively charged Arg to a negatively charged Glu. T54E or T54V mutations meant to decrease or increase the hydrophobicity of the cavity, respectively. A32E mutation is unique in that it is the only mutation listed here that involves a mutation of a residue on the heilx. The change induced by this mutation is also unique in that the whole 012 helix in MolB of apo-KL-A32E mutant is completely disordered. The Leu121 is not a conserved residue. The mutations, L121E or L121D, both introduce negatively charged residues deep inside the cavity replacing a neutral residue, which decreases the hydrophobicity of the cavity. Generally, combinations of these mutations do not change the overall fold of the CRABPII greatly, except for the R132K:R111LzA32E (KL-A32E) mutant where the 012 helix is completely disordered in MolB. 3.2.2. The structure of apo-R132K:R111L:L121E:R59E-CRABPII (apo- KLE-R59E) The structure of apo-KLE-R59E is solved at 1.85 A with final Rwork and Rfi-ee of 18.6% and 24.8%, respectively. The Ramachandran plot indicated that all of the residues except Asp126 are in the favored/allowed region. Asp126 in both MolA and M018 is in the disallowed region. However, the electron density map suggests the correct assignment of the residue. The same residue has always been found in the disallowed region in previously solved CRABPII structures (22), which added to our confidence for 102 the assignment for this residue. The overall fold of this mutant is approximately the same as WT-CRABPII (PDB ID: 1CBS, 2FR3) with RMSD of 1.117 for MolA (Figure III-2A) and 0.463 for MolB (Figure III-2B). The MolBs of the two molecules are very similar which is also obvious from the RMSD value of 0.463. The two MolAs are, however, experiencing greater differences. The main difference lies in the region where the C- terrninal of the (12 helix meets the BC-BD loop. A partial unwinding of the 012 helix is clearly observed in MolA of the apo-KLE-R59E mutant structure, which is not observed in MolB and in either of the molecules of the WT-CRABPII structure. To accommodate this movement, the [SC-8D loop in apo-KLE-R59E correspondingly adopts an outward movement, compared to the WT-CRABPII MolA. Similar movements have been observed in previously solved apo-KLE-CRABPII mutant structure (PDB ID: 3D97). The Arg59 in the WT-CRABPII resides on the (BC-8D loop and guards the entrance of the protein. In the WT-CRABPII structure, Arg59 interacts with Gln74 through two water-mediated hydrogen bonding (Figure III-2C), which is disrupted upon the R59E mutation in the apo-KLE-R59E mutant (Figure III-2D). Clearly, the disruption of this interaction does not affect the overall rigidity of the CRABPII. The other three mutations, R132K:R111L:L121E, are all deep inside the active site. A buffer molecule, bis-tris-propane (BTP), is present inside the binding cavity in both MolA and MolB (Figure III-2E). In both molecules, BTP is close to Trp109 and partly fills the hydrophobic hole created by R111L mutation as the overlay between this mutant and the WT-CRABPII shows that Arg] 11 side chain would compete with the BTP for the same space. The atomic positions of BTP in MolA are well defined with an average B factor of 37.8 A2, reasonably higher than the overall B factor (28.9 A2). The BTP in MolB, 103 Figure III-2. Structure of the apo-KLE-RS9E-CRABPII. (a) Overlay of MolA of apo— KLE—R59E-CRABPII (gray) and MolA of apo—WT-CRABPII (green); (b) Overlay of MolB of apo-KLE—R59E-CRABPII (purple) and MolB of apo-WT-CRABPII (cyan); (c) Salt bridge between Arg59 and Gln74 in apo-WT-CRABPII MolA; (d) Disrupted salt bridge in apo-KLE—R59E-CRABPII MolA; (e) 2Fo-Fc map calculated at 1.0 0 for bis- tris-propane bound in apo-KLE-R59E-CRABPII MolA; (f) Water-mediated interactions in MolA of apo-KLE-R59E-CRABPII. 104 however, is poorly defined (average B: 65.2 A2) even though the bulk electron density clearly suggests its presence. Overlaying the MolA and MolB shows that the two BTP molecules do not overlap with each other. In MolA, the BTP makes three hydrogen bonding interactions with the nearby protein residues. In MolB, however, the hydrogen bonding interactions are harder to assign due to our lack of confidence for each atomic position of the molecule. A water-mediated network is clearly shown, involving polar residues, BTP and the ordered water molecules and extending from deep down in the cavity to the 012 helix at the entrance (Figure III-2F). Similar water-mediated interactions have been observed in previously solved CRABPII mutant structures and have been proposed to be important in maintaining the structural integrity of CRABPII by acting as a pillar inside the huge, hydrophobic binding cavity of CRABPII. 3.2.3. The structure of apo-R132K:Y134F:R111 L:L121E-CRABPII (apo- KFLE) The structure of apo-KFLE was solved at a resolution of 1.78 A with Rwork and Rfree of 18.0% and 23.8%, respectively. The Ramachandran plot indicates again that Asp126 is the only residue sitting outside of the allowed region in both molecules of the asymmetric unit. The protein also maintains the rigid overall structure. Overlay of this mutant structure with the WT-CRABPII gives RMSD of 1.137 for MolA (Figure III-3A) and 0.450 for MolB (Figure III-3B), which is similar to the ones observed for KLE- 105 2'9st 3_45 K132 " ‘ 2.78 2.8.4... J -W ' ..... .o‘” ~ ‘8 .~ .' 2.56 ‘ w96 .g3.‘90’ 2.7 2.93 ‘ 3.26 E121 ' ”2.74 "=: 337 mu ,/ 2.77 "223 W260 Figure III-3. Structure of the apo-KFLE-CRABPII. (a) Overlay of MolA of apo- KFLE-CRABPII (purple) and MolA of apo-WT-CRABPII (green); (b) Overlay of MoIB of apo-KFLE-CRABPII (orange) and MolB of apo-WT-CRABPII (cyan); (c) Water- mediated interaction in apo-KFLE—CRABPII MolA. R59E. In fact, this apo-KFLE mutant structure greatly resembles that of the KLE-R59E mutant including the same partial unwinding at the 012 helix region in MolA. The RMSD between the two structures is 0.200 for MolA and 0.276 for MolB. This resemblance is not very surprising. Except for the three common mutations, R132K:R111L:L121E, that are shared between KF LE and KLE-R59E, the fourth mutation in the KFLE mutant, Y134F, is a conserved mutation inside the binding site. Therefore, minimal overall 106 structural changes should be expected upon this mutation. There is no BTP bound in either of the molecules even though the growth condition for the crystal also contains BTP. In MolA of apo-KFLE, the water-mediated interactions are as follows: Glu121- W260-Lys132-W274-W223-carbonyl group of Ala32/carbonyl group of Ser37 (Figure III-3C). Figure III-4. Structure of the apo-KFLD-CRABPII. (a) Overlay of MolA of apo- KFLD—CRABPII (magenta) and MolA of apo-WT-CRABPII (green); (b) Overlay of MolB of apo-KFLD-CRABPII (yellow) and MolB of apo-WT-CRABPII (cyan). 3.2.4. The structure of apo-R132K:Y134F:R111L:L121D-CRABPII (apo- KFLD) The structure of apo-KFLD was solved at 1.94 A with Rwork and Rfree of 21.0% and 27.7%, respectively. Again, the overall structure of this mutant resembles those of KLE-R59E (0.271 for MolA and 0.256 for MolB) and KFLE (0.210 for MolA and 0.188 for MolB). The same partial unwinding is also observed at the C-terminal of the 012 helix 107 region in MolA (Figure 111-4). The helical structures are well maintained in MolB. The crystal was also grown in conditions with BTP as the buffer and yet no BTP was found in the crystal structure of this mutant. Surprisingly, the water-mediated interactions are not observed in this mutant. 3.2.5. The structure of apo-R132K:Y134F:R111L:L121D:T54V-CRABPII (apo-KFLD-T54V) The structure of apo-KFLD-T54V was solved at 1.51 A. Final Rwork and Rfi-ee after refinement are 17.1% and 21.2%, respectively. The Ramachandran plot showed again that Asp126 is the only residue that is outside of the favored/allowed region. There is no buffer molecule bound inside the binding cavity. Interestingly, no partial unwinding is observed for the 012 helix in MolA of this mutant (Figure III-5A,B). Since this KFLD-T54V mutant only differs from the KFLD mutant by a single mutation, T54V, and this mutation did not change the size, shape or conformation of the residue, it may suggest that the appearance of the partial unwinding of the 012 helix C-terrninal is random. Due to the fully maintained helical turns, the structure of this mutant now shares more structural similarity with the apo-WT-CRABPII structure (RMSD: 0.302 for MolA and 0.363 for MolB) than with that of the apo-KFLD mutant (RMSD: 1.136 for MolA and 0.344 for M018). The water-mediated interactions are shown in Figure III-5C. 108 Figure [II-5. Structure of the apo-KFLD-T54V-CRABPII. (a) Overlay of MolA of apo-KFLD-T54V-CRABPII (pink) and MolA of apo-WT-CRABPII (green); (b) Overlay of MolB of apo-KFLD-T54V-CRABPII (purple) and MolB of apo-WT-CRABPII (cyan); (c) Water-mediated interactions in the apo-KFLD-T54V-CRABPII. 3.2.6. The structure of apo-R132K:R111L:T54E-CRABPII (apo-KL- T54E) The structure of apo-KL-T54E was solved at 1.85 A resolution. Final Rwork and Rfree after refinement are 18.8% and 24.4%, respectively. Asp126 is still the only residue that sits outside of the favored/allowed region in the Ramachandran plot. There is no partial unwinding on the 012 helix in MolA. The RMSD values between this mutant structure and that of the WT-CRABPII are 0.369 for MolA (Figure III-6A) and 0.344 for M013 (Figure III-6B). BTP molecules are found in the cavities of both MolA and MolB 109 of this mutant structure (Figure HI-6C). It need to be pointed out that the B factors for both BTP molecules in this structure (56.3 A2 in MolA and 84.0 in MolB) are much higher than the average B factor of the whole structure (24.0 A2). Long-range water- mediated interactions are absent in this mutant probably due to the absence of charged residues (Argl 11 or Glu121) deep inside the cavity. Short-ranged water-mediated interactions are found between the Glu54 and the 012 helix: carboxylate group of Glu54- W187-W172-carbonyl group of Ala36 (Figure III-6D). The Glu54 also forms hydrogen bonds with the bound BTP. Figure [II-6. Structure of the apo-KL-T54E-CRABPII. (a) Overlay of MolA of apo- KL-T54E-CRABPII (hot pink) and MolA of apo-WT-CRABPII (green); (b) Overlay of MolB of apo-KL—T54E—CRABPII (yellow) and MolB of apo—WT-CRABPII (cyan); (c) 2Fo-Fc map of the BTP bound calculated at 1.0 o; (d) Water-mediated interactions in MolA of the apo-KL-T54E-CRABPII. 110 3.2.7. The structure of apo-R132K:R111L:A32E-CRABPII (apo-KL- A32E) K132 2.76 '1" 3.02 \2137 U s‘ 2.25, ..~‘zs1 . ‘ W277 ‘~ W331 ‘35 Figure [II-7. Structure of the apo-KL-ASZE-CRABPII. (a) Overlay of MolA of apo- KL—A32E-CRABPII (purpleblue) and MolA of apo-WT-CRABPII (green); (b) Salt bridge between Arg59 and Glu32 in the apo-KL-A32E—CRABPII MolA; (c) Water- mediated interactions in MolA of the apo-KL—A32E—CRABPII. The crystal structure of apo-KL-A32E was solved at 1.56 A resolution. Final Rwork and Rfi-ee after refinement are 15.3% and 19.9%, respectively. The Ramachandran plot showed that Asp126 is the only residue outside of the favored/allowed region as observed before. The helical structure in MolA is well maintained with no partial 111 unwinding at the C-terrninal of the 012 helix. The RMSD between this MolA and MolA of the WT-CRABPII is calculated to be 0.414, consistent with the observation that these two structures are very similar (Figure III-7A). The introduced Glu32 on the (12 helix forms a salt bridge with Arg59 on the BC-BD loop (Figure III-7B). In MolB of the apo- KL-A32E, however, the 012 helix is completely disordered. The rest of the KL-A32E MolB still folds normally and overlays perfectly with MolB of other apo-CRABPII mutants. Given the fact that the structures listed earlier all contain R132K:R111L mutations and are all capable of maintaining the structural integrity of the protein, A32E mutation in KL-A32E is clearly detrimental to the helical structure of the 012 helix. No buffer molecule is found in the cavity in either MolA or MolB. The water-mediated network in MolA is shown in Figure III-7C. 3.3. Discussion 3.3.1. Rigid overall fold of the apo-CRABPII mutant structures By overlaying all six apo-CRABPII mutant structures together (Figure III-8), it is evident that most of the B-barrel portion of the protein remains rigid in every structure (red structures in Figure 111-8). The flexibility can be seen for the helix-turn-helix region, which serves as a lid governing the entry and exit of the ligand, and the BC-BD loop that is next to the 012 helix. This is consistent with what have been observed in the NMR solution structure (26) of apo-WT-CRABPII and in the crystal structures of other CRABPII mutants (2 7). 112 unwinding unwinding Figure [II-8. Stereoview of the overlay of all apo structures together. Red: MolBs of WT, KLE-R59E, KFLE, KFLD, KFLD-T54V, and KL—T54E; green: MolAs of WT, KFLD-T54V, KL—TS4E, and KL—A32E; blue: MolAs of KLE-R59E, KFLE, and KFLD. Three different conformations of the helices are present in different molecules. The first conformation is shared by all of the MolBs, except for that of the apo-KL-A32E mutant structure which lost its 012 helix completely. The MolBs are shown in red colors in Figure III-8. All of these MolB molecules adopt almost an identical conformation with the RMSD against WT-CRABPII MolB between 0.347 and 0.484. Some of the MolAs adopt the second conformation (WT, KFLD-T54V, KL-A32E and KL-T54E), which is shown in the green structures in Figure III-8. In this conformation, the helices are clearly shifted compared to the MolB structures, together with the neighboring B-tum regions. The third conformation is present in the rest of the MolA molecules (KLE-R59E, KFLE, KFLD, blue colors in Figure 111-8). The interesting feature of this conformation is the partial unwinding at the C-terrninal of the 012 helix. Similar unwinding has also been observed in crystal structures of other CRABPII mutants, such as apo-KLE-CRABPII (PDB ID: 3D97) mutant and apo-KE-CRABPII. 113 The occurrence of multiple conformations of the helices results from their unique flexibility as discussed previously. The reason that all these MolBs display a uniform conformation is probably because of the tight crystal packing. In other words, the tighter interactions from the neighboring molecules that the MolB helices are experiencing probably constrain the helices in these MolBs and leave them with only one favorable conformation to adopt. The helices of the MolAs, however, are in a much less constrained environment in the crystals and therefore are able to adopt different and more relaxed conformations given their flexible nature. It is not certain what causes the partial unwinding at the C-terrninal of the 012 helix. It seems that the R132K and L121E/D mutations are present in all of the structures that share the unwinding (KE, KLE, KLE-R59E, KFLE, and KFLD). However, at the same time there are other mutants that share the same two mutations and yet do not have the unwinding, such as KFLD-T54V and KFLE-T54V (PDB ID: 2FSO). Therefore the partial unwinding is probably favored by the combination of R132K and L121E/D mutations, but not guaranteed. Also, the T54V mutation seems to be able to reverse the unwinding through some unclear interactions. 3.3.2. A32E mutation is detrimental to the 012 helix in apo-KL-A32E Considering that the 012 helices in the apo-CRABPII mutant structures reported here are all well ordered except for the KL-A32E mutant, we are able to see that the A32E mutation is probably responsible for the loss of helical structure in MolB of the apo-KL-A32E CRABPII mutant. Similarly, the 012 helix has been seen to be disordered in a previously solved CRABPII mutant structure, apo-FISW-CRABPII (PDB ID: 2F RS) 114 (22). Phe15 is located at the 011 helix. The F15W mutation clearly destabilizes the 012 helix and results in a totally disordered 012 helix in MolA of the apo-F15W-CRABPII mutant. Steric hindrance between Trp15 and the neighboring 012 helix residues is believed to be responsible for the structural loss of the helix. When it comes to the KL- A32E mutant, the loss of ordered helical structural is probably not because of steric reasons since the Glu32 is close to the entrance and can clearly swing its side chains to avoid bumping into neighboring residues. More possibly, the difference in favored phi angles between Glu and Ala might be the reason for the detrimental effects of Glu32. It is also interesting to notice that the 012 helix in MolA of the apo-KL-A32E mutant is well preserved and the average B factor for the MolA 012 helix is 20.6 A2, which is even lower than the overall B factor of 26.3 A2. In MolA, a salt bridge is established between Glu32 and Arg59 (Figure III-7B). The same salt bridge is maintained in another crystal structure of the same KL-A32E mutant when complexed with a retinoid analog, suggesting a preference of this salt bridge of this mutant. When overlaying the MolA and MolB of different apo mutant structures together, we can see that the BC-BD loop in M013 is shifted further away from the 012 helix than that in MolA (Figure III-8), which causes the Arg59 to move away from the 012 helix. In the structure of MolB of KL-A32E, this movement might cause disruption of the salt bridge between A32E and Arg59. We then speculate that, while the A32E mutation is detrimental to the structural integrity of the 012 helix, the salt bridge between Arg59 and Glu32 observed in MolA of apo-KL—A32E possibly compensate for the reduced structural integrity introduced by having Glu32 on the 012 helix and hold the ordered structure of the helix. 115 The disruption of the salt bridge in MolB due to crystal packing then lead to the disordered 012 helix. 3.3.3. Water-mediated network acts as a pillar inside the cavity The water-mediated interactions have been observed in other apo-CRABPII mutant structures and were suggested to play an essential role in preserving the structural integrity of the protein. Similar situations are seen in the apo-mutant structures reported here. Water-mediated interactions are observed in all six apo structures, except for apo- KFLD. The appearance of these ordered waters that constitute these hydrogen bonding networks in all of these crystal structures again suggest their possible structural roles. If we overlay the water-mediated interactions together, we are able to see that these different network patterns do not overlap with each other. This is due to the different mutations that each mutant is carrying. However, at places where no mutation is involved, we could be able to see some "conserved" water molecules in different mutant structures. However, contrary to what have been suggested before, the lack of the continuous water-mediated network in apo-KFLD does not seem to break down the overall fold of the mutant. Similarly, the lack of long-range water-mediated interactions in apo-KL- A32E and apo-KL-T54E due to the absence of charged residues inside the cavities of the two mutants does not lead to collapsed structures either. It is therefore more reasonable to postulate that the ordered water molecules appear due to the stabilizing effect of the polar residues inside the cavity of CRABPII. 116 3.3.4. The hydrophobic cavity allows binding of BTP BTP contains six hydroxy groups with three on each end of the molecule, which are linked through a hydrocarbon chain. Out of the total of six apo-CRABPII mutant structures reported here, two of them, apo-KLE-R59E and apo-KL-T54E, showed binding of BTP in the cavity of the protein. These binding situations differ in their average B factors, indicating a difference in terms of how comfortable the molecule is bound inside a specific mutant. By overlaying the structures that have BTP in the binding sites, it is clear that these BTP molecules do not overlap. The freedom of binding for BTP is due to lack of highly specific interactions between BTP and the binding environment, which is consistent with the relatively high B-factors observed for some of the bound BTPs. All BTP molecules occupy approximately the same hydrophobic hole created by the R1 1 1L mutation and at the same time tend to adopt multiple conformations with no particular preference energetically. That could possibly explain the less definitive electron densities of specific atoms of BTP in apo-KL-T54E. There are also cases where there is no BTP bound even when the crystal also grew in conditions that contain BTP, indicating a low binding affinity of BTP toward CRABPII mutants. Binding of BTP is not believed to alter the conformation of the protein, which is apparent from the fact that different mutants display minimal structural changes with or without BTP in the binding sites, such as BTP-bound KL-T54E vs. WT-CRABPII and BTP- bound KLE-R59E vs KFLE-CRABPII. 117 3.5. Conclusion The structural information deduced from six high resolution apo-CRABPII mutant structures are presented and discussed. The B-barrel portion of CRABPII is again demonstrated to be very rigid and resistant to multiple mutations involving either conserved or non-conserved residues. The helix-turn-helix is proven to be very flexible, with three conformations observed in different mutants. A partial unwinding at the C- terminal of the 012 helix is seen in three mutant structures. This unwinding is believed to be caused by a combination of helices' flexibility as well as mutations of R132K and L121E/D. Water-mediated interactions linking residues deep inside the cavity with the ones at the helical entrance are found in almost all mutants, indicating their important roles in maintaining the overall fold of these apo-CRABPII mutants. 118 Reference 10. 11. 12. 13. Chambon, P. (1996) A decade of molecular biology of retinoic acid receptors, F aseb J 10, 940-954. Chomienne, C., F enaux, P., and Degos, L. (1996) Retinoid differentiation therapy in promyelocytic leukemia, Faseb J 10, 1025-1030. DeLuca, L. M. (1991) Retinoids and their receptors in differeciation, embryogenesis, and neoplasia., Faseb J 5, 2924-2933. Lotan, R. (1980) Effects of Vitamin-a and Its Analogs (Retinoids) on Normal and Neoplastic-Cells, Biochimica Et Biophysica Acta 605, 33-91. Fisher, G. J ., and Voorhees, J. J. (1996) Molecular mechanisms of retinoid actions in skin, Faseb J 10, 1002-1013. Hong, W. K., Itri, L. M. (1994) Retinoids and human cancer, in The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. B., Roberts, A. B., Goodman, D. S., Ed.) 2nd ed., pp 597-630, Raven Press, New York. Giguere, V. (1994) Retinoic Acid Receptors and Cellular Retinoid-Binding Proteins - Complex Interplay in Retinoid Signaling, Endocrine Reviews 15, 61 -79. Kliewer, S. A., Umesono, K., Mangelsdorf, D. J ., and Evans, R. M. (1992) Retinoid X-Receptor Interacts with Nuclear Receptors in Retinoic Acid, Thyroid- Horrnone and Vitamin-D3 Signaling, Nature 355, 446-449. Mangelsdorf, D. J ., Thummel, C., Beato, M., Herrlich, P., Schutz, G., Umesono, K., Blumberg, B., Kastner, P., Mark, M., Chambon, P., and Evans, R. M. (1995) The Nuclear Receptor Superfamily - the 2nd Decade, Cell 83, 835-839. Morrisskay, G. (1993) Retinoic Acid and Craniofacial Development - Molecules and Morphogenesis, Bioessays 15, 9-15. Bailey, J. S., and SIU, C. H. (1988) Purification and Partial Characterization of a Novel Binding-Protein for Retinoic Acid from Neonatal Rat, J Biol Chem 263, 9326-9332. ONG, D. E., and Chytil, F. (1975) Retinoic Acid-Binding Protein in Rat Tissue - Partial-Purification and Comparison to Rat Tissue Retinol-Binding Protein, J Biol Chem 250, 6113-6117. Sani, B. P. H., D.L. (1974) Retinoic acid-binding-protein in chick-embryo metatarsal skin, Biochem. Biophys. Res. Commun. 61 , 1276-1282. 119 14. 15. 16. 17. l8. 19. 20. 21. 22. 23. 24. Ong, D. E., Newcomer, M. E., & Chytil, F. (1994) Cellular retinoid-binding proteins, in The Retinoids: Biology, Chemistry, and Medicine (Sporn, M. B., Roberts, A. B., Goodman, D. S., Ed.) 2nd ed., pp 283-318, Raven Press, New York. Budhu, A., Gillilan, R., and Noy, N. (2001) Localization of the RAR interaction domain of cellular retinoic acid binding protein-II, Journal of Molecular Biology 305, 939-949. Budhu, A. S., and Noy, N. (2002) Direct channeling of retinoic acid between cellular retinoic acid-binding protein 11 and retinoic acid receptor sensitizes mammary carcinoma cells to retinoic acid-induced growth arrest, Molecular and Cellular Biology 22, 2632-2641. Boylan, J. F., and Gudas, L. J. (1991) Overexpression of the Cellular Retinoic Acid Binding Protein-I (Crabp-I) Results in a Reduction in Differentiation- Specific Gene-Expression in F9 Teratocarcinoma Cells, J Cell Biol 112, 965-979. Boylan, J. F., and Gudas, L. J. (1992) The Level of Crabp-I Expression Influences the Amounts and Types of All-Trans-Retinoic Acid Metabolites in F9 Teratocarcinoma Stem-Cells, J Biol Chem 26 7, 21486-21491. Napoli, J. L. (2000) Retinoic acid: Its biosynthesis and metabolism, Prog Nucleic Acid Re 63, 139-188. Dong, D., Ruuska, S. E., Levinthal, D. J ., and Noy, N. (1999) Distinct roles for cellular retinoic acid-binding proteins I and II in regulating signaling by retinoic acid, J Biol Chem 274, 23695-23698. Kleywegt, G. J ., Bergfors, T., Senn, H., Lemotte, P., Gsell, B., Shudo, K., and Jones, T. A. (1994) Crystal-Structures of Cellular Retinoic Acid-Binding Protein-I and Protein-Ii in Complex with All-Trans-Retinoic Acid and a Synthetic Retinoid, Structure 2, 1241-1258. Vaezeslami, S., Mathes, E., Vasilelou, C., Borhan, B., and Geiger, J. H. (2006) The structure of apo-wild-type cellular retinoic acid binding protein 11 at 1.4 A and its relationship to ligand binding and nuclear translocation, Journal of Molecular Biology 363, 687-701. Thompson, J. R., Bratt, J. M., and Banaszak, L. J. (1995) Crystal-Structure of Cellular Retinoic Acid-Binding Protein-I Shows Increased Access to the Binding Cavity Due to Formation of an Intermolecular Beta-Sheet, J. Mol. Biol. 252, 433- 446. Scapin, G., Gordon, J. 1., and Sacchettini, J. C. (1992) Refinement of the Structure of Recombinant Rat Intestinal Fatty Acid-Binding Apoprotein at 1.2-a Resolution, .1. Biol. Chem. 267, 4253-4269. 120 25. 26. 27. Gunasekaran, K., Hagler, A. T., and Gierasch, L. M. (2004) Sequence and structural analysis of cellular retinoic acid-binding proteins reveals a network of conserved hydrophobic interactions, Proteins 54, 179-194. Wang, L. C., Li, Y., Abildgaard, F ., Markley, J. L., and Yan, H. G. (1998) NMR solution structure of type [1 human cellular retinoic acid binding protein: Implications for ligand binding, Biochemistry- Us 3 7, 12727-12736. Wang, L. C., Li, Y., and Yan, H. G. (1997) Structure-function relationships of cellular retinoic acid-binding proteins - Quantitative analysis of the ligand binding properties of the wild-type proteins and site-directed mutants, J Biol Chem 272, 1541-1547. 121 CHAPTER lV. Investigation of the Interaction between Cellular Retinoic Acid Binding Protein ll and Retinoic Acid Receptor Gamma Ligand Binding Domain 4.1. Background As stated earlier, the effects of retinoic acid (RA) are mediated by at least two families of proteins, nuclear hormone receptors and cellular retinoic acid binding proteins (CRABPs). The transcriptional activities of RA are mediated by two members of the nuclear hormone receptor superfamily: Retinoic Acid Receptor (RAR) and Retinoid X Receptor (RXR). This superfamily is composed of nuclear ligand-activated transcriptional regulators that also include steroid hormone, thyroid hormone and vitamin D3 receptors. The RARs and the RXRs each consist of three isotypes (or, B, and y) encoded by separate genes. Each isotype has several isoforrns, arising from the differential usage of two promoters and alternative splicing (I, 2). The RAR family (RARa, B, and y) binds to and is therefore activated by all-trans (tRA) and 9-cis RA (9cRA), whereas the RXR family (RXRot, B, and y) is activated only by 9cRA (Figure IV-l). l \ \ \ O \ \ \ \ OH I \ HO O all-trans-retinoic acid 9-cis-retinoic acid Figure IV-l. All-trans retinoic acid and 9-cis retinoic acid. 122 N-terminal Hinge C-terminal domain region domain DNA binding Ligand binding ' domain (DBD) doman (LDB) 3D Figure [V-2. Structural organization of nulear receptors (3). The RARs and the RXRs exhibit the conserved modular structure of nuclear receptors and their amino acid sequence can be divided into six regions (A-F) based on homology among themselves and with other members of the nuclear receptor superfamily (Figure IV-2) (4). Region C, consisting of 66 amino acids, contains two zinc finger motifs and is the sequence specific DNA binding domain (DBD). Within a given species, region C is highly conserved between the three RAR isotypes (94-97%) and between the three RXR isotypes (91-97%). Region B consists of 220 amino acids and is the ligand binding domain (LBD) which is also highly conserved between RAR isotypes (84-90%) and RXR isotypes (88-95%). This region is functionally complex, including the ligand binding domain, the ligand-inducible transcriptional activation function AF -2, and a dimerization surface. The interspecies conservation of the DBD and the LED of a given RAR type or of a given RXR type is even greater than the similarity found when 123 comparing the three RAR types or the three RXR types within a given species, which suggests that the DNA and ligand binding properties of the various RARs and RXRs could exhibit subtle functional differences. Region D is located as a hinge between the DBD and the LBD and is conserved both among RARs or RXRs within a given species and is conserved between species for a given RAR or RXR. The C-terminal F domain, which exists in the RARs, is missing in the RXRs and its function is still unknown. However, the amino end A/B regions of RARs and RXRs have both been shown to contain a second transcriptional activation function AF - 1 . Figure [V-3. Comparison of the crystal structures of the apo-RXRa (A, PDB: lLBD; (5)) and holo-RARy (B, PDB: 2LBD; (6)) ligand-binding domains reveals the ligand-induced conformational change. N-terminal is labeled in blue and C-tenninal in red in the rainbow representation in both structures. 124 The crystal structures of the LBDs of both RARs and RXRs have been determined (Figure 3) (5, 6). The LBDs are formed by 12 conserved alpha-helices and a beta-tum (between H5 and H6) which are folded into a three-layered, anti-parallel helical sandwich with H4, H5, H8, H9 and H11 situated between H1, H2 and H3 on one side and H6, H7 and H10 on the other. In this structure, the C-terrninal helix H12, which encompasses the AF-2 activation domain, points away from the LBD core. By comparing the crystal structures of apo-RXRa LBD and holo-RARy LBD, the major conformational change before and after ligand binding seems to be the relocation of H12, which after ligand binding folds back towards the LBD core and “seals” the ligand entry site. In the absence of the ligand, retinoic acid receptors are found primarily in the nucleus. They bind as asymmetric, oriented RAR/RXR heterodimers to specific DNA sequences, RA response elements (RAREs), composed typically of two direct repeats of a core hexameric motif, (A/G)G(G/T)TCA (2, 7). The classical RARE is a 5-bp-spaced direct repeat (DRS). However, RAR/RXR heterodimers also bind to direct repeats separated by 1 bp (DRl) or 2 bp (DR2). Unliganded and DNA-bound retinoid receptors repress transcription through recruitment of the corepressors NcoR and SMRT (8, 9). After ligand binding, the resulting conformational changes cause the release of the corepressors and recruit coactivators which upon association with larger complexes with chromatin modifying and remodeling activities will decompact repressive chromatin and facilitate the positioning of the transcriptional machinery at the promoter region (9-11). CRABP I and II, are the other family of proteins that mediate the biological effects of RA (12-15). CRABPs belong to a family of proteins called intracellular lipid- 125 binding proteins (iLBP), including fatty acid-binding proteins, lipid-binding proteins, and cellular retinoid-binding proteins (16). CRABPs are responsible for modulating the intracellular existence of RA. The sequences of CRABPs are highly conserved in vertebrates (17, 18). CRABP I and II are closely related, with an identity of 76% in human. CRABPs are believed to act as carriers of RA, serving to solubilize and protect RA in the aqueous environment of the cytosol. In addition to this general role, CRABPs have also been shown to facilitate RA modification upon interaction with other proteins (12-15). Despite the sequence similarity between CRABPI and CRABPII, the two proteins have been shown to have distinct roles in regulating RA (17). Professor Noy and colleagues demonstrated that the rate constant for movement of RA from CRABPII, but not CRABPI, to RAR strongly depends on the concentration of RAR (19). They therefore suggested that transfer of RA from CRABPI to RAR involves dissociation of RA from CRABPI, followed by association of RA with the acceptor. In contrast, movement of RA from CRABPII to RAR is facilitated by a mechanism involving direct interactions between CRABPII and RAR. However, the interaction between the two is short-lived according to their observation as no stable complex could be trapped in spite of several attempts including chemical crosslinking, electrophoresis under denaturing conditions, fluorescence anisotropy titrations and electrophoretic mobility shift assays. Their later investigations revealed three spatially aligned residues E75Q, K81P, and E102K (CRABPI/CRABPII), displaying dramatic changes in electrostatic surface potential of the two proteins (1 7). Substituting the corresponding CRABPII residues onto CRABPI conferred upon the protein the ability to channel RA to RAR and to enhance the 126 transcriptional activity of RAR in cells. Conversely, substituting the three residues on CRABPII into their CRABP I counterparts resulted in loss of the ability to interact with RAR and to enhance the activity of the receptor. Professor Chomienne and co-workers also worked on the interaction between RAR and CRABPII. By attaching CRABPII with fluorescence proteins they have demonstrated the distribution of CRABPII in the nucleus (20). However, contradictory to what had been suggested by Professor Noy, they have shown that CRABPII was associated with MR0 and RXRa in a ligand-independent manner through a GST pull- down assay using GST-tagged RARa or RXRor. Moreover, they have demonstrated that, in the presence of retinoids that bind both the nuclear receptors and CRABPII, transcription by RARot-RXRot heterodimers is enhanced by the presence of CRABPII, suggesting CRABP II as a novel transcriptional regulator involved in RA signaling. COOH Figure [V-4. The synthetic retinoid analog CD270. To further investigate the problem, Budhu and Noy performed a similar in vitro pull-down assay using immobilized His-tagged RARor-LBD incubated with CRABPII in the presence of RA or a synthetic analog CD270 (Figure IV-4) which had been reported to interact with CRABPII with a high affinity but to be a poor ligand for RARor (18). 127 The results are consistent with their earlier statement that CRABPII interacts with RAR in a ligand-dependent manner and that the complex is not stable. No CRABPII was found to co-precipitate with RARor-LBD in the absence of ligand or in the presence of RA. However, CD270 stabilized the complex in a dose-dependent manner and CRABPII was co-precipitated with RARy-LBD in the presence of CD270. As we have shown earlier, among our mutants of CRABPII, the triple mutant KLE-CRABPII shows an excellent retinal-binding property (Kd = 1.4 :1: 4.9 nM). In this particular protein, retinal forms a protonated Schiff base with the lysine residue in position 132, which can be followed by the UV red shift from 380 run (free retinal) to 447 nm (protonated Schiff base). This covalent protein-ligand complex gave us an opportunity that resembles the situation in Noy’s research involving the synthetic CD270. In both cases, ligands prefer to stay with the CRABPII proteins instead of being transferred into RAR, which has been suggested by Noy and colleagues as key in order to achieve a stable complex between the two proteins. We therefore set out to detect the protein-protein interaction between CRABPII and RAR employing the all-trans-retinal- bound KLE-CRABPII mutant. 4.2 Results and discussion 4.2.1. In vitro binding assay. With both RARy-LBD and CRABPIIs (WT-CRABPII and KLE-CRABPII) in hand, the binding assay was performed. His-tagged RARy-LBD was first immobilized on Ni-NTA resin. WT-CRABPII and KLE-CRABPII were then incubated with the 128 immobilized RARy-LBD either in the absence of ligand or in the presence of either RA (for wild type) or retinal (for triple mutant) at 4°C for 1 h. The resin was then washed three times with Tris buffer (10 mM Tris, 100 mM NaCl, pH 8.0). SDS gel loading buffer was then added to the beads and proteins attaching to the resin were analyzed by SDS electrophoresis. The results are shown in Figure IV-S. For wild-type CRABPII, no interaction between RARy-LBD and WT-CRABP II was observed either in the absence (lane 6) or in the presence of RA (lane 7). The same results were obtained for the triple mutant CRABP II: no KLE-CRABPII was co-precipitated with RARy-LBD either in the absence (lane 8) or in the presence of RA (lane 9) or retinal (lane 10). A. 12345678910 Figure IV-5. In vitro binding assay. 1. MW; 2. WT-CRABPII; 3. KLE-CRABPII; 4. Ni-NTA beads + WT-CRABPII; 5. Ni-NTA beads + KLE-CRABPII; 6. Ni-NTA beads + His6-RAR-LBD + WT-CRABPII; 7. Ni-NTA beads + His6-RAR-LBD + WT-CRABPII + RA; 8. Ni-NTA beads + His6-RAR-LBD + KLE-CRABPII; 9. Ni-NTA beads + His6- RAR-LBD + KLE-CRABPII + RA; 10. Ni-NTA beads + His6-RAR-LBD + KLE- CRABPII + retinal. A. Coomassie blue stain; B. Silver stain. 129 4.2.2. Following the protonated Schiff base formation by UV. As mentioned before, there is a major red shift in the UV spectrum from 380 nm to 447 run before and after retinal binds to KLE-CRABPII. In order to investigate whether RARy-LBD will prevent the formation of the protonated Schiff base, RARy- LBD was added into the complex between KLE-CRABPII and retinal. A UV scan was performed and it turned out that the protonated Schiff base is no longer existent after RAR-LBD was added, based on the fact that the UV absorbance went back to around 375 nm. Two possibilities were proposed: one is that the Schiff base might have been hydrolyzed and retinal was transferred into RARy-LBD; the other possibility is that the Schiff base is not completely hydrolyzed, but instead of being protonated, the Schiff base is now unprotonated, which could also lead to an absorption at 375 nm. 4.2.3. In vitro binding assay using reductively aminated Schiff base complex. Since a secondary amine base is much more stable than a protonated Schiff base in the aqueous environment, a reductive amination was conducted using Na(CN)BH3. The excess of retinal in the reaction was removed from the reduced protein-ligand complex by buffer exchange using a concentrator. Similar in vitro binding assay was performed. However, again, no co-precipitation with RARy-LBD was observed for reduced KLE-CRABPII-retinal complex. 4.2.4. Presence of Tween20 leads to hydrolysis of the protonated Schiff base. 130 Tween20 (0.1%) was present in the purified RARy-LBD solution. It was introduced during the dialysis after Ni-NTA purification (see detailed procedures in Chapter VI). The function of this detergent is to prevent aggregation of the protein. Since detergent is well known to be able to change the solubility of proteins in buffers, it is suspected that the presence of Tween20 might lead to certain conformational change of CRABPII, which will eventually lead to the hydrolysis of the PSB. Abs 300 400 500 600 Wavelength (nm) Figure IV-6. Detergent titration of the protonated Schiff base formed between KLE- CRABPII and retinal. By adding the Tween20 containing buffer into the Schiff base complex, it was observed that the absorption went back to 375 nm. A detergent titration experiment was also performed (Figure IV-6). To the Schiff base formed complex was added Tris buffer 131 containing different concentrations of Tween20. The concentration of Tween20 that starts to destroy the protonated Schiff base was found out to be between 0.001% and 0.005% (v/v). 4.2.5. Purification of RARy-LBD without detergent. Since the detergent was the reason for disrupting the protonated Schiff base, RARy-LBD was then purified without detergent. After Ni column, the protein solution was dialyzed against 10 mM Tris, 100 mM NaCl, pH 8.0. It was then concentrated, followed by gel filtration purification. However, the spectrum of the gel filtration looked quite different from the previous run when Tween20 was used in the buffer. RARy-LBD purified this way came off the column quite early and covered a much broader range of fractions. What’s more, the UV of RARy-LBD after purification does not look normal. Absorption at 260 nm is strongly interfering the protein absorption at 280 nm, suggesting possible nucleotide contamination. The fact that RARy-LBD came off the gel filtration column earlier as well as the fact that the UV spectrum of RARy-LBD showed an elevated baseline indicate possible aggregation of the protein. To test this possibility, detection for inclusion bodies was performed and insoluble RARy-LBD was detected in the cell pellet. To reduce the inclusion bodies and aggregation problems, several things were modified in the RARy- LBD purification. The optimized purification included the following steps. After OD600 reached 0.6, IPTG was added and the growth was continued at 16°C for 20 h (previously 5 h at r.t.). The proteins were eluted off the Ni-NTA column by 250 mM imidazole, 100 mM NaCl, 10 mM Tris, pH 8.0 (previously 100 mM EDTA, 500 mM NaCl, 20 mM Tris, 132 pH 8.0). After the Ni column, each individual fraction was examined by UV and the best fraction(s) were immediately loaded onto the Sephadex G-50 gel filtration column. It was then eluted with 10 mM Tris, 100 mM NaCl, 10% glycerol, pH 8.0. The fractions were again examined under UV and it turned out that most of the fractions obtained this way showed maximum absorption very close to or exactly at 280 nm, suggesting much less or even no contamination from nucleotides. 4.2.6. Circular dichroism of RARy-LBD and fluorescence titration of RARy-LBD with retinoic acid. The purified RARy-LBD was then characterized first with CD. Both apo- and holo-RARy-LBD were measured and the obtained spectra resemble the literature data (21 ). a-Helix content was indicated by the negative ellipticity with minima at 222 and 208 nm. There are no major changes before and after tRA binding, which is also consistent with the published data. 800000 750000 g 700000 - g 650000 « 8 E 600000 . 550000 - 500000 " ___AT TIA T I I I l 0 0.0000002 0.0000004 0.0000006 0.0000008 0.000001 Concentration of RA Figure [V-7. Fluorescence titration of RARy-LBD with all-trans-retinoic acid. 133 0.2 A. —-0 ——0.1 eq 0.3 B. --0.1 eq 0.25 _o.2 eq 0.3 eq — 0.4 eq ——0.5 eq —0.6 eq —0.7 eq b‘ —0.8 eq // \ —-0.9 eq 1.0 eq 250 300 350 400 450 500 550 600 um I Figure IV-8. Titration of KLE-CRABPII with retinal. A. Using KLE-CRABPII purified earlier; B. Using fully active KLE-CRABPII. 134 Fluorescence titration of RARy-LBD with all-trans-retinoic acid was also performed to verify the ligand binding ability of the purified protein. The extinction coefficient of RARy-LBD was measured to be 11,972 chM-l. Titration data is shown in Figure lV-7. The calculated Kd is 0.5 nM, which is very close to the 0.6 nM binding affinity reported in the literature. 4.2.7. Titration of KLE-CRABP II with retinal followed by UV. A perfect KLE-CRABPII (R132K:R111L:L121E) should bind retinal up to 1 equivalent and still maintain the maximum absorption at around 440 nm corresponding to the protonated Schiff base. However, the KLE-CRABPII that was purified in the previous batch failed to give a decent titration (Figure IV-8A). The retinal could only be added up to 0.5 eq and after that the absorption started to blue shift. Later the same titration using KLE-CRABPII purified in a different batch turned out to be fully active (Figure IV-8B). This batch of protein was later used in the in vitro binding assays. 4.2.8. In vitro binding assay using RARy-LBD directly from Ni-NTA resin. Since we had concerns on loss of activity of RARy-LBD due to dialysis and other handling of the protein, we decided to try the binding assay one more time using the RARy-LBD directly from the Ni resin. This way we would save the effort of eluting the protein off the Ni resin and later on binding it again onto the resin and also exert 135 minimum handling of the protein. Moreover, since the purity of the protein was proved to be fairly good after Ni-NTA purification, this seemed to be a method worth trying. Besides RARy-LBD, retinal solution was also freshly prepared. The KLE- CRABPII used here is the one proved to be fully functional based on the titration experiments described earlier. In this assay the interaction between KLE-CRABPII and RARy-LBD was investigated and that between WT-CRABPII and RARy-LBD was not. The procedure is the same as described earlier in the first in vitro binding assay. The results are shown in Figure IV-9. KLE-CRABPII showed up together with RARy-LBD on the gel either in the absence (lane 3) or presence of retinal (lane 4), however, in the same intensity as the control (lane 2), indicating again lack of interaction between the two proteins. Lane 5, 6 and 7 showed the protein content of the supematants corresponding to lane 2, 3 and 4 respectively. As shown, the same amount of KLE-CRABPII appeared on the gel, suggesting the same degree of binding for KLE-CRABPII both in control and in real assays and therefore no interaction between the two proteins. Lane12345678910 Figure [V-9. In vitro binding assay using RARy-LBD directly from Ni-NTA resin. 1. RAR-LBD on Ni beads; 2. Ni-NTA beads + KLE-CRABPII; 3. Ni-NTA beads + Hisé-RAR-LBD + KLE-CRABPII; 4. Ni-NTA beads + His6-RAR-LBD + KLE-CRABPII + retinal 5. Supernatant of lane 2; 6. Supernatant of lane 3; 7. Supernatant of lane 4; 8. Washed off lane 2; 9. Washed off lane 3; 10. Washed off lane 4. 136 4.3. Conclusions Despite several efforts, no in vitro protein-protein interaction between CRABPII and RAR-LBD is detected. CD270 should be used to set up a positive control for future binding experiments. 137 References 10. 11. 12. Gronemeyer, H., and Laudet, V. (1995) Transcription factors .3. Nuclear receptors, Protein Profile 2, 1173-1308. Leid, M., Kastner, P., and Chambon, P. (1992) Multiplicity Generates Diversity in the Retinoic Acid Signaling Pathways, Trends Biochem Sci 1 7, 427-433. Hoffmeier. http://cn.wikipcdia.org/wiki/Image:Nuclcar Receptor Structurepng. Steinmetz, A. C. U., Renaud, J. P., and Moras, D. (2001) Binding of ligands and activation of transcription by nuclear receptors, Annu Rev Bioph Biom 30, 329- 359. Renaud, J. P., Roche], N., Ruff, M., Vivat, V., Chambon, P., Gronemeyer, H., and Moras, D. (1995) Crystal-Structure of the Rar-Garnma Ligand-Binding Domain Bound to All-Trans-Retinoic Acid, Nature 378, 681-689. Bourguet, W., Ruff, M., Chambon, P., Gronemeyer, H., and Moras, D. (1995) Crystal-Structure of the Ligand-Binding Domain of the Human Nuclear Receptor er-Alpha, Nature 375, 377-382. Mangelsdorf, D. J., and Evans, R. M. (1995) The er Heterodimers and Orphan Receptors, Cell 83, 841-850. Aranda, A., and Pascual, A. (2001) Nuclear hormone receptors and gene expression, Physiol Rev 81, 1269-1304. Glass, C. K., and Rosenfeld, M. G. (2000) The coregulator exchange in transcriptional functions of nuclear receptors, Gene Dev 14, 121-141. Nagy, L., Kao, H. Y., Love, J. D., Li, C., Banayo, E., Gooch, J. T., Krishna, V., Chatterjee, K., Evans, R. M., and Schwabe, J. W. R. (1999) Mechanism of corepressor binding and release from nuclear hormone receptors, Gene Dev 13, 3209-3216. Perissi, V., Staszewski, L. M., Mclnemey, E. M., Kurokawa, R., Krones, A., Rose, D. W., Lambert, M. H., Milburn, M. V., Glass, C. K., and Rosenfeld, M. G. (1999) Molecular determinants of nuclear receptor-corepressor interaction, Gene Dev 13, 3198-3208. Boylan, J. F ., and Gudas, L. J. (1992) The Level of Crabp-I Expression Influences the Amounts and Types of All-Trans-Retinoic Acid Metabolites in F9 Teratocarcinoma Stem-Cells, J Biol Chem 26 7, 21486-21491. 138 13. 14. 15. 16. 17. 18. 19. 20. 21. F iorella, P. D., and Napoli, J. L. (1991) Expression of Cellular Retinoic Acid Binding-Protein (Crabp) in Escherichia-Coli - Characterization and Evidence That Holo-Crabp Is a Substrate in Retinoic Acid Metabolism, J Biol Chem 266, 16572- 16579. Fiorella, P. D., and Napoli, J. L. (1994) Microsomal Retinoic Acid Metabolism - Effects of Cellular Retinoic Acid-Binding Protein (Type-I) and C18- Hydroxylation as an Initial Step, J Biol Chem 269, 10538-10544. Napoli, J. L., Boerrnan, M. H. E. M., Chai, X., Zhai, Y., and Fiorella, P. D. (1995) Enzymes and Binding-Proteins Affecting Retinoic Acid Concentrations, J Steroid Biochem 53, 497-502. Ong, D. E., Newcomer, M. E., & Chytil, F. (1994) Cellular retinoid-binding proteins, in The Retinoids: Biology, Chemistry, and Medicine (Spom, M. B., Roberts, A. B., Goodman, D. S., Ed.) 2nd ed., pp 283-318, Raven Press, New York. Budhu, A., Gillilan, R., and Noy, N. (2001) Localization of the RAR interaction domain of cellular retinoic acid binding protein-II, Journal of Molecular Biology 305, 939-949. Budhu, A. S., and Noy, N. (2002) Direct channeling of retinoic acid between cellular retinoic acid-binding protein 11 and retinoic acid receptor sensitizes mammary carcinoma cells to retinoic acid-induced growth arrest, Molecular and Cellular Biology 22, 2632-2641. Dong, D., Ruuska, S. E., Levinthal, D. J., and Noy, N. (1999) Distinct roles for cellular retinoic acid-binding proteins I and II in regulating signaling by retinoic acid, J Biol Chem 274, 23695-23698. Delva, L., Bastie, J. N., Rochette-Egly, C., Kraiba, R., Balitrand, N., Despouy, G., Chambon, P., and Chomienne, C. (1999) Physical and functional interactions between cellular retinoic acid binding protein [1 and the retinoic acid-dependent nuclear complex, Molecular and Cellular Biology 1 9, 7158-7167. Lupisella, J. A., Driscoll, J. E., Metzler, W. J., and Reczek, P. R. (1995) The Ligand-Binding Domain of the Human Retinoic Acid Receptor-Gamma Is Predominantly Alpha-Helical with a Trp Residue in the Ligand-Binding Site, J Biol Chem 270, 24884-24890. 139 CHAPTER V. Preliminary X-Ray Crystallographic Studies of Cellular Retinaldehyde Binding Protein 5.1. Background As stated earlier, the lack of a three-dimensional structure of the cellular retinaldehyde binding protein (CRALBP) has hindered many research efforts in addressing important biological questions associated with this protein. So we decided to tackle this problem. 5. 2. Crystallization of the full-length CRALBP The CRALBP (rCRALBP) gene was generously provided by professor John W. Crabb. The rCRALBP gene was then cloned into both His-tagged (pETl9b) and non- fusion (pET17b) plasmids and each was expressed respectively in BL21(DE3) cells. 5.2.1. His-tagged full-length CRALBP crystallization. The His-tagged rCRALBP was purified using Ni-NTA affinity chromatography to homogeneity (Figure V-l). The binding property of this His-tagged rCRALBP was tested by observing UV shifts while titrating the protein with one of its natural ligands, l l-cis-retinal. The purified His-rCRALBP was then tested at room temperature for crystallization. Primary screen of the apo protein at different concentrations (25 mg/mL, 140 10 mg/mL, and 5 mg/mL) were performed and results obtained were compared. None of these trials gave protein crystals. Figure V-l. Pure fractions of His-rCRALBP eluted after Ni-NTA purification. Since room temperature crystallization did not work well. Crystallization in the cold room (4°C) was then pursued. Primary screen of this protein (9 mg/mL) at 4°C gave several hits, which later on were all proven to be salt crystals. Since the apo protein failed to crystallize, we turned to the holo-protein based on the knowledge that ligand-bound CRALBP was more soluble than the protein itself (I). The protein was concentrated to 10 mg/mL and ll-cis-retinal was added up to 2.5 eq. The primary screen was set up in the cold room (4°C) under dim red light. Boxes were wrapped with aluminum foil and kept in the dark. Unfortunately, no crystal was observed in any conditions. 141 5.2.2. Non-fusion full-length CRALBP crystallization. Since the His-tagged protein failed to give any crystals either in the apo form or bound to ll-cis-retinal, we decided to try to crystallize the protein without the His-tag. CRALBP was expressed using the pETl7b vector in BL21(DE3) codon+ cells as a tag- free protein. The purification largely follows the procedure developed by Crabb and colleagues (I). In the first step of purification, a DEAE column was employed. The roughly purified protein mixture was then applied to a hydroxyapatite (HTP) column and eluted with a gradient of sodium phosphate buffer. This step eliminated most of the impurities and gave moderately pure CRALBP. The last step of purification employed the FPLC source Q column, after which the CRALBP was purified to homogeneity (Figure V-2). l I v V i 1‘ Figure V-2. Non-fusion CRALBP purified. This gel displays the fractions after the FPLC Q column purification. Fractions containing CRALBP are 38 through 43. Lane 1. Molecular Weight Standard; 2. fraction 29; 3. fraction 30; 4. fraction 31; 5. fraction 38; 6. fraction 39; 7. fraction 40; 8. fraction 41; 9. fraction 42; 10. fraction 43. 142 The binding activity of purified CRALBP with its natural ligand 11-cis-retinal was tested using the UV titration method. When low equivalents of ligand (up to 0.2 eq) were added, the complex showed a distinct red shift from 380 nm to ~420 nrn. However, when more ligand was added, the absorption started to shift back toward 380 nm (Figure V-3). By looking thoroughly in the original paper of Crabb, I found that the protein they purified was also not fully active (1). The only graph that showed the red shift of the bound chromophore in the paper is actually looking very similar to our titration with 0.2 eq of ligand being added. Primary crystallization trials of the apo form of this protein were performed at 4°C. However, again, no crystals were observed for this non-fusion full-length CRALBP. — 1.0 eq ~—0.9 eq ——-0.8 eq —0.7 eq -—0.6 eq -——0.5 eq —0.4 eq 0.3 eq ——0.2 eq —0.1 eq —0 eq 500 Figure V-3. UV titration of non-fusion CRALBP with ll-cis-retinal. 143 5.2.3. Full-length CRALBP with a cleavable His-tag. In previous attempts, the purified non-firsion CRALBP showed reduced activity, probably due to the prolonged purification steps. In order to overcome this possible problem, we decided to introduce a cleavable tag to the protein so that the purification could be done in much shorter time and hopefully with similar or even better purity. A. Overnight cutting B. 3 hour cutting -. . - w~~~-mmw‘ 4 Figure V-4. Full length CRALBP expressed in pET28a and purified with his-tag cleaved. A. Overnight cutting followed by FPLC resource Q purification; B. 3 hour cutting followed by FPLC resource Q purification. The full-length gene encoding CRALBP was successfully cloned into the pETQ8a expression vector. The expressed protein contains a poly-histidine tag at the N-terminal followed by a thrombin-cutting site. The full length CRALBP was first over-expressed and purified by Ni-column. The purified His-tagged protein was then treated with thrombin to cleave the tag. It turned out that this step could be non-specific if the cutting time is too long. The overnight cutting gave two bands, which could not be separated after resource Q FPLC or gel filtration (Figure V-4A). Cutting for 3 hours gave a much better result. The full length protein turned out to be the major component, with very 144 little over-cut protein present (Figure V-4B). However, crystallization of this construct did not yield any crystals either. 5.3. Crystallization of the N-terminal-truncated CRALBP. Attempts to crystallize the native, full-length CRALBP all turned out to be not successful. This then lead us to question whether floppy domains exist in native CRALBP, which hinder the crystallization of this protein. According to the literature, the N-temimal of CRALBP is subjected to limited trypsin digestion (2). When we looked back at our trypsin cutting experiment (Figure V-4), we could also identify that the over- digested peptide is only slightly smaller than the full-length protein. This result possibly suggests that a floppy N-terminal exists in CRALBP. What's more, one of the family members of CRALBP, a-tocopherol transfer protein (aTTP), was crystallized with a N- terminal truncation (3). These facts all suggest that N-terminal deletion of CRALBP might help the crystallization of the protein. 5.3.1. Truncation at 43144 on the N-terminal of CRALBP. It is vital to make the right decision on where to make the truncation. Certainly, cutting should not disrupt the globular structure of the protein. It is also important that not a lot of unstructured amino acids are left hanging at the N-terminus after the deletion, otherwise they will still hamper the crystal packing of the protein. To make the decision on the cutting position, I first looked at the corresponding position where others have made deletion to help crystallize aTTP. The first 20 amino acids were actually cleaved off aTTP, while its first on helix starts at position 27 based on its crystal structure. 145 Another family member, Sec14, has its first or helix start at position 33 (4). Besides, based on the model for CRALBP, the proposed first helix starts at around 66 (5). Two sequence alignment methods, ClustalW and MAP, were used to compare the corresponding position 20 of aTTP on CRALBP. However, the two disagree with each other to a large extent especially at the N—tennini of this family of proteins. By comparing the starting position of the first a-helix of each protein based on the secondary structure prediction generated with the two methods, MAP seemed to give more accurate estimations (1AA difference) over ClustalW (5AA difference). Therefore, from the comparison of the MAP sequence alignment, position 43 on CRALBP turned out to be the corresponding position 20 on aTTP. Since 43 is still far enough from position 66, where the first a-helix starts according to the model built for CRALBP, it should be a safe place to make the deletion. Therefore, it was decided that the truncated protein would start from position 44. The truncated gene was generated through PCR and cloned into the pET28a vector. However, multiple attempts to express this truncated CRALBP resulted in no expression at all. The inclusion body was later detected in the pellet of cells after sonication, indicating misfolded proteins. To improve the situation, lower temperature growth (16°C) and less IPTG induction were investigated. Fortunately, this time small amounts of the target protein were obtained after Ni column purification (Figure V-5). After cutting with thrombin, the protein was further purified by resource Q F PLC. The protein was mostly pure except for a contaminant with a higher MW. 146 Lane 1 2 3 4 5 6 7 8 Figure V-S. N-43/44-truncated CRALBP expressed from pET28a and purified by Ni-column. 1. MW; 2. Flow through; 3. First wash; 4. Second wash; 5~8. Elution fractions. However, the 43/44 truncation made at the N-terminus of CRALBP was later proven to be not active for binding to ll-cis-retinal. The truncated protein failed to bind to ll-cis-retinal based on the fact that no red shift of the absorption maxima was observed when the ligand was introduced to the truncated protein. Although red shift of the absorption maxima does not necessarily equal to ligand binding, it is still safe to say that cleavage of 43 amino acids off the N-terminal is too much, which either results in a completely non-functional protein or removes important protein-ligand interactions that lead to no red shift of the bound chromophore. 5.3.2. Discovery of the ideal N-terminal truncation site. To look further into truncation options, we decided to make several truncations at the N-terminus (A10, A15, A20, A25, A30, A35). These mutants were successfully cloned, expressed and purified using the same systems that were employed for the 43/44 truncation construct earlier. During expression of these truncation mutants, it turned out 147 that, as expected, not all the mutants could be expressed. Deletion of up to 25 amino acids was able to maintain good expression of the protein. Deletion of 30 amino acids or more resulted in very poor or no expression (Figure V-6). MW A10 A15 A20 A25 A30 H Figure V-6. Expression of the truncated CRALBPs. To further test whether the well expressed proteins are all active, UV titration experiments were carried out for each of the four mutants: A10, A15, A20 and A25. By continuously adding 0.1 equivalent of the ligand ll-cis-retinal into the protein solutions, red shifts from 380 nm to 410 nm were observed in all cases. The first three proteins (A10, A15 and A20) bind to the ligand almost as well as the full length CRALBP. For these three truncations, red shifts continued until up to 0.4~0.5 eq was added for all three truncations (0.5~0.6 eq for the full length protein). Afterwards, the absorption started going back toward 380 nm. However, the last one, A25, could only take up to 0.3~0.4 eq 148 of ligand before the absorption started going back toward 380 nm (Figure V-7). The result suggested that A25 might not be as active as the other three truncations. A. FL-CRALBP titrated wlth 11-cis-retinal 1.1 eq 1.0 eq 0.9 eq 0.8 eq —0.7 eq —-O.6 eq —-0.5 eq —0.4 eq -0.3 eq — 0.2 eq 0.1 eq --0 eq — 11-cis-retlnal B. 10-CRALBP titrated with 11-cls-retinal -~‘0.6 eq 0.5 eq —0.4 eq —0.3 eq —0.2 eq 0.1 eq —0 eq ABS .° p—c i N U1 0 Figure V-7. UV titration of full-length and truncated CRALBPs with ll-cis-retinal. 149 Figure V-7 (cont’d). C. 15-CRALBP titrated with 11-cis-retlnal D. 20-CRALBP titrated with 11-cis-retinal 150 —0.8 eq --0.7 eq —-~ 0.6 eq ~ , 0.5 eq —0.4 eq —0.3 eq -—-O.2 eq 0.1 eq —0 eq —0.7 eq -.._ 0.6 eq 0.5 eq --0.4 eq —0.3 eq —0.2 eq 0.1 eq —0 eq Figure V-7 (cont’d). E. 25-CRALBP titrated with 11-cis—retinal —0.5 eq —0.4 eq 2nd ~ 0.4 eq 0.3 eq 2nd —0.3 eq --0.2 eq —0.1 eq 2nd 0.1 eq —0 eq F. 30-CRALBP titrated with 11-cis-retinal 0.1 eq —0eq 250 350 450 nm 151 Further proof of A25 being different from other active truncations came from gel filtration experiments. Gel filtration purification for each of the full length CRALBP as well as the four truncations (A10, A15, A20, A25) were carried out, respectively, to further purify the proteins as well as to observe whether they would purify as monomers or dimers or even other association forms. A10, A15, and A20 all gave similar profiles to the wild-type CRALBP for their gel filtration run (a huge peak at the void, a smaller, sharper peak fractions later and a shoulder in between, Figure V-8), while A25 only gave a peak at the void. Figure V-8. Gel filtration separation of full-length-CRALBP. 152 A. Fractions 13-16 titrated with 11-cis-retinal —0.5 eq — 0.4 eq — 0.3 eq 0.2 eq --O.1 eq ABS B. Fractions 17-20 titrated with 11-cis-retlnal — 1.0 eq —-0.9 eq —O.8 eq ---0.7 eq .._ -0.6 eq --0.5 eq —O.4 eq —0.3 eq —0.2 eq 0.1 eq ..._0 eq ABS Figure V-9. UV titration of different fractions from gel filtration separation of A20- CRALBP. 153 Figure V-9 (cont’d). C. Fractions 21-23 titrated with 11-cis-retinal 0.04 U! n < -o.o1 . , A 250 350 450 nm This observation, together with the previous observation that A25 was not as active as the other truncations, also led to the suspicion that the more active form of the protein might reside in the fractions that showed up later in the gel filtration runs of A10, A15, and A20. To test this suspicion, UV titration experiments were carried out for different gel filtration fractions, respectively (Figure V-9). It turned out that the most active protein did come from the later fractions corresponding to the smaller, sharper peak with an approximate molecular weight of 120 KD (0.8 eq, tetramer?). The protein coming out of the void had the worst binding property (0.2~0.3 eq) and the shoulder fractions showed an average binding activity (~0.6 eq). 154 A. CD of void (13-16) fractions from 20-CRALBP gel filtration 150 q 100 50 j 190 200 210 220 230 240 250 B. CD of middle (17-20) fractions from 20-CRALBP gel filtration 190 200 210 220 230 240 250 Figure V-10. CD spectra of different fractions from gel filtration separation of A20- CRALBP. 155 Figure V-10 (cont’d). C. CD of late (21-23) fractions from 20-CRALBP gel filtration 30 J l -20 .. l -70 l l l -120 éae — ——e _. —.— Te —“» ,._,_._ .g. -ej - —.A —. _~»—~— ,. 190 200 210 220 230 240 250 The circular dichroism (CD) absorptions of A20 out of different gel filtration fractions were also measured (Figure V-IO). The CD spectrum of the later fractions resembles the reported CD spectrum of active full length CRALBP. The CD of the void fractions only showed a signal at ~220 nm and was missing the signal at ~208 nm. The middle fractions showed a CD signal as an average of the previous two. Based on the above results, it turned out that A20-CRALBP is the ideal construct, which is hopefully free of floppiness and, at the same time, proven to be able to express well as a functional protein. 156 Figure V-ll. Gel filtration purification of the A20-CRALBP grown at dim-rent temperatures. A. 37°C; B. 16°C. 157 5.3.3. Enhancing the expression and purification of functional CRALBP. As shown in Figure V-8, the gel filtration purification of the A20-CRALBP results in an elution profile with a majority of the protein being eluted at the void. This suggests possible aggregation of the overexpressed protein. To tackle this problem, different growth temperatures were tried and yielded exciting results. It turned out that lower temperature did greatly favor the functional A20-CRALBP over the aggregated form (Figure V-l l). The trypsin cutting conditions were also optimized to remove the tag after the Ni-NTA purification (Figure V-12). The complete procedure for optimized growth and purification of CRALBP (both truncated and the full length) is described in Chapter VI. Full length CRALBP A20 CRALBP Trypsin:prot l :50 l :500 1550 15500 -. . 52% fl ‘ ‘4 fl #5. Q t g ' . ‘Cb.“.”‘ ‘ " I- a c... unused MW STD 05“ 1h 2" 4" 05h ”1 2“ 4*! MW STD 0.511 1h 2h 4h 0.5h lh 2h 4h Figure V-12. Trypsin digestion performed for both full length and A20-CRALBP under 4°C. 158 5.3.4. Crystallization trials. The crystallization of the pure, active apo-A20-CRALBP (10 mg/ml and 20 mg/ml, respectively) at both room temperature and 4°C did not yield any hits. Figure V-13. Crystals of merocyanin-bound proteins. A. merocyanin-bound A20— CRALBP; B. merocyanin-bound KLE-R59W-CRABPII. The crystallization efforts then turned to the holo-CRALBP complexes. Merocyanin has been proven to be an excellent substrate for CRALBP. The merocyanin- A20-CRALBP complex was then used in crystallization. Very interestingly, purple, crystal-like entities were observed in a particular condition (Figure V-l3A). Based on our previous experience with the merocyanin-bound-KLE-RS9W-CRABPII mutant crystals (Figure V-l3B), we know that the purple color can result from the formation of the protonated Schiff base between a protein Lys residue and the aldehyde fimctional group on the ligand. Therefore, we are hopeful that this entity is formed by proteins instead of salts. Unfortunately, this crystal-like entity failed to diffract X-rays, which probably indicates its lack of a more ordered arrangement of the molecules microscopically. Further efforts toward optimization of the quality of the crystals did not 159 improve the situation. However, the discovery so far gives us confidence that the A20- CRALBP finally showed some tendency toward crystallization and that we might be at the very edge of obtaining the first crystal of the CRALBP protein. The synthesis of 11- cis-retinal and ring-locked-l l-cis-retinal is underway and we hope we will have better luck trying to crystallize CRABLP bound with these new ligands. 5.4. Conclusion. The crystallization of CRALBP in both full length and the N-terminally truncated version has been performed. Despite various efforts, no crystals have been obtained so far. However, by introducing retinoid analogs into the protein, we seem to be able to improve the protein's tendency toward crystallization. It seems very promising that future crystallization trials probably with new ligands might lead us to obtaining the first crystal of the CRALBP. 160 Reference 1. Crabb, J. W., Chen, Y., Goldflam, S., West, K., Kapron, J. . (1998) Methods for producing recombinant human cellular retinaldehyde-binding protein, Methods in Molecular Biology 89, 91-104. 2. Crabb, J. W., Gaur, V. P., Garwin, G. G., Marx, S. V., Chapline, C., Johnson, C. M., and Saari, J. C. (1991) Topological and Epitope Mapping of the Cellular Retinaldehyde-Binding Protein from Retina, J Biol Chem 266, 16674-16683. 3. Min, K. C., Kovall, R. A., and Hendrickson, W. A. (2003) Crystal structure, of human alpha-tocopherol transfer protein bound to its ligand: Implications for ataxia with vitamin E deficiency, P Natl Acad Sci USA 100, 14713-14718. 4. Sha, B. D., Phillips, S. E., Bankaitis, V. A., and Lao, M. (1998) Crystal structure of the Saccharomyces cerevisiae phosphatidylinositol-transfer protein, Nature 391, 506-510. 5. Wu, Z. P., Hasan, A., Liu, T. Y., Teller, D. C., and Crabb, J. W. (2004) Identification of CRALBP ligand interactions by photoaffinity labeling, hydrogen/deuterium exchange, and structural modeling, J Biol Chem 279, 27357- 27364. 161 CHAPTER VI. Experimental 6.1. Cloning, mutagenesis, expression and purification 6.1.1. Cloning, mutagenesis, expression and purification of CRABPII mutants. Mutagenesis. Site-directed mutagenesis was performed using the CRABPII-pET17b plasmid following Stratagene’s Quikchange® protocol. The PCR products were transfected into J M109 competent Escherichia coli cells for plasmid maintenance. MW 18 19 21 23 25 28 31 33 35 Figure VI-l. R132K:Y134F:R111L:L121E-CRABPII (KFLE) purified after source- 15Q anion exchange column. The corresponding fraction number is listed on top of each lane. Protein Expression and Purification. Human recombinant CRABPII was expressed and purified as described (1 ). The target gene (CRABPII/pETl7b), as isolated from JM109 162 using Qiagen’s Maxi Prep® DNA isolation kit, was transformed into E. coli strain BL21(DE3)pLysS cells (Stratagene), according to standard protocols. The transformed cells were grown at 37°C in an LB agar plate containing both ampicillin (100 ug/mL) and chloramphenicol (25 ug/mL). A single colony was inoculated in 50 mL of LB containing the same amount of the two antibiotics and grown overnight with shaking at 250 rpm at 37°C. This culture was then transferred to 1 L of LB with the same antibiotics. The culture was incubated at 37°C until A600 reached between 0.6 and 0.8. The expression was induced by addition of 0.5 mM isopropyl-l-B-D-galactopyranoside (IPTG). The growth was continued for another 5 h at 30°C. The cells were harvested using centrifugation (5000 rpm, 10 min) and frozen at -80°C. The frozen cells from 2 L of growth culture were thawed and resuspended in 50 mL of 10 mM Tris (pH 8.0). The suspended cells were lysed by four l-minute bursts of sonication (probe sonicator, Biologics Inc., 60% power) on ice and centrifuged for 30 min at 4°C at 7000 rpm. The supernatant was subjected to FastQ (Q-sepharose Fast Flow Resin) column chromatography (100 mL bed volume; Amersham Biosciences). The column was washed with 10 mM Tris (pH 8.0) and proteins eluted with 100 mM of NaCl in 10 mM Tris (pH 8.0). The fractions were analyzed by SDS 16% polyacrylamide gel electrophoresis and those of the highest purity were combined. The protein solution was desalted by concentration using Centriprep 3 concentrators (Amicon, MWCO 3000), followed by dilution with 10 mM Tris (pH 8.0). The proteins were further purified on a BioRad system (BioLogics Duo Flow, BioRad) using a SourceISQ (Amersham Biosciences) anion exchange column. Fractions were analyzed by UV absorption at 280 nm as well as SDS 16% PAGE. All the steps of purification were performed at 4°C. The 163 most pure fractions were combined and concentrated to ~20 mg/mL, as determined by A280. The pure and concentrated protein solutions were aliquoted and stored at -80°C. 6.1.2. Cloning, expression and purification of RARy-LBD. The pDONR vector (gentamycin resistant) containing the RARy-LBD gene (178- 423) with a hexa-His tag at the amino end was obtained from Dr. Marc Ruff in professor Pierre Chambon’s lab. The gene was subcloned into the NdeI-BamHI sites of pETle and the resulting plasmid was amplified in DHSor. Transformation into E. coli BL21(DE3) provided the construct ready for overexpression. M a. ...-.... Figure VI-2. Elation fractions of RARy-LBD from Ni—NTA purification. Cells were grown in LB medium containing ampicillin as described (2, 3). Cell culture from 6 L shake flasks was centrifuged and the cell pellet was resuspended in 50 mL buffer A (5 mM imidazole, 500 mM NaCl, 20 mM Tris, pH 8.0), followed by sonication. The cell lysate, after centrifirgation, was mixed with Ni-NTA resin (20 mL) 164 and shaken at 4°C for about 1 hour. The flow through was collected. The resin was then washed twice with 50 mL of buffer B (50 mM imidazole, 500 mM NaCl, 20 mM Tris, pH 8.0). The proteins were eluted with buffer C (100 mM EDTA, 500 mM NaCl, 20 mM Tris, pH 8.0) as 5 mL fractions, yielding 14 mg of protein (Figure VI-2). The protein solution was dialyzed against Dialysis buffer (10 mM DTT, 500 mM NaCl, 10 mM Tris, 0.1% Tween20, pH 8.0). It was then concentrated to 2.5 mg/mL, followed by gel-filtration chromatography (Superdex200, Pharmacia) using the same Dialysis buffer. The RARy-LBD (28 KD) was obtained mainly as a homodimer as expected. 6.1.3. Cloning, mutagenesis, expression and purification of CRALBP. 6.1.3.1. Initial constructs of CRALBP. The rCRALBP gene was generously provided by professor John W. Crabb. The rCRALBP gene was then cloned into both His-tagged (pETl9b) and non-fusion (pETl7b) plasmids and expressed respectively in BL21(DE3) cells. The His-tagged rCRALBP was purified using Ni-NTA affinity chromatography to homogeneity (Figure VI-3). The binding of this His-tagged rCRALBP was tested by observing UV shifts while titrating the protein with its natural ligand ll-cis-retinal. Non-fusion CRALBP was expressed from the pETl7b vector in BL21(DE3) codon® cells as a tag-free protein. The purification largely follows the procedure developed by Crabb and colleagues (4). In the first step of purification, a DEAE column was employed using a gradient of 10 mM to 400 mM of NaOAc in 25 mM Tris buffer with 0.1 mM DTT, pH 7.0. The roughly purified protein mixture was then applied onto a 165 hydroxyapatite (HTP) chromatography and eluted with a gradient of sodium phosphate (0-80mM) buffer (25 mM Tris, 300 mM NaOAc, 0.1 mM DTT, pH 7.0). This step eliminated most of the impurities and gave moderately pure CRALBP. The last step of purification employed the FPLC Q column (gradient from 50 mM to 500 mM of NaOAc in 20 mM MOPS and 0.1 mM DTT, pH 7.0), after which the CRALBP was purified to homogeneity (Figure VI-4). Figure VI-3. Pure fractions of the his-tagged rCRALBP eluted off the Ni column. 12345678910 .-,. w 1! g I I c} 6 ""‘~ r. mm... her. Figure VI-4. Non-fusion CRALBP after purification. This gel displays the fractions after the FPLC Q column purification. Fractions containing CRALBP are 38 through 43. Lane 1. Molecular Weight Standard; 2. fraction 29; 3. fraction 30; 4. fraction 31; 5. fraction 38; 6. fraction 39; 7. fraction 40; 8. fraction 41; 9. fraction 42; 10. fraction 43. 166 6.1.3.2. Cloning, mutagenesis, expression and purification of the newer constructs. Molecular cloning. Inserts of full length, A10, A15, A20, A25, A30, A35-CRALBP (all N-terrninal deletions) were prepared using sequential digestion with BamHI and HindIII. The primers used for generating different deletions are listed in Figure VI-5. The digested genes were then successfully ligated into pET28a vectors, respectively. 10 AAGGATCCgtacctgaagaggaacaggag 15 AAGGATCCcaggagctccgtgcccaa 20 AAGGAchcaactggagcagctcac 25 AAGGATCCacaaccaaggaccatggacc 30 AAGGATCCggacctgtctttggcccgt 35 AAGGATCcccgtgcagccagctgc Figure VI-S. Primers used for constructing different N -terminal deletion CRALBP. Capital letters represent non-matching bases. The BamHl cutting sites together with the overhang bases are in purple. Expression and purification. The plasmids were transformed into E. coli BL21 codon69 cells, respectively. Cells were plated on the LB plates containing both kanamycin and chloramphenicol and grown at 37°C for about 12 hrs. Single colonies were used to inoculate 50 mL LB/kan/chlor medium and grown overnight at 37°C. The overnight culture was then used to inoculate 1L LB/kan/chlor medium. The cells were allowed to grow first at 37°C till OD600 reached between 0.5 and 0.6 (roughly 1.5 hrs). The culture was then taken out of the shaker and cooled down in the cold room (4°C) for 2-3 hrs. IPTG was then added to 0.5 mM and the growth was continued at 16°C for ~20 hrs. The 167 cells were harvested after centrifugation at 4,000 rpm for 10 min at 4°C. If not used immediately, cells were usually frozen using liquid nitrogen and stored at -20°C. Cells were defrosted and re-suspended in 40 mL of buffer A (same as described in RAR-LBD purification). Protease inhibitor cocktail tablet (complete, EDTA-free, Roche) was then added. The cell suspension was lysed using four 1-minute bursts of sonication (probe sonicator, Biologics Inc., 60% power) on ice and centrifuged for 30 min at 4°C at 7000 rpm. The supernatant was loaded onto Ni-NTA resin (bed volume of ~ 15 mL) and the flow-through was collected. The resin was then washed with 2x50 mL of buffer B (same as used in RAR-LBD purification). The protein of interest was then eluted with 5x10 mL of the elution buffer (250 mM imidazole, 500 mM NaCl, 20 mM Tris, pH 8.0). The pure fractions were combined and thrombin was added to a molar ratio of ~ 1:500 (thrombin : protein). The solution was dialyzed against 20 mM Tris, 50 mM NaCl, pH 8.0 for a total period of 6-9 hours, during which the buffer was changed 3 times. After dialysis, the protein solution was concentrated using Centriprep 3 concentrators (Amicon, MWCO 3000) to a volume of 1-2 mL. The sample was then placed in an eppendorf tube and centrifuged at 14,000 rpm at -4°C for 10 min to remove precipitate. The supernatant was further purified by a size-exclusion column (HiPrepTM 16/60, SephacrleM S-300, Pharmacia Biotech) on F PLC (Pharmacia Biotech) and eluted with the same buffer as used in the dialysis. The elution profile features two peaks, with one at the void and one that immediately follows (Figure VI-6). Although both peaks show very pure CRALBP, only the later peak was verified to contain fully active protein and therefore is collected. The pure, active protein (Figure VI-7) was then concentrated to the desired concentration, frozen in liquid nitrogen and stored at -80°C. 168 Figure VI-6. Typical size-exclusion purification profile of CRALBP represented by A20-CRALBP grown at 16°C. Figure VI-7. CRALBP after purification. A. Full length CRALBP; B. A20-CRALBP. 169 6.2. Protein crystallization 6.2.1. Co-crystallization of CRABPII mutants bound with ligands. Preparation of complexes of the C15-a1dehyde bound to the CRABPII mutants. The light sensitive C15-aldehyde was synthesized by a colleague (5), dissolved in ethanol, stored at -20°C, and always protected from light. The complex was prepared freshly by adding the Cl 5-aldehyde stock solution (0.245 M in ethanol) to the protein solution (~ 20 mg/mL) in a volume ratio of 1:16 (CU-aldehyde in ethanol : protein solution) to a final molar ratio of ~10:1 (C15-aldehyde : protein). The complex was first incubated at room temperature for 30 min and then 4°C for l~2 hrs before being used for crystallization. The complex was only exposed to dim red light when necessary (crystallization, flash freezing or under the microscope) and was kept in the dark for the rest of the time. Preparation of complexes of the merocyanin bound to the CRABPII mutants. The light sensitive merocyanin was synthesized by a colleague, dissolved in ethanol, stored at - 20°C, and always protected from light. The complex was prepared by adding the merocyanin stock solution (0.032 M) to the protein solution (~ 20 mg/mL) in a volume ratio of 1:10 (merocyanin in ethanol : protein solution) to a final molar ratio of 2.5:1 (merocyanin : protein). The complex was first incubated at 4°C for over 1 hr and then at room temp for several more hrs before being used for crystallization. The complex was only exposed to dim red light when necessary (crystallization, flash freezing or under the microscope) and was kept in the dark for the rest of the time. 170 Figure VI-8. Pictures of crystals of the ligand-bound CRABPII mutants. A. C15- aldehyde-bound KL-A32E-CRABP11; B. C‘s-aldehyde-bound KL-T54E-CRABPII; C. C15-aldehyde-bound KFLDV-CRABPII; D. merocyanin—bound KLE-R59W—CRABPII; E. merocyanin—bound KLE—R59W—CRABPII, taken while the crystal was being cryo protected by the nitrogen stream when sitting on the goniometer at the LS-CAT at APS. 171 Crystallization and Cryoprotection. The hanging drop vapor diffusion method was used. The crystallization was performed at 4°C under dim red light and the crystallization boxes were covered with aluminum foil to avoid light exposure. Boxes were kept at 4°C. Crystals of holo-KLE-R59W were grown in 0.1M bis-tris-propane (BTP), pH 7.5, and 30% PEG 400. Crystals appeared within 24 hrs. On the fourth day crystals were taken out of the drop and flash frozen by immersion in liquid N2. Crystals of holo-KL-A32E were grown in 0.1M Na citrate, pH 7.5, 0.2M ammonium acetate, and 26% PEG 4000 (Figure VI-8A). Crystals appeared within 24 hrs and were immediately transferred to a soaking drop containing 4 uL of the reservoir solution (0.1M Na citrate, pH 7.5, 0.2M ammonium acetate, and 20% PEG 4000), l uL of protein solution (22 mg/mL), and 0.5 uL of ligand stock solution (0.245M in ethanol). After 42 hrs of soaking, crystals were briefly immersed in a cryo solution composed of 0.1M Na citrate, pH 7.5, 0.2M ammonium acetate, 26% PEG 4000, and 20% glycerol and flash frozen in liquid N2. Crystals of holo-KL-T54E were grown in 0.1M NaOAc, pH 5.8, and 18% PEG 6000 (Figure VI-8B). Crystals appeared after 1 week. On the same day crystals were soaked briefly in 0.1M NaOAc, pH 5.8, 18% PEG 6000, and 30% glycerol and flash frozen. Crystals of halo-KFLDV were grown and frozen in the same way as holo-KL- T54E (Figure VI-8C). Crystals appeared after 40 hrs and were flash frozen right after appearance. Crystals of merocyanin-bound KLE-R59W were grown in 0.1M MES, pH 6.3, 0.2M ammonium sulfate, and 30% polyethylene glycol mono-methyl-ether (PEGMME) 172 5000 (Figure VI-8D,E). Crystals displayed a distinctive purple color and grew bigger gradually for two weeks. Crystals were soaked briefly in 0.1M MES, pH 6.3, 0.2M ammonium sulfate, and 30% PEGMME, and 20% glycerol and flash frozen. 6.2.2. Crystallization of apo-CRABPII mutants. The hanging drop vapor diffusion method was used. The droplet consisted 1 ul of concentrated protein (~20 mg/ml) and 1 ul of the reservior. The crystallization was performed at room temperature. All crystals appeared within 24 hours and were frozen within a week after appearance. Cryoprotectants were prepared with either 20% or 30% glycerol in addition to the components of the corresponding growth conditions in the same concentrations. All of these apo CRABPII mutant structures are in the P 1 spacegroup. Crystals of the apo-R132K:R111L:A32E-CRABPII (KL-A32E) were grown in 0.1M Tris, pH 8.5, 0.2M NaOAc, and 26% PEG 4000 (Figure VI-9A). The crystals were briefly immersed in a cryo solution containing 20% glycerol before they were flash frozen in liquid nitrogen. Crystals of the apo-R132K:R111L:T54E-CRABPII (KL-T54E) were grown in 0.1M bis-tris-propane (BTP), pH 7.5, and 30% PEG 4000. Two microliters of the cryoprotectant (20% glycerol) was added to the crystal-containing droplet and the crystals were then taken out of the drop after ~30 seconds, followed by flash freezing in liquid nitrogen. Crystals of the apo-R132K:Yl34F:R111L:L121D:T54V-CRABPII (KFLD-T54V or KFLDV) were grown in 0.1M Tris, pH 7.5, 0.2M NaOAc, and 24% PEG 4000 (Figure 173 VI-9B). Two microliters of the cryoprotectant (20% glycerol) was added to the crystal- containing droplet and the crystals were incubated in the new drop for ~30 sec before being flash frozen in liquid nitrogen. Figure VI-9. Pictures of crystals of the apo-CRABPII mutants. A. apo-KL-A32E- CRABPII; B. apo-KFLDV-CRABPII; C. apo-KFLD—CRABPII; D. apo-KFLE—CRABPII. The apo-R132K:R111L:L121E:R59E-CRABPII (KLE-R59E) was crystallized in 0.1M BTP, pH 7.5, and 22% PEG 8000. Two microliters of the cryoprotectant (30% glycerol) was added to the crystal-containing droplet and the crystals were incubated in the new drop for ~30 sec before flash frozen in liquid nitrogen. 174 The apo-R132K2Yl34FzR111L:L121D-CRABPII (KFLD) was crystallized in 0.1M BTP, pH 8.1, and 22% PEG 4000 (Figure VI-9C). Cryoprotection and flash freezing were performed in the same way as for apo-KLE-R59E. The apo-R132K:Y134F:Rll1L:L121E-CRABPII (KFLE) was crystallized in 0.1M BTP, pH 7.0-8.5, and 20-30% PEG 4000 (Figure VI-9D). Cryoprotection and flash freezing were performed in the same way as for apo-KLE-R59E. 6.3. Data processing and structure determination 6.3.1. Structures of the C15-aldehyde-bound CRABPII mutants. Data Collection. Datasets were collected at the Advanced Photon Source LS-CAT 21-ID at Argonne National Laboratory. Processing and scaling was done using HKL2000 (6). For the dataset of KLE-R59W (1.95 A, P3121), the crystal to detector distance was 200 mm and 120° of data were collected with an oscillation of 1°. For KL-A32E (1.22 A, P21), the crystal to detector distance was 75 mm and 200° of data were collected with an oscillation of 1°. For KL-T54E (2.00 A, P6122), the crystal to detector distance was 200 mm and 150° of data were collected with an oscillation of 05°. For KFLDV (1.54 A, P212121), the crystal to detector distance was 150 mm and 360° of data were collected with an oscillation of 1° (only the first 120° were scaled). All datasets were collected at the Advanced Photon Source at Argonne National Laboratory. Datasets for KL-A32E, KL-T54E, KFLD-T54V and KLE-R59E were collected at LS-CAT 21-ID. The dataset for KFLD was collected at SBC-CAT 19-BM. 175 The dataset for KFLE was collected at COM-CAT 32-ID. All datasets were processed with HKL2000 (6). Structure solution and refinement. Structures of KLE-R59W and KL-A32E were solved using Molrep (7) in CCP4 (8) with KLE-CRABPII as the search model (PDB ID:ZG7B). Structures of KL-T54E and KFLDV were solved using Phaser (9) with the same model. Model building was accomplished in both TURBO-FRODO (10) and COOT (11) for each of the four structures. REFMACS (12-15) within the CCP4 suite of programs was used to refine the structures against 95% of the data, while 5% of data was chosen randomly for cross validation, Rfree. For KL-A32E, SHELXL-97 (16) was used in the later stages of refinement. For KL-T54E, TLSMD analysis (17, 18) was performed and TLS refinement in REFMACS with 3 groups for Chain A and 4 groups for Chain B resulted in improved B factors and R/Rfree. Asp126 was the only residue outside of the allowed region in the Ramachandran plot in all four structures after refinement. In holo-KL-A32E it resides in the generously allowed region, while in the other three structures in the disallowed region. The same Asp126 was found to reside in disallowed regions in all previously determined CRABPII structures (19). The good electron density maps at this position greatly enhanced our confidence in the non-favorable geometry of this residue. 6.3.2. Structure of the merocyanin-bound KLE-R59W-CRABPII mutant. Data Collection. The dataset was collected at the Advanced Photon Source LS-CAT 21- ID at Argonne National Laboratory. Processing and scaling was done using HKL2000 176 (6). The dataset was determined to be in P3121 spacegroup with the highest resolution at 2.60 A. The crystal to detector distance was 250 mm and 180° of data were collected with an oscillation of 1°. Structure solution and refinement. Structure of the merocyanin-bound KLE-R59W and was solved using Molrep (7) in CCP4 (8) with KLE-CRABPII as the search model (PDB ID:2G7B). Model building was accomplished in both TURBO-FRODO (10) and COOT (11). REFMACS (12-15) within the CCP4 suite of programs was used to refine the structure against 95% of the data, while 5% of data was chosen randomly for cross validation, Rfree. TLSMD analysis (I 7, 18) was performed and TLS refinement in REFMACS with 5 groups resulted in improved overall B-factor and R-factors. 6.3.3. Structures of the apo-CRABPII mutants. All structures were solved using either Molrep (7) or rigid body refinement in RefmacS (12-14) in CCP4 with apo-WT-CRABPII MolA as the search model (PDB ID: 2FS6). Model building was accomplished in both COOT (11) and TURBO-FRODO (10) for every structure. Restrained refinement in Refinac5 was used to refine the structures against 90-95% of the date, while 5-10% of data were chosen randomly for cross validation, Rfree. Anisotropic refinement for temperature factors was employed for KL- A32E and KFLD-T54V (13). For KL-T54E, TLSMD analysis was performed and TLS refinement with 9 groups for both chain A and chain B resulted in improved B factors as W611 as Rwork/Rfree (15, 17, 18). 177 dosm E5288 “mg 05 8 Home“ momoficocam E mos—«.5 838 888 88.8 488 88.8 488 88.8 £88 83.8 888 3 352m 8:; 3; 8.3V :8 88: 8.3 88b 28 8.5 8.3 $822868 :8 83. 8:2 as? 3.: assume 83:: 2 am $88 :88 $32 $58 eczema a3 8.: a? 86 2... 8.3 3m 88 3 8d 2:. 0: omega. 8.8 No 8.9 3. 8.3 8.2 8.3 3 8i 3 858583. 83-38 2%. 7%.: 83-8.8 82-2. 5 88. 786 8.3%.? 3. 72.8 8.35.? mm. 72.3 3. 23a :3 owes 5:283. 8388 888.8 $888 8388 888.8 9: 58233 8.82 888 8.82 8.8 882 C s 8.8 8.8 8.8 28 8.8 C a 8.88 8.8 888 888 8.8 C a $2 8.8 on a 2m 32 :3 o 8.3 8.3 m: m9. 2% 2V 8 8.3 3... 3» 3m 3m 9: a mcommcoEmw :00 “ED n N N _ s 85 .aeafioev N a: €12 .888 8a ~22 9.888% 3833.9 3an 9.329 maid. 382-me .5585: 8883.20 8883.20 8802920 8883-20 can? 23:8 53.5 225.2%: 2: a 8:88.... 838:8 as. 5.x .3.» 2%... 178 8888 8.888 9888 8808 8.888 888 858 8.8 8.8 8.8 8 8.8 8.88 88388885 8 8 8 8.8 8 888 883888 888288888 8.88 3. 88 8.8 8.8 8.88 88322 8.88 8.88 8.88 8.88 8.88 8.88 82888 882 SE gégnomsam 888.8 888._ 888.8 8.8 8888 C 2888 8888 888.8 38.8 88.8 :88 88.8 88 5882 8888 88888 888 9828 8888 88888 88: 8 88888 8882 8888 8888882 8888 88 888 88 88 _8_ sea 88 .882 00.8 8.88 2.8 8.88 88.8_ _.88 8.88 88 .53 8.88 8.8_ 8.8_ 88.: 8.88 8.88 s. 8 888.88 888.88 8 8.88 888.88 888.88 :8 8928.88 88282 3888-858 >388 8888.29 888.838 3888-838 888888282 88888288 _ o 8888288 J 88888288 G 8888288 J 8.58.8.5 .8385 58.55 8:88.858: 85 88 8888838 88858588.. 8.5258 8.5 8.88.8 179 £288 8532082 88— 05 8 88.888 88885583 5 8823/ 2.88: 8.8 8.88: _.8 8.88: _.8 8.88: 8.8 2.8: 8.8 8.88: 8.8 8.8: seem 8.88: 8.88 8.88: 8.88 8.88: 8.88 88.88: 8.88 2.88: 8.88 8.88: 8.88 8888328880 88888 8888_ 88888 8888_ 88888 888 88888882 88885 :888_ 88888 8882 88888 :88: 88888 8888882 as... 88.8: 8.88 8.8: 8.88 8.8: 8.88 88.8: 8.8_ 88.8: 8.88 8.8: 8.88 88 8882.8. 8.8: 8.8 8.: 8.8 8.8: 8.8 8. : 8.8 8.8: 8.8 8.8: 8.8 8888888888 888-88.: 88.88.: 28.788: 888.788.: 888-88.: 88:88.: 8. 78.88 8. 78.88 8.8-8.88 8. 8.8.8 8. _.8.88 8.78.88 9: 8882 888288 8888.8 88888.8 88888.8 8888._ 888._ 88888.8 9: 58828883 8.88 8.88 8.88 8.88 8.88 8.88 C 8 8.82 8.88 8.88 8.88 8.888 8.88 C 8 8.88_ 8.88 8.88 8.88 _.88_ 8.88 C 8 8.88 8.88 888 8.88 8.88 8.88 8: 8 _.88 8.88 8.88 8.88 8.88 8.88 a: 8 8.88 8.88 8.88 8.88 8.88 8.88 9: 8 mzommaoemfl :00 3:3 8 8 8 8 8 8 888 88888888: 8 8 8 _ 8 Z _ 8 Z _ 8 88888888 8.88.838 8888-88 >388 388 88.88 8888.85. 8.8.88.5 Esau 5885.888 85 88 888838 88882.8 388 8838 8.5 8.888 180 888 858 2858 8888 8888 8.8.88 8.8.88 8E8 8.8 8.8 8.8 8.8 8.8 8.8 8.8: 88388885 8.88 8.8 8.8 8.8 8 8 8 883888 8888888: 8.8 8.8 8.8 8.8 8.8 8.8 88: 883888. 8.88 8.88 8.88 8.88 8.88 8.88 88: 82288 8882 SE 5585858838 88._ 88._ 88.8 82 88.8 88._ C 2888 888 88.8 888.8 88.8 888.8 88.8 888.8 9: 5882 888m 88882 858 8828 88888 88: 88:8 88888 88888 88888 8888 8888882 .988 888 :8 888 88_ 888 888 as; 88 .682 00.. 8.8_ 8.88 8. 88 8.88 8.88 8.88 88: 88 .53 8.88 8.8_ 88. 8.88 8.8_ 8.88 8.8: 8 m 8.88 8.88 8.88 _.88 8.88 8.88 8: 8388.88 882838 888838 8.88888 >588 3.88 8388 8888-858 838.83 8:55:— =mm